Next Article in Journal
Experimental Investigation on the Surface Formation Mechanism of NdFeB during Diamond Wire Sawing
Next Article in Special Issue
The Effect of Substitution of Mn by Pd on the Structure and Thermomagnetic Properties of the Mn1−xPdxCoGe Alloys (Where x = 0.03, 0.05, 0.07 and 0.1)
Previous Article in Journal
A Contrast Analysis of Deformation Characteristics and Critical Dynamic Stress of Natural and Fiber-Binder Reinforced Subgrade Filler after Different Freeze-Thaw Cycles
Previous Article in Special Issue
Entalpy of Mixing, Microstructure, Structural, Thermomagnetic and Mechanical Properties of Binary Gd-Pb Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Nd Substitution on the Phase Constitution in (Zr,Ce)Fe10Si2 Alloys with the ThMn12 Structure

1
Institute of Molecular Physics, Polish Academy of Sciences, Mariana Smoluchowskiego 17, 60-179 Poznań, Poland
2
NanoBioMedical Centre, Adam Mickiewicz University, Wszechnicy Piastowskiej 3, 61-614 Poznań, Poland
3
Institut des Molécules et Matériaux du Mans (IMMM, UMR CNRS 6283), Le Mans University, Avenue Olivier Messiaen, CEDEX 09, 72085 Le Mans, France
4
Institute of Materials Engineering, Faculty of Science and Technology, University of Silesia, 75 Pułku Piechoty 1a, 41-500 Chorzów, Poland
5
Department of Physics, Faculty of Science, University of Hradec Králové, Rokitanského 62, 500 03 Hradec Králové, Czech Republic
6
CNRS, ICMPE, University Paris Est Creteil, UMR 7182, 2 Rue Henri Dunant, 94320 Thiais, France
*
Author to whom correspondence should be addressed.
Materials 2023, 16(4), 1522; https://doi.org/10.3390/ma16041522
Submission received: 11 January 2023 / Revised: 6 February 2023 / Accepted: 9 February 2023 / Published: 11 February 2023

Abstract

:
Iron-based compounds with a ThMn12-type structure have the potential to bridge the gap between ferrites and high performance Nd2Fe14B magnets. From the point of view of possible applications, the main advantage is their composition, with about 10 wt.% less rare earth elements in comparison with the 2:14:1 phase. On the other hand, the main issue delaying the development of Fe-rich alloys with a ThMn12-type structure is their structural stability. Therefore, various synthesis methods and stabilizing elements have been proposed to stabilize the structure. In this work, the influence of increasing Nd substitution on the phase constitution of Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys was analyzed. X-ray diffraction and 57Fe Mössbauer spectrometry were used as the main methods to derive the stability range and destabilization routes of the 1:12 structure. For the arc-melted samples, an increase in the lattice parameters of the ThMn12-type structure was observed with the simultaneous growth of bcc-(Fe,Si) content with increasing Nd substitution. After isothermal annealing, the ThMn12-type structure (and the coexisting bcc-(Fe,Si)) were stable over the whole composition range. While the formation of a 1:12 phase was totally suppressed in the as-cast state for x = 0.3, further heat treatment resulted in the growth of about 45% of the ThMn12-type phase. The results confirmed that the stability range of ThMn12-type structure in the Nd-containing alloys was well improved by other substitutions and the heat treatment, which in turn, is also needed to homogenize the ThMn12-type phase. After further characterization of the magnetic properties and optimization of microstructure, such hard/soft magnetic composites can show their potential by exploiting the exchange spring mechanism.

1. Introduction

Currently, permanent magnets are widely used and the most known among them, with the highest energy product BH|max|, are based on rare earth elements (REE), such as Sm or Nd [1,2], with well-known representatives in Nd2Fe14B [3], Sm2Co17 [4], and Sm2Fe17N3 [5]. The range of possible applications is broad and is still growing [6]. With the development of new phases and the improvement of intrinsic and extrinsic properties, permanent magnets are still implemented in a rather trivial way, as a replacement of electromagnets, in loudspeakers and actuators, but also find more sophisticated applications in magnetic resonance imaging, in information storage devices, magnetic levitation, as bearings, or as miniaturized magnetic field sensors [6,7]. The demand is still growing, as huge quantities of permanent magnets are used in, e.g., the automotive industry or wind turbine production [8]. Their criticality (growing demand with unstable supply chains) has led to the extensive research of new materials [9], where two important directions can be distinguished: (i) the decrease of REE content with preservation of magnets’ magnetic performance and (ii) the search for novel REE-free hard magnetic materials, e.g., L10-FeNi [10] known from meteorites [11]. Moreover, there is a renewed interest in the tetragonal ThMn12-type (1:12 stoichiometry) phase, which has already been known for decades [12]. This tetragonal structure is related to that of SmCo5 of CaCu5-type, where half of the rare earth atoms are replaced with transition metal dumbbells [7]. Even if its BH|max| would be slightly lower compared with high performance magnets, it would be still placed in the performance gap between cheaper ferrite-based and more expensive Nd-based magnets with 2:14:1 structure [13]. Nevertheless, another issue that has to be solved with these compounds is their structural stability. For most Fe- and REE-based compounds, ThMn12-type phase is less stable than, for example, the hexagonal 2:17-type structure [14]. Therefore, stabilizing elements, such as Ti, V, Si or Al, have to be used [15,16,17]. For example, the SmFe12 compound forms as a thin film, but is unstable in bulk [18]. In such cases, large transition metal atoms are needed to stabilize the structure, as in SmFe11Ti [19]. The simultaneous substitution of other elements, for example, Zr on rare earth sites and Co on Fe sites, allowed a decrease in Ti content in (Sm0.8Zr0.2) (Fe0.75Co0.25)11.5Ti0.5 and the optimization of the magnetic properties [20]. Additionally, there are several further issues that must be taken into account, such as microstructure and twinning [18,21]. One of the largest groups of alloys with the ThMn12 structure is based on Sm [22,23,24,25]. Nonetheless, for already known and investigated 1:12-type compounds, the maximum effort was made to develop Ce-based compounds due to the low criticality of this element [26,27,28]. Recently, the formation of the ThMn12 structure (almost 100% of volume fraction) has been confirmed in Zr0.4Ce0.6Fe10Si2 [29]. The Ce substitution increased the anisotropy field from 16.9 kOe for ZrFe10Si2 to 24 kOe for Zr0.4Ce0.6Fe10Si2. Replacement of Zr by Nd in ZrFe10Si2 and further nitrogenation of this alloy have also been investigated [30] and are believed to be very promising in the optimization of hard magnetic properties. Nd0.4Zr0.6Fe10-Si2 has been also examined to confirm whether the formation of the ThMn12-type phase is possible and to determine its magnetic and structural properties [31]. It has also been shown than the replacement of Fe atoms by Co improved magnetic performance, with the enhancement of the Curie temperature as a main benefit [31]. These results raised a question: how would substitution of Zr by Nd in Zr0.4Ce0.6Fe10Si2 affect formation of the 1:12 structure [29]? Nd is believed to enhance the magnetic performance of this alloy [32]. As-cast samples were also isothermally annealed at 1373 K, the temperature which has been reported before [33], to lie in the stability region for ThMn12 phase formation. Therefore, on the basis of structural and spectroscopic measurements, the structural stability of the ThMn12-type phase in the Nd-substituted Zr0.4Ce0.6Fe10Si2 alloy was determined in this paper.

