Next Article in Journal
Effect of Diffusion Annealing Temperature on the Formation Process and Properties of a Carbon–Aluminum Composite Layer on Pure Titanium
Previous Article in Journal
Laboratory Compaction Study and Mechanical Performance Assessment of Half-Warm Mix Recycled Asphalt Mixtures Containing 100% RAP
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ferromagnetic ε-Fe2MnN: High-Pressure Synthesis, Hardness and Magnetic Properties

1
Max-Planck-Institut für Chemische Physik fester Stoffe, Nöthnitzer Straße 40, 01187 Dresden, Germany
2
School of Materials Science and Engineering, Shanghai University, Shanghai 200444, China
3
Institut für Anorganische Chemie, Universität Stuttgart, Pfaffenwaldring 55, 70569 Stuttgart, Germany
4
Present address Departamento de Química, Facultad de Ciencias, Universidad Católica del Norte, Casa Central. Angamos 0610, 1240000 Antofagasta, Chile
*
Author to whom correspondence should be addressed.
Materials 2019, 12(12), 1993; https://doi.org/10.3390/ma12121993
Submission received: 31 May 2019 / Revised: 17 June 2019 / Accepted: 19 June 2019 / Published: 21 June 2019
(This article belongs to the Section Materials Chemistry)

Abstract

:
The iron manganese nitride Fe2MnN was obtained by high-pressure–high-temperature synthesis from ζ-Fe2N and elemental Mn at 15(2) GPa and 1573(200) K. The phase crystallizes isostructural to binary ε-Fe3N. In comparison to the corresponding binary iron nitride, the microhardness of ε-Fe2MnN is reduced to 6.2(2) GPa. Above about 800 K the ternary compound decomposes exothermally under loss of nitrogen. ε-Fe2MnN is ferromagnetic with a Curie temperature of roughly 402 K.

Graphical Abstract

1. Introduction

During recent years, high-manganese austenitic steels have been regarded an object of fundamental interest for applications in the field of advanced construction materials, because of their exceptional mechanical properties with regard to strain and strength leading to an excellent balance between flow stresses and ductility [1,2,3,4]. These impressive properties originate from a transformation of the metastable austenite to either hcp (hexagonal close packed) or bcc (body centred cubic) martensite under mechanical load and the formation of twins during deformation, both of which can be influenced by the Mn and C content. Thus, the mechanical properties of manganese-rich steels, with relation to chemical compositions, are currently intensely investigated [1,2,3,4,5,6].
For hardening the surfaces of iron and steel equipment, the formation of iron nitrides is of broad technological importance as it improves fatigue endurance. Moreover, nitrides like γ′-Fe4N and ε-Fe3N exhibit an improved corrosion resistance against atmospheres containing oxygen or water [7]. In addition, the magnetic properties of several binary iron nitrides (α″-Fe16N2, γ′-Fe4N and ε-Fe3N) have motivated numerous investigations, both experimental (e.g., [7,8,9,10,11,12,13]) and theoretical (e.g., [13,14,15,16]), for potential applications in high-density data recording. Many efforts were devoted to modifying or improving the magnetic properties by substituting iron with other transition metals, mostly resulting in varieties of cubic γ′-Fe4N [17,18,19,20,21,22].
ε-Fe3N is known to exhibit an exceptionally broad compositional width from about Fe3N0.75 to Fe3N1.5, depending on the temperature [23]. In the idealized crystal structure of ε-Fe3N, nitrogen atoms occupy one third of the octahedral holes of the hcp of iron (basically a ε-Fe-type array) in an ordered arrangement, resulting in exclusively vertex-sharing coordination octahedra around N. All edge- and face-sharing octahedral holes relative to the resulting NFe6/2-framwork remain empty [12,24,25,26]. Figure 1 depicts a section of the idealized fully ordered crystal structure of ε-Fe3N. Excess nitrogen content, up to Fe3N1.5, is realized via the occupation of edge-sharing octahedral voids. However, even for samples with compositions close to the ideal Fe3N some entropy-driven transfer of nitrogen from the ideal position to those additional octahedral holes is observed, depending on temperature and thermal history of the sample [12,24,27].
Reports on substituted ε-Fe3N are comparably rare, partly due to the metastable nature of the products. Nanoparticles of Co-, Ni- and Cr-substituted ε-Fe3N were prepared by ammonolysis, which typically provides samples with degrees of substitution below unity [28,29,30,31,32,33,34,35,36,37,38,39,40]. According to Mössbauer spectroscopy, the cobalt and nickel atoms randomly replace Fe in the nitride. Additionally, superparamagnetic properties have been revealed. However, the magnetic properties of these nanoparticles apparently are highly dominated by the surface states of the spins. Larger particle size ε-Fe3–xCoxN were obtained by a high-pressure metathesis reaction and studied by neutron diffraction [39,40]. ε-Fe3–xMnxN with x below unity was claimed to form in nanoscale particles from iron chloride and ethylendiamine with subsequent annealing in nitrogen [41,42]. We recently developed a high-pressure–high-temperature synthesis strategy to access such metastable materials, which successfully yielded the ternary compounds Fe2CoN, Fe2NiN and Fe2IrN0.24 [43,44]. Here, we present the synthesis of a ternary manganese iron nitride, ε-Fe2MnN, by reacting ζ-Fe2N and Mn at high-pressure–high-temperature conditions before quenching to room temperature.

