Next Article in Journal
Evaluation of Fire Blight Resistance of Eleven Apple Rootstocks Grown in Kazakhstani Fields
Previous Article in Journal
ESH: Design and Implementation of an Optimal Hashing Scheme for Persistent Memory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Influence of Preparation Parameters on the Morphology and Magnetic Properties of Fe-N Powders Obtained by the Gas Atomization Method

National Institute of Research and Development for Technical Physics, 47 Prof. Dimitrie Mangeron Blvd., 700050 Iasi, Romania
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2023, 13(20), 11529; https://doi.org/10.3390/app132011529
Submission received: 25 September 2023 / Revised: 11 October 2023 / Accepted: 19 October 2023 / Published: 20 October 2023

Abstract

:
α″-Fe16N2 materials are of increasing interest for their applications in products such as rare earth-free permanent magnets. The lack of a method of mass production for powders as raw materials delays the preparation of such magnets. Through employing the gas atomization method, we managed to prepare α″-Fe16N2 powders whose morphology and magnetic properties were tailored by the preparation parameters. As a result of optimizing the preparation parameters (ejection temperature and pressure, ejection nozzle diameter, and atomization pressure), we managed to prepare powders with a size of about 30 μm and a content of 31% α″-Fe16N2 phase. The value of the saturation magnetization (234.8 emu/g), the reasonable coercivity value (970 Oe) presented by the prepared powders, and the opportunity of scaling up approaches based on the preparation of powders via gas atomization support the feasibility of preparing α″-Fe16N2 powders at an industrial level.

1. Introduction

Permanent magnets play a crucial role in various aspects of modern life for applications in “green energy” (wind turbines, electric–hybrid cars), automation (and robotics), information technology, magnetic sensors, and magnetic microdevices, and their influence is constantly growing [1,2,3,4,5]. At the moment, rare earth-based magnets dominate the global magnet market due to their outstanding magnetic performance. However, some of these rare earths, such as neodymium (Nd), praseodymium (Pr), dysprosium (Dy), and terbium (Tb), are listed as critical materials by the US Department of Energy, the International Renewable Energy Agency, and other international institutions [6,7,8,9], and their future use for permanent magnets production is uncertain. Also, the rare earth extraction and separation process is very pollutive, although new separation and extraction methods are constantly being proposed [10,11,12,13,14]. According to the Motion Control & Motor Association (MCMA), the world’s population is expected to reach 9 billion by 2050, and by this time, every human on Earth will be served by three service robots. Furthermore, each robot is expected to contain more than 100 motors. As a result, the demand for permanent magnets is expected to increase significantly in the coming years. This, coupled with increasing concerns about environmental degradation due to the exploitation of rare earths, increasing costs, and availability problems of rare earths, has led to worldwide efforts to search for rare earth free magnetic materials that can be used for the preparation of a new generation of permanent magnets. The ideal permanent magnet should have the following characteristics: an abundant and environmentally friendly element composition, a high saturation magnetization, a high energy product, fairly high coercivity, and a high Curie temperature. The compound α″-Fe16N2 fulfills almost all of these requirements, as its constituent materials Fe and N are abundant and environmentally friendly, it possesses a giant saturation magnetization of ~2.9 T (~305 emu/g) [15,16], its theoretical energy product is estimated to be two times higher than that of the strongest rare earth magnet [17], it has a Curie temperature of about 540 °C [18], and the reported coercivity of the α″-Fe16N2 compound is about 2 kOe [19]. After the first report by Kim and Takahashi on the giant saturation magnetization of about 290 emu/g of the α″-Fe16N2 phase in the evaporated Fe-N thin film, many researchers have made efforts to reproduce these results. However, it was not until nearly 20 years later that researchers from Hitachi in Japan reported a close value for the saturation magnetization obtained using single-crystal α″-Fe16N2 films grown via molecular beam epitaxy [18,20]. Encouraged by the promising magnetic properties demonstrated by thin films, researchers are interested in preparing bulk samples with the same promising magnetic properties as potential precursors for permanent magnets. Thus, magnetic materials containing the α″-Fe16N2 phase have been prepared in the form of thin films, foils, rods, ribbons, nanocones, nanopowders, and powders [21,22,23,24,25,26,27,28,29,30]. However, the most suitable precursors for the preparation of permanent magnets are materials in the form of powders or even nanopowders. In the case of the α″-Fe16N2 compound, compaction is also a challenge due to the fact that the α″-Fe16N2 compound decomposes at temperatures over 200 °C [31]. Therefore, high-temperature sintering or spark plasma sintering, the usual methods for compacting Nd-Fe-B permanent magnets, are not feasible methods for preparing these types of permanent magnets. To date, the only feasible method is the compaction of powders or nanopowders at lower temperatures in the presence of a binder. However, in the case of nanoparticles, the presence of the binder can reduce quite a lot of the saturation flux density because of the decreased phase purity, consequently reducing the energy product, the (BH)max of the α″-Fe16N2 magnets. In addition, due to their small size, α″-Fe16N2 nanoparticles are highly susceptible to oxidation. These problems can be mitigated by the use of powders, for which the binder–magnetic material ratio would be lower than in the approach based on nanoparticles. For this study, α″-Fe16N2 powders were prepared via two methods: (i) the mechanical ball milling of iron-based materials with a source of solid nitrogen (ammonium nitrate [30,32,33]) and (ii) by keeping the iron powders up to 660–670 °C in ammonia–hydrogen mixtures and rapidly quenching them in water with a long consecutive tempering at 120–150 °C [34] (a method inspired by the pioneering work of Jack [35]). Although preparation via ball milling is a mass preparation method used to prepare powders, the powders obtained via this method showed a high degree of oxidation that contributed to the degradation of the magnetic properties of the resulting particles. In the case of the second approach, (ii), the sample magnetization was about 189 emu/g, lower than the saturation magnetization of iron. Therefore, a mass production method for α″-Fe16N2 powders with good magnetic properties is still lacking, which is why the preparation of permanent magnets based on α″-Fe16N2 is delayed. On the other hand, it is well known one of the most used methods for the mass preparation of powders is the gas atomization method [36,37]. Therefore, in this work, we propose the preparation of Fe-N powders via the gas atomization method and study the influence of some preparation parameters on the morphology and magnetic properties of the obtained Fe-N powders.

