Open access peer-reviewed chapter

Bandgap-Engineered Iron Oxides for Solar Energy Harvesting

Written By

Munetoshi Seki

Submitted: 25 July 2017 Reviewed: 18 December 2017 Published: 26 January 2018

DOI: 10.5772/intechopen.73227

From the Edited Volume

Iron Ores and Iron Oxide Materials

Edited by Volodymyr Shatokha

Chapter metrics overview

1,421 Chapter Downloads

View Full Metrics

Abstract

Epitaxial films of Rh-substituted α-Fe2O3 were fabricated by a pulsed laser deposition technique, and their photoelectrochemical characteristics were investigated for the development of visible light-responsive photoanodes for water splitting. The photocurrent in the films upon irradiation in the visible region was significantly enhanced after Rh substitution. Moreover, a near-infrared photocurrent was clearly observed for Rh:Fe2O3 photoanodes, whereas no photoresponse could be detected for the α-Fe2O3 films. These improved photoelectrochemical properties are attributed to the increased light absorption due to the hybridization of Rh-4d states and O-2p states at the valence band maximum. Moreover, Rh substitution also strongly influences the photocarrier transport properties of the films. The electrical conductivity of Rh:Fe2O3 is higher than that for α-Fe2O3 by two orders of magnitude, which is possibly due to the extended 4d orbitals of the Rh3+ ions. Thus, the improved electrical properties may lead to an increased photocurrent by lowering the recombination rate of photogenerated carriers.

Keywords

  • solar water splitting
  • pulsed laser deposition
  • photoelectrochemical cell
  • iron oxides
  • bandgap engineering

