Zero thermal expansion with high Curie temperature in Ho2Fe16Cr alloy

Shovan Dana, S. Mukherjee*a, Chandan Mazumdarb and R. Ranganathanb
aDepartment of Physics, The University of Burdwan, Burdwan-713104, India. E-mail: sanseb68@yahoo.co.in
bCondensed Matter Physics Division, Saha Institute of Nuclear Physics, 1/AF, Bidhannagar, Kolkata-700064, India

Received 10th August 2016 , Accepted 23rd September 2016

First published on 23rd September 2016


Abstract

We report the observation of zero thermal expansion with a high Curie temperature in a Ho2Fe16Cr alloy. Among the R2Fe17−xCrx (R = rare earth elements) series of alloys, Ho2Fe16Cr shows not only enhancement of the Curie temperature (TC) to 415 K in comparison with the parent compound Ho2Fe17 (330 K), but also shows zero thermal expansion (ZTE) in the wide temperature range 13–330 K due to reduction of the magneto-volume effect. We believe that such a single component ZTE magnetic material with a TC higher than room temperature is of practical importance.


1 Introduction

Recently, ZTE materials with a very low coefficient of thermal expansion (α) have attracted considerable interest as a particular class of functional materials.1–6 These materials appear to be of immense practical importance to improve the performance of various electronic and optical devices. Several materials: YbGaGe,1 Fe[Co(CN)6],2 In(HfMg)0.5Mo3O12,3 0.8PbTiO3–0.2Bi(Ni0.5Ti0.5)O3,4 antiperovskite manganese nitride Mn3AN (A = Cu/Sn, Zn/Sn),5 (Al0.3(HfMg)0.85)(WO4)3,6 Invar alloys,7,8 Sc0.05Ga0.05Fe0.1F3,9 La(Fe,Si)13,10 La(Fe,Al)13,11 carbon doped La(Fe,Si)13 (ref. 12) etc. have been identified as ZTE materials, in different ranges of temperature.

Materials showing negative thermal expansion (NTE) play a key role in producing ZTE materials.13,14 In general, to achieve ZTE materials, attempts have been made to form composites by dispersing particles of a NTE material like ZrW2O8 into an isotropic matrix of positive thermal expansion (PTE) material like copper (Cu) or aluminium (Al).13,15,16 ZTE composites have also been synthesized using antiperovskite NTE materials like Mn3Cu0.5A0.5N (A = Ni, Sn),17 Mn3Zn0.5Sn0.5N18 and Cu by optimizing their mass ratios. LaFe10.5Co1.0Si1.5/Cu19 has been recently found to show a tailoring thermal expansion property. In such composites, the NTE material is used to compensate for the thermal expansion of the matrix of PTE material to obtain a desired low α. However, the actual values of α do not match the desired values, as the effects of the interfaces could not be estimated in advance accurately.13 Among the composites, α as low as 3.5 × 10−6 K−1 in the range of T = 320–355 K, has been obtained in the case of a metal matrix composite Mn3Cu0.5Sn0.5N1−δ/AC8A.13 In composites, the large difference in α between the matrix and the compensator causes a stress at the interfaces or grain boundaries, and thereby, may degrade the composite functionalities by forming micro-cracks.13 Therefore, as a functional material, a single-component ZTE material13 will be more effective than a composite one.

A single component ZTE material can be formed as a solid solution of a PTE material and a NTE material. In(HfMg)0.5Mo3O12[thin space (1/6-em)]3 is such a solid solution of a PTE material HfMgMo3O12 and a NTE material In2Mo3O12. It shows ZTE in the temperature range 500–900 K with an average linear intrinsic α = −0.4 × 10−6 K−1, and an average bulk α = 0.4 × 10−6 K−1. A similar single-phase ceramic material, (Al2x(HfMg)1−x) (WO4)3,6 formed by combining a NTE material (HfMg)(WO4)3 and a PTE material Al2(WO4)3, behaves like a ZTE material for x = 0.15, between room temperature and 800 °C.

