Open Access
4 October 2016 Photoelectrochemical and theoretical investigations of spinel type ferrites (MxFe3xO4) for water splitting: a mini-review
Dereje H. Taffa, Ralf Dillert, Anna C. Ulpe, Katharina C. L. Bauerfeind, Thomas Bredow, Detlef W. Bahnemann, Michael Wark
Author Affiliations +
Abstract
Solar-assisted water splitting using photoelectrochemical cells (PECs) is one of the promising pathways for the production of hydrogen for renewable energy storage. The nature of the semiconductor material is the primary factor that controls the overall energy conversion efficiency. Finding semiconductor materials with appropriate semiconducting properties (stability, efficient charge separation and transport, abundant, visible light absorption) is still a challenge for developing materials for solar water splitting. Owing to the suitable bandgap for visible light harvesting and the abundance of iron-based oxide semiconductors, they are promising candidates for PECs and have received much research attention. Spinel ferrites are subclasses of iron oxides derived from the classical magnetite (FeIIFe2IIIO4) in which the FeII is replaced by one (some cases two) additional divalent metals. They are generally denoted as MxFe3xO4 (M=Ca, Mg, Zn, Co, Ni, Mn, and so on) and mostly crystallize in spinel or inverse spinel structures. In this mini review, we present the current state of research in spinel ferrites as photoelectrode materials for PECs application. Strategies to improve energy conversion efficiency (nanostructuring, surface modification, and heterostructuring) will be presented. Furthermore, theoretical findings related to the electronic structure, bandgap, and magnetic properties will be presented and compared with experimental results.

1.

Introduction

Solar energy is considered as one of the promising clean energy sources which is abundantly available (120,000 TW)1 and can be utilized to meet the increasing energy demand of the world if it is coupled with a suitable means of capturing and conversion technology. One such avenue for energy storage is photoelectrochemical cells (PECs) for the production of hydrogen. With that, the energy is stored in the form of chemical bonds and is later available on demand.210 PEC systems, in principle, are analogs to commercial electrolyzers, but the voltage required to electrolyze water comes directly from the photovoltage generated from absorbed solar photons. The photoabsorbing material (mostly semiconductors) should produce photovoltages sufficient to drive the water splitting reaction into hydrogen and oxygen, which necessitates band positions of the semiconducutor straddling the water reduction and oxidation potentials. Furthermore, the photoabsorbing material should be stable under solar irradiation, cheap and abundant, and should possess enough catalytic activity and selectivity. Thus, finding suitable semiconductors which satisfy these criteria poses a primary challenge for this technology to be used as an economical viable route for renewable energy.

Numerous inorganic compounds from a variety of material classes have been studied in regard to their suitability as electrode materials for PEC application.2,8,10 Metal oxides are one of the promising materials for PEC water splitting. The vast amount of research on metal oxide semiconductors for PEC is driven by (1) the cost reduction arising from the abundance of metal oxides, (2) stability under PEC operating conditions, and (3) the facile synthetic procedures which do not require vacuum facilities. Spinel ferrites, with the general empirical formula, MxFe3xO4, are widely investigated as magnetic materials and used as spin filters in spintronics.11,12 They also find wide applications in the area of electrochemical energy storage (batteries and electrochemical capacitors) and photocatalysis.1316 Spinel ferrites exhibit attractive photoelectrochemical activities originating from the (1) narrow optical bandgap (<2.2  eV) for efficiently harvesting light of the visible solar spectrum and (2) multiple oxidation states stabilized by the spinel structure with individual transition metals of known catalytic properties. Furthermore, as the constitute transition metals are abundant and are of low cost, they are promising candidates for PEC scaleup applications.

In the following sections of this mini-review, we cover a brief introduction of PEC to provide an overview of the working principles and the reactions involved. Then the structural properties and some fundamental aspects of spinel ferrite materials (MxFe3xO4) will be presented. Recent advances in the photoelectrochemical application of spinel ferrite for solar-assisted electroreduction or oxidation of water follow. Finally, theoretical investigations focusing on the magnetic, electronic, and optical properties of spinel ferrites will be summarized.

1.1.

Basic Working Principle of Photoelectrochemical Cells

Photoelectrochemical water splitting into hydrogen and oxygen is thermodynamically an up-hill process with the free-energy barrier of G=+237  kJ/mole or E0=1.23  eV under standard conditions. The photoelectrolysis of water consists of five main steps, as depicted in Fig. 1: (1) absorption of photons with energy greater than the bandgap of the semiconductor, (2) photogeneration of electron–hole pairs, (3) band bending at the semiconductor–electrolyte interface leading to separation of charge carriers, (4) diffusion of charge carriers toward the semiconductor–electrolyte interface, and (5) redox reaction involving charge carriers and solution species (oxidation of water to O2 and reduction of water to H2). Additionally, the protons are transported across the electrolyte through proton permeable membranes and the electrons travel to the external circuit to complete the process. The overall solar splitting of water involves two half reactions taking place at the photoanode and photocathode simultaneously. For a PEC system comprising an n-type photoanode and p-type photocathode material, the two half reactions in alkaline media (pH=14) can be expressed as follows:

Eq. (1)

Reduction half recation:2H2O+2eH2+2OHERedo=0.82  versusNHE,

Eq. (2)

Oxidation half recation:2HOH2O+1/2+O2++2eEoxdo=0.41  versusNHE,

Eq. (3)

Over all reaction:H2OH2+1/2O2ΔE=1.23.
Three types of PEC configuration are demonstrated:9 (1) “photoanode coupled with metal electrode,” (2) “photocathode coupled with metal electrode,” and (3) “photoanode coupled with photocathode” [Fig. 1(a)]. The latter configuration allows the use of only solar photons for the overall splitting of water without using precious metal electrodes (typically, Pt for H2 evolution and Ru and Ir oxides for O2 evolution are used) and can operate without external bias.5,17 It can be arranged in stack or side-by-side layouts [Figs. 1(b) and 1(c)]. In a stack, the electrode with the smaller bandgap would be placed behind the electrode with the larger bandgap so that the high energy photons would be absorbed first and the lower energy photons transmitted to the back electrode. Using a combination of the two semiconductors with bandgaps of 1.84 eV for the top and 1.23 eV for the bottom electrode, a theoretical solar-to-hydrogen (STH) efficiency of 23% is predicted.17 The side-by-side configuration is preferable when the two photoelectrodes have similar bandgaps (but different band alignments each optimized for water oxidation or reduction reactions), allowing each to have access to full solar illumination. In this case, the highest STH efficiency is reduced to 16% when the bandgap for the two semiconductors is 1.59 eV. Additionally, the bandgaps have to be close to each other to absorb the same number of solar photons in order to match the photocurrents for hydrogen evolution reaction (HER) and oxygen evolution reaction (OER).18 Furthermore, this particular configuration allows the use of smaller bandgap materials which otherwise cannot be used for stacked configuration and can drive the two half reactions with a smaller photovoltage than that required for a single photoabsorber material.17 Note that in both approaches, the STH efficiency prediction takes into account the major potential losses associated with (1) free-energy loss (the difference between the acquired photovoltage and the bandgap) ranging from 0.5 to 0.6 V for highly crystalline semiconductors but expected to be higher for oxide based semiconductors, (2) kinetic overpotentials for the HER and OER (highly dependent on the choice of the catalyst, e.g., 0.05 V for Pt and 0.4 V for RuO2),19 and (3) other potential losses including resistive and charge transport. Notably, the required bandgap is well above the thermodynamic potential of 1.23 eV and should be sufficient to split water and overcome the electrochemical overpotentials (Fig. 3). However, most real systems have higher potential losses (entropy, reflection) even under the assumption of ideal band alignments, and the minimal bandgap required to split water can be as high as 2.0 to 2.3 eV.20,21 Several excellent reviews appeared recently on topics related to solar water splitting with comprehensive coverage on the general principles of PEC and cell design and configurations, and readers are referred to these reviews for in depth discussions of the topic.5,9,22,23

Fig. 1

Schematic representations of (a) the band diagram of PEC in which both the photoanode and the photocathode have the required ideal band positions for spontaneous water splitting. The total photovoltage generated is large enough to split water and to overcome the electrochemical overpotentials ηox and ηred for the oxidation and reduction of water, respectively. Note that the energy scale is V versus NHE at pH=14. The two types of PEC configurations: (b) dual stack and (c) dual side-by-side.

JPE_7_1_012009_f001.png

1.2.

Structural Properties of Spinel Ferrites

Spinel ferrites are ternary transition metal oxides which are represented by a general formula MxFe3xO4, where M refers to the divalent metal ions (M=Ni, Co, Zn, Ca, Mg, Mn, and so on). They are best known as magnetic materials and photocatalysts, with the most common photocatalytic application being the degradation of pollutants.12,24 The structure of spinel ferrites is derived from the mineral spinel, MgAl2O4, by replacing the trivalent Al atom with Fe3+ and Mg atom by other divalent metal ions.

In the spinel ferrites, the oxide anions are arranged in a cubic close-packed lattice and the cations M and Fe occupy two different crystallographic sites, namely, the tetrahedral (A) and octahedral (B) sites.25 The cubic unit cell consists of 56 atoms, 32 oxygen anions, and 24 cations, 8 of them occupying tetrahedral sites and the other 16 being located at the octahedral sites. Although the charges of M and Fe in the prototypical spinel structure (x=1) are +2 and +3, respectively, other combinations are also possible. The type and the distribution of the divalent metal cations govern the final ferrite structure and dictate the chemistry of ferrites. The main factors governing the preference of the individual ions for the two crystallographic sites are the ionic radii and coordination chemistry of the ions. For example, Zn2+ and Cd2+ preferentially occupy the tetrahedral sites, whereas Ni2+ and Cr3+ have strong preference for octahedral sites. When the tetrahedral sites are occupied by divalent metal ion M2+ and the octahedral sites by Fe3+, the resulting ferrites are called normal spinels (e.g., ZnFe2O4) [Fig. 2(a)]. If the Fe3+ cations fully occupy the tetrahedral sites and octahedral sites are occupied evenly by M2+ and Fe3+, this leads to an inverse spinel [Fig. 2(b)]; examples of these classes include but are not limited to Fe3O4, NiFe2O4, and CoFe2O4. However, there exists a certain degree of inversion in most ferrites which is determined by the fraction of M2+ ions in the octahedral sites.

Fig. 2

Crystal structures of spinel ferrites: (a) normal spinel, (b) inverse spinel, and (c) orthorhombic each demonstrating the three crystallographic sites.

JPE_7_1_012009_f002.png

Thus, in general, the spinel ferrites can be represented by [M1yFey]A[MyFe2y]BO4, where the superscripts A and B identify the tetrahedral and octahedral sites, respectively, and y corresponds to the degree of inversion (0y1). Mixed spinel structures are reported for Mn ferrites and Mn-Zn ferrites. Some spinel ferrites, such as MgFe2O4, CaFe2O4, and BaFe2O4, are known also to form orthorhombic phases [Fig. 2(c)].2628 Other ferrites, such as CuFe2O4, form crystalline solids with cubic or tetragonal unit cells depending on the synthetic conditions.29

Ferrites are regarded to be chemically and thermally stable in aqueous systems.15 Considering the Pourbiax diagrams, most ferrites are stable in the alkaline or near neutral media in which most PEC investigations are carried out, however, they suffer from corrosion in acidic media.30,31 Most of them are semiconductors with bandgap energies allowing the excitation by visible light and possess energetic positions of the conduction and the valence band suitable for either reduction of protons and/or oxidation of water (Fig. 3). The theoretical and experimental bandgap energies Eg are presented in Table 1, and the energetic positions of the valence band EVB and the conduction band ECB of some simple ferrites are shown in Fig. 3. The variation of M1, M2, x, and y in Mx1My2Fe3xyO4 is known to affect the resistivity (conductivity),4144 the optical properties (reflectivity), bandgap energy, and the p/n-type behavior45,46 of the semiconductor. Also, the ability to catalyze thermal reactions is affected by the chemical nature and the magnitude of x present in an MxFe3xO4 compound.47,48

Fig. 3

Band positions of spinel ferrites in contact with aqueous solution referenced with NHE (rightpH=14 and leftpH=0) relative to the standard potentials for the reduction and oxidation of water. Note that the variations of the band positions for some of the spinel ferrites data were collected from references cited in this article.

JPE_7_1_012009_f003.png

Table 1

PEC performances of ferrite-based photoelectrodes.

