Skip Navigation
Skip to contents

Journal of Powder Materials : Journal of Powder Materials

OPEN ACCESS
SEARCH
Search

Articles

Page Path
HOME > J Powder Mater > Volume 31(6); 2024 > Article
Critical Review
Epsilon Iron Oxide (ε-Fe2O3) as an Electromagnetic Functional Material: Properties, Synthesis, and Applications
Ji Hyeong Jeong, Hwan Hee Kim, Jung-Goo Lee, Youn-Kyoung Baek*
Journal of Powder Materials 2024;31(6):465-479.
DOI: https://doi.org/10.4150/jpm.2024.00290
Published online: December 31, 2024

Nano Materials Research Division, Korea Institute of Materials Science, 797 Changwondaero, Changwon 51508, Republic of Korea

*Corresponding Author: Youn-Kyoung Baek, TEL: +82-55-280-3605, FAX: +82-55-280-3391, E-mail: ykbaek@kims.re.kr
• Received: September 19, 2024   • Revised: November 20, 2024   • Accepted: December 5, 2024

© The Korean Powder Metallurgy & Materials Institute

This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/4.0/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

  • 105 Views
  • 6 Download
next
  • Iron oxide (ε-Fe2O3) is emerging as a promising electromagnetic material due to its unique magnetic and electronic properties. This review focuses on the intrinsic properties of ε-Fe2O3, particularly its high coercivity, comparable to that of rare-earth magnets, which is attributed to its significant magnetic anisotropy. These properties render it highly suitable for applications in millimeter wave absorption and high-density magnetic storage media. Furthermore, its semiconducting behavior offers potential applications in photocatalytic hydrogen production. The review also explores various synthesis methods for fabricating ε-Fe2O3 as nanoparticles or thin films, emphasizing the optimization of purity and stability. By exploring and harnessing the properties of ε-Fe2O3, this study aims to contribute to the advancement of next-generation electromagnetic materials with potential applications in 6G wireless telecommunications, spintronics, high-density data storage, and energy technologies.
Iron oxide (Fe2O3) exists in a variety of structural polymorphs, each of which is utilized in a wide range of applications due to its distinct physical and chemical properties[16]. Representative polymorphs of Fe2O3, include the alpha (α), beta (β), gamma (γ), and epsilon (ε) phases, with their distinctive properties summarized in Table 1 [4, 5, 710]. α-Fe2O3 (hematite), the most thermodynamically stable phase, has a corundum structure (space group = R3¯c) and is widely utilized in pigments and photocatalysts. β-Fe2O3, a metastable phase with a bixbyite structure (space group = Ia3¯), has potential applications in advanced energy materials [11, 12] and environmental technologies[13, 14]. γ-Fe2O3 (maghemite), which adopts a spinel structure (space group = P412121), is primarily utilized in the form of nanoparticles as a soft magnetic material for magnetic recording media and biomedical applications. Among the polymorphs, epsilon iron oxide (ε-Fe2O3) is distinguished by its orthorhombic structure (space group = Pna21) and its entirely unique electromagnetic properties, which recently garnered significant attention as a potential electromagnetic functional material (Fig. 1) [3, 5].
ε-Fe2O3 exhibits unique magnetic and electronic properties including extremely high magnetic anisotropy and coercivity (Hc) comparable to or exceeding rare earth-based magnets [6, 15]. These properties enable the utilization of ε-Fe2O3 in advanced magnetic applications such as millimeter-wave absorbers, high-density magnetic recording media, and spintronics [3, 16, 17]. Additionally, ε-Fe2O3 also exhibits semiconductor properties with a bandgap that allows it to absorb visible light, suggesting its potential as a photocatalyst for hydrogen production [9]. These characteristics suggest that ε-Fe2O3 is a versatile material capable of performing various functions beyond simple magnetic materials. However, to maximize the electromagnetic functionality of the iron oxides, it is essential to it is essential to synthesize high purity ε phase. As ε-Fe2O3 is an intermediate phase between α-Fe2O3 and γ-Fe2O3, it is only appears in the form of nanoparticles with diameters ranging from 8 to 50 nm, which requires more demanding synthesis conditions than other polymorphs [6]. Recent advances in wet/dry powder preparation and advanced deposition techniques have enabled ε-Fe2O3 materials to be stably prepared not only as nanoparticles but also as nanofilms, greatly expanding their potential for electromagnetic applications [18, 19].
In this paper, we will discuss the electromagnetic properties of ε-Fe2O3 among the many polymorphs of Fe2O3, and review the latest synthesis methods and applications to fully utilize its properties. Through this, we will revisit the scientific and technological significance of ε-Fe2O3 and investigate its potential as a next-generation electromagnetic functional material.
2.1. Structure and Properties of ε-Fe2O3
Among the polymorphs of iron oxide, ε-Fe2O3 exhibits distinct structural and magnetic properties that set it apart as a promising material for advanced applications. Notably, ε-Fe2O3 is characterized by its unique orthorhombic crystal structure, where Fe ions have four sites: three octahedral sites (FeAO6, FeBO6, FeCO6) and one tetrahedral site (FeDO4) located at the center of the crystal, with a spin of S=5/2[5].
Molecular-field theory calculations suggest that the magnetism of ε-Fe2O3 at room temperature arises from super-exchange interactions between Fe ions mediated by O. The spins of FeB and FeC are aligned upward along the a-axis, while FeA and FeD exhibit Neel P-type ferrimagnetism, with their spins aligned downward along the a-axis [3, 20]. ε-Fe2O3 exhibits a coercivity (Hc) of over 2 T at room temperature[6], which is fourfold greater than of conventional M-type ferrites and nearly twice that of NdFeB. This ultra-high coercivity can be explained by two factors. First, ε-Fe2O3 is only formed in a specific range of particle size conditions (8 nm to 50 nm), which corresponds to the single magnetic domain particle size that exhibits the maximum coercivity [5]. The actual iron oxide crystal phase undergoes a phase transition in the sequence of γ → ε → β → α phase depending on the size of the powder. This transition is driven by the relationship of the free energies of the nanopowders as shown in (Eq. 1).
(1)
Gi=GBi+6VmdGsi
Where G(i) is the free energy of the nanopowder, GB(i) is the free energy of the bulk state, Vm is the molar volume, d is the diameter of the nanopowder, GS(i) is the surface free energy, and i represents the phase type. Since the magnitudes of bulk and surface free energies vary across different crystalline phases (GB(γ)>GB(ε)>GB(β)>GB(α), GS(γ)<GS(ε)<GS(β)<GS(α), the relationship between the free energy G(i) of the nanopowder and the diameter (d) of the powder is given by (Fig. 2) [3]. Specifically, within a certain particle size range, the free energy G(ε) of the ε crystalline phase is the lowest, indicating that the ε-phase forms with high purity in the single magnetic domain particle size range (8 nm to 50 nm). Furthermore, the high coercivity of ε-Fe2O3 is attributed to its magnetic anisotropy fields (Ha). Ha is proportional to the crystal magnetic anisotropy constant (K1) and inversely proportional to the saturation magnetization (Ms) value as expressed in (Eq. 2).
(2)
Ha=2K1Ms
The crystal magnetic anisotropy constant is a physical parameter that quantifies the anisotropy of energy in the magnetization direction, which depends on the crystal structure of the material. A higher value indicates greater stability of the magnetization direction. The crystal magnetic anisotropy constant (K1) of epsilon iron oxide is calculated to be 2×106 erg/cm3, which is significantly larger than the K1 value of γ-Fe2O3 (~ 104 erg/cm3) and the K1 value of α-Fe2O3 (~ 105 erg/cm3) [5]. This large K1 of ε-Fe2O3 has been reported to be due to the strong orbital hybridization of Fe and O [20]. Generally, Fe3+ has a d5 electron configuration, resulting in zero orbital angular momentum (L). However, in the case of ε-Fe2O3, the orbital hybridization of Fe and O alters the Fe ion’s electron configuration to d5+q, instead of the standard d5 electron configuration. This increases the orbital angular momentum of Fe, leading to stronger spin-orbit coupling and consequently, a high coercive force [21].