A single component ZTE material can also be achieved by lowering the NTE coefficient of a NTE material, by weakening the inherent mechanism responsible for NTE. The underlying mechanism responsible for the observed NTE in different reported materials is not unique; rather, they are widely different.13,20 For example, a flexible network causes NTE in ZrW2O8,21 LiAlSiO4,22 Cd(CN)2,23 ReO3,24 siliceous faujasite,25 metal nitroprussides,26 HfScMo2VO12 (ref. 27) etc. whereas atomic radius contraction is responsible for NTE in Bi0.95La0.05NiO3.28 The site anisotropy results in NTE in GdPd3B0.25C0.75,29 and tetragonality introduces NTE in PbTiO3.4 The magneto-volume effect (MVE) causes NTE in different magnetic materials like Invar alloys,30,31 YMn2,32 pure33 and substituted R2Fe17 (R = rare earth element) compounds,34–36 manganese antiperovskites.37 A ZTE multiferroic compound 0.8PbTiO3–0.2Bi(Ni0.5Ti0.5)O3 (ref. 4) with α = 0.4 × 10−7 K−1 between room temperature and 500 °C has been achieved by weakening the tetragonality of the parent compound PbTiO3, by using the dopant Bi(Ni0.5Ti0.5)O3. A similar study has also been done on Nd/Sm substituted 0.5PbTiO3–0.5BiFeO3.38 Optimization of the heat treatment and the chemical composition lowers the NTE coefficient of antiperovskite Mn3AN(A = Cu/Sn, Zn/Sn).5 ZTE can be achieved by weakening the MVE in magnetic materials showing NTE. R2Fe17 compounds are such magnetic materials, where on increasing the temperature, the lattice parameter c decreases faster than the lattice parameter a due to MVE, and thereby, results in NTE.

R2Fe17 compounds with a hexagonal Th2Ni17 structure for heavy rare earths (rhombohedral Th2Zn17 structure for light rare earths) show NTE due to MVE. Cr-substitution in R2Fe17 compounds can weaken the NTE behavior as well as increase TC.34–36 Thus, the study of Cr-substitution in R2Fe17 compounds is important in searching for ZTE magnetic materials with a high TC, although previous studies in Cr-substituted R2Fe17 compounds (R = Tb, Tm, Gd) do not show the existence of ZTE in their studied range of temperature. Such studies are restricted only in a temperature range above room temperature. Tm2Fe17−xCrx (x > 0) shows NTE near the TC (around 430 K) with a minimum NTE coefficient of volume expansion, αν = −9.15 × 10−6 K−1 for the x = 0.5 compound.36 Tb2Fe16Cr possesses a minimum value of αν = −5.28 × 10−6 K−1 in the temperature range 292–511 K,35 whereas Gd2Fe16Cr has minimum αν = −7.03 × 10−6 K−1 in the temperature range 294–454 K.34 Possessing a lower NTE coefficient33 and having TC (330 K) higher than the room temperature among the R2Fe17 compounds, Ho2Fe17 may be a good starting material for finding a magnetic ZTE material operative at room temperature. In this paper, we study the effect of Cr-substitution on the NTE and the magnetic behavior of a R2Fe17-type compound, Ho2Fe17. Our study suggests that Ho2Fe16Cr, with a high TC, is a ZTE material having a low coefficient of thermal expansion (αν = 1.3 × 10−6 K−1) over a wide range of temperature (13–330 K).

2 Experimental procedure

Ho2Fe17–xCrx (x = 0, 1, 2) compounds were prepared by the method of arc-melting (in an argon atmosphere) with at least 99.9% pure starting materials. The ingots were re-melted five to six times to ensure homogeneity. The samples were annealed in a vacuum sealed quartz tube at 1173 K for 7 days, followed by quenching in ice water. The room temperature powder X-ray diffraction (XRD) patterns of the samples were taken using CuKα radiation in a TTRAX III diffractometer (M/S Rigaku Corp., Japan). The XRD patterns at different temperatures (13–515 K) were recorded using the same instrument, with a very low scan speed (0.01° steps, and 0.4° min−1) for a better statistical average. This is necessary in our case as our sample contains more than 85% Fe and the Cu Kα radiation is in the absorption edge of Fe. The magnetization was measured using SQUID VSM and PPMS evercool-II (M/S Quantum Design, Inc., USA) from 4–380 K. High temperature VSM (Model EV9, M/S MicroSense, LLC Corp., USA) was employed to measure the magnetization from 300–550 K.