PhotoelectrodeSupportSynthesis methodPhotoresponseElectrolyte
ZnFe2O4FTO/Al-ZnOCBDIPCE 23% at 400 nm [1.23 V versus reversible hydrogen electrode (RHE)]0.1 M Na2SO432
1.72  mAcm2 (1.23 V versus RHE; 100  mWcm2)
ZnFe2O4FTOAACVDIPCE 10% at 400 nm (1.23 V versus RHE)1.0 M NaOH33
0.35  mAcm2 (1.23 V versus RHE)
ZnFe2O4FTOHT and SC0.32  mAcm2 (1.23 V versus RHE)1.0 M NaOH34
ZnFe2O4FTO/ATOALD0.26  mAcm2 (1.23 V versus RHE; 100  mWcm2)0.1 M NaOH35
IPCE 8% at 400 nm (1.23 V versus RHE)
CaFe2O4FTOPLD0.18  mAcm2 (0.21 V versus RHE; 300 W Xe)0.1 M Na2SO436
CaFe2O4/Ca2Fe2O5PtSpptIPCE 1% at 400  nm2 (0.56 V versus RHE)0.1 M NaOH37
with n-TiO2 (Voc=1.09  V, jsc=0.52  mAcm2, 500 W Xe)a
CaFe2O4PtSpptIPCE 3% at 400 nm (0.2 V versus RHE)0.1 M
with n-TiO2 (Voc=0.97  V, jsc=0.22  mAcm2, 500 W Xe)aNaOH38
CoFe2O4FTOEDWith Pt (jsc=0.3  μAcm2, 30  mWcm2)a0.1 M NaS239
NiFe2O4S-G methodWith Pt (Voc=0.43  V, jsc=0.71  mAcm2, 50  mWcm2)a0.5 M KCl40
Note: CBD, chemical bath deposition; AACVD, aerosol-assisted chemical vapor deposition; HT, hydrothermal; SC, solution casting; ALD, atomic layer deposition; PLD, pulsed laser deposition; ED, electrochemical deposition; Sppt, solution precipitation; S-G, Sol-gel.

aTwo electrode configuration.

2.

Photoelectrochemical Application of Spinel Ferrites

Owing to the varied chemical composition, multiple valency states and choice of metal cation ferrites have attractive photoelectrochemical and catalytic activities. The improved electrical conductivity compared to the corresponding single component metal oxide (iron oxides) is mainly attributed to the presence of different metal cations which facilitate the electron transport process and/or support rich redox chemistry and also has significant importance in designing efficient photoelectrodes for PECs.49 The early works on the fundamental (photo) electrochemical investigations of ferrite electrodes date back to the late 1970s and early 1980s. Kung et al.50 reported one of the first ferrites employed for PEC, CdFe2O4, and in the following the years, other types of ferrites were investigated including Li0.5Fe2.5O4,51 MgFe2O4,52,53 Fe3O4,54 p-CaFe2O4,3638,5559 TixFe3xO4,60 p- and n-type CoxFe3xO4 and CoTixFe2xO4,45 p-CoFe2O4,39 p- and n-type NiFe2O4,40,46 n-ZnFe2O4,41,51,33,61 and ZnxTiyFe3xyO4.41 Most of the early works focused on the basic properties of PECs, such as the determination of the flat band potentials and the energetic positions of the valance and conduction bands, however, the reported photocurrents and efficiencies were very low. Some of the fundamental semiconducting properties of these ferrites are summarized in Fig. 3 and the PEC performances are presented in Table 1.

2.1.

p-Type Spinel Ferrites

As with many other materials investigated for photoelectrochemical reduction of water, spinel ferrites are also studied as photocathodes for HER. In order to reduce protons to H2, the conduction band edge of the photocathode must be more negative than the hydrogen redox potential. Considering the band diagrams in Fig. 3, most of the spinel ferrites meet this criterion, however, only p-type CaFe2O4,3638,5559 CoFe2O4,39 and NiFe2O440 have been studied in PECs. Furthermore, the mechanism of HER is pH dependent; in acidic media, the reaction mainly involves proton reduction while the reduction of water to hydroxide ions is the primary route in alkaline solutions. Hence, PEC studies involving spinel ferrites are preferably conducted in neutral or basic solutions, as most spinel ferrites are not stable in acidic media.

One of the most investigated p-type ferrite is CaFe2O4. Matsumoto et al.55 reported a p-type CaFe2O4 for the photoelectrochemical reduction of water. The CaFe2O4 electrodes are prepared as pressed pellets and sintered at 1200°C followed by oxidation under O2 at 1000°C. The CaFe2O4 photocathode exhibits p-type behavior from Mott–Schottky analysis. However, the cathodic photocurrent for HER is negligibly low and when the electrode is coupled with an n-type Zn1.2Fe1.8O4, photoelectrolysis of water without external bias results in an STH conversion efficiency of <0.01%. Matsumoto49 summarized his results on ferrites and other oxide semiconductors, compiled data on the bandgap, and formulated an empirical relation between the bandgap, the conduction band edge, and the valence band edge.

Cao et al.36 have investigated the visible light-induced water splitting reaction employing a p-CaFe2O4 photocathode. The photocathodes have been fabricated by depositing CaFe2O4 thin films on fluorine-doped tin oxide (FTO) coated glass employing a pulsed laser deposition method. A hydrogen evolution rate of 4.8  μmolm2h1 was observed under visible light irradiation (300 W Xe) using a Pt metal electrode without applying any additional bias. In a three-electrode configuration, a cathodic photocurrent was observed at values more negative than +0.64  V. A photocurrent density of 0.117  mAcm2 at 0.06  V was reported as being significantly larger than the values reported by Matsumoto et al.55,56 for metal-loaded CaFe2O4 photoelectrodes, probably due to shorter electron transfer distances in the thinner films and higher electric conductivity.36

Furthermore, Ye et al. have compared the photoelectrochemical properties of p-CaFe2O4, n-ZnFe2O4, p-CaFe2O4/n-ZnFe2O4, and multiple p-n junction CaFe2O4/ZnFe2O4 photoelectrodes. The electrodes have been prepared by a pulsed laser deposition method using CaFe2O4 and ZnFe2O4 pellets as the targets and FTO as the substrate. The authors observed a photocathodic current assigned to the reduction of water on a single-layer p-CaFe2O4 thin film, and a photoanodic current due to the oxidization of water on a single-layer n-ZnFe2O4 thin film. An FTO/ZnFe2O4/CaFe2O4 photoelectrode exhibited a negative photocurrent and a positive open circuit photovoltage (+0.025  V, λ=430  nm, 118  μWcm2) indicating that this electrode with a p-CaFe2O4 layer at the surface contacting the electrolyte acts as a photocathode. The photovoltage generated in such a system is controlled by the built-in junction potential (determined by the number of junctions) and the open circuit voltage (controlled by the type of semiconductor in contact with the electrolyte), as illustrated in Figs. 4(a) and 4(b). Investigating the photoelectrochemical properties of four multiple-junction FTO/(ZnFe2O4/CaFe2O4)x photoelectrodes with the same single-layer thickness of 10 to 15 nm, but with an increasing number of layers x (x=10, 15, 20, and 25), has a remarkable effect on the photocurrent density and the onset potential was observed. The 20-junction photoelectrode showed the highest photocurrent density (0.025  mAcm2 at +0.4  V) and the most positive onset potential (+1.3  V) of all four samples [Fig. 4(c)]. Furthermore, the 20-junction photoelectrode-based PEC exhibited a high open circuit photovoltage of up to +0.97  V, which was much higher than that for a cell having a single junction photoelectrode exhibiting only an open circuit photovoltage of +0.13  V.58

Fig. 4

(a) Band structure of CaFe2O4/ZnFe2O4 junction immediately after light irradiation. (b) Band structure of CaFe2O4/ZnFe2O4 junction under the light irradiation after reaching steady-state conditions. In both cases, the Voc=Voc+Vpn. (c) JV curves of different multilayer p-CaFe2O4/n-ZnFe2O4 junction electrodes (0.1 M Na2SO4, 500 W Xe lamp). Reprinted with permission from Ref. 52. © American Chemical Society 2013.

JPE_7_1_012009_f004.png

The quantum efficiency of a pristine p-CaFe2O4 electrode was found to be relatively low, which is presumably due to the poor mobility of the photogenerated charge carriers.56 Therefore, efforts have been made to improve the conductivity of CaFe2O4 electrodes by doping the material. Doping with Na and Mg was performed by Matsumoto et al. yielding oxides of the type Ca1xNaxFe2yMgyO4. The authors suggested that as the ionic radii of Na+ and Mg2+ are similar to Ca2+ and Fe3+, respectively, Na+ will substitute for Ca2+ and Mg2+ will replace Fe3+. This will create acceptor levels in the bandgap leading to higher electronic conductivity; however, the reported photocurrents are still very small. The authors suggested that the formation of oxygen vacancies at higher amounts of Na (x>0.2) lead to the decrease of conductivity, but the formation of oxygen vacancies is not supported experimentally. In an attempt to improve the low quantum efficiency for the light-induced water splitting reaction, Sekizawa et al. have prepared various metal-doped CaFe2O4 electrodes by radio frequency magnetron cosputtering onto glass substrates coated with antimony-doped tin oxide (ATO) followed by postannealing at a low temperature. However, the doping metals were aggregated in the films after annealing as revealed by scanning transmission electron microscopy. Doping of CaFe2O4 with Au and Ag resulted in an enhancement of the photocurrent without affecting the p-type conductivity. Doping with Ag resulted in an improvement of the carrier mobility together with a red-shift of the photoabsorption. Ag-doped CaFe2O4 showed a 23-fold higher photocurrent than undoped CaFe2O4.59 It is worth mentioning that the enhanced photoresponse may originate from the dopant metals (Ag, Au, Cu/CuO) as they are known as HER cocatalysts.

In a series of papers, Ida et al.37,38,57 reported the light-induced water splitting employing p-CaFe2O4-based cathode coupled with different n-type semiconductors. In their first report, a suspension of presynthesized p-CaFe2O4 powder was coated on a Pt substrate and annealed at 1100°C to 1200°C.38 The sample annealed at 1200°C resulted in a flat, well-adhered crystalline film of p-CaFe2O4 oriented in (320) and (420) planes. Under illumination (500 W Xe lamp), the PEC consisting of p-CaFe2O4 and an n-TiO2 photoanode generated a photovoltage of 0.97 V and short-circuit photocurrent density of 0.22  mAcm2, respectively. The obtained photovoltage is close to the difference between the two onset potentials of p-CaFe2O4 (0.31 V) and n-TiO2 (0.75  V). The authors demonstrated generation of hydrogen under irradiation with visible light without applying a bias, however, the stoichiometric ratio H2/O2 was an order of magnitude different than the expected value of 2. The slow rate of O2 generation is accounted for from the weak absorption of O2 on TiO2 and slow water oxidation kinetics (indicating the need for an OER catalyst). The PEC configuration, the band diagram, and the amounts of gases generated as a function of time in this study are presented in Fig. 5.

Fig. 5

(a) Reaction and band model in PECs using p-type and n-type semiconductor electrodes. (b) Current–potential curve of a photocell with p-CaFe2O4 (2  cm2) and nTiO2 (0.5  cm2) electrodes and model structure of measurement cell and (c) amount of hydrogen and oxygen gases generated from the photocell short-circuited by connecting the p-CaFe2O4 and n-TiO2 electrodes as a function of illumination time. Measurements were carried out in 0.1 M NaOH (aq) under illumination (500 W Xe lamp). Reprinted with permission from Ref. 48. © American Chemical Society 2010.

JPE_7_1_012009_f005.png

Later, the same authors reported the presence of the Ca2Fe2O5 impurity phase in the p-CaFe2O4, enhancing the short circuit photocurrent density (0.55  mAcm2) and slightly increasing the photovoltage (1.09 V).37 Additionally, the O2 formation is enhanced, decreasing the H2/O2 ratio to 3.7, which is, however, still higher than the theoretical value of 2. n-ZnO was also tested as a photoanode together with p-CaFe2O4 and a photovoltage of 0.82 V was generated. When employing this PEC for unbiased water splitting, only H2 gas was detected.57 The photodissolution of n-ZnO is partly responsible for the absence of O2 gas, but there is no obvious correlation between the amount of dissolved Zn2+ ions and the H2 gas evolved (which is much higher than the Zn2+ ion concentration detected).

Very few additional spinel ferrites other than p-CaFe2O4 are studied as photoathodes. Yang et al. have investigated the photoelectrochemical performance of porous CoFe2O4 nanosheets on FTO. The electrodes have been prepared from an aqueous solution of Co and Fe nitrate through a template-free electrochemical deposition followed by a heat treatment at 933 K. The electrodes exhibited only a small cathodic photocurrent of 0.3  μAcm2 in 0.1 M aqueous Na2S solution at zero bias voltage under visible light illumination (λ390  nm, 30  mWcm2).39

The photoelectrochemical properties of p-NiFe2O4 pellets prepared by sintering sol–gel synthesized particles at 850°C were investigated by Rekhila et al. The open-circuit voltage of and short-circuit current of a two electrode configuration consisting of Pt and p-NiFe2O4 in a 0.5 M KCl cell were reported to be 0.43 V and 0.71  mAcm2 under irradiation with visible light (50  mWcm2). A photon-to-electron conversion efficiency of 0.28 was calculated. However, corrosion of the semiconductor electrode was observed under illumination as well as in the dark.40

Considering the bandgaps presented in Fig. 3, less attention is given for other p-type ferrites (CoFe2O4, NiFe2O4, CuFe2O4) and more tests and investigations are still required. Sometimes, even the fundamental optical properties are controversial. For example, most report the optical indirect bandgap of CoFe2O4 as 1.3 to 1.4 eV,62,63 but Xiong et al. reported a bandgap much lower than these values (0.8 eV).64 In the authors’ opinion, one of the main reasons is that these p-type ferrites are known to exist in a completely or partially inverted spinel structure.24,63 Such a degree of inversion should be well controlled and require a rational design of synthetic strategy to accurately determine the optical properties and to enhance the PEC performance. In addition, they are prone to photocorrosion particularly in acidic media, which can be alleviated by applying suitable protective layers (TiO2, Al2O3).

