Next Article in Journal
Analyzing Discrepancies in Chemical-Shift Predictions of Solid Pyridinium Fumarates
Previous Article in Journal
Pros and Cons of the Cannabinoid System in Cancer: Focus on Hematological Malignancies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanochemical Synthesis and Nitrogenation of the Nd1.1Fe10CoTi Alloy for Permanent Magnet

by
Hugo Martínez Sánchez
1,*,
George Hadjipanayis
2,
Germán Antonio Pérez Alcázar
1,3,
Ligia Edith Zamora Alfonso
1,3 and
Juan Sebastián Trujillo Hernández
1,4
1
Departamento de Física, Universidad del Valle, Meléndez, Cali A.A. 25360, Colombia
2
Department of Physics and Astronomy, University of Delaware, 217 Sharp Lab, Newark, DE 19716, USA
3
Centro de Excelencia en Nuevos Materiales (CENM), Universidad del Valle, Meléndez, Cali A.A. 25360, Colombia
4
Facultad de Ciencias Naturales y Matemáticas, Universidad de Ibagué, Ibagué 730007, Colombia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(13), 3854; https://doi.org/10.3390/molecules26133854
Submission received: 27 May 2021 / Revised: 13 June 2021 / Accepted: 18 June 2021 / Published: 24 June 2021

Abstract

:
In this work, the mechanochemical synthesis method was used for the first time to produce powders of the nanocrystalline Nd1.1Fe10CoTi compound from Nd2O3, Fe2O3, Co and TiO2. High-energy-milled powders were heat treated at 1000 °C for 10 min to obtain the ThMn12-type structure. Volume fraction of the 1:12 phase was found to be as high as 95.7% with 4.3% of a bcc phase also present. The nitrogenation process of the sample was carried out at 350 °C during 3, 6, 9 and 12 h using a static pressure of 80 kPa of N2. The magnetic properties Mr, µ0Hc, and (BH)max were enhanced after nitrogenation, despite finding some residual nitrogen-free 1:12 phase. The magnetic values of a nitrogenated sample after 3 h were Mr = 75 Am2 kg–1, µ0Hc = 0.500 T and (BH)max = 58 kJ·m–3. Samples were aligned under an applied field of 2 T after washing and were measured in a direction parallel to the applied field. The best value of (BH)max ~ 114 kJ·m–3 was obtained for 3 h and the highest µ0Hc = 0.518 T for 6 h nitrogenation. SEM characterization revealed that the particles have a mean particle size around 360 nm and a rounded shape.

1. Introduction

The high demand for Nd2Fe14B permanent magnets, with maximum energy product (BH)max up to 450 kJ·m–3 [1], in applications such as electric vehicles and wind turbines, implies a high production cost owing to the considerable content of Nd. The ferromagnetic R(Fe,M)12 compounds, where R is a rare earth element and M is a stabilizing element (V, Ti, Mo, Cr, W, Al or Si), are currently being investigated as alternatives to Nd-Fe-B due to their excellent magnetic properties and their much lower R content [2,3]. It has been found that the ferromagnetic R(Fe,M)12 compounds crystalize in the tetragonal ThMn12-type structure (space group I4/mmm) with R= Y, Nd, Sm, Gd, Tb, Dy, Ho, Er, Tm or Lu [4], and that these compounds exhibit strong exchange interactions for the light rare earths. The Sm(Fe,M)12 compounds are characterized by a strong uniaxial anisotropy [3] and develop a coercive force µ0Hc as high as 1.17 T [5]. The second type of attractive compounds are the Nd(Fe,M)12Nx and Pr(Fe,M)12Nx with M = Ti, V or Mo [6]. The highest coercivity values for these compounds with a subsequent nitrogenation, µ0Hc up to 1.3 T, have been achieved in nanocrystalline quasi-isotropic alloys with M = Mo [7,8,9,10]; somewhat lower µ0Hc values, up to 0.9 T, were reported for nanocrystalline alloys with M = V [11]. Although it is most favorable for a high coercivity, the microstructure of randomly oriented nanograins—usually generated through melt-spinning or high-energy ball-milling—dramatically lowers the attainable (BH)max due to isotropic nature of the samples. The same alloys (M = Mo, V) also allowed for the synthesis of anisotropic R(Fe,M)12Nx powders with µ0Hc of 0.35–0.58 T [12,13,14,15,16]. No significant coercivity, however, could be developed in the NdFe11TiNx compound, which is particularly attractive due to the high Ms of 1.48 T [17]; even the µ0Hc reported for the nanocrystalline alloys did not exceed 0.23 T [11]. The mechanochemical synthesis was demonstrated to yield single-crystal submicron and nanoparticles of hard magnetic materials, including the R(Fe,M)12 alloys, which exhibit coercivity values typical of nanocrystalline powders [18,19]. This synthesis technique also requires less-expensive oxides, rather than metals, as the raw materials, and it may be expected to reduce local demagnetization fields due to more regular shapes of the particles [20]. The authors are not aware of any reports on mechanochemically synthesized R(Fe,M)12Nx compounds. By reducing nanoscale oxide precursors Su et al. [21] prepared anisotropic NdFe10Mo2Nx particles, which were a few microns in size and exhibited an µ0Hc of 0.35 T. The aim of this work was to prepare a Nd(Fe,Co,Ti)12Nx permanent magnet material through mechanochemical synthesis and subsequent nitrogenation of the Nd(Fe,Co,Ti)12 compound and to study its extrinsic magnetic properties as a function of the temperature and magnetic alignment.