2. Materials and Methods

Fe ingots of about 20–25 g were prepared via the arc melting technique from commercial iron lumps (Alfa Aesar) with a purity higher than 99.99%. The resulting Fe ingots were transformed into powders via the gas atomization method using a homemade atomization plant. The schematic structure of the gas atomization plant is shown in Figure 1.
The Fe ingot to be melted was placed in a quartz crucible provided in the lower part with a circular ejection nozzle and heated via induction to temperatures up to 50 °C above the liquidus temperature of Fe (1538 °C) in order to obtain appropriate viscosity and surface tension values that allowed for the steady flow of the liquid jet. A photocell located at the top of the quartz crucible was used to measure the surface temperature of the melted iron after it had been previously calibrated with a high-precision pyrometer. Both the atomizing chamber and the melting chamber were previously vacuumed to a pressure of 103 mbar and flushed 4 times with nitrogen gas and refilled with nitrogen to reduce the oxygen content. Before being ejected, the melt was maintained for five minutes under a nitrogen pressure between 1 and 3 bar. Induction melting creates currents of liquid that facilitate the continuous movement of the material, resulting in the exposure of the liquid iron to the nitrogen gas that fills the melting chamber. The flow of the liquid metal is initiated by the sudden decrease in pressure in the atomizing chamber to 0.2 bar. At the same time, the atomization gas, the pressure of which varied between 10 and 20 bars, hit the molten iron jet, resulting in the disintegration of the melt into droplets. The atomizing gas exit nozzle was in the form of a continuous annular slot with a width of 300 µm and concentric with the melt ejection nozzle (right Figure 1). The diameter of the ejection nozzle varied between 100 and 300 µm. The angle between the direction of the liquid jet and that of the gas jet was kept constant at 45° in all experiments. The distance between the ejection nozzle and the particle collector was tested and established as the optimal value of 120 cm. After atomization, the collected powders were immediately transferred to a nitrogen-filled glovebox without exposure to air to prepare the samples for structural, morphological, and magnetic measurements. The structural analysis was carried out via X-ray diffraction (XRD) in a Bruker AXS D8—Advance diffractometer over an angular range 20–90° with a step of 0.02° and a counting time of 2 s per step using conventional Cu–Kα incident radiation. The Rietveld refinements of the XRD patterns were performed using MAUD software [38]. The microstructure and morphology of the sample was investigated using a scanning electron microscope, FIB/FE—SEM Cross—Beam Carl Zeiss NEON 40 EsB (Oberkochen, Germany), equipped with an energy dispersive X-ray spectroscopy (EDS) module. The free software Image J (https://imagej.net/ij/, accessed on 24 September 2023) was used to characterize the size distribution of the particles [39]. Magnetic measurements were performed using a Vibrating Sample Magnetometer (VSM) (Lake Shore VSM 7410, Westerville, OH, USA) in a maximum applied field of 20 kOe at room temperature.

3. Results

In an attempt to investigate the influence of the parameters of the atomization process, such as the ejection temperature, the ejection pressure, the diameter of the ejection nozzle, and the atomization pressure, on the morphology and magnetic properties of the atomized powders, they were modified one by one.