3 Results and discussion

Fig. 1 [left panel] shows the room temperature XRD data of the Ho2Fe17−xCrx compounds. The data show that all the samples form in a single phase with a hexagonal Th2Ni17-type structure (space group: P63/mmc). Miller indices (hkl) for the major set of crystallographic planes have been shown in the figure. Fig. 1 [right panel] shows the XRD pattern of Ho2Fe16Cr at three temperatures, namely, 13 K, 300 K, and 453 K. The XRD data suggest that the compound remains in a single phase with the Th2Ni17-type structure in the temperature range of measurement. This is consistent for other Ho2Fe17−xCrx compounds (x = 0, 2). Lattice parameters (a, c) of each sample at different temperatures have been estimated by analyzing XRD patterns.
image file: c6ra20216k-f1.tif
Fig. 1 [Left panel] Room temperature XRD pattern of Ho2Fe17−xCrx (x = 0, 1, 2) compounds. [Right panel] XRD patterns of Ho2Fe16Cr at T = 13 K, 300 K and 453 K. Red circles depict the experimentally observed data points, the black lines are data generated using FullProf software, and the blue lines are the differences between the estimated and experimentally observed data points. The olive bars are the Bragg positions allowed by the space group.

Fig. 2 [left panel] shows the effect of Cr substitution on the lattice parameters (a, c) and [right panel] the lattice volume v of the Ho2Fe17−xCrx (x = 0, 1, 2) compounds. We observe that the thermal expansion of any of the Ho2Fe17−xCrx compounds is anisotropic like the parent R2Fe17 compounds33 and their derivatives,34–36 i.e., a increases while c decreases with increasing temperature. The decrease in c with increasing temperature results from MVE. In R2Fe17 compounds, with a hexagonal Th2Ni17 structure for heavy rare earths (rhombohedral Th2Zn17 structure for light rare earths), the crystallographic sites for the Fe atoms are 4f(6c), 6g(9d), 12j(18f), and 12k(18h) sites.39 As the distance between the Fe atoms at 4f(6c) sites is very short (<0.244 nm)36 (Fig. 3), the direct exchange between these atoms gives rise to antiferromagnetic (AFM) coupling, while the rest of the atoms are coupled ferromagnetically. Therefore, a large amount of magnetic energy is stored along the Fe–Fe distance between the atoms at 4f(6c) sites, and this magnetic energy can be reduced by increasing said distance, which is dependent only on the parameter c.40 Finally, a compromise is obtained between the magnetic energy and the elastic energy by increasing the lattice volume for the magnetic state at a lower temperature. If Cr atoms are substituted for Fe atoms in R2Fe17 compounds, Cr atoms prefer to occupy 4f(6c) sites.36 As the magnetic moment of a Cr atom is less than that of an Fe atom, such substitution will reduce the effective MVE, and consequently weaken the NTE below TC.


image file: c6ra20216k-f2.tif
Fig. 2 [Left panel] Lattice parameters a and c and [Right panel] unit cell volume v of Ho2Fe17−xCrx (x = 0, 1, 2) compounds (vertical line in both left and right panel corresponds to T/TC = 1). Data for the parent compound (x = 0) has been taken from ref. 41.

image file: c6ra20216k-f3.tif
Fig. 3 Th2Ni17 – type crystal structure (space group: P63/mmc) of Ho2Fe17 compound. Short distances between Fe atoms at 4f sites (highlighted by red boxes) are responsible for the magneto-volume effect.

