Next Article in Journal
Disintegration of Six Different Quartz Types during Heating to 1600 °C
Previous Article in Journal
Strain Energy Dissipation Characteristics and Neural Network Model during Uniaxial Cyclic Loading and Unloading of Dry and Saturated Sandstone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Self-Reduction during the Thermal Decomposition of an Ammonium Molybdate

1
Department of Energy & Resources Engineering, Korea Maritime & Ocean University, Busan 49112, Republic of Korea
2
Department of Environmental Science, Keimyung University, Daegu 42601, Republic of Korea
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(2), 133; https://doi.org/10.3390/min13020133
Submission received: 15 November 2022 / Revised: 4 January 2023 / Accepted: 14 January 2023 / Published: 17 January 2023
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)

Abstract

:
In the hydrometallurgical process of molybdenum using ammonia solution, ammonium paramolybdate tetrahydrate (APT: (NH4)6Mo7O24·4H2O) is produced as an intermediate product after a crystallization step. ATP is then thermally decomposed at a high temperature to give MoO3, which is reduced by hydrogen gas in a two-stage process to reduce molybdenum metal powder as the final product. If APT is pre-dried at an appropriately low temperature to remove the crystal water corresponding to 4 mol per mol of APT, it changes into (NH4)4Mo5O17, and the content of residual ammonia, which can be utilized as a reductant, in the ammonium molybdate increases. In this regard, the self-reducing potential of (NH4)4Mo5O17 was examined in this study through the effectiveness analysis of the residual ammonia component as a reductant for the primary hydrogen reduction step. In a series of experimental work on the thermal decomposition of (NH4)4Mo5O17 in an inert atmosphere, a maximum self-reduction degree of 18% was achieved. Based on this result, it can be expected that the metal powder can be manufactured in a more effective way than conventional processes in terms of hydrogen consumption and reaction time.

1. Introduction

Molybdenum is regarded as one of the key metals in Republic of Korea due to its high melting point, high strength at high temperatures, high thermal conductivity, and adequate resistance to corrosion [1,2]. Although there is an operating molybdenum mine in Republic of Korea, the self-sufficiency ratio of molybdenum in Republic of Korea remains under 2% [1]. Most molybdenum resources such as sulfides are imported, and therefore, there have been attempts to improve the molybdenum production technologies and processes in Republic of Korea [3].
In the production of molybdenum metal powder of a high purity, ammonium paramolybdate tetrahydrate (APT: (NH4)6Mo7O24·4H2O) is the first intermediate, and then it is roasted and converted into molybdenum trioxide, MoO3, which is the second intermediate. When the ammonium molybdate is heated, it decomposes in a stepwise procedure and gives off NH3 and H2O, leaving MoO3 as the final product. The stepwise decomposition of ammonium molybdate is reported in the literature as follows: (NH4)6Mo7O24·4H2O → (NH4)4Mo5O17 → (NH4)2Mo4O13 → (NH4)2Mo14O43 → (NH4)2Mo22O67 → MoO3 [4,5,6].
According to the analytical results of TGA, DTA, and XRD from the literature, the stepwise decomposition was observed in the range of 110~380 °C, and finished around 400 °C. Although the steps of the thermal decomposition might be changed by heating conditions, they are not as affected by the oxygen potential, especially in the early part, as shown in Figure 1. If APT is pre-dried at an appropriately low temperature to remove the crystal water corresponding to 4 mol per mol of APT, it changes into (NH4)4Mo5O17, and the content of residual ammonia, which can be utilized as a reductant, in the ammonium molybdate increases. Although the self-reduction by the residual ammonia was reported briefly in the U.S. Patent 2,385,843 [7], no satisfactory investigation on the self-reduction of (NH4)4Mo5O17 has been reported yet. Therefore, the self-reducing potential of (NH4)4Mo5O17 was extensively examined in this study through the analyses of the effectiveness of this residual ammonia component in terms of decreasing hydrogen consumption and reaction time.

