Open access

Unusual Temperature Dependence of Zero-Field Ferromagnetic Resonance in Millimeter Wave Region on Al-Substituted ε-Fe2O3

Written By

Marie Yoshikiyo, Asuka Namai and Shin-ichi Ohkoshi

Submitted: 06 October 2012 Published: 31 July 2013

DOI: 10.5772/55779

From the Edited Volume

Ferromagnetic Resonance - Theory and Applications

Edited by Orhan Yalcin

Chapter metrics overview

3,629 Chapter Downloads

View Full Metrics

1. Introduction

Insulating magnetic materials absorb electromagnetic waves. This absorption property is one of the important functions of magnetic materials, which is widely applied in our daily life as electromagnetic wave absorbers to avoid electromagnetic interference problems [1-5]. For example, spinel ferrites are used as absorbers for the present Wi-Fi communication, which uses 2.4 GHz and 5 GHz frequency waves. With the development of information technology, the demand is rising for sending heavy data such as high-resolution images at high speed. Recently, high-frequency electromagnetic waves in the frequency range of 30–300 GHz, called millimeter waves, are drawing attention as a promising carrier for the next generation wireless communication. For example, 76 GHz is an important frequency, which is beginning to be used for vehicle radars. There are also new audio products coming to use, applying millimeter wave communication in the 60 GHz region [6,7]. However, there had been no magnetic material that could absorb millimeter waves above 80 GHz before our report on ε-Fe2O3.

Well-known forms of Fe2O3 are α-Fe2O3 and γ-Fe2O3, commonly called as hematite and maghemite, respectively. However, our research group first succeeded in preparing a pure phase of ε-Fe2O3, which is a rare phase of iron oxide Fe2O3 that is scarcely found in nature [8–10]. Since then, its physical properties have been actively studied, and one of the representative properties is the gigantic coercive field (Hc) of 20 kilo-oersted (kOe) at room temperature [11–18]. We have also reported metal-substituted ε-Fe2O3 (ε-MxFe2–xO3, M = In, Ga, Al, and Rh), and showed that this series absorb millimeter waves from 35–209 GHz at room temperature due to zero-field ferromagnetic resonance (so called natural resonance) [19-29]. ε-Fe2O3 based magnet is expected to be a leading absorbing material for the future wireless communication using higher frequency millimeter waves.

In this chapter, we first introduce the synthesis, crystal structure, magnetic properties, and the formation mechanism of the original ε-Fe2O3 [8–10]. Then we report the physical properties of Al-substituted ε-Fe2O3, mainly focusing on its millimeter wave absorption properties due to zero-field ferromagnetic resonance. The resonance frequency was widely controlled from 112–182 GHz by changing the aluminum substitution ratio [23]. Furthermore, from a scientific point of view, temperature dependence of zero-field ferromagnetic resonance was investigated and was found to show an anomalous behavior caused by the spin reorientation phenomenon [28].

Advertisement

2. ε-Fe2O3

This section introduces the synthesis, crystal structure, magnetic properties, and the formation mechanism of ε-Fe2O3. ε-Fe2O3 had only been known as impurity in iron oxide materials, and its properties were clarified for the first time after our success in the synthesis of single-phase ε-Fe2O3 in 2004 [8].

2.1. Synthesis, crystal structure, and magnetic properties of ε-Fe2O3

Single-phase ε-Fe2O3 nanoparticles are synthesized by a chemical method, combining reverse-micelle and sol-gel techniques (Figure 1) [8−10,16]. In the reverse-micelle step, two reverse-micelle systems, A and B, are formed by cetyl trimethyl ammonium bromide (CTAB) and 1-butanol in n-octane. Reverse-micelle A contains aqueous solution of Fe(NO3)3 and Ba(NO3)2, and reverse-micelle B contains NH3 aqueous solution. These two microemulsion systems are mixed under rapid stirring. Tetraethoxysilane (C2H5O)4Si is then added to this solution, which forms SiO2 matrix around the Fe(OH)3 nanoparticles through 20 hours of stirring. The precipitation is separated by centrifugation and sintered in air at 1000˚C for 4 hours. The SiO2 matrix is removed by stirring in NaOH solution at 60˚C for 24 hours.