The Curie temperature (TC) for each of the Ho2Fe17−xCrx (x = 0, 1, 2) compounds has been determined from the magnetization (M) curve plotted as a function of temperature at a magnetic field of H = 0.05 T (Fig. 4 [left panel]). The data shows that the TC of Ho2Fe16Cr is 415 K, higher than that (326 K) of the parent compound, whereas the TC (402 K) of Ho2Fe15Cr2 is lower than that of Ho2Fe16Cr. The Cr-substitution enhances TC substantially for a lower concentration of Cr, and then TC decreases at higher concentrations. Similar behavior has also been observed for other Cr-substituted R2Fe17 compounds.34–36 The modification of TC due to Cr substitution has been explained by the preference of the magnetically weaker Cr atoms to replace Fe atoms at 4f(6c) sites.36 The strength of the Fe(Cr)–Fe(Cr) interactions in the 3d-sublattice determines the value of TC in the R2Fe17−xCrx compounds.36 In the parent compound R2Fe17, the Fe atoms at the 4f(6c) sites are coupled antiferromagnetically, while the rest of the atoms are coupled ferromagnetically. The Cr atoms at 4f(6c) sites reduce the interactions between 4f(6c) and other crystal sites (4f(6c), 6g(9d), 12j(18f), and 12k(18h)). At lower concentrations of Cr, the strength of the AFM interactions between 4f(6c)–4f(6c) sites reduces much more strongly than the ferromagnetic (FM) interactions between 4f(6c)-other sites.36 This results in an enhancement in the total FM interactions in the 3d-sublattice and a subsequent increase in TC. For higher concentrations of Cr, the reduction of FM interactions has been argued to surpass that of AFM interactions causing a decrease in total FM strength of the 3d-sublattice, and an associated lowering of TC. A similar feature of TC dependence on the Cr-concentration has also been observed in Fe–Cr alloys,42 where through experimental as well as computational studies, it was shown that the TC of the Fe–Cr alloy (up to 6 at% Cr-substitution) is higher than that of pure Fe. The coercive field (Hc) also increases with increasing Cr concentration. Materials with a high coercive field possess a high value of maximum energy product (BHmax), a necessary criterion for permanent magnetic materials. The values of TC, MS and Hc of Ho2Fe17−xCrx (x = 0, 1, 2) are shown in Table 1.


image file: c6ra20216k-f4.tif
Fig. 4 [Left panel] Magnetization as a function of temperature at a magnetic field of H = 0.05 T, [Right panel] magnetization as a function of magnetic field at T = 5 K for Ho2Fe17−xCrx (x = 0, 1, 2) compounds. The lower inset of [Right panel] shows the virgin curves with the fit of eqn (3), and the upper inset of [Right panel] shows the central part of each MH curve.
Table 1 TC, MS and Hc of Ho2Fe17−xCrx (x = 0, 1, 2)
Sample name TC (in K) MS (in μB per f.u) Hc (in Oe)
Ho2Fe17 330 23.5 175
Ho2Fe16Cr 415 16.4 720
Ho2Fe15Cr2 402 10.8 2300


According to the Stoner–Edwards–Wohlfarth (SEW) theory,43 in the absence of any external magnetic field, the volume strain (ωS) arising from MVE at a temperature T is

 
image file: c6ra20216k-t1.tif(1)
where M0 is the spontaneous magnetic moment, κ is the compressibility, and C is the magneto-volume coupling constant. Moriya and Usami modified eqn (1) by including the contribution of spin fluctuations to the free energy as43
 
image file: c6ra20216k-t2.tif(2)
where ξ2(T) represents the average of the squared thermal spin fluctuation amplitude. The volume strain due to MVE is proportional to the square of M0. The Cr-substitution reduces the value of M0. This is evident from Fig. 4 [right panel], where we observe that the saturation magnetization (MS), which is a measure of M0, decreases with increasing Cr-concentration. MS has been estimated by fitting the virgin MH curve of each sample with the expression44
 
image file: c6ra20216k-t3.tif(3)
that describes the law of approach to saturation (where A, B and χ are constants). As M0 decreases with increasing Cr-concentration, so also ωS decreases.