2. Materials and Methods

High-purity APT and MoO3 powders of reagent grade (99.9 wt%) were used in this study. The APT was dried using an ON-11E natural convection oven (Jeio Tech. Co. Ltd., Daejeon, Republic of Korea) at a temperature range of 70 to 100 °C. Then, 10 g of the sample was loaded in an alumina boat and placed in the middle of a horizontal tube furnace (ID 60 × L 1000). Ar and H2 gases were used in the self-reduction stage and in the primary reduction stage, respectively. The Ar gas flow rate was fixed at 0.3 L/min, and H2 gases at 0.1 or 0.2 L/min. The temperature was fixed at 575 °C because it was the lowest temperature in this study to achieve 100% conversion of MoO3 → MoO2, and the residence time range was from 30 to 60 min. The conversion rate was calculated by the weight loss, and the phase change and the morphology were characterized by X-ray diffraction (XRD, D/Max 2500, Rigaku, Tokyo, Japan) and by scanning electron (SEM, MIRA-3, Tescan Co., Brno, Czech Republic) at the Eco-friendly Shipbuilding Core Research Support Center, respectively.

3. Results and Discussion

3.1. Thermal Decomposition of Ammonium Molybdates

3.1.1. Drying of APT

As APT, (NH4)6Mo7O24·4H2O, is thermally decomposed in a stepwise manner, the weight and the composition of ammonium molybdate change by the equivalent amount of gaseous loss, as can be seen in Table 1. In the first step, which can be referred to as the drying stage, (NH4)6Mo7O24·4H2O is transformed to anhydrous ammonium molybdate, (NH4)4Mo5O17, and all of the water molecules, including the crystal ones, are removed together with a small amount of ammonia, as presented by Equation (1):
5 (NH4)6Mo7O24·4H2O → 7 (NH4)4Mo5O17 + 2 NH3 + 21 H2O
From Table 1, it is clear that (NH4)4Mo5O17 is the most favorable state of ammonium molybdate compared to other anhydrous ammonium molybdates in terms of the ratios of NH3/O and H2/O. Based on the data, experimental work was carried out to obtain the optimal conditions for the production of (NH4)4Mo5O17. The identification of (NH4)4Mo5O17 was performed by the analyses of XRD and weight loss.
The results are shown in Figure 2, Figure 3 and Figure 4. When (NH4)6Mo7O24·4H2O was dried at 90 °C, the crystal water could be completely removed after 720 min. When it was dried at 100 °C, the transformation was observed after 60 min, and a considerable amount of MoO3 phase was observed after 120 min. An optimal temperature, therefore, can be found in the range of 90~100 °C. The SEM images for the phase transform are shown in Figure 4. Entangled fiber networks were observed in the structure of (NH4)4Mo5O17 after the transformation.

3.1.2. Self-Reduction of (NH4)4Mo5O17

After the transformation of (NH4)6Mo7O24·4H2O into (NH4)4Mo5O17, the contents of residual ammonia and hydrogen, which can be utilized as reductants, in the ammonium molybdate increased, as shown in Table 1. The reactions related to this utilization can be listed as follows:
(NH4)4Mo5O17 → 5 MoO3 + 4 NH3 (g) + 2 H2O(g)
3 (NH4)4Mo5O17 → 15 MoO2 + 5 N2(g) + 21 H2O(g) + 2 NH3(g)
12 MoO3 + 2 NH3(g) → 3 Mo4O11 + N2(g) + 3 H2O(g)
15 MoO3 + 10 NH3(g) → 15 MoO2 + 5 N2(g) + 15 H2O(g)
This utilization was thermodynamically examined using Outokumpu HSC Chemistry 6 software, and the results are presented in Figure 5, which shows the equilibrium molar amounts of molybdenum oxides and gaseous components when the reaction “5 MoO3 + (0~4) NH3 (g) + 2 H2O(g)” takes place at 575 °C with a stepwise increase in internal NH3 from 0 mol to 4 mol. Based on the graphs shown in Figure 5, it is theoretically expected that MoO3 can be completely converted to MoO2 at 3.4 mol of internal NH3, which means MoO3 can be completely reduced to MoO2 by the internal ammonia without the need for any external supply of ammonia. Experimental work was also carried out to examine the self-reduction, and the results are presented in Figure 6 and Figure 7. The highest degree of reduction achieved in this case was about 18% based on the analysis of weight loss, as follows:
Reduction   Degree   ( % ) = Removed   oxygen   amount Initial   oxygen   amount   in   form   of   MoO 3  
Unlike the theoretical expectation, a perfect reduction degree could not be obtained in this study, and this difference can be ascribed to the technical difficulties in keeping conditions of standard state in the experiment. Practically, the NH3 produced during the thermal decomposition cannot have sufficient time to react with Mo oxide powder before leaving the reactor, because it passes through the powder layer as soon as it is produced. The SEM images for the phase transform are shown in Figure 6. The material exhibited crystallinity, and several shapes of crystal, plates, needles, and irregular bulk shapes were observed after the self-reduction. Molybdenum trioxide (MoO3) usually exhibits three polymorphic structures, orthorhombic (α-MoO3), monoclinic (β-MoO3), and hexagonal phase (h-MoO3), among which the former is the only thermodynamically stable phase [8,9,10]. The changes in particle size and morphology of MoO3 powder with sintering temperature have been reported in the rage of 100 °C to 700 °C [11]. Meanwhile, molybdenum dioxide (MoO2) crystallizes in a monoclinic cell as a distorted rutile crystal structure, and the crystal structure of the metastable phase Mo4O11 can be found either as a low-temperature monoclinic η-phase or high-temperature orthorhombic γ-phase [12,13,14]. MoO3, Mo4O11, and MoO2 phases in Figure 7 were not identified. The only differences between orthorhombic and monoclinic crystals are the size and the distortion, and thus, they were not easy to distinguish. In addition, there have been reports that MoO2 is likely to retain the general morphological habit of the MoO3 from which it was fully reduced in solid state [15,16]. Moreover, the size and the morphology change with the sintering temperature.