Similar to other iron oxide polymorphs ε-Fe2O3 exhibits distinct band gap characteristics and semiconductor properties. The valence band in ε-Fe2O3 primarily consists of O2p orbitals, while the conduction band is dominated by Fe3d orbitals. Density of states (DOS) analysis from previous studies[21] shows that in ε-Fe2O3, the occupied Fe3d band located between -8.3 and -6.6 eV, the O2p band spans from -6.3 to -0.8 eV, and the unoccupied Fe3d band extends from +0.7 to +3.2 eV relative to the Fermi level (Fig. 3(a)) [21]. The resulting band gap is estimated to be 1.6-1.9 eV for ε-Fe2O3, compared to the 2.0 to 2.2 eV band gap of α-Fe2O3, as shown in Table. 1. This band structure characterizes ε-Fe2O3 as a charge-transfer insulator, where electronic transitions involve electron transfer from O2p to Fe3d states [21, 22]. Such charge-transfer transitions enable efficient electron excitation under visible light, making ε-Fe2O3 suitable for photocatalytic applications. Density functional theory (DFT) calculations further reveal that the heterostructure of ε-Fe2O3 and α-Fe2O3 forms a type-III broken band-gap alignment, with the band edges of α-Fe2O3 positioned higher in energy than those of ε-Fe2O3 [22]. This alignment facilitates spontaneous spatial separation of photogenerated charge carriers, with electrons transferring from α-Fe2O3 to ε-Fe2O3 and holes moving in the opposite direction (Fig. 3(b)) [22]. Such charge separation reduces electron-hole recombination, enhancing photocatalytic efficiency compared to using a single material. These electronic and magnetic properties highlight the potential of ε-Fe2O3 as a multifunctional material. Its tunable electromagnetic properties enable diverse applications, as explored in subsequent sections.
2.2. Fabrication Methods for ε-Fe₂O3
The unique crystal structure and excellent electromagnetic properties of ε-Fe2O3 show that it has significant potential as an electromagnetic functional material. However, unlike α-Fe2O3, which is abundant in nature, and γ-Fe2O3, which is easily synthesized, ε-Fe2O3 is strongly influenced by the surface free energy, and the ε crystalline phase is only forms under conditions that maintain a particle morphology with diameters in tens of nanometers. In 2004, high purity epsilon iron oxide was artificially synthesized by the Ohkoshi group using the reverse micelle sol-gel method [6]. In this process, surfactants such as cetyltrimethylammonium bromide (CTAB) and cetyltrimethylammonium chloride (CTAC) were used above the critical micelle concentration to create reverse micelles in oil. Then, a water-soluble iron salt solution was added to trap Fe ions in the micelles and disperse them in a lipophilic solvent, followed by the addition of silica precursors to initiate a sol-gel reaction [17, 23, 24]. This process involved emulsifying Fe ions with surfactants, trapping them in a silica matrix, which restricted the growth of iron oxide particles during subsequent heat treatment, leading to the formation of nanometer-sized ε crystalline phases. However, the use of surfactants and oils to create reverse micelle, combined with 12-hour sol-gel reaction is inefficient in terms of yield and cost. Accordingly, efforts have been made to reduce the overall reaction time of the silica sol-gel reaction to 2 hours, though a multi-step batch process [15]. Additionally, studies have been conducted on synthesizing ε-Fe2O3 using cost-effective raw materials and scalable production methods. The clay mineral nontronite, an iron-rich silicate, is inexpensive, readily available, and requires no additional pretreatment, facilitating its use in low-cost synthesis [25]. Similarly, the ball milling process, a scalable mechanochemical technique for producing fine powders, has been utilized to synthesize ε-Fe2O3 [26]. While both methods effectively produce the ε-Fe2O3, they may have limitations related to secondary phase formation and elemental substitution for Fe.
Recently, a study reported the continuous preparation of ε-Fe2O3 nanoparticles using a spray-drying method although this method only produced micrometer-sized powders (Fig. 4) [27]. In this approach, a water-soluble iron salt and silica precursor are homogeneously mixed to form a clear solution, which is continuously injected through a nozzle, and atomized in a high-temperature reactor. As the solution dries instantaneously, the silica forms a micrometer-sized xerogels, within which the Fe precursor is trapped, resulting in a composite powder. During subsequent heat treatment, when Fe ions inside the silica xerogel combine with oxygen to transition to the iron oxide crystalline phase. The silica matrix restricts particle growth, allowing the final particle size to be reduced 50 nm or less, leading to high purity (92.7%) formation of the ε crystalline phase. The aerosol method has been a widely used continuous manufacturing process that has been widely used in the industrial preparation of fine powders for pharmaceuticals, food, and chemicals, can produce powders in a single step without additional washing. Moreover, the method is versatile in combining different compounds into particles and allows the synthesis of composites containing dispersed Fe and Si elements without the use of surfactants. Therefore, it is expected that a commercialized technology capable of producing pure ε-phase at high yields will soon be realized. These advancements in cost-effective and scalable synthesis methods for ε-Fe2O3 hold significant potential for applications requiring ferromagnetic nanoparticles with high magnetic anisotropy. In particular, these nanoparticles are anticipated to find extensive use in fields such as magnetic recording media and electromagnetic wave absorbers, as further detailed in the following sections.
The ε-Fe2O3 crystalline phase, which occurs on the scale of a few to several tens of nanometers, has recently actively researched for the fabrication of thin films and their application in spintronic devices [28]. Spintronics devices, such as spin valves, magnetic tunnel junctions, and spin current devices, are integral to magnetoresistive random access memory (MRAM), a spintronic memory based on these devices, has been utilized as a non-volatile memory [2932]. The operation principle of MRAMs is based on the tunnel magnetoresistance (TMR) effect, where the spin of electrons tunnels through an insulating layer located between two ferromagnetic layers, causing a change in electrical resistance depending on magnetization direction of the two layers. The change in resistance is used to store data as 0 and 1, and the magnetization direction is retained even when the power is turned off, ensuring non-volatility. One of the ferromagnetic layers has a fixed magnetization direction, requiring an anisotropic material, while the other is composed of a soft magnetic material to allow its magnetization direction to change in response to an external magnetic field or current[30]. In addition, to improve data stability and the efficiency of reading and writing, thin film fabricated in the form of single crystal rather than polycrystals, where magnetization directions may be distributed in multiple directions are more suitable for application in magnetic recording device.
Although ε-Fe2O3 exhibits high magnetic anisotropy, its crystalline phase, which only appears in nanoparticle form, has not been suitable for application in substrate-based devices. In previous studies, researchers attempted to immobilize a mixture of ε-Fe2O3 nanorods composites onto wafers using chemical vapor deposition (CVD) [33], but a significant reduction in magnetic properties was observed compared to the nanoparticle form. This reduction was speculated to result from decreased ordering of the crystal structure [18]. Therefore, it is crucial to study the preparation of ε-Fe2O3 in thin film form while preserving its unique magnetic properties, and various preparation methods have been tried. Gich group used pulsed laser deposition (PLD) to form single crystal thin films, stabilizing the ε crystalline phase through epitaxial modification on SrTiO3 single crystal substrates, or by using a GaFeO3 buffer layer with the same structure as ε-Fe2O3 to promote the growth of ε-Fe2O3 thin films. These thin film exhibited a coercivity of about 8 kOe [34]. Also, in 2017, Corbellini group successfully achieved the growth of (001)-oriented ε-Fe2O3 epitaxial thin films on single crystal zirconia (YSZ) (100) substrates using PLD (Fig. 5) [18]. This study is the result of growing ε-Fe2O3 thin films directly on the YSZ substrate, which are commonly used to grow oxides on silicon wafers in the (100) orientation. This suggests the potential for directly growth of ε-Fe2O3 thin films on silicon wafers, an important milestone for expanding applications to large-area devices. The Curie temperature of the prepared thin film was 460 K, showed a similar level to the Curie temperature observed in the nanoparticle form (approximately 490 K). However, the PLD has limitations, including the needs for high deposition temperature (700-900 K) and lattice-matched substrate. For the practical application of ε-Fe2O3 thin films, it is essential to develop methods that can form thin films at lower temperatures or to explore new fabrication technique compatible with a wider range of substrates.