2. Materials and Methods

The precursor ζ-Fe2N was prepared in microcrystalline form by reacting iron powder (99.9%, Johnson Matthey/Alfa, Ward Hill, MA, USA) with NH3 (99.98%) as described earlier [43]. Subsequent handling was performed in Ar-filled gloveboxes (MBraun, Garching, Germany) with oxygen and water levels below 0.1 ppm. Manganese powders were obtained by grinding bulk metal Mn (99.99%, Alfa, Ward Hill, MA, USA) followed by selecting the fraction which passed a 50 μm sieve. In a typical high-pressure–high-temperature synthesis, a mixture of ζ-Fe2N and Mn powders (ca. 35 mg), in a molar ratio of 1:1.02, was sealed into crucibles machined from hexagonal boron nitride. The slight excess of Mn was utilized for compensating losses due to reactions of Mn with BN at the sample–crucible contact under high-pressure–high-temperature conditions. The crucibles were placed in graphite sleeves enclosed by zirconia parts for thermal insulation. These assemblies were transferred to MgO/Cr2O3 octahedra with an edge length of 14 mm and the graphite tubes were furnished with circular molybdenum disks for electrical contact. The octahedra were compressed to a pressure of 15(2) GPa and heated to 1573(200) K. After maintaining the desired temperature for 30 minutes, the samples were quenched to room temperature by disconnecting the electrical current. Finally, the pressure was released to ambient pressure within 13.5 h. The obtained product shows a metallic luster and conchoidal fracture.
The high-pressure–high-temperature samples were analyzed by powder X-ray diffraction for phase identification and crystal structure determination by using a Huber G-670 diffractometer (HUBER Diffraktionstechnik, Rimsting, Germany) (CoKα1 radiation, λ = 1.788965 Å) in the 2θ range from 3° to 100° with a step size of 0.005°. The lattice parameters were determined by using the software package WinCSD, with Si powder as an internal standard [45]. Homogeneity and composition of polished samples on a micrometer scale were examined by metallographic measurements using optical microscopes (Zeiss Axioplan 2) (Jena, Germany) combined with energy dispersive X-ray spectroscopy (Philips XL30 scanning electron microscope with a Phoenix V. 5.29 analytic unit, EDXS). Simultaneous thermogravimetry and differential thermal analysis (TG/DTA) were performed by a Netzsch STA449C instrument (Netzsch Gerätebau, Selb, Germany) in the temperature range from 300 to 1270 K, with a heating rate of 10 K min−1 in an argon atmosphere. Microhardness was determined on a metallographically prepared microstructure with a Vickers indenter (MHT 10, Anton Paar, Graz, Austria) (manufacturer, city, country), which was attached to an optical microscope (Zeiss Axioplan 2). Magnetization data were recorded by a SQUID magnetometer (MPMS XL-7, Quantum Design, San Diego, CA, USA) in the temperature range between 1.8 and 600 K and magnetic fields up to 7 T. The sample used for establishing magnetic properties of Fe2MnN was a polycrystalline pellet.