3.2. Direct Hydrogen Reduction of Ammonium Molybdates

In a conventional process, (NH4)6Mo7O24·4H2O is first roasted to produce MoO3, which is then reduced by hydrogen gas in two stages: MoO3 → MoO2 and MoO2 → Mo. MoO3 is easily reduced by either hydrogen or other reducing gas, and several studies have reported morphology changes during the reduction [17,18,19,20,21]. In the process proposed in this study, (NH4)6Mo7O24·4H2O is first dried to (NH4)4Mo5O17 and then is directly reduced, without roasting, by hydrogen gas in two stages, so that the ammonia constituent in (NH4)4Mo5O17 can be utilized as a reductant in the first stage to produce MoO2. The experimental results carried out for this purpose are presented in Figure 8, Figure 9 and Figure 10. In the experiments, (NH4)4Mo5O17 was reduced by hydrogen gas of two flow rates, 0.1 L/min and 0.2 L/min at 575 °C, which is a typical temperature for the first stage, and the conversion rates of (NH4)4Mo5O17 were compared with those of MoO3. As can be seen in Figure 8 and Figure 9, the conversion rates of (NH4)4Mo5O17 were found to be higher than those of MoO3, especially in the time range of 20~30 min. By the time the conversion rate (i.e., the reduction degree) of (NH4)4Mo5O17 had reached 100%, that of MoO3 reached 80% at the hydrogen flow rate of 0.2 L/min. This phenomenon also supports the self-reduction of (NH4)4Mo5O17 by the internal ammonia. The morphology changes during the direct reduction are presented in Figure 10. As the direct reduction proceeded, the kernel shape of the ammonium molybdate transformed into platelet shapes of molybdenum oxides such as Mo4O11 and MoO2, and finally all the ammonium molybdates were changed into MoO2 of platelet shapes. This appearance agrees with the model suggested by Dang et al. for hydrogen reduction of MoO3 to MoO2 at 773 to 829 K [22]. The SEM images of the internal structure of the directly H2-reduced samples are shown in Figure 11. SEM images of that for (NH4)4Mo5O17 could not be obtained because of technical difficulties. As can be seen, as the direct reduction proceeded, more pores and cracks were observed within the internal structures. This aspect is similar to what was observed during the hydrogen reduction of MoO3 [3].

