Development of ternary iron vanadium oxide semiconductors for applications in photoelectrochemical water oxidation

Harahari Mandala, Sanjib Shyamala, Paramita Hajraa, Aparajita Beraa, Debasis Sariketa, Sukumar Kundub and Chinmoy Bhattacharya*a
aDepartment of Chemistry, Indian Institute of Engineering Science & Technology, (IIEST), Shibpur, Howrah – 711103, West Bengal, India. E-mail: cbhattacharya.besus@gmail.com
bDepartment of Metallurgy & Materials Engineering, Indian Institute of Engineering Science & Technology, (IIEST), Shibpur, Howrah – 711103, West Bengal, India

Received 28th October 2015 , Accepted 22nd December 2015

First published on 23rd December 2015


Abstract

Herein, we report the synthesis of Fe–V-oxides via the drop casting of metal precursor solutions in different proportions onto an indium tin oxide (ITO) coated glass followed by annealing in air at 500 °C for 3 h. UV-vis spectroscopy of the Fe–V-oxides indicates absorption due to ‘direct’ and ‘indirect’ band gaps, although Fe-oxide shows a direct band gap nature. Scanning electron microscopy-energy dispersive X-ray (SEM-EDX) and X-ray diffraction (XRD) studies reveal different surface morphologies with variable crystalline phases for the Fe2O3, FeVO4, FeV2O4 and Fe2VO4 semiconductors. The photoelectrochemical (PEC) water oxidation reaction over the different materials reveals that the FeV2O4 semiconductor exhibits the maximum photocurrent of 0.18 mA cm−2 at an applied bias of +1.0 V (vs. Ag/AgCl) under the illumination of 100 mW cm−2 compared to the other Fe2O3, FeVO4 and Fe2VO4 semiconductors. Electrochemical impedance spectroscopic (Mott–Schottky) analysis confirms n-type semiconductivity for all the materials with highest donor density, in the order of 2.7 × 1020 cm−3, for the FeV2O4 thin film, and PL spectra are useful for measuring the separation efficiency of the photo-generated charge carriers. Chronoamperometric studies under constant illumination of the best semiconductor (FeV2O4) indicate the significant stability of the material, and photoelectrochemical action spectra demonstrate 22% incident photon to current conversion efficiency (IPCE) and 60% absorbed photon to current conversion efficiency (APCE).


1. Introduction

We prepared a stable and low cost Fe-based semiconductor (SC), which can absorb a large portion of solar photons while having a minimum band gap energy. Most stable semiconductors absorb almost exclusively in the UV region of the solar spectrum.1,2 Iron(III) oxide (Fe2O3), which has a band gap of ∼2.2 eV, can absorb most of the visible light (300 to 560 nm), of solar radiation, which comprises 38% of the photons of sunlight in air mass (AM) 1.5 spectra. Even though the band gap of Fe2O3 is suitable to allow absorption of a significant amount of sunlight photons, its photoresponse is quite low, which is mainly due to its low electrical conductivity and high recombination rate of photogenerated electron–hole pairs.3,4 To minimize these limitations, n-type iron(III) oxide SC has been modified by incorporating different proportions of vanadium5,6 into its matrix with the aim to enhance its photoelectrochemical water oxidation1,7,8 behavior. The goal of this work is to synthesize Fe–V-oxide films of different compositions using a drop cast method and determine their efficacy toward the photoelectrochemical water splitting process.9,10 Since solar hydrogen is a sustainable and environment-friendly energy carrier, it is considered to replace fossil fuels in the near future. It can be generated by the splitting of water under solar light illumination.11–13 The search for stable, efficient and affordable PEC electrode materials14–17 (photoanodes and photocathodes) is an on-going quest. Hematite, α-Fe2O3, remains an intriguing choice as an oxygen-evolving photoanode material because of its environmentally benign chemical stability,18,19 low price and visible absorptivity character.20 Recent theoretical and computational works support that hematite remains a promising parent material for artificial photosynthesis21 by undergoing water oxidation and CO2 reduction reactions. Its bulk and surface electronic structure has been under scrutiny for many decades and is considered for the improvement of efficiency.18

Herein, we report the synthesis of a visible light responsive V incorporated Fe-oxide compound SC. This composite demonstrates photocatalytic activity for oxygen evolution from water, under UV-vis illumination. Fe–V-oxide22 compound theoretical calculations predict a decrease of ∼2.8 eV in its band gap with the Fermi level in the middle of the valence band maximum and conduction band minimum.

