Next Article in Journal
Research on Control Strategy of Stainless Steel Diamond Plate Pattern Height Rolling Based on Local Constraints
Previous Article in Journal
Comprehensive Physicochemical Analysis of Polyphosphate-Modified Alginate Matrices: Synthesis, Structural Analysis, and Calcium Ion Release Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanisms Underlying Phase Transition and Regulation of Tantalum Powder Properties During Magnesium Thermal Reduction of Ta2O5 in a Molten Salt Medium

1
School of Material Science and Engineering, Zhengzhou University, Zhengzhou 450001, China
2
Zhongyuan Critical Metals Laboratory, Zhengzhou University, Zhengzhou 450001, China
3
School of Thermal Engineering, Shandong Jianzhu University, Jinan 250101, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(5), 1115; https://doi.org/10.3390/ma18051115
Submission received: 26 January 2025 / Revised: 21 February 2025 / Accepted: 24 February 2025 / Published: 1 March 2025

Abstract

:
Magnesium reduction of Ta2O5 (tantalum pentoxide) is a metallurgical process widely used to extract metallic tantalum powder from its oxide form using magnesium as a reducing agent in a molten salt medium. This study explores the mechanisms and patterns of phase transformation during the magnesium reduction of Ta2O5 in a molten salt medium, focusing on the influence of temperature and time on the physical and chemical properties of the resulting tantalum powder. The results reveal that under various reaction conditions in a molten salt medium, the magnesium reduction of Ta2O5 follows four distinct pathways: Ta2O5 → Ta, Ta2O5 → MgTa2O6 → Ta, Ta2O5 → MgTa2O6 → Mg4Ta2O9 → Ta, and Ta2O5 → Mg4Ta2O9 → Ta. Each pathway significantly affects the physical and chemical properties of the resulting tantalum powder. Using a uniform mixing method, the reaction proceeds directly from Ta2O5 to Ta powder in a single step. As the reaction temperature increases from 600 °C to 900 °C, the average particle size of the tantalum powder enlarges from 30 nm to 150 nm, with the product phase transitioning from a mixture of Ta and Ta2O to a single Ta phase. Additionally, prolonged holding time improves the uniformity of the tantalum powder’s particle distribution. This study accomplishes directional control over the phase transformation and the properties of tantalum powder during the reduction process, thus offering valuable guidance for the preparation of high-performance tantalum powder.

Graphical Abstract

1. Introduction

Tantalum, a strategic critical metal, is known for its high density, high melting point, and high thermal conductivity [1]. As a valve metal, tantalum forms a stable oxide film on its surface, providing excellent corrosion resistance and a high dielectric constant. Owing to these unique properties, it is widely used in industries such as chemical engineering, electronics, steel smelting, aerospace, and biomedical applications [2,3,4]. The primary use of tantalum is in the production of tantalum capacitors, which account for approximately 50–60% of its consumption [5,6,7]. Compared to other types, tantalum capacitors offer advantages such as large capacity, compact size, wide operating temperature range, and high reliability, making them indispensable in electronic devices, communication equipment, military applications, and aerospace industries [8,9,10,11].
The performance of tantalum capacitors is directly influenced by the physical and chemical properties of tantalum powder, which are closely related to its preparation process. The tantalum powder can be synthesized through various methods, including sodium reduction of potassium fluorotantalate [12,13,14], magnesium reduction of tantalum oxide [15,16,17], carbothermal reduction [18,19,20], FFC electrolysis [21,22], chloride reduction [23,24], and homogeneous reduction [25]. Of these methods, only the sodium reduction of potassium fluorotantalate and the magnesium reduction of tantalum oxide have been industrialized. Currently, the sodium reduction method is widely used, where potassium fluorotantalate is reduced using liquid sodium as the reducing agent in the presence of a halide salt diluent, producing tantalum powder with a relatively high specific surface area. However, this method has limitations of a lengthy process, difficulty in controlling the reaction, uneven particle size of tantalum powder, high impurity content, and environmental concerns due to the toxic and harmful fluorides produced, making it unsuitable for meeting the high-pressure and high-reliability requirements of tantalum capacitors [26,27,28].
For the synthesis of metallic tantalum powder, researchers have mainly used gaseous magnesium reduction of tantalum oxide and the magnesiothermic self-propagating method. Shekhter et al. [29,30]. conducted extensive research on the preparation of tantalum powder via gaseous magnesium reduction of tantalum oxide and found challenges of controlling magnesium vapors during the reaction process and imposing stringent equipment requirements, which make it difficult to ensure a homogeneous reaction environment, thereby limiting its widespread application. Hwang et al. discovered that in the gaseous magnesium reduction process, tantalum oxide undergoes a transition from Ta2O5 to Ta2O to Ta, where single oxide particles are gradually reduced from the outside in [15]. In 1998, H.C. Starck, a German company, began researching gaseous magnesium reduction tantalum powder preparation technology and realized mass production in 2006, monopolizing the global high-pressure, high-capacitance tantalum powder market. The developed magnesium-reduced tantalum powder particles are uniform and the formation voltage can be up to 400 V. In the same excitation voltage, the magnesium-reduced tantalum powder has a higher specific capacitance than the sodium-reduced tantalum powder by roughly 20–30%, which is of significant superiority [31,32], while the domestic has not yet overcome the technology [33]. Professor Zhang’s team at Northeastern University found that the magnesiothermic self-propagating method is a highly exothermic reaction, leading to rapid and incomplete reduction reactions, resulting in low-valent oxides and composite oxides in the product that cannot be easily removed by wet processing [34]. Orlov et al. reported that magnesium tantalate often forms during the magnesiothermic self-propagating process and tantalum powder derived from magnesium tantalate reduction has a larger specific surface area than that obtained from direct tantalum oxide reduction [35].
In order to produce high-performance tantalum powder with controllable particle size and uniform particles, based on these methods, for the difficult problem of difficult to regulate the magnesium thermal reaction process, our team, in collaboration with Ningxia Orient Tantalum Industry Co., Ltd. (Shizuishan, China), proposed and developed a method of magnesium reduction of tantalum oxide in a molten salt medium. This method effectively absorbs the exothermic reactions’ heat, offering advantages in reaction control and powder properties regulation.
At present, due to the lack of relevant basic research, the evolution of the physical phase in the process of magnesium reduction of tantalum oxide is not clear, which causes great difficulties in the removal of oxides in tantalum powder and the regulation of uniformity [34,35]. In this paper, the production of tantalum powder by reduction of tantalum oxide using magnesium as a reducing agent in a molten salt medium is studied, and the reduction rate is controlled by varying material distribution methods. This research systematically explores phase evolution during the magnesium reduction of tantalum oxide, playing a crucial role in understanding the magnesium reduction mechanism and controlling the reduction process and the final properties of the resulting tantalum powder.