Figure 1.

Synthetic procedure of ε-Fe2O3 nanomagnets using a combination method of reverse-micelle and sol-gel techniques. The inset is a transmission electron microscopy image of ε-Fe2O3 nanorods.

With this synthesis method, rod-shaped ε-Fe2O3 is obtained due to the effect of Ba2+ ions, which adsorb on particular planes of ε-Fe2O3, inducing growth towards one direction. Spherical ε-Fe2O3 nanoparticles can also be synthesized by a different method without Ba2+ ions, which is an impregnation method using mesoporous silica nanoparticles [17,29,30]. Methanol and water solution containing Fe(NO3)3 is immersed into mesoporous silica and heated in air at 1200°C for 4 hours. The etching process is the same as above.

The crystal structure of ε-Fe2O3 is shown in Figure 2a. It has an orthorhombic crystal structure (space group Pna21) with four non-equivalent Fe sites, A, B, C, and D sites. A, B, and C sites are six-coordinated octahedral sites, and D site is a four-coordinated tetrahedral site. ε-Fe2O3 exhibits spontaneous magnetization at a Curie temperature (TC) of 500 K. Figure 2b presents magnetization versus external magnetic field curve at 300 K, which shows a huge Hc value of 20 kOe. Before this finding, the largest Hc value among metal oxide was 6 kOe of barium ferrite, BaFe12O19 [31], which indicates that the Hc of ε-Fe2O3 is over three times larger. The magnetic structure has been investigated using molecular field theory, which indicated that B and C sites have positive sublattice magnetizations, and A and D sites have negative sublattice magnetizations [32]. This result was consistent with the experimental results from neutron diffraction measurements, Mössbauer spectroscopy measurements, etc. [13,14], and was also consistent with first-principles calculation results [33].

Figure 2.

a) Crystal structure of ε-Fe2O3. Dark blue, purple, light blue, and pink polyhedrons indicate A, B, C, and D sites, respectively. (b) Magnetization versus external magnetic field curve of ε-Fe2O3 at 300 K. Inset is a schematic illustration of the sublattice magnetizations of each site.

2.2. Formation mechanism of ε-Fe2O3

Here we discuss the formation mechanism of ε-Fe2O3 from the viewpoint of phase transformation. By changing the sintering temperature in the present synthesis, a phase transformation of γ-Fe2O3 → ε-Fe2O3 → α-Fe2O3 was observed accompanied by an increase of particle size. γ- and α-Fe2O3 are very common phases of Fe2O3, and it has been well known that γ-Fe2O3 transforms directly into α-Fe2O3 in a bulk form. In the present case, it is considered that ε-Fe2O3 appeared as a stable phase at an intermediate size region due to the large surface energy effect. Free energy of each i-phase (Gi, i = γ, ε, or α) is expressed as a sum of chemical potential (μi) and surface energy (Aiσi):

Gi=μi+Aiσi,E1

where Ai is molar surface area and σi is surface free energy of a particle. Since, Ai is equal to 6Vm,i/d, where Vm,i and d represent the molar volume and particle diameter, respectively, the free energy per molar volume is expressed as

Gi/Vm,i=μi/Vm,i+6σi/d.E2

This equation indicates that the contribution of the surface energy increases with the decrease of particle diameter. When the parameters satisfy the following three conditions, μγ>με>μα, σγ<σε<σα, and(σεσγ)/(σασε)<(μεμγ)/(μαμε), the free energy curve for each phase, Gγ/Vm,γ, Gε/Vm,ε, and Gα/Vm,α intersect to form ε-Fe2O3 as the most stable phase at an intermediate d value (Figure 3). Such nanosize effect has also been reported for other metal oxide materials, e.g. Al2O3 [34,35] and Ti3O5 [36].

Figure 3.

Representation of free energy per volume (G/V) versus particle diameter (d) for the three phases of Fe2O3. Green, blue, and red lines indicate the G/V curves for γ-Fe2O3, ε-Fe2O3, and α-Fe2O3, respectively. Below the graph are the crystal structures of γ-Fe2O3, ε-Fe2O3, and α-Fe2O3 from the left to right.