2. Materials and Methods

2.1. Sample Preparation

The alloy with nominal composition Nd1.1Fe10CoTi was obtained by mechanochemical synthesis, using powders of Nd2O3 (nominal purity of 99.9%), Fe2O3 (particles ≤5 µm and 99% purity), TiO2 (≤5 µm and 99% purity) and Co (1.6 µm and 99.8% purity) as the raw materials. The reactants were mixed with 1 mm granules of Ca (99.5% purity) serving as a reducing agent and CaO powder acting as a dispersant. The amounts of the precursor powders were optimized using a mole ratio (Fe+Co+Ti)/Nd of 6.62 (necessitated by incomplete reduction of the rare earth oxide), and a Ca/O ratio of 2.0 (not counting the CaO dispersant). The mass of the CaO dispersant was three times the mass of the reactants other than Ca. The reactants were first mixed thoroughly with Ca and then with the dispersant. After that, a 5 g charge of the mixed powders was mechanically activated with a Spex-8000 high-energy (University of Delaware, Department of Physics and Astronomy, Sharp Lab, Newark DE, USA) mill by milling for 4 h in argon atmosphere, with six 12 mm steel balls. The subsequent annealing at temperatures from 950 to 1100 °C during 10 min was done in argon-filled quartz capsules, without exposing the as-milled powder to the air. A temperature of 1000 °C was found to be optimal for the formation of the ThMn12-type structure. For nitrogenation, the annealed powders were sealed in quartz capsules under 80 kPa of a 99.999%-pure N2 and held at 350 °C for 3, 6, 9 and 12 h. Finally, a multistep procedure of repeated washing with deionized water, glycerol, acetic acid, and ethanol [19] was used to collect the ferromagnetic nitrogenated particles, and to remove the CaO, rare earth oxides and unreacted Ca.

2.2. Powder Characterization

The powders were characterized by X-ray diffraction (XRD) in a Rigaku Ultima IV diffractometer (University of Delaware, Department of Physics and Astronomy, Sharp Lab, Newark, DE, USA) with Cu-Kα radiation, and the diffraction patterns were measured in the 2θ range from 25 to 60 degrees. The analysis of XRD patterns was done using the Maud program. Scanning electron microscopy (SEM) was used to study the particle shape and size distribution, using a JEOL JSM-6335F scanning electron microscope (University of Delaware, Department of Physics and Astronomy, Sharp Lab, Newark, DE, USA). The magnetic properties were measured using a Quantum Design VersaLab vibrating sample magnetometer (University of Delaware, Department of Physics and Astronomy, Sharp Lab, Newark DE, USA); the particles were immobilized with paraffin wax, in some cases while applying a 2 T orienting field (hereinafter, “aligned” powders). Correction for self-demagnetization was done using demagnetization factors experimentally determined for similarly prepared Ni powder.

3. Result and Discussion

3.1. Structural and Magnetic Characterization before Washing

After the mechanochemical activation and annealing at 1000 °C for 10 min, the powders were characterized by X-ray diffraction and vibrating sample magnetometry. The X-ray diffraction pattern of Figure 1 shows the peaks of the crystalline phases present before washing the powder. The highest peaks correspond to the CaO phase with cubic crystal structure (space group Fm-3m); the phase with ThMn12-type crystal structure (space group I4/mmm) or “1:12”, is present as a minority phase. The cubic bcc phase (space group Im-3m); and some peaks of the hexagonal Nd2O3 phase (space group P–63/mmc), as well as of the orthorhombic CaTiO3 phase (space group Pnma) were also identified. The peak at 2θ = 27.59 deg possibly corresponds to a rare earth oxide isotypical to Pr24O44 (space group P–1).
The hysteresis loop of the annealed and not washed powder is shown in Figure 2. Magnetization values at 3 T of 15.67 Am2kg−1, remanence Mr = 4.43 Am2kg−1, and µ0Hc = 0.121 T were obtained by analysis of this measurement. The magnetization values are low due to the predominance of the diamagnetic CaO diluting the ferromagnetic phase-(s). The low Mr and µ0Hc values are consistent with the weak anisotropy of nitrogen-free Nd(Fe,M)12.