2. Experimental Section

2.1. Materials

In this study, tantalum pentoxide (≥99.5%; Jiujiang Nonferrous Metals Smelting Co., Ltd., Jiujiang, China) was used as the raw material, magnesium metal (99.8%; Tangshan Weihe Magnesium Powder Co., Ltd., Tangshan, China) as the reducing agent, and potassium chloride (99.8%; Shanghai Macklin Biochemical Technology Co., Ltd., Shanghai, China) as the diluent. The microstructure and composition analysis of tantalum oxide, as depicted in Figure 1, reveal that the primary phase of the raw material is Ta2O5, with particle diameters ranging from 150 to 300 nm. These particles are bonded together, forming a honeycomb-like agglomerated structure.
The phase composition of the raw materials and reaction products was analyzed using an X-ray diffractometer (XRD) (PANalytical Empyrean, Almelo, The Netherlands) with a step size of 0.02° over the 2θ range of 10–90°. The oxygen content in the products was determined by an oxygen/nitrogen/hydrogen analyzer (NCS, Beijing, China), employing a pulse heating-infrared absorption method with ppm-level accuracy. The microstructure and elemental composition of the raw materials and synthesized products were analyzed using scanning electron microscopy (SEM) (ZEISS, Oberkochen, Germany) equipped with an energy-dispersive spectroscopy (EDS) system (resolution: 0.8 nm). Particle size distribution of tantalum powder was analyzed after 10 min of ultrasonic dispersion using a laser particle size analyzer (Malvern, Malvern, UK) with a measuring range of 0.02–2000 μm.

2.2. Methods

To control the reaction rate and analyze intermediate phases, three distinct material placement methods were designed, as illustrated in Figure 2. In placement method A, the crucible is sequentially layered from top to bottom with magnesium powder, potassium chloride, and tantalum pentoxide (Figure 2a). In placement method B, the lower section of the crucible contains magnesium powder, the upper layer contains tantalum pentoxide, and potassium chloride is placed between them (Figure 2b). In placement method C, all three raw materials are uniformly mixed together (as shown in Figure 2c).
Using three distinct material placement methods, specific amounts of magnesium, potassium chloride, and tantalum pentoxide were weighed and loaded into an alumina crucible (30 mm in diameter and 60 mm in height). The alumina crucible containing the materials was then placed into a tube furnace. The tube furnace was evacuated multiple times and subsequently heated in an argon atmosphere at a rate of 10 °C/min to a predetermined temperature, where it was held from 0 to 5 h. After the holding time, the furnace was cooled to room temperature.
The product was thoroughly washed with distilled water (multiple times) to remove potassium chloride from the reduction products. The washed solid product was then filtered and separated, followed by stirring and soaking in a 10% nitric acid solution for 1 h to eliminate magnesium oxide impurities. Finally, the acid-washed solid product was rinsed with distilled water until the rinse water was neutral and then dried at 60 °C for 12 h.

3. Results and Discussion

3.1. Thermomechanical Analysis

Due to the multiple valence states of tantalum, various chemical reactions can occur during the magnesium reduction of tantalum pentoxide, as expressed in Equation (1) through (6). The standard Gibbs free energy change and standard enthalpy change for these reactions were calculated at different temperatures using the reaction module of the thermodynamic calculation software Factsage 8.3, as summarized in Figure 3. Figure 3a shows that in a temperature range of 25 to 1500 °C, both tantalum pentoxide and the lower valence tantalum oxides (TaO2 and TaO) can be reduced to tantalum, whereas the reduction of Ta2O5 to TaO2 and TaO does not occur spontaneously. As the temperature increases, the standard Gibbs free energies of magnesium reduction of tantalum oxides of different valence states are closer to the 0 scale; the lower reaction temperatures are favorable for the reduction reaction.
T a 2 O 5 + Mg = 2 Ta O 2 + MgO
T a 2 O 5 + 3 Mg = 2 TaO + 3 MgO
T a 2 O 5 + 5 Mg = 2 Ta + 5 MgO
Ta O 2 + Mg = TaO + MgO
Ta O 2 + 2 Mg = Ta + 2 MgO
TaO + Mg = Ta + MgO
Figure 3b shows that the standard enthalpy values for reactions (3) through (6) are negative, indicating that these reactions are exothermic in nature. As the temperature rises, the energy states of the reactants and products change, leading to increased exothermic heat. To improve reaction efficiency and mitigate the negative effects of uneven tantalum powder growth caused by the exothermic reaction, this study comprehensively investigates the magnesium reduction of tantalum pentoxide at temperatures ranging from 500 to 1000 °C.
Additionally, the effect of potassium chloride and magnesium content on the adiabatic temperature of the Ta2O5-Mg-KCl system was analyzed using Factsage, as illustrated in Figure 4.
When the potassium chloride content is zero and the magnesium content is present at the theoretical amount, the system’s adiabatic temperature reaches its highest value of 2825 °C, which is the melting point of magnesium oxide. As the potassium chloride content increases with the magnesium, the system’s maximum adiabatic temperature gradually decreases, due to the addition of a lower melting point constituent. Due to the latent heat of phase transitions, two significant temperature plateaus occur at the boiling points of potassium chloride (1475 °C) and magnesium (1095 °C). Notably, at the boiling point of magnesium, changes in the adiabatic temperature of the reaction system occur only with substantial variations in the amounts of magnesium and potassium chloride. Considering the adiabatic temperature and the amounts of magnesium and potassium chloride, a system with an adiabatic temperature of 1095 °C was selected for the study of magnesium reduction process in detail.