This study further demonstrates the improvement in its visible-light photocatalytic behavior for both water splitting and solar-energy conversion.

2. Experimental section

2.1. Reagents

Iron(III) nitrate nonahydrate [Fe(NO3)3·9H2O] and ammonium metavanadate (NH4VO3) were purchased from Sigma-Aldrich. NaH2PO4, Na2HPO4, Na2SO4, Na2SO3, and ethylene glycol (EG) were purchased from Merck (AR quality) and used as received. All metal precursor solutions were prepared in ethylene glycol (EG) at a 0.1 M concentration. Milli-Q grade water was used in all experiments to prepare solutions. 2.0 × 1.5 cm2 ITO-coated glass slides (Xinyan technology Ltd, Hong Kong) were thoroughly cleaned via sonication in successive solutions of soap water followed by ethanol and finally rinsed with Milli-Q water.

2.2. Preparation of thin film electrodes through the drop casting technique

Appropriate solutions with Fe[thin space (1/6-em)]:[thin space (1/6-em)]V (v/v) ratios of 1[thin space (1/6-em)]:[thin space (1/6-em)]0, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 2[thin space (1/6-em)]:[thin space (1/6-em)]1 were prepared in EG in separate glass vials, and the mixtures were kept in an ultrasonication bath for ∼1 hour. 500 μL of each mixtures was dropped cast onto a cleaned ITO substrate to prepare semiconductor films. The thin films were annealed in air at 500 °C with a heating ramping rate of 1 °C min−1 followed by soaking for 3 h to obtain uniform, well-adhered thin films. The thickness (t) of the films was measured gravimetrically using the following relation:
 
t = m/ρA (1)
where m and A stand for mass and area of the deposited film and ρ the density of the Fe2O3 and Fe–V-oxide films. The thickness was found to be 260 nm for Fe2O3, 396 nm for FeVO4, 480 for FeV2O4 nm and 481 nm for Fe2VO4.

2.3. Optical characterization

Absorption spectra of the prepared thin films were obtained using a V-630 UV-vis spectrophotometer (Jasco, Japan) within the wavelength range of 350–700 nm, and the band gap energy was calculated from the absorption edge on the wavelength scale.

2.4. Photoluminescence study

Photoluminescence spectra of the Fe2O3 and Fe–V-oxide thin film semiconductors were obtained using a PerkinElmer LS 55 Fluorescence Spectrometer at an excitation wavelength of 360 nm.

2.5. Surface characterization via SEM, XRD and EDX measurements

The surface morphology and quality of the film on the ITO coated glass substrate were determined via scanning electron microscopic (SEM) analysis using a HITACHI S3400N Instrument. Elemental composition was further determined by energy dispersive X-ray (EDX) spectroscopic analyses of the samples23 using a 7021-H HORIBA EMAX Instrument installed with the SEM. The crystalline behavior of the prepared semiconductor was evaluated using the powder X-ray diffraction (XRD) technique on a Rigaku MiniFlex II, operated with a Cu Kα target (λ = 1.54 Å) within the 2θ range from 20° to 80° at a fixed scan rate of 2° min−1.

2.6. Photoelectrochemical measurement

Photocurrent density measurements were performed by linear sweep voltammetry (LSV) using a CHI-650 potentiostat (CH Instrument, Austin, TX) in a standard three-electrode configuration cell (0.27 cm2 geometric area exposed, using one O-ring of the same inner area) with a scan rate of 10 mV s−1. The as prepared Fe–V-oxide films served as the photoanode, Pt wire as the counter electrode and Ag/AgCl as the reference electrode and all the potentials reported herein are with respect to this reference electrode. Measurements were carried out either in 0.1 M Na2SO4 with 0.1 M Na2SO3 as a sacrificial electron donor within the potential range from −0.1 to 0.6 V or in the presence of 0.1 M Na2SO4 in 0.1 M phosphate buffer solution (PBS) for water oxidation within the range from 0.0 to 1.2 V. 100 mW cm−2 illumination was maintained using a 300 W Xe arc lamp (Excelitas USA) as the light source in this study. Visible-light responsiveness of the semiconductors towards the photo-oxidation of water was determined by holding a UV cut-off filter (λ > 420 nm) in front of the Xe lamp. PEC measurements were conducted by irradiating the prepared electrodes through the electrolyte/electrode interface.