3.1. Ejection Temperature

Through maintaining the ejection pressure at 2 bar, the atomization pressure at 15 bar, and the diameter of the ejection nozzle at 200 μm, atomized powders were prepared, for which the melt was superheated by 10, 20, and 50 °C, respectively, above the melting temperature of Fe (1538 °C). Figure 2 shows SEM images of the atomization results of the powders for the constant atomization parameters specified above and different ejection temperatures.
The melt ejection temperature is primarily reflected in the melt viscosity. A high melt temperature leads to a low viscosity, while a low temperature leads to a high melt viscosity. Atomization of the high viscosity melt (low temperature, 1548 °C) generally results in a large amount of fibers/ligaments and less spherical particles, as can be seen in Figure 2a. With the increase in temperature to 1568 °C, the viscosity drops; thus, it is possible to disintegrate the melt flow into particles, mostly spherical ones (Figure 2b). The further increase in the ejection temperature to 1588 °C leads to uniformity in the morphology of the formed particles; thus, for the parameters specified above, we obtained mostly spherical powders with a diameter between 20 and 120 μm (Figure 2c). Therefore, we can state that a further increase in melting/ejection temperatures would lead to finer and more uniform particles. However, such an increase in temperature is not possible due to technological limitations related to the fact that the melting temperature of iron is close to the softening temperature of the quartz tube in which it is heated, meaning that further melt heating would lead to the flow of the quartz, prohibiting the atomization process.

3.2. Ejection Pressure (Pej)

Through maintaining the ejection temperature at 1588 °C (the optimal temperature at which the most spherical particles were obtained), the atomization pressure at 15 bar, and the diameter of the ejection nozzle at 200 μm, atomized powders were prepared for different ejection pressures between 1 and 3 bar. Figure 3 shows SEM images of the powders obtained via gas atomization at different ejection pressures whilst the rest of the atomization parameters were constant.
An analysis of the SEM images revealed that the size of the powders increases with the ejection pressure. Thus, for pej = 1 bar, the average particle diameter is 117 μm, and for pej = 2 bar, the average particle diameter is 163 μm, while for pej = 3 bar, the average particle diameter increases to approximately 192 μm. These results can be explained by the link between the mass flow rate of the melt and the median diameter of the particles. Lubanska [40] established the correlation between the average diameter (dm) of the gas-atomized powders and the atomization parameters using the following equation:
d m = D K [ v m v g 1 W e ( 1 + m m m g ) ] 1 / 2
where W e = ( Δ U ) 2 ρ m D σ is the Weber number and represents the ratio of the product of the square of the relative velocity between the melt and the gas ( Δ U ), the density of the metal melt ( ρ m ), the diameter of the ejection nozzle (D), and the surface tension of the melt (σ). K is a constant specific to each atomization system, v m and v g are the kinematic viscosities of the melt and the gas, respectively, and m m / m g is the ratio between the mass flow rate of the melt and the gas flow rate. The increase in ejection pressure is reflected in the increase in the mass flow rate of the melt. According to Equation (1), the increase in mass flow leads to an increase in the diameter of the particles. Therefore, to obtain small-sized particles, the mass flow rate should be kept low.

3.3. The Diameter of the Ejection Nozzle

Figure 4 shows the distributions of the diameters of the atomized particles as a function of the diameter of the ejection nozzle. The rest of the parameters, namely the ejection temperature (1588 °C), ejection pressure (2 bar), and atomization pressure (15 bar), were kept constant.
Generally, gas atomization techniques are characterized by a broad particle size distribution. It can be observed that as the diameter of the ejection nozzle decreases, the particle size distribution curve becomes narrower and moves to the small particle size region. Median particle diameter values were extracted from the particle size distributions and plotted against the ejection nozzle diameter (D), as shown in Figure 4d. A ~50% decrease in median particle diameter is observed when the diameter of the melt nozzle is reduced from D = 300 μm to D = 100 μm, indicating that the diameter values of the ejection nozzle greatly influence the particle size distribution. However, reducing the diameter of the melt nozzle leads to an increase in the resistance to the flow of the melt through the nozzle, and further reduction leads to the possibility of clogging the ejection nozzle.

3.4. Atomizing Pressure

In Figure 5, SEM images of the atomized powders prepared at different atomization pressures while the rest of the parameters, namely the ejection temperature (1588 °C), the ejection pressure (2 bar), and ejection nozzle diameter (200 μm), were kept constant are shown.
It can be observed that the diameter of the particles decreases as the atomization pressure increases. For a constant value of the atomization nozzle section, as it is in our case (300 µm), as the atomization pressure increases, the force exerted by the gas on the melt jet also increases, which leads to its easier disintegration. According to Equation (1), the dependence between particle diameter and gas flow rate is inversely proportional; on the other hand, there is a direct proportional relationship between gas flow rate and pressure. Therefore, increasing the gas flow and the atomization pressure, respectively, leads to a decrease in the diameter of the atomized powder. Thus, for an atomization pressure of 20 bar, the average diameter of the atomized powders is 32 µm. Thus, we can conclude that atomization pressure is the most influential parameter involved in particle size reduction. This observation agrees with the results reported by Silva et al. in [41].