3.2. Evolution of the Physical Phase During Magnesium Reduction

The phase evolution at 900 °C was investigated under different material placement methods and holding times. Figure 5 presents the phase composition of the primary reduction products and the acid-washed products under placement method A.
As shown in the XRD patterns in Figure 5a, extending the holding time gradually eliminates the Ta2O5 phase in the product, and after 5 h, the developed phases include Ta, MgTa2O6, KCl, and MgO. A comparison of the products before and after acid washing under the same conditions reveals that KCl and MgO are washed out in acid, while Mg4Ta2O9 emerges along with existing MgTa2O6. This indicates that after KCl dissolves in water and MgO dissolves in acid, the relative content of Mg4Ta2O9 increases, leading to its stronger diffraction peaks. It also suggests that tantalum oxides such as Mg4Ta2O9, MgTa2O6, and Ta2O5 are insoluble in dilute nitric acid. In Figure 5b, it is evident that without holding at 900 °C, the reduction products consist of Ta, MgTa2O6, and Ta2O5, where a part of the tantalum oxide is reduced to tantalum, and magnesium oxide combines with unreacted Ta2O5 to form MgTa2O6. In placement method A, the limited contact between magnesium and tantalum oxide slows the reaction rate, leaving a significant amount of tantalum oxide unreduced even after 5 h. The formation of magnesium tantalate can be explained by the possible reactions expressed in Equations (7)–(9) [36,37,38]. Due to the lack of thermodynamic data for magnesium tantalate, the feasibility of these reactions cannot be verified through thermodynamic calculations. However, it can be shown by Baskin et al. that the Mg4Ta2O9 and MgTa2O6 phases can exist stably, and the two intermediate phases of magnesium tantalate produced during the reduction process in this paper are stable phases [39].
T a 2 O 5 + MgO = MgT a 2 O 6
T a 2 O 5 + 4 MgO = M g 4 T a 2 O 9
MgT a 2 O 6 + 3 MgO = M g 4 T a 2 O 9
Figure 6 shows the SEM microstructure of the acid-washed products obtained at different holding times for placement method A. The EDS analysis of the products under different conditions reveals that at 900 °C without any holding time, the acid-washed products consist of large plate-like MgTa2O6, smaller particles of tantalum, and tantalum oxide, with the tantalum obtained after reduction largely retaining the original morphology of the Ta2O5. As the holding time increases, the magnesium tantalate grows to some extent, while the tantalum powder particles show minor changes.
Figure 7 presents the crystallinity and phase composition of the products obtained under different holding times for placement method B. Similar to placement method A, the water and acid washing effectively remove magnesium oxide and potassium chloride. Without holding time, the reduction products consist of unreacted Ta2O5, Ta, MgTa2O6, and Mg4Ta2O9. When the holding time is increased to 1 h, all Ta2O5 converts to Ta, Mg4Ta2O9, and MgTa2O6, with Mg4Ta2O9 being the product of a further combination of tantalum oxide and magnesium oxide. The presence of Ta2O after washing suggests slight surface oxidation of Ta during the acid-washing process. Extending the holding time beyond 3 h results in only the tantalum phase remaining, indicating that magnesium tantalates (Mg4Ta2O9 and MgTa2O6) can also be reduced to tantalum, as indicated by the possible reactions in Equations (10) and (11) [35,40].
Mg Ta 2 O 6 + 5 Mg = 2 Ta + 6 MgO
M g 4 T a 2 O 9 + 5 Mg = 2 Ta + 9 MgO
Compared to placement method A, method B reveals higher reduction efficiency. It is conceived that in method B, tantalum pentoxide particles easily move downward to contact magnesium after passing through the molten salt layer, whereas in method A, liquid magnesium, hindered by surface tension, struggles to penetrate the molten salt layer to react with tantalum pentoxide, resulting in losses due to vaporization.
Figure 8 presents the microstructure of the products obtained under different holding times for placement method B. Unlike method A, where MgTa2O6 appears in large, plate-like structures, MgTa2O6 in method B is blocky and smaller in size. Additionally, in method B, Ta2O5 and Ta particles have similar sizes. When the holding time exceeds 3 h, both magnesium tantalate phases are completely reduced to tantalum, and the size of the reduced tantalum powder particles remains largely unaffected by a further increase in holding time.
Figure 9 illustrates the phase transition at 900 °C with different holding times under placement method C. At this temperature, the products consist solely of the pure metallic tantalum phase, indicating that the tantalum pentoxide has been fully reduced with no intermediate phases observed, suggesting that the intermediate phase is unstable or does not have the conditions for formation in this reduction method. Investigation into these possible mechanisms is ongoing.
Figure 10 and Table 1 summarize a comparison of the particle size distribution characteristics of tantalum powder obtained at 900 °C under different holding times for placement methods B and C. As in method B, a minimum of 3 h is required for a stable Ta phase formation (however, a minimum of 0 h is sufficient in case of uniform mixing (method C)); therefore, it is compared the particle size distribution of Ta powder successfully synthesized in methods B and C at different holding times.
The SPAN value calculated using Equation (12) represents the uniformity of the tantalum powder particle size, and the smaller SPAN value represents the more uniform powder size.
SPAN = D 90 - D 10 D 50
In the equation, D10, D50, and D90 represent the particle sizes at which 10%, 50%, and 90% of the total powder volume, respectively, have a particle size smaller than these values, with the units being micrometers (μm).
As is evident from the figures and tables, under the same holding time, the SPAN value of tantalum powder obtained under uniform mixing conditions is 0.46, which is significantly smaller than 0.60 when magnesium is placed in the lower layer, indicating a more concentrated particle size distribution. When magnesium is in the lower layer, multiple phase evolution pathways exist (Ta2O5 → Ta and Ta2O5 → MgTaxOy → Ta), resulting in more small and large particles, which leads to an uneven particle size distribution. At the same temperature, by extending the holding time, the SPAN value of tantalum powder under uniform mixing conditions decreases from 0.46 to 0.43, and the SPAN value of tantalum powder obtained by reduction with magnesium in the lower layer decreases from 0.60 to 0.57. This results in a more concentrated particle size distribution and improved uniformity of the tantalum powder, although the change in particle size is relatively small. The increase in tantalum powder size mainly depends on the nucleation and growth mechanisms during the reduction process.
Based on the experimental results, the phase transformation patterns during the magnesium reduction of tantalum pentoxide are summarized as illustrated in Figure 11. As magnesium diffuses to contact Ta2O5, uneven diffusion creates regions with varying concentrations, some with an excess and others with a deficiency. In regions with excess magnesium, Ta2O5 is completely reduced to Ta and MgO, whereas in magnesium-deficient regions, only part of Ta2O5 is reduced to tantalum, and unreacted Ta2O5 combines with MgO to form magnesium tantalate (MgTa2O6 and Mg4Ta2O9). As magnesium continues to diffuse, sufficient magnesium in these regions allows the further reduction of magnesium tantalate and any remaining Ta2O5 to tantalum. Considering the formation of magnesium tantalate, four possible reaction pathways during the magnesium reduction of tantalum oxide are shown in Figure 11: Ta2O5 → Ta; Ta2O5 → MgTa2O6 →Ta; Ta2O5 → MgTa2O6 → Mg4Ta2O9 → Ta; and Ta2O5 → Mg4Ta2O9 → Ta.
Notably, all four reaction pathways are observed in placement methods A and B, whereas placement method C is dominated exclusively by pathway 1 (Ta2O5 → Ta). According to the findings of Orlov et al., ensuring the consistency of the reaction pathway in the magnesium reduction process is crucial for achieving a uniform particle size distribution of tantalum powder [35,41]. In placement methods A and B, both tantalum powder obtained from the direct reduction of tantalum oxide and the reduction of magnesium tantalate contribute to the non-uniformity of particle size. In contrast, placement method C, which is free from these additional influences, is more effective in ensuring the uniformity of the resulting tantalum powder.