3. Results and Discussion

The obtained product exhibits pronounced similarities to the ε-type binary and ternary iron nitride samples earlier obtained by this technique [24,43,44]. The material is stable against moist air without any visible changes for at least several days.
Powder X-ray diffraction measurements reveal the formation of a ε-type product (Figure 2) with pronounced disorder of the nitrogen atoms (Table 1 and Table 2 as well as Figure 1). Overall, the values for the lattice parameters a = 4.71872(5) Å and c = 4.41982(7) Å are close to those of binary ε-Fe3N1.08 (a = 4.7241(2) Å, c = 4.3862(2) Å [24]), which is consistent with the negligible difference in the atomic radii of manganese and iron. Scatter of the lattice parameters for different samples and metallographic analysis evidences a significant homogeneity range of the ε-phase—a sample with 5% higher nitrogen content according to chemical analysis, Fe2MnN1.05, yielded lattice parameters a = 4.7344(4) Å and c = 4.4264(5) Å. This elongation of the unit cell parameters by about 0.33% in a and 0.15% in c upon increase of the nitrogen content by 5% is close to the composition dependent changes in the unit cell observed for pure ε-Fe3N1+x [46]. Full profile Rietveld refinements in space group P6322 (No. 182) were carried out on a pattern obtained from a sample with nominal composition Fe2MnN and reveal nearly equal distribution of nitrogen over the two accessible octahedral voids (avoiding face-sharing with the further present octahedral voids) within the hcp of metal atoms, with only little preference for the site occupied in the idealized arrangement assigned to ε-type Fe3N (occupation factor of about 60%, see Table 2). This finding is attributed to the high synthesis temperature followed by rapid cooling during synthesis. A possible (partial) order of iron and manganese within the hcp motif of metal atoms cannot be accessed with standard powder X-ray diffraction techniques due to the close diffraction power of both elements. Fe–Mn-disorder, as observed for nanoparticles, is also analogously present in bulk ε-Fe2MN (M = Co, Ni) according to Mössbauer spectroscopy and neutron diffraction [33,39], and may assist nitrogen transfer between accessible sites.
Figure 3 shows elemental mapping obtained on a polished metallographic surface of an as-cast ε-Fe2MnN sample. It reveals a homogeneous distribution of iron, manganese and nitrogen, with the atomic metal ratio of n(Fe):n(Mn) = 2.06(1):0.94(1) determined from the average of three independent EDXS measurements within the central part of the polished sample. Before characterizing thermal and magnetic properties, the sample and crucible were separated and residual boron nitride, as well as the inhomogeneous surface and small inclusions, were carefully removed.
Thermal analysis at ambient pressure (Figure 4) shows the onset of an endothermal signal around 900 K, assigned to the decomposition of the sample into a (Fe,Mn):N alloy as indicated by the associated mass loss. The total release amounts to 7.32% according to the thermal analysis, which is in fair agreement with the calculated nitrogen content of 7.75 wt% in ε-Fe2MnN with ideal composition. However, for a valid calculation of the weight loss, the solubility of N in (Fe,Mn) alloy has to be taken into account [23]. Interestingly, ε-Fe2MnN appears thermally stable in somewhat higher temperatures than the respective binary iron nitride and its Ni- and Co-containing analogues (compare Table 3).
Figure 5 shows a typical micrograph of the Vickers indentation for a metallographically treated sample. The microhardness measurements at six independent indentations yield the average value of 636 ± 21 HV, which corresponds to 6.2(2) GPa. Due to the brittle nature of the material, some cracking is typically observed at the corners of the indentation, as can be seen in Figure 5. Thus, the derived hardness value is a minimum border. However, the finding shows that the hardness of the iron manganese nitride sample is slightly below that of ε-Fe3Nx (7.4(10) GPa) [24], but is still in the range of tool steels (up to 9 GPa) and other nitrogen-hardened steels (∼5‒15 GPa) [47,48,49,50].
In an earlier investigation, the magnetic properties of ε-Fe2CoN and ε-Fe2NiN were characterized by experimental data combined with spin-polarized DFT calculations [43]. The loss of ferromagnetic moments in comparison to ε-Fe3N was attributed to the additional valence electrons of Co and Ni, which remain essentially localized in the minority-spin 3d channels of the transition metal atoms. Considering that Mn has one valence electron less than Fe and assuming pure ferromagnetic coupling, the total magnetic moment of ε-Fe2MnN would simply correspond to the sum of the magnetic moments of Fe and Mn, i.e., a larger saturation magnetization would be expected for ε-Fe2MnN than for ε-Fe3N.
Figure 6 shows the magnetization of ε-Fe2MnN as a function of temperature. On cooling, a magnetic phase transition from paramagnetism to a ferromagnetically ordered state is revealed. The Curie temperature TC = 402 ± 5 K is determined from the critical behavior of the low-field magnetization. Applying the Curie–Weiss law χ = CCW/(T – θp) to the 3.5 T data above 550 K results in a high paramagnetic Curie temperature θp = +448 K and a Curie constant CCW corresponding to an effective paramagnetic moment µeff/f.u. = 5.0 µB. The rather rounded shape of the transition may be indicative of the presence of micrograins. The inverse susceptibility and the fit is shown in the bottom inset of Figure 6. At low temperatures, soft magnetic behavior is observed (Figure 6, upper inset). The extrapolated spontaneous magnetic moment Msp amounts to 3.88 µB. The magnetic behavior of ε-Fe2MnN is well described as an itinerant d-electron system. The ordered moment is significantly lower than the maximum moment expected from the Slater–Pauling curve [51,52], which predicts an average magnetic moment of ≈1.8 µB per d-metal atom for the (3d + 4s) electron count 7.67. Table 3 compares selected magnetic characteristics found on ε-type compounds ε-Fe2MN (M = Mn, Fe, Co, Ni) obtained by the applied high-pressure–high-temperature synthesis.