Fig. 2 [left panel] shows that a(T) is lower for a higher concentration (x) of Cr. This observation can be explained by the smaller ionic size of Cr than Fe,36 and the fact that MVE does not alter a much. However, Cr is positioned in the left side of the same row of Fe in the periodic table, and it suggests that Cr possesses a larger crystal or ionic radius than that of Fe.45 It may be pointed out that such understanding holds well when both the ions possess the same valence as well as identical spin state (high-spin or low-spin).45 Moreover, in two different crystal environments, if the site coordination numbers differ, the same element in a particular valence state may assume two different ionic radii.45 Reduction of the lattice parameter a has also been found for other members of the R2Fe17−xCrx series.34–36,46–48 Even for Mn, placed also like Cr on the left side of Fe in the periodic table, similar reduction in a has been found with increasing x, in R2Fe17−xMnx.49,50 A better understanding of the fact can be obtained from the neutron diffraction study of Nd2Fe17−δCrδ (δ = 0, 0.5, 1, 1.9).47 The study suggests that Cr prefers to occupy 6c, and with the introduction of Cr, 6c–6c, 6c–18h, and 6c–18f bond lengths reduce continuously with increasing δ, while 6c–9d bond length remains almost constant. Considering such reduction in different bond lengths with the introduction of Cr, one may conclude that the ionic size of Cr is less than Fe in the same site of the R2Fe17−xCrx lattice. The reduction of the mentioned bond lengths is an experimental observation, but the underlying reason lies within the valence as well as the spin states of two (Fe, Cr) ions and finally the crystalline environment. Here, MVE does not alter a much. The lattice parameter c(T) for Ho2Fe16Cr is higher than that of the parent compound even above TC. This occurs because the strain along the c-axis arising from MVE is not only higher in the parent compound than Ho2Fe16Cr, but also more than the change in the c parameter obtained by just replacing one Fe atom by a Cr atom in Ho2Fe17. The higher value of c(T) for Ho2Fe16Cr, even above TC, suggests that the effect of spin fluctuation is important in this system. A similar conclusion can also be drawn from the observation of NTE in the parent compound below 365 K,41 which is higher than TC = 330 K (Fig. 2 [right panel]). For Ho2Fe15Cr2, the reduction of c(T) due to further replacement of one more Fe atom by a Cr atom overcompensates the change in the same due to reduction of MVE, and therefore, c(T) reduces compared to Ho2Fe16Cr.

Fig. 2 [right panel] shows the unit cell volume v(T) of Ho2Fe17−xCrx (x = 0, 1, 2). The parent compound Ho2Fe17 shows NTE in the temperature range 295–365 K with αν lying in the range (1.3–3.7) × 10−5 K−1.41 The parent compound appears to be a strong magnetostrictive material. In comparison with the parent compound, in Ho2Fe16Cr, Cr-substitution increases TC and weakens NTE. For Ho2Fe16Cr, NTE is observed in the range T = 330–425 K with αν = −4.3 × 10−6 K−1. However, the most interesting feature of Ho2Fe16Cr is the negligible thermal expansion of the unit cell over a wide range of temperature (13–330 K) including room temperature. In this temperature range, the magneto-volume strain compensates for the lattice volume strain, and Ho2Fe16Cr behaves as a single component ZTE material with αν = 1.3 × 10−6 K−1. By further increasing the Cr concentration NTE disappears, i.e., we do not observe any NTE for Ho2Fe15Cr2.

The temperature dependence of the spontaneous linear magnetostrictive deformation in the basal plane, λa = (amap)/ap, that along the c-axis λc = (cmcp)/cp, and the spontaneous volume magnetostrictive deformation ωS = (vmvp)/vp36 are important parameters for identifying the strength and the nature of magnetoelastic coupling. am, cm, and vm are the experimentally measured values of a, c and v, whereas ap, cp, and vp are the corresponding values obtained by extrapolation36 from the paramagnetic state using the Grüneisen relation, image file: c6ra20216k-t4.tif, image file: c6ra20216k-t5.tif, where θD is the Debye temperature and [scr R, script letter R] is the molar gas constant. The value of θD has been taken as 400 K, estimated earlier51 for R2Fe17 compounds other than Y2Fe17. The parameters am, cm, and vm along with ap, cp, and vp are presented in Fig. 5 for the compound Ho2Fe16Cr. λa, λc and ωS have been plotted as a function of temperature in Fig. 6(a)–(c) respectively.