3.3. Effect of Reduction Temperature and Time on the Physicochemical Properties of Tantalum Powder

In placement method C, where the materials are uniformly mixed, the reaction proceeds more homogeneously, and no intermediate phases are grown. This study further investigates the influence of temperature and time on the physical and chemical properties of tantalum powder synthesized under uniform mixing conditions (method C).
Figure 12 illustrates the effect of temperature on phase transformation during the reduction process. The results indicate that no reaction occurs at 500 °C. However, at 600 °C, the tantalum phase begins to form in the system without any magnesium tantalates. A comparison of the phases before and after acid washing reveals that the diffraction peaks of the tantalum phase split into double peaks corresponding to Ta and Ta2O. This double peak effect becomes more pronounced at lower reduction temperatures, indicating incomplete deoxygenation. Due to the smaller particle size and higher surface activity, tantalum powder is more prone to oxygen absorption, resulting in Ta2O (with an oxygen mass fraction of 4.23%) formation during water washing. As the reduction temperature increases, the double peak phenomenon disappears, and only the tantalum phase remains in the washed products.
Figure 13a shows the variation of oxygen content at different reaction temperatures. As can be seen from the figure, the oxygen content in the tantalum powder gradually decreases as the reaction temperature increases. The activation energy of the reaction was calculated from the change in oxygen content at different reaction temperatures using Equation (13), as shown in Figure 13b.
k = Ae E a / ( R T )
where k denotes the reaction rate constant, A denotes the constant, R denotes the gas constant, T denotes the absolute temperature, and Ea denotes the activation energy. The oxygen content in tantalum pentoxide is 18.1%, and assuming that the reduction in oxygen content corresponds to the reaction rate constant k, i.e., the value of k is proportional to the reduction in oxygen content, the reduction in oxygen content at each temperature is taken as the reaction rate constant k. From this, the activation energy of the reaction is calculated to be 3.856 kJ/mol. As the reaction is exothermic, the reaction itself releases heat, providing the molecules of the reactants with additional energy, making it easier for them to reach an active state and react chemically, resulting in a lower activation energy for the reaction.
Figure 14 shows the morphology of the products before and after acid washing synthesized under uniform mixing conditions (placement method C) at different reaction temperatures.
In Figure 14a,c,e,g, the interaction between magnesium and tantalum pentoxide initiates surface reactions and nucleation, forming tantalum powder particles. As the reaction temperature increases, both tantalum powder and magnesium oxide particles grow significantly. At 600 °C, fine tantalum powder adhering to the surface of magnesium oxide is formed. With a further increase in temperature, the number and size of tantalum powder particles increase markedly, with magnesium oxide particles growing more prominently into regular block shapes.
Figure 14b,d,f,h compares the morphology of the products after water washing. At 600 °C, the tantalum powder obtained is fine and weakly agglomerated, with a uniform morphology and a particle size of around 30 nm. As the reduction temperature increases, the tantalum powder particles grow significantly, reaching a size range of 50 to 150 nm at 900 °C. The synthesized particles of tantalum powder exhibit more regular shapes, better crystallinity, and improved dispersion. This improvement may be attributed to the decreased viscosity of the molten salt at higher temperatures, which reduces resistance to mass transfer and facilitates particle coalescence in the molten salt.
Figure 15 presents the microstructure of tantalum powder before and after washing with water at 900 °C with different holding times. As shown in Figure 15a,c, extending the holding time from 1 h to 5 h leads to significant growth in the magnesium oxide particles with good crystallinity. Figure 15b,d show that with longer holding times, the number of small particles in the tantalum powder decreases, resulting in enhanced uniformity. This indicates that regulating the temperature and the holding time contributes to the growth of tantalum powder with higher uniformity. Additionally, measurements of the oxygen content in tantalum powder synthesized at different holding times show that longer holding times help reduce the oxygen content. For instance, when held for 3 h, the oxygen content in the tantalum powder is 1.79%. Therefore, by carefully controlling the reaction temperature and holding time, it is possible to effectively regulate the oxygen content, uniformity, particle size, and particle size distribution of high-purity tantalum powder in the magnesium reduction reaction.