1. Introduction

Iron oxides are well known to have various physical properties depending on their composition and crystal structures (see Table 1). They have been the subject of extensive investigation over the past decades from both fundamental and practical perspectives. For example, magnetite (Fe3O4) has been one of the most widely investigated oxides in various research fields owing to its high magnetic transition temperature (~585°C) and high spin polarization of carriers [1, 2, 3]. Numerous Fe3O4-based ferromagnetic semiconductors and related spintronics devices have been reported. Another simple iron oxide, wüstite (FeO) has attracted much attention in various fields such as Earth sciences, oxide electronics, spintronics, and chemical engineering [4, 5, 6]. Moreover, multifunctional bismuth ferrite (BiFeO3, BFO) has been of great interest owing to its potential applications in numerous room temperature multiferroic devices [7, 8, 9]. BFO is also considered to be a good candidate for use in solar energy conversion systems because of its electrical polarization-induced photovoltaic effects [10]. The triangular antiferromagnet RFe2O4 (R = Ho, Y, Yb, Lu, and In) is a multilayered oxide and was discovered in the 1970s by Kimizuka et al. [11]. RFe2O4 is composed of alternating hexagonal Fe─O and R─O layers stacked along the c-axis, and Fe2+/Fe3+ charge order occurs in the Fe─O layers below 320 K, which is followed by magnetic ordering below ~240 K [12]. Recently, a number of studies on RFe2O4 have been stimulated by the discovery of the giant magnetoelectric response in LuFe2O4 and its application to multiferroic devices is currently the subject of extensive investigations [13, 14]. A great number of investigations on the magneto-optical (MO) properties of garnet-type ferrites (R3Fe5O12) have been carried out for applications in the field of optical communications. They are currently recognized as the most promising materials in magnonics and related areas. Especially, they are widely used in ferromagnetic resonance experiments and magnon-based Bose-Einstein-condensates owing to their extremely low damping [15, 16, 17, 18]. Furthermore, there has been much interest in hexaferrites, MFe12O19 (M = Ba and Sr) [19, 20], which are commonly applied in a wide variety of data storage and recording devices. One of the most favorable characteristics of the above iron oxides is their chemical stability, and they are also nontoxic. Moreover, iron and oxygen are abundant in the Earth. These features imply that iron oxides are favorable materials for applications in environmentally friendly electronics, spintronics, and magnonics. The author focuses on α-Fe2O3 commonly referred as a hematite, which is known as a promising candidate for semiconductor photoanodes for photoelectrochemical (PEC) water splitting [21, 22, 23]. A schematic of a PEC cell is shown in Figure 1. They consist of a photoactive electrode and a metal counter electrode immersed in a suitable electrolyte solution. The photogenerated electron-hole pairs are split by the electric field in the space-charge region at the surface of photoelectrodes. Since Honda and Fujishima’s pioneered work on PEC water splitting with a TiO2 photoelectrode [24], there has been worldwide research focused on the solar generation of hydrogen as a renewable and clean energy source. Many kinds of materials including TiO2 have been investigated for their application as photoelectrodes. However, most of them are wide-gap semiconductors, and only a small fraction of the solar spectrum can be utilized by the PEC cells based on these materials. A high PEC responsivity to visible (VIS) and near-infrared (near-IR) light is required to harvest the lower energy region of the solar spectrum. From this viewpoint, α-Fe2O3 has attracted much attention because of its promising properties for application as a photoanode in a solar water splitting cell. It possesses a narrow bandgap energy (Eg) of 2.1 eV that allows for the absorption of up to 40% of solar spectrum. However, the reported efficiencies for PEC water splitting using α-Fe2O3-based photoelectrodes are significantly low. This poor PEC property of α–Fe2O3can be attributed to the short diffusion length of the photogenerated holes. For α–Fe2O3-based PEC cells, only the holes generated near the electrolyte/photoanode interface can oxidize water [25, 26]. That is, most of the photogenerated electron-hole pairs recombine before reaching the photoelectrode surface. The hematite lattice is composed of an alternating stack of Fe bilayers and O layers along c-axis as illustrated in Figure 2. Spins of Fe3+ ions within each bilayer are parallel, whereas adjacent Fe bilayers have opposite spins. 3d electrons of Fe can move by hopping via the change in the Fe2+/Fe3+ valence within the Fe bilayers, whereas the exchange of electrons between neighboring Fe bilayers is spin forbidden [27, 28, 29]. Therefore, the orientation of a highly conducting (001) plane vertical to the substrate will facilitate the collection of photogenerated carriers and suppress their recombination. The author employed a Ta-doped SnO2 (TTO) layer grown on α–Al2O3 (110) single-crystal substrates for the epitaxial growth of α–Fe2O3 films along the [110] direction. As shown in Figure 3(a), the SnO2 (101) plane matches the α–Fe2O3 (110) plane with a lattice mismatch of approximately 1.3%, which is favorable for the epitaxial growth of hematite along the [110] direction on the SnO2(101)/α–Al2O3(101) substrate [30]. Another issue regarding α-Fe2O3 concerns its low responsivity to near-IR light. It is well known that the photocurrent in α-Fe2O3 is maximized at a wavelength (λ) of ~350 nm, exhibits a significant decrease with increasing λ, and approaches zero at approximately 600 nm, corresponding to its bandgap [31]. An improvement of the PEC responsivity in VIS and near-IR regions by controlling the bandgap would be useful for solar energy harvesting. Unfortunately, there exist few reports on such bandgap engineering in α-Fe2O3. From this viewpoint, the author focused on Rh-substituted α-Fe2O3 (FRO). α-Rh2O3 has a bandgap Eg of 1.2–1.4 eV [32] and the same corundum-type crystal structure as α-Fe2O3. Therefore, the bandgap of α-Fe2O3 could be narrowed by Rh substitution in the films [33, 34]. Figure 3(b) shows a schematic of the band alignment of FRO [35, 36]. α-Fe2O3 is a charge transfer-type insulator with a bandgap between the Fe 3d state (upper Hubbard band) and the fully occupied O 2p state. In contrast, the bandgap of α-Rh2O3 originates from the ligand field splitting of the Rh 4d orbitals. The Rh 4d (t2g) band in α-Rh2O3 lies near the O 2p band, and they effectively hybridize at the valence band maximum (VBM) [32, 37, 38]. In this chapter, the PEC characteristics of FRO photoanodes fabricated using pulsed laser deposition (PLD) are discussed in association with their electronic structures.

Table 1.

Various types of iron oxide and their physical properties (ρ: electrical resistivity, TN: Néel temperature, and Eg: bandgap energy).

Figure 1.

(a) Schematic of the photoelectrochemical (PEC) cell for solar water splitting and (b) electronic band structure of the PEC cell. Adapted by permission from Springer: Correlated Functional Oxides edited by H. Nishikawa, N. Iwata, T. Endo, Y. Takamura, G. Lee, and P. Mele (2017).

Figure 2.

(a) Corundum-type crystal structure of α-Fe2O3. (b) The transport of electron in the Fe 3d band is schematically illustrated. (c) Antiferromagnetic spin coupling in α-Fe2O3. [Copyright (2014), The Japan Society of Applied Physics].

Figure 3.

(a) Top: Crystal structures of α-Fe2O3 (corundum type) and SnO2 (rutile type). Bottom: Schematic showing the in-plane atomic configuration of α-Fe2O3 and SnO2. [Copyright (2012), The Japan Society of Applied Physics].