Advertisement

3. Al-substituted ε-Fe2O3

In this section, synthesis, crystal structure, and various physical properties of Al-substituted ε-Fe2O3, ε-AlxFe2–xO3, is discussed. Especially, the millimeter wave absorption property by zero-field ferromagnetic resonance is focused.

3.1. Synthesis of Al-substituted ε-Fe2O3

ε-AlxFe2–xO3 samples (x = 0.06, 0.09, 0.21, 0.30, 0.40) were synthesized by the same method as the original ε-Fe2O3, using the combination of reverse-micelle and sol-gel techniques. Reverse-micelle A contained aqueous solution of Fe(NO3)3 and Al(NO3)3, and the mixing ratio was adjusted to obtain the different samples, x = 0.06, 0.09, 0.21, 0.30, and 0.40. The sintering temperature was 1050°C for x = 0.06, 0.09, 0.30, and 0.40, and 1025°C for x = 0.21. Only the sample for x = 0 was prepared by an impregnation method using mesoporous silica nanoparticles. The SiO2 matrices for all samples were etched by NaOH solution. The morphology and size of the obtained samples were examined using transmission electron microscopy (TEM), which showed spherical nanoparticles with an average particle size between 20-50 nm (Figure 4).

Figure 4.

Transmission electron microscopy images of ε-AlxFe2–xO3 samples. The black bars indicate the scale.

3.2. Al-substitution effect in crystal structure and magnetic properties

X-ray diffraction (XRD) patterns indicated the samples to have the same orthorhombic crystal structure as the original ε-Fe2O3. The Rietveld analyses of the XRD patterns showed a constant decrease in the lattice constants with the degree of Al-substitution. The analysis results also indicated that the Al3+ ions introduced in the samples have site selectivity in the substitution. For example, in the x = 0.21 sample, the Al3+ substitution ratio of each Fe site was 0%, 3%, 8%, and 30% for A, B, C, and D site, respectively. This tendency for the Al3+ ion to prefer D site was consistent with all of the Al-substituted samples (Figure 5). This site selectivity can be understood by the smaller ion radius of Al3+ (0.535 Å) compared to Fe3+ (0.645 Å) [37]. The Al3+ ions prefer to occupy the smaller tetrahedral D site than the octahedral A, B, and C sites.

Figure 5.

Al3+ occupancy ratio for A, B, C, and D site. Square, diamond, circle, and triangle plots represent A, B, C, and D site, respectively. Inset is the crystal structure of ε-Fe2O3.

The magnetic properties of the samples are shown in Table 1. The field-cooled magnetization curves under an external magnetic field of 10 Oe showed that the TC value decreased from 500 K to 448 K with the increase of Al-substitution (Figure 6, upper right). From the magnetization versus external magnetic field measurements, gradual change of the hysteresis loops was also observed. The obtained hysteresis loops of x = 0, 0.21, and 0.40 samples are shown in Figure 6. With Al-substitution, the Hc value decreased from 22.5 kOe to 10.2 kOe, and saturation magnetization (Ms) value increased. These changes in the magnetic properties can be explained by the metal replacement of Fe3+ magnetic ions (3d5, S = 5/2) by non-magnetic Al3+ ions (3d0, S = 0). As mentioned previously, ε-Fe2O3 is a ferrimagnet with positive sublattice magnetizations at B and C sites and negative sublattice magnetizations at A and D sites. With the substitution of D site Fe3+ ions with non-magnetic Al3+, the total magnetization increases, leading to the increase of Ms value. In addition, the non-magnetic Al3+ ions reduce the superexchange interaction between the magnetic sites, resulting in a decrease of TC [32]. In this way, the magnetic properties can be widely controlled by Al-substitution.

xTC (K)Hc (kOe)Ms (emu/g)
050022.514.9
0.0649619.115.1
0.0949017.514.6
0.2148014.917.0
0.3046613.820.3
0.4044810.219.7

Table 1

Magnetic properties of ε-AlxFe2–xO3.

Figure 6.

Magnetization versus external magnetic field curve at 300 K for the samples x = 0, 0.21, and 0.40 (left), Curie temperature (TC) versus x value plot (upper right), and coercive field (Hc) versus x value plot (lower right).