4. Conclusions

This study systematically investigated the phase transformation process during the magnesium reduction of tantalum pentoxide by varying material placement methods to adjust the reaction rate. The influence of different process parameters on the properties of tantalum powder was then studied under placement method C, based on the observed phase transformation patterns.
(1)
During the magnesium reduction process, no lower-valence tantalum oxides were formed, consistent with thermodynamic calculations. However, in placement methods A and B, regions with insufficient magnesium were present during the reaction, where the by-product magnesium oxide was easily combined with tantalum pentoxide to form magnesium tantalate (MgTa2O6 and Mg4Ta2O9). As the reaction proceeded in magnesium-rich conditions, these magnesium tantalates (MgTa2O6 and Mg4Ta2O9) were further reduced to tantalum, resulting in multiple reaction pathways (Ta2O5 → Ta and Ta2O5 → MgTaxOy → Ta). In contrast, in placement method C, tantalum pentoxide was directly reduced to high-purity metallic tantalum, following a single reaction pathway (Ta2O5 → Ta).
(2)
Tantalum powder obtained via the single pathway (placement method C) had a more concentrated particle size distribution, with fewer small particles and better particle uniformity, compared to the powder obtained via multiple pathways (placement method B). To ensure uniformity in tantalum powder particles, maintaining consistency in the reaction pathway during the reduction process is essential.
(3)
Under placement method C, the magnesium reduction of Ta2O5 occurred only at temperatures above 600 °C. At lower temperatures, the resulting particles were finer and had higher activity, with increased oxygen content after water washing, resulting in the presence of Ta2O and Ta phases instead of pure Ta. At 900 °C, a pure tantalum phase was obtained. As the reduction temperature increased from 600 °C to 900 °C, the particle size of the tantalum powder gradually increased from 30 nm to 150 nm, but the particle uniformity decreased. However, extending the holding time improved the uniformity of the tantalum powder morphology, and the oxygen content in the tantalum powder decreased to 1.79% after a holding time of 3 h.
(4)
The in-depth study of the phase transformation laws during the magnesium reduction of tantalum pentoxide and the influence of process parameters on the properties of tantalum powder will provide an important theoretical basis and technical support for the production and application of tantalum powder. In addition, this research can serve as a reference for studies on other metal reduction processes. Furthermore, by focusing on the magnesium tantalate formed during the reaction, our team aims to further synthesize a single phase of magnesium tantalate by solid-state synthesis, refine its thermodynamic parameters, and reduce it in order to further investigate its effect on the properties of tantalum powder.

Author Contributions

Conceptualization, Y.C. (Yusi Che) and J.H.; methodology, Y.C. (Yusi Che), J.H., Y.C. (Yi Chen) and Z.H.; software, T.L. and C.Z.; formal analysis, C.Z.; investigation, Y.C. (Yi Chen), T.L. and Z.H.; visualization, Y.C. (Yi Chen); resources, R.W.; data curation, R.W. and Y.C. (Yusi Che); writing—Y.C. (Yi Chen); writing—review and editing, R.W. and Y.C. (Yusi Che); project administration, J.H.; funding acquisition, R.W. and Y.C. (Yusi Che). All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by the National Natural Science Foundation of China (Grant No. U23A20609 and 52304371), Project of Zhongyuan Critical Metals Laboratory (Grant No. GJJSGFJQ202305), China National Postdoctoral Program for Innovative Talent (Grant No. BX20230327), and Key Scientific Research Project of Colleges and Universities in Henan Province (Grant No. 24A450002).

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

The authors declare no conflict of interest.