3.2. SEM Characterization

Figure 3 shows the SEM images of the washed particles nitrogenated for different times.
The ferromagnetic particles tend to have rounded shape and a log-normal size distribution with a positive asymmetry (Fisher asymmetric coefficient γ1 > 0). The log-normal fit showed mean particle sizes of 0.29, 0.37, 0.35, 0.39, and 0.41 µm for the 0, 3, 6, 9, and 12 h nitrogenated samples, respectively. These results show that the particles tend to grow with the increase of nitrogenation time. It is worth noting that the small particle size of about 360 nm and the rounded shape of the particles are important factors for the decrease of the demagnetized field [20] and the improvement of the coercive force.

3.3. XRD after Washing

In Figure 4, the XRD diffraction patterns of the washed powders are shown. For the sample without nitrogenation (0 h), the pattern of the peaks of the 1:12 phase, and a single peak of the bcc phase at 2θ = 44.62 deg are present. The refinement of the pattern gave a volume fractions of 95.7 and 4.3% for the 1:12 and the bcc phases, respectively. The refined lattice parameters of the 1:12 phase were a = 8.589 Å and c = 4.808 Å. The average crystallite sizes found from the XRD refinement were 269.7(15.2) nm and 46.4(6.2) nm for the 1:12 and bcc phases, respectively. The mean crystallite size of the 1:12 phase is very close to the value of 0.29 µm obtained for the mean particle size (SEM), indicating that the particles of this phase are single crystals. These single crystal particles lead an improved μ0Hc because of their submicron size.
The XRD diffractions patterns for the samples nitrogenated for 3, 6, 9 and 12 h do not feature the characteristic shift of the peaks to the left. Instead, the peaks are broadened to their left side as it can be seen in Figure 4. This, apparently, indicates that in all the samples a certain part of the 1:12 phase remains nitrogen-free, even though most of the phase does absorb nitrogen. Thus, the resulting powders may be characterized by a gradient in the N content. In order to verify this hypothesis, Le Bail [22] analysis was carried out, assuming one bcc phase and three distinct 1:12 phases 1:12(a), 1:12(b) and 1:12(c). This analysis allowed for a good refinement of the XRD patterns with the parameters which are presented in Table 1.
The 1:12(a) phase is inherited from the parent, nitrogen-free alloy, because its lattice parameters and unit cell volume are similar, see Table 1. Its XRD peaks are in the right side of the XRD spectrum, as they do not shift to the left upon the nitrogenation. The 1:12(b) and 1:12(c) phases have bigger lattice parameters and therefore bigger volumes of nitrides, indicating that they absorbed different amounts of nitrogen. At the same time, the intensity of the bcc phase peak increases with the nitrogenation time (see Figure 4), indicating that the volume fraction of this phase increases from 4.3 to 8.5% when the nitrogenation time increases from 0 to 12 h.