2.2.

n-Type Spinel Ferrites

The photoelectrochemical oxidation of water to O2 requires an n-type semiconductor with the valence band located more positive than the H2O/O2 oxidation potential (1.23 V versus NHE). In contact with an aqueous electrolyte, such a semiconductor results in an upward band bending which drives the holes toward the surface leading to the oxidation of water to O2. Additionally, good electrical properties and stability under water oxidation conditions are needed. Among the spinel ferrites, n-ZnFe2O4 is the promising candidate and is the only n-type photoanode material reported for PEC application.

Systematic investigation of the photoelectrochemical activity ZnFe2O4 was reported by Tahir and Wijayantha33 and Tahir et al.61 The electrodes were prepared by aerosol-assisted chemical vapor deposition (AACVD) of alcoholic solutions of a bimetallic precursor ([Fe2(acac)4(dmaeH2)2][ZnCl4]) on FTO. The thickness, morphology, and nanostructure of the electrode were controlled by altering the solvent for dissolution of the bimetallic precursor and physical deposition parameters.33,61 The photocurrents were found to be dependent on the solvent, as well as on the deposition temperature and the deposition time. A maximum photocurrent density of 0.35  mAcm2 at 1.23 V versus RHE was obtained with a ZnFe2O4 electrode synthesized using a 0.1 M solution of the bimetallic precursor in ethanol, the optimum deposition temperature of 450°C, and a deposition time of 35 min. This electrode showed an incident-photon-to-electron conversion efficiency of 13.5% at 350 nm and an applied potential of 1.23 V versus RHE.33 In the AACVD process, the aerosol droplet size (controlled by the solvent) and the enthalpy of combustion determine the decomposition pathway (homogeneous versus heterogeneous), thus varying the methanol/ethanol ratio of the solvent resulted in a change in the texture of the ZnFe2O4 electrode. A compact ZnFe2O4 film composed of hexagonal like particles was obtained in pure methanol, but the structure transformed in to porous ZnFe2O4 nanorod films when ethanol was used as the solvent. Intermediate structures were obtained by varying the methanol/ethanol ratio. The textured electrodes exhibited a significantly higher photocurrent under AM1.5 illumination compared to their compact counterparts. The authors attributed this behavior to the improved collection of the photogenerated minority carriers at the ZnFe2O4/electrolyte interface as the average feature size gradually decreased from 500  nm (methanol) to 100  nm (ethanol).61

In general, ferrites as photoelectrodes need high temperatures to crystallize (>1000°C). This limits the choice of support materials and poses a critical challenge to maintain the desired electrode material properties such as surface area and porosity. Recently, Kim et al. introduced a hybrid microwave annealing (HMA) postsynthetic heat treatment with graphite powder as the susceptor being compatible to most transparent conducting glasses. They treated solution processed β-FeOOH nanorods with a Zn nitrate solution and obtained ZnFe2O4 nanorods after thermal treatment at 550°C for 3 h. Some unwanted ZnO on the nanorods was removed in NaOH. Subsequently, the ZnFe2O4 nanorods were subjected to a second heating step at 800°C (20 min) or to HMA (5 min) to increase the crystallinity. The HMA-treated ZnFe2O4 nanorods exhibited at 1.23 V versus RHE (1 M NaOH) and AM 1.5G illumination a photocurrent of 0.240  mAcm2, which was a 10- to 15-fold increase in comparison to conventional thermally treated electrodes and was stable for at least 3 h. The authors claimed that stoichiometric amounts of H2 and O2 can be measured with Faradaic efficiencies (= actual gas evolution rate/rate expected from current) of 90% to 100%. The improved performance after the HMA treatment was attributed to better crystallinity and reduced surface defects as evinced by electrochemical impedance spectroscopy.65 Extending their work, the same authors recently reported the influence of the composition of the annealing atmosphere on the photoelectrochemical behavior of ZnFe2O4 (Fig. 6).34 ZnFe2O4 nanorod films were first treated at 800°C (20 min in air) followed by a mild temperature treatment at 200°C (2 h) either under vacuum, air, or hydrogen atmosphere. The hydrogen and vacuum post-thermal treatment enhanced the photoactivity about 20-fold [Fig. 6(b)]. The increased activity is attributed to oxygen vacancies created in the ZnFe2O4 lattice due to the limited oxidation environment as proven by O 1s XPS. Optimal oxygen vacancy concentrations increase the majority carrier density and lead to improved charge separation. The highest photocurrent was observed for the hydrogen treated sample, 0.320  mAcm2 at 1.23 V versus RHE under AM 1.5G illumination. The authors suggested two types of mechanisms of how the lattice O (OL) is replaced by oxygen vacancies (V¨O) under controlled hydrogen and vacuum atmosphere [Fig. 6(a)]. In hydrogen atmosphere, the OL reacts with H2 and leaves the lattice as water molecules:

Eq. (4)

H2+OLH2O+V¨O+2e,
in vacuum, the oxygen removal is represented by Eq. (5) and the concentration of the oxygen vacancies is given in Eq. (6):

Eq. (5)

OL=1/2O2(g)+V¨O+2e,K,

Eq. (6)

[V¨O]=Kn2[O2]1/2,
where n represents the electron density and [O2] is the partial pressure of the oxygen.

Fig. 6

(a) Schematic representation of different mechanisms of generating oxygen vacancies by post-treatments under hydrogen or vacuum conditions, (b) JV characteristics of ZnFe2O4 photoanodes before and after post-treatment [hydrogen (H), vacuum (V), and atmospheric air (A)] under AM 1.5G illumination (100  mWcm2) in 1 M NaOH electrolyte (scan rate=10  mVs1). Reprinted with permission from Ref. 60. © Royal Society of Chemistry 2015.

JPE_7_1_012009_f006.png

In general, for most iron based n-type semiconductor oxide materials, the diffusion length of the minority carrier is very short leading to an inherently high charge recombination rate, which limits the efficiency of the PECs. Several strategies to address this issue were developed including nanostructuring33,61,65 and doping.59 Another attractive nanostructuring strategy is to use a structured transparent conductive oxide current collector to capture and tunnel the photogenerated electrons readily while the large interfacial area allows efficient transfer of the holes to the solution. This strategy was recently demonstrated for ZnFe2O4 decorated Al-doped ZnO (AZO) nanowire films.32 The Al:ZnO nanowires were grown on FTO substrate hydrothermally at 88°C and treated with an ethanolic solution of FeCl3. Subsequent annealing at 550°C leads to ZnFe2O4-coated Al:ZnO nanowires [Figs. 7(a)7(d)]. Depending on the time of FeCl3 exposure, the nanowires can be converted to nanotubes due to the dissolution of Al:ZnO in acidic FeCl3 solution. The photoanode shows outstanding photoelectrochemical performance with low onset potential (0.38 V versus RHE) with a photocurrent density of 1.72  mAcm2 at 1.23 V versus RHE [Fig. 7(e)]. The synergy of high conductivity of Al:ZnO, the nanowire morphology for charge separation, and the visible light absorption of ZnFe2O4 coating are attributed to be responsible for the high photoelectrochemical performance.

Fig. 7

SEM images of (a) 0.5% AZO and 0.5% AZO–ZFO with treatment times of (b) 1 min, (c) 3 min, (d) 7 min, (e) LSV plots of 0.5% AZO and 0.5% AZO–ZFO photoanodes under chopped illumination and (e) the corresponding ABPE plots. Insets show the schematic illustration of morphology evolution for the AZO and AZO–ZFO composite photoanodes with different treatment times. Reprinted with permission from Ref. 61. © Royal Society of Chemistry 2016.

JPE_7_1_012009_f007.png

Recently, Hufnagel et al.35 prepared mesoporous ZnFe2O4 thin films on a macroporous ATO scaffold using atomic layer deposition (ALD). The photoresponse of the electrodes was tested in three electrode PECs and the electrodes exhibited 4- to 5-fold higher photocurrent density (0.26  mAcm2 at 1.23 versus RHE) compared to nonstructured ZnFe2O4 films prepared in a similar way (0.05  mAcm2 at 1.23 versus RHE). Additionally, the authors show that such electrodes have more negative photocurrent onsets (0.9 V versus RHE) compared to reported values.33

Spinel ferrites were recently investigated for construction of heterojunction electrodes to improve the photoelectrochemical performance of other widely used semiconductors. In this regard, heterojunction electrodes such as ZnFe2O4/Fe2O3,6668 CaFe2O4/Fe2O3,69,70 CaFe2O4/TaON,71 and CaFe2O4/BiVO472 as photoanodes for the OER were studied. Table 2 presents the PEC performance of ferrite heterojunction photoelectrodes. Borse et al. prepared ZnFe2O4/Fe2O3 layers on stainless steel by depositing an aqueous solution of Zn and Fe salts employing a plasma spray method and investigated the photoelectrochemical activity of the ZnFe2O4/Fe2O3 electrode. Under simulated solar light (AM1.5G, 100  mWcm2) with a bias of 1.4 V versus RHE, a photocurrent of 0.1  mAcm2 is measured, which is fivefold higher than for pristine ZnFe2O4. The authors also reported hydrogen production in a two electrode set-up employing graphite as the counter electrode. Again, the composite photoanode exhibited a significantly higher photoactivity than a bare ZnFe2O4 photoelectrode. The rates of HER at the ZnFe2O4 and the ZnFe2O4/Fe2O3 photoanode were calculated to be 46.3 and 99.0  μmolcm2h1, respectively, resulting in STH conversion efficiencies of 0.06 and 0.0125, respectively. However, no data were given for the formation of molecular oxygen. The results of electrochemical impedance spectroscopy evinced a significantly lower interfacial charge transfer resistance of the ZnFe2O4/Fe2O3 composite electrode than the ZnFe2O4 electrode.67

Table 2

PEC performance of ferrite based composite photoelectrodes.

PhotoelectrodeSupportSynthesis methodPhotoresponseElectrolyte
Fe2O3/ZnFe2O4FTOHT and SC0.44  mAcm2 (1.5  V versus RHE; 100  mWcm2)0.5 M NaOH + 0.1 M
IPCE 17% at 400 nm (1.5 V versus RHE)Glucose68
Fe2O3/ZnFe2O4FTOED and SC0.4  mAcm2 (1.4 V versus RHE, 100  mWcm2), Al treated1 M NaOH66
TiO2/ZnFe2O4FTOHT and ALD0.7  mAcm2 (1.0 V versus RHE, 100  mWcm2)1 M KOH73
CaFe2O4/Fe2O3FTOHTIPCE 10% at 420 nm 1.23 V versus RHE)1.0 M NaOH69
0.53  mAcm2 (1.0 V versus RHE; 100  mWcm2)
TaON/CaFe2O4FTOEPDIPCE 30% at 400 nm 1.23 V versus RHE0.5 M NaOH71
1.23  mAcm2 (1.23 V versus RHE; 300  WXe>420  nm)
BiVO4/CaFe2O4FTOCBD and EPDIPCE 20% at 420 nm (1.23 V versus RHE)0.5 M Na2SO474
0.96  mAcm2 (1.23 V versus RHE; 100  mWcm2)
Note: EPD, electrophoretic deposition.

ZnFe2O4/Fe2O3 nanorod composite photoanodes have also been prepared using hydrothermally grown FeOOH nanorods and subsequent treatment with different concentrations of Zn precursor. After calcinations at 750°C, a ZnFe2O4/Fe2O3 composite electrode was obtained.68 The photocurrent density for the composite electrode was 0.44  mAcm2 at 1.2  V versus RHE, which was almost twice as high as that for a Fe2O3 electrode (0.24  mAcm2). McDonald et al. employed the electrodeposition route and obtained photoelectrodes composed of an α-Fe2O3 (hematite) core and a ZnFe2O4 shell as confirmed by XRD.66 The electrodeposited β-FeOOH films on FTO were converted into α-Fe2O3 by heat treatment and subsequent treatment with Zn-containing solution on top of α-Fe2O3 film yielded a Zn-rich top layer on α-Fe2O3 after annealing. The highest photocurrent was obtained with a composite electrode exhibiting a ZnFe2O4/Fe2O3 ratio of 1. The increase in the photocurrent of the heterojunction electrodes compared to the bare α-Fe2O3 electrode was explained as being due to the enhanced electron hole separation at the ZnFe2O4/Fe2O3 interface. A further enhancement in photocurrent was obtained by a treatment of the composite electrodes with an Al3+ solution yielding thin layers of a solid solution (ZnFe2xAlxO4 or Fe2xAlxO3) after heat treatment. With this, the number of surface states that serve as electron–hole recombination centers is probably reduced. But it was also observed that both the formation of a ZnFe2O4 layer and the incorporation of Al3+ into the surface made the surface less catalytic for the OER. However, when Co2+ was introduced into the surface of the ZnFe2O4/Fe2O3 composite electrodes as oxygen evolution catalysts, the onset of the photocurrent was shifted to more negative voltage and the overall photocurrent was improved.66 Furthermore, ZnFe2O4/TiO2 heterostructures are also prepared using a combination of solution-phase materials growth techniques and ALD.73 The TiO2 nanowires were hydrothermally grown on FTO substrate and later infiltrated by Zn precursor using ALD. The ZnFe2O4/TiO2 composite electrodes show an enhanced visible light photoresponse compared to bare TiO2 photoelectrodes. The authors suggest that the extended visible light absorption by the ZnFe2O4, the nanowire morphology of the TiO2, and favorable band edge positions are the main reasons for the increased photoresponse.