References

  1. Buckman, R.W. New applications for tantalum and tantalum alloys. JOM 2000, 52, 40–41. [Google Scholar] [CrossRef]
  2. Mohsan, A.U.H.; Wei, D. Advancements in Additive Manufacturing of Tantalum via the Laser Powder Bed Fusion (PBF-LB/M): A Comprehensive Review. Materials 2023, 16, 6419. [Google Scholar] [CrossRef] [PubMed]
  3. Ogi, T.; Horiuchi, H.; Makino, T.; Arif, A.F.; Okuyama, K. Simple, Rapid, and Environmentally Friendly Method for Selectively Recovering Tantalum by Guanidine-Assisted Precipitation. ACS Sustain. Chem. Eng. 2018, 6, 9585–9590. [Google Scholar] [CrossRef]
  4. Wang, Y.; Qin, X.; Lv, N.; Gao, L.; Sun, C.; Tong, Z.; Li, D. Microstructure Optimization for Design of Porous Tantalum Scaffolds Based on Mechanical Properties and Permeability. Materials 2023, 16, 7568. [Google Scholar] [CrossRef] [PubMed]
  5. He, J.L.; Zhang, X.Q.; Yang, G.Q.; Zheng, A.G. The Key Technique and R&D of Capacitor Grade Tantalum Powder. Materials China 2014, 33, 545–553. [Google Scholar]
  6. Niu, B.; Chen, Z.; Xu, Z. Recovery of Valuable Materials from Waste Tantalum Capacitors by Vacuum Pyrolysis Combined with Mechanical-Physical Separation. ACS Sustain. Chem. Eng. 2017, 5, 2639–2647. [Google Scholar] [CrossRef]
  7. Yang, Q.; Furigay, M.H.; Chaudhuri, S.; Gau, M.R.; Schatz, G.C.; Schelter, E.J. Toward Redox- and Photoredox-Based Niobium/Tantalum Separations. ACS Sustain. Chem. Eng. 2024, 12, 8503–8511. [Google Scholar] [CrossRef]
  8. Freeman, Y.; Lessner, P. Evolution of Polymer Tantalum Capacitors. Appl. Sci. 2021, 11, 5514. [Google Scholar] [CrossRef]
  9. Micheau, C.; Arrachart, G.; Turgis, R.; Lejeune, M.; Draye, M.; Michel, S.; Legeai, S.; Pellet-Rostaing, S. Ionic Liquids as Extraction Media in a Two-Step Eco-Friendly Process for Selective Tantalum Recovery. ACS Sustain. Chem. Eng. 2020, 8, 1954–1963. [Google Scholar] [CrossRef]
  10. Pan, L.T.; Li, B.; Zheng, A.G.; Lu, J.B.; Li, H.J.; Wang, Q.Y. Application of tantalum to lsic. Chin. J. Rare Met. 2003, 27, 28–34. [Google Scholar]
  11. Niu, B.; Chen, Z.; Xu, Z. Recovery of Tantalum from Waste Tantalum Capacitors by Supercritical Water Treatment. ACS Sustain. Chem. Eng. 2017, 5, 4421–4428. [Google Scholar] [CrossRef]
  12. Lee, Y.; Kim, H.; Kwon, N.; Sim, J.; Lee, M.; Oh, S.; Seo, S.; Shin, J.; Park, K. Effects of reduction temperature and diluent content on the properties of high-purity tantalum powder prepared using the Hunter process. Int. J. Refract. Met. Hard Mater. 2023, 112, 106128. [Google Scholar] [CrossRef]
  13. Sim, J.-J.; Choi, S.-H.; Lee, Y.-K.; Heo, S.G.; Kim, T.-S.; Seo, S.-J.; Park, K.-T. Consideration of Diluents Selection and Input Amounts of the Hunter Process for Tantalum Production. Met. Mater. Int. 2021, 27, 1980–1987. [Google Scholar] [CrossRef]
  14. Yoon, J.S.; Kim, B.I. Characteristics and production of tantalum powders for solid-electrolyte capacitors. J. Power Sources 2007, 164, 959–963. [Google Scholar] [CrossRef]
  15. Hwang, S.-M.; Park, S.-J.; Wang, J.-P.; Park, Y.-H.; Lee, D.-W. Preparation of tantalum metal powder by magnesium gas reduction of tantalum pentoxide with different initial particle size. Int. J. Refract. Met. Hard Mater. 2021, 100, 105620. [Google Scholar] [CrossRef]
  16. Ryu, H.S.; Nersisyan, H.; Park, K.T.; Lee, J.H. Porous tantalum network structures exhibiting high electrochemical performance as capacitors. J. Energy Storage 2021, 34, 102222. [Google Scholar] [CrossRef]
  17. Lee, Y.-K.; Sim, J.-J.; Byeon, J.-S.; Lee, Y.-T.; Cho, Y.-W.; Kim, H.-C.; Heo, S.-G.; Lee, K.-A.; Seo, S.-J.; Park, K.-T. Production of high-purity tantalum metal powder for capacitors using self-propagating high-temperature synthesis. Arch. Metall. Mater. 2021, 66, 935–939. [Google Scholar] [CrossRef]
  18. Brar, L.K.; Singla, G.; Kaur, N.; Pandey, O.P. Thermal stability and structural properties of Ta nanopowder synthesized via simultaneous reduction of Ta2O5 by hydrogen and carbon. J. Therm. Anal. Calorim. 2015, 119, 175–182. [Google Scholar] [CrossRef]
  19. Liu, D.; Liu, J.; Ye, P.; Zhang, H.; Zhang, S. Low-Temperature, Efficient Synthesis of Highly Crystalline Urchin-like Tantalum Diboride Nanoflowers. Materials 2022, 15, 2799. [Google Scholar] [CrossRef]
  20. Hwang, S.-M.; Hong, J.-W.; Park, Y.-H.; Lee, D.-W. Synthesis of Tantalum Carbide Using Purified Hexane by Titanium Powder. Materials 2022, 15, 7510. [Google Scholar] [CrossRef] [PubMed]
  21. Kong, Y.P.; Xiao, Y.; Liang, X.M.; Fan, Y.Y.; Wang, L.Q.; Yang, S.; He, J.L.; Chen, Y.R. Facile synthesis of nanoscale tantalum powder by electro-deoxidation of Ta2O5. Int. J. Refract. Met. Hard Mater. 2023, 114, 106269. [Google Scholar] [CrossRef]
  22. Wang, Y.T.; Xue, Y.D.; Zhang, C.H. Electrochemical product engineering towards sustainable recovery and manufacturing of critical metals. Green Chem. 2021, 23, 6301–6321. [Google Scholar]
  23. Park, K.Y.; Kim, H.J.; Suh, Y.J. Preparation of tantalum nanopowders through hydrogen reduction of TaCl5 vapor. Powder Technol. 2007, 172, 144–148. [Google Scholar] [CrossRef]
  24. Niu, B.; Chen, Z.; Xu, Z. Method for Recycling Tantalum from Waste Tantalum Capacitors by Chloride Metallurgy. ACS Sustain. Chem. Eng. 2017, 5, 1376–1381. [Google Scholar] [CrossRef]
  25. Zhu, H.M.; Sadoway, D.R. Synthesis of nanoscale particles of Ta and Nb3Al by homogeneous reduction in liquid ammonia. J. Mater. Res. 2001, 16, 2544–2549. [Google Scholar]
  26. Nersisyan, H.; Ryu, H.S.; Lee, J.H.; Suh, H.; Won, H.I. Tantalum network nanoparticles from a Ta2O5+kMg system by liquid magnesium controlled combustion. Combust. Flame 2020, 219, 136–146. [Google Scholar] [CrossRef]
  27. Zhang, X.; Guo, S.; Yang, G.; Xie, Q.; Zuo, J.; Luo, W. Microstructure and Electrical Properties of Tantalum Powder with Different Calcining Technology. Chin. J. Rare Met. 2018, 42, 67–74. [Google Scholar]
  28. Yue, R.; Zhou, L. Effect of Refiner Addition on Particle Size and Electrical Properties of FTB-48 Capacitor Tantalum Powder. Rare Met. Cem. Carbides 2020, 48, 18–21. [Google Scholar]
  29. Shekhter, L.N.; Tripp, T.B.; Lanin, L.L.; Conlon, A.M. Metalothermic Reduction of Refractory Metal Oxides. U.S. Patent 2004/0163491, 26 August 2004. [Google Scholar]
  30. Shekhter, L.N.; Tripp, T.B.; Lanin, L.L.; Reichert, K.; Thomas, O.; Vieregge, J. Metal Powders Produced by the Reduction of the Oxides with Gaseous Magnesium. U.S. Patent 6558447, 6 May 2003. [Google Scholar]
  31. Hagymási, M.; Otterstedt, R.D.; Schnitter, C.; Thomas, O.; Reifenheiser, R. Pushing Tantalum capacitors to the limit: View to 300 V anodisations and beyond. In Proceedings of the Space Passive Component Days (SPCD), Noordwijk, The Netherlands, 9–12 October 2018. [Google Scholar]
  32. Hagymási, M.; Schnitter, C.; Reifenheiser, R.; Thomas, O. Highest capacitance at higher voltages: Pushing the limits of tantalum high voltage capacitors. In Proceedings of the Space Passive Component Days (SPCD), Noordwijk, The Netherlands, 12–14 October 2016. [Google Scholar]