image file: c6ra20216k-f5.tif
Fig. 5 Temperature dependent lattice parameters a, c and unit cell volume v of the compound Ho2Fe16Cr, extracted from the fitted temperature dependent XRD patterns (denoted by suffix m), and the same parameters extrapolated from the paramagnetic region (denoted by suffix p).

image file: c6ra20216k-f6.tif
Fig. 6 (a and b) Linear magnetostrictive deformations λa, and λc, (c) spontaneous volume magnetostrictive deformation ωS as a function of temperature (T) for the Ho2Fe17−xCrx (x = 1, 2) compounds. (d) ωS as a function of MS2 for the compound Ho2Fe16Cr.

Although it has been suggested that the free energy in the magnetic state can be reduced by changing the dimension along the c-axis only,52 our experimental data show a finite λa for both the samples Ho2Fe17−xCrx (x = 1 or 2) i.e., the magneto-elastic effect also changes a. Moreover, the larger value of λc than λa in each case suggests that the magnetoelastic effect along the c-axis is stronger than that in the basal plane. As evident from Fig. 6(c), the value of ωS is considerably high at low temperatures, and it is non-zero above TC. Thus Ho2Fe17−xCrx is a strong magnetostrictive material in which spin fluctuation plays an important role. Fig. 6(d) shows that ωS varies linearly with the square of MS. This is in accordance with eqn (1) and (2), and has been suggested to be a direct experimental proof of the fact that ωS is related to the MVE.53

4 Conclusion

In the search for magnetic ZTE materials, our study shows that Cr-substitution for Fe atoms in Ho2Fe17 weakens the NTE of the parent compound and alters the TC. Ho2Fe16Cr behaves as a single component ZTE material in the temperature range 13–330 K. Moreover, the TC of Ho2Fe16Cr is 415 K, considerably higher than room temperature. Cr-substitution also increases the coercivity compared to that of the parent compound. Therefore, Ho2Fe16Cr with a high TC, moderate coercivity and very low thermal expansion is a potential material for a permanent magnet.

Acknowledgements

Shovan Dan and S. Mukherjee thank UGC for financial support in the form of major project (MRP-MAJOR-PHYS-2013-16282). The work at SINP was carried out under CMPID-DAE Project.