4. Conclusions

Ammonium paramolybdate tetrahydrate (APT: (NH4)6Mo7O24·4H2O) was pre-dried at an appropriately low temperature to remove the crystal water corresponding to 4 mol per mol of APT, and then it was transformed into (NH4)4Mo5O17, of which the residual ammonia content increased. The utilization of the ammonia content as a reductant was extensively examined for the first stage of hydrogen reduction to produce MoO2. The results can be summarized as follows:
(1)
As the direct hydrogen reduction of (NH4)4Mo5O17 proceeded, the kernel shape of the ammonium molybdate was transformed into platelet shapes of molybdenum oxides such as Mo4O11 and MoO2, and finally all the ammonium molybdates were changed into MoO2 of platelet shapes. The conversion rates of (NH4)4Mo5O17 were found to be higher than those of MoO3, especially in the time range of 20~30 min. By the time the conversion rate of (NH4)4Mo5O17 had reached 100%, that of MoO3 reached 80% at the hydrogen flow rate of 2 L/min.
(2)
The best reduction degree obtained in this case was about 18% based on the analysis of weight loss. Although a perfect degree could not be obtained in this study due to the technical difficulties to keep conditions of standard state in the experiment, the effectiveness of this residual ammonia content in (NH4)4Mo5O17 was confirmed in terms of decreasing the hydrogen consumption and reaction time.

Author Contributions

Conceptualization, K.Y. and H.K.; methodology, W.B.K. and H.K.; data curation, K.Y., H.K. and S.-h.L.; writing—original draft preparation, K.Y. and H.K.; writing—review and editing, H.K. and S.-h.L.; project administration, K.Y.; funding acquisition, K.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (or Industrial Strategic Technology Development Program) (20011286, Development of 4N5 grade ultrahigh purity of molybdenum with high melting point by smelting and refining technology for semiconductor applications from northern country resources) funded by the Ministry of Trade, Industry, and Energy (MOTIE, Republic of Korea).

Data Availability Statement

Not applicable.