A related photoanode has been prepared by anisotropic growth of a β-FeOOH film on FTO from an aqueous solution containing Fe and Ca ions followed by two-step thermal annealing at 550°C and 800°C. The authors suggested that this procedure induces the formation of a p-CaFe2O4/n-Fe2O3 heterojunction photoanode. The presence of Ca in the Fe2O3 film, leading to the formation of CaFe2O4, has been proven by XPS measurements. Under illumination (AM 1.5G, 100  mWcm2), the heterojunction photoanode exhibits a photocurrent density of 0.53  mAcm2 at 1.23 V versus RHE, which is a 100% higher photocurrent response than that obtained using a bare α-Fe2O3 electrode. Based on electrochemical impedance spectroscopy, the photocurrent enhancement has again been attributed to an enhanced charge carrier separation and a reduced resistance of the interfacial charge transfer between the electrolyte and the electrode.69

Kim et al.71,74 reported the preparation and characterization of p-CaFe2O4 modified TaON and BiVO4 p-n heterojunction photoanodes. Both n-type semiconductors are known to be suitable anode materials for solar-driven water splitting in PECs. The valence band of p-CaFe2O4 is more positive than the water oxidation potential, and both semiconductors TaON and BiVO4 form staggered relative band positions with the ferrite as required for an effective heterojunction photoanode. In both cases, p-CaFe2O4, which has been synthesized by a conventional solid state reaction, was deposited on top of the n-semiconductor/FTO electrode by electrophoresis. The pristine TaON electrode showed an anodic photocurrent density of 0.230  mAcm2 at 1.23 V versus RHE (0.5 M NaOH, λ>420  nm). The CaFe2O4 layer on the surface of a TaON electrode resulted in a significant increase of the photocurrent density (1.26  mAcm2). The observed photocurrent was found to be a result of overall water splitting yielding H2 and O2 in a ratio of 1.5, accompanied, however, by a deterioration of the TaON. Impedance spectroscopic analysis indicated that the formation of the heterojunction increased the photocurrent density by reducing the resistance of the charge carrier transport and, consequently, enhancing the electron–hole separation.71 Anodic photocurrents have likewise been observed for both BiVO4 and CaFe2O4/BiVO4 electrodes (0.5 M Na2SO4, AM 1.5G 100  mWcm2); the bare BiVO4 electrode showed a photocurrent density of 0.58  mAcm2 at 1.23 V versus RHE V while the CaFe2O4/BiVO4 heterojunction photoanode exhibited 0.96  mAcm2, comprising an increase of 65% over that measured at the BiVO4 electrode. The formation of the heterojunction was found to reduce the recombination of the photogenerated charge carriers on the electrode surface with little effect on bulk recombination as evinced by an investigation of the interfacial transfer of charge carriers using hydrogen peroxide as an electron donor.74

In the above demonstrated heterojunction electrodes, efficient charge separation is the key factor for improved photoelectrochemical activity. This is demonstrated in the energy diagrams of the composite electrodes in Fig. 8. The ferrites (CaFe2O4 and ZnFe2O4) have valence band energies located between the water oxidation potential and that of the valence band of the second semiconductor. Thus, holes generated at the more positive valence band are extracted to the ferrite valence band reducing bulk recombination and allowing successful injection of the holes to the electrolyte. On the other hand, as the conduction band of the ferrites is situated at a more negative potential, the photogenerated electrons will be easily transferred to the conduction band of the second semiconductor for collection at the back contact.

Fig. 8

Schematic representation of the band positions of heterojunction electrodes showing the flow of photogenertaed charge carriers: (a) the CaFe2O4 /TaON/ (p-n junction) and (b) the Fe2O3/ZnFe2O4 (nn junction).

JPE_7_1_012009_f008.png

Furthermore, the modification of the heterojunction photoanodes by depositing OER cocatalysts results in higher photocurrent densities by decreasing the onset potential and facilitating the interfacial charge transfer. In this regard, “cobalt phosphate” (CoPi) is used widely as an OER catalyst.7578 For example, in the CaFe2O4/TaON71 heterojunction electrode system, after deposition of CoPi and with an applied bias of 1.23 V versus RHE, H2 and O2 were generated with nearly of stoichiometric ratio of 2.1 (123  μmol H2 and 59  μmol O2 were produced within 3 h of illumination with λ400  nm). The STH efficiency was 0.053% at 1.0 V versus RHE, but reached 0.55% when a PV device is coupled in tandem configuration (assuming the applied voltage is zero). However, the initial current decreased within 3 h to about 50%. The beneficial role of the CoPi cocatalyst was revealed by performing the gas evolution experiment with a CaFe2O4/TaON photoanode in the absence of CoPi. Under these experimental conditions, no constant photocurrent was obtained, the faradaic efficiencies decreased (50% to 70%), and the H2/O2 ratio became less than stoichiometric (1.51) due to the self-oxidation of TaON.71 Similar results have been obtained with CaFe2O4/BiVO4 composite photoanodes. In this case, the CoPi modified CaFe2O4/BiVO4 electrode exhibited a lower photocurrent density when compared to the CaFe2O4/BiVO4 heterojunction electrode, but an improved stability of the current density was observed indicating that the presence of the OER cocatalyst is beneficial for the stabilization of the CaFe2O4/BiVO4 heterojunction. The evolution of H2 and O2 during the photoelectrochemical water splitting reaction was measured in a three electrode set-up in phosphate buffer (pH=7) under applied bias. The total amounts of H2 and O2 evolved within 2 h of illumination with visible light (λ400  nm) were 297 and 140  μmol, respectively. The resulting H2/O2 ratio of 2.1 confirmed that the generation of the photocurrent was mainly due to the water splitting reaction. The Faradic efficiency during this reaction was reported to be about 78% to 88%. The photocurrent density was initially 4  mAcm2, but dropped rapidly during the first 30 min of illumination and decreased slowly afterward.74

In general, the above few examples of spinel ferrite-based PECs demonstrated the attractive photoelectrochemical properties of these materials. The measured photocurrents of PECs with ferrite photoelectrodes are still low, but different strategies proved that higher photocurrents can be achieved. The main contributors for low photoelectrochemical performances are identified as (1) slow interfacial charge carrier transfer, (2) inherently high charge carrier recombination rates, and (3) loss of interfacial area due to the high thermal treatment. Strategies to address these bottle necks, including (1) nanostructuring, (2) forming heterojunction structures, (3) cocatalyst coating, and (4) control of the defect chemistry are shown to enhance the photoelectrochemical performance.

It is worth mentioning that the reported experimental band positions of some of the spinel ferrites related to HER and OER potentials are still arguable. Based on the energetic band diagrams of ferrites presented in Fig. 3, ZnFe2O4 and MgFe2O4 are, in principle, capable of producing hydrogen. However, to the best of the authors’ knowledge, there is no literature report proving this. For example, the reported flatband potentials of ZnFe2O4 (0.6 to 0.8 V versus RHE)33,35 suggest the conduction band lies slightly below the HER potential. We also found that the conduction band edge is slightly positive (100  mV) to that of the proton reduction potential, but we are still in the process of verifying this. Thus, more experimental investigations are required to refine the existing photoelectrochemical data. Furthermore, MgFe2O4 is also interesting for OER, but there are no reports except the fundamental investigation reported by Benco and Koffyberg.52 However, there are a few reports on photocatalytic H2 production using MgFe2O4 which to some extent show the potential of this material.53,72

3.

Theoretical Investigations

Considering the above-mentioned PEC applications of spinel ferrites, several questions concerning the water splitting process arise that cannot be easily answered. Theoretical investigations allow an insight to material on an atomistic level that the experiment cannot provide in general. For example, theory allows one to study water and hydroxyl adsorption on different sites of differently terminated ferrite surfaces, and its effect on the degree of inversion, the magnetic structure, the bandgaps, and band positions. It is well known that the accuracy of theoretical predictions of electronic properties strongly depends on the choice of the method. During the last years, several quantum chemical protocols have been developed which go beyond the traditional independent-particle model at density-functional theory (DFT) level, the most refined being self-consistently vertex corrected GW.79 A considerable number of theoretical investigations were devoted to the electronic, magnetic, structural, and energetic properties of ferrites. With a few exceptions, the quantum-chemical studies were performed at DFT level, in most cases employing the local (spin) density approximation (LSDA) or the generalized-gradient approximation (GGA). As is well known, the electronic properties of transition metal oxides are not accurately described within the LSDA or GGA, thus hybrid approaches combining DFT with unrestricted Hartree–Fock theory (HFT) are preferable. Global DFT-HFT hybrids have been demonstrated to provide accurate lattice constants, atomization energies, and bandgaps for a wide range of compounds80 but are computationally much more demanding than GGA methods. Therefore, in most studies, a semiempirical on-site correction, the so-called LSDA+U or GGA+U approach,81 is applied. In this method, the effective parameter U replaces one-center exchange and Coulomb integrals from unrestricted HFT. In principle, the value of U for each transition metal can be derived from exact theory, but in practical calculations, it is usually treated as an empirical parameter to adjust certain properties to experimental reference data.

Absolute band positions thus, the work functions and fundamental bandgaps can only be obtained from two-dimensional slab model calculations of surfaces. However, recently, very accurate results for both properties were obtained for binary systems (e.g., ZnO, CdO, GaAs, GaP, InP) from self-consistent GW calculations.79 So far, most theoretical investigations of ferrite surfaces, however, have focused on the adsorption of water and other small adsorbates. The following sections provide a survey of recent theoretical studies on bulk and surface properties of ferrites.

3.1.

Electronic Structure of the Bulk

The effect of charge ordering in the octahedral sites of Fe3O4 and CoFe2O4 on their electronic structure was investigated using DFT+U.82 A precise description of charge ordering was found to be crucial in determining the bandgaps of the compounds. GGA+U calculations of the electronic structure of antiferromagnetic CaFe2O4 yield an indirect bandgap of 1.9  eV.83 The ionicity of Fe3O4 has been determined using DFT calculations.84 Furthermore, a new developed quantum mechanical estimation method for the ionicity of spinel ferrites has been proposed and tested. On the basis of this, the ionicities of the spinel ferrites MFe2O4 (M: Co, Cu, Fe, Mn, Ni) were calculated. The electronic structure of NiFe2O4 has been investigated using LSDA+U and hybrid-DFT in 2012.85 According to the theoretical results, the system is an indirect gap material in one of the minority channels and slightly larger direct bandgaps can be found both in the minority and majority channels. The electronic structure of MFe2O4 (M: Ca, Mg, Zn) was investigated in a combined experimental and theoretical study.86 The DFT calculations reveal that the M-ion controllably affects the density of states of the Fe d-orbitals near the Fermi level. The electronic structure of ZnFe2O4 was studied using GGA+U.87 Taking the effect of spin arrangement on symmetry into account, ZnFe2O4 was classified as a semiconductor. The impact of cation distribution in CuFe2O4 on electronic structure and magnetic properties has been investigated by Feng et al.88 The lattice structure was optimized on the GGA level and the electronic structure was calculated with GGA+U. The calculated density of states shows that the distribution of Cu ions significantly impacts the electronic structure. Multilayer bispinel composites, in which one member is Fe3O4 and the other is MFe2O4 (M: Co, Mg, Mn, Ni), were modeled using GGA+U by Wells et al.89 It was found that substitution of the transition metal sites in the supercell produces cation charge transfers and magnetization modulation. Band shifts and gap modulation were comparable to the chemically similar bulk compounds. Two different distributions for the octahedral-site cations in ZnFe2O4 and CdFe2O4 have been investigated using LDA and GGA, as well as LDA+U and GGA+U.90 It was shown that a different octahedral-site distribution impacts the density of states as well as the bandgaps in both the normal and inverse spinel configurations of these compounds. Magnetic properties and the electronic structure of NiFe2O4 have been studied using hybrid-DFT.91 The calculated density of states suggests that NiFe2O4 is an insulator. The electronic structure of normal and inverse spinel ferrites MFe2O4 (M: Co, Fe, Mn, Ni) was investigated by self-interaction corrected LSDA.92 For both structures, all studied compounds were found to be insulating but with smaller gaps in the normal spinel structure. The calculated spin magnetic moments and exchange splitting of the conduction bands were dramatically increased when moving from the inverse spinel structure to the normal spinel. A first principle investigation of the electronic structure of MFe2O4 (M: Co, Fe, Mn, Ni) compares the performance of LSDA and LSDA+U.93 For the LSDA+U approach, the charge ordering is stable in contrast to a metallic state given by the LSDA approach. Calculated x-ray absorption spectra as well as the x-ray magnetic circular dichroism spectra were in good agreement with the experiment. The electrical and magnetic properties of the normal and inverse spinel structures of MnFe2O4 were calculated with DFT by Zuo and Vittoria.94 The calculated bandgap suggests that MnFe2O4 is a complex insulator, in contrast to earlier LSDA and GGA calculations which suggest a half-metallic behavior. MnFe2O4 has been investigated theoretically at DFT level.95 The calculated band structure shows a low carrier density half-metal in the fully ordered state, in contrast to experimental characterizations. The computations yield a strong coupling of the energy bands at the Fermi energy to the internal structural parameter u as well as strong effects on the electronic structure upon partial interchange of Fe and Mn atoms.