3.5. Structural Characterization of the Atomized Powders

It is well known that nitrogen has limited solubility in iron-based alloys, and the saturation limit depends on the temperature of the melt, as well as the partial pressure of nitrogen above the melt. According to Sievert’s law, the amount of nitrogen dissolved in the melt is directly proportional to the square root of pressure [42]. It is important to remember that the ejection pressure is the pressure at which the melt is held for 5 min in the crucible before the effective atomization. Figure 6 shows the Rietveld refined XRD diffractograms of Fe-N powders of the atomized powders prepared at an ejection temperature of 1588 °C, an atomization pressure of 20 bar, an ejection nozzle diameter of 100 μm, and different ejection pressures (between 1 and 3 bar). The reliability factors of the Rietveld refinement and the percentage concentration of the phases resulting from the XRD analysis are listed in Table 1.
It can be observed that the reliability factors of the Rietveld refinement, such as the goodness of fit (χ2) and the residual factor of the weighted profile (Rwp), have values <2 and <15%, respectively, indicating a good agreement of refinement. The structures of the powders consist of a mixture of α-Fe, γ-F4N, α′-Fe8N, and α″-Fe16N2 crystalline phases. In order to explain the formation of these phases, we will refer to the Fe-N phase diagram [43]. According to the phase diagram, the formation of the austenite γ-phase optimally occurs with the combination of nitrogen potential and temperatures above 592 °C. The austenite γ-phase is a face-centered cubic (fcc) arrangement of Fe atoms, with N randomly occupying up to one in ten of the octahedral interstices [44]. When the γ-phase is quenched (what happens to the melt droplets immediately after formation), a martensitic transformation takes place: the Fe atom arrangement changes from a fcc arrangement to a body-centered cubic (bcc) arrangement [44,45]. Due to the rapid cooling of the γ-phase, part of the nitrogen atoms that do not have enough energy cannot diffuse to the desired position, remaining stuck in the previous positions, so a transformation without diffusion takes place, and the α′-Fe8N phase is formed. The α″-Fe16N2 has a structure that is very similar to that of α′-Fe8N, except for the arrangement of nitrogen atoms: in the α′-Fe8N phase, the N atoms are disordered, while in the α″-Fe16N2, they are ordered. This means that during the fcc–bcc transformation, some of the N atoms have enough time and energy to move from the interstices to the desired positions, managing to arrange themselves and form the α″-Fe16N2 phase. Therefore, both nitrogen atom positions and nitrogen concentration are key points for the phases generated after quenching. The results of our XRD pattern analysis indicate that increasing the ejection pressure leads to an increase in the percentage of Fe-N-based phases (to the detriment of the α-Fe phase). Thus, increasing the ejection pressure from 1 to 3 bar led to an increase in the percentage of the α″-Fe16N2 phase from 14.3 to 26.7%. This can be explained by the fact that a higher nitrogen pressure facilitates the diffusion of a larger amount of nitrogen atoms in the iron lattice. The increased content of the γ-F4N phase for powders obtained at 3 bar is most likely related to the insufficiently fast quenching of these larger-sized powders. The cooling rate of the large-sized atomized droplets is much slower than that of the small-sized droplets. The results from our analysis of the X-ray diffraction patterns (not shown here) corresponding to the Fe-N powders atomized at different ejection temperatures did not indicate significant modifications to the structure.

3.6. Magnetic Properties of the Atomized Fe-N Powders

Figure 7b–d show the hysteresis loops of the atomized powders prepared at an ejection temperature of 1588 °C, an atomization pressure of 20 bar, an ejection nozzle diameter of 100 μm, and different ejection pressures (between 1 and 3 bars). In order to compare the magnetic properties of the Fe-N powders atomized in nitrogen with those of some Fe powders, a reference sample of the Fe powders was atomized in argon gas, both for ejection and for atomization. Figure 7a shows the hysteresis loops corresponding to the reference powders.
The Fe reference powders (atomized in argon) showed a saturation magnetization of Ms = 214.5 emu/g and a coercivity of Hc = 29 Oe. The saturation magnetization is lower than the theoretical value of 218 emu/g, most likely due to the superficial oxidation of the powders, although the XRD investigations did not reveal the existence of oxides. Both the coercivity and saturation magnetization of the Fe-N powders are superior to those of the Fe reference powders, and they increased with increasing ejection pressure, as well as with increasing Fe-N phase content. Since the γ-F4N phase has a lower saturation magnetization than Fe (Ms = 188 emu g [46]) and the α′-Fe8N phase has a saturation magnetization comparable to that of Fe alone (Ms = 220 emu/g [47]), the phase responsible for increasing the saturation magnetization of the powders is the α″-Fe16N2 phase. Thus, the powders obtained at an ejection pressure of 1 bar, with a α″-Fe16N2 phase content of 14.3%, show a 9.2 emu/g increase in saturation magnetization compared to the Fe powders, and the coercivity also significantly increases from 29 Oe to 568 Oe. For the powders prepared at 2 bar, for which the phase content was increased to 22.1% (according to Table 1), the saturation magnetization increased to 232.3 emu/g, and the coercivity increased to 675 Oe. However, for the powders obtained at 3 bar, the saturation magnetization decreased to 229.4 emu/g due to the significant increase in the γ-F4N phase content. To overcome this impediment, after maintaining a pressure of 3 bar above the melt for 5 min before atomization, the pressure was reduced, and ejection was performed at 1 bar. This approach allowed for the preparation of Fe-N powders with a size of about 30 μm, an increased α″-Fe16N2 content of 31% (without the γ-F4N phase), and a superior saturation magnetization (Ms) of about 234.8 emu/g and a coercivity (Hc) of about 970 Oe.