3.4. Magnetic Characterization of Aligned and Non-Aligned Powders

Figure 5 presents the hysteresis loops of the randomly oriented Nd1.1Fe10CoTiNx powders for 0, 3, 6, 9, and 12 h of nitrogenation. The M3T, Mr and µ0Hc magnetic properties for the 0 h sample increase after washing from 15.67 to 121.92 Am2kg−1, 4.43 to 57.79 Am2kg−1, and 0.121 to 0.276 T, respectively, due to the removal of non-magnetic CaO and the rare earth oxides. Moreover, the Mr and the µ0Hc values increase after nitrogenation, as it can be seen in the Figure 5 and are listed in Table 2 (in parenthesis). This enhancement of the magnetic properties is due to the interstitial modification of those 1:12 crystallites which absorb nitrogen, leading to enhancement of their magnetocrystalline anisotropy.
The values of µ0Hc decrease when the nitrogenation is increased from 3 to 12 h, apparently, because of a nearly doubled amount of the soft magnetic bcc phase. The Mr = 75.27 Am2kg−1, µ0Hc = 0.500 T, and (BH)max = 58.38 kJ·m−3 values for 3 h of nitrogenation are comparable to those reported in [9,23], for isotropic nanocrystalline Nd(Fe,Mo)12Nx alloys, even though 1:12 alloys stabilized with Mo, are known to exhibit a higher coercivity than the 1:12 alloys stabilized with Ti.
In Figure 6, the hysteresis loops for the anisotropic ferromagnetic particles aligned with an external field are shown; the corresponding M3T, Mr, µ0Hc, and (BH)max dates are listed in Table 2 (without parenthesis). All the hard magnetic properties were found to be enhanced by the alignment. It is obvious that the increase of M3T and Mr is the immediate result of the easy axes of the crystallites being aligned with the direction of the measurements. The improvement of µ0Hc is caused by the increase of the magnetocrystalline anisotropy due to the nitrogenization. The reduced remanence Mr/M3T increases upon nitrogenation and the alignment; for 3, 6 and 9 h of nitrogenation, a value of 0.90 was obtained despite the soft magnetic bcc phase. The possibility to align the easy axes of magnetization in the Nd(Fe,M)12Nx particles is important for the development of high (BH)max values, as it is shown in Table 2. A (BH)max = 113.74 kJ·m−3 that was obtained for the aligned particles after nitrogenation for 3 h is the highest (BH)max value reported so far for the Nd(Fe,M)12Nx compounds, and it contributes to closing the gap between the hexaferrites and the “RE-lean” magnets [1]. Moreover, this (BH)max value is better than those reported for NdFe9.4Co1.6MoNx [9], Mn-Al-C [24], and MnBi [25], that are 56.50, 73.21, and 70.82 kJ·m−3, respectively.

3.5. Magnetic Characterization at Low Temperatures

To study the evolution of M3T, Mr, and µ0Hc with the temperature, the hysteresis loops were measured in the first and the second quadrant, for the sample nitrogenated at 6 h, as can see in Figure 7. The 6 h nitrogenation was selected because it exhibited one of the best magnetic properties after nitrogenation without alignment. The strong uniaxial anisotropy of the R(Fe,Ti)12 compounds has been associated with the strong coupling of both rare earth and transition metal magnetic sublattices, which favor an axial orientation [26]. However, experimental and theoretical studies have shown that for the NdFe11Ti(N or H) compounds, a rotation of the easy axis anisotropy to a canted or basal orientation with the decrease of the temperature is obtained, and this behavior is known as spin reorientation [4,27,28].
Temperature dependence of magnetization of the NdFe11Ti compound reported a peak at ~200 K related to the spin reorientation of this compound [4]. In aligned powders, a maximum of magnetization was observed at temperatures higher than the spin reorientation temperature [29]. The decrease of M(H) as a function of temperature shown in Figure 7 can be associated with the spin reorientation. The jumps of magnetization emerging in the 1st M(H) quadrant with the decrease of temperature can also be attributed to a metamagnetic spin reorientation transition [30]. The M3T, Mr, and µ0Hc values are shown in Figure 8 as a function of the temperature.
For the µ0Hc, a tendency to increase with the decrease in the temperature is shown in Figure 8; A maximum value of µ0Hc = 1.259 T was obtained at 100 K. This fact indicates that the spin-reorientation transition occurs around this temperature, in agreement with the reported spin-reorientation temperature for the NdFe11TiH compound (100 K [27]). On the other hand, this change in the µ0Hc with the decrease of the temperature was also observed in the (Nd0.2Ce0.8)2Fe14B compound at 100 K, where it was similarly associated with the spin-reorientation temperature [31]. For the lower temperature of 50 K, the µ0Hc decreases to 1.171 T as can be seen in Figure 8.

4. Conclusions

The mechanochemical synthesis has been demonstrated to be a powerful technique to synthetize submicron single crystal Nd(Fe,Co,Ti)12Nx particles with rounded shape, and reasonably good magnetic properties. The nitrogenated Nd1.1Fe10CoTiNx compounds exhibited a significant improvement of their magnetic properties after being washed, and aligned, even though according to the Le Bail analysis of the XRD patterns about 40 vol.% of the original 1:12 phase, did not adsorb nitrogen in this experiment. The increase in the magnetic properties Mr, µ0Hc and (BH)max upon nitrogenation results in values of 75 Am2kg−1, 0.500 T and 58 kJ·m−3, respectively, for the sample with 3 h of nitrogenation. Comparison of the XRD analysis of the crystallite size and SEM characterization of the particles revealed that the washed 1:12 particles were monocrystalline. The alignment of the samples with an external magnetic field appears to be a very important factor to significantly enhance the magnetic properties. The best (BH)max ~ 114 kJ·m−3 was measured for the sample with 3 h of nitrogenation, thus significantly improving the values previously reported for other systems. The best µ0Hc = 0.518 T was obtained in a sample with 6 h of nitrogenation. Low temperature M(H) curves show magnetization jumps, which indicate a first order magnetization process.