Calculations of the K-edge x-ray absorption near-edge structure (XANES) in elemental iron and MFe2O4 (M: Mg, Mn, Ni, Zn) were carried out by Safontseva and Nikiforov.96 It was shown that the Fe K-edge energy shift found experimentally occurs upon the transition from elemental iron to the spinel ferrites. This shift was demonstrated to be identically directed for ferrites with a normal and inverted spinel structure. A computational study of ferrimagnetic ZnxNi1xFe2O4 compounds using the pseudofunction method was carried out in 1996.97 Substitution of Ni with Zn enhances the localization of the 3d states of Fe on the octahedral sites, so that the O 2p-Fe 3d hybridized states can be resolved into two distinct twofold and threefold features. Normal and inverse MnFe2O4 was investigated with HF level of theory in 1996.98 From Mulliken population analysis and net spin density distributions, it was concluded that the charge states of Mn and Fe in the ground state show no evidence of charge transfer leading to Fe2+ at A sites and Mn3+ at B sites in the inverse spinel structure AB2O4. An early computational study of the band structure and magnetic moments of ferrites MFe2O4 (M: Co, Fe, Mn, Ni, Zn) on LSDA level99 only covered the metallic, high-temperature phase in the case of M=Co, Fe, Mn, Zn. In contrast, NiFe2O4 was described as an insulator.

Table 3 compares experimental bandgaps of selected ferrites with calculated values gained by different theoretical protocols. It is obvious that plain DFT without any further corrections systematically underestimates the bandgap of the considered systems. The theory states that using a self-interaction corrected or a hybrid DFT approach reduces the error. Using a DFT+U framework provides results that show good agreement with the experimental data. This is easily explained by the added potential U, which is an empirical parameter that can be explicitly chosen to fit the experimental bandgap. There is no universal choice for the U parameter.

Table 3

Comparison of experimental and calculated bandgap energies of some selected MFe2O4 spinel ferrites.

Spinel ferriteExperimental bandgaps (Eg/eV)bCalculated bandgaps (Eg/eV)
Fe3O40.21000.75b (Ref. 101)
0.87c
0.2d (Ref. 102)
0.14e (Ref. 103)
0.0d (Refs. 104 and 105)
0.35f (Ref. 106)
0.19e (Ref. 93)
0.08 to 0.53b (Ref. 82)
MgFe2O42.072
1.74d53
CaFe2O41.955,561.9b (Ref. 83)
1.9472,107
BaFe2O41.85 to 1.90108
CoFe2O41.39±0.31i, 2.31±0.28d621.171b (Ref. 109)
1.42i, 1.95d630.9b (Ref. 110)
0.63e (Ref. 93)
0.75 to 1.41b (Ref. 82)
NiFe2O41.52±0.08i, 2.3d, 2.74d621.63b (Ref. 101)
1.56d, 1.99i401.675b (Ref. 101)
1.3g (Ref. 111)
0.97b (Ref. 110)
0.99e (Ref. 93)
1.1e, 2c (Ref. 85)
CuFe2O41.421120.016 to 1.2b (Ref. 88)
1.54i, 1.96d29
ZnFe2O41.9411.68b (Ref. 101)
1.901130.837g, 0.930g, 0.787g, 0.869g, 0.874h (Ref. 116) 1.821b (Ref. 109)
1.92i114
1.83i, 1.93d115
1.81i, 1.90d
CdFe2O42.350
MnFe2O40.4b (Ref. 117)
0.075l (Ref. 92)

aThe values of the indirect and direct bandgap are labeled by i and d, respectively.

bGGA+U

cHybrid

dLDA+U

eLSDA+U

fSIC-LSDA

gGGA

hLDA

3.2.

Magnetic Properties

Since the magnetic structure of ferrites (strongly) affects the calculated bandgap,92 the magnetic properties also have to be taken into account when discussing electronic properties. The magnetic properties of MFe2O4 (M: Co, Fe, Mn, Ni, Zn) were studied with DFT methods.118 The theoretically obtained magnetizations were consistent with experimental results in the absence of an external field. GGA+U was applied to investigate the electronic structure and magnetic properties of MnFe2O4.117 The calculations account for a cubic structure with ordered spins and insulating behavior. It was found that the high-spin state is favorable for the two cations Mn and Fe. The position of magnesium ions in Mg2+-doped lithium ferrite of the composition Li0.50.5xMgxFe2.50.5xO4 has been investigated by interatomic potential and DFT calculations.119 The lowest energy structure was found for Mg2+ ions evenly replacing Li+ and Fe3+ ions on octahedral sites. This occupation affects a decrease in magnetization for the Mg2+-doped ferrite relative to the undoped lithium ferrite. A computational study of the spinel ferrites CoFe2O4 and NiFe2O4 shows that LSDA+U and GGA+U allow for a good quantitative description of these materials.110 The effect of epitaxial strain on the magnetocrystalline anisotropy was investigated and the results are in good agreement with experimental observations. The structure of partially inverse spinel CoFe2O4 as well as its electronic and magnetic properties has been investigated by the GGA+U approach.120 It was found that the Co and Fe ions prefer their high-spin configurations with higher spin moments at octahedral sites. Certain investigated inversion degrees show half-metallic behavior.

3.3.

Structural and Thermodynamic Properties

A computational study of the inversion thermodynamics and electronic structure of (thio) spinels FeM2X4 (M: Co, Cr, Mn, Ni; X: O, S) was published in 2015.121 The analysis of the configurational free energies shows that FeCr2X4 and FeMn2S4 are fully normal, FeNi2X4 and FeCo2S4 are intermediate, and FeCo2O4 and FeMn2O4 are fully inverted. The calculations illustrate that FeCr2X4, FeMn2X4, FeCo2O4, and FeNi2O4 are half metals in the ferrimagnetic state when Fe is in tetrahedral positions. When M is filling the tetrahedral positions, the Cr-containing compounds and FeMn2O4 are shown to be half-metallic systems, whereas the Co and Ni spinels are shown to be insulators. Yao et al.116 investigated the structure and electronic properties of normal spinel ZnFe2O4 using GGA and LDA. They suggest that the GGA functional RPBE combined with ultrasoft pseudopotentials is a good method for predicting the crystal structure of the compound. The computational results indicate that ZnFe2O4 is a direct gap semiconductor and that there is a strong hybridization between the Fe 3d states and the O 2p states as well as between the Zn 3d states and O 2p states. DFT calculations at the GGA+U level were performed on AB2O4 (A: Fe, Ni, Zn; B: Fe, Cr) spinel oxides in order to determine thermodynamic properties.101 Calculated mixing energies quantitatively reproduce experimental data. Reactions leading to an excess of A or B, respectively, were found to be slightly exothermic in a number of spinel compounds. A set of effective chemical potentials (ECPs) that connect energies of MFe2O4 (M: Co, Fe, Ni, Zn) spinels and oxides calculated at 0 K from DFT to free energies at high temperature and pressure in the presence of water was derived and tested.109 The ECPs were used to calculate free energies of low index stoichiometric surfaces of nickel ferrite in water, predicting surface denuding at high temperatures. A computational study compares the performance of GGA-DFT and hybrid-DFT (B3LYP) for the equilibrium structure of Fe3O4.104 The ground state calculated by GGA-DFT is metallic with Fd-3m symmetry while the hybrid level of theory yields a charge ordered semiconducting state with P2/c symmetry. Phonon frequency calculations showed that charge ordering causes symmetry breaking of force constants on symmetry lowering from the cubic unit cell to the monoclinic unit cell.

3.4.

Surfaces and Adsorption

The structure, electronic properties, and energetics of the NiFe2O4(001) surface and its interaction with water both in the absence and in the presence of surface oxygen vacancies have been studied using DFT+U.122 It was shown that water adsorbs dissociatively on the surface oxygen vacancies leading to the formation of surface hydroxyls. Furthermore, it was found that at high temperature, water desorbs leaving a surface containing oxygen vacancies. The reactivity of the NiFe2O4(111) surface has been studied using DFT+U.123 The surface reactivity is significantly higher in comparison with the Fe3O4(111) surfaces. Dissociation of water was found to be highly favorable on the NiFe2O4(111) surfaces. The activation barrier for the dissociation of a single water molecule was dependent on the termination of the surface. The electronic properties of CuFe2O4 and the adsorption behavior of an NO molecule on the CuFe2O4 (100) surface were studied using DFT+U.124 The authors suggest that the ground state of CuFe2O4 bulk has an inverse spinel structure and is a magnetic semiconductor. The NO molecule prefers to adsorb on the top site of the Fe atom on the (100) surface, forming an N-Fe bond. DFT+U calculations of the adsorption behavior of Ni and Ti on the Fe3O4(001) surface have been carried out by Bliem et al.125 For both atoms, an incorporation in an octahedral Fe site of the force-relaxed Fe3O4(001) surface is energetically favorable. Boron adsorption on an Fe3O4(100) surface was studied by GGA calculations in 2015.126 It was shown that B adsorption induces half-metallicity at the Fe3O4(100) surface. The adsorption of group IV atoms (C, Si, Ge, Sn) on the Fe3O4(100) surface has been investigated using GGA.127 The results show that all these atoms prefer to bind on the surface oxygen atom, which has no tetrahedral Fe neighbor. The adsorption structures and energies of a single Au atom on six different terminations of the Fe3O4(111) surface were computed using GGA+U.128 It was found that the Au-atom adsorption energy decreases with increasing stability of the surface. Furthermore, the results indicate that the Au atom is reduced and has a negative charge on the iron-terminated surfaces, whereas it is oxidized and has a positive charge on the oxygen-terminated surfaces. Van Natter et al.129 investigated with cluster models possible active sites on the (100), (110), and (111) surfaces of Fe3O4. Adsorption energies of oxygen adatoms located on exposed cation sites were calculated on hybrid-DFT (B3LYP) level of theory. The computed energies vary proportionally to the number of oxygen atoms missing from the normal octahedral coordination of the cation adsorption sites. A theoretical investigation of bare and water terminated NiFe2O4 surfaces was carried out in 2014 using GGA+U.130 It was found that surfaces that have more metal cations exposed are more stable. The most stable surfaces are shown to be along the (111) planes. Water adsorption on the NiFe2O4 surfaces was found to be an exothermic process. In 2014, a DFT investigation of the NiFe2O4(001) surface reported an overpotential of 0.42 V for the OER.131 It was concluded that Fe-doped β-NiOOH and NiFe2O4 could be the phases responsible for the enhanced OER activity of NiOx when it is doped with Fe.

4.

Conclusion

As many ferrites are composed of metals with known electrocatalytic properties for the OER,41 the development of cheap and varied electrode materials for PECs is a reachable target without the use of precious OER catalysts like IrO2 or RuO2. However, the existing photoelectrochemical performances need to be enhanced for economically viable PEC applications. Though the fundamental semiconductor properties are satisfactorily understood, the mechanism of charge transport and the water oxidation at the surface is not well known yet. The improvement of new synthesis strategies enabling the formation of high crystalline materials at relatively low temperature (e.g., hydrothermal and microwave synthesis) is crucial to obtain high surface area materials. In addition, thin under layers or over layers of TiO2 and Al2O3 (but also other metal oxides) are expected to enhance the photocurrent either through adjusting band alignments or passivating surface recombination centers as was revealed for other iron-based photoelectrodes.132 The formation of nanostructures with well-defined morphology, shape, and orientation will enhance the photoactivity by providing a high density of surface reaction sites and by reducing charge recombination as the size of nanostructures approaches the width of the space charge layer. Another strategy which can be exploited further is the use of heterojunctions to effectively separate the photogenerated charge carries with properly matched valence and conduction band edges. Furthermore, existing computational methods adequately predict the electronic and magnetic properties of spinel ferrites. The calculated bandgaps for some of them reasonably agree with experimental values, but for others, more refinements are needed. Most importantly, the relative band positions are key properties to determine the oxidation or reduction power, thus developing computational tools to predict the band positions is essential to understand the very scattered experimentally reported band positions of spinel ferrites.

Acknowledgments

This work is supported by the Deutsche Forschungsgemeinschaft (DFG) under the program SPP 1613 (Wa 1116/28, BR 1768/9-1, BA 1137/22-1)

References

1. 

N. S. Lewis and D. G. Nocera, “Powering the planet: chemical challenges in solar energy utilization,” Proc. Natl. Acad. Sci., 103 (43), 15729 –15735 (2006). http://dx.doi.org/10.1073/pnas.0603395103 Google Scholar

2. 

A. Kudo and Y. Miseki, “Heterogeneous photocatalyst materials for water splitting,” Chem. Soc. Rev., 38 (1), 253 –278 (2009). http://dx.doi.org/10.1039/B800489G CSRVBR 0306-0012 Google Scholar

3. 

K. Maeda and K. Domen, “Photocatalytic water splitting: recent progress and future challenges,” J. Phys. Chem. Lett., 1 (18), 2655 –2661 (2010). http://dx.doi.org/10.1021/jz1007966 JPCLCD 1948-7185 Google Scholar

4. 

S. Choudhary et al., “Nanostructured bilayered thin films in photoelectrochemical water splitting—a review,” Int. J. Hydrogen Energy, 37 (24), 18713 –18730 (2012). http://dx.doi.org/10.1016/j.ijhydene.2012.10.028 IJHEDX 0360-3199 Google Scholar

5. 

M. S. Prevot and K. Sivula, “Photoelectrochemical tandem cells for solar water splitting,” J. Phys. Chem. C, 117 (35), 17879 –17893 (2013). http://dx.doi.org/10.1021/jp405291g JPCCCK 1932-7447 Google Scholar

6. 

F. E. Osterloh, “Inorganic nanostructures for photoelectrochemical and photocatalytic water splitting,” Chem. Soc. Rev., 42 (6), 2294 –2320 (2013). http://dx.doi.org/10.1039/C2CS35266D CSRVBR 0306-0012 Google Scholar

7. 