2. Methods

Samples with Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) compositions were prepared by arc-melting in an argon atmosphere. The melting procedure was repeated several times to ensure homogeneity. Additionally, residual oxygen was gathered by melting a Zr getter. Afterwards, samples wrapped in Ta foil were placed in quartz capsules, which were then evacuated three times with the use of turbomolecular pump (down to 10−4 mbar), refilled with Ar up to 500 mbar and sealed. Samples were then annealed isothermally at 1373 K in the Carbolite MTF 12/25/250 tube furnace for 72 h and subsequently quenched in cold water. For microscopic measurements, the powdered sample was suspended in isopropyl alcohol, and the resulting material, after dispersion in an ultrasonic bath for 30 min, was deposited on a Cu grid with an amorphous carbon film standardized for transmission electron microscope (TEM) observations. Structural analysis was performed using X-ray powder diffraction (XRPD). Patterns were recorded with a PANalytical X’Pert Pro diffractometer in Bragg–Brentano geometry (CoKα radiation). A cobalt anode was chosen to avoid X-ray fluorescence of iron compounds. Data were collected in the range of 2θ from 20 to 110° with a 0.0167° step and 400 s per step for a total acquisition time of 5 h. X-ray diffraction patterns were analyzed using MAUD software (Rietveld method combined with Fourier analysis) [34] to determine the phase constitution, volume fractions and lattice parameters of various phases. Transmission 57Fe Mössbauer effect experiments were performed at room temperature (RT) using a constant acceleration conventional spectrometer with a 57Co source. Samples were powdered using an Fe-free tool. The powder obtained was collected and placed in an altuglass sample holder with a content of about 5 mg Fe/cm2. The hyperfine structure was modeled by least square refinement involving magnetic and quadrupolar components with Lorentzian lines. The values of the isomer shift are quoted relative to those of bcc-Fe at RT. Initial measurements were made in the range from −12 to 12 mm/s, to check the presence of oxides. Final spectra with improved statistics were measured from −8 to 8 mm/s. Microstructure analysis of the selected samples was carried out using JEOL JEM-3010 high-resolution transmission electron microscope with 300 kV acceleration voltage, equipped with a Gatan 2k × 2k Orius™ 833 SC200D CCD camera. Recorded selected area electron diffraction (SAED) patterns, as well as the calculated Fast Fourier Transforms (FFT) from high-resolution images were indexed using ElDyf [35] computer software.