3.3. Electromagnetic wave absorption of Al-substituted ε-Fe2O3 by zero-field ferromagnetic resonance

Zero-field ferromagnetic resonance is a resonance phenomenon caused by the gyromagnetic effect induced by an electromagnetic wave irradiation under no magnetic field (Figure 7). This phenomenon is observed in ferromagnetic materials with magnetic anisotropy. When the magnetization is tilted away from the easy-axis by the magnetic component of the electromagnetic wave, precession of the magnetization occurs around the easy-axis due to gyromagnetic effect. Resonance is observed when this precession frequency coincides with the electromagnetic wave frequency, resulting in electromagnetic wave absorption at the particular frequency [38]. This resonance frequency (fr) is proportional to the magnetocrystalline anisotropy (Ha) and can be expressed as

fr=(ν/2π)Ha,E3

where ν is the gyromagnetic ratio. If the sample is consisted of randomly oriented particles with uniaxial magnetic anisotropy, the Ha value is proportional to Hc. Therefore, electromagnetic wave absorption at high frequencies is expected with insulating materials exhibiting large coercivity, which is the case for ε-Fe2O3 based magnets.

Figure 7.

a) A schematic illustration of zero-field ferromagnetic resonance (natural resonance), resulting in electromagnetic wave absorption due to the precession of magnetization around the easy-axis. The M and E of the electromagnetic wave indicate the magnetic and electric components, respectively. (b) An illustration indicating that larger coercive field (Hc) results in higher resonance frequency (fr).

With the general electromagnetic wave absorption measurement using free space absorption measurement system, the absorption frequencies of the present ε-AlxFe2-xO3 samples exceeded the measurement range, where the maximum is 110 GHz. Therefore, the absorption measurements were conducted using terahertz time domain spectroscopy (THz-TDS) at room temperature. The THz-TDS measurement system is shown in Figure 8. A mode-locked Ti:sapphire femtosecond pulse laser with a time duration of 20 fs at a repetition rate of 76 MHz was used. The output was divided into a pump and probe beam for the time-domain system. For THz wave emitter and detector, dipole type and bowtie type low-temperature-grown GaAs photoconductive antennas were used, respectively. The sample was set on a sample holder, which was inserted between a set of paraboloidal mirrors concentrating the THz wave at the location of the sample. The temporal waveforms of the electric component of the transmitted THz pulse waves were obtained by changing the delay time between the pump and probe pulses. The temporal waves were Fourier transferred to obtain the frequency dependence, and the absorption spectra were calculated using the following equation:

(Absorption)=10log|t(ω)|2(dB),E4

where t(ω) is the complex amplitude transmittance. An absorption of 20 dB indicates 99% absorption.

The electromagnetic wave absorption spectra are shown in Figure 9. Absorption peaks were observed at 112 GHz (x = 0.40), 125 GHz (x = 0.30), 145 GHz (x = 0.21), 162 GHz (x = 0.09), 172 GHz (x = 0.06), and 182 GHz (x = 0). The fr value decreased with Al-substitution, consistent with the behavior of the Hc value (Figure 6, lower right). The observed electromagnetic wave absorption due to zero-field ferromagnetic resonance at exceptional high frequencies was achieved by the large Ha value of this series with large coercivity.

Figure 8.

A schematic diagram of the terahertz time domain spectroscopy measurement system.

Figure 9.

Electromagnetic wave absorption spectra of ε-AlxFe2–xO3 (left). Red, orange, yellow, green, light blue, and blue lines are the absorption spectra for x = 0.40, 0.30, 0.21, 0.09, 0.06, and 0, respectively. Right is the zero-field ferromagnetic resonance frequency (fr) versus x value plot.