Author Contributions

Conceptualization, H.M.S., G.H., G.A.P.A.; methodology, H.M.S., G.H., L.E.Z.A., J.S.T.H.; software, H.M.S., J.S.T.H.; validation, H.M.S., G.H., G.A.P.A., L.E.Z.A., J.S.T.H.; formal analysis, H.M.S.; investigation, H.M.S.; writing—original draft preparation, H.M.S.; writing—review and editing, H.M.S., G.H., G.A.P.A., L.E.Z.A., J.S.T.H.; supervision, G.H., G.A.P.A., L.E.Z.A., J.S.T.H.; project administration, G.A.P.A., G.H.; funding acquisition, J.S.T.H., G.A.P.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was sponsored in part by the ARMY RESEARCH LABORATORY and was accomplished under Cooperative Agreement Number W911NF-19-2-0030, in part by COLCIENCIAS under Contract 110671250407, and in part by UNIVERSIDAD DEL VALLE under project CI 71181.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the results of this research are available when the editorial office of this journal requires it.

Acknowledgments

The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of the Army Research Laboratory or the U.S. Government. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation herein. The authors would like to thank Anit Giri, from the U.S. Army Research Laboratory for a critical discussion of the results. The authors give the thanks to Aleksandr Gabay, from Department of Physics and Astronomy, University of Delaware, for your support about mechanochemical synthesis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Madugundo, R.; Rama Rao, N.V.; Schönhöbel, A.M.; Salazar, D.; El-Gendy, A.A. Recent Developments in Nanostructured Permanent Magnet Materials and Their Processing Methods. In Magnetic Nanostructured Materials: From Lab to Fab; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  2. Tukker, A. Rare earth elements supply restrictions: Market failures, not scarcity, hamper their current use in high-tech applications. Environ. Sci. Technol. 2014, 48, 9973–9974. [Google Scholar] [CrossRef] [PubMed]
  3. Schönhöbel, A.M.; Madugundo, R.; Vekilova, O.Y.; Eriksson, O.; Herper, H.C.; Barandiarán, J.M.; Hadjipanayis, J.C. Intrinsic magnetic properties of SmFe12-xVx alloys with reduced V-concentration. J. Alloys Compd. 2019, 786, 969–974. [Google Scholar] [CrossRef] [Green Version]
  4. Hu, B.P.; Li, H.S.; Gavigan, J.P.; Coey, J.M.D. Intrinsic magnetic properties of the iron-rich ThMn12-structure alloys R(Fe11Ti); R=Y, Nd, Sm, Gd, Tb, Dy, Ho, Er, Tm and Lu. J. Phys. Condens. Matter 1989, 1, 755–770. [Google Scholar] [CrossRef]
  5. Schultz, L.; Schnitzke, K.; Wecker, J. High coercivity in mechanically alloyed Sm-Fe-V magnets with a ThMn12 crystal structure. Appl. Phys. Lett. 1990, 56, 1376–1378. [Google Scholar] [CrossRef]
  6. Yang, J.; Yang, Y. Magnetic properties and interstitial atom effects in the R(Fe,M)12 compounds. In Handbook of Advanced Magnetic Materials; Springer: Boston, MA, USA, 2006; pp. 1414–1451. [Google Scholar]
  7. Yang, J.; Oleinek, P.; Müller, K.H. Hard magnetic properties of melt-spun R(Fe,M)12 nitrides (R= Nd or Pr; M= Mo or V). J. Appl. Phys. 2000, 88, 988–992. [Google Scholar] [CrossRef]
  8. Zhang, X.D.; Cheng, B.P.; Yang, Y.C. High coercivity in mechanically milled ThMn12-type Nd–Fe–Mo nitrides. Appl. Phys. Lett. 2000, 77, 4022–4024. [Google Scholar] [CrossRef]
  9. Liu, S.; Han, J.; Du, H.; Wang, C.; Yang, Y.; Yang, J.; Chang, H. High energy product in mechanically alloyed ThMn12-type compound with exchange coupling effect. J. Magn. Magn. Mater. 2007, 312, 449–452. [Google Scholar] [CrossRef]