3. Results and Discussion

The X-ray diffraction patterns for Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys synthesized by the arc-melting method are shown in Figure 1. The Zr0.4Ce0.6Fe10Si2 alloy contains about 90% of the ThMn12-type structure with the addition of about 10% of bcc-(Fe,Si) phase. The presence of Si in bcc structure was expected, taking into account changed lattice parameter values (Table 1), and was confirmed by spectroscopic measurements. By replacing 25 at.%, 50 at.% and 75 at.% of Zr by Nd, we obtained nominally Zr0.3Nd0.1Ce0.6Fe10Si2, Zr0.2Nd0.2Ce0.6Fe10Si2, and Zr0.1Nd0.3Ce0.6Fe10Si2 alloys. In the first case, about 82% of the ThMn12-type structure and 18% of bcc-(Fe,Si) phase formed. Substitution of 50% of Zr by Nd led to further suppression of ThMn12-type phase formation and facilitated the growth of highly disordered hexagonal phase, nominally (REE)Fe5. About 52% of ThMn12-type phase formed along with 25% of the bcc-(Fe,Si) phase and an insignificant amount of α-Zr. One must bear in mind that the fitting procedure for an α-Zr phase is based just on two reflections. Therefore, this finding should be treated with a caution. Nevertheless, for clarity reasons, we use this designation throughout the paper. In Zr0.1Nd0.3Ce0.6Fe10Si2, the ThMn12-type structure was not observed, while the presence of hexagonal-type phase, bcc-(Fe,Si) phase and α-Zr was confirmed. The bcc-(Fe,Si) lattice parameter does not vary much depending on the Nd content, and the bcc lattice contained about 8–10 at.% of Si in every investigated sample. The Nd influence on the values of ThMn12-type phase lattice parameters and lattice volume was noticed, where all parameters (a = b and c) increased with the growth of Nd content (Table 1). Along with the influence of thermodynamic parameters (e.g., enthalpies of formation), lattice expansion could be one of the main reasons for the instability of the ThMn12-type structure for large substitutions of Zr by Nd. Detailed results are presented in Figure 2. Zirconium and silicon play a significant role in the stabilization of the ThMn12-type structure in Fe-based alloys [29,30]. The influence of Nd, Ce and Zr on the enthalpies of formation of various phases in the Fe-Si-based alloys has been also analyzed by us lately [36]. Beneficial influence of Zr on the formation of the ThMn12-type phase was confirmed, but the enthalpy of formation of other competing phases (i.e., bcc solid solution and amorphous phase) also decreased significantly with increased Zr content. When we analyzed enthalpy values only, it has been reported that Ce and Nd should have deteriorating impact on the stability of crystalline phases. Nevertheless, with small rare earth element substitutions, other quantities such as a mismatch entropy [36] or site preference [30] play a role in the formation of ThMn12-type phase.
Mössbauer spectrometry was used to support the structural investigations. Several sextets with different hyperfine field parameters, suggesting a multiphase character of investigated samples and numerous magnetically distinguishable surroundings of Fe atoms were observed. As already mentioned, the spectrum initially assigned in the fitting procedure to be bcc-Fe was non-typical, as it was composed of at least three sub-spectra. The determined hyperfine parameters values were confronted with the literature data [37]. We considered the values of specific hyperfine field parameters and applied them in a fitting model according to wt.% of Si in Fe [37]. The results suggested, along with the lattice parameter value of bcc phase, that the observed bcc-Fe phase was in fact the bcc-(Fe,Si). Part of Si diffuses into the bcc lattice and is not directly involved in the ThMn12-type structure stabilization. Mössbauer spectra are shown in Figure 3, while hyperfine parameters are listed in Table 2. Notably, more precise estimation of Si content in bcc lattice was possible, as the lattice parameter of bcc structure and Si content were linearly dependent on each other [38]. Therefore, it was possible to extrapolate refined lattice parameter to confirm the content of Si diffusing into the bcc structure.
Afterwards, isothermal annealing of arc-melted alloys at 1373 K for 72 h (conditions favorable for ThMn12-type structure formation [33]) was performed. X-ray diffraction patterns are shown in Figure 4. After heat treatment, higher precipitation of bcc-(Fe,Si) was observed for the parent Zr0.4Ce0.6Fe10Si2 alloy, with about 17% of total volume fraction, when compared to 10% in the arc-melted sample. A small amount of 2:17 structure formation and an insignificant amount of ZrFeSi2 phase were also noticed. In contrast to as-cast samples, the ThMn12-type phase was present in the whole composition range in the annealed samples. As observed in arc-melted samples already, Nd-substitution led to a decreasing volume fraction of the ThMn12-type structure. The inverse dependence was observed for the bcc-(Fe,Si) phase. The amount of hexagonal 2:17-type phase and ZrFeSi2 phase did not exceed 5 at.% in the whole composition range. Silicon content in the bcc lattice was determined by estimating the lattice parameter [37,38]. Its content is in the range of 9–10 at.%, in accordance with the previous results for the as-cast samples. Lattice parameters of ThMn12-type phase increased with Nd content (Table 1). As the refinement of the lattice parameters was made using the Rietveld method, the estimated standard deviation was the only measure of the uncertainty of the lattice parameters which did not exceed 0.001 Å.
The Mössbauer spectrometry results showed that the ThMn12-type structure obtained after arc-melting was clearly not homogeneous. As it was not possible to describe three subspectra for 8i, 8j and 8f positions (Figure 3), a distribution of parameters was used by similarity to these already known from previous investigations, where Ce substitution for Nd in similar alloys was reported to be responsible for lowering of the hyperfine field (Bhf) [39,40,41]. Consequently, the mean value of the hyperfine field parameter was checked and estimated to be equal to 28.7 ± 0.5, 28.7 ± 0.5, 29.9 ± 0.5 and 30.6 ± 0.5 T for the Zr0.4Ce0.6Fe10Si2, Zr0.3Nd0.1Ce0.6Fe10Si2, Zr0.2Nd0.2Ce0.6Fe10Si2, and Zr0.1Nd0.3Ce0.6Fe10Si2, respectively. This slight increase was connected with the growth of a disordered hexagonal phase replacing the ThMn12-type one. All hyperfine parameters determined from the spectra shown in Figure 3 are presented in Table 2. We observed the evolution of spectra with Nd substitution, which is in accordance with X-ray diffraction results, and we concluded that the content of pure bcc-Fe phase (with no Si in the nearest neighborhood) increases from 3% for Zr0.4Ce0.6Fe10Si2, to 6, 11 and 17% for Zr0.3Nd0.1Ce0.6Fe10Si2, Zr0.2Nd0.2Ce0.6Fe10Si2, and Zr0.1Nd0.3Ce0.6Fe10Si2, respectively. The same trend occurred for the bcc phase sub spectrum that was assumed to contain 12 at.% of Si (from 2, through 3, 7 up to 11% with increasing Nd content). The content of another sub-spectrum (bcc phase with 26 at.% of Si in Fe) is equal to about 6% for each sample. Small inaccuracies are possible because of overlapping spectra with hyperfine field parameters of about 24 T, coming from both the ThMn12-type structure and bcc phase with 26 at.% of Si in Fe. Annealed samples revealed much better homogeneity than arc-melted alloys, as shown by more consistent positions and narrower peaks observed in Mössbauer spectra (Figure 5). The spectra for 8i, 8j and 8f positions for the ThMn12-type phase were fully described. The hyperfine field parameter of Fe in 8i and 8j positions increased with the substitution of Nd for Zr (Table 3). Variation of Bhf was not so evident for the 8f position. It changed from 19.8 ± 0.5, through 19.4 ± 0.5, 19.5 ± 0.5 to 19.4 ± 0.5 T with the increasing Nd content. There is also a doublet visible in the middle of the Mössbauer spectra of all annealed samples, whose contribution increased with Nd substitution. It was assumed to originate from (Zr,Nd,Ce)FeSi2 phase, as its content also follows the annealed alloys’ constitution determined on the basis of X-ray diffraction. One must bear in mind the presence of grain boundaries (GB), which may play a large role while analyzing the Mössbauer spectra. Their contribution was estimated to be equal to about 10%, with a Bhf mean value varying from 19.8 ± 0.5 T for Zr0.4Ce0.6Fe10Si2, through 18.7 ± 0.5, 18.3 ± 0.5, down to 18.2 ± 0.5 T, for Zr0.3Nd0.1Ce0.6Fe10Si2, Zr0.2Nd0.2Ce0.6Fe10Si2, and Zr0.1Nd0.3Ce0.6Fe10Si2, respectively. Their presence can have a great influence on the magnetic properties of nanocomposite [40,41].
To examine the microstructure of the samples obtained and to confirm the phase composition obtained by X-ray diffraction, transmission electron microscopy (TEM) was performed on two selected samples: arc melted and annealed Zr0.4Ce0.6Fe10Si2. The TEM measurements were performed mainly to confirm the existence of grain boundaries, which were presumed mainly on the basis of Mössbauer spectrometry results. We took advantage of the use of this method to confirm simultaneously that, in fact, we deal mainly with ThMn12-type and (Fe,Si)-bcc phases. Nevertheless, we observed that the phase constitution and the mean crystallite size differ for various pieces of the same sample. Therefore, we omitted a more detailed analysis as it would not bring the results, which are representative for the sample as a whole. Nonetheless, the sample after arc-melting showed coarse grains. Electron diffraction measurements confirmed the presence of the tetragonal ThMn12-type structure and the bcc-(Fe,Si) phase (Figure 6). TEM observations of the annealed sample were in good agreement with the X-ray diffraction results. The tetragonal ThMn12-type phase and bcc-(Fe,Si) phase were also observed in the annealed sample. It was not able to determine the crystallite sizes of ThMn12-type phase precisely. For the bcc-(Fe,Si) phase, the recorded high-resolution images revealed a well-crystallized nanocrystalline structure with a crystallite size of about 9 nm (Figure 7). Some grain boundaries and/or structurally disordered regions are also visible, as suggested before by Mössbauer spectrometry.