3.4. Temperature dependence of zero-field ferromagnetic resonance in Al-substituted ε-Fe2O3

Among the ε-AlxFe2–xO3 samples discussed in the previous section, we focused on the x = 0.06 sample and measured the temperature dependence of zero-field ferromagnetic resonance. The Al3+ substitution ratios of each Fe site in ε-Al0.06Fe1.94O3 are 3%, 0%, 0%, and 11% for A, B, C, and D site, respectively. The magnetic properties of ε-Al0.06Fe1.94O3 are shown in Figure 10. The field-cooled magnetization curve under 10 Oe external magnetic field showed a TC value of 496 K and a cusp at 131 K (= Tp). The cusp in the magnetization is due to the spin reorientation phenomenon, which is known to occur in this temperature region [11,12]. The magnetization versus external magnetic field curve exhibited an Hc value of 19.1 kOe at 300 K.

Figure 10.

Magnetic properties of ε-Al0.06Fe1.94O3. (a) Field-cooled magnetization curve under an external magnetic field of 10 Oe. (b) Magnetization versus external magnetic field curve at 300 K.

For the THz-TDS measurement, ε-Al0.06Fe1.94O3 powder sample was pressed into a pellet-form. The absorption spectra at different temperatures are shown in Figure 11a. These absorption spectra versus frequency were obtained by calibration of the background noise. They were also fitted by Lorentz function. At 301 K, the fr value was 172 GHz, consistent with the result in the previous section. With the decrease of temperature, the fr value gradually increased to 186 GHz at 204 K, and turned to an abrupt decrease down to 147 GHz at 77 K. The fr value continued to decrease with lowering the temperature, and at 21 K, the fr value was 133 GHz (Figure 11b). Temperature dependence was also observed in the linewidth of the absorption spectra. The full width at half maximum (Δf) value increased from 5 GHz at 301 K to 19 GHz at 77 K with decreasing temperature, and then, decreased to 16 GHz at 21 K (Figure 11c).

Temperature dependencies of magnetic hysteresis loop and ac magnetic susceptibility was studied in order to understand the anomalous temperature dependencies of fr and Δf. As mentioned in Figure 10a, the field-cooled magnetization curve shows an increase below TC, but a cusp appears at Tp = 131 K, where the magnetization turns to a decrease. The Hc value increased from 19.4 kOe at 300 K to 22.6 kOe at 200 K, and then decreased to 4.5 kOe at 70 K with the decrease of temperature. The Hc versus temperature plot indicates a sigmoid decrease in a wide temperature range of 200–60 K centered at Tp (i.e., ±70 K from the center temperature, Tp = 131 K) (Figure 12a). In other words, the beginning and ending temperatures of the spin reorientation are about 200 K and 60 K, respectively, with decreasing temperature. The temperature region of the sigmoid decrease of fr almost corresponds to the temperature range of the spin reorientation. The sigmoid increase of Δf was also observed in the spin reorientation temperature region. Figure 12b is the ac magnetic susceptibility versus temperature with frequency of 10 Hz under field amplitude of 1 Oe. As the temperature decreased, the real part of the ac magnetic susceptibility (χ’) gradually increased to a maximum value of 3.8 × 10–4 emu/g Oe at 60 K and then decreased. The imaginary part (χ’’) showed similar temperature dependence with a maximum around 70 K. These temperature dependencies of ac magnetic susceptibility correspond to that of Δf [39,40].

Figure 11.

a) Electromagnetic wave absorption spectra of ε-Al0.06Fe1.94O3 at different temperatures. Black lines and red lines indicate the observed spectra and fitted Lorentz function. (b) Temperature dependence of zero-field ferromagnetic resonance frequency (fr). (c) Temperature dependence of full width at half maximum (Δf) of the absorption spectra. Dotted lines are to guide the eye.

Figure 12.

a) Temperature dependence of coercive field (Hc). Dotted line is to guide the eye. (b) Temperature dependence of ac magnetic susceptibility (real part χ’ and imaginary part χ’’) measured at 10 Hz and 1 Oe field amplitude.

As mentioned previously, the fr value is proportional to the Ha value, and in this case with randomly oriented samples, fr is also related to the Hc value. Therefore, the observed anomalous temperature dependence of fr in ε-Al0.06Fe1.94O3 was understood by the temperature dependence of Hc. The sigmoid decrease centered at Tp originates from the disappearance of magnetic anisotropy due to the spin reorientation phenomenon [11–13].