2.7. Chronoamperometric measurement

The stability of the thin-film semiconductor electrode undergoing a solar driven water oxidation reaction was evaluated via continuous photocurrent measurements for 30 min in 0.1 M Na2SO4–PBS when the electrode was held at a constant potential of 1.0 V under 100 mW cm−2 illumination. A monochromator (Oriel) was used in combination with a power meter and a silicon detector (Newport) to measure the photocurrent spectrum. From the electrochemical action spectra, incident photon to current conversion efficiency (IPCE) and absorbed photon to current conversion efficiency (APCE) were calculated.

2.8. Electrochemical impedance spectroscopic analysis of the semiconductor–electrolyte interface

Electrochemical impedance spectroscopic (EIS) measurements were conducted using a similar experiment setup as stated above, at a constant applied potential of 1.0 V within the frequency range from 20 mHz to 100 kHz with the oscillation amplitude of 10 mV. Experiments were conducted in dark and illuminated (UV-vis and only visible) conditions. A Mott–Schottky experiment was conducted in 0.1 M Na2SO4 in 0.1 M phosphate buffer solution (PBS) using different frequencies of 200, 500 and 1000 Hz at an applied potential ranging from −0.4 to 0.8 V, maintaining an ac amplitude of 10 mV at each of the potentials.

3. Results and discussion

3.1. Physicochemical characterization

Thin film Fe–V-oxide semiconductors were synthesized via drop-casting using Fe3+ and V5+ as precursors in the EG solution. The colors of the individual precursor solution and the corresponding semiconductor thin films developed on ITO coated glass substrates are presented in Fig. S1(a and b) (ESI).

It has been observed that with the gradual addition of V to the pure Fe2O3 matrix, the color of the solution as well as that of the films changes from deep reddish to light yellow. Fig. 1a represents the UV-visible absorption spectra of the different thin film semiconductors, and Fig. 1b (inset) represents the Tauc plot for the calculation of band gap energy as well as to determine the nature of the band-to-band transition pattern. The Tauc plot indicates absorption characteristics due to the distinct ‘direct’ band gap at ∼1.9 eV and ‘indirect’ band gap at 2.6–2.7 eV for the Fe–V-oxides, which impart yellow color to the film, whereas the presence of a pure ‘direct’ band gap ∼2.0 eV for the Fe-oxide causes the film to be reddish in color. The thickness of the different materials is found to vary as 260 nm for Fe2O3, 396 nm for FeVO4, 480 for FeV2O4 nm and 481 nm for Fe2VO4.


image file: c5ra22586h-f1.tif
Fig. 1 (a) Absorption spectra for the determination of the band gap energy of the Fe–V-oxide compound thin film semiconductor prepared with 500 μL solution on the glass substrate. (b) Inset: Tauc plot of same film semiconductors showing ‘direct’ and ‘indirect’ gaps.

Photoluminescence (PL) emission spectra characterize the recombination of free carriers, and thus highlight the separation efficiency of the photo-generated charge carriers in a semiconductor.24–26 The higher the PL intensity, the greater the probability of charge carrier recombination.27 A comparison of the PL spectra (excited at 360 nm) of Fe2O3 and the Fe–V-oxide compounds at room temperature is presented in Fig. 2. The PL spectrum indicates a strong peak at ∼480 nm, which corresponds to the high-level transition in Fe–V-oxide. Among the different samples, FeV2O4, which has the lowest intensity in that particular region, demonstrates that the recombination process can be effectively controlled inside the matrix. Therefore, FeV2O4 is very effective for separating the photo-generated charge carriers, which in turns triggers the charge transfer process at a faster rate than the other semiconductor thin films.


image file: c5ra22586h-f2.tif
Fig. 2 PL spectra of the Fe–V-oxide semiconductors excited at 360 nm.