4. Conclusions

Several Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys mainly composed of a ThMn12-type structure were synthesized by arc-melting and subsequent annealing. The presence of the bcc phase, which was found to contain Si with a rather stable content throughout the whole series of samples (about 9–10 at.% of Si in bcc-Fe lattice), was confirmed for all samples. Moreover, a facilitated formation of bcc-(Fe,Si) was observed with increasing Nd content. The ThMn12-type structure became less stable with increasing Nd content, also due to a slight increase in lattice parameters/volume. The 57Fe Mössbauer spectrometry results confirmed that the ThMn12-type structure in the pre-annealed samples was highly disordered, and that the distribution of hyperfine parameters had to be used to describe the spectra correctly. For an equiatomic Zr/Nd composition, the crystallization of a highly disordered hexagonal phase, nominally (REE)Fe5, was observed. This process resulted in a simultaneous decrease in the driving force for the formation of a ThMn12-type structure (totally suppressed for x = 0.3). In contrast, the ThMn12-type phase was present in the whole composition range, along with the bcc-(Fe,Si), in the annealed samples. In addition, the growth of the 2:17 hexagonal phase was observed, whose content increased with Nd substitution (from 2 at.% up to 5 at.%). Importantly, the ThMn12 structure became more homogeneous after annealing. The homogeneity, as well as the improved stability, of the ThMn12-type phase (even for high Nd substitution) confirms that the preconceived synthesis route and the choice of stabilizing elements (Si, Zr, Nd, Ce) were appropriate for the future development of these alloys. It was evident that Nd adversely affected stability of 1:12 structure, but it was confirmed that the preservation of its high content was still possible (with Zr, Ce and Si substitution). Further experiments are needed to determine the influence of Nd-substitution on magnetic properties. Moreover, attention should be drawn to the possibilities of exploitation of the exchange spring effect, due to the presence of a soft magnetic bcc phase. By optimizing the content of hard and soft magnetic phases and their microstructure [42], one can expect improved values of coercivity and remanent magnetization, leading to the maximization of overall magnet performance.