  33. Shekhter, L.N.; Simkins, L.F.; Greville, H.P.; Lanin, L. Method of Preparing Primary Refractory Metal. U.S. Patent 7399335, 15 July 2008. [Google Scholar]
  34. Yan, J.P.; Dou, Z.H.; Zhang, T.A.; Yan, J.S.; Zhou, X.Y. Primary Reduction Mechanism in the Process of Multi-stage Reduction to Prepare Titanium Powder. Rare Met. Mater. Eng. 2022, 51, 1470–1477. [Google Scholar]
  35. Orlov, V.M.; Kryzhanov, M.V. Magnesiothermic Reduction of Mg4Ta2O9 in the Combustion Regime. Inorg. Mater. 2018, 54, 910–914. [Google Scholar] [CrossRef]
  36. Kryzhanov, M.V.; Orlov, V.M.; Sukhorukov, V.V. Thermodynamic modeling of magnesiothermic reduction of niobium and tantalum from pentoxides. Russ. J. Appl. Chem. 2010, 83, 379–383. [Google Scholar] [CrossRef]
  37. Wu, H.T.; Yang, C.H.; Wu, W.B.; Yue, Y.L. Study on synthesis and evolution of nanocrystalline Mg4Ta2O9 by aqueous sol-gel process. Surf. Rev. Lett. 2012, 19, 1250024. [Google Scholar]
  38. Dang, M.Z.; Lin, H.X.; Yao, X.G.; Ren, H.S.; Xie, T.Y.; Peng, H.Y.; Tan, Z.Y. Effects of B2O3 and MgO on the microwave dielectric properties of MgTa2O6 ceramics. Ceram. Int. 2019, 45, 24244–24247. [Google Scholar]
  39. Baskin, Y.; Schell, D.C. Phase Studies in the Binary System MgO–Ta2O5. J. Am. Ceram. Soc. 1963, 46, 174–177. [Google Scholar] [CrossRef]
  40. Orlov, V.M.; Kryzhanov, M.V. Production of tantalum powders by the magnesium reduction of tantalates. Russ. Metall. Met. 2015, 2015, 590–593. [Google Scholar] [CrossRef]
  41. Orlov, V.M.; Kiselev, E.N. Magnesium Vapor Reduction of Tantalum Oxide Compounds in the Temperature Range 540–680 °C. Inorg. Mater. 2022, 58, 799–805. [Google Scholar]
Figure 1. SEM-EDS results (a) and XRD patterns (b) of experimental tantalum pentoxide.
Figure 1. SEM-EDS results (a) and XRD patterns (b) of experimental tantalum pentoxide.
Materials 18 01115 g001
Figure 2. Spreading of raw materials in the crucible; (a) Tantalum pentoxide on the bottom; (b) Tantalum pentoxide on the top; (c) homogeneous mixture.
Figure 2. Spreading of raw materials in the crucible; (a) Tantalum pentoxide on the bottom; (b) Tantalum pentoxide on the top; (c) homogeneous mixture.
Materials 18 01115 g002
Figure 3. Standard Gibbs free energy change (a) and standard molar enthalpy change (b) for the magnesium reduction of tantalum oxide as a function of temperature.
Figure 3. Standard Gibbs free energy change (a) and standard molar enthalpy change (b) for the magnesium reduction of tantalum oxide as a function of temperature.
Materials 18 01115 g003
Figure 4. Effect of content of magnesium and potassium chloride on the adiabatic temperature of the system: (a) 3D surface view; (b) top view; (c) main view; and (d) side view.
Figure 4. Effect of content of magnesium and potassium chloride on the adiabatic temperature of the system: (a) 3D surface view; (b) top view; (c) main view; and (d) side view.
Materials 18 01115 g004
Figure 5. XRD patterns of the products before and after acid washing (placement method A). (a) after reduction; (b) after acid washing.
Figure 5. XRD patterns of the products before and after acid washing (placement method A). (a) after reduction; (b) after acid washing.
Materials 18 01115 g005
Figure 6. SEM images showing morphology of pickling products (of placement method A) at different holding time conditions: (a) 0 h; (b) 1 h; (c) 3 h; and (d) 5 h.
Figure 6. SEM images showing morphology of pickling products (of placement method A) at different holding time conditions: (a) 0 h; (b) 1 h; (c) 3 h; and (d) 5 h.
Materials 18 01115 g006
Figure 7. XRD patterns of the products before and after acid washing (placement method B). (a) after reduction; (b) after acid washing.
Figure 7. XRD patterns of the products before and after acid washing (placement method B). (a) after reduction; (b) after acid washing.
Materials 18 01115 g007
Figure 8. SEM images showing morphology of pickling products (of placement method B) at different holding time conditions: (a) 0 h; (b) 1 h; (c) 3 h; and (d) 5 h.
Figure 8. SEM images showing morphology of pickling products (of placement method B) at different holding time conditions: (a) 0 h; (b) 1 h; (c) 3 h; and (d) 5 h.
Materials 18 01115 g008
Figure 9. XRD patterns of the products obtained after acid washing (in placement method C).
Figure 9. XRD patterns of the products obtained after acid washing (in placement method C).
Materials 18 01115 g009
Figure 10. Comparison of particle size distribution (measured using Laser particle size analyzer) of pure metallic tantalum powder synthesized at 900 °C under different holding times, in placement methods B and C.
Figure 10. Comparison of particle size distribution (measured using Laser particle size analyzer) of pure metallic tantalum powder synthesized at 900 °C under different holding times, in placement methods B and C.
Materials 18 01115 g010
Figure 11. Principle of phase transition during magnesium reduction of tantalum pentoxide. Different routes highlight the diffusion of magnesium for complete reduction of Ta2O5 to Ta.
Figure 11. Principle of phase transition during magnesium reduction of tantalum pentoxide. Different routes highlight the diffusion of magnesium for complete reduction of Ta2O5 to Ta.
Materials 18 01115 g011
Figure 12. XRD patterns of the products: (a) before and (b) after pickling at different reduction temperatures.
Figure 12. XRD patterns of the products: (a) before and (b) after pickling at different reduction temperatures.
Materials 18 01115 g012
Figure 13. Change in oxygen content at different reaction temperatures (a); Plot of ln k vs. 1/T×103 for the estimation of the activation energy of the reduced Ta powder (b).
Figure 13. Change in oxygen content at different reaction temperatures (a); Plot of ln k vs. 1/T×103 for the estimation of the activation energy of the reduced Ta powder (b).
Materials 18 01115 g013
Figure 14. SEM images showing the morphology of the products: (a,c,e,g) before and (b,d,f,h) after pickling at different reduction temperatures.
Figure 14. SEM images showing the morphology of the products: (a,c,e,g) before and (b,d,f,h) after pickling at different reduction temperatures.
Materials 18 01115 g014
Figure 15. SEM images showing morphology of the products: (a,c) before and (b,d) after pickling at 900 °C for different holding times of (a,b) 1 h and (c,d) 5 h.
Figure 15. SEM images showing morphology of the products: (a,c) before and (b,d) after pickling at 900 °C for different holding times of (a,b) 1 h and (c,d) 5 h.
Materials 18 01115 g015
Table 1. Laser particle size variation of tantalum powder under different process conditions.
Table 1. Laser particle size variation of tantalum powder under different process conditions.
Placement MethodsTime (h)D10 (μm)D50 (μm)D90 (μm)SPAN
B30.1270.1700.2290.60
B50.1290.1700.2260.57
C10.1350.1690.2130.46
C30.1360.1690.2090.43
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