4. Conclusions

For this work, we prepared iron–nitride powders containing α-Fe, γ-F4N, α′-Fe8N, and α″-Fe16N2 crystalline phases via the gas atomization method. We experimentally demonstrated that by tuning the preparation parameters (ejection temperature and pressure, ejection nozzle diameter, and atomization pressure), we managed to increase the content of the α″-Fe16N2 phase in the atomized powders and improve their magnetic properties. It was also found that the preparation parameters significantly influence the morphology and dimensions of the powders. The best magnetic properties (Ms = 234.8 emu/g and Hc = 970 Oe) were obtained in Fe-N powders with a content of 31% α″-Fe16N2 and a size of approximately 30 μm. Although this experimental work verified the feasibility of using the gas atomization method to produce powders containing α″-Fe16N2 phase, overcoming some of the existing limitations at our home facility by increasing the melting temperature (using a crucible with high-temperature softening) or increasing the nitrogen pressure in the ejection chamber (for the time preceding the actual atomization) could lead to an improvement of the magnetic properties of the powders obtained.

Author Contributions

Conceptualization, funding acquisition, resources, investigation, result discussion, writing original draft, M.G.; Sample preparation, experimental design, data analysis, writing original draft, M.L.; Data curation, formal analysis, investigation, equipment assistance, G.S.; experimental design, data analysis, equipment assistance, G.A.; investigation, data analysis, results discussion, M.P.; Supervision, methodology, results discussion N.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Ministry of Research, Innovation and Digitization, CNCS—UEFISCDI, project number PN-III-P4-PCE-2021-0298, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data supporting the reported results are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coey, J.M.D. Perspective and Prospects for Rare Earth Permanent Magnets. Engineering 2020, 6, 119–131. [Google Scholar] [CrossRef]
  2. McCallum, R.W.; Lewis, L.H.; Skomski, R.; Kramer, M.J.; Anderson, I.E. Practical Aspects of Modern and Future Permanent Magnets. Annu. Rev. Mater. Res. 2014, 44, 451–477. [Google Scholar] [CrossRef]
  3. Cui, J.; Kramer, M.; Zhou, L.; Liu, F.; Gabay, A.; Hadjipanayis, G.; Balasubramanian, B.; Sellmyer, D. Current progress and future challenges in rare-earth-free permanent magnets. Acta Mater. 2018, 158, 118–137. [Google Scholar] [CrossRef]
  4. Nakamura, H. The current and future status of rare earth permanent magnets. Scr. Mater. 2018, 154, 273–276. [Google Scholar] [CrossRef]
  5. Mohapatra, J.; Liu, J.P. Chapter 1: Rare-Earth-Free Permanent Magnets: The Past and Future. In Handbook of Magnetic Materials; Brück, E., Ed.; Department of Radiation, Science & Technology, Delft University of Technology: Delft, The Netherlands, 2018; Volume 27, pp. 1–57. [Google Scholar] [CrossRef]
  6. Bauer, D.; Diamond, D.; Li, J.; Sandalow, D.; Telleen, P.; Wanner, B.; Mintzer, I. US Department of Energy Critical Materials Strategy. 2010. Available online: https://www.energy.gov/sites/prod/files/edg/news/documents/criticalmaterialsstrategy.pdf (accessed on 24 September 2023).
  7. Guyonnet, D.; Lefebvre, G.; Menad, N. “Rare Earth Elements and High-Tech Products” Prepared for CEC4EUROPE (Circular Economy Coalition for Europe). 2018. Available online: https://www.cec4europe.eu/wp-content/uploads/2022/01/Chapter_3_3_Guyonnet_et_al_Rare_earth_elements_and_high_tech_products.pdf (accessed on 24 September 2023).
  8. Seo, Y.; Morimoto, S. Comparison of dysprosium security strategies in Japan for 2010–2030. Resour. Policy 2014, 39, 15–20. [Google Scholar] [CrossRef]
  9. Popov, V.V.; Grilli, M.L.; Koptyug, A.; Jaworska, L.; Katz-Demyanetz, A.; Klobčar, D.; Balos, S.; Postolnyi, B.O.; Goel, S. Powder Bed Fusion Additive Manufacturing Using Critical Raw Materials: A Review. Materials 2021, 14, 909. [Google Scholar] [CrossRef]