4. Summary

In summary, ε-Fe2MnN has been synthesized under high-pressure–high-temperature conditions of 15(2) GPa and 1573(200) K. Upon heating without application of external pressure, the phase decomposes under loss of nitrogen. The substitution of iron by manganese causes a decrease of the magnetic moment to 3.88 µB and the Curie temperature to 402 K. The compound exhibits a microhardness of about 6.2(2) GPa, which is below that of ε-Fe3N, but still in the range of values for nitrided steels. Information on phases in this ternary system may aid the development of high-manganese austenitic steels, particularly in surface-hardened materials manufactured from these. Although such phases may appear during surface treatment, little is known on the formation and stability of the corresponding relevant nitrides.

Author Contributions

Conceptualization of the study was done by U.S. and R.N.; synthesis and sample characterization by K.G., W.C. and R.C.; magnetization measurements by M.B; metallography by U.B.; refinement of X-ray diffraction data by L.A.; data curation, U.S.; writing—original draft preparation, U.S., R.N., K.G. and W.C.; writing—review and editing, R.N.; visualization, R.N. and U.S.; supervision, U.S. and R.N.

Funding

This research received additional funding from the CAS/SAFEA International Partnership Program for Creative Research Teams (51121064) and the priority program “Synthesis, in situ characterization, and quantum mechanical modelling of Earth Materials, oxides, carbides and nitrides at extremely high pressures and temperatures” of the Deutsche Forschungsgemeinschaft (DFG) (SPP 1236). K.G. acknowledges support from the Foundation of China under Project No. 21501118, 111 project (D16002) and the Young Eastern Scholar Project of Shanghai Municipal Education Commission (QD2015031).