Acknowledgments

We thank Dong Nam Shin for his help in the preparation of TGA/DTA analyses and for his comments during the revision process of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chung, K.W. Proposal on how to Activate Recycling of Critical Minerals. J. Korean Soc. Miner. Energy Resour. Eng. 2022, 59, 489–497. [Google Scholar] [CrossRef]
  2. Nguyen, T.N.H.; Nguyen, T.T.H.; Lee, M.S. Leaching of Molybdenite by Hydrochloric Acid Solution Containing Sodium Chlorate. Resour. Recycl. 2022, 31, 26–33. [Google Scholar] [CrossRef]
  3. Koo, W.B.; Yoo, K.; Kim, H. The Hydrogen Reduction Behavior of MoO3 Powder. Resour. Recycl. 2022, 31, 29–36. [Google Scholar] [CrossRef]
  4. Yin, Z.; Zhao, Q.; Chen, S. Thermal decomposition of ammonium molybdate mixture. Trans. Nonferrous Met. Soc. China 1991, 4, 46–56. [Google Scholar]
  5. Murugan, R.; Chang, H. Thermo-Raman investigation on thermal decomposition of (NH4)6Mo7O24·4H2O. J. Chem. Soc. Dalton Trans. 2001, 20, 3125–3132. [Google Scholar] [CrossRef]
  6. Kim, Y.H.; Kim, H.G. Two-stage Fluidized-bed Systems for the Metal Powder production of molybdenum and nickel. In Proceedings of the 26th International Conference on Metallurgy and Materials, Hotel Voronez I, Brno, Czech Republic, 24–26 May 2017; pp. 1763–1768. [Google Scholar]
  7. Fredrik, R.R. Reduction of Ammonium Molybdate. U.S. Patent 2,385,843, 2 October 1945. [Google Scholar]
  8. Hou, X.; Huang, J.; Liu, M.; Li, X.; Hu, Z.; Feng, Z.; Zhang, M.; Luo, J. Single-Crystal MoO3 Micrometer and Millimeter Belts Prepared from Discarded Molybdenum Disilicide Heating Elements. Sci. Rep. 2018, 8, 16771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Lou, X.W.; Zeng, H.C. Complex α-MoO3 nanostructures with external bonding capacity for self-assembly. J. Am. Chem. Soc. 2003, 125, 2697–2704. [Google Scholar] [CrossRef] [PubMed]
  10. Cai, L.L.; Rao, P.M.; Zheng, X.L. Morphology-controlled flame synthesis of single, branched, and flower-like α-MoO3 nanobelt arrays. Nano Lett. 2011, 11, 872–877. [Google Scholar] [CrossRef] [PubMed]
  11. Chiang, T.H.; Yeh, H.C. The Synthesis of α-MoO3 by Ethylene Glycol. Materials 2013, 6, 4609–4625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. De Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalantar-Zadeh, K. Molybdenum Oxides—From Fundamentals to Functionality. Adv. Mater. 2017, 29, 1701619. [Google Scholar] [CrossRef] [PubMed]
  13. Van Pham, D.; Patil, R.; Yang, C.-C.; Yeh, W.-C.; Liou, Y.; Ma, Y.-R. Impact of the crystal phase and 3d-valence conversion on the capacitive performance of one-dimensional MoO2, MoO3, and Magnéli-phase Mo4O11 nanorod-based pseudocapacitors. Nano Energy 2018, 47, 105–114. [Google Scholar] [CrossRef]
  14. Van Pham, D.; Patil, R.A.; Lin, J.-H.; Lai, C.-C.; Liou, Y.; Ma, Y.-R. Doping-free bandgap tuning in one-dimensional Magnéli-phase nanorods of Mo4O11. Nanoscale 2016, 8, 5559–5566. [Google Scholar] [CrossRef] [PubMed]
  15. Burch, R. Preparation of high surface area reduced molybdenum oxide catalysts. J. Chem. Soc. Faraday Trans. 1978, 74, 2982. [Google Scholar] [CrossRef]
  16. Ellefson, C.A.; Marin-Flores, O.; Ha, S.; Norton, M.G. Synthesis and applications of molybdenum (IV) oxide. J. Mater. Sci. 2012, 47, 2057–2071. [Google Scholar] [CrossRef]
  17. Xia, T.; Li, Q.; Liu, X.; Meng, J.; Cao, X. Morphology-Controllable Synthesis and Characterization of Single-Crystal Molybdenum Trioxide. J. Phys. Chem. B 2006, 110, 2006–2012. [Google Scholar] [CrossRef] [PubMed]
  18. Fang, L.; Shu, Y.; Wang, A.; Zhang, T. Green synthesis and characterization of anisotropic uniform single-crystal α-MoO3 nanostructures. J. Phys. Chem. C 2007, 111, 2401–2408. [Google Scholar] [CrossRef]
  19. Lou, X.W.; Zeng, H.C. Hydrothermal Synthesis of α-MoO3 Nanorods via Acidification of Ammonium Heptamolybdate Tetrahydrate. Chem. Mater. 2002, 14, 4781–4789. [Google Scholar] [CrossRef]
  20. Li, X.L.; Liu, J.F.; Li, Y.D. Low-temperature synthesis of large-scale single-crystal molybdenum trioxide (MoO3) nanobelts. Appl. Phys. Lett. 2002, 81, 4832. [Google Scholar] [CrossRef]
  21. Chen, Y.; Lu, C.; Xu, L.; Ma, Y.; Hou, W.; Zhu, J.J. Single-crystalline orthorhombic molybdenum oxide nanobelts: Synthesis and photocatalytic properties. CrystEngComm 2010, 12, 3740. [Google Scholar] [CrossRef]