  10. Zhou, L.; Ge, J. Estimating the environmental cost of mixed rare earth production with willingness to pay: A case study in Baotou, China. Extr. Ind. Soc. 2021, 8, 340–354. [Google Scholar] [CrossRef]
  11. Schreiber, A.; Marx, J.; Zapp, P.; Hake, J.-F.; Voßenkaul, D.; Friedrich, B. Environmental Impacts of Rare Earth Mining and Separation Based on Eudialyte: A New European Way. Resources 2016, 5, 32. [Google Scholar] [CrossRef]
  12. Li, J.; Li, M.; Zhang, D.; Gao, K.; Xu, W.; Wang, H.; Geng, J.; Ma, X.; Huang, L. Clean production technology of selective decomposition of Bayan Obo rare earth concentrate by NaOH. J. Clean. Prod. 2019, 236, 117616. [Google Scholar] [CrossRef]
  13. Yale Environment 360. Available online: https://e360.yale.edu/features/china-wrestles-with-the-toxic-aftermath-of-rare-earth-mining (accessed on 2 July 2019).
  14. Filho, W.L. Chapter 17: An Analysis of the Environmental Impacts of the Exploitation of Rare Earth Metals. In Rare Earths Industry, Technological, Economic, and Environmental Implications; De Lima, I.B., Filho, W.L., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 269–277. [Google Scholar] [CrossRef]
  15. Ji, N.; Liu, X.; Wang, J.-P. Theory of giant saturation magnetization in α″-Fe16N2: Role of partial localization in ferromagnetism of 3d transition metals. N. J. Phys. 2010, 12, 063032. [Google Scholar] [CrossRef]
  16. Wang, J.P.; Ji, N.; Liu, X.; Xu, Y.; Sánchez-Hanke, C.; Wu, Y.; De Groot, F.M.F.; Al-Lard, L.F.; Lara-Curzio, E. Fabrication of Fe16N2 films by sputtering process and experimental investigation of origin of giant saturation magnetization in Fe16N2. IEEE Trans. Magn. 2012, 48, 1710–1717. [Google Scholar] [CrossRef]
  17. Wang, J.-P. Environment-friendly bulk Fe16N2 permanent magnet: Review and prospective. J. Magn. Magn. Mater. 2020, 497, 165962. [Google Scholar] [CrossRef]
  18. Sugita, Y.; Mitsuoka, K.; Komuro, M.; Hoshiya, H.; Kozono, Y.; Hanazono, M. Giant magnetic moment and other magnetic properties of epitaxially grown Fe16N2 single-crystal films. J. Appl. Phys. 1991, 70, 5977–5982. [Google Scholar] [CrossRef]
  19. Jiang, Y.; Mehedi, M.A.; Fu, E.; Wang, Y.; Allard, L.F.; Wang, J.-P. Synthesis of Fe16N2 compound Free-Standing Foils with 20 MGOe Magnetic Energy Product by Nitrogen Ion-Implantation. Sci. Rep. 2016, 6, 25436. [Google Scholar] [CrossRef] [PubMed]
  20. Sugita, Y.; Takahashi, H.; Komuro, M.; Igarashi, M.; Imura, R.; Kambe, T. Magnetic and electrical properties of single phase, single-crystal Fe16N2 films epitaxially grown by molecular beam epitaxy. J. Appl. Phys. 1996, 79, 5576–5581. [Google Scholar] [CrossRef]
  21. Ji, N.; Allard, L.F.; Lara-Curzio, E.; Wang, J.-P. N site ordering effect on partially ordered Fe16N2. Appl. Phys. Lett. 2011, 98, 092506. [Google Scholar] [CrossRef]
  22. Li, X.; Yang, M.Y.; Jamali, M.; Shi, F.Y.; Kang, S.S.; Jiang, Y.F.; Zhang, X.W.; Li, H.S.; Okatov, S.; Faleev, S.; et al. Heavy-metal-free, low-damping, and non-interface perpendicular Fe16N2 thin film and magnetoresistance device. Phys. Status Solidi Rapid Res. Lett. 2019, 13, 1900089. [Google Scholar] [CrossRef]
  23. Hang, X.; Zhang, X.; Ma, B.; Lauter, V.; Wang, J.-P. Epitaxial Fe16N2 thin film on nonmagnetic seed layer. Appl. Phys. Lett. 2018, 112, 192402. [Google Scholar] [CrossRef]
  24. Zhang, X.; Nomura, K.; Wang, J.P. New insight on the mössbauer spectra for Fe16N2 thin films with high saturation magnetization. Jpn. J. Appl. Phys. 2019, 58, 120907. [Google Scholar] [CrossRef]
  25. Liu, J.; Guo, G.; Zhang, X.; Zhang, F.; Ma, B.; Wang, J.-P. Synthesis of α″-Fe16N2 foils with an ultralow temperature coefficient of coercivity for rare-earth-free magnets. Acta Mater. 2020, 184, 143–150. [Google Scholar] [CrossRef]
  26. Liu, J.; Guo, G.; Zhang, F.; Wu, Y.; Ma, B.; Wang, J.P. Synthesis of α″-Fe16N2 ribbons with a porous structure. Nanoscale Adv. 2019, 1, 1337. [Google Scholar] [CrossRef] [PubMed]
  27. Li, Y.; Kuang, Q.; Men, X.; Wang, S.; Li, D.; Choi, C.; Zhang, Z. Anisotropic Growth and Magnetic Properties of α″-Fe16N2@C Nanocones. Nanomaterials 2021, 11, 890. [Google Scholar] [CrossRef]