SEM analysis reveals the surface morphology of the thin-film semiconductors. Fig. 3a–d represent the SEM images for the Fe–V-oxide compounds grown over ITO coated glass substrates. For the pure iron oxide material, as presented in Fig. 3a, the surfaces are found to be covered with agglomerated particles. With the addition of 33% V to the matrix, the surface morphology (Fig. 3b) changes significantly with the appearance of ‘cubic’ particles of approximate size of 250–300 nm. Upon the addition of 50% V, the morphology changes to a ‘3D triclinic’ structure of average length 800–1000 nm along with presence of some cubic structures (Fig. 3c). With the further addition of to V (70% V), the surface was found to be covered with small particles along with ‘cylindrical’ shaped particles with an average diameter in the order of 300–400 nm (Fig. 3d).


image file: c5ra22586h-f3.tif
Fig. 3 (a–d) SEM images of (a) Fe2O3; (b) Fe2VO4; (c) FeVO4 and (d) FeV2O4 films on ITO coated glass substrates annealed at 500 °C for 3 h.

The EDX analysis of the different semiconductor thin films measures the individual elemental compositions (atomic percentage, %) of the constituents. Fig. 4a–d represent the elemental mapping of the Fe–V-oxide compounds, whereas Fig. 4e (bar plot) indicates the relative proportions of the individual elements in the semiconductor matrix. The composition analysis corresponds well to the prepared semiconductors developed via the drop-casting techniques.


image file: c5ra22586h-f4.tif
Fig. 4 (a–e) EDX elemental mapping of (a) Fe2O3, (b) Fe2VO4, (c) FeVO4, and (d) FeV2O4. (e) Atomic percentage (%) of Fe (red), V (yellow) and O (green).

The crystalline nature of the as-prepared samples was examined via the XRD technique. Fig. 5 represents the XRD pattern for pure Fe-oxide, which is identified as α-Fe2O3 (hematite) compounds with a ‘rhombohedral’ geometry (JCPDS file 79-1741). The addition of different levels of vanadium to the Fe-oxide leads to a change in the crystallinity of the semiconductors to a significant extent. The material with 33% V added has been identified as Fe2VO4 using JCPDS file 75-1519, which has a ‘cubic’ structure, whereas the 50% V added film is identified as FeVO4 with JCPDS file 71-1592, which shows a preferential ‘triclinic’ nature. The compound containing 70% of V shows a ‘cubic’ structure, as indicated by JCPDS file 75-0317 and has been identified as FeV2O4. The different peaks in the XRD pattern for all the materials have been indexed with the corresponding (hkl) values compared with the standard JCPDS data file.


image file: c5ra22586h-f5.tif
Fig. 5 XRD patterns of Fe–V-oxide compound semiconductor films annealed at 500 °C for 3 h on an ITO conducting glass.

The crystallize size, D, (in Å), of the strongest peaks for each of the XRD patterns was analyzed using the Scherrer formula:

 
D = (0.9λ)/(β[thin space (1/6-em)]cos[thin space (1/6-em)]θ) (2)
where ‘λ’ is the wavelength of X-ray (1.54 Å), ‘β’ FWHM (full width at half maxima, in degrees), and ‘θ’ is the diffraction angle (in degrees). The value of d, which is the interplanar spacing between the atoms, is calculated using the Bragg's law.
 
= 2d[thin space (1/6-em)]sin[thin space (1/6-em)]θ (3)

The average sizes of the crystallites corresponding to the strongest peaks for each of the Fe2O3 (104), FeVO4 (311), FeV2O4 (311) and Fe2VO4 (112) were calculated as 21 nm, 23 nm, 23 nm and 26 nm, respectively.

