Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

The synergistic effect of cation mixing in mesoporous BixFe1−xVO4 heterojunction photoanodes for solar water splitting

Mengyuan Zhang a, Huu Khue Pham a, Yanan Fang a, Ying Fan Tay a, Fatwa F. Abdi *b and Lydia H. Wong *a
aSchool of Materials Science and Engineering, Nanyang Technological University, Nanyang Avenue, Singapore 639798. E-mail: lydiawong@ntu.edu.sg
bInstitute for Solar Fuels, Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, Hahn-Meitner-Platz 1, Berlin 14109, Germany. E-mail: fatwa.abdi@helmholtz-berlin.de

Received 17th February 2019 , Accepted 16th May 2019

First published on 17th May 2019


Abstract

Mixed metal vanadates are an interesting class of materials due to their favorable bandgap for visible light absorption and their catalytic activity. Here, we report a novel BixFe1−xVO4 mixture system fabricated by electrospinning, which upon annealing is composed of triclinic FeVO4 and monoclinic BiVO4 phases. The mixture demonstrates extended optical absorption and a clear bandgap shift as compared with a pure BiVO4. This is also accompanied by an extended wavelength range for its photoactivity as evident from the incident photon-to-current efficiency. In addition, the mixture with a Bi/(Bi + Fe) ratio of 0.5 (i.e., x = 0.5) shows superior charge transfer and charge separation efficiency. The improved charge transfer efficiency is attributed to the higher catalytic activity of the mixed cation, while the presence of a BiVO4/FeVO4 heterojunction enhances the charge separation. The formation of the heterojunction is verified through detailed microscopic investigations revealing BiVO4 particles intimately surrounded by FeVO4. Our results demonstrate the advantage of establishing a mixture of complex metal oxides in extending optical absorption and boosting the photoelectrochemical performance.


Introduction

In the field of photoelectrochemical (PEC) water splitting, bismuth vanadate (BiVO4) has emerged as one of the highest performing metal oxide photoanodes. With recent progress and success of achieving 90% of its theoretical maximum photocurrent, it is clear that further improvement in the solar-to-hydrogen (STH) efficiency of bismuth vanadate (BiVO4) is limited by its bandgap (2.4–2.5 eV).1,2 Numerous attempts have been reported on either identifying novel narrow bandgap photoanode materials or decreasing the bandgap of BiVO4. Some emerging materials, such as CuWO4 (2.1 eV),3 SnWO4 (2.1 eV),4,5 ZnFe2O4 (2.0 eV),6,7 CuxVyOz (1.8–2.0 eV),8,9 and FeVO4 (2.06 eV),10,11 have been reported to be promising photoanode candidates. However, these novel complex metal oxides still show much lower performance than expected, and further developments on a relatively longer timescale are necessary. Sulfurization or nitridation has been demonstrated to effectively reduce the bandgap of BiVO4, but its PEC stability may be a concern.2,12 As an alternative, the bandgap of BiVO4 can also be modified by incorporating another metal cation to form a metal vanadate mixture. The combination of a new metal cation—if chosen properly—could possibly raise the valence band of BiVO4 without much effect on the conduction band or form a favorable heterojunction with extended light absorption. The effective bandgap, as a result, will be reduced.

Studies on mixed metal vanadates have been reported from a decade ago for photocatalytic applications (e.g., BiCu2VO6 and BiZn2VO6).13–18 Combining two metal vanadates (Cu and Bi) into a solid solution to lower the bandgap of BiVO4 has been experimentally demonstrated.13,15 The absorption of BiCu2VO6 extends to 600 nm, which is 80 nm longer than that of BiVO4. It was proposed that the bandgap reduction is due to band modulation through the incorporation of two metal cations. The additional advantage reported is that BiCu2VO6 and BiZn2VO6 possess higher photocatalytic activity than pristine BiVO4.15,17 Similarly, a Bi0.75Fe0.25VO4 mixture demonstrated higher photocatalytic efficiency in the degradation of methylene blue than FeVO4 or BiVO4.19 However, the detailed mechanism on why mixed vanadates can offer higher catalytic activity or improved photoelectrochemical performance is not well understood. In addition, a high performance mixed vanadate photoanode has not been reported in the literature.