Advertisement

4. Conclusion

In this chapter, a rare phase of diiron trioxide, ε-Fe2O3, and its Al-substituted series were introduced. The synthesis, crystal structure, and its exceptional physical properties were discussed, especially its huge magnetic anisotropy exhibiting a gigantic coercive field, which enables electromagnetic wave absorption due to zero-field ferromagnetic resonance at high frequencies in the millimeter wave region. Al-substitution effect was observed in the ε-AlxFe2–xO3 series, widely controlling the magnetic properties and the zero-field ferromagnetic resonance frequency: ε-AlxFe2–xO3 absorbed millimeter waves from 112–182 GHz at room temperature. Temperature dependence of zero-field ferromagnetic resonance was also investigated for ε-Al0.06Fe1.94O3 sample, and an anomalous behavior was observed due to spin reorientation phenomenon.

Since ε-AlxFe2–xO3 is composed of very common and low costing elements, it is friendly to the environment and can be economically produced. Its chemical stability is also an advantage in the viewpoint of industrial applications, such as electromagnetic wave absorbers in the near future, where high-frequency millimeter waves are likely to be used in order to transport heavy data at high speed.

Acknowledgement

The present research was supported partly by the Core Research for Evolutional Science and Technology (CREST) program of the Japan Science and Technology Agency (JST), a Grant-in-Aid for Young Scientists (S) from Japan Society for the Promotion of Science (JSPS), DOWA Technofund, the Asahi Glass Foundation, Funding Program for Next Generation World-Leading Researchers from JSPS, a Grant for the Global COE Program “Chemistry Innovation through Cooperation of Science and Engineering”, Advanced Photon Science Alliance (APSA) from the Ministry of Education, Culture, Sports, Science and Technology of Japan (MEXT), the Cryogenic Research Center, The University of Tokyo, and the Center for Nano Lithography & Analysis, The University of Tokyo, supported by MEXT Japan. M. Y. is grateful to Advanced Leading Graduate Course for Photon Science (ALPS) and JSPS Research Fellowships for Young Scientists. A. N. is grateful to JSPS KAKENHI Grant Number 24850004 and Office for Gender Equality, The University of Tokyo. We are grateful to Dr. S. Sakurai of The University of Tokyo. We also thank Prof. M. Nakajima and Prof. T. Suemoto for support in THz-TDS measurements, Mr. Y. Kakegawa and Mr. H. Tsunakawa for collecting the TEM images, and Mr. K. Matsumoto, Mr. M. Goto, Mr. S. Sasaki, Mr. T. Miyazaki, and Mr. T. Yoshida of DOWA Electronics Materials Co., Ltd. for the valuable discussions.