Advertisement

2. Experimental procedures

The FRO films were grown using a PLD technique with an argon fluoride (ArF) excimer laser (λ = 193 nm). The laser pulse frequency was 5 Hz. The fluence remained constant at 1.1 J/cm2. The typical growth rate of the films was 0.5 nm/min. After deposition, the FRO films were annealed in air at 700°C to improve their crystallinity. The author employed two types of bottom electrodes, viz., TTO deposited onto an α-Al2O3 (110) substrate and polycrystalline fluorine-doped SnO2 (FTO) formed on a soda-lime glass substrate. An Fe2-xRhxO3 (x = 0.0–2.0) pellet prepared by a solid-state reaction was used as a target for PLD. The growth temperature was kept at 700 and 800°C for the FRO and TTO films, respectively. The crystal structures of the samples were confirmed using X-ray diffraction (XRD). In the PEC measurement, the I-V properties were measured using an electrochemical analyzer under the illumination of Xe lamp (500 W). Optical measurements were conducted using a Vis-UV spectrometer. X-ray photoemission spectroscopy (XPS) was performed to evaluate the structure of the valence band (VB) in the FRO films.

Advertisement

3. Crystal structures

The XRD 2 theta-omega scan of the FeRhO3 films is shown in Figure 4(a). For the as-deposited sample, broad peaks are observed at 35 and 75°, which are ascribed to the (110) and (220) reflections of corundum-type FRO, respectively. This indicates that the films grown along [110] despite their low crystalline quality. Sharp peaks appear after thermal annealing, suggesting an improvement in the crystallinity. The in-plane epitaxial relationship was evaluated to be TTO [010]//FRO [001] by in-plane XRD measurements. This result agrees with the atomic configurations in Figure 3(a) [30]. The lattice constants obey Vegard’s law, implying that Fe had been appropriately substituted with Rh. In contrast to the films deposited onto the sapphire substrates, the films deposited on the glass substrates are polycrystalline in nature, as shown in Figure 4(b) and (c).

Figure 4.

(a) XRD 2θ/ω scan of the FRO film (x = 1.0) grown on TTO/α-Al2O3(110). (b) XRD pattern of the FRO film (x = 0.2) grown on the FTO substrate. (c) Magnified image of the XRD pattern of (b). [Copyright (2012 and 2014), The Japan Society of Applied Physics].

Advertisement

4. Optical properties

Figure 5(a) shows the light absorption spectra of the films. The fundamental absorption edge of α-Fe2O3 is related to charge transfer from O 2p states to the upper Hubbard band promoted by photons [denoted by TCT in Figure 3(b)]. For films with a higher Rh content, a broadband appears at 1.5–4.5 eV that is possibly related to α-Rh2O3. The optical transition in α-Rh2O3 is unclear; its absorption edge is considered to be associated with the d-d transition of Rh3+, judging from the bandgap structure [31, 39]. Figure 5(b) shows the values of an indirect bandgap (Eg), which were derived from the Tauc relation, αhν ∝ ( − Eg)2 (α: optical absorption coefficient and : photon energy). Eg decreases as the content of Rh in the films increases according to the above discussion. Value of Eg of 2.1 and 1.2 eV were obtained for α-Fe2O3 and α-Rh2O3, respectively. These values are almost identical to those reported for polycrystalline films [40].

Figure 5.

(a) Optical absorption coefficients as a function of the wavelength for FRO films on α-Al2O3(110) substrate at 298 K. For clarity, each spectrum is offset, with a spacing scaled to the composition. The peaks of α-Fe2O3 (x = 0.0) are assigned according to Ref. 19. (b) Compositional dependence of the indirect bandgap energy Eg and the absorption coefficients α at λ = 500 and 800 nm. The photographs of α-Fe2O3 (x = 0.0) and FRO (x = 0.2) films are inset. [Copyright (2012), The Japan Society of Applied Physics].