3.2. Photocurrent measurements

The three-electrode cell setup with an area of 0.27 cm2 exposed SC surface as the working electrode (WE) was used to study the photoelectrochemical (PEC) behavior of the compounds under periodic chopped UV-visible illumination of 100 mW cm−2. Fig. 6a represents the linear sweep voltammograms for photo oxidation of the sacrificial reagent (SO32−) over the different SC surfaces at a scan rate of 10 mV s−1, within the potential range from −0.1 to 0.6 V. Fig. 6b inset represents the bar plot, i.e. variation of photocurrent as measured from LSV (Fig. 6a) at a potential of 0.5 V, for the compounds which indicate that FeV2O4 exhibits the highest photocurrent compared to others.
image file: c5ra22586h-f6.tif
Fig. 6 (a) Linear sweep voltammograms of Fe–V-oxide semiconductor electrodes in 0.1 M SO32−–SO42− solution in the presence of UV-vis light (intensity: 100 mW cm−2). (b, inset) Variation of photocurrent (IPh) at 0.5 V for the different materials.

The water oxidation behavior of the SC materials was evaluated through LSV plots. The plot in Fig. 7a was obtained in 0.1 M SO42− and 0.1 M phosphate buffer solution (pH 7) within the potential range from 0.0 to 1.2 V and the variation of photocurrent for the different materials at 1.2 V is presented in the bar plot (Fig. 7b, inset). Similar to that observed for sacrificial oxidation, in this case, the highest photocurrent is also observed for the FeV2O4 compound SC, which attains a value of 0.18 mA cm−2 at 1.2 V. The visible-light responsiveness of the semiconductors towards water oxidation was determined by holding a UV cut-off filter (λ > 420 nm) in front of the Xe lamp. Fig. S2 (ESI) represents the comparative behavior of the FeV2O4 film undergoing the water oxidation process, wherein the material retains ∼30% photo-performance when exposed to periodic UV-vis and visible illumination.


image file: c5ra22586h-f7.tif
Fig. 7 (a) Linear sweep voltammograms of Fe–V-oxide semiconductor electrodes in 0.1 M SO42− phosphate buffer solution (for water oxidation) in the presence of UV-vis light (intensity: 100 mW cm−2). (b, inset) Variation of photocurrent (IPh) at 1.2 V for the different materials.

The stabilities of the semiconductor electrodes were verified via chronoamperometry by holding a fixed potential of 1.0 V vs. Ag/AgCl for the photoelectrochemical oxidation of water under the constant illumination of 100 mW cm−2. Fig. 8 represents the chronoamperometric (current–time) diagram of all the semiconductors, wherein FeV2O4 indicates fairly stable water oxidation behavior for at least 30 minutes. The gradual decrease in photocurrent for much longer exposure may be ascribed to the steady recombination of the photo-generated electrons and holes within the bulk of the SC matrix.


image file: c5ra22586h-f8.tif
Fig. 8 Current–time response (chronoamperometry) curves of the Fe–V-oxide semiconductors at an applied potential of 1.0 V in 0.1 M Na2SO4 with a 0.1 M phosphate buffer solution (pH 7) under UV-vis light irradiation.

Electrochemical action spectra were obtained under periodic chopped monochromatic illumination. Fig. 9 represents the photoelectrochemical action spectra for all the semiconductors in 0.1 M Na2SO4 in the presence of PBS, which were measured at 1.0 V with a gradual change in monochromatic light wavelength from 250 to 620 nm. The corresponding photocurrent at each wavelength was calculated from the action spectra, which is presented in Fig. S3 (ESI). The onset of photocurrent for the different semiconductors (Fe2O3, Fe2VO4, FeVO4 and FeV2O4) was found to be 560, 580, 600 and 610 nm, which correspond to the ‘true’ or photoelectrochemical band gap of the materials of 2.21, 2.13, 2.06 and 2.03 eV, respectively.


image file: c5ra22586h-f9.tif
Fig. 9 Photoelectrochemical action spectra of the Fe–V-oxide semiconductors, which were calculated from the photocurrent spectra of the material.