In our previous report, a series of BixFe1−xVO4 (0 ≤ x ≤ 1) films were fabricated by spray pyrolysis.20 Initial results indicated the possibility of maintaining good mobility of BiVO4 and extending visible light absorption by mixing FeVO4 with BiVO4. The present study adopts electrospinning to deposit mixed vanadate films and higher annealing temperature to improve the crystallinity. The combination of Fe and Bi vanadate shows higher catalytic efficiency and increased visible light absorption; in particular, the Fe0.5Bi0.5VO4 composition demonstrates the highest PEC performance. Microscopic observations show that the Fe0.5Bi0.5VO4 consists of adjacent FeVO4 and BiVO4 nanoparticles, which results in a favorable heterojunction to facilitate charge separation. Our study presents the synergistic effect of a Bi/Fe vanadate heterojunction and gives insights to further improve the performance of vanadate-based photoanodes.

Results and discussion

Physical characterization and chemical compositions

We prepared a series of BixFe1−xVO4 (0 ≤ x ≤ 1) films by electrospinning (Scheme S1) (the expression BixFe1−xVO4 here is determined by the composition of precursors and does not necessarily indicate the formation of a solid solution). For brevity, the different films are denoted by the percentage amount of Bi in each film, i.e., Bi0 corresponds to a FeVO4 film, Bi50 corresponds to Bi0.5Fe0.5VO4, and Bi100 corresponds to BiVO4. Prior to other measurements, EDX spectroscopy was used to confirm that the Fe and Bi ratio follows the initial amount added to the precursor (Fig. 1a). EDX mapping confirms the uniform distribution of Fe and Bi throughout the film (Fig. S1). Fig. 1b shows the X-ray diffraction (XRD) patterns of the various BixFe1−xVO4 films. At each extreme of the composition (i.e., Bi0 and Bi100), a pure phase of triclinic FeVO4 and monoclinic BiVO4 was obtained, respectively. For the mixed vanadate, both FeVO4 and BiVO4 peaks were detected from Bi5 to Bi50. From Bi65 and above, the FeVO4 phase can no longer be detected by XRD, but only the BiVO4 peak intensity changes with increasing Bi content (Fig. S2). This may be caused by the relatively small amount of FeVO4 phase in the films. Further investigations on the crystal structure by transmission electron microscopy are discussed later in the paper. Compared with the peak intensity of BiVO4, the FeVO4 peak intensity is relatively lower. This is due to the fact that the films were annealed at 550 °C while annealing at 650 °C is typically required to obtain good crystallinity of FeVO4.10,20 This lower temperature was chosen as a compromise because V loss occurs at temperatures above 450 °C.21 Fig. S3 indeed shows that all of our films are slightly V-poor. The effect of V loss has been discussed elsewhere and is not the focus of this report.21
image file: c9ta01791g-f1.tif
Fig. 1 (a) Bi ratio in the annealed films calculated from the EDX results compared with the Bi ratio in the precursor. (b) XRD patterns of all compositions annealed at 550 °C in air.

The chemical composition and oxidation states in the films were analyzed by XPS. All the spectra were calibrated using the adventitious carbon peak at 284.8 eV. Fe, Bi, V and O core level XPS spectra of FeVO4, Bi50 and BiVO4 films are shown in Fig. S4. Peaks at 710–714 eV can be observed in FeVO4 and Bi50 films, which can be assigned to Fe 2p3/2 multiplets with an oxidation state of +3.22,23 The peak at 716 eV, which is present for all films, corresponds to the Sn 3p3/2 core level, which is caused by the exposed FTO.24 XPS peaks at 158.7 eV and 164.1 eV confirm the existence of Bi3+ in Bi50 and BiVO4.25,26 The vanadium is present as V5+ in all films, as evident from the V 2p3/2 peak at ∼516.9 eV.27,28 The binding energy of O 1s at 530 eV matches with the oxygen lattice peak in the oxide.27 The peak positions do not shift for all films, suggesting that the chemical environment of the mixture remains unchanged.

A gradual color change from orange (FeVO4) to bright yellow (BiVO4) can be observed in the visual appearance of the films (Fig. S5), similar to our previous report.20 The porous appearance of the films is due to the addition of PVP to the electrospinning precursor. The absorption spectra of the mixed vanadate films obtained from UV-vis spectroscopy are shown in Fig. 2a and b. Other than the pure BiVO4 film, the absorption edge of the mixed vanadate films extends beyond 520 nm (absorption edge of BiVO4), validating the effect of Fe incorporation on increasing the visible light absorption range. For Bi10 to Bi50 films, a higher Bi ratio increases the absorption at around 480 nm, which is a characteristic absorption shoulder of BiVO4 (Fig. 2a). Beyond this Bi concentration (i.e., 50% and above), visible light absorption at the 520–600 nm range decreases with higher Bi content (Fig. 2b), possibly due to the difference of absorption coefficients of BiVO4 and FeVO4.20,29 Although the films were detected to be mixtures of BiVO4 and FeVO4, the ‘apparent bandgap’ of the mixture, instead of the actual bandgap, can still be determined. Thus, the direct and indirect bandgaps of all the films can be calculated using Tauc analysis; the results are shown in Fig. 2c. The direct bandgap only shows minor differences for all compositions (2.45–2.5 eV), but the indirect bandgap increases with increasing Bi content.