Acknowledgments

The authors would like to thank Philipp Marasas, Julia Hübner, Liudmila Muzica and Susanne Leipe for assistance with the high-pressure–high-temperature experiments as well as Marcus Schmidt and Susann Scharsach for the TG/DTA measurements. We gratefully acknowledge their support. Helpful discussions with Walter Schnelle and Yuri Grin are also gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Allain, S.; Chateau, J.-P.; Bouaziz, O.; Migot, S.; Guelton, N. Correlations between the calculated stacking fault energy and the plasticity mechanisms in Fe-Mn-C alloys. Mater. Sci. Eng. A 2004, 387–389, 158–162. [Google Scholar] [CrossRef]
  2. Scott, C.; Allain, S.; Faral, M.; Guelton, N. The development of a new Fe-Mn-C austenitic steel for automotive applications. Rev. Metall. Int. J. Metall. 2006, 103, 293–302. [Google Scholar] [CrossRef]
  3. Jin, J.-E.; Lee, Y.-K. Strain hardening behavior of a Fe-18Mn-0.6C-1.5Al TWIP steel. Mater. Sci. Eng. A 2009, 527, 157–161. [Google Scholar] [CrossRef]
  4. Saeed-Akbari, A.; Imlau, J.; Prahl, U.; Bleck, W. Derivation and Variation in Composition-Dependent Stacking Fault Energy Maps Based on Subregular Solution Model in High-Manganese Steels. Metall. Mater. Trans. A 2009, 40A, 3076–3090. [Google Scholar] [CrossRef]
  5. Grässel, O.; Krüger, L.; Frommeyer, G.; Meyer, L.W. High strength Fe-Mn-(Al, Si) TRIP/TWIP steels development-properties-application. Int. J. Plast. 2000, 16, 1391–1409. [Google Scholar] [CrossRef]
  6. Hamada, A.S.; Karjalainen, L.P.; Somani, M.C. The influence of aluminum on hot deformation behavior and tensile properties of high-Mn TWIP steels. Mater. Sci. Eng. A 2007, 467, 114–124. [Google Scholar] [CrossRef]
  7. Spies, H.J.; Liedtke, D. Wärmebehandlung von Eisenwerkstoffen II: Nitrieren und Nitrocarburieren; Kontakt & Studium Band 686; Expert Verlag: Tübingen, Germany, 2014; ISBN 978-3-8169-3282-6. [Google Scholar]
  8. Komuro, M.; Kozono, Y.; Hanazono, M.; Sugita, Y. Epitaxial growth and magnetic properties of Fe16N2 films with high saturation magnetic-flux density. J. Appl. Phys. 1990, 67, 5126–5130. [Google Scholar] [CrossRef]
  9. Coey, J.M.D.; Smith, P.A.I. Magnetic nitrides. J. Magn. Magn. Mater. 1999, 200, 405–424. [Google Scholar] [CrossRef]
  10. Frazer, B.C. Magnetic structure of Fe4N. Phys. Rev. 1958, 112, 751–754. [Google Scholar] [CrossRef]
  11. 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]
  12. Leineweber, A.; Jacobs, H.; Hüning, F.; Lueken, H.; Kockelmann, W. Nitrogen ordering and ferromagnetic properties of ε-Fe3N1+x (0.10 ≤ x ≤ 0.39) and ε-Fe3(N0.80C0.20)1.38. J. Alloys Compd. 2001, 316, 21–38. [Google Scholar] [CrossRef]
  13. Widenmeyer, M.; Shlyk, L.; Senyshyn, A.; Mönig, R.; Niewa, R. Structural and Magnetic Characterization of Single-phase Sponge-like Bulk α″-Fe16N2. Z. Anorg. Allg. Chem. 2015, 641, 348–354. [Google Scholar] [CrossRef]
  14. Peltzer y Blancá, E.L.; Desimoni, J.; Christensen, N.E.; Emmerich, H.; Cottenier, S. The magnetization of γ’-Fe4N: Theory vs. experiment. Phys. Status Solidi B 2009, 246, 909–928. [Google Scholar] [CrossRef]
  15. Sakuma, A. Self-consistent calculations for the electronic structures of iron nitrides, Fe3N, Fe4N and Fe16N2. J. Magn. Magn. Mater. 1991, 102, 127–134. [Google Scholar] [CrossRef]
  16. Tanaka, H.; Harima, H.; Yamamoto, T.; Katayama-Yoshida, H.; Nakata, Y.; Hirotsu, Y. Electronic band structure and magnetism of Fe16N2 calculated by the FLAPW method. Phys. Rev. B 2000, 62, 15042–15046. [Google Scholar] [CrossRef]
  17. Matar, S.; Siberchicot, B.; Pénicaud, M.; Demazeau, G. The electronic and magnetic properties of Fe3N. J. Phys. I Fr. 1992, 2, 1819–1831. [Google Scholar] [CrossRef]
  18. Xue, D.; Li, F.; Yang, J.; Kong, Y.; Gao, M. Effects of substitutional atoms on the properties of γ’-(Fe1-xTMx)4N (TM = Co, Ni) compounds. J. Magn. Magn. Mater. 1997, 172, 165–172. [Google Scholar] [CrossRef]
  19. Kuhnen, C.A.; de Figueiredo, R.S.; dos Santos, A.V. Mössbauer spectroscopy, crystallographic, magnetic and electronic structure of ZnFe3N and InFe3N. J. Magn. Magn. Mater. 2000, 219, 58–68. [Google Scholar] [CrossRef]
  20. Houben, A.; Müller, P.; von Appen, J.; Lueken, H.; Niewa, R.; Dronskowski, R. Synthesis, crystal structure, and magnetic properties of the semihard itinerant ferromagnet RhFe3N. Angew. Chem. Int. Ed. 2005, 44, 7212–7215. [Google Scholar] [CrossRef]