Recently, a method for fabricating high-purity ε-Fe2O3 thin films without a lattice-matched substrate at a low temperature of 280°C using the atomic layer deposition (ALD) has been devised [19]. ALD is a chemical vapor deposition method that uses chemical precursors under relatively low vacuum (10−2 to 10 mbar) and low temperature (< 400 °C), enabling atomically precise thickness control and fine composition control. The iron oxide thin films prepared by ALD are ε crystalline phase, with only 2.5% α crystalline phase impurities. As shown in the M-H curve (Fig. 6) [35], there was no kink resulting from soft magnetic phase (typically γ-Fe2O3) admixture, confirming the formation of high-purity ε-Fe2O3 thin films. Furthermore, 57Fe Mössbauer measurements identified the four distinct Fe sites inherent to the ε crystalline phase. While the low-temperature transition observed in nanoparticles between 110 K and 150 K was shifted to the higher temperature range in the thin film, the transition was still present. This shift in transition temperature, characteristic of the thin film, was attributed to grain boundaries and substrate-induced strain. The use of ALD for ε-Fe2O3 thin film holds significant potential for broader applications, as it does not require a lattice-matched substrate. The development represents an advancement in expanding the application range of ε-Fe2O3. Furthermore, it is expected to play a critical role in advanced technological applications, such as photocatalytic devices that require its semiconducting properties and efficient charge carrier transport, as discussed in subsequent sections.
3.1 Millimeter-Wave Absorption Applications
The high magnetic anisotropy of ε-Fe2O3 has attracted attention for its potential application as a millimeter wave (30-300 GHz) absorber and is currently under active investigation. Electromagnetic wave absorbers are classified into conduction loss, dielectric loss, and magnetic loss materials, which attenuate the incident electromagnetic waves. Among these, magnetic loss materials offer the advantage of more effective electromagnetic wave absorption in high frequency bands, due to magnetic loss caused by the imaginary component of permeability, allowing them to selectively absorb electromagnetic signals in specific frequency bands. This phenomenon is attributed to the ferromagnetic resonance of magnetic materials, where ferromagnetic materials in a magnetic field selectively absorb specific frequencies. When a ferromagnet is exposed to electromagnetic waves, its spins precess around the magnetization axis, absorbing electromagnetic waves at a frequency that matches the precession. The natural resonance frequency (fr) is proportional to the anisotropic field (Ha) as shown in the equation (3) below [35], where v is the gyromagnetic constant.
(3)
fr=v2π×Ha
Therefore, to adjust the frequency band to be absorbed, the magnetic anisotropy of the magnetic material must be adjusted, which can be achieved through control of the material composition. ε-Fe2O3 has been reported to exhibit ferromagnetic resonance absorption at approximately 180 GHz due to its high Ha [17]. The crystalline magnetic anisotropy of ε-Fe2O3 can also be controlled by substituting transition metal element at the Fe sites, and studies have been conducted to adjust the absorption band accordingly (Fig. 7(a)-(b)) [17, 36]. It has been reported that substitution Fe3+ (S = 5/2) with non-magnetic ion such as Ga3+ and Al3+ (S = 0) can modulate fr over a wide range by substituting FeD or FeA sites. When substituted with Ga (ε-GaxFe2-xO3) and Al (ε-AlxFe2-xO3), it has been reported that fr can be adjusted from 35 GHz (x = 0.67) to 147 GHz (x = 0.10)[17] and from 182 GHz (x = 0.04) to 112 GHz (x = 0.4), respectively, depending on the content (x = 0.10 ~ 0.67) and (x = 0.04 ~ 0.4)[36]. Conversely, fr was found to increase when the spin-orbit interaction was enhanced by substituting Rh3+, which has a larger orbital angular momentum than Fe3+ [37]. With Rh substitution, Hc increased up to 3.5 T [16], and for ε-Rh0.14Fe1.86O3, a Hc of 2.7 T and fr of 209 GHz were observed (Fig. 7(c)) [37].
In addition, complex substitution of various elements on ε-Fe2O3 is also being actively studied. A representative example is the co-substitution of Co2+(S=3/2), Ti4+ and Ga3+ (ε-GaxTiyCoyFe2-x-yO3). This combination is notable because the large magnetic anisotropy in the a-axis direction, caused by the orbital angular momentum of Fe3+ at the FeB site, is counteracted by the magnetic anisotropy of Co2+, which substitutes for FeD along the c-axis direction (Fig. 7(d)-(e)) [38]. As a result, the crystalline magnetic anisotropy is significantly reduced, lowering the coercivity to below 3 kOe [38]. In addition, to utilize this material as a more efficient millimeter-wave absorber, a study was conducted in which ceramic nanopowders with high conductivity (Ti4O7) were applied to achieve both dielectric loss and magnetic loss effect [16, 38]. This resulted in a high dielectric constant, as electron transfer on the Ti4O7 surface was blocked at the iron oxide nanoparticle interface, inhibiting the current flow. When prepared as a thin film (216 ㎛), it exhibited 99.8% absorption capacity (RL = -27.2 dB) at 80 GHz with a broadband width of 16.2 GHz (Fig. 8) [39]. ε-Fe2O3 can serve as an effective absorber in the millimeter-wave band due to a large magnetic anisotropy, and the absorption band can be tuned by controlling the substitutional composition. This demonstrates its potential as a multifunctional absorber capable of responding to wide band of electromagnetic waves, making it suitable for 6G mobile communication applications.
3.2 High-Density Magnetic Recording Media Applications
In the era of big data, the need for high-density magnetic recording media capable of reliably storing and managing large amounts of data over long periods become increasingly important [4043]. Currently, magnetic recording tapes are mainly composed of magnetic materials such as spindle Co-Fe alloy nanoparticles or barium ferrite, which play an crucial role in long-term data storage across various fields such as insurance, finance, broadcasting, and web services [4447]. To achieve high-density recording, it is essential to reduce the size of the magnetic filler particles used in these tapes. However, the reduction in size lead to several side effects. In the case of metallic alloys, a decrease in particle size increases the surface area, which raises the potential for oxidation and can result in pyrophoric character. On the other hand, iron oxides such as barium ferrite and magnetite do not exhibit pyrophoric character, but as their particle size decreases, they may lose magnetic ordering, potentially leading to a transition to superparamagnetism. As a solution to this issue, ε-Fe2O3 particles have been found to have a superparamagnetic limit at a particle size of 7.5 nm, allowing them to maintain a smaller size compared to barium ferrite (20-25 nm) [48]. This indicates that ε-Fe2O3 has significant potential as an ultra-high-density magnetic recording media.
Furthermore, magnetic particles in the form of multiple magnetic domains, rather than single magnetic domain are likely to have their spins only partially aligned due to their magnetic anisotropy. While this partial alignment does not affect data playback but it raises concerns about potential noise generation. Therefore, ε-Fe2O3 with its low superparamagnetic limit, well-maintained magnetic arrangement can be synthesized with single magnetic domain sphere sizes ranging from 8 to 30 nm, making it a suitable candidate for ultra-high-density magnetic recording media. However, ε-Fe2O3 exhibits a high coercivity of more than 20 kOe at room temperature, which can decrease the efficiency of information storage if used in its original form. To address this issue, research has been conducted to adjust the coercivity by substituting Fe with transition metal elements. Since magnetic fillers suitable for magnetic recording media typically require a coercivity of around 3 kOe, the Ohkoshi group has worked on modifying the coercivity by substituting element such as Ga, Ti and Co [16, 38, 49, 50].