image file: c9ta01791g-f2.tif
Fig. 2 (a) and (b) Absorption spectra and (c) calculated direct and indirect bandgaps for all films with various Bi ratios.

Our results clearly show that a higher Fe content leads to a larger visible absorption range. However, the appearance of the absorption shoulder at 480 nm and the varying absorption spectra at ∼520–600 nm suggest that an optimal Bi/Fe ratio may exist in order to achieve the maximum total absorption. By integrating all the absorption spectra with the AM1.5 photons, the theoretical absorbed photocurrent density (Jabs) can be readily calculated for each composition and the results are shown in Fig. S6. All the films have a Jabs of ∼3–4 mA cm−2. With these absorption changes brought about by the Bi/Fe ratio, a maximum is observed for the Bi10 film (4 mA cm−2).

Improved charge carrier efficiencies

The AM1.5 photocurrent density of all films was measured in 0.1 M phosphate buffer (KPi, pH = 7). Most films show higher photocurrent under back-side illumination as compared to front illumination (Fig. S7). 30 min deposition time (electrospinning) is found to be optimal to ensure sufficient film thickness and avoid delamination of films during annealing. As the PEC performance is similar for 30 min and 45 min deposited films (Fig. S8), only films electrospun for 30 minutes will be discussed below (the original photocurrent curves are shown in Fig. S9). As shown in Fig. 3a, the photocurrent of pristine FeVO4 and BiVO4 is quite comparable to the previously reported sprayed FeVO4 and electrospun BiVO4.20,27 The comparison of the photocurrents for the electrospun and spray pyrolysed samples are shown in Table S1. Among all the compositions, the Bi50 film shows the highest photocurrent (0.47 mA cm−2 at 1.23 V vs. RHE). The peak photocurrent trend at a Bi 50% ratio is even more obvious at higher potential, while other Bi rich or Fe rich composition films only generate moderate photocurrents. From Bi0 to Bi50, the photocurrent increases almost linearly with increasing Bi content; for Bi50 to Bi100, the photocurrent instead decreases. To exclude any surface limitation (due to slow charge transfer or surface recombination), the photocurrent was also measured in KPi with added H2O2 as a hole scavenger. The hole scavenger photocurrent trend is similar to that in KPi (Fig. 3b), with a Bi50 film showing the highest photocurrent of up to 1.82 mA cm−2 at 1.23 V vs. RHE.
image file: c9ta01791g-f3.tif
Fig. 3 (a) Photocurrent as a function of Bi ratio for all compositions, measured in 0.1 M KPi buffer (pH = 7) (a) and in 0.5 M H2O2 + 0.1 M KPi (b) under back-side illumination. (c) Charge transfer efficiency (ηtrans) and separation efficiency (ηsep) at 1.23 V vs. RHE as a function of Bi ratio. (d) Onset potential taken from chopped current–voltage curves in KPi buffer.

The charge transfer and separation efficiencies can be calculated from the photocurrent density with and without a hole scavenger (definition and calculation steps are described in the ESI). The transfer efficiency (ηtrans) of all compositions is shown in Fig. S10 and the overall trend at 1.23 V vs. RHE is shown in Fig. 3c. It is interesting to note that the charge transfer efficiencies of all Bi/Fe mixed vanadate films are more or less constant and higher than those of pure FeVO4 and BiVO4 over the whole potential range. This is in good agreement with the observation that mixed vanadates have higher catalytic activity than a single metal vanadate.15,17,19 In addition, a closer look at the trend shows that the ηtrans of the films with higher Fe content (Bi5 to Bi50) is slightly higher than those of the Bi rich films (Bi65 to Bi100). We attribute this to the higher catalytic activity of Fe sites for oxygen evolution as compared to Bi. Indeed, the dark water oxidation current of FeVO4 and Bi50 films is higher than that of BiVO4 (Fig. S11).