  28. Dirba, I.; Schwöbel, C.A.; Diop, L.V.B.; Duerrschnabel, M.; Molina-Luna, L.; Hofmann, K.; Komissinskiy, P.; Kleebe, H.-J.; Gutfleisch, O. Synthesis, morphology, thermal stability and magnetic properties of α″-Fe16N2 nanoparticles obtained by hydrogen reduction of γ-Fe2O3 and subsequent nitrogenation. Acta Mater. 2017, 123, 214–222. [Google Scholar] [CrossRef]
  29. Tobise, M.; Saito, S. Synthesis of α″-(Fe,M)16N2 nanoparticles obtained by hydrogen reduction and subsequent nitridation starting from α-(Fe,M)OOH (M=Co,Al). IEEE Trans. Magn. 2022, 58, 1–5. [Google Scholar] [CrossRef]
  30. Li, J.; Yuan, W.; Peng, X.; Yang, Y.; Xu, J.; Wang, X.; Hong, B.; Jin, H.; Jin, D.; Ge, H. Synthesis of fine a α″-Fe16N2 powders by low-temperature nitridation of α-Fe from magnetite nanoparticles. AIP Adv. 2016, 6, 125104. [Google Scholar] [CrossRef]
  31. Widenmeyer, M.; Hansen, T.C.; Niewa, R. Formation and Decomposition of Metastable α′′-Fe16N2 from in situ Powder Neutron Diffraction and Thermal Analysis. Z. Anorg. Allg. Chem. 2013, 639, 2851–2859. [Google Scholar] [CrossRef]
  32. Jiang, Y.F.; Liu, J.M.; Suri, P.K.; Kennedy, G.; Thadhani, N.N.; Flannigan, D.J.; Wang, J.P. Preparation of an α″-Fe16N2 magnet via a ball milling and shock compaction approach. Adv. Eng. Mater. 2016, 18, 1009–1016. [Google Scholar] [CrossRef]
  33. Kim, J.; Hwang, J.; Yi, S. Iron nitride based magnetic powder synthesized by mechanical alloying of Fe-based glassy powders and solid nitrogen compounds. J. Magn. Magn. Mater. 2021, 539, 168329. [Google Scholar] [CrossRef]
  34. Huang, M.Q.; Wallace, W.E.; Simizu, S.; Pedziwiatr, A.T.; Obermyer, R.T.; Sankar, S.G. Synthesis and characterization of Fe16N2 in bulk form. J. Appl. Phys. 1994, 75, 6574–6576. [Google Scholar] [CrossRef]
  35. Jack, K.H. The synthesis, structure, and characterization of α″-Fe16N2 (invited). J. Appl. Phys. 1994, 76, 6620–6625. [Google Scholar] [CrossRef]
  36. Fritsching, U.; Uhlenwinkel, V. Hybrid Gas Atomization for Powder Production. In Powder Metallurgy; Katsuyoshi, K., Ed.; InTech.: Houston, TX, USA, 2012; pp. 99–124. ISBN 978-953-51-0071-3. [Google Scholar]
  37. Anderson, I.E.; White, E.M.H.; Dehoff, R. Feedstock powder processing research needs for additive manufacturingDevelopment. Curr. Opin. Solid State Mater. 2018, 22, 8–15. [Google Scholar] [CrossRef]
  38. MAUD. Available online: https://luttero.github.io/maud/ (accessed on 24 September 2023).
  39. Kumari, R.; Rana, N. Particle size and shape analysis using ImageJ with customized tools for segmentation of particles. Int. J. Eng. Res. Technol. 2015, 4, 247–250. [Google Scholar] [CrossRef]
  40. Lubanska, H. Correction of spray ring data for gas atomization of liquid metals. JOM 1970, 22, 45–49. [Google Scholar] [CrossRef]
  41. Da Silva, F.C.; de Lima, M.L.; Colombo, G.F. Evaluation of a Mathematical Model Based on Lubanska Equation to Predict Particle Size for Close-Coupled Gas Atomization of 316L Stainless Steel. Mater. Res. 2022, 25, e20210364. [Google Scholar] [CrossRef]
  42. Sieverts, A.; Zapf, G. Eisen und Stickstoff. Z. Phys. Chem. 1935, A172, 314–315. [Google Scholar] [CrossRef]
  43. Van Voorthuysen, E.H.D.M.; Boerma, D.O.; Chechenin, N.C. Low-temperature extension of the Lehrer diagram, and the iron-nitrogen phase diagram. Metall. Mater. Trans. 2002, A33, 2593. [Google Scholar] [CrossRef]
  44. Jack, K.H. The occurrence and the crystal structure of α″-iron nitride; a new type of interstitial alloy formed during the tempering of nitrogen-martensite. Proc. R. Soc. Lond. Ser. A 1951, 208, 216–224. [Google Scholar]
  45. Jack, K.H. The synthesis and characterization of bulk α″-Fe16N2. J. Alloys Compd. 1995, 222, 160–166. [Google Scholar] [CrossRef]
  46. Yamaguchi, T.; Sakita, M.; Nakamura, M.; Kobira, T. Synthesis and characteristics of Fe4N powders and thin films. J. Magn. Magn. Mater. 2000, 215–216, 529–531. [Google Scholar] [CrossRef]
  47. Ji, N.; Wu, Y.; Wang, J.-P. Epitaxial high saturation magnetization FeN thin films on Fe(001) seeded GaAs(001) single crystal wafer using facing target sputterings. J. Appl. Phys. 2011, 109, 07B767. [Google Scholar] [CrossRef]
Figure 1. The sketch of the gas atomization plant (left) and illustration of the atomization process (right).