  21. Houben, A.; Burghaus, J.; Dronskowski, R. The Ternary Nitrides GaFe3N and AlFe3N: Improved Synthesis and Magnetic Properties. Chem. Mater. 2009, 21, 4332–4338. [Google Scholar] [CrossRef]
  22. Burghaus, J.; Kleemann, J.; Dronskowski, R. The Ternary Nitrides InxFe4-xN (0 ≤ x ≤ 0.8): Synthesis, Magnetic Properties, and Theoretical Considerations. Z. Anorg. Allg. Chem. 2011, 637, 935–939. [Google Scholar] [CrossRef]
  23. Wriedt, H.A.; Gokcen, N.A.; Nafziger, R.H. The Fe-N (Iron-Nitrogen) System. Bull. Alloy Phase Diagr. 1987, 8, 355–377. [Google Scholar] [CrossRef]
  24. Niewa, R.; Rau, D.; Wosylus, A.; Meier, K.; Hanfland, M.; Wessel, M.; Dronskowski, R.; Dzivenko, D.A.; Riedel, R.; Schwarz, U. High-Pressure, High-Temperature Single-Crystal Growth, Ab initio Electronic Structure Calculations, and Equation of State of epsilon-Fe3N1+x. Chem. Mater. 2009, 21, 392–398. [Google Scholar] [CrossRef]
  25. Jack, K.H. The Iron–Nitrogen System: The Crystal Structures of ε-Phase Iron Nitrides. Acta Crystallogr. 1952, 5, 404–411. [Google Scholar] [CrossRef]
  26. Jacobs, H.; Rechenbach, D.; Zachwieja, U. Structure determination of γ’-Fe4N and ε-Fe3N. J. Alloys Compd. 1995, 227, 10–17. [Google Scholar] [CrossRef]
  27. Leineweber, A.; Jacobs, H.; Hüning, F.; Lueken, H.; Schilder, H.; Kockelmann, W. ε-Fe3N: Magnetic structure, magnetization and temperature dependent disorder of nitrogen. J. Alloys Compd. 1999, 288, 79–87. [Google Scholar] [CrossRef]
  28. Gajbhiye, N.S.; Ningthoujam, R.S.; Bhattacharyya, S. Magnetic properties of Co and Ni substituted ε-Fe3N nanoparticles. Hyperfine Interact. 2005, 164, 17–26. [Google Scholar] [CrossRef]
  29. Gajbhiye, N.S.; Bhattacharyya, S. Spin-glass-like ordering in ε-Fe3-xNixN (0.1 ≤ x ≤ 0.8) nanoparticles. Mater. Chem. Phys. 2008, 108, 201–207. [Google Scholar] [CrossRef]
  30. Gajbhiye, N.S.; Bhattacharyya, S.; Sharma, S. Observation of exchange bias and spin-glass-like ordering in ε-Fe2.8Cr0.2N nanoparticles. PRAMANA J. Phys. 2008, 70, 367–373. [Google Scholar] [CrossRef]
  31. Mishra, P.P.; Raja, M.M.; Panda, R.N. Enhancement of magnetic moment in Co substituted nanocrystalline ε-CoxFe3–xN (0.2 ≤ x ≤ 0.4) synthesized by modified citrate precursor route. Mater. Res. Bull. 2016, 75, 127–133. [Google Scholar] [CrossRef]
  32. Gajbhiye, N.S.; Bhattacharyya, S. Mössbauer and magnetic studies for the coexistence of ε-Fe3–xNixN and γ’-Fe4–yNiyN phases in Fe-Ni-N nanoparticles. Indian J. Pure Appl. Phys. 2007, 45, 834–838. [Google Scholar]
  33. Gajbhiye, N.S.; Ningthoujam, R.S.; Weissmuller, J. Mössbauer Study of Nanocrystalline ε-Fe3–xCoxN System. Hyperfine Interact. 2004, 156, 51–56. [Google Scholar] [CrossRef]
  34. Gajbhiye, N.S.; Bhattacharyya, S. Mössbauer studies of ε-Fe3–xNixN and γ’-Fe4–yNiyN nanoparticles. Hyperfine Interact. 2005, 165, 147–151. [Google Scholar] [CrossRef]
  35. Gajbhiye, N.S.; Bhattacharyya, S. Synthesis and structural investigation of ε-Fe3–xNixN (0.0 ≤ x ≤ 0.8) nanoparticles. Prog. Cryst. Growth Charact. 2006, 52, 132–141. [Google Scholar] [CrossRef]
  36. Ningthoujam, R.S.; Gajbhiye, N.S. Magnetization studies on ε-Fe2.4Co0.6N nanoparticles. Mater. Res. Bull. 2010, 45, 499–504. [Google Scholar] [CrossRef]
  37. Zhao, Y.S.; Wang, M.; Ma, Y.Q. Effects of nitriding temperature on the structure and magnetic properties of CoFe2 alloy. J. Mater. Sci. 2018, 29, 20071–20080. [Google Scholar] [CrossRef]
  38. Lei, X.; Zhang, P.; Wang, X.; Wang, W.; Yang, H. (Fe1–xNix)3N nanoparticles: The structure, magnetic and photocatalytic properties for water splitting. RSC Adv. 2016, 6, 44641–44645. [Google Scholar] [CrossRef]
  39. Lei, L.; Zhang, L.; Gao, S.; Hu, Q.; Fang, L.; Chen, X.; Xia, Y.; Wang, X.; Ohfuji, H.; Kojima, Y.; et al. Neutron diffraction study of structural and magnetic properties of ε-Fe3N1.098 and ε-Fe2.322Co0.678N0.888. J. Alloys Compd. 2018, 752, 99–105. [Google Scholar] [CrossRef]
  40. Zhang, L.; Gao, S.; Hu, Q.; Hu, Q.; Qi, L.; Feng, L.; Lei, L. Synthesis and characterization of spherical-bulk ε-Fe3–xCoxN (x = 0.0, 0.25, 1.95). Mater. Chem. Phys. 2017, 197, 94–99. [Google Scholar] [CrossRef]