In particular, the Ga0.31Ti0.05Co0.05Fe1.59O3 case exhibits a magnetization value of 23.4 emu/g at 7 T, which represents a 44% increase from the original ε-Fe2O3 value of 16.2 emu/g. This increase is attributed to the substitution of Fe3⁺ ions (S = 5/2) with Ga3⁺ (S = 0), Ti4⁺ (S = 0), and Co2⁺ (S = 3/2) at rates of 48%, 10%, and 10%, respectively. The total magnetization of ε-Fe2O3 is determined by the sum of the positive sublattice magnetization (MB, MC > 0) at the FeB and FeC sites and the negative sublattice magnetization (MA, MD < 0) at the FeA and FeD sites. Ga substitution reduces the negative sublattice magnetization at FeD, thereby increasing the overall magnetization value. Furthermore, the coercivity (Hc) was adjusted to below 2.69 kOe through element substitution. This reduction in Hc is attributed to the compensating effect of the single-ion magnetic anisotropy of Fe3⁺ and Co2⁺. First-principles calculations indicate the ε-Fe2O3 crystal phase originally exhibits a strong magnetic anisotropy along the a-axis, which arises from the orbital angular momentum resulting from the hybridization between Fe3⁺ and O2⁻ions at the FeB site. However, the substitution of Co2⁺ ions at FeD sites introduces magnetic anisotropy along the c-axis, which appears to offset the magnetic anisotropy along the a-axis, thereby reducing the coercivity to below 3 kOe [38]. The resulting ε-Fe2O3 nanoparticles were dispersed in a urethane resin and vinyl chloride polymer and then coated on a non-magnetic polyethylene (PE) film while an external magnetic field was applied (Fig. 9) [38]. During this process, the ε-Fe2O3 nanoparticles were oriented vertically, forming a coated magnetic tape. Compared to commercial CoFe alloys from Dowa Electric Materials, the magnetic recording tape produced exhibited a sharper power spectrum and a higher signal-to-noise ratio (S/N) than existing commercial products. This indicates the high potential of ε-Fe2O3 nanoparticles as a magnetic recording media. This study demonstrates that magnetic recording media utilizing the properties of ε-Fe2O3 can outperform current commercial materials, highlighting its potential application as a next-generation high-density magnetic recording media.
3.3 Visible light catalytic Applications
Solar hydrogen production is an environmentally friendly method to produce hydrogen by utilizing clean, renewable energy sources. It has gained significant attention as a key technology for building a sustainable hydrogen economy. The main approaches include photoelectrochemical (PEC) water splitting, photocatalytic water splitting, and photo-reforming using oxygenated organic compounds (OOCs), all of which aim to minimize greenhouse gas emissions and maximize energy efficiency. Recently, research has focused on the potential of ε-Fe2O3, which can absorb visible light.
Currently, PEC water splitting has attracted considerable attention from researchers, and it has been found that ferroelectric materials can play an important role as PEC catalysts. This is due to the internal electric field induced by their spontaneous electric polarization, which can enhance the separation of electron-hole pairs generated by sunlight[51-53]. Conventional ferroelectrics such as BaTiO3 and Pb(Zrx,Ti1-x)O3 have bandgaps greater than 3 eV, making them unsuitable for use in the visible light range. In contrast, BiFeO3 possesses a relatively narrow bandgap (≈2.2–2.8 eV) and exhibits the largest spontaneous polarization (100 μC cm-2) [54]. However, the solar energy conversion efficiency of BiFeO3 remains limited due to the rapid recombination rate of photoexcited charge carriers[50]. On the other hand, Fe2O3 is a well-known photocatalyst for solar hydrogen production due to its visible light absorption (bandgap ≈ 2.1 eV), non-toxicity, and low cost [55-58]. However, its primary drawbacks low electrical conductivity, hole diffusion length of 2-4 nm, and low absorption coefficient have resulted in solar-to-hydrogen conversion values that are lower than theoretically predicted, with photocurrent reaching up to 12 mA cm-2 [59, 60]. To overcome the limitations of these single photocatalysts and enhance solar energy conversion efficiency, a study on nanocomposite photocatalysts based on heterojunction of ε-Fe2O3 and BiFeO3 was published[61]. In this study, a photoelectrode for solar water splitting was fabricated based on a heterostructure in which BiFeO3 nanopillars were vertically embedded into an ε-Fe2O3 matrix thin film using the PLD (Fig. 10) [61]. By optimizing the photodegradation performance through the adjustment of the ratio between the two materials, the BiFeO3-ε-Fe2O3 photoelectrode containing 9% ε-Fe2O3 exhibited more than double the photocurrent (0.19 mA cm-2) compared to the single-material sample. Additionally, it maintained high stability under continuous light irradiation for 25 hours. This result can be explained by the charge separation effect caused by band-bending between the two materials. When the two materials are contacted, free charge carriers are redistributed across the interface until equilibrium is achieved, leading to the realignment of the Fermi (Fig. 10(g)-(h)) [61]. Since the work function of BiFeO3 (approximately 4.95 eV) is smaller than that of ε-Fe2O3 (approximately 5.72 eV) [61-63], electrons migrate from BiFeO3 to ε-Fe2O3. This migration forms a charge region at the heterointerface and induces band-bending, which results in hole accumulation (electron deficiency) at the BiFeO3 interface. Due to this band-bending, photoexcited electrons can easily migrate from BiFeO3 to ε-Fe2O3, while holes can migrate from ε-Fe2O3 to BiFeO3, significantly inhibiting electron-hole recombination. In other words, ε-Fe2O3, with a bandgap corresponding to visible light, can be utilized as a photocatalytic material. Furthermore, it is a promising material for enhancing photocurrent by inducing charge separation when combined with existing materials.
ε-Fe2O3 has also been actively studied as a photocatalyst for the photoreforming of oxygen-containing organic compounds such as ethanol, glycerol, and glucose [9]. This approach to hydrogen production via biomass photoreforming offers significant environmental benefits by simultaneously removing pollutants and processing waste from the biomass industry [64-66]. Although α-Fe2O3 has traditionally been used as visible light catalyst due to its low fabrication cost and ease of processing into various nanostructures, its performance has been constrained by factors such as a low absorption coefficient, high recombination losses of charge carriers, and a short diffusion length of charge carriers [2, 67]. To retain the advantages of α-Fe2O3 while overcoming its disadvantages, ε-Fe2O3 and β-Fe2O3 polymorphs of Fe2O3 were prepared by chemical vapor deposition (CVD) and their performance was compared in this study. The catalytic activity of the Fe2O3 polymorphs increased in the order of α-Fe2O3 < ε-Fe2O3 < β-Fe2O3 with the average hydrogen production rate of ε-Fe2O3 reaching 125 mmol h-1 m-2, which was significantly higher than that of α-Fe2O3 [9]. Additionally, ε-Fe2O3 demonstrated better photocorrosion resistance than α-Fe2O3, indicating its ability to maintain photocatalytic activity over an extended period. Stable hydrogen generation was observed for at least 20 hours of continuous irradiation, demonstrating the robustness and reliability of ε-Fe2O3 as a photocatalyst. This study shows that ε-Fe2O3 is a promising alternative material that can overcome the limitations of α-Fe2O3 and make an important contribution to the development of high-performance photocatalysts.
Epsilon iron oxide (ε-Fe2O3) has attracted attention for various advanced applications, including millimeter-wave absorbers, high-density magnetic storage media, and photocatalysis, due to its unique magnetic and electronic properties. Specifically, its high magnetic anisotropy and coercivity set ε-Fe2O3 apart from conventional magnetic materials, enhancing its potential as an electromagnetic wave absorber in the millimeter-wave band and as a high-density magnetic storage media. In addition, its semiconducting property of visible light absorption suggests potential applications as a catalyst for solar hydrogen production. To maximize the electromagnetic functionality of ε-Fe2O3, it is crucial to synthesize it in high purity and stable form. Various synthesis methods have been developed, successfully yielding nanoparticles and thin films. Advances in the continuous manufacturing process of ε phase nanoparticles, which are particularly challenging to prepare, are expected to further promote the practical applications of ε-Fe2O3 across various applications.
Future research will need to focus on optimizing the properties of ε-Fe2O3 for each application and developing structural design and fabrication techniques to enhance performance. For example, magnetic property optimization and composite structuring to enhance absorption in high-frequency band, control of thin film structure for spintronics device applications, and design of heterostructure to improve efficiency as photocatalysts will become a key area of research. With continued research, ε-Fe2O3 is expected to become a core material in various industries including 6G wireless communication, spintronic devices, high-density magnetic storage, and hydrogen production, serving as a next-generation electromagnetic functional material.