References

  1. 1. ZhouZ. G.Magnetic Ferrite Materials. Beijing: Science Press; 1981
  2. 2. YoshidaS.SatoM.SugawaraE.ShimadaY.Permeability and Electromagnetic-Interference Characteristics of Fe-Si-Al Alloy Flakes-Polymer Composite. J. Appl. Phys. 199985846364638
  3. 3. NaitoY.SuetakeK.Application of Ferrite to Electromagnetic Wave Absorber and its CharacteristicsIEEE Trans. Microwave Theory Tech. 1971MT1916572
  4. 4. YusoffA. N.AbdullahM. H.AhmedS. H.JusohS. F.MansorA. A.Hamid SAA. Electromagnetic and Absorption Properties of Some Microwave AbsorbersJ. Appl. Phys. 2002922876882
  5. 5. AdamJ. D.DavisL. E.DionneG. F.SchloemannE. F.StitzerS. N.Ferrite Devices and MaterialsIEEE Trans. Microwave Theory Tech. 2002503721737
  6. 6. FedericiJ.MoellerL.Review of Terahertz and Subterahertz Wireless CommunicationsJ. Appl. Phys. 2010111101 EOF
  7. 7. KhatibM.Wireless Communications. Rijeka: Intech; 2011
  8. 8. JinJ.OhkoshiS.HashimotoK.Giant Coercive Field of Nanometer-Sized Iron Oxide.Adv. Mater. 20041614851
  9. 9. OhkoshiS.SakuraiS.JinJ.HashimotoK.The Addition Effects of Alkaline Earth Ions in the Chemical Synthesis of e-Fe2O3 Nanocrystals that Exhibit a Huge Coercive Field. J. Appl. Phys. 2005K312.
  10. 10. JinJ.HashimotoK.OhkoshiS.Formation of Spherical and Rod-Shaped e-Fe2O3 Nanocrystals with a Large Coercive Field. J. Mater. Chem. 2005151910671071
  11. 11. SakuraiS.JinJ.HashimotoK.OhkoshiS.Reorientation Phenomenon in a Magnetic Phase of e-Fe2O3 Nanocrystal. J. Phys. Soc. Jpn. 200574719461949
  12. 12. KurmooM.RehspringerJ.HutlovaA.D.OrleansC.VilminotS.EstournesC.NiznanskyD. Formation of Nanoparticles of e-Fe2O3 from Yttrium Iron Garnet in a Silica Matrix: An Unusually Hard Magnet with a Morin-Like Transition below 150 KChem. Mater. 200517511061114
  13. 13. GichM.FronteraC.RoigA.TaboadaE.MolinsE.RechenbergH. R.ArdissonJ. D.Macedo WAA, Ritter C, Hardy V, Sort J, Skumryev V, Nogués J. High- and Low-Temperature Crystal and Magnetic Structures of e-Fe2O3 and their Correlation to its Magnetic PropertiesChem. Mater. 2006181628892897
  14. 14. TucekJ.OhkoshiS.ZborilR.Room-Temperature Ground Magnetic State of e-Fe2O3: In-Field Mössbauer Spectroscorpy Evidence for Collinear Ferrimagnet. Appl. Phys. Lett. 2011
  15. 15. SakuraiS.ShimoyamaJ.HashimotoK.OhkoshiS.Large Coercive Field in Magnetic-Field Oriented e-Fe2O3 Nanorods. Chem. Phys. Lett. 2008
  16. 16. SakuraiS.TomitaK.HashimotoK.YashiroH.OhkoshiS.Preparation of the Nanowire Form of e-Fe2O3 Single Crystal and a Study of the Formation Process. J. Phys. Chem. C 2008112341309513098
  17. 17. SakuraiS.NamaiA.HashimotoK.OhkoshiS.First Observation of Phase Transformation of All Four Fe2O3 Phases (? ? e ? ß ? a-Phase). J. Am. Chem. Soc. 2009131511829918303
  18. 18. TucekJ.ZborilR.NamaiA.OhkoshiS. e-F. e.O3: An Advanced Nanomaterial Exhibiting Giant Coercive Field, Millimeter-Wave Ferromagnetic Resonance, and Magnetoelectric Coupling. Chem. Mater. 2010222464836505
  19. 19. SakuraiS.KurokiS.TokoroH.HashimotoK.OhkoshiS.Synthesis, Crystal Structure, and Magnetic Properties of e-InxFe2-xO3 Nanorod-Shaped MagnetsAdv. Funct. Mater. 2007171422782282
  20. 20. NamaiA.SakuraiS.OhkoshiS.Synthesis, Crystal Structure, and Magnetic Properties of e-GaIIIxFeIII2xO3 Nanorods. J. Appl. Phys. 2009B516.
  21. 21. YamadaK.TokoroH.YoshikiyoM.YorinagaT.NamaiA.OhkoshiS.Phase Transition of e-InxFe2xO3 Nanomagnets with a Large Thermal Hysteresis Loop. J. Appl. Phys. 2012B506.