  10. Lin, Z.; Han, J.; Liu, S.; Xing, M.; Yang, Y.; Yang, J.; Wang, C.; Dua, H.; Yang, Y. Phase composition, microstructures and magnetic properties of melt-spun Nd1.2Fe10.5Mo1.5 ribbons and their nitrides. J. Magn. Magn. Mater. 2012, 324, 196–199. [Google Scholar] [CrossRef]
  11. Tang, S.L.; Wu, C.H.; Jin, X.M.; Wang, B.W.; Li, G.S.; Ding, B.Z.; Chuang, Y.C. Phase formation and magnetic properties of annealed mechanical alloying Nd-Fe-V-Ti alloys and their nitrides; J. Alloys Compd. 1998, 264, 240–243. [Google Scholar] [CrossRef]
  12. Yang, J.; Mao, W.; Yang, Y.; Ge, S.; Zhao, Z.; Li, F. Nitrogenation of the magnetic compound R(Fe,M)12. J. Appl. Phys. 1998, 83, 1983–1987. [Google Scholar] [CrossRef]
  13. Mao, W.; Zhang, X.; Ji, C.; Chang, H.; Cheng, B.; Yang, Y.; Du, H.; Xue, Y.; Zhang, B.; Wang, L.; et al. Structural and magnetic properties of PrFe12-xVx and their nitrides. Acta Mater 2001, 49, 721–728. [Google Scholar] [CrossRef]
  14. Han, J.; Liu, S.; Xing, M.; Lin, Z.; Kong, X.; Yang, J.; Wang, C.; Du, H.; Yang, Y. Preparation of anisotropic Nd(Fe,Mo)12N1.0 magnetic materials by strip casting technique and direct nitrogenation for the strips. J. Appl. Phys. 2011, 109, 07A738. [Google Scholar] [CrossRef]
  15. Yang, Y.; Yang, J.; Han, J.; Wang, C.; Liu, S.; Du, H. Research and development of interstitial compounds. IEEE Trans. Magn. 2015, 51, 2103806. [Google Scholar] [CrossRef]
  16. Fu, J.B.; Yu, X.; Qi, Z.Q.; Yang, W.Y.; Liu, S.Q.; Wang, C.S.; Du, H.L.; Han, J.Z.; Yang, Y.C.; Yang, J.B. Magnetic properties of Nd(Fe1-xCox)10.5M1.5 (M = Mo and V) and their nitrides. AIP Adv. 2017, 7, 056202. [Google Scholar] [CrossRef] [Green Version]
  17. Akayama, M.; Fujii, H.; Yamamoto, K.; Tatami, K. Physical properties of nitrogenated RFe11Ti intermetallic compounds (R = Ce, Pr and Nd) with ThMn12-type structure. J. Magn. Magn. Mater. 1994, 130, 99–107. [Google Scholar] [CrossRef]
  18. Gabay, A.M.; Hadjipanayis, G.C. Mechanochemical synthesis of magnetically hard anisotropic RFe10Si2 powders with R representing combination of Sm, Ce and Zr. J. Magn. Magn. Mater. 2017, 422, 43–48. [Google Scholar] [CrossRef] [Green Version]
  19. Gabay, A.M.; Martín-Cid, A.; Barandiaran, J.M.; Salazar, D.; Hadjipanayis, G.C. Low-cost Ce1-xSmx(Fe,Co,Ti)12 alloys for permanent magnets. AIP Adv. 2016, 6, 056015. [Google Scholar] [CrossRef] [Green Version]
  20. Dirba, I.; Li, J.; Sepehri-Amin, H.; Ohkubo, T.; Schrefl, T.; Hono, K. Anisotropic, Single-crystalline SmFe12-based microparticles with high roundness fabricated by jet-milling. J. Alloys Compd. 2019, 804, 155–162. [Google Scholar] [CrossRef]
  21. Su, M.Z.; Liu, S.F.; Qian, X.L.; Lin, J.H. An alternative approach to the finely crystalline powder of rare earth-transition metal alloys: J. Alloys Compd. 1997, 249, 229–233. [Google Scholar] [CrossRef]
  22. Le Bail, A.; Duroy, H.; Fourquet, J.L. Ab-initio structure determination of LiSbWO6 by X-ray powder diffraction. Mater. Res. Bull. 1988, 23, 447–452. [Google Scholar] [CrossRef]
  23. Aubert, A.; Madugundo, R.; Schönhöbel, A.M.; Salazar, D.; Garitaonandia, J.S.; Barandiaran, J.M.; Hadjipanayis, G. Structural and magnetic properties of Nd-Fe-Mo-(N) melt-spun ribbons with ThMn12 structure. Acta Mater. 2020, 195, 519–526. [Google Scholar] [CrossRef]