Advertisement

5. XPS spectroscopy

The results of XPS are presented in Figure 6. In the spectra of Fe 2p core level (Figure 6(a)), main peaks are at around 710 and 723 eV and are assigned to Fe 2p2/3 and 2p1/3 orbitals of α-Fe2O3, respectively [41, 42, 43]. These core level peaks become weaker as the Rh content increases. In turn, new distinct peaks appears at approximately 310 and 315 eV, which are assigned to Rh 3d3/2 and 3d5/2 orbitals, respectively [44, 45]. As seen in Figure 6(c), the VBM of α-Fe2O3 is estimated to be 0.65 eV. In contrast, the VBM of α-Rh2O3 is located near the Fermi level (~0.0 eV). Three distinct peaks are observed in the VB spectrum of the films. The bands centered at 1, 2, and 3 eV in the VB spectrum of α-Fe2O3 are assigned to the Fe 3eg, 2tg, and 2eg orbitals, respectively [46]. The crystal field splitting energy between the Fe 3eg and 2tg orbitals was estimated to be 2.5 eV in a previous study [47], which agrees with the experimental value (2.4 eV) well. In the VB spectrum of α-Rh2O3, three distinct peaks similarly appear. Unfortunately, there are hardly any reports of VB spectra for α-Rh2O3. However, by comparing with results obtained by ultraviolet photoemission spectroscopy (UPS) for ZnRh2O4 [48, 49], the peak observed at 1 eV is attributed to the t2g orbitals of the RhO6 octahedra in α-Rh2O3. The peaks at 2 and 3 eV are assigned to the Rh 4d, 5 s, and 5p mixed states [48]. The VBM and Eg exhibit a similar dependences on the Rh content, as shown in Figure 6(d). In addition, the change in the VBM (0.7 eV) for x = 1.0 is close to the bandgap decrease (0.8 eV) for x = 1.0. From these results, we can conclude that the bandgap decrease by Rh substitution is due to the hybridization of the O 2p valence band with the Rh t2g band at the VBM.

Figure 6.

XPS spectra of (a) Fe 2p and (b) Rh 3d core levels for FRO films. (c) Valence band XPS spectra of FRO films. The inset shows the enlarged VB spectra near the Fermi level. (d) Compositional dependence of VBM and bandgap energy. [Copyright (2012), The Japan Society of Applied Physics].

Advertisement

6. Photoelectrochemical properties

The current-potential curves of the films are shown in Figure 7(a) and (b). For α-Fe2O3, the photocurrent is 2.87 μA/cm2 at 0.5 V under irradiation with VIS light (λ = 400–700 nm). As shown in Figure 7(a), the VIS photocurrent is remarkably increased after Rh substitution (17.3 μA/cm2 at 0.5 V for x = 0.2). The effect of Rh substitution becomes more obvious with near-IR irradiation (λ = 700–900 nm). As shown in Figure 7(b), for x = 0.2, a near-IR photocurrent is evidently observed, whereas a photocurrent is hardly detected for the α-Fe2O3 film. These improved PEC properties of the FRO films may be caused by the increased light absorption in these wavelength regions. Furthermore, the electrical properties of the films also affect their PEC performance. The electrical conductivity for x = 0.2 (σ = 3.8 × 10−4 Ω−1 cm−1 at 300 K) is significantly larger than that for x = 0.0 (σ = 2.6 × 10−6 Ω−1 cm−1 at 300 K). This is possibly due to the extended profile of the Rh 4d state [50]. Thus, the improved electrical conductivity possibly causes an increased photocurrent by lowering the recombination rate of the photocarriers as in the case for Ti- or Si-doped α-Fe2O3 [22, 38, 40]. In Figure 7(c), the spectra of the incident photon-to-current efficiency (IPCE) are shown. The IPCE was estimated using the following relationship: IPCE (%) = 100 × [hc/e] × I/[P × λ], where h, c, e, I(mA/cm2), and P(mW/cm2) denote the Planck constant, the light velocity, the elementary charge, the photocurrent, and the power of the illumination per unit area, respectively [22]. The IPCE for x = 0.1 and 0.2 is much higher than that of α-Fe2O3 in the 340–850 nm wavelength region. For α-Fe2O3, the IPCE decreases to zero when the wavelength exceeds 610 nm, corresponding to its bandgap. On the other hand, for x = 0.2, the IPCE is 2.35% at 610 nm and gradually decreased to 0.11% at 850 nm as shown in the inset of Figure 7(c). The IPCE decreases drastically when x exceeds 0.2 as shown in Figure 7(c), and a photocurrent is hardly detected for x ≥ 0.75 in the 340–900 nm wavelength region. These results possibly reflect the drastic change in the optical transition process caused by Rh substitution. On the one hand, the photogenerated carriers in α-Fe2O3 diffuse through the bands related to the Fe 3d and O 2p states [22]. On the other hand, the recombination probability of the carriers generated in α-Rh2O3 following the d-d transition in Rh3+ is significantly high [32, 51]. This nature impairs the PEC performance for a higher Rh content. We note that that the rate of decrease in Eg is drastically decreased when x exceeds 0.2. This result suggests that the band-edge electronic structure is not strongly influenced by Rh substitution; therefore, the d-d transition predominantly occurs for x > 0.2. The IPCE peak wavelength shifts from 350 to 430 nm as Rh content increases from x = 0.0 to 0.2. This is a desirable feature for energy harvesting, because the peak of the solar spectrum is at ~475 nm. The PEC properties of the polycrystalline film (x = 0.2) are shown in Figure 7(d). The photocurrent of the single-crystal FRO grown on a TTO/Al2O3 (110) substrate is higher than that of the polycrystalline FRO grown on FTO glass. This result can be explained by the electronic transport properties of the films. The conductivity of the (110)-oriented single-crystalline film along the out-of-plane direction (σ = 3.4 × 10−4 Ω−1 cm−1) is much larger than that of the polycrystalline film (σ = 8.8 × 10−6 Ω−1 cm−1). The improvement in the electrical conductivity in the single-crystal films may accelerate the collection of photocarriers, resulting in their enhanced PEC properties. In Figure 7(e), the Mott-Schottky plots are shown. The density of donors N is expressed as follows:

N=2/e0ε0εd1/C2/dV1E1

Figure 7.

Chopped I-V curves under illumination with (a) VIS light (λ = 400–700 nm, 100 mW/cm2) and (b) NIR light (λ = 700–900 nm, 640 mW/cm2) for x = 0.0 and 0.2. (c) IPCE as a function of wavelength for the FRO films at 0.55 V vs. Ag/AgCl in an aqueous electrolytic solution containing 1.0 M NaOH. The inset shows the magnified IPCE spectra at λ = 600–900 nm (d) I-V curves of α-Fe2O3 films deposited on TTO/α-Al2O3(110) (blue) and FTO glass (black) substrates. (e) Mott-Schottky plots for FRO films. [Copyright (2012 and 2014), The Japan Society of Applied Physics].

where e0 represents an electron charge, and ε0 and ε are the vacuum and relative electric permittivities, respectively. By employing the reported value of ε = 80 for hematite, the donor densities are calculated to be 4.2 × 1017 (x = 0.2) and 5.3 × 1017 cm−3 for x = 0.0 (hematite) and 0.2, respectively. Thus, the donor density does not significantly change after Rh substitution, in contrast to the case for Ti- or Si-doped α-Fe2O3. It is considered that the number of Fe2+ ions, which acts as electron donors, increases when Fe3+ is substituted with Ti4+ or Si4+ owing to the charge neutrality. In contrast, in the FRO films, the valence of the Rh ions is +3, and hence, the content of Fe2+ does not increase after Rh substitution. Nevertheless, the conductivity for x = 0.2 is two orders of magnitude larger than that for x = 0.0. Therefore, it is considered that the carrier mobility is remarkably increased after Rh substitution probably owing to the extended nature of the Rh 4d states.

Advertisement

7. Summary

The Rh-substituted α-Fe2O3 photoelectrodes were successfully fabricated using a pulsed laser deposition method. Their bandgap narrowed as the Rh content increased. XPS analyses revealed that the bandgap narrowing is brought by the hybridization of the Rh 4d state with the O2p–Fe 3d states at the VBM of α-Fe2O3. The photocurrent was significantly enhanced for lower Rh contents. Moreover, the PEC properties were improved by the control of crystal growth condition. These results will be utilized in the development of high-efficiency solar energy conversion devices based on iron oxides.

Advertisement

Acknowledgments

This work was supported by JSPS Core-to-Core Program, A. Advanced Research Networks, and JSPS KAKENHI grant numbers JP16K14226 and JP15H03563. The author would like to thank Prof. H. Tabata, Prof, H. Matsui, and Dr. H. Yamahara for their support and helpful discussion.