IPCE measurement showed the clear wavelength dependence of the photocurrent generation. The incident photon to current conversion efficiency (IPCE) measurement was performed with monochromatic irradiation employing a 300 W Xe arc lamp attached with an optical power meter (Oriel). The IPCE (%) was calculated from the following equation:

 
IPCE (%) = (1240/λ) × (IPh/Pin) × 100 (4)
where λ is the wavelength of the incident light in nm, IPh is the photocurrent in mA cm−2 and Pin is the power of the incident beam in 100 mW cm−2 of monochromatic light. The applied potential was monitored against an Ag/AgCl reference, and the absorbed photon to current conversion efficiency (APCE) was further calculated by considering the absorbance (Aλ) of the material at that particular wavelength using the following equation:
 
APCE (%) = (IPCE/1 − 10Aλ) × 100 (5)

Fig. 10a and b represent the variation of % IPCE and % APCE, respectively, for the different materials, and it has been observed that the FeV2O4 semiconductor exhibits the highest photoconversion efficiency of IPCE as 22%, and corresponding APCE as 60%, compared to the other Fe–V-oxides.


image file: c5ra22586h-f10.tif
Fig. 10 Variation of (a) % IPCE and (b) % APCE of Fe–V-oxide semiconductors in 0.1 M Na2SO4–PBS (pH 7) at applied potential 1.0 V vs. Ag/AgCl reference electrode.

3.3. Capacitance measurement

The Mott–Schottky (M–S) plot, which is the linear variation of 1/C2 (space charge capacitance in F cm−2) with respect to the applied potential, can be expressed using the relation shown in eqn (6).
 
image file: c5ra22586h-t1.tif(6)
where εs is the dielectric constant of the semiconductor, ε0 is the permittivity of free space, e is the electronic charge in C, ND is the donor density (per cm3), E is the applied potential in V, Efb is the flat band potential in V, kB is the Boltzmann constant and T is the temperature in absolute scale. The flat band potential of the semiconductor–electrolyte interface can be obtained from the intercept on the potential axis, whereas the slope of the straight line is inversely related to the carrier concentration, ND image file: c5ra22586h-t2.tif The positive slope of the M–S plots, as shown in Fig. 11, confirms the n-type semiconductivity of all the materials and the estimated flat band potentials for Fe2O3, Fe2VO4, FeVO4 and FeV2O4 are obtained from the intercept of the tangent line of the M–S plots on the potential axis as −0.15, −0.10, 0.08 and 0.07 V, respectively. The positive shifting of the flat band potentials with the gradual addition of % V to the matrix may favor the overall charge transfer kinetics across the semiconductor–electrolyte interface. The donor densities of the Fe–V-oxide compound semiconductors of different compositions were calculated from the slope of the respective M–S plots using the respective dielectric constants (εs) of the materials to be Fe2O3: 14.2;28 FeVO4: 14.78 (ref. 29) and FeV2O4: 7.0,30 and the value for Fe2VO4 is considered to be the same as that for FeVO4 due to the unavailability of data. The order of the slopes of the M–S plots is found to be quite similar to the values reported for similar Fe-oxide based semiconductor materials.18 A relatively higher carrier concentration of FeV2O4 matrix (2.7 × 1020 cm−3) may lead to a reduction in the ohmic resistivity of the film, which thereby improves the photoelectrochemical water oxidation performance, as observed in the LSV plots. M–S plots measured at 200, 500 and 1000 Hz using FeV2O4 are presented in Fig. S4 (ESI), which indicate a frequency independent behavior from the analysis.

image file: c5ra22586h-f11.tif
Fig. 11 Capacitance–voltage profile for Fe–V-oxide compounds recorded at 1000 Hz with an ac amplitude of 10 mV at each potential.

The energy band diagram for the different Fe–V-oxide compounds is presented in Fig. 12, which shows the conduction band (CB) position near their flat band positions and the corresponding valence band (VB), considering their individual band gap energies. The redox potential of the photo-generated holes (+1.8–1.9 V vs. Ag/AgCl) is high enough to drive the water oxidation reaction (H2O → O2, E0 +0.59 V vs. Ag/AgCl at pH 7), and the material exhibits excellent photoelectrochemical oxidation behavior.


image file: c5ra22586h-f12.tif
Fig. 12 Schematic band diagrams of Fe–V-oxides semiconductor–electrolyte interface.