Since the Jabs values of all films are relatively close to each other (Fig. S6), the charge carrier separation efficiency (ηsep) (Fig. 3c and S12) follows a similar trend to the hole scavenger photocurrent. The incorporation of Bi/Fe at around a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio (i.e., Bi50) results in an optimal charge separation. The ηsep of Bi50 at 1.23 V vs. RHE is 10-fold and 1.7-fold higher as compared to that of FeVO4 and BiVO4, respectively. It is therefore clear that the photocurrent improvement in the mixed Bi–Fe-vanadate films can be mainly attributed to the more efficient charge carrier separation.

The much higher ηsep can be explained by the formation of FeVO4 and BiVO4 heterojunctions with favorable band alignment to facilitate efficient charge carrier separation.23 As the two vanadates are homogeneously distributed, indeed the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio is expected to be optimal to ensure enhanced charge separation at all parts of the film. A BiVO4/FeVO4 heterojunction in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio has also been reported to obtain the highest photocatalytic activity for pollutant degradation,30 which agrees very well with our results and interpretation.

Electrical impedance measurements were performed to shed light on the charge transfer processes. The Nyquist plots for all compositions derived from EIS measurements under illumination are shown in Fig. S13. The different radii of semicircles can reflect the overall resistance. It is clear that Bi50 has the smallest semicircle; either increasing or decreasing the Fe content only results in a larger radius of the semicircle. An equivalent circuit is shown in Fig. S14, which is typically used for the reported heterojunction structure and porous electrodes (the electrospun films show some extent of porosity as evidenced by SEM images). Here, Rs is the serial resistance, Csc is the space charge layer capacitance, Rct is the charge transfer resistance and CPE is the constant phase element.31–33 From FeVO4 to Bi50, Rct reduces with increasing Bi ratio; however Rct increases again from Bi50 to BiVO4. Such a trend similar to the changes of the semicircle radius in the Nyquist plot implies that Bi50 has the lowest charge transfer resistance, which can be properly explained with the presence of a BiVO4/FeVO4 heterojunction.

The heterojunction explanation is also supported by the trend in the onset potential as a function of Bi ratio. The addition of BiVO4 to FeVO4 leads to a dramatic onset potential drop from 0.94 V vs. RHE (Bi0, FeVO4) to ∼0.4 V vs. RHE (Bi80–Bi95) (Fig. 3d). The cathodic onset potential shift is a typical effect of the heterojunction, which has been reported in TiO2, Fe2O3, and BiVO4 photoanodes.34–38 While elemental doping or surface passivation can also induce a cathodic onset potential shift, we have ruled out the presence of Bi or Fe doping in our mixed vanadates as no obvious peak shift was observed (see Fig. S2) and the mixed vanadates show similar surface charge transfer efficiency (Fig. S10). Further observations by SEM show that all films are composed of nanoparticles with similar morphologies (Fig. S15). Capacitive current measurements show no apparent trend of the electrochemical active surface area of all films (Fig. S16). Cross-sectional view images also confirm similar particle sizes across the depth profile (∼150–170 nm) for different compositions (Fig. S17).

To further confirm the formation of a heterojunction, TEM and high resolution TEM (HRTEM) images of the Bi50 sample were taken as shown in Fig. 4. A clear phase contrast can be observed between the bright and dark regions (Fig. 4c and d), where the dark regions are surrounded by the bright regions. Clear lattice fringes can be observed in the bright regions, with d-spacings of 0.85 nm and 0.34 nm corresponding to the (001) and (021) planes of FeVO4, respectively (Fig. 4a and b). The dark regions, however, do not show any clear fringes. To confirm the composition in the dark regions, Energy Dispersive X-ray Analysis (EDAX) was carried out at the sample locations shown in Fig. 4c and d, and the elemental mapping is shown in Fig. S18. The dark regions are verified to be BiVO4 particles while FeVO4 particles (bright regions) are more dispersed. From the mapping images, the microscopic structure of BiVO4 particles intimately surrounded by FeVO4 particles can be proposed. This structure also explains why only the FeVO4 phase can be clearly seen in the HRTEM images (bright regions), while the dark regions do not show lattice fringes due to the convolution between the BiVO4 and FeVO4 structure. This is further supported by the higher surface Fe/Bi ratio of the Bi50 film (2.2 as revealed from XPS, see Table S2) as compared to the bulk Fe/Bi ratio of 1 (see Fig. 1a). Such an intricate structure shows clear evidence for the heterojunction, which enhances the charge carrier separation efficiency in the mixed vanadate films, particularly in the case of Bi50. A core–shell BiVO4/FeVO4 structure has indeed been reported to enhance charge separation.23


image file: c9ta01791g-f4.tif
Fig. 4 (a) and (b) HRTEM images of the Bi50 sample at two different places, showing d-spacings of the FeVO4 (001) and (021) planes to be 0.85 nm and 0.34 nm, respectively. (c) and (d) TEM images of the Bi50 sample at two different locations. Note the different scale-bars between the two images.