A. A. Ismail and D. W. Bahnemann, “Photochemical splitting of water for hydrogen production by photocatalysis: a review,” Sol. Energy Mater. Sol. Cells, 128 85 –101 (2014). http://dx.doi.org/10.1016/j.solmat.2014.04.037 SEMCEQ 0927-0248 Google Scholar

8. 

R. Marschall, “Semiconductor composites: strategies for enhancing charge carrier separation to improve photocatalytic activity,” Adv. Funct. Mater., 24 (17), 2421 –2440 (2014). http://dx.doi.org/10.1002/adfm.201303214 AFMDC6 1616-301X Google Scholar

9. 

T. Hisatomi, J. Kubota and K. Domen, “Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting,” Chem. Soc. Rev., 43 (22), 7520 –7535 (2014). http://dx.doi.org/10.1039/C3CS60378D CSRVBR 0306-0012 Google Scholar

10. 

S. J. A. Moniz et al., “Visible-light driven heterojunction photocatalysts for water splitting—a critical review,” Energy Environ. Sci., 8 (3), 731 –759 (2015). http://dx.doi.org/10.1039/C4EE03271C EESNBY 1754-5692 Google Scholar

11. 

M. Sugimoto, “The past, present, and future of ferrites,” J. Am. Ceram. Soc., 82 (2), 269 –280 (1999). http://dx.doi.org/10.1111/j.1551-2916.1999.tb20058.x JACTAW 0002-7820 Google Scholar

12. 

D. S. Mathew and R. S. Juang, “An overview of the structure and magnetism of spinel ferrite nanoparticles and their synthesis in microemulsions,” Chem. Eng. J., 129 (1–3), 51 –65 (2007). http://dx.doi.org/10.1016/j.cej.2006.11.001 Google Scholar

13. 

C. Z. Yuan et al., “Mixed transition-metal oxides: design, synthesis, and energy-related applications,” Angew. Chem. Int. Ed., 53 (6), 1488 –1504 (2014). http://dx.doi.org/10.1002/anie.201303971 Google Scholar

14. 

M. S. Park et al., “Porous nanoarchitectures of spinel-type transition metal oxides for electrochemical energy storage systems,” Phys. Chem. Chem. Phys., 17 (46), 30963 –30977 (2015). http://dx.doi.org/10.1039/C5CP05936D Google Scholar

15. 

E. Casbeer, V. K. Sharma and X. Z. Li, “Synthesis and photocatalytic activity of ferrites under visible light: a review,” Sep. Purif. Technol., 87 1 –14 (2012). http://dx.doi.org/10.1016/j.seppur.2011.11.034 Google Scholar

16. 

R. Dillert et al., “Research update: photoelectrochemical water splitting and photocatalytic hydrogen production using ferrites (MFe2O4) under visible light irradiation,” APL Mater., 3 (10), 104001 (2015). http://dx.doi.org/10.1063/1.4931763 Google Scholar

17. 

B. A. Pinaud et al., “Technical and economic feasibility of centralized facilities for solar hydrogen production via photocatalysis and photoelectrochemistry,” Energy Environ. Sci., 6 (7), 1983 –2002 (2013). http://dx.doi.org/10.1039/c3ee40831k EESNBY 1754-5692 Google Scholar

18. 

M. Woodhouse and B. A. Parkinson, “Combinatorial approaches for the identification and optimization of oxide semiconductors for efficient solar photoelectrolysis,” Chem. Soc. Rev., 38 (1), 197 –210 (2009). http://dx.doi.org/10.1039/B719545C CSRVBR 0306-0012 Google Scholar

19. 

M. F. Weber and M. J. Dignam, “Efficiency of splitting water with semiconducting photoelectrodes,” J. Electrochem. Soc., 131 (6), 1258 –1265 (1984). http://dx.doi.org/10.1149/1.2115797 JESOAN 0013-4651 Google Scholar

20. 

A. B. Murphy et al., “Efficiency of solar water splitting using semiconductor electrodes,” Int. J. Hydrogen Energy, 31 (14), 1999 –2017 (2006). http://dx.doi.org/10.1016/j.ijhydene.2006.01.014 IJHEDX 0360-3199 Google Scholar

21. 

J. R. Bolton, S. J. Strickler and J. S. Connolly, “Limiting and realizable efficiencies of solar photolysis of water,” Nature, 316 (6028), 495 –500 (1985). http://dx.doi.org/10.1038/316495a0 Google Scholar

22. 

M. G. Walter et al., “Solar water splitting cells,” Chem. Rev., 110 (11), 6446 –6473 (2010). http://dx.doi.org/10.1021/cr1002326 CHREAY 0009-2665 Google Scholar

23. 

M. Gratzel, “Photoelectrochemical cells,” Nature, 414 (6861), 338 –344 (2001). http://dx.doi.org/10.1038/35104607 Google Scholar

24. 

A. L. Tiano et al., “Correlating size and composition-dependent effects with magnetic, mossbauer, and pair distribution function measurements in a family of catalytically active ferrite nanoparticles,” Chem. Mater., 27 (10), 3572 –3592 (2015). http://dx.doi.org/10.1021/acs.chemmater.5b00767 CMATEX 0897-4756 Google Scholar

25. 

A. Goldman, Modern Ferrite Technology, Springer, Pittsburgh, Pennsylvania (2006). Google Scholar

26. 

D. Levy et al., “Equation of state, structural behaviour and phase diagram of synthetic MgFe2O4, as a function of pressure and temperature,” Phys. Chem. Miner., 31 (2), 122 –129 (2004). http://dx.doi.org/10.1007/s00269-004-0380-4 PCMIDU 0342-1791 Google Scholar

27. 

R. A. Candeia et al., “Synthesis and characterization of spinel pigment CaFe2O4 obtained by the polymeric precursor method,” Mater. Lett., 58 (5), 569 –572 (2004). http://dx.doi.org/10.1016/S0167-577X(03)00563-9 MLETDJ 0167-577X Google Scholar

28. 

R. A. Candeia et al., “Monoferrite BaFe2O4 applied as ceramic pigment,” Ceram. Int., 33 (4), 521 –525 (2007). http://dx.doi.org/10.1016/j.ceramint.2005.10.018 Google Scholar

29. 

N. Helaili et al., “Synthesis and physical properties of the CuFe2xMnxO4 (0<=x<=2) solid solution,” Mater. Chem. Phys., 148 (3), 734 –743 (2014). http://dx.doi.org/10.1016/j.matchemphys.2014.08.042 MCHPDR 0254-0584 Google Scholar

30. 

W. T. Thompson et al., Uhlig’s Corrosion Handbook, 3rd ed.John Wiley & Sons, Inc., Hoboken, New Jersey (2011). Google Scholar

31. 

Y. Hemmi et al., “Electrochemical considerations regarding general corrosion of materials in a bwr primary circuit,” J. Nucl. Sci. Technol., 31 (11), 1202 –1213 (1994). http://dx.doi.org/10.1080/18811248.1994.9735277 JNSTAX 0022-3131 Google Scholar

32. 

Y.-F. Xu et al., “In situ formation of zinc ferrite modified Al-doped ZnO nanowire arrays for solar water splitting,” J. Mater. Chem. A, 4 (14), 5124 –5129 (2016). http://dx.doi.org/10.1039/C5TA10563C Google Scholar

33. 

A. A. Tahir and K. G. U. Wijayantha, “Photoelectrochemical water splitting at nanostructured ZnFe2O4 electrodes,” J. Photochem. Photobiol. A, 216 (2–3), 119 –125 (2010). http://dx.doi.org/10.1016/j.jphotochem.2010.07.032 JPPCEJ 1010-6030 Google Scholar

34. 

J. H. Kim et al., “Defective ZnFe2O4 nanorods with oxygen vacancy for photoelectrochemical water splitting,” Nanoscale, 7 (45), 19144 –19151 (2015). http://dx.doi.org/10.1039/C5NR05812K NANOHL 2040-3364 Google Scholar

35. 

A. G. Hufnagel et al., “Zinc ferrite photoanode nanomorphologies with favorable kinetics for water-splitting,” Adv. Funct. Mater., 26 (25), 4435 –4443 (2016). http://dx.doi.org/10.1002/adfm.v26.25 AFMDC6 1616-301X Google Scholar

36. 

J. Y. Cao et al., “Fabrication of p-type CaFe2O4 nanofilms for photoelectrochemical hydrogen generation,” Electrochem. Commun., 13 (3), 275 –278 (2011). http://dx.doi.org/10.1016/j.elecom.2011.01.002 ECCMF9 1388-2481 Google Scholar

37. 

S. Ida et al., “Photoelectrochemical hydrogen production from water using p-type and n-type oxide semiconductor electrodes,” Electrochim. Acta, 82 397 –401 (2012). http://dx.doi.org/10.1016/j.electacta.2012.03.174 ELCAAV 0013-4686 Google Scholar

38. 

S. Ida et al., “Preparation of p-Type CaFe2O4 photocathodes for producing hydrogen from water,” J. Am. Chem. Soc., 132 (49), 17343 –17345 (2010). http://dx.doi.org/10.1021/ja106930f JACSAT 0002-7863 Google Scholar

39. 

H. Yang et al., “Electrochemical synthesis of CoFe2O4 porous nanosheets for visible light driven photoelectrochemical applications,” New J. Chem., 37 (10), 2965 –2968 (2013). http://dx.doi.org/10.1039/c3nj00627a NJCHE5 1144-0546 Google Scholar

40. 

G. Rekhila, Y. Bessekhouad and M. Trari, “Visible light hydrogen production on the novel ferrite NiFe2O4,” Int. J. Hydrogen Energy, 38 (15), 6335 –6343 (2013). http://dx.doi.org/10.1016/j.ijhydene.2013.03.087 IJHEDX 0360-3199 Google Scholar

41. 

Y. Matsumoto et al., “Photoelectrochemical properties of the Zn-Ti-Fe spinel oxides,” J. Electrochem. Soc., 133 (4), 711 –716 (1986). http://dx.doi.org/10.1149/1.2108660 JESOAN 0013-4651 Google Scholar

42. 

Z. Simsa et al., “Optical and magneto-optical properties of magnetite and manganese ferrites,” J. Magn. Magn. Mater., 15–18 775 –776 (1980). http://dx.doi.org/10.1016/0304-8853(80)90757-X JMMMDC 0304-8853 Google Scholar

43. 

S. Balaji et al., “Combustion synthesis and characterization of Sn4+ substituted nanocrystalline NiFe2O4,” Mater. Sci. Eng. B, 119 (2), 119 –124 (2005). http://dx.doi.org/10.1016/j.mseb.2005.01.021 Google Scholar

44. 

K. N. Harish et al., “Synthesis, enhanced optical and photocatalytic study of Cd-Zn ferrites under sunlight,” Catal. Sci. Technol., 2 (5), 1033 –1039 (2012). http://dx.doi.org/10.1039/c2cy00503d Google Scholar

45. 

M. D. Archer, G. C. Morris and G. K. Yim, “Electrochemical approaches to solar-energy conversion—a brief overview and preliminary-results obtained with n-type cobalt ferrite,” J. Electroanal. Chem., 118 89 –100 (1981). http://dx.doi.org/10.1016/S0022-0728(81)80534-7 JECHES 0022-0728 Google Scholar

46. 

M. S. Antonious et al., “Photoelectrochemical characteristics of p-type and n-type polycrystalline Ni-ferrite electrodes in aqueous-solutions,” Mater. Res. Bull., 21 (12), 1515 –1523 (1986). http://dx.doi.org/10.1016/0025-5408(86)90093-0 MRBUAC 0025-5408 Google Scholar

47. 

C. G. Ramankutty and S. Sugunan, “Surface properties and catalytic activity of ferrospinels of nickel, cobalt and copper, prepared by soft chemical methods,” Appl. Catal. A, 218 (1–2), 39 –51 (2001). http://dx.doi.org/10.1016/S0926-860X(01)00610-X Google Scholar

48. 

G. K. Reddy et al., “Cr- and Ce-doped ferrite catalysts for the high temperature water-gas shift reaction: TPR and Mossbauer spectroscopic study,” J. Phys. Chem. C, 115 (4), 920 –930 (2011). http://dx.doi.org/10.1021/jp102959p JPCCCK 1932-7447 Google Scholar

49. 

Y. Matsumoto, “Energy positions of oxide semiconductors and photocatalysis with iron complex oxides,” J. Solid State Chem., 126 (2), 227 –234 (1996). http://dx.doi.org/10.1006/jssc.1996.0333 JSSCBI 0022-4596 Google Scholar

50. 

H. H. Kung et al., “Semiconducting oxide anodes in photoassisted electrolysis of water,” J. Appl. Phys., 48 (6), 2463 –2469 (1977). http://dx.doi.org/10.1063/1.324010 JAPIAU 0021-8979 Google Scholar

51. 

L. G. J. Dehaart and G. Blasse, “Photoelectrochemical properties of ferrites with the spinel structure,” J. Electrochem. Soc., 132 (12), 2933 –2938 (1985). http://dx.doi.org/10.1149/1.2113696 JESOAN 0013-4651 Google Scholar

52. 

F. A. Benko and F. P. Koffyberg, “The effect of defects on some photoelectrochemical properties of semiconducting MgFe2O4,” Mater. Res. Bull., 21 (10), 1183 –1188 (1986). http://dx.doi.org/10.1016/0025-5408(86)90045-0 MRBUAC 0025-5408 Google Scholar

53. 