Author Contributions

Conceptualization, M.K. and Z.Ś.; Data curation, M.K., J.-M.G., M.Z. and Z.Ś.; Formal analysis, M.K., J.-M.G. and Z.Ś.; Investigation, M.K., S.A., J.-M.G. and M.Z.; Methodology, M.K., J.-M.G. and Z.Ś.; Supervision, Z.Ś.; Validation, M.K. and Z.Ś.; Visualization, M.K., M.Z. and Z.Ś.; Writing—Original draft, M.K. and M.Z., Writing—review & editing, M.K., J.-M.G., S.A., B.I., L.B. and Z.Ś. All authors have read and agreed to the published version of the manuscript.

Funding

M.K. was financially supported by the project “Środowiskowe interdyscyplinarne studia doktoranckie w zakresie nanotechnologii” (“Environmental interdisciplinary doctoral studies in nanotechnology”) No. POWR.03.02.00-00-I032/16 under the European Social Fund –Operational Programme Knowledge Education Development, Axis III Higher Education for Economy and Development, Action 3.2 PhD Programme. Financial support of B.I. investigations by the Polish Academy of Sciences in the framework of the Polish-French cooperation between PAS and CNRS in the form of long-term visit is also acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors thank Agnieszka Grabias for discussions and critical revision.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pflüger, J.; Heinen, W.; Gudat, W. On the thermal stability of the SmCo5 magnet structure for the BESSY undulator. Nucl. Instrum. Methods Phys. Res. Sect. B Accel. Spectrom. Detect. Assoc. Equip. 1987, 262, 528–533. [Google Scholar] [CrossRef]
  2. Sagawa, M.; Fujimura, S.; Togawa, N.; Yamamoto, H.; Matsuura, Y. New material for permanent magnets on a base of Nd and Fe (invited). J. Appl. Phys. 1984, 55, 2083–2087. [Google Scholar] [CrossRef]
  3. Sagawa, M.; Fujimura, S.; Yamamoto, H.; Matsuura, Y.; Hiraga, K. Permanent magnet materials based on the rare earth-iron-boron tetragonal compounds. IEEE Trans. Magn. 1984, 20, 1584–1589. [Google Scholar] [CrossRef]
  4. Kumar, K. RETM5 and RE2TM17 permanent magnets development. J. Appl. Phys. 1988, 63, R13–R57. [Google Scholar] [CrossRef]
  5. Iriyama, T.; Kobayashi, K.; Imaoka, N.; Fukuda, T.; Kato, H.; Nakagawa, Y. Effect of nitrogen content on magnetic properties of Sm2Fe17Nx (0 < x < 6). IEEE Trans. Magn. 1992, 28, 2326–2331. [Google Scholar] [CrossRef]
  6. Coey, J.M.D. Permanent magnet applications. J. Magn. Magn. Mater. 2002, 248, 441–456. [Google Scholar] [CrossRef]
  7. Hirosawa, S.; Nishino, M.; Miyashita, S. Perspectives for high-performance permanent magnets: Applications, coercivity, and new materials. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 013002. [Google Scholar] [CrossRef]
  8. Takahashi, Y.; Sepehri-Amin, H.; Ohkubo, T. Recent advances in SmFe12-based permanent magnets. Sci. Technol. Adv. Mater. 2021, 22, 449–460. [Google Scholar] [CrossRef]
  9. Skomski, R.; Manchanda, P.; Kumar, P.; Balasubramanian, B.; Kashyap, A.; Sellmyer, D. Predicting the Future of Permanent-Magnet Materials. Trans. Magn. 2013, 49, 3215. [Google Scholar] [CrossRef]
  10. Paulevé, J.; Chamberod, A.; Krebs, K.; Bourret, A. Magnetization Curves of Fe–Ni (50–50) Single Crystals Ordered by Neutron Irradiation with an Applied Magnetic Field. J. Appl. Phys. 1968, 39, 989–990. [Google Scholar] [CrossRef]
  11. Lewis, L.H.; Mubarok, A.; Poirier, E.; Bordeaux, N.; Manchanda, P.; Kashyap, A.; Skomski, R.; Goldstein, J.; Pinkerton, F.E.; Mishra, R.K. Inspired by nature: Investigating tetrataenite for permanent magnet applications. J. Phys. Condens. Matter 2014, 26, 064213. [Google Scholar] [CrossRef]
  12. Solzi, M.; Pareti, L.; Moze, O.; David, W.I.F. Magnetic anisotropy and crystal structure of intermetallic compounds of the ThMn12 structure. J. Appl. Phys. 1988, 64, 5084–5087. [Google Scholar] [CrossRef]
  13. Coey, J.M.D. Permanent magnets: Plugging the gap. Scr. Mater. 2012, 67, 524–529. [Google Scholar] [CrossRef]
  14. Harashima, Y.; Fukazawa, T.; Kino, H.; Miyake, T. Effect of R-site substitution and the pressure on stability of RFe12: A first-principles study. J. Appl. Phys. 2018, 124, 163902. [Google Scholar] [CrossRef]
  15. 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, 868–870. [Google Scholar] [CrossRef]
  16. Buschow, K.H.J. Permanent magnet materials based on tetragonal rare earth compounds of the type RFe12−xMx. J. Magn. Magn. Mater. 1991, 100, 79–89. [Google Scholar] [CrossRef]
  17. 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. 2016, 7, 056202. [Google Scholar] [CrossRef]