3.4. Impedance measurements

Frequency dispersive impedance data were collected for all the materials and a typical Nyquist plot showing the variation of real and imaginary part of the impedance, under UV-vis illumination, is presented in Fig. 13a. The Nyquist plots were analyzed based on the equivalent circuit model (Fig. 13b), which includes a solution resistance (Rs), charge-transfer resistance (Rct) and double-layer capacitance (Cct) associated with the charge transfer process. In the present case, the charge-transfer resistance (Rct) values of the different materials are found to vary for Fe2O3, FeVO4, Fe2VO4 and FeV2O4 as 102, 45, 25 and 23 kohm, respectively. In fact, a material with a low Rct and high Cct value is generally considered suitable for use in PEC cells as a photo-electrode due to its efficiency in the transfer of charge carriers across the semiconductor–electrolyte interface and better semiconductivity of the film matrix.
image file: c5ra22586h-f13.tif
Fig. 13 (a) Nyquist plot of Fe–V-oxide semiconductor films in 0.1 M Na2SO4 with PBS aqueous solution under a UV-vis light source. (b, inset) Equivalent circuit diagram to analyze the Nyquist plot.

In the present case, the minimum value of Rct and relatively higher Cct for the FeV2O4 compound semiconductor support the superior photoelectrochemical oxidation behavior of this material. Fig. S5 (ESI) represents the impedance data of FeV2O4 in dark, visible and UV-visible conditions, which indicates the facile charge transfer reaction in the presence of UV-vis light.

4. Conclusion

Iron vanadium oxide semiconductor thin films of different compositions have been prepared on an ITO coated glass substrate at 500 °C. EDX analysis further confirms the presence of iron, vanadium and oxygen in the films, almost in the same order as that prepared through the drop-casting technique. The semiconductor electrodes are found to remain almost stable, at least for half an hour in aqueous electrolytes at pH 7 under light illumination of 100 mW cm−2. The Fe–V-oxides show a direct band gap at ∼1.9 eV and indirect gap at ∼2.6–2.7 eV, which result in the significant visible-light responsiveness of the materials for water oxidation reaction. The FeV2O4 semiconductor, which has a distinct surface morphology, exhibits the highest photocurrent due to its high carrier concentration, low charge transfer resistance and higher double layer capacitance. The photoconversion efficiency of 22% IPCE and 60% APCE has been recorded with the FeV2O4 semiconductor material toward the O2 evolution reaction from water.

Acknowledgements

Financial support through the sponsored project grant of the Govt. of India from SERB-DST (Sanction Order No. SB/S1/PC-042/2013, dt. 28-05-2014), the BRNS-DAE (Sanction Order No. 2013/37C/61/BRNS, dt. 03-12-2013) and the DST, Govt. of West Bengal, (Sanction order no. 902(Sanc.)/ST/P/S & T/4G – 1/2013, dt. 08/01/2015) to the Department of Chemistry, IIEST, Shibpur is gratefully acknowledged. Instrumental support from the MHRD, UGC-SAP to the Department of Chemistry, IIEST, Shibpur is also gratefully acknowledged.