References

  1. 1. Bibes M, Berthelemy A. Oxide Spintronics. IEEE Transactions on Electron Devices. 2007;54:1003-1023. DOI: 10.1109/TED.2007.894366
  2. 2. Zhang Z, Satpathy S. Electron states, magnetism, and the Verwey transition in magnetite. Physical Review B. 1991;44:13319-13331. DOI: 10.1103/PhysRevB.44.13319
  3. 3. Seki M, Hossain AKM, Kawai T, Tabata H. High-temperature cluster glass state and photomagnetism in Zn and Ti-substituted NiFe2O4 films. Journal of Applied Physics. 2005;97:083541(1)-(6). DOI: 10.1063/1.1863422
  4. 4. Seki M, Takahashi M, Adachi M, Yamahara H, Tabata H. Fabrication and characterization of wüstite-based epitaxial thin films: p-Type wide-gap oxide semiconductors composed of abundant elements. Applied Physics Letters. 2014;105:113105(1)-(4). DOI: 10.1063/1.4896316
  5. 5. Bowen HK, Adler D, Auker BH. Electrical and optical properties of FeO. Journal of Solid State Chemistry. 1975;12:355-359. DOI: 10.1016/0022-4596(75)90340-0
  6. 6. Srinivasan G, Seehra MS. Variation of magnetic properties of FexO with nonstoichiometry. Journal of Applied Physics. 1984;55:2327-2329. DOI: 10.1063/1.333651
  7. 7. Wang J, Neaton JB, Zheng H, Nagarajan V, Ogale SB, Liu B, Viehland D, Vaithyanathan V, Scholm DG, Waghmare UV, Spaldin NA, Rabe KM, Wutting M, Ramesh R. Epitaxial BiFeO3 multiferroic thin film heterostructures. Science. 2003;299:1719-1722. DOI: 10.1126/science.1080615
  8. 8. Polomska M, Kaczmarek W, Pajak Z. Electric and magnetic properties of (Bi1−xLax)FeO3 solid solutions. Physica Status Solidi A. 1974;23:567-574. DOI: 10.1002/pssa.2210230228
  9. 9. Sonowska I, Neumaier TP, Steichele E. Spiral magnetic ordering in bismuth ferrite. Journal of Physics C: Solid State Physics. 1982;15:4835-4846. DOI: 10.1088/0022-3719/15/23/020
  10. 10. Yang SY, Martin LW, Byrnes SJ, Conry TE, Basu SR, Paran D, Reichertz L, Ihlefeld J, Adamo C, Melville A, Chu Y-H, Yang C-H, Musfeldt JL, Sholm DG, Ager III JW, Ramesh R. Photovoltaic effects in BiFeO3. Applied Physics Letters. 2009;062909(1)-(3). DOI: 10.1063.1.3204695
  11. 11. Kimizuka N, Katsura T. Standard free energy of formation of YFeO3, Y3Fe5O12, and a new compound YFe2O4 in the Fe-Fe2O3-Y2O3 system at 1200°C. Journal of Solid State Chemistry. 1975;13:176-181. DOI: 10.1016/0022-4596(75)90116-4
  12. 12. Ikeda N, Ohsumi H, Ohwada K, Ishii K, Imai T, Kakurai K, Murakami Y, Yoshii K, Mori S, Horibe Y, Kito H. Ferroelectricity from ion valence ordering in the charge frustrated system LuFe2O4. Nature. 2005;436:1136-1138. DOI: 10.1038/nature04039
  13. 13. Xiang HJ, Whangbo M-H. Charge order and the origin of giant magnetocapacitance in LuFe2O4. Physical Review Letters. 2007;98:246403(1)–(4). DOI: 10.1103/PhysRevLett.98.246403
  14. 14. Seki M, Konya T, Inaba K, Tabata H. Epitaxial thin films of InFe2O4 and InFeO3 with two-dimensional triangular lattice structures grown by pulsed laser deposition. Applied Physics Express. 2010;3:105801(1)-(3). DOI: 10.1143/APEX.3.105801
  15. 15. Kumar N, Kim NG, Park YA, Hur N, Jung JH, Ham KJ, Yee KJ. Epitaxial growth of terbium iron garnet thin films with out-of-plane axis of magnetization. Thin Solid Films. 2008;516:7753-7757. DOI: 10.1016/j.tsf.2008.05.032
  16. 16. Krockenberger Y, Matsui H, Hasegawa T, Kawasaki M, Tokura Y. Solid phase epitaxy of ferromagnetic Y3Fe5O12 garnet thin films. Applied Physics Letters. 2008;93:092505(1)-(3). DOI: 10.1063/1.2976747
  17. 17. Yamahara H, Mikami M, Seki M, Tabata H. Epitaxial strain-induced magnetic anisotropy in Sm3Fe5O12 thin films grown by pulsed laser deposition. Journal of Magnetism and Magnetic Materials. 2011;323:3143-3146. DOI: 10.1016/j.jmmm.2011.06.074