  41. Lei, X.; Ye, Z.; Qie, Y.; Fan, Z.; Chen, X.; Shi, Z.; Yang, H. The synthesis, morphology and magnetic properties of (Fe1–xMnx)3N nanoparticles. J. Mater. Sci. 2019, 30, 277–283. [Google Scholar] [CrossRef]
  42. Lei, X.; Pan, X.; Ye, Z.; Yang, X.; Chen, X.; Shi, Z.; Yang, H. Synthesis, Structure and Properties Comparison of Fe3N Doped with Ni, Mn and Co. ChemistrySelect 2019, 4, 5945–5949. [Google Scholar] [CrossRef]
  43. Guo, K.; Rau, D.; Toffoletti, L.; Müller, C.; Burkhardt, U.; Schnelle, W.; Niewa, R.; Schwarz, U. Ternary Metastable Nitrides ε-Fe2TMN (TM = Co, Ni): High-Pressure, High-Temperature Synthesis, Crystal Structure, Thermal Stability, and Magnetic Properties. Chem. Mater. 2012, 24, 4600–4606. [Google Scholar] [CrossRef]
  44. Guo, K.; Rau, D.; Schnelle, W.; Burkhardt, U.; Niewa, R.; Schwarz, U. High-Pressure High-Temperature Synthesis of ε-Fe2IrN0.24. Z. Anorg. Allg. Chem. 2014, 640, 814–818. [Google Scholar] [CrossRef]
  45. Akselrud, L.; Grin, Y. WinCSD: Software package for crystallographic calculations (Version 4). J. Appl. Crystallogr. 2014, 47, 803–805. [Google Scholar] [CrossRef]
  46. Liapina, T.; Leineweber, A.; Mittemeijer, E.J.; Kockelmann, W. The lattice parameters of ε-iron nitrides: Lattice strains due to a varying degree of nitrogen ordering. Acta Mater. 2004, 52, 173–180. [Google Scholar] [CrossRef]
  47. Brandes, E.A.; Brook, G.B. Smithells Metals Reference Book, 7th ed.; Butterworth-Heinemann: Oxford, UK, 1997; ISBN 9780080517308. [Google Scholar]
  48. Oberg, E.; Jones, F.D.; Horton, H.L.; Ryffel, H.H. Machinery’s Handbook, 27th ed.; Industrial Press: New York, NY, USA, 2004. [Google Scholar]
  49. Weber, T.; de Wit, L.; Saris, F.W.; Königer, A.; Rauschenbach, B.; Wolff, G.K.; Krauss, S. Hardness and corrosion resistance of single-phase nitride and carbide on iron. Mater. Sci. Eng. 1995, A199, 205–210. [Google Scholar] [CrossRef]
  50. Guemmaz, M.; Mosser, A.; Grob, J.J.; Stuck, R. Sub-surface modifications induced by nitrogen ion implantation in stainless steel (SS316L). Correlation between microstructure and nanoindentation results. Surf. Coat. Technol. 1998, 100, 353–357. [Google Scholar] [CrossRef]
  51. Pauling, L. The nature of the interatomic forces in metals. Phys. Rev. 1938, 54, 899–904. [Google Scholar] [CrossRef]
  52. Slater, J.C. Electronic structure of alloys. J. Appl. Phys. 1937, 8, 385–390. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of ε-Fe2MnN. Grey spheres represent Fe/Mn, green and blue spheres represent nitrogen. Exclusive occupation of the green octahedra (site 2c) results in the ideal ordered crystal structure of ε-Fe3N. Entropy-driven nitrogen transfer from site 2c to partially occupied 2b (blue octahedral) is typically observed.
Figure 1. Crystal structure of ε-Fe2MnN. Grey spheres represent Fe/Mn, green and blue spheres represent nitrogen. Exclusive occupation of the green octahedra (site 2c) results in the ideal ordered crystal structure of ε-Fe3N. Entropy-driven nitrogen transfer from site 2c to partially occupied 2b (blue octahedral) is typically observed.
Materials 12 01993 g001
Figure 2. Powder X-ray diffraction pattern (CoKα1 radiation) of ε-Fe2MnN. Black symbols represent measured intensities, the dark cyan line represents calculated intensities, and the light gray line below the diagram shows the difference. Intensities are given on a linear scale.
Figure 2. Powder X-ray diffraction pattern (CoKα1 radiation) of ε-Fe2MnN. Black symbols represent measured intensities, the dark cyan line represents calculated intensities, and the light gray line below the diagram shows the difference. Intensities are given on a linear scale.
Materials 12 01993 g002
Figure 3. Element mappings of a polished ingot for (a) iron (pink), (b) manganese (yellow) and (c) nitrogen (blue). The metallographic specimen of the target phase Fe2MnN exhibits homogeneous color distributions and, thus, chemical composition.
Figure 3. Element mappings of a polished ingot for (a) iron (pink), (b) manganese (yellow) and (c) nitrogen (blue). The metallographic specimen of the target phase Fe2MnN exhibits homogeneous color distributions and, thus, chemical composition.
Materials 12 01993 g003
Figure 4. TG-DTA data of ε-Fe2MnN. The endothermic effect at about 900 K (onset) is attributed to the decomposition into a (Fe, Mn):N alloy. The mass loss of 7.32 wt% up to 1270 K is in accordance with the calculated nitrogen content in ε-Fe2MnN (7.75 wt%).