Funding

This research was supported by the Basic Research Program (PNK9960) of Korea Institute of Materials Science.

Conflict of Interest Declaration

The authors declare no relevant conflicts of interest.

Author Information and Contribution

Jihyeong Jeong : Writing a manuscript and researching previous studies, Hwanhee Kim and Jung-Goo Lee: data organization and researching previous studies, Youn-Kyoung Baek: Supervision, reviewing and editing.

Acknowledgement

None.

Fig. 1.
(a) Graphical representations of the crystal structures of α-Fe2O3, ε-Fe2O3, and γ-Fe2O3. Adapted with permission from [5] Copyright 2009 American Chemical Society. (b) Crystallographic structure of the ε-Fe2O3 phase represented by the cation polyhedral. Adapted with permission from [3] Copyright 2010 American Chemical Society.
jpm-2024-00290f1.jpg
Fig. 2.
Stability of individual polymorphs of Fe2O3 based on the calculated dependence of the free energy per volume (i.e., G/V) on the size (d) of the iron(III) oxide nanoparticles of a particular polymorph. Reproduced with permission from [3] Copyright 2010 American Chemical Society.
jpm-2024-00290f2.jpg
Fig. 3.
(a) Density of states (DOS) of ε-Fe2O3, with the total DOS (black), iron DOS (red), and oxygen DOS (blue). Adapted with permission from [21] Copyright 2012 American Chemical Society, (b) Schematic diagram of band bending and the charge flow at the interface of ε-Fe2O3 and α-Fe2O3 heterostructure after connection. Adapted with permission from [22] Copyright 2020 Royal Society of Chemistry.
jpm-2024-00290f3.jpg
Fig. 4.
(a) Schematic illustration of the preparation of ε-Fe2O3 nanoparticles via spray drying. (b) and (c) Scanning electron microscopy (SEM) images of as-spray-dried precursor particles with a precursor molar ratio (Fe/Si) of 0.4:1. (d) and (e) Field-emission transmission electron microscopy (FE-TEM) images of ε-Fe2O3 nanoparticles embedded in SiO2 particles after annealing at 1180°C for 4 h. (f) and (g) TEM images of the corresponding ε-Fe2O3 NPs after SiO2 removal. Reproduced with permission from [27] Copyright 2022 Royal Society of Chemistry.
jpm-2024-00290f4.jpg
Fig. 5.
(a) STEM image of an approximately 100-nm-thick film of ε-Fe2O3 on YSZ (100) highlighting the formation of pillar-like twins, (b) details of the interface between the substrate and the film, evidencing the formation of “bubbles” of a foreign phase (most likely Fe3O4) at the interface. Reproduced with permission from [18] Copyright 2017 Nature.
jpm-2024-00290f5.jpg
Fig. 6.
(a) Magnetization hysteresis loop at 300 K of ε-Fe2O3/AlFeO3//Nb:STO(111). (b) Ferroelectric hysteresis loop measured at 300 K and 10 Hz with 100 ms of delay time. Reproduced with permission from [34] Copyright 2014 Wiley-VCH.
jpm-2024-00290f6.jpg
Fig. 7.
Absorption spectra of (a) ε-GaxFe2-xO3. Adapted with permission from [17]. Copyright 2007 Wiley-VCH. (b) ε-AlxFe2-xO3. Adapted with permission from [36]. Copyright 2008 American Chemical Society. (c) ε-RhxFe2-xO3. Adapted with permission from [37]. Copyright 2012 Nature. Magnetic structure of (d) ε-Fe2O3 and (e, left) ε -Ga0.31Ti0.05Co0.05Fe1.59O3. Red and blue arrows denote the sublattice magnetizations of Fe3+ and Co2+, respectively. Black arrows show the total magnetization. (e, Right) The direction of the single-ion anisotropies (K, red and blue arrows) for Fe3+ at the B site along a-axis and Co2+ at the D site along c-axis. Adapted with permission from [38] Copyright 2016 Wiley-VCH.
jpm-2024-00290f7.jpg
Fig. 8.
(a) Schematic illustration of phase matching for absorption, (b) photograph, and (c) observed RL spectrum of the flexible millimeter-wave–absorbing ultrathin film composed of the Ti4O7@ε-Ga0.21Ti0.05Co0.05Fe1.69O3 composite and acrylic acid ester polymer on a copper foil. Adapted with permission from [39] Copyright 2021 Wiley-VCH.
jpm-2024-00290f8.jpg
Fig. 9.
(a) Photograph of the manufactured magnetic recording tape composed of ε-Ga0.31Ti0.05Co0.05Fe1.59O3. (b) Schematic illustration of the LTO-3 AMR head. (c) Cross-section SEM image of the manufactured magnetic tape. (d) Power spectrum of the signal of the ε-Ga0.31Ti0.05Co0.05Fe1.59O3 tape (red line) and cobalt–iron alloy tape (gray line) measured with the LTO-3 AMR head. Reproduced with permission from [38] Copyright 2016 Wiley-VCH.
jpm-2024-00290f9.jpg
Fig. 10.
(a) Cross-sectional TEM image of BFO-FO heterostructures on STO substrate. (b) High-resolution TEM image showing the interfaces between films and STO substrates. The corresponding FFT patterns of (c) BFO and (d) ε-FO. (e) Top view. (f) Schematic of the self-assembled BFO-FO vertical heterostructures with BFO pillars embedded in the ε-FO matrix. (g), (h) Schematic diagrams illustrating the energy band alignment and the expected charge flow at the BFO-FO heterojunction under light excitation. Adapted with permission from [61] Copyright 2016 Wiley-VCH.