References

  1. J. R. Salvador, F. Guo, T. Hogan and M. G. Kanatzidis, Nature, 2003, 425, 702–705 CrossRef CAS PubMed.
  2. S. Margadonna, K. Prassides and A. N. Fitch, J. Am. Chem. Soc., 2004, 126, 15390–15391 CrossRef CAS PubMed.
  3. K. J. Miller, C. P. Romao, M. Bieringer, B. A. Marinkovic, L. Prisco and M. A. White, J. Am. Ceram. Soc., 2013, 96, 561–566 CAS.
  4. P. Hu, J. Chen, J. Deng and X. Xing, J. Am. Chem. Soc., 2010, 132, 1925–1928 CrossRef CAS PubMed.
  5. K. Takenaka and H. Takagi, Appl. Phys. Lett., 2009, 94, 131904 CrossRef.
  6. T. Suzuki and A. Omote, J. Am. Ceram. Soc., 2006, 89, 691–693 CrossRef CAS.
  7. M. van Schilfgaarde, I. A. Abrikosov and B. Johansson, Nature, 1999, 400, 46–49 CrossRef CAS.
  8. M. Shiga, Curr. Opin. Solid State Mater. Sci., 1996, 1, 340–348 CrossRef CAS.
  9. L. Hu, J. Chen, L. Fan, Y. Ren, Y. Rong, Z. Pan, J. Deng, R. Yu and X. Xing, J. Am. Chem. Soc., 2014, 136, 13566–13569 CrossRef CAS PubMed.
  10. W. Wang, R. Huang, W. Li, J. Tan, Y. Zhao, S. Li, C. Huang and L. Li, Phys. Chem. Chem. Phys., 2015, 17, 2352–2356 RSC.
  11. W. Li, R. Huang, W. Wang, Y. Zhao, S. Li, C. Huanga and L. Li, Phys. Chem. Chem. Phys., 2015, 17, 5556–5560 RSC.
  12. S. Li, R. Huang, Y. Zhao, W. Wanga and L. Li, Phys. Chem. Chem. Phys., 2015, 17, 30999–31003 RSC.
  13. K. Takenaka, Sci. Technol. Adv. Mater., 2012, 13, 013001–013012 CrossRef.
  14. H. Li, S. Liu, L. Chen, J. Zhao, B. Chen, Z. Wang, J. Meng and X. Liu, RSC Adv., 2015, 5, 1801–1807 RSC.
  15. C. Verdon and D. C. Dunand, Scr. Mater., 1997, 36, 1075–1080 CrossRef CAS.
  16. A. Matsumoto, K. Kobayashi, T. Nishio and K. Ozaki, Mater. Sci. Forum, 2003, 426–432, 2279–2284 CrossRef CAS.
  17. L. Ding, C. Wang, Y. Na, L. Chu and J. Yan, Scr. Mater., 2011, 65, 687–690 CrossRef CAS.
  18. X. Yan, J. Miao, J. Liu, X. Wu, H. Zou, D. Sha, J. Ren, Y. Dai, J. Wang and X. Cheng, J. Alloys Compd., 2016, 677, 52–56 CrossRef CAS.
  19. X. Shan, R. Huang, Y. Han, C. Huang, X. Liu, Z. Lu and L. Li, J. Alloys Compd., 2016, 662, 505–509 CrossRef CAS.
  20. G. D. Barrera, J. A. O. Bruno, T. H. K. Barron and N. L. Allan, J. Phys.: Condens. Matter, 2005, 17, R217–R252 CrossRef CAS.
  21. T. A. Mary, J. S. O. Evans, T. Vogt and A. W. Sleight, Science, 1996, 272, 90–92 CAS.
  22. F. H. Gillery and E. A. Bush, J. Am. Ceram. Soc., 1959, 42, 175–177 CrossRef CAS.
  23. A. E. Phillips, A. L. Goodwin, G. J. Halder, P. D. Southon and C. J. Kepert, Angew. Chem., Int. Ed., 2008, 47, 1396–1399 CrossRef CAS PubMed.
  24. T. Chatterji, T. C. Hansen, M. Brunelli and P. F. Henry, Appl. Phys. Lett., 2009, 94, 241902 CrossRef.
  25. M. P. Attfield, M. Feygenson, J. C. Neuefeind, T. E. Proffen, T. C. A. Lucas and J. A. Hriljac, RSC Adv., 2016, 6, 19903–19909 RSC.
  26. T. Matsuda, J. Kimb and Y. Moritomo, RSC Adv., 2011, 1, 1716–1720 RSC.
  27. Y. Cheng, Y. Liang, X. Ge, X. Liu, B. Yuan, J. Guo, M. Chao and E. Lianga, RSC Adv., 2016, 6, 53657–53661 RSC.
  28. M. Azuma, W. Chen, H. Seki, M. Czapski, S. Olga, K. Oka, M. Mizumaki, T. Watanuki, N. Ishimatsu, N. Kawamura, S. Ishiwata, M. G. Tucker, Y. Shimakawa and J. P. Attfield, Nat. Commun., 2011, 2, 347 CrossRef PubMed.