  24. Kubo, T.; Ohtani, T.; Kojima, S.; Kato, N.; Kojima, K.; Sakamoto, Y.; Tsukahara, M. Machinable Anisotropic Permanent Magnets of Mn-Al-C Alloys. U.S. Patent 3, 976, 519, 24 August 1976. [Google Scholar]
  25. Yang, J.; Yang, W.; Shao, Z.; Liang, D.; Zhao, H.; Xia, Y.; Yang, Y. Mn-based permanent magnets. Chin. Phys. B. 2018, 27, 117503. [Google Scholar] [CrossRef]
  26. Boltich, E.B.; Ma, B.M.; Zhang, L.Y.; Pourarian, F.; Malik, S.K.; Sankar, S.G.; Wallace, W.E. Spin reorientations in RTiFe11 systems (R = Tb, Dy and Ho). J. Magn. Magn. Mater. 1989, 78, 364–370. [Google Scholar] [CrossRef]
  27. Piquer, C.; Grandjean, F. A magnetic and Mössbauer spectral study of the spin reorientation in NdFe11Ti and NdFe11TiH. J. Appl. Phys. 2004, 95, 6308–6316. [Google Scholar] [CrossRef]
  28. Guslienko, K.Y.; Kou, X.C.; Grössinger, R. Magnetic anisotropy and spin-reorientation transitions in RFe11Ti (R= Nd, Tb, Dy, Er) rare-earth intermetallics, J. Magn. Magn. Mater. 1995, 150, 383–392. [Google Scholar] [CrossRef]
  29. Luong, N.H.; Thuy, N.P.; Tal, L.T.; Hoang, N.V. Effect of Y substitution of the spin reorientation in NdFe11Ti compound. Phys. Status Solidi 1990, 121, 607–610. [Google Scholar] [CrossRef]
  30. Pateti, L.; Paoluzi, A.; Albertini, F. Magnetic anisotropy and magnetization processes in 3:29 and 1:12 Nd(FeTi)-based compounds. J. Appl. Phys. 1994, 76, 7473–7477. [Google Scholar] [CrossRef]
  31. Kolchugina, N.B.; Zheleznyi, M.V.; Savchenko, A.G.; Menushenkov, V.P.; Burkhanov, G.S.; Koshkid’ko, Y.S.; Ćwik, J.; Dormidontov, N.A.; Skotnicova, K.; Kursa, M.; et al. Simulating the hysteretic characteristics of hard magnetc materials based on Nd2Fe14B and Ce2Fe14B intermetallics. Crystals 2020, 10, 518. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction pattern of the Nd1.1Fe10CoTi powder annealed at 1000 °C for 10 min.
Figure 1. X-ray diffraction pattern of the Nd1.1Fe10CoTi powder annealed at 1000 °C for 10 min.
Molecules 26 03854 g001
Figure 2. Hysteresis loop of the Nd1.1Fe10CoTi powder annealed at 1000 °C for 10 min.
Figure 2. Hysteresis loop of the Nd1.1Fe10CoTi powder annealed at 1000 °C for 10 min.
Molecules 26 03854 g002
Figure 3. SEM micrographs of mechanochemical synthesized of Nd1.1Fe10CoTi particles nitrogenized for (a) 0 h, (b) 3 h, (c) 6 h, (d) 9 h and (e) 12 h.
Figure 3. SEM micrographs of mechanochemical synthesized of Nd1.1Fe10CoTi particles nitrogenized for (a) 0 h, (b) 3 h, (c) 6 h, (d) 9 h and (e) 12 h.
Molecules 26 03854 g003
Figure 4. X-ray diffraction patterns of Nd1.1Fe10CoTi alloys nitrogenated from 0 to 12 h followed by washing.
Figure 4. X-ray diffraction patterns of Nd1.1Fe10CoTi alloys nitrogenated from 0 to 12 h followed by washing.
Molecules 26 03854 g004
Figure 5. Hysteresis loops of the randomly oriented Nd1.1Fe10CoTiNx alloys after washing as a function of nitrogenation time.
Figure 5. Hysteresis loops of the randomly oriented Nd1.1Fe10CoTiNx alloys after washing as a function of nitrogenation time.
Molecules 26 03854 g005
Figure 6. Hysteresis loops of the Nd1.1Fe10CoTiNx alloys nitrogenized for the indicated time, washed, and magnetically oriented.
Figure 6. Hysteresis loops of the Nd1.1Fe10CoTiNx alloys nitrogenized for the indicated time, washed, and magnetically oriented.
Molecules 26 03854 g006
Figure 7. Demagnetization curves of Nd1.1Fe10CoTiNx powder nitrogenated during 6 h and washed. The specimen was not magnetically oriented.