jpm-2024-00290f10.jpg
Table 1.
Physical and magnetic properties of various polymorphs of iron oxide [4,5,7-10]
Properties Iron Oxide Molecular Formula
α-Fe2O3 β-Fe2O3 γ-Fe2O3 ε-Fe2O3
Structural type Corundum structure Bixbyite structure Spinel structure Orthorhombic structure
Space group R3¯ c Ia3¯ P4332 (cubic), P412121 (tetragonal) Pna21
Transition temperature TC = 956 K TN = 119 K TC = 820-986 K TC = 495 K
Type of magnetism Weak ferromagnetic or antiferromagnetic Antiferromagnetic Ferrimagnetic Ferrimagnetic
Density (g/cm3) 5.26 - 4.87 4.78
Crystallographic system Rhombohedral, hexagonal Cubic Cubic or tetrahedral Orthorhombic
Lattice parameter (Å) aRh=5.427, a=9.393 aCubic=8.3474, a=5.095,
aHex=5.034, aTetra=8.347, b=8.789,
cHex=13.75 cTetra=25.01 c=7.437
Fe site - FeA site, FeA site, FeA site,
FeB site FeB site FeB site,
FeC site,
FeD site
Particle size ≥50 nm ≤ 50nm ≤ 8nm ≤ 50nm
Band gap 2.0 – 2.2 eV 1.7 – 1.9 eV 2.3 eV 1.6 – 1.9 eV
  • 1. S. Piraman, S. Sundar, R. Mariappan, Y. Y. Kim and K. Min: Sens. Actuators, B Chem., 234 (2016) 386.Article
  • 2. K. Sivula, F. Le Formal and M. Grätzel: ChemSusChem, 4 (2011) 432.Article
  • 3. J. Tuček, R. Zbořil, A. Namai and S. I. Ohkoshi: Chem. Mater., 22 (2010) 6483.Article
  • 4. L. MacHala, J. Tuček and R. Zbořil: Chem. Mater., 23 (2011) 3255.Article
  • 5. S. Sakurai, A. Namai, K. Hashimoto and S. I. Ohkoshi: J. Am. Chem. Soc., 131 (2009) 18299.Article
  • 6. J. Jin, S. I. Ohkoshi and K. Hashimoto: Adv. Mater., 16 (2004) 48.Article
  • 7. A. S. Teja and P. Y. Koh: Prog. Cryst. Growth Charact. Mater., 55 (2009) 22.Article
  • 8. I. Ahamed, R. Pathak, R. Skomski and A. Kashyap: AIP Adv., 8 (2018) 148.
  • 9. G. Carraro, C. MacCato, A. Gasparotto, T. Montini, S. Turner, O. I. Lebedev, V. Gombac, G. Adami, G. Van Tendeloo, D. Barreca and P. Fornasiero: Adv. Funct. Mater., 24 (2014) 372.Article
  • 10. O. Malina, J. Tuček, P. Jakubec, J. Kašlík, I. Medřík, H. Tokoro, M. Yoshikiyo, A. Namai, S. I. Ohkoshi and R. Zbořil: RSC Adv., 5 (2015) 49719.Article
  • 11. C. Liu, N. Zhang, Y. Li, R. Fan, W. Wang, J. Feng, C. Liu, J. Wang, W. Hao, Z. Li and Z. Zou: Nat. Commun., 14 (2023) 1.
  • 12. G. Carraro, D. Barreca, M. Cruz-Yusta, A. Gasparotto, C. MacCato, J. Morales, C. Sada and L. Sánchez: ChemPhysChem, 13 (2012) 3798.Article
  • 13. M. M. Rahman, A. Jamal, S. B. Khan and M. Faisal: Superlattices Microstruct., 50 (2011) 369.Article
  • 14. M. Li, J. Wu and G. Shen: Catal. Sci. Technol., 12 (2022) 2659.Article
  • 15. E. Gorbachev, M. Soshnikov, M. Wu, L. Alyabyeva, D. Myakishev, E. Kozlyakova, V. Lebedev, E. Anokhin, B. Gorshunov, O. Brylev, P. Kazin and L. Trusov: J. Mater. Chem. C, 9 (2021) 6173.Article
  • 16. H. Tokoro, A. Namai and S. I. Ohkoshi: Dalt. Trans., 50 (2021) 452.ArticlePDF
  • 17. S. Ohkoshi, S. Kuroki, S. Sakurai, K. Matsumoto, K. Sato and S. Sasaki: Angew. Chem., 119 (2007) 8544.Article
  • 18. L. Corbellini, C. Lacroix, C. Harnagea, A. Korinek, G. A. Botton, D. Ménard and A. Pignolet: Sci. Rep., 7 (2017) 3712.Article
  • 19. T. Jussila, A. Philip, J. Lindén and M. Karppinen: Adv. Eng. Mater., 25 (2023) 2201262.Article
  • 20. S. I. Ohkoshi, A. Namai and S. Sakurai: J. Phys. Chem. C, 113 (2009) 11235.Article
  • 21. M. Yoshikiyo, K. Yamada, A. Namai and S. I. Ohkoshi: J. Phys. Chem. C, 116 (2012) 8688.Article
  • 22. I. Ahamed, N. Seriani, R. Gebauer and A. Kashyap: RSC Adv., 10 (2020) 27474.Article
  • 23. S. Sakurai, K. Tomita, K. Hashimoto, H. Yashiro and S. I. Ohkoshi: J. Phys. Chem. C, 112 (2008) 20212.Article
  • 24. S. Sakurai, J. I. Shimoyama, K. Hashimoto and S. I. Ohkoshi: Chem. Phys. Lett., 458 (2008) 333.Article
  • 25. K. Kelm and W. Mader: Z. Anorg. Allg. Chem., 631 (2005) 2383.Article
  • 26. Y. Zhao and G. Wen: J. Magn. Magn. Mater., 512 (2020) 167039.Article
  • 27. G. R. Jo, M. B. Yun, Y. Hun Son, B. Park, J. G. Lee, Y. G. Kim, Y. G. Son and Y. K. Baek: Chem. Commun., 58 (2022) 11442.Article
  • 28. S. M. Suturin, A. M. Korovin, S. V. Gastev, M. P. Volkov, A. A. Sitnikova, D. A. Kirilenko, M. Tabuchi and N. S. Sokolov: Phys. Rev. Mater., 2 (2018) 1.Article
  • 29. S. Bhatti, R. Sbiaa, A. Hirohata, H. Ohno, S. Fukami and S. N. Piramanayagam: Mater. Today, 20 (2017) 530.Article
  • 30. S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. Von Molnár, M. L. Roukes, A. Y. Chtchelkanova and D. M. Treger: Science, 294 (2001) 1488.Article
  • 31. S. A. Wolf, J. Lu, M. R. Stan, E. Chen and D. M. Treger: IEEE, 98 (2010) 2155.Article
  • 32. A. Fert: Rev. Mod. Phys., 80 (2008) 1517.Article