Figure 4. TG-DTA data of ε-Fe2MnN. The endothermic effect at about 900 K (onset) is attributed to the decomposition into a (Fe, Mn):N alloy. The mass loss of 7.32 wt% up to 1270 K is in accordance with the calculated nitrogen content in ε-Fe2MnN (7.75 wt%).
Materials 12 01993 g004
Figure 5. Indentation in a polished sample of ε-Fe2MnN for the determination of Vickers microhardness (optical micrograph, bright field contrast).
Figure 5. Indentation in a polished sample of ε-Fe2MnN for the determination of Vickers microhardness (optical micrograph, bright field contrast).
Materials 12 01993 g005
Figure 6. Main panel: Temperature dependence of the magnetization of Fe2MnN measured in high (3.5 T) and low (0.1 T) magnetic fields. Upper inset: Field dependence of the magnetization at T = 1.8 K. Lower inset: Inverse high-temperature susceptibility measured at 3.5 T. The black line shows a linear fit of the data above 550 K.
Figure 6. Main panel: Temperature dependence of the magnetization of Fe2MnN measured in high (3.5 T) and low (0.1 T) magnetic fields. Upper inset: Field dependence of the magnetization at T = 1.8 K. Lower inset: Inverse high-temperature susceptibility measured at 3.5 T. The black line shows a linear fit of the data above 550 K.
Materials 12 01993 g006
Table 1. Data collection 1, structure refinement and crystallographic information for Fe2MnN. A second sample with a 5% excess of nitrogen according to chemical analysis resulted in lattice parameters a = 4.7344(4) Å and c = 4.4264(5) Å.
Table 1. Data collection 1, structure refinement and crystallographic information for Fe2MnN. A second sample with a 5% excess of nitrogen according to chemical analysis resulted in lattice parameters a = 4.7344(4) Å and c = 4.4264(5) Å.
CompositionFe2MnN
Space groupP6322 (No. 182)
Structure typeε-Fe3N
Unit cell
a4.71875(5)
c4.41981(7)
Volume/Å385.229(3)
Formula units, Z2
Radiation 1CoKα1
Measurement range, deg.3.5 ≤ 2θ ≤ 95.465
Measured points18394
Measured reflections17
Refined parameters8
R(P)/R(wP)0.034/0.0056
1 λ = 1.788965 Å.
Table 2. Wyckoff positions, site occupancy factors (SOF), relative atomic coordinates and isotropic displacement parameters Biso (in 104 pm2).
Table 2. Wyckoff positions, site occupancy factors (SOF), relative atomic coordinates and isotropic displacement parameters Biso (in 104 pm2).
AtomSiteSOF 1xyzBiso
Fe, Mn6g2/3, 1/30.3413(3)01/21.0(2)
N2c0.571(8)2/31/33/41
N2b0.429001/41
1 The sums of the SOFs of Fe and Mn, as well as that of nitrogen occupying different positions, were constrained to 1.
Table 3. Magnetic moments, Curie temperatures TC, unit cell parameters and thermal decomposition temperatures TD observed for ε-type nitrogen compounds Fe2MN (M = Mn, Fe, Co, Ni). Any comparison of the unit cell parameters should take into account slight variations of Fe/M ratios and particularly the nitrogen content of the samples.
Table 3. Magnetic moments, Curie temperatures TC, unit cell parameters and thermal decomposition temperatures TD observed for ε-type nitrogen compounds Fe2MN (M = Mn, Fe, Co, Ni). Any comparison of the unit cell parameters should take into account slight variations of Fe/M ratios and particularly the nitrogen content of the samples.
MMsp/µBTC/KacTD
Mn3.884024.71875(5)4.41981(7)900this work
Fe6.05754.6982(3)4.3789(4)750[24,26]
Co4.34884.6759(5)4.3776(5)750[43]
Ni3.12344.6698(4)4.3699(4)750[43]

Share and Cite

MDPI and ACS Style

Schwarz, U.; Guo, K.; Clark, W.P.; Burkhardt, U.; Bobnar, M.; Castillo, R.; Akselrud, L.; Niewa, R. Ferromagnetic ε-Fe2MnN: High-Pressure Synthesis, Hardness and Magnetic Properties. Materials 2019, 12, 1993. https://doi.org/10.3390/ma12121993

AMA Style

Schwarz U, Guo K, Clark WP, Burkhardt U, Bobnar M, Castillo R, Akselrud L, Niewa R. Ferromagnetic ε-Fe2MnN: High-Pressure Synthesis, Hardness and Magnetic Properties. Materials. 2019; 12(12):1993. https://doi.org/10.3390/ma12121993

Chicago/Turabian Style

Schwarz, Ulrich, Kai Guo, William P. Clark, Ulrich Burkhardt, Matej Bobnar, Rodrigo Castillo, Lev Akselrud, and Rainer Niewa. 2019. "Ferromagnetic ε-Fe2MnN: High-Pressure Synthesis, Hardness and Magnetic Properties" Materials 12, no. 12: 1993. https://doi.org/10.3390/ma12121993

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