  18. Gabay, A.M.; Hadjipanayis, G.C. Recent developments in RFe12-type compounds for permanent magnets. Scr. Mater. 2018, 154, 284–288. [Google Scholar] [CrossRef]
  19. Isnard, O.; Guillot, M.; Miraglia, S.; Fruchart, D. High field magnetization measurements of SmFe11Ti and SmFe11TiH1−δ. J. Appl. Phys. 1996, 79, 5542. [Google Scholar] [CrossRef]
  20. Kuno, T.; Suzuki, S.; Urushibata, K.; Kobayashi, K.; Sakuma, N.; Yano, M.; Kato, A.; Manabe, A. (Sm,Zr)(Fe,Co)11.0-11.5Ti1.0-0.5 compounds as new permanent magnet materials. AIP Adv. 2016, 6, 025221. [Google Scholar] [CrossRef] [Green Version]
  21. Ener, S.; Skokov, K.P.; Palanisamy, D.; Devillers, T.; Fischbacher, J.; Eslava, G.G.; Maccari, F.; Schäfer, L.; Diop, L.V.B.; Radulov, I.; et al. Twins—A weak link in the magnetic hardening of ThMn12-type permanent magnets. Acta Mater. 2021, 214, 116968. [Google Scholar] [CrossRef]
  22. Matsumoto, M.; Hawai, T.; Ono, K. (Sm,Zr)Fe12−xMx (M = Zr, Ti, Co) for Permanent-Magnet Applications: Ab Initio Material Design Integrated with Experimental Characterization. Phys. Rev. Appl. 2020, 13, 064028. [Google Scholar] [CrossRef]
  23. Landa, A.; Söderlind, P.; Moore, E.E.; Perron, A. Thermodynamics and Magnetism of SmFe12 Compound Doped with Co and Ni: An Ab Initio Study. Appl. Sci. 2022, 12, 4860. [Google Scholar] [CrossRef]
  24. Fukazawa, T.; Harashima, Y.; Miyake, T. First-principles study on the stability of (R, Zr)(Fe, Co, Ti)12 against 2-17 and unary phases (R = Y, Nd, Sm). Phys. Rev. Mater. 2022, 6, 054404. [Google Scholar] [CrossRef]
  25. Gabay, A.M.; Hadjipanayis, G.C. Mechanochemical synthesis of magnetically hard anisotropic RFe10Si2 powders with R representing combinations of Sm, Ce and Zr. J. Magn. Magn. Mater. 2017, 422, 43–48. [Google Scholar] [CrossRef]
  26. Zhou, C.; Pinkerton, F.E. Magnetic hardening of CeFe12−xMox and the effect of nitrogenation. J. Alloys Compd. 2013, 583, 345–350. [Google Scholar] [CrossRef]
  27. Bhandari, C.; Paudyal, D. Enhancing stability and magnetism of ThMn12-type cerium-iron intermetallics by site substitution. Phys. Rev. Res. 2022, 4, 023012. [Google Scholar] [CrossRef]
  28. Snarski-Adamski, J.; Werwiński, M. Effect of transition metal doping on magnetic hardness of CeFe12-based compounds. J. Magn. Magn. Mater. 2022, 554, 169309. [Google Scholar] [CrossRef]
  29. Gabay, A.M.; Hadjipanayis, G.C. ThMn12-type structure and uniaxial magnetic anisotropy in ZrFe10Si2 and Zr1−xCexFe10Si2 alloys. J. Alloys Compd. 2016, 657, 133–137. [Google Scholar] [CrossRef]
  30. Barandiaran, J.M.; Martin-Cid, A.; Schönhöbel, A.M.; Garitaonandia, J.S.; Gjoka, M.; Niarchos, D.; Makridis, S.; Pasko, A.; Aubert, A.; Mazaleyrat, F.; et al. Nitrogenation and sintering of (Nd-Zr)Fe10Si2 tetragonal compounds for permanent magnets applications. J. Alloys Compd. 2019, 784, 996–1002. [Google Scholar] [CrossRef] [Green Version]
  31. Gjoka, M.; Sarafidis, C.; Giannopoulos, G.; Niarchos, D.; Hadjipanayis, G.; Tabares, J.A.; Alcázar, G.A.P.; Zamora, L.E. Effect of cobalt substitution on structure and magnetic properties of Nd0.4Zr0.6Fe10−xCoxSi2 (x = 0–3) alloys and their ribbons. J. Rare Earths 2019, 37, 1096–1101. [Google Scholar] [CrossRef]
  32. Martin-Cid, A. Development of New High Anisotropy Phases for Permanent Magnet Applications. PhD Thesis, Universidad del Pais Vasco, Pais Vasco, Spain, 2018. [Google Scholar]
  33. Kobayashi, K.; Furusawa, D.; Suzuki, S.; Kuno, T.; Urushibata, K.; Sakuma, N.; Yano, M.; Shoji, T.; Kato, A.; Manabe, A.; et al. High-Temperature Stability of ThMn12 Magnet Materials. Mater. Trans. 2018, 59, 1845–1853. [Google Scholar] [CrossRef]
  34. Lutterotti, L. Total pattern fitting for the combined size–strain–stress–texture determination in thin film diffraction. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. Atoms 2009, 268, 334–340. [Google Scholar] [CrossRef]
  35. Gigla, M.; Pączkowski, P. ElDyf for Windows; Version 2.1; Institute of Materials Science, University of Silesia in Katowice: Katowice, Poland, 2005. [Google Scholar]
  36. Kołodziej, M.; Śniadecki, Z. Thermodynamic Modeling of Formation Enthalpies of Amorphous and Crystalline Phases in Zr, Nd, and Ce Substituted Fe-Si System. Appl. Sci. 2023, 13, 1966. [Google Scholar] [CrossRef]
  37. Randrianantoandro, N.; Gaffet, E.; Mira, J.; Greneche, J.-M. Magnetic hyperfine temperature dependence in Fe–Si crystalline alloys. Solid State Commun. 1999, 111, 323–327. [Google Scholar] [CrossRef]