  22. Dang, J.; Zhang, G.H.; Chou, K.C. Phase Transitions and Morphology Evolutions during Hydrogen Reduction of MoO3 to MoO2. High Temp. Mater. Process. 2014, 33, 305–312. [Google Scholar] [CrossRef]
Figure 1. TGA and DTA results for the thermal decomposition of ammonium molybdate (a) under N2 atmosphere, (b) under air atmosphere.
Figure 1. TGA and DTA results for the thermal decomposition of ammonium molybdate (a) under N2 atmosphere, (b) under air atmosphere.
Minerals 13 00133 g001
Figure 2. Weight losses of APT with drying time at 90 °C and 100 °C.
Figure 2. Weight losses of APT with drying time at 90 °C and 100 °C.
Minerals 13 00133 g002
Figure 3. XRD analyses of APT with drying time at 90 °C and 100 °C.
Figure 3. XRD analyses of APT with drying time at 90 °C and 100 °C.
Minerals 13 00133 g003
Figure 4. SEM images of ammonium molybdates before and after drying (at 90 °C for 720 min).
Figure 4. SEM images of ammonium molybdates before and after drying (at 90 °C for 720 min).
Minerals 13 00133 g004
Figure 5. Equilibrium molar amounts when the reaction “5 MoO3 + (0~4) NH3 (g) + 2 H2O (g)” takes place at 575 °C with a stepwise increase in NH3 from 0 mol to 4 mol (created using Outokumpu HSC Chemistry 6 software).
Figure 5. Equilibrium molar amounts when the reaction “5 MoO3 + (0~4) NH3 (g) + 2 H2O (g)” takes place at 575 °C with a stepwise increase in NH3 from 0 mol to 4 mol (created using Outokumpu HSC Chemistry 6 software).
Minerals 13 00133 g005
Figure 6. XRD analysis of (NH4)4Mo5O17 after heating at 575 °C for 60 min in Ar atmosphere.
Figure 6. XRD analysis of (NH4)4Mo5O17 after heating at 575 °C for 60 min in Ar atmosphere.
Minerals 13 00133 g006
Figure 7. SEM images before and after the transformation of (NH4)4Mo5O17 to molybdenum oxides by self-reduction at 575 °C for 30 min in Ar atmosphere.
Figure 7. SEM images before and after the transformation of (NH4)4Mo5O17 to molybdenum oxides by self-reduction at 575 °C for 30 min in Ar atmosphere.
Minerals 13 00133 g007
Figure 8. Variations in the rates of conversion to MoO2 with hydrogen reduction time in the first stage of hydrogen reduction at 575 °C for (NH4)4Mo5O17 and MoO3.
Figure 8. Variations in the rates of conversion to MoO2 with hydrogen reduction time in the first stage of hydrogen reduction at 575 °C for (NH4)4Mo5O17 and MoO3.
Minerals 13 00133 g008
Figure 9. XRD patterns and conversion rates obtained after the first stage of hydrogen reduction at 575 °C for (NH4)4Mo5O17 and MoO3.
Figure 9. XRD patterns and conversion rates obtained after the first stage of hydrogen reduction at 575 °C for (NH4)4Mo5O17 and MoO3.
Minerals 13 00133 g009
Figure 10. SEM images of the (NH4)4Mo5O17 before and after the direct hydrogen reduction at 575 °C.
Figure 10. SEM images of the (NH4)4Mo5O17 before and after the direct hydrogen reduction at 575 °C.
Minerals 13 00133 g010
Figure 11. SEM images of the (NH4)4Mo5O17 after the direct hydrogen reduction at 575 °C.
Figure 11. SEM images of the (NH4)4Mo5O17 after the direct hydrogen reduction at 575 °C.
Minerals 13 00133 g011
Table 1. Weight and composition changes in ammonium molybdates in stepwise thermal decomposition.
Table 1. Weight and composition changes in ammonium molybdates in stepwise thermal decomposition.
CompoundH2O/NH3NH3/OO/MoNH3/MoH2/OWeight Loss (%)
(NH4)6Mo7O24·4H2O 0.670.253.430.860.500.00
(NH4)4Mo5O170.000.243.400.800.476.67
(NH4)2Mo4O130.000.153.250.500.3111.09
(NH4)2Mo14O430.000.053.070.140.0916.35
(NH4)2Mo22O670.000.033.050.090.0617.11
MoO3 0.003.000.000.0018.45
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

Yoo, K.; Koo, W.B.; Kim, H.; Lee, S.-h. The Self-Reduction during the Thermal Decomposition of an Ammonium Molybdate. Minerals 2023, 13, 133. https://doi.org/10.3390/min13020133

AMA Style

Yoo K, Koo WB, Kim H, Lee S-h. The Self-Reduction during the Thermal Decomposition of an Ammonium Molybdate. Minerals. 2023; 13(2):133. https://doi.org/10.3390/min13020133

Chicago/Turabian Style

Yoo, Kyoungkeun, Won Beom Koo, Hanggoo Kim, and Sang-hun Lee. 2023. "The Self-Reduction during the Thermal Decomposition of an Ammonium Molybdate" Minerals 13, no. 2: 133. https://doi.org/10.3390/min13020133

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