  22. 22. OhkoshiS.KurokiS.SakuraiS.MatsumotoK.SatoK.SasakiS.A Millimeter-Wave Absorber Based on Gallium-Substituted e-Iron Oxide Nanomagnets. Angew. Chem. Int. Ed. 2007464483928395
  23. 23. NamaiA.SakuraiS.NakajimaM.SuemotoT.MatsumotoK.GotoM.SasakiS.OhkoshiS.Synthesis of an Electromagnetic Wave Absorber for High-Speed Wireless Communication.J. Am. Chem. Soc. 2009131311701173
  24. 24. NamaiA.KurahashiS.HachiyaH.TomitaK.SakuraiS.MatsumotoK.GotoT.OhkoshiS.High Magnetic Permeability of e-GaxFe2xO3 Magnets in the Millimeter Wave Region. J. Appl. Phys. 2010A955.
  25. 25. NakajimaM.NamaiA.OhkoshiS.SuemotoT.Ultrafast Time Domain Demonstration of Bulk Magnetization Precession at Zero Magnetic Field Ferromagnetic Resonanace Induced by Terahertz Magntic Field. Opt. Express 201018171826018268
  26. 26. AfsarM. N.KorolevK. A.NamaiA.OhkoshiS.Measurements of Complex Magnetic Permeability of Nano-Size e-AlxFe2-xO3 Powder Materials at Microwave and Millimeter Wavelengths. IEEE Trans. Magn. 2012481127692742
  27. 27. NamaiA.KurahashiS.GotoT.OhkoshiS.Theoretical Design of a High-Frequency Millimeter Wave Absorbing Sheet Composed of Gallium Substituted e-Fe2O3 Nanomagnet. IEEE Trans. Magn. 2012481143864389
  28. 28. YoshikiyoM.NamaiA.NakajimaM.SuemotoT.OhkoshiS.Anomalous Behavior of High-Frequency Zero-Field Ferromagnetic Resonance in Aluminum-Substituted e-Fe2O3. J. Appl. Phys. 2012A726.
  29. 29. NamaiA.YoshikiyoM.YamadaK.SakuraiS.GotoT.YoshidaT.MiyazakiT.NakajimaM.SuemotoT.TokoroH.OhkoshiS.Hard Magnetic Ferrite with a Gigantic Coercivity and High Frequency Millimetre Wave RotationNat. Commun. 2012
  30. 30. MöllerK.KoblerJ.BeinT.Collidal Suspensions of Nanometer-Sized Mesoporous Silica. Adv. Funct. Mater. 2007174605612
  31. 31. HanedaK.MiyakawaC.KojimaH.Preparation of High-Coercivity BaFe12O19J. Am. Ceram. Soc. 1974578354357
  32. 32. OhkoshiS.NamaiA.SakuraiS.The Origin of Ferromagnetism in e-Fe2O3 and e-GaxFe2-xO3 Nanomagnets. J. Phys. Chem. C 2009113261123511238
  33. 33. YoshikiyoM.YamadaK.NamaiA.OhkoshiS.Study of the Electronic Structure and Magnetic Properties of e-Fe2O3 by First-Principles Calculation and Molecular Orbital Calculations. J. Phys. Chem. C 20121161586888691
  34. 34. MchaleJ. M.AurouxA.PerrottaA. J.NavrotskyA.Surface Energies and Thermodynamic Phase Stability in Nanocrystalline AluminasScience19972775327788791
  35. 35. WenH. L.ChenY. Y.YenF. S.HuangC. H.Size Characterization of ?- and a-Al2O3 Crystallites During Phase TransformationNanostruct. Mater. 199911189101
  36. 36. OhkoshiS.TsunobuchiY.MatsudaT.HashimotoK.NamaiA.HakoeF.TokoroH.Synthesis of a Metal Oxide with a Room-Temperature Photoreversible Phase Transition.Nat. Chem. 201027539545
  37. 37. ShannonR. D.Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and ChalcogenidesActa Cryst. 1976A32(Sep1) 751 EOF767 EOF
  38. 38. ChikazumiS.Physics of Ferromagnetism. New York: Oxford University Press; 1997
  39. 39. AlgarabelP. A.MoralA.IbarraM. R.ArnaudasJ. I.Spin Reorientation in ReCo5 Compounds: A.C. Susceptibility and Thermal ExpansionJ. Phys. Chem. Solids 1988492213222
  40. 40. MalikS. K.AdrojaD. T.MaB. M.BoltichE. B.SohnJ. G.SankarS. G.WallaceW. E.Spin Reorientation Phenomenon in Nd0.5Er1.5Fe14-xMxB (M = Al and Co), as Determined by AC Susceptibility Measurements. J. Appl. Phys. 199067945894591

Written By

Marie Yoshikiyo, Asuka Namai and Shin-ichi Ohkoshi

Submitted: 06 October 2012 Published: 31 July 2013