H. Zazoua et al., “Enhanced photocatalytic hydrogen production under visible light over a material based on magnesium ferrite derived from layered double hydroxides (LDHs),” Int.J. Energy Res., 38 (15), 2010 –2018 (2014). http://dx.doi.org/10.1002/er.v38.15 IJERDN 0363-907X Google Scholar

54. 

M. Buchler et al., “Comparison of the semiconductive properties of sputter-deposited iron oxides with the passive film on iron,” J. Electrochem. Soc., 145 (2), 378 –385 (1998). http://dx.doi.org/10.1149/1.1838272 JESOAN 0013-4651 Google Scholar

55. 

Y. Matsumoto et al., “New photocathode materials for hydrogen evolution—CaFe2O4 and Sr7Fe10O22,” J. Phys. Chem., 91 (3), 577 –581 (1987). http://dx.doi.org/10.1021/j100287a018 JPCHAX 0022-3654 Google Scholar

56. 

Y. Matsumoto, K. Sugiyama and E. I. Sato, “Improvement of CaFe2O4 photocathode by doping with Na and Mg,” J. Solid State Chem., 74 (1), 117 –125 (1988). http://dx.doi.org/10.1016/0022-4596(88)90337-4 JSSCBI 0022-4596 Google Scholar

57. 

S. Ida et al., “Photoelectrochemical hydrogen production from water using p-type CaFe2O4 and n-Type ZnO,” Electrochemistry, 79 (10), 797 –800 (2011). http://dx.doi.org/10.5796/electrochemistry.79.797 ECHMBU 0305-9979 Google Scholar

58. 

J. Y. Cao et al., “Photoelectrochemical properties of nanomultiple CaFe2O4/ZnFe2O4 pn junction photoelectrodes,” Langmuir, 29 (9), 3116 –3124 (2013). http://dx.doi.org/10.1021/la304377z LANGD5 0743-7463 Google Scholar

59. 

K. Sekizawa et al., “Structural improvement of CaFe2O4 by metal doping toward enhanced cathodic photocurrent,” ACS Appl. Mater. Interfaces, 6 (14), 10969 –10973 (2014). http://dx.doi.org/10.1021/am502500y AAMICK 1944-8244 Google Scholar

60. 

B. T. Chang et al., “Photoelectrochemical study of a spinel-type titanomagnetite,” J. Solid State Chem., 72 (2), 201 –208 (1988). http://dx.doi.org/10.1016/0022-4596(88)90023-0 JSSCBI 0022-4596 Google Scholar

61. 

A. A. Tahir et al., “A new route to control texture of materials: nanostructured ZnFe2O4 photoelectrodes,” Int. J. Hydrogen Energy, 38 (11), 4315 –4323 (2013). http://dx.doi.org/10.1016/j.ijhydene.2013.01.130 IJHEDX 0360-3199 Google Scholar

62. 

K. Dileep et al., “Probing optical band gaps at the nanoscale in NiFe2O4 and CoFe2O4 epitaxial films by high resolution electron energy loss spectroscopy,” J. Appl. Phys., 116 (10), 103505 (2014). http://dx.doi.org/10.1063/1.4895059 JAPIAU 0021-8979 Google Scholar

63. 

C. Himcinschi et al., “Optical and magneto-optical study of nickel and cobalt ferrite epitaxial thin films and submicron structures,” J. Appl. Phys., 113 (8), 084101 (2013). http://dx.doi.org/10.1063/1.4792749 JAPIAU 0021-8979 Google Scholar

64. 

P. Xiong et al., “Ternary titania-cobalt ferrite-polyaniline nanocomposite: a magnetically recyclable hybrid for adsorption and photodegradation of dyes under visible light,” Ind. Eng. Chem. Res., 52 (30), 10105 –10113 (2013). http://dx.doi.org/10.1021/ie400739e Google Scholar

65. 

J. H. Kim et al., “Awakening solar water-splitting activity of ZnFe2O4 nanorods by hybrid microwave annealing,” Adv. Energy Mater., 5 (6), (2015). http://dx.doi.org/10.1002/aenm.201401933 ADEMBC 1614-6840 Google Scholar

66. 

K. J. McDonald and K. S. Choi, “Synthesis and photoelectrochemical properties of Fe2O3/ZnFe2O4 composite photoanodes for use in solar water oxidation,” Chem. Mater., 23 (21), 4863 –4869 (2011). http://dx.doi.org/10.1021/cm202399g CMATEX 0897-4756 Google Scholar

67. 

R. Dom et al., “Eco-friendly ferrite nanocomposite photoelectrode for improved solar hydrogen generation,” RSC Adv., 3 (35), 15217 –15224 (2013). http://dx.doi.org/10.1039/c3ra42051e Google Scholar

68. 

Y. H. Guo et al., “Photoelectrochemical activity of ZnFe2O4 modified alpha-Fe2O3 nanorod array films,” RSC Adv., 4 (70), 36967 –36972 (2014). http://dx.doi.org/10.1039/C4RA05289G Google Scholar

69. 

M. G. Ahmed et al., “Enhanced photoelectrochemical water oxidation on nanostructured hematite photoanodes via p-CaFe2O4/n-Fe2O3 heterojunction formation,” J. Phys. Chem. C, 119 (11), 5864 –5871 (2015). http://dx.doi.org/10.1021/jp512804p JPCCCK 1932-7447 Google Scholar

70. 

J. Y. Kim et al., “Single-crystalline, wormlike hematite photoanodes for efficient solar water splitting,” Sci. Rep., 3 2681 (2013). http://dx.doi.org/10.1038/srep02681 Google Scholar

71. 

E. S. Kim et al., “Fabrication of CaFe2O4/TaON heterojunction photoanode for photoelectrochemical water oxidation,” J. Am. Chem. Soc., 135 (14), 5375 –5383 (2013). http://dx.doi.org/10.1021/ja308723w JACSAT 0002-7863 Google Scholar

72. 

H. G. Kim et al., “Fabrication of CaFe2O4/MgFe2O4 bulk heterojunction for enhanced visible light photocatalysis,” Chem. Commun., 39 5889 –5891 (2009). http://dx.doi.org/10.1039/b911805e Google Scholar

73. 

X. L. Zheng et al., “ZnFe2O4 leaves grown on TiO2 trees enhance photoelectrochemical water splitting,” Small, 12 (23), 3181 –3188 (2016). http://dx.doi.org/10.1002/smll.v12.23 SMALBC 1613-6810 Google Scholar

74. 

E. S. Kim et al., “Improved photoelectrochemical activity of CaFe2O4/BiVO4 heterojunction photoanode by reduced surface recombination in solar water oxidation,” ACS Appl. Mater. Interfaces, 6 (20), 17762 –17769 (2014). http://dx.doi.org/10.1021/am504283t AAMICK 1944-8244 Google Scholar

75. 

M. W. Kanan and D. G. Nocera, “In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+,” Science, 321 (5892), 1072 –1075 (2008). http://dx.doi.org/10.1126/science.1162018 Google Scholar

76. 

M. W. Kanan, Y. Surendranath and D. G. Nocera, “Cobalt-phosphate oxygen-evolving compound,” Chem. Soc. Rev., 38 (1), 109 –114 (2009). http://dx.doi.org/10.1039/B802885K CSRVBR 0306-0012 Google Scholar

77. 

Y. Surendranath, M. W. Kanan and D. G. Nocera, “Mechanistic studies of the oxygen evolution reaction by a cobalt-phosphate catalyst at neutral pH,” J. Am. Chem. Soc., 132 (46), 16501 –16509 (2010). http://dx.doi.org/10.1021/ja106102b JACSAT 0002-7863 Google Scholar

78. 

G. M. Carroll, D. K. Zhong and D. R. Gamelin, “Mechanistic insights into solar water oxidation by cobalt-phosphate-modified alpha-Fe2O3 photoanodes,” Energy Environ. Sci., 8 (2), 577 –584 (2015). http://dx.doi.org/10.1039/C4EE02869D EESNBY 1754-5692 Google Scholar

79. 

A. Gruneis et al., “Ionization potentials of solids: the importance of vertex corrections,” Phys. Rev. Lett., 112 (9), 096401 (2014). http://dx.doi.org/10.1103/PhysRevLett.112.096401 PRLTAO 0031-9007 Google Scholar

80. 

M. Marsman et al., “Hybrid functionals applied to extended systems,” J. Phys: Condens. Matter, 20 (6), 064201 (2008). http://dx.doi.org/10.1088/0953-8984/20/6/064201 Google Scholar

81. 

S. L. Dudarev et al., “Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study,” Phys. Rev. B, 57 (3), 1505 –1509 (1998). http://dx.doi.org/10.1103/PhysRevB.57.1505 Google Scholar

82. 

D. Odkhuu et al., “A first-principles study of magnetostrictions of Fe3O4 and CoFe2O4,” J. Appl. Phys., 115 (17), 17A916 (2014). http://dx.doi.org/10.1063/1.4863811 JAPIAU 0021-8979 Google Scholar

83. 

K. Obata et al., “Electronic structure of CaFe2O4 with antiferromagnetic spin ordering,” J. Ceram. Soc. Jpn., 121 (1417), 766 –769 (2013). http://dx.doi.org/10.2109/jcersj2.121.766 JCSJEW 0914-5400 Google Scholar

84. 

D. H. Ji et al., “Quantum mechanical method for estimating ionicity of spinel ferrites,” J. Magn. Magn. Mater., 326 197 –200 (2013). http://dx.doi.org/10.1016/j.jmmm.2012.09.016 JMMMDC 0304-8853 Google Scholar

85. 

Q. C. Sun et al., “Optical band gap hierarchy in a magnetic oxide: electronic structure of NiFe2O4,” Phys. Rev. B, 86 (20), 205106 (2012). http://dx.doi.org/10.1103/PhysRevB.86.205106 Google Scholar

86. 

R. Dom et al., “Synthesis of solar active nanocrystalline ferrite, MFe2O4 (M: Ca, Zn, Mg) photocatalyst by microwave irradiation,” Solid State Commun., 151 (6), 470 –473 (2011). http://dx.doi.org/10.1016/j.ssc.2010.12.034 SSCOA4 0038-1098 Google Scholar

87. 

S. Soliman et al., “Electronic structure calculations for ZnFe2O4,” Phys. Rev. B, 83 (8), 085205 (2011). http://dx.doi.org/10.1103/PhysRevB.83.085205 Google Scholar

88. 

M. Feng et al., “Ab initio study on copper ferrite,” J. Appl. Phys., 107 (9), 09A521 (2010). http://dx.doi.org/10.1063/1.3338905 JAPIAU 0021-8979 Google Scholar

89. 

D. M. Wells et al., “Local electronic and magnetic structure of mixed ferrite multilayer materials,” Phys. Rev. B, 81 (17), 174422 (2010). http://dx.doi.org/10.1103/PhysRevB.81.174422 Google Scholar

90. 

C. Cheng and C.-S. Liu, “Effects of cation distribution in ZnFe2O4 and CdFe2O4: ab initio studies,” J. Phys.: Conf. Ser, 145 012028 (2009). http://dx.doi.org/10.1088/1742-6596/145/1/012028 JPCSDZ 1742-6588 Google Scholar

91. 

X. Zuo et al., “A computational study of nickel ferrite,” J. Magn. Magn. Mater., 303 (2), E432 –E435 (2006). http://dx.doi.org/10.1016/j.jmmm.2006.01.102 JMMMDC 0304-8853 Google Scholar

92. 

Z. Szotek et al., “Electronic structures of normal and inverse spinel ferrites from first principles,” Phys. Rev. B, 74 (17), 174431 (2006). http://dx.doi.org/10.1103/PhysRevB.74.174431 Google Scholar

93. 

V. N. Antonov, B. N. Harmon and A. N. Yaresko, “Electronic structure and x-ray magnetic circular dichroism in Fe3O4 and Mn-, Co-, or Ni-substituted Fe3O4,” Phys. Rev. B, 67 (2), 024417 (2003). http://dx.doi.org/10.1103/PhysRevB.67.024417 Google Scholar

94. 

X. Zuo and C. Vittoria, “Calculation of exchange integrals and electronic structure for manganese ferrite,” Phys. Rev. B, 66 (18), 184420 (2002). http://dx.doi.org/10.1103/PhysRevB.66.184420 Google Scholar

95. 

D. J. Singh, M. Gupta and R. Gupta, “First-principles investigation of MnFe2O4,” Phys. Rev. B, 65 (6), 064432 (2002). http://dx.doi.org/10.1103/PhysRevB.65.064432 Google Scholar

96. 

N. Y. Safontseva and I. Y. Nikiforov, “On the shape of iron K absorption edges for monoferrites with a Me(Mg, Mn, Ni, Zn) Fe2O4 spinel structure,” Phys. Solid State, 43 (1), 61 –64 (2001). http://dx.doi.org/10.1134/1.1340188 PSOSED 1063-7834 Google Scholar

97. 

W. F. Pong et al., “Oxygen 1s x-ray-absorption near-edge structure of Zn-Ni ferrites: a comparison with the theoretical calculations,” Phys. Rev. B, 54 (23), 16641 –16645 (1996). http://dx.doi.org/10.1103/PhysRevB.54.16641 Google Scholar

98. 

W. C. Mackrodt and E. A. Simson, “Cation valence charge states of MnFe2O4: an ab initio Hartree-Fock study,” J. Chem. Soc. Faraday Trans., 92 (12), 2043 –2047 (1996). http://dx.doi.org/10.1039/FT9969202043 Google Scholar

99. 