References

  1. V. Nivoix and B. Gillot, Chem. Mater., 2000, 12, 2971–2976 CrossRef CAS.
  2. H. Mandal, S. Shyamal, P. Hajra, B. Samanta, P. Fageria, S. Pande and C. Bhattacharya, Electrochim. Acta, 2014, 141, 294–301 CrossRef CAS.
  3. K. Sivula, F. L. Formal and M. Gratzel, Chem. Mater., 2009, 21, 2862–2867 CrossRef CAS.
  4. A. B. Murphy, Sol. Energy Mater. Sol. Cells, 2007, 91, 1326–1337 CrossRef CAS.
  5. S. K. Biswas and J. O. Baeg, Int. J. Hydrogen Energy, 2013, 38, 14451–14457 CrossRef CAS.
  6. A. Dixit, P. Chen, G. Lawes and J. L. Musfeld, Appl. Phys. Lett., 2011, 99, 141908 CrossRef.
  7. X. Zhou, G. Wu, G. Gao, J. Wang, H. Yang, J. Wu, J. Shen, B. Zhou and Z. Zhang, J. Phys. Chem. C, 2012, 116, 21685–21692 CAS.
  8. K. Maeda and K. Domen, J. Phys. Chem. C, 2007, 111, 7851–7861 CAS.
  9. X. Wang, K. R. C. Heier, L. K. Stern and R. Poeppelmeier, Inorg. Chem., 1998, 37, 6921–6927 CrossRef CAS PubMed.
  10. A. Kudo, Int. J. Hydrogen Energy, 2006, 31, 197–202 CrossRef CAS.
  11. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS PubMed.
  12. O. Khaselev and J. A. Turner, Science, 1998, 280, 425–427 CrossRef CAS PubMed.
  13. S. Shyamal, P. Hajra, H. Mandal, J. K. Singh, A. K. Satpati, S. Pande and C. Bhattacharya, ACS Appl. Mater. Interfaces, 2015, 7, 18344–18352 CAS.
  14. L. Vayssieres, N. Beermann, S. E. Lindquist and A. Hagfeldt, Chem. Mater., 2001, 13, 233–235 CrossRef CAS.
  15. X. Wang, D. A. V. Griend, C. L. Stern and K. R. Poeppelmeier, Inorg. Chem., 2000, 39, 136–140 CrossRef CAS PubMed.
  16. C. Bhattacharya, H. C. Lee and A. J. Bard, J. Phys. Chem. C, 2013, 117, 9633–9640 CAS.
  17. H. Xu, H. Li, L. Xu, C. Wu, G. Sun, Y. Xu and J. Chu, Ind. Eng. Chem. Res., 2009, 48, 10771–10778 CrossRef CAS.
  18. X. Zhang, H. Li, S. Wang, F. F. Fan and A. J. Bard, J. Phys. Chem. C, 2014, 118, 16842–16850 CAS.
  19. J. S. Jang, J. Lee, H. Ye, F. F. Fan and A. J. Bard, J. Phys. Chem. C, 2009, 113, 6719–6724 CAS.
  20. K. G. Schrantz, S. Tymen, F. Boudoire, R. Toth, D. K. Bora, W. Calvet, M. Gratzel, E. Constable and A. Braun, Phys. Chem. Chem. Phys., 2013, 15, 1443–1451 RSC.
  21. Y. Lin, S. Zhou, S. W. Sheehan and D. Wang, J. Am. Chem. Soc., 2011, 133, 2398–2401 CrossRef CAS PubMed.
  22. C. D. Morton, I. J. Slipper, M. J. K. Thomas and B. D. Alexander, J. Photochem. Photobiol., A, 2010, 216, 209–214 CrossRef CAS.
  23. M. A. Mahadik, P. S. Shinde, M. Cho and J. S. Jang, J. Mater. Chem. A, 2015, 3, 23597–23606 CAS.
  24. J. W. Tang, Z. G. Zou and J. H. Ye, J. Phys. Chem. B, 2003, 107, 14265–14269 CrossRef CAS.
  25. J. Zhan, X. Yang, D. Wang, S. Li, Y. Xie, Y. Xia and Y. T. Qian, Adv. Mater., 2000, 12, 1348–1351 CrossRef CAS.
  26. K. Fujihara, S. Izumi, T. Ohno and M. Matsumura, J. Photochem. Photobiol., A, 2000, 132, 99–104 CrossRef CAS.
  27. P. Hazra, A. Jana, M. Hazra and J. Datta, RSC Adv., 2014, 4, 33662–33671 RSC.
  28. S. S. Kulkarni and C. D. Lokhande, Mater. Chem. Phys., 2003, 82, 151–156 CrossRef CAS.
  29. L. Zhao, M. P. Y. Wu, K. W. Yeh and M. K. Wu, J. Solid State Chem., 2011, 151, 1728–1732 CrossRef CAS.
  30. H. Takei, T. Suzuki and T. Katsufuji, Appl. Phys. Lett., 2007, 91, 072506 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra22586h

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