Figure 1. The sketch of the gas atomization plant (left) and illustration of the atomization process (right).
Applsci 13 11529 g001
Figure 2. SEM images of gas-atomized powders ejected at different ejection temperatures. T = 1548 °C (a), T = 1568 °C (b), and T = 1588 °C (c).
Figure 2. SEM images of gas-atomized powders ejected at different ejection temperatures. T = 1548 °C (a), T = 1568 °C (b), and T = 1588 °C (c).
Applsci 13 11529 g002
Figure 3. SEM images of gas-atomized powders ejected at different ejection pressures. pej = 1 bar (a), pej = 2 bar (b), and pej = 3 bar (c).
Figure 3. SEM images of gas-atomized powders ejected at different ejection pressures. pej = 1 bar (a), pej = 2 bar (b), and pej = 3 bar (c).
Applsci 13 11529 g003
Figure 4. Distribution of the diameters of the atomized powders vs. the diameter of the ejection nozzle—D = 300 μm (a), D = 200 μm (b), D = 100 μm (c)—and median particle diameter vs. ejection nozzle diameter (dm) (d).
Figure 4. Distribution of the diameters of the atomized powders vs. the diameter of the ejection nozzle—D = 300 μm (a), D = 200 μm (b), D = 100 μm (c)—and median particle diameter vs. ejection nozzle diameter (dm) (d).
Applsci 13 11529 g004
Figure 5. SEM images of gas-atomized powders at different atomization pressures. pat = 10 bar (a), pat = 15 bar (b), and pat = 20 bar (c).
Figure 5. SEM images of gas-atomized powders at different atomization pressures. pat = 10 bar (a), pat = 15 bar (b), and pat = 20 bar (c).
Applsci 13 11529 g005
Figure 6. The Rietveld refinement results of the Fe-N powders atomized at different ejection pressures—pej = 1 bar (a), pej = 2 bar (b), and pej = 3 bar (c). The experimentally observed profile (black color), calculated profile (red color), the difference between the calculated and experimental data (blue color), and the contribution of specific phases are shown.
Figure 6. The Rietveld refinement results of the Fe-N powders atomized at different ejection pressures—pej = 1 bar (a), pej = 2 bar (b), and pej = 3 bar (c). The experimentally observed profile (black color), calculated profile (red color), the difference between the calculated and experimental data (blue color), and the contribution of specific phases are shown.
Applsci 13 11529 g006
Figure 7. The hysteresis loops of the Fe reference powders (a) and of the atomized Fe-N powders at different ejection pressures: 1 bar (b), 2 bar (c), and 3 bar (d).
Figure 7. The hysteresis loops of the Fe reference powders (a) and of the atomized Fe-N powders at different ejection pressures: 1 bar (b), 2 bar (c), and 3 bar (d).
Applsci 13 11529 g007
Table 1. The reliability factors of the Rietveld refinement and the evolution of the phase composition vs. ejection pressure of the atomized Fe-N powders.
Table 1. The reliability factors of the Rietveld refinement and the evolution of the phase composition vs. ejection pressure of the atomized Fe-N powders.
pej (Bar)χ2Rwp (%)Phase Volume Fraction (%)
α-Feγ-F4Nα′-Fe8Nα″-Fe16N2
11.4512.5374.5-11.214.3
21.6713.24554.318.622.1
31.3211.8140.512.420.326.7
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

Grigoras, M.; Lostun, M.; Stoian, G.; Ababei, G.; Porcescu, M.; Lupu, N. The Influence of Preparation Parameters on the Morphology and Magnetic Properties of Fe-N Powders Obtained by the Gas Atomization Method. Appl. Sci. 2023, 13, 11529. https://doi.org/10.3390/app132011529

AMA Style

Grigoras M, Lostun M, Stoian G, Ababei G, Porcescu M, Lupu N. The Influence of Preparation Parameters on the Morphology and Magnetic Properties of Fe-N Powders Obtained by the Gas Atomization Method. Applied Sciences. 2023; 13(20):11529. https://doi.org/10.3390/app132011529

Chicago/Turabian Style

Grigoras, Marian, Mihaela Lostun, George Stoian, Gabriel Ababei, Marieta Porcescu, and Nicoleta Lupu. 2023. "The Influence of Preparation Parameters on the Morphology and Magnetic Properties of Fe-N Powders Obtained by the Gas Atomization Method" Applied Sciences 13, no. 20: 11529. https://doi.org/10.3390/app132011529

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