M. Penicaud et al., “Calculated electronic band-structure and magnetic-moments of ferrites,” J. Magn. Magn. Mater., 103 (1–2), 212 –220 (1992). http://dx.doi.org/10.1016/0304-8853(92)90255-M JMMMDC 0304-8853 Google Scholar

100. 

K. Jordan et al., “Scanning tunneling spectroscopy study of the electronic structure of Fe3O4 surfaces,” Phys. Rev. B, 74 (8), 085416 (2006). http://dx.doi.org/10.1103/PhysRevB.74.085416 Google Scholar

101. 

D. A. Andersson and C. R. Stanek, “Mixing and non-stoichiometry in Fe-Ni-Cr-Zn-O spinel compounds: density functional theory calculations,” Phys. Chem. Chem. Phys., 15 (37), 15550 –15564 (2013). http://dx.doi.org/10.1039/c3cp50312g PPCPFQ 1463-9076 Google Scholar

102. 

H. T. Jeng, G. Y. Guo and D. J. Huang, “Charge-orbital ordering and Verwey transition in magnetite,” Phys. Rev. Lett., 93 (15), 156403 (2004). http://dx.doi.org/10.1103/PhysRevLett.93.156403 PRLTAO 0031-9007 Google Scholar

103. 

I. Leonov et al., “Electronic structure of charge-ordered Fe3O4 from calculated optical, magneto-optical Kerr effect, and OK-edge x-ray absorption spectra,” Phys. Rev. B, 74 (16), 165117 (2006). http://dx.doi.org/10.1103/PhysRevB.74.165117 Google Scholar

104. 

A. D. Rowan, C. H. Patterson and L. V. Gasparov, “Hybrid density functional theory applied to magnetite: crystal structure, charge order, and phonons,” Phys. Rev. B, 79 (20), 205103 (2009). http://dx.doi.org/10.1103/PhysRevB.79.205103 Google Scholar

105. 

G. K. H. Madsen and P. Novak, “Charge order in magnetite. An LDA+U study,” Europhys. Lett., 69 (5), 777 –783 (2005). http://dx.doi.org/10.1209/epl/i2004-10416-x EULEEJ 0295-5075 Google Scholar

106. 

Z. Szotek et al., “Ab initio study of charge order in Fe3O4,” Phys. Rev. B, 68 (5), 054415 (2003). http://dx.doi.org/10.1103/PhysRevB.68.054415 Google Scholar

107. 

H. G. Kim et al., “Photocatalytic nanodiodes for visible-light photocatalysis,” Angew. Chem. Int. Ed., 44 (29), 4585 –4589 (2005). http://dx.doi.org/10.1002/(ISSN)1521-3773 Google Scholar

108. 

P. H. Borse et al., “Synthesis of barium ferrite for visible light photocatalysis applications,” J. Korean Phys. Soc., 58 (6), 1672 –1676 (2011). http://dx.doi.org/10.3938/jkps.58.1672 KPSJAS 0374-4884 Google Scholar

109. 

C. J. O’Brien, Z. Rak and D. W. Brenner, “Free energies of (Co, Fe, Ni, Zn) Fe2O4 spinels and oxides in water at high temperatures and pressure from density functional theory: results for stoichiometric NiO and NiFe2O4 surfaces,” J. Phys.: Condens. Matter, 25 (44), 445008 (2013). http://dx.doi.org/10.1088/0953-8984/25/44/445008 Google Scholar

110. 

D. Fritsch and C. Ederer, “Epitaxial strain effects in the spinel ferrites CoFe2O4 and NiFe2O4 from first principles,” Phys. Rev. B, 82 (10), 104117 (2010). http://dx.doi.org/10.1103/PhysRevB.82.104117 Google Scholar

111. 

Z. Rak, C. J. O’Brien and D. W. Brenner, “First-principles investigation of boron defects in nickel ferrite spinel,” J. Nucl. Mater., 452 (1–3), 446 –452 (2014). http://dx.doi.org/10.1016/j.jnucmat.2014.05.031 JNUMAM 0022-3115 Google Scholar

112. 

A. Kezzim et al., “Visible light induced hydrogen on the novel hetero-system CuFe2O4/TiO2,” Energy Convers. Manage., 52 (8–9), 2800 –2806 (2011). http://dx.doi.org/10.1016/j.enconman.2011.02.014 ECMADL 0196-8904 Google Scholar

113. 

P. H. Borse et al., “Photocatalytic hydrogen generation from water-methanol mixtures using nanocrystalline ZnFe2O4 under visible light irradiation,” J. Korean Phys. Soc., 55 (4), 1472 –1477 (2009). http://dx.doi.org/10.3938/jkps.55.1472 KPSJAS 0374-4884 Google Scholar

114. 

S. Boumaza et al., “Visible light induced hydrogen evolution on new hetero-system ZnFe2O4/SrTiO3,” Appl. Energy, 87 (7), 2230 –2236 (2010). http://dx.doi.org/10.1016/j.apenergy.2009.12.016 Google Scholar

115. 

R. Dom et al., “Synthesis of a hydrogen producing nanocrystalline ZnFe2O4 visible light photocatalyst using a rapid microwave irradiation method,” RSC Adv., 2 (33), 12782 –12791 (2012). http://dx.doi.org/10.1039/c2ra21910g Google Scholar

116. 

J. H. Yao et al., “Density functional theory investigations on the structure and electronic properties of normal spinel ZnFe2O4,” Integr. Ferroelectr., 145 (1), 17 –23 (2013). http://dx.doi.org/10.1080/10584587.2013.788310 IFEREU 1058-4587 Google Scholar

117. 

A. Elfalaky and S. Soliman, “Theoretical investigation of MnFe2O4,” J. Alloys Compd., 580 401 –406 (2013). http://dx.doi.org/10.1016/j.jallcom.2013.05.197 JALCEU 0925-8388 Google Scholar

118. 

M. Padervand et al., “An experimental and theoretical study on the structure and photoactivity of XFe2O4 (X=Mn, Fe, Ni, Co, and Zn) structures,” Russ. J. Phys. Chem. A, 88 (13), 2451 –2461 (2014). http://dx.doi.org/10.1134/S0036024414130184 Google Scholar

119. 

H. M. Widatallah et al., “Atomistic simulation and ab initio study of the defect structure of spinel-related Li0.50.5xMgxFe2.50.5xO4,” Mater. Res. Bull., 47 (12), 3995 –4000 (2012). http://dx.doi.org/10.1016/j.materresbull.2012.08.048 MRBUAC 0025-5408 Google Scholar

120. 

Y. H. Hou et al., “Structural, electronic and magnetic properties of partially inverse spinel CoFe2O4: a first-principles study,” J. Phys. D: Appl. Phys., 43 (44), 445003 (2010). http://dx.doi.org/10.1088/0022-3727/43/44/445003 JPAPBE 0022-3727 Google Scholar

121. 

D. Santos-Carballal et al., “First-principles study of the inversion thermodynamics and electronic structure of FeM2X4 (thio) spinels (M=Cr, Mn, Co, Ni; X=O, S),” Phys. Rev. B, 91 (19), 195106 (2015). http://dx.doi.org/10.1103/PhysRevB.91.195106 Google Scholar

122. 

X. Shi et al., “Structure of the NiFe2O4(001) surface in contact with gaseous O-2 and water vapor,” Surf. Sci., 640 73 –79 (2015). http://dx.doi.org/10.1016/j.susc.2015.03.012 SUSCAS 0039-6028 Google Scholar

123. 

P. V. Kumar et al., “High surface reactivity and water adsorption on NiFe2O4 (111) surfaces,” J. Phys. Chem. C, 117 (11), 5678 –5683 (2013). http://dx.doi.org/10.1021/jp309434a JPCCCK 1932-7447 Google Scholar

124. 

Z. Jiang et al., “Adsorption of NO molecule on spinel-type CuFe2O4 surface: a first-principles study,” J. Phys. Chem. C, 115 (26), 13035 –13040 (2011). http://dx.doi.org/10.1021/jp203492j JPCCCK 1932-7447 Google Scholar

125. 

R. Bliem et al., “Adsorption and incorporation of transition metals at the magnetite Fe3O4(001) surface,” Phys. Rev. B, 92 (7), 075440 (2015). http://dx.doi.org/10.1103/PhysRevB.92.075440 Google Scholar

126. 

X. Sun, A. Pratt and Y. Yamauchi, “Half-metallicity induced by boron adsorption on an Fe3O4(100) surface,” Phys. Chem. Chem. Phys., 17 (23), 15386 –15391 (2015). http://dx.doi.org/10.1039/C5CP02466H PPCPFQ 1463-9076 Google Scholar

127. 

X. Sun et al., “Significant variation of surface spin polarization through group IV atom (C, Si, Ge, Sn) adsorption on Fe3O4(100),” Phys. Chem. Chem. Phys., 16 (1), 95 –102 (2014). http://dx.doi.org/10.1039/C3CP53272K PPCPFQ 1463-9076 Google Scholar

128. 

X. H. Yu et al., “Single gold atom adsorption on the Fe3O4(111) surface,” J. Phys. Chem. C, 116 (19), 10632 –10638 (2012). http://dx.doi.org/10.1021/jp301313u Google Scholar

129. 

R. M. Van Natter, J. S. Coleman and C. R. F. Lund, “DFT models for active sites on high temperature water-gas shift catalysts,” J. Mol. Catal. A: Chem., 292 (1–2), 76 –82 (2008). http://dx.doi.org/10.1016/j.molcata.2008.07.015 Google Scholar

130. 

C. J. O’Brien, Z. Rak and D. W. Brenner, “Calculated stability and structure of nickel ferrite crystal surfaces in hydrothermal environments,” J. Phys. Chem. C, 118 (10), 5414 –5423 (2014). http://dx.doi.org/10.1021/jp5002308 JPCCCK 1932-7447 Google Scholar

131. 

Y. F. Li and A. Selloni, “Mechanism and activity of water oxidation on selected surfaces of pure and Fe-doped NiOx,” ACS Catal., 4 (4), 1148 –1153 (2014). http://dx.doi.org/10.1021/cs401245q Google Scholar

132. 

L. Steier et al., “Understanding the role of underlayers and overlayers in thin film hematite photoanodes,” Adv. Funct. Mater., 24 (48), 7681 –7688 (2014). http://dx.doi.org/10.1002/adfm.v24.48 AFMDC6 1616-301X Google Scholar

Biography

Dereje H. Taffa is a postdoctoral research associate in the Institute of Chemistry at the Carl von Ossietzky University Oldenburg. He studied chemistry at Addis Ababa University in Ethiopia and received his BSc degree in chemistry and MSc degree in physical chemistry. He earned his PhD in chemistry from Osnabrueck University. His research interest includes preparation and characterization of oxide based semiconductors thin films for energy storage and conversion applications.

Ralf Dillert is working since 2006 as a chemist at the Institute of Technical Chemistry of the Leibniz University Hannover. His actual research interests comprise photocatalytic water treatment, photocatalytic removal of air pollutants, self-cleaning surfaces, and the interfacial electron transfer at the electrolyte/semiconductor interface.

Anna C. Ulpe is a PhD student of chemistry in the Mulliken Center for Theoretical Chemistry at the University of Bonn. Her current field of research is the computational investigation of optical properties of semiconductors.

Katharina C. L. Bauerfeind studied chemistry and is working towards her PhD in the Mulliken Center for Theoretical Chemistry at the University of Bonn. Currently she is investigating the photocatalytic properties of binary and ternary transition metal oxides with computational means.

Thomas Bredow studied chemistry and received his MSc degree and PhD in chemistry from the Leibniz University of Hannover. He is currently a professor of theoretical chemistry in the Mulliken Center for Theoretical Chemistry at the University of Bonn. His research interest includes quantum-chemical modeling of solids and surfaces, defect formation and migration, heterogeneous catalysis, and semiempirical methods.

Detlef W. Bahnemann is a supernumerary professor and head of Photocatalysis and Nanotechnology Research Unit at the Institute of Technical Chemistry, Leibniz University of Hannover and the director of Megagrant Laboratory “Photoactive Nanocomposite Materials” at St. Petersburg State University in Russia. He received his MSc degree and PhD in chemistry from the Technical University Berlin. His research focuses on photocatalysis and ultrafast photocatalytic processes, solar water splitting and photoelectrochemical solar cells.

Michael Wark is a professor of technical chemistry and the head of the research group of photocatalysis and sustainable feedstock utilization in the Institute of Chemistry at Carl von Ossietzky University Oldenburg. He studied chemistry and obtained his MSc degree and his PhD in chemistry from the University of Bremen. His research interest includes materials for renewable energy applications, micro-and mesoporous structures, ordered mesoporous thin films and inorganic-organic hybrid structures.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Dereje H. Taffa, Ralf Dillert, Anna C. Ulpe, Katharina C. L. Bauerfeind, Thomas Bredow, Detlef W. Bahnemann, and Michael Wark "Photoelectrochemical and theoretical investigations of spinel type ferrites (MxFe3xO4) for water splitting: a mini-review," Journal of Photonics for Energy 7(1), 012009 (4 October 2016). https://doi.org/10.1117/1.JPE.7.012009
Published: 4 October 2016
Lens.org Logo
CITATIONS
Cited by 120 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Iron

Spinel

Electrodes

Semiconductors

Water splitting

Metals

Oxides

Back to Top