  • 33. D. Barreca, G. Carraro, D. Peeters, A. Gasparotto, C. Maccato, W. M. M. Kessels, V. Longo, F. Rossi, E. Bontempi, C. Sada and A. Devi: Chem. Vap. Depos., 20 (2014) 313.Article
  • 34. M. Gich, I. Fina, A. Morelli, F. Sánchez, M. Alexe, J. Gàzquez, J. Fontcuberta and A. Roig: Adv. Mater., 26 (2014) 4645.Article
  • 35. C. Kittel: Phys. Rev., 71 (1947) 270.Article
  • 36. A. Namai, S. Sakurai, M. Nakajima, T. Suemoto, K. Matsumoto, M. Goto, S. Sasaki and S. I. Ohkoshi: J. Am. Chem. Soc., 131 (2008) 1170.Article
  • 37. A. Namai, M. Yoshikiyo, K. Yamada, S. Sakurai, T. Goto, T. Yoshida, T. Miyazaki, M. Nakajima, T. Suemoto, H. Tokoro and S. I. Ohkoshi: Nat. Commun., 3 (2012) 1035.Article
  • 38. S. ichi Ohkoshi, A. Namai, M. Yoshikiyo, K. Imoto, K. Tamazaki, K. Matsuno, O. Inoue, T. Ide, K. Masada, M. Goto, T. Goto, T. Yoshida and T. Miyazaki: Angew. Chem. Int. Ed., 55 (2016) 11403.ArticlePDF
  • 39. R. Kinugawa, K. Imoto, Y. Futakawa, S. Shimizu, R. Fujiwara, M. Yoshikiyo, A. Namai and S. I. Ohkoshi: Adv. Eng. Mater., 23 (2021) 2001473.Article
  • 40. G. Cherubini, C. C. Chung, W. C. Messner and S. O. R. Moheimani: IEEE Trans. Control Syst. Technol., 20 (2012) 296.Article
  • 41. H. Nishio and H. Yamamoto: IEEE Trans. Magn., 46 (2010) 3747.Article
  • 42. S. I. Ohkoshi, K. Imoto, A. Namai, M. Yoshikiyo, S. Miyashita, H. Qiu, S. Kimoto, K. Kato and M. Nakajima: J. Am. Chem. Soc., 141 (2019) 1775.Article
  • 43. S. ichi Ohkoshi, M. Yoshikiyo, K. Imoto, K. Nakagawa, A. Namai, H. Tokoro, Y. Yahagi, K. Takeuchi, F. Jia, S. Miyashita, M. Nakajima, H. Qiu, K. Kato, T. Yamaoka, M. Shirata, K. Naoi, K. Yagishita and H. Doshita: Adv. Mater., 32 (2020) 139821.
  • 44. I. Tudosa, C. Stamm, A. B. Kashuba, F. King, H. C. Siegmann, J. Stöhr, G. Ju, B. Lu and D. Weeler: Nature, 428 (2004) 831.ArticlePDF
  • 45. D. Alloyeau, C. Ricolleau, C. Mottet, T. Oikawa, C. Langlois, Y. Le Bouar, N. Braidy and A. Loiseau: Nat. Mater., 8 (2009) 940.ArticlePDF
  • 46. M.L. Plumer, J. van Ek and D. Weller, The Physics of Ultra-High-Density Magnetic Recording, Springer Science and Business Media, 41(2001).
  • 47. S. U. T Hayashi, S. Hirono and M. Tomita: Nature, 381 (1996) 772.ArticlePDF
  • 48. S. I. Ohkoshi, A. Namai, K. Imoto, M. Yoshikiyo, W. Tarora, K. Nakagawa, M. Komine, Y. Miyamoto, T. Nasu, S. Oka and H. Tokoro: Sci. Rep., 5 (2015) 14414.Article
  • 49. M. Yoshikiyo, A. Namai, K. Imoto, H. Tokoro and S. I. Ohkoshi: Eur. J. Inorg. Chem., 2018 (2018) 847.ArticlePDF
  • 50. M. Yoshikiyo, Y. Futakawa, R. Shimoharai, Y. Ikeda, J. MacDougall, A. Namai and S. Ohkoshi: Chem. Phys. Lett., 803 (2022) 139821.Article
  • 51. W. Yang, Y. Yu, M. B. Starr, X. Yin, Z. Li, A. Kvit, S. Wang, P. Zhao and X. Wang: Nano Lett., 15 (2015) 7574.Article
  • 52. D. Cao, Z. Wang, L. Wen, Y. Mi and Y. Lei: Angew. Chem., 126 (2014) 11207.Article
  • 53. W. Ji, K. Yao, Y. F. Lim, Y. C. Liang and A. Suwardi: Appl. Phys. Lett., 103 (2013) 103.Article
  • 54. C. B. E., R. R. T. Zhao, A. Scholl, F. Zavaliche, K. Lee, M. Barry, A. Doran, M. P. Cruz, Y. H. Chu, C. Ederer, N. A. Spaldin, R. R. Das, D. M. Kim and S. H. Baek: Nat. Mater., 5 (2006) 823.
  • 55. Y. Ling, G. Wang, D. A. Wheeler, J. Z. Zhang and Y. Li: Nano Lett., 11 (2011) 2119.Article
  • 56. G. Wang, Y. Ling, D. A. Wheeler, K. E. N. George, K. Horsley, C. Heske, J. Z. Zhang and Y. Li: Nano Lett., 11 (2011) 3503.Article
  • 57. D. A. Wheeler, G. Wang, Y. Ling, Y. Li and J. Z. Zhang: Energy Environ. Sci., 5 (2012) 6682.Article
  • 58. A. R. H. Dotan, O. Kfir, E. Sharlin, O. Blank, M. Gross, I. Dumchin and G. Ankonina: Nat. Mater., 12 (2013) 158.ArticlePDF
  • 59. M. S. Prévot and K. Sivula: J. Phys. Chem. C, 117 (2013) 17879.Article
  • 60. K. Sivula, R. Zboril, F. Le Formal, R. Robert, A. Weidenkaff, J. Tucek, J. Frydrych and M. Grätzel: J. Am. Chem. Soc., 132 (2010) 7436.Article
  • 61. L. T. Quynh, C. N. Van, Y. Bitla, J. W. Chen, T. H. Do, W. Y. Tzeng, S. C. Liao, K. A. Tsai, Y. C. Chen, C. L. Wu, C. H. Lai, C. W. Luo, Y. J. Hsu and Y. H. Chu: Adv. Energy Mater., 6 (2016) 1600686.Article
  • 62. S.J. Clark and J. Robertson: Appl. Phys. Lett., 90 (2007) 132903.Article
  • 63. Y. Xu and M. A. A. Schoonen: Am. Mineral., 85 (2000) 543.Article
  • 64. M. Cargnello, A. Gasparotto, V. Gombac, T. Montini, D. Barreca and P. Fornasiero: Eur. J. Inorg. Chem., 28 (2011) 4309.Article
  • 65. D. I. Kondarides, V. M. Daskalaki, A. Patsoura and X. E. Verykios: Catal. Letters, 122 (2008) 26.ArticlePDF
  • 66. A. Patsoura, D. I. Kondarides and X. E. Verykios: Catal. Today, 124 (2007) 94.Article
  • 67. M. J. Katz, S. C. Riha, N. C. Jeong, A. B. F. Martinson, O. K. Farha and J. T. Hupp: Coord. Chem. Rev, 256 (2012) 2521.Article