image file: c9ta01791g-f5.tif
Fig. 5 UPS spectra of (a) BiVO4 and (b) FeVO4 with a bias of 10 eV. Secondary electron cut-off and valence band maximum positions have been indicated.

image file: c9ta01791g-f6.tif
Fig. 6 IPCE of FeVO4, Bi50 and BiVO4 measured in (a) 0.1 M KPi and (b) 0.1 M KPi + 0.5 M H2O2 at 1.5 V vs. RHE. Both insets show the partially zoomed graphs to visualize the onset wavelength.

Mott–Schottky measurements were carried out to obtain the flat-band position of our films, which can be further used to estimate the Fermi level position. The flat-band potential of BiVO4 and FeVO4 is extrapolated to be ∼0.13 V and ∼0.72 V vs. RHE, respectively (Fig. S19), which agree very well with reported values.20,39 For a conclusive confirmation of band alignments, UPS was conducted and the spectra are presented in Fig. 5. The work function can be determined using the secondary electron cut-off. The position of the valence band maximum (EV) with respect to the Fermi level (EF) can be determined from the low kinetic energy cut-off of the spectrum.39 As shown in Fig. 5, the work function for BiVO4 and FeVO4 can be determined to be 4.5 eV and 4.9 eV, respectively, while the EV with respect to the EF is determined to be 2.36 eV and 2.06 eV, respectively. Although the measured work function of BiVO4 and FeVO4 is slightly higher than the reported values (BiVO4: 4.56 eV39 and FeVO4: 5.1 eV),20 the difference of the work function position is definitely clear. The band alignments of BiVO4 and FeVO4, therefore, are depicted in Scheme 1. The conduction band position is calculated from the optical bandgap and valence band position. Scheme 1 clearly shows that BiVO4 and FeVO4 can form a heterojunction structure. Photogenerated holes can thus be more efficiently extracted for water oxidation due to larger band bending in the heterojunction, which is in agreement with the higher ηsep for Bi50 as compared with BiVO4 and FeVO4.


image file: c9ta01791g-s1.tif
Scheme 1 Schematic diagram of band alignments of BiVO4 and FeVO4, confirming the heterojunction alignment. Note that the band bending at the interfaces is not depicted for simplicity.

We also note that the optimal charge carrier separation efficiency in the Bi50 film can be caused by the improved charge carrier transport property (i.e., carrier mobility). This is, however, unlikely to be the case. While we have no direct evidence of the carrier mobility in our electrospun films—time-resolved microwave conductivity (TRMC) measurements of our films are not possible since electrospinning the films on quartz (which is required for TRMC) results in mechanically unstable films—our previous study on sprayed mixed Bi–Fe-vanadates have shown that the carrier mobility continuously decreases with decreasing Bi content.20

Influence of the vanadate mixture on visible light photoactivity

Finally, we investigated the photoactivity of the films under different illumination wavelengths. The incident photon-to-current conversion efficiency (IPCE) curves of all films are shown in Fig. S20. Although the absorption edge of FeVO4 is ∼600 nm, which is considerably longer than that of BiVO4 (520 nm), the IPCE of FeVO4 at wavelengths above 525 nm is negligibly small (∼0) (Fig. 6a). The mismatch of the optical absorption onset and photoelectrochemical onset wavelength of FeVO4 can be attributed to the low carrier mobility and indirect absorption energy loss in a long wavelength range.20,40 Similar to the observation in UV-vis spectra, the IPCE at all wavelengths was enhanced for the mixed vanadates as compared with both FeVO4 and BiVO4 (Fig. S20), with Bi50 showing the highest IPCE (Fig. 6a). Further IPCE measurements in H2O2 were carried out to better illustrate the onset wavelength for photoactivity (Fig. 6b). The Bi50 film clearly exhibits an extended PEC response to its absorption edge at 565 nm (corresponds to its indirect bandgap of 2.2 eV in Fig. 2c), while FeVO4 and BiVO4 displayed the same onset wavelength as in KPi (∼520 nm, Fig. 6b). More direct evidence of the photo response of Bi50 at 550 nm is shown in Fig. S21. It is clear that Bi50 generated measurable photocurrent under monochromatic illumination at 550 nm.

Since