  38. Varga, L.K.; Mazaleyrat, F.; Kovac, J.; Greneche, J.M. Structural and magnetic properties of metastable Fe1-xSix (0.15 < x < 0.34) alloys prepared by a rapid-quenching technique. J. Phys. Condens. Matter 2002, 14, 1985–2000. [Google Scholar] [CrossRef]
  39. Gjoka, M.; Psycharis, V.; Devlin, E.; Niarchos, D.; Hadjipanayis, G. Effect of Zr substitution on the structural and magnetic properties of the series Nd1−xZrxFe10Si2 with the ThMn12 type structure. J. Alloys Compd. 2016, 687, 240–245. [Google Scholar] [CrossRef]
  40. Zhao, L.Z.; Yu, H.Y.; Guo, W.T.; Zhang, J.S.; Zhang, Z.Y.; Hussain, M.; Liu, Z.W.; Greneche, J.M. Phase and Hyperfine Structures of Melt-spun Nanocrystalline (Ce1–xNdx)16Fe78B6 Alloys. IEEE Trans. Magn. 2017, 53, 1–5. [Google Scholar] [CrossRef]
  41. Zhao, L.; Li, C.; Hao, Z.; Liu, X.; Liao, X.; Zhang, J.; Su, K.; Li, L.; Yu, H.; Greneche, J.-M.; et al. Influences of element segregation on the magnetic properties in nanocrystalline Nd-Ce-Fe-B alloys. Mater. Charact. 2018, 148, 208–213. [Google Scholar] [CrossRef]
  42. Idzikowski, B.; Wolf, M.; Handstein, A.; Nenkov, K.; Engelmann, H.J.; Stobiecki, F.; Muller, K.-H. Inverse giant magnetoresistance in granular Nd/sub 2/Fe/sub 14/B//spl alpha/-Fe. IEEE Trans. Magn. 1997, 33, 3559–3561. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys. ThMn12-type structure, α-(Fe,Si), α-Zr and highly disordered hexagonal structure were identified and marked with respective symbols.
Figure 1. X-ray diffraction patterns of arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys. ThMn12-type structure, α-(Fe,Si), α-Zr and highly disordered hexagonal structure were identified and marked with respective symbols.
Materials 16 01522 g001
Figure 2. Refined values of lattice parameters for (a) ThMn12 and (b) bcc-(Fe,Si) structures according to Nd content for arc-melted (white points) and annealed at 1373 K for 72 h (black points) Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys.
Figure 2. Refined values of lattice parameters for (a) ThMn12 and (b) bcc-(Fe,Si) structures according to Nd content for arc-melted (white points) and annealed at 1373 K for 72 h (black points) Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys.
Materials 16 01522 g002
Figure 3. Mössbauer spectra in arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys.
Figure 3. Mössbauer spectra in arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys.
Materials 16 01522 g003
Figure 4. X-ray diffraction patterns of Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys annealed at 1373 K for 72 h, mainly composed of ThMn12-type structure.
Figure 4. X-ray diffraction patterns of Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys annealed at 1373 K for 72 h, mainly composed of ThMn12-type structure.
Materials 16 01522 g004
Figure 5. Mössbauer spectra of Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys annealed at 1373 K for 72 h.
Figure 5. Mössbauer spectra of Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) alloys annealed at 1373 K for 72 h.
Materials 16 01522 g005
Figure 6. TEM images of the arc melted Zr0.4Ce0.6Fe10Si2 sample: (a) bright field image of the particle with predominantly ThMn12-type phase with (b) corresponding selected area electron diffraction pattern and (c) bright field image of the particle with predominantly bcc-(Fe,Si)-like phase with (d) corresponding selected area electron diffraction pattern. Red rings on the SAED patterns indicate theoretical Bragg positions for both phases.
Figure 6. TEM images of the arc melted Zr0.4Ce0.6Fe10Si2 sample: (a) bright field image of the particle with predominantly ThMn12-type phase with (b) corresponding selected area electron diffraction pattern and (c) bright field image of the particle with predominantly bcc-(Fe,Si)-like phase with (d) corresponding selected area electron diffraction pattern. Red rings on the SAED patterns indicate theoretical Bragg positions for both phases.
Materials 16 01522 g006
Figure 7. TEM images of the Zr0.4Ce0.6Fe10Si2 sample annealed at 1373 K for 72 h: (a) bright field image of the particle with predominantly ThMn12-type phase with (b) corresponding selected area electron diffraction pattern and (c) bright field image of the particle with predominantly bcc-(Fe,Si)-like phase with (d) corresponding selected area electron diffraction pattern. High-resolution images of the ThMn12-like phase and bcc-(Fe,Si)-like phase are shown in (e,f), respectively. Red rings on the SAED patterns indicate theoretical Bragg positions for both phases.