  18. 18. Adachi M, Seki M, Yamahara H, Tabata H. Long-term potentiation of magnonic synapses by photocontrolled spin current mimicked in reentrant spin-glass garnet ferrite Lu3Fe5-2xCoxSixO12 thin films. Applied Physics Express. 2015;91:085118(1)-(7). DOI: 10.7567/APEX.8.043002
  19. 19. Xu P, Han X, Wang M. Synthesis and magnetic properties of BaFe12O19 Hexaferrite nanoparticles by a reverse microemulsion technique. Journal of Physical Chemistry C. 2007;111:5866-5870. DOI: 10.1021/jp068955c
  20. 20. Zi ZF, Sun YP, Zhu XB, Yang ZR, dai JM, Song WH. Structural and magnetic properties of SrFe12O19 hexaferrite synthesized by a modified chemical co-precipitation method. Journal of Magnetism and Magnetic Materials. 2008;320:2746-2751. DOI: 10.1016/j.jmmm.2008.06.009
  21. 21. Chernomordik BD, Russell HB, Cvelbar U, Jasinski JB, Kumar V, Dautsch T, Sunkara MK. Photoelectrochemical activity of as-grown, α-Fe2O3 nanowire array electrodes for water splitting. Nanotechnology. 2012;23:194009(1)-(9). DOI: 10.1088/0957-4484/23/19/194009
  22. 22. Kay A, Cesar I, Grätzel M. New benchmark for photooxidation by nanostructured α-Fe2O3 films. Journal of American Chemical Socety. 2006;128:15714-15721. DOI: 10.1021/ja0643801
  23. 23. Thimsen E, Formal FL, Grätzel M, Warren SC. Influence of plasmonic Au nanoparticles on the photoactivity α-Fe2O3 electrodes for water splitting. Nano Letters. 2011;11:35-43. DOI: 10.1021/nl1022354
  24. 24. Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972;238:37-38. DOI: 10.1038/238038a0
  25. 25. Sanchez C, Sieber KD, Somorjai GA. The photoelectrochemistry of niobium doped α-Fe2O3. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 1988;252:269-290. DOI: 10.1016/0022-0728(88)80216-X
  26. 26. Campbell AS, Schwertmann U, Stanjek H, Friedl J, Kyek A, Campbell PA. Si incorporation into hematite by heating Si-Fe ferrihydrite. Langmuir. 2002;18:7804-7809. DOI: 10.1021/la011741w
  27. 27. Benjelloun D, Bonnet J-P, Doumerc J-P, Launay J-C, Onillon M, Hagenmuller P. Anisotropie des proprieties electriques de I’oxyde de fer Fe2O3α. Materials Chemistry and Physics. 1984;10:503-518. DOI: 10.1016/0254-0584(84)90001-4
  28. 28. Nakau T. Electrical conductivity of α-Fe2O3. Journal of the Physical Society of Japan. 1960;15:727-727. DOI: 10.1143/JPSJ.15.727
  29. 29. Iordanova N, Dupuis M, Rosso KM. Charge transport in metal oxides: A theoretical study of hematite alpha-Fe2O3. Journal of Chemical Physics. 2005;122:144305(1)-(10). DOI: 10.1063/1.1869492
  30. 30. Kwon J-H, Choi Y-H, Kim DH, Yang M, Jang J, Kim TW, Hong S-H, Kim M. Orientation relationship of polycrystalline Pd-doped SnO2 thin film deposits on sapphire substrates. Thin Solid Films. 2008;517:550-553. DOI: 10.1016/j.tsf.2008.06.074
  31. 31. Ling Y, Wang G, Wheeler DA, Zhang JZ, Li Y. Sn-doped hematite nanostructures for photoelectrochemical water splitting. Nano Letters. 2011;11:211-2125. DOI: 10.1021/nl200708y
  32. 32. Koffyberg JP. Optical bandgaps and electron affinities of semiconducting Rh2O3(I) and Rh2O3(III). Journal of Physics and Chemistry of Solids. 1992;53:1285-1288. DOI: 10.1016/0022-3697(92)90247-B
  33. 33. Seki M, Yamahara H, Tabata H. Enhanced photocurrent in Rh-substituted α-Fe2O3 thin films grown by pulsed laser deposition. Applied Physics Express. 2012;5:115801(1)-(3). DOI: 10.1143/APEX.5.115801
  34. 34. Seki M, Takahashi M, Ohshima T, Yamahara H, Tabata H. Solid-liquid-type solar cell based on α-Fe2O3 heterostructures for solar energy harvesting. Japanese Journal of Applied Physics. 2014;53:05FA07(1)–(6). DOI: 10.7567/JJAP.53.05FA07