Figure & Data

References

    Citations

    Citations to this article as recorded by  

      • PubReader PubReader
      • ePub LinkePub Link
      • Cite this Article
        Cite this Article
        export Copy Download
        Close
        Download Citation
        Download a citation file in RIS format that can be imported by all major citation management software, including EndNote, ProCite, RefWorks, and Reference Manager.

        Format:
        • RIS — For EndNote, ProCite, RefWorks, and most other reference management software
        • BibTeX — For JabRef, BibDesk, and other BibTeX-specific software
        Include:
        • Citation for the content below
        Epsilon Iron Oxide (ε-Fe2O3) as an Electromagnetic Functional Material: Properties, Synthesis, and Applications
        J Powder Mater. 2024;31(6):465-479.   Published online December 31, 2024
        Close
      • XML DownloadXML Download
      Figure
      • 0
      • 1
      • 2
      • 3
      • 4
      • 5
      • 6
      • 7
      • 8
      • 9
      Epsilon Iron Oxide (ε-Fe2O3) as an Electromagnetic Functional Material: Properties, Synthesis, and Applications
      Image Image Image Image Image Image Image Image Image Image
      Fig. 1. (a) Graphical representations of the crystal structures of α-Fe2O3, ε-Fe2O3, and γ-Fe2O3. Adapted with permission from [5] Copyright 2009 American Chemical Society. (b) Crystallographic structure of the ε-Fe2O3 phase represented by the cation polyhedral. Adapted with permission from [3] Copyright 2010 American Chemical Society.
      Fig. 2. Stability of individual polymorphs of Fe2O3 based on the calculated dependence of the free energy per volume (i.e., G/V) on the size (d) of the iron(III) oxide nanoparticles of a particular polymorph. Reproduced with permission from [3] Copyright 2010 American Chemical Society.
      Fig. 3. (a) Density of states (DOS) of ε-Fe2O3, with the total DOS (black), iron DOS (red), and oxygen DOS (blue). Adapted with permission from [21] Copyright 2012 American Chemical Society, (b) Schematic diagram of band bending and the charge flow at the interface of ε-Fe2O3 and α-Fe2O3 heterostructure after connection. Adapted with permission from [22] Copyright 2020 Royal Society of Chemistry.
      Fig. 4. (a) Schematic illustration of the preparation of ε-Fe2O3 nanoparticles via spray drying. (b) and (c) Scanning electron microscopy (SEM) images of as-spray-dried precursor particles with a precursor molar ratio (Fe/Si) of 0.4:1. (d) and (e) Field-emission transmission electron microscopy (FE-TEM) images of ε-Fe2O3 nanoparticles embedded in SiO2 particles after annealing at 1180°C for 4 h. (f) and (g) TEM images of the corresponding ε-Fe2O3 NPs after SiO2 removal. Reproduced with permission from [27] Copyright 2022 Royal Society of Chemistry.
      Fig. 5. (a) STEM image of an approximately 100-nm-thick film of ε-Fe2O3 on YSZ (100) highlighting the formation of pillar-like twins, (b) details of the interface between the substrate and the film, evidencing the formation of “bubbles” of a foreign phase (most likely Fe3O4) at the interface. Reproduced with permission from [18] Copyright 2017 Nature.
      Fig. 6. (a) Magnetization hysteresis loop at 300 K of ε-Fe2O3/AlFeO3//Nb:STO(111). (b) Ferroelectric hysteresis loop measured at 300 K and 10 Hz with 100 ms of delay time. Reproduced with permission from [34] Copyright 2014 Wiley-VCH.
      Fig. 7. Absorption spectra of (a) ε-GaxFe2-xO3. Adapted with permission from [17]. Copyright 2007 Wiley-VCH. (b) ε-AlxFe2-xO3. Adapted with permission from [36]. Copyright 2008 American Chemical Society. (c) ε-RhxFe2-xO3. Adapted with permission from [37]. Copyright 2012 Nature. Magnetic structure of (d) ε-Fe2O3 and (e, left) ε -Ga0.31Ti0.05Co0.05Fe1.59O3. Red and blue arrows denote the sublattice magnetizations of Fe3+ and Co2+, respectively. Black arrows show the total magnetization. (e, Right) The direction of the single-ion anisotropies (K, red and blue arrows) for Fe3+ at the B site along a-axis and Co2+ at the D site along c-axis. Adapted with permission from [38] Copyright 2016 Wiley-VCH.
      Fig. 8. (a) Schematic illustration of phase matching for absorption, (b) photograph, and (c) observed RL spectrum of the flexible millimeter-wave–absorbing ultrathin film composed of the Ti4O7@ε-Ga0.21Ti0.05Co0.05Fe1.69O3 composite and acrylic acid ester polymer on a copper foil. Adapted with permission from [39] Copyright 2021 Wiley-VCH.
      Fig. 9. (a) Photograph of the manufactured magnetic recording tape composed of ε-Ga0.31Ti0.05Co0.05Fe1.59O3. (b) Schematic illustration of the LTO-3 AMR head. (c) Cross-section SEM image of the manufactured magnetic tape. (d) Power spectrum of the signal of the ε-Ga0.31Ti0.05Co0.05Fe1.59O3 tape (red line) and cobalt–iron alloy tape (gray line) measured with the LTO-3 AMR head. Reproduced with permission from [38] Copyright 2016 Wiley-VCH.
      Fig. 10. (a) Cross-sectional TEM image of BFO-FO heterostructures on STO substrate. (b) High-resolution TEM image showing the interfaces between films and STO substrates. The corresponding FFT patterns of (c) BFO and (d) ε-FO. (e) Top view. (f) Schematic of the self-assembled BFO-FO vertical heterostructures with BFO pillars embedded in the ε-FO matrix. (g), (h) Schematic diagrams illustrating the energy band alignment and the expected charge flow at the BFO-FO heterojunction under light excitation. Adapted with permission from [61] Copyright 2016 Wiley-VCH.
      Epsilon Iron Oxide (ε-Fe2O3) as an Electromagnetic Functional Material: Properties, Synthesis, and Applications
      Properties Iron Oxide Molecular Formula
      α-Fe2O3 β-Fe2O3 γ-Fe2O3 ε-Fe2O3
      Structural type Corundum structure Bixbyite structure Spinel structure Orthorhombic structure
      Space group R3¯ c Ia3¯ P4332 (cubic), P412121 (tetragonal) Pna21
      Transition temperature TC = 956 K TN = 119 K TC = 820-986 K TC = 495 K
      Type of magnetism Weak ferromagnetic or antiferromagnetic Antiferromagnetic Ferrimagnetic Ferrimagnetic
      Density (g/cm3) 5.26 - 4.87 4.78
      Crystallographic system Rhombohedral, hexagonal Cubic Cubic or tetrahedral Orthorhombic
      Lattice parameter (Å) aRh=5.427, a=9.393 aCubic=8.3474, a=5.095,
      aHex=5.034, aTetra=8.347, b=8.789,
      cHex=13.75 cTetra=25.01 c=7.437
      Fe site - FeA site, FeA site, FeA site,
      FeB site FeB site FeB site,
      FeC site,
      FeD site
      Particle size ≥50 nm ≤ 50nm ≤ 8nm ≤ 50nm
      Band gap 2.0 – 2.2 eV 1.7 – 1.9 eV 2.3 eV 1.6 – 1.9 eV
      Table 1. Physical and magnetic properties of various polymorphs of iron oxide [4,5,7-10]


      Journal of Powder Materials : Journal of Powder Materials
      TOP