Figure 7. TEM images of the Zr0.4Ce0.6Fe10Si2 sample annealed at 1373 K for 72 h: (a) bright field image of the particle with predominantly ThMn12-type phase with (b) corresponding selected area electron diffraction pattern and (c) bright field image of the particle with predominantly bcc-(Fe,Si)-like phase with (d) corresponding selected area electron diffraction pattern. High-resolution images of the ThMn12-like phase and bcc-(Fe,Si)-like phase are shown in (e,f), respectively. Red rings on the SAED patterns indicate theoretical Bragg positions for both phases.
Materials 16 01522 g007
Table 1. Refined values of the lattice parameters of the two phases identified for Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples after arc-melting and annealing at 1373 K for 72 h: bcc-(Fe,Si) and ThMn12-type structure.
Table 1. Refined values of the lattice parameters of the two phases identified for Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples after arc-melting and annealing at 1373 K for 72 h: bcc-(Fe,Si) and ThMn12-type structure.
Nd Contentx = 0
Arc-Melted
x = 0
Annealed
x = 0.1
Arc-Melted
x = 0.1
Annealed
x = 0.2
Arc-Melted
x = 0.2
Annealed
x = 0.3
Arc-Melted
x = 0.3
Annealed
PhaseLattice Parameter
bcc-(Fe,Si)a = b = c
[±0.001]
2.8622.8632.8622.8632.8602.8612.8642.861
ThMn12a = b
[±0.001]
8.3898.3858.4018.4048.4058.415-8.428
c
[±0.001]
4.7444.7364.7454.7414.7524.742-4.750
Table 2. Refined values of the hyperfine parameters of the phases identified for arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples.
Table 2. Refined values of the hyperfine parameters of the phases identified for arc-melted Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples.
Phase Bhf (T)
[±0.5]
δ (mm/s) [±0.01]2ε (mm/s) [±0.01]Area %
[±2]
Zr0.4Ce0.6Fe10Si2
ThMn12 <28.7><−0.13><0.06>88
bcc-(Fe,Si)bcc-Fe32.3−0.1703
Fe—12% Si 30.4−0.102
Fe—26% Si24.5−0.107
Zr0.3Nd0.1Ce0.6Fe10Si2
ThMn12 <28.7><0.05><0.05>83
bcc-(Fe,Si)bcc-Fe33.1−0.1506
Fe—12% Si 30.8−0.1103
Fe—26% Si24.60.0708
Zr0.2Nd0.2Ce0.6Fe10Si2
ThMn12 + hex <29.9><−0.14><0.0>76
bcc-(Fe,Si)bcc-Fe32.8−0.14011
Fe—12% Si 30.3−0.0907
Fe—26% Si24.10.1806
Zr0.1Nd0.3Ce0.6Fe10Si2
Nd(Zr, Ce)Fe hex <30.6> <−0.14><0.02>66
bcc-(Fe,Si)bcc-Fe32.6−0.12017
Fe—12% Si 30.1−0.1011
Fe—26% Si24.20.306
Table 3. Refined values of the hyperfine parameters of the two phases (bcc-(Fe,Si) and ThMn12-type structure) identified for Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples annealed at 1373 K for 72 h.
Table 3. Refined values of the hyperfine parameters of the two phases (bcc-(Fe,Si) and ThMn12-type structure) identified for Zr0.4−xNdxCe0.6Fe10Si2 (0 ≤ x ≤ 0.3) samples annealed at 1373 K for 72 h.
Phase Bhf (T)
[±0.5]
δ (mm/s) [±0.01]2ε (mm/s) [±0.01]Area %
[±2]
Zr0.4Ce0.6Fe10Si2
ThMn128i24.4−0.060.0923
8j21.9−0.10.0926
8f19.8−0.10.0922
bcc-(Fe,Si)bcc-Fe33.1−0.12−0.035
Fe—4.5% Si 31.3−0.08−0.033
Fe—12% Si29.1−0.16−0.036
Fe—26% Si 24.70.2−0.035
Zr0.3Nd0.1Ce0.6Fe10Si2
ThMn128i24.9−0.10.0620
8j22.5−0.160.0624
8f19.4−0.220.0618
bcc-(Fe,Si)bcc-Fe33−0.180.174
Fe—4.5% Si 32.3−0.080.172
Fe—12% Si30−0.130.179
Fe—26% Si 240.250.179
Zr0.2Nd0.2Ce0.6Fe10Si2
ThMn128i25−0.10.0415
8j22.7−0.150.0421
8f19.5−0.210.0415
bcc-(Fe,Si)bcc-Fe33−0.030.066
Fe—4.5% Si 32.3−0.150.068
Fe—12% Si29.7−0.120.0610
Fe—26% Si 240.270.0610
Zr0.1Nd0.3Ce0.6Fe10Si2
ThMn128i25.2−0.12−0.0315
8j22.9−0.14−0.0318
8f19.4−0.24−0.0312
bcc-(Fe,Si)bcc-Fe33.3−0.06−0.0312
Fe—4.5% Si 32.3−0.13−0.037
Fe—12% Si30−0.07−0.0312
Fe—26% Si 240.32−0.039
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

Kołodziej, M.; Grenèche, J.-M.; Auguste, S.; Idzikowski, B.; Zubko, M.; Bessais, L.; Śniadecki, Z. Influence of Nd Substitution on the Phase Constitution in (Zr,Ce)Fe10Si2 Alloys with the ThMn12 Structure. Materials 2023, 16, 1522. https://doi.org/10.3390/ma16041522

AMA Style

Kołodziej M, Grenèche J-M, Auguste S, Idzikowski B, Zubko M, Bessais L, Śniadecki Z. Influence of Nd Substitution on the Phase Constitution in (Zr,Ce)Fe10Si2 Alloys with the ThMn12 Structure. Materials. 2023; 16(4):1522. https://doi.org/10.3390/ma16041522

Chicago/Turabian Style

Kołodziej, Mieszko, Jean-Marc Grenèche, Sandy Auguste, Bogdan Idzikowski, Maciej Zubko, Lotfi Bessais, and Zbigniew Śniadecki. 2023. "Influence of Nd Substitution on the Phase Constitution in (Zr,Ce)Fe10Si2 Alloys with the ThMn12 Structure" Materials 16, no. 4: 1522. https://doi.org/10.3390/ma16041522

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