Chen, Y.; Han, Z.; Li, T.; Wang, R.; Zhang, C.; Che, Y.; He, J. Mechanisms Underlying Phase Transition and Regulation of Tantalum Powder Properties During Magnesium Thermal Reduction of Ta2O5 in a Molten Salt Medium. Materials 2025, 18, 1115. https://doi.org/10.3390/ma18051115

AMA Style

Chen Y, Han Z, Li T, Wang R, Zhang C, Che Y, He J. Mechanisms Underlying Phase Transition and Regulation of Tantalum Powder Properties During Magnesium Thermal Reduction of Ta2O5 in a Molten Salt Medium. Materials. 2025; 18(5):1115. https://doi.org/10.3390/ma18051115

Chicago/Turabian Style

Chen, Yi, Zhenghao Han, Tianchen Li, Ruifang Wang, Chao Zhang, Yusi Che, and Jilin He. 2025. "Mechanisms Underlying Phase Transition and Regulation of Tantalum Powder Properties During Magnesium Thermal Reduction of Ta2O5 in a Molten Salt Medium" Materials 18, no. 5: 1115. https://doi.org/10.3390/ma18051115

APA Style

Chen, Y., Han, Z., Li, T., Wang, R., Zhang, C., Che, Y., & He, J. (2025). Mechanisms Underlying Phase Transition and Regulation of Tantalum Powder Properties During Magnesium Thermal Reduction of Ta2O5 in a Molten Salt Medium. Materials, 18(5), 1115. https://doi.org/10.3390/ma18051115

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