Figure 7. Demagnetization curves of Nd1.1Fe10CoTiNx powder nitrogenated during 6 h and washed. The specimen was not magnetically oriented.
Molecules 26 03854 g007
Figure 8. Saturation magnetization, remanence and coercivity Nd1.1Fe10CoTiNx alloy nitrogenated during 6 h and as a function of temperature. The specimen was not magnetically oriented.
Figure 8. Saturation magnetization, remanence and coercivity Nd1.1Fe10CoTiNx alloy nitrogenated during 6 h and as a function of temperature. The specimen was not magnetically oriented.
Molecules 26 03854 g008
Table 1. Lattice parameters, unit cell volume V, volume fraction and density ρ, for phases in the Nd1.1Fe10CoTiNx alloys nitrogenated for 0, 3, 6, 9 and 12 h. All data were obtained through refinement of patterns.
Table 1. Lattice parameters, unit cell volume V, volume fraction and density ρ, for phases in the Nd1.1Fe10CoTiNx alloys nitrogenated for 0, 3, 6, 9 and 12 h. All data were obtained through refinement of patterns.
Nitrogenation Time (h)Phase a (Å)c (Å) V (Å3)Fraction (vol. %)ρ (g/cm3)
01:128.5894.808354.795.77.579
bcc2.870 4.37.848
31:12 (a)8.5854.804354.143.77.593
1:12(b)8.6314.843360.817.77.452
1:12(c)8.6424.953369.934.27.268
bcc2.870 4.47.843
61:12(a)8.5834.805354.042.37.595
1:12(b)8.6164.849360.022.47.469
1:12(c)8.6384.971370.929.57.249
bcc2.871 5.87.840
91:12(a)8.5854.804354.134.37.593
1:12(b)8.6314.843360.810.07.452
1:12(c)8.7004.977376.748.97.136
bcc2.873 6.87.820
121:12(a)8.5844.802353.836.57.597
1:12(b)8.6424.841361.59.57.436
1:12(c)8.7414.950378.245.57.107
bcc2.871 8.57.840
Note: Uncertainties for the lattice parameters are ±0.001 Å and for unit cell volume is ±0.16 Å3.
Table 2. Magnetic properties of aligned and randomly oriented (in parenthesis) Nd1.1Fe10CoTiNx powders nitrogenated for the indicated time.
Table 2. Magnetic properties of aligned and randomly oriented (in parenthesis) Nd1.1Fe10CoTiNx powders nitrogenated for the indicated time.
Nitrogenation Time (h)M3T
(Am2kg−1)
Mr (Am2kg−1)μ0Hc (T)(BH)max (kJ·m−3)
0122.83 (121.92)108.21 (57.79)0.294 (0.276)81.07
(30.44)
3128.80
(114.17)
116.34 (75.27)0.504 (0.500)113.74
(58.38)
6128.97 (106.90)116.14 (62.16)0.518 (0.457)111.41
(39.89)
9126.38 (112.45)113.24 (68.28)0.395 (0.371)96.40
(44.02)
12126.97 (115.12)112.87 (71.15)0.316 (0.312)87.85
(45.74)
Note: Uncertainties for M3T, Mr, μ 0Hc and (BH)max are ±0.20 Am2kg−1, ±0.14 Am2kg−1, ±0.001 T and ±0.35 kJ·m−3, respectively.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínez Sánchez, H.; Hadjipanayis, G.; Pérez Alcázar, G.A.; Zamora Alfonso, L.E.; Trujillo Hernández, J.S. Mechanochemical Synthesis and Nitrogenation of the Nd1.1Fe10CoTi Alloy for Permanent Magnet. Molecules 2021, 26, 3854. https://doi.org/10.3390/molecules26133854

AMA Style

Martínez Sánchez H, Hadjipanayis G, Pérez Alcázar GA, Zamora Alfonso LE, Trujillo Hernández JS. Mechanochemical Synthesis and Nitrogenation of the Nd1.1Fe10CoTi Alloy for Permanent Magnet. Molecules. 2021; 26(13):3854. https://doi.org/10.3390/molecules26133854

Chicago/Turabian Style

Martínez Sánchez, Hugo, George Hadjipanayis, Germán Antonio Pérez Alcázar, Ligia Edith Zamora Alfonso, and Juan Sebastián Trujillo Hernández. 2021. "Mechanochemical Synthesis and Nitrogenation of the Nd1.1Fe10CoTi Alloy for Permanent Magnet" Molecules 26, no. 13: 3854. https://doi.org/10.3390/molecules26133854

Article Metrics

Back to TopTop