  35. 35. Mashiko H, Oshima T, Otomo A. Band-gap narrowing in α-(CrxFe1-x)2O3 solid-solution films. Applied Physics Letters. 2011;99:241904(1)–(3). DOI: 10.1063/1.3669704
  36. 36. Moore EA. First-principles study of the mixed oxide α-FeCrO3. Physical Review B. 2007;76:195107(1)–(7). DOI: 10.1103/PhysRevB.76.195107
  37. 37. Uemoto K, Wentzcovitch RM. Effect of the d electrons on phase transitions in transition metal sesquioxides. Physics and Chemistry of Minerals. 2011;38:387-395. DOI: 10.1007/s00269-010-0412-1
  38. 38. Sartoretti CJ, Alexander BD, Solarska R, Rutkowska IA, Augustynski J, Cerny R. Photochemical oxidation of water at transparent ferric oxide film electrodes. Journal of Physical Chemistry B. 2005;109:13685-13692. DOI: 10.1021/jp051546g
  39. 39. Beatham N, Orchard AF, Thornton G. X-ray and UV photoelectron spectra of the metal sesquioxides. Journal of Physics and Chemistry of Solids. 1981;42:1051-1055. DOI: 10.1016/0022-3697(81)90129-3
  40. 40. Tang H, Matin MA, Wang H, Deutsh T, Al-Jassim M, Turner J, Yan Y. Synthesis and characterization of titanium-alloyed hematite thin film for photoelectrochemical water splitting. Journal of Applied Physics. 2011;110:123511(1)–(7). DOI: 10.1063/1/3671414
  41. 41. Temesghen W, Sherwood P. Analytical utility of valence band X-ray photoelectron spectroscopy of iron and its oxides, with spectral interpretation by cluster and band structure calculations. Analytical and Bioanalytical Chemistry. 2002;373:601-608. DOI: 10.1007/s00216-002-1362-3
  42. 42. Fujii T, Groot FMF, Sawatzky GA, Voogt FC, Hibma T, Okada K. In situ XPS analysis of various iron oxide films grown by NO2-assisted molecular beam epitaxy. Physical Review B. 1999;59:3195-3202. DOI: 10.1103/PhysRevB.59.3195
  43. 43. McIntyre NS, Zetaruk DG. X-ray photoelectron spectroscopic studies of iron oxides. Analytical Chemistry. 1977;49:1521-1529. DOI: 10.1021/ac50019a016
  44. 44. Abe Y, Kato K, Kawamura M, Sakai K. Rhodium and rhodium oxide thin films characterized by XPS. Surface Science Spectra. 2001;8:117-125. DOI: 10.1116/11.20010801
  45. 45. Blomberg S, Lundgren E, Westerström R, Erdogan E, Martin NM, Mikkelsen A, Andersen JN. Structure of the Rh2O3 (0001) surface. Surface Science. 2012;606:1416-1421
  46. 46. Uekawa N, Kaneko K. Nonstoichiometric properties of nanoporous iron oxide films. Journal of Physical Chemistry B. 1998;102:8719-8724. DOI: 10.1021/jp982249x
  47. 47. Tossell JA, Vaughan DJ, Johnson KH. Electronic structure of ferric iron octahedrally coordinated to oxygen. Nature Physical Science. 1973;244:42-45. DOI: 10.1038/physci244042a0
  48. 48. Mizoguchi H, Hirano M, Fujitsu S, Takeuchi T, Ueda K, Hosono H. ZnRh2O4: A p-type semiconducting oxide with a valence band composed of a low spin state of Rh3+ in a 4d6 configuration. Applied Physics Letters. 2002;80:1207-1209. DOI: 10.1063/1.1450252
  49. 49. Dekkers M, Rijnders G, Blank DHA. ZnIr2O4, a p-type transparent oxide semiconductor in the class of spinel zinc-d6-transition metal oxide. Applied Physics Letters. 2007;90:021903(1)-(3). DOI: 10.1063/1.2431548
  50. 50. Lee YS, Lee JS, Noh TW, Byun DY, Yoo KS, Yamaura K, Takayama-Muromachi E. Systematic trends in the electronic structure parameters of the 4d transition-metal oxides SrMO3 (M = Zr, Mo, Ru, and Rh). Physical Review B. 2003;67:113101(1)–(4). DOI: 10.1103/PhysRevB67.113101
  51. 51. Walter MG, Warren L, McKone JR, Boettcher SW, Mi Q, Santori EA, Lewis NS. Solar water splitting cells. Chemical Reviews. 2010;110:6446-6473. DOI: 10.1021/cr1002326

Written By

Munetoshi Seki

Submitted: 25 July 2017 Reviewed: 18 December 2017 Published: 26 January 2018