  29. A. Pandey, C. Mazumdar, R. Ranganathan, S. Tripathi, D. Pandey and S. Dattagupta, Appl. Phys. Lett., 2008, 92, 261913 CrossRef.
  30. M. Hayase, M. Shiga and Y. Nakamura, J. Phys. Soc. Jpn., 1973, 34, 925–933 CrossRef CAS.
  31. K. Sumiyama, M. Shiga, M. Morioka and Y. Nakamura, J. Phys. F: Met. Phys., 1979, 9, 1665–1677 CrossRef CAS.
  32. H. Nakamura, H. Wada, K. Yoshimura, M. Shiga, Y. Nakamura, J. Sakurai and Y. Komura, J. Phys. F: Met. Phys., 1988, 18, 981–991 CrossRef CAS.
  33. P. A. Alanso, Doctoral Thesis, University of Oviedo, Spain, 2011 Search PubMed.
  34. H. Yan-Ming, T. Ming, W. Wei and W. Fang, Chin. Phys. B, 2010, 19, 067502 CrossRef.
  35. Y. Hao, M. Zhao, Y. Zhou and J. Hu, Scr. Mater., 2005, 53, 357–360 CrossRef CAS.
  36. Y. Hao, X. Zhang, B. Wang, Y. Yuang and F. Wang, J. Appl. Phys., 2010, 108, 023915 CrossRef.
  37. K. Takenaka and H. Takagi, Appl. Phys. Lett., 2005, 87, 261902 CrossRef.
  38. X. Peng, J. Chen, K. Lin, L. Fan, Y. Rong, J. Deng and X. Xing, RSC Adv., 2016, 6, 32979–32982 RSC.
  39. J. L. Wang, S. J. Campbell, O. Tegus, C. Marquina and M. R. Ibarra, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 174423 CrossRef.
  40. D. Givord and R. Lemaire, IEEE Trans. Magn., 1974, 10, 109–113 CrossRef CAS.
  41. J. L. Wang, A. J. Studer, S. J. Kennedy, R. Zeng, S. X. Dou and S. J. Campbell, J. Appl. Phys., 2012, 111, 07A911 Search PubMed.
  42. M. Y. Lavrentiev, K. Mergia, M. Gjoka, D. Nguyen-Manh, G. Apostolopoulos and S. L. Dudarev, J. Phys.: Condens. Matter, 2012, 24, 326001 CrossRef PubMed.
  43. Y. Takahashi, Spin Fluctuation Theory of Itinerant Electron Magnetism, Springer Tracts in Modern Physics, Springer-Verlag, Berlin Heidelberg, 2013, p. 253 Search PubMed.
  44. B. D. Cullity and C. D. Graham, Introduction to Magnetic Materials, IEEE Press, Wiley, 2008 Search PubMed.
  45. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751–767 CrossRef.
  46. X. C. Kou, F. R. de Boer, R. Grssinger, G. Wiesinger, H. Suzuki, H. Kitazawa, T. Takamasu and G. Kido, J. Magn. Magn. Mater., 1998, 177–181, 1002–1007 CrossRef CAS.
  47. E. Girt, Z. Altounian and J. Yang, J. Appl. Phys., 1997, 81, 5118–5120 CrossRef CAS.
  48. I. Nehdi, L. Bessais, C. D. -Mariadassou, M. Abdellaoui and H. Zarrouk, J. Alloys Compd., 2003, 351, 24–30 CrossRef CAS.
  49. P. C. Ezekwenna, G. K. Marasinghe, W. J. James, O. A. Pringle, G. J. Long, H. Luo, Z. Hu, W. B. Yelon and P. I. Hritier, J. Appl. Phys., 1997, 81, 4533 CrossRef CAS.
  50. Y. Wang, F. Yang, C. Chen, N. Tang, P. Lin and Q. Wang, J. Appl. Phys., 1998, 84, 6229–6232 CrossRef CAS.
  51. A. V. Andreev, A. V. Deryagin, S. M. Zadvorkin, N. V. Kudrevatykh, R. H. Levitin, V. N. Moskalev, Y. F. Popov and R. Y. Yumaguzhin, Fizika Magnitnykh Materialov, Physics of Magnetic Materials, ed. D. D. Mishin, Kalinin University, Kalinin, USSR, 1985, p. 21, in Russian Search PubMed.
  52. D. Givord, R. Lemaire, W. J. James, J.-M. Moreau and J. S. Shah, IEEE Trans. Magn., 1971, 7, 657–659 CrossRef CAS.
  53. J. Chen, L. Hu, J. Deng and X. Xing, Chem. Soc. Rev., 2015, 44, 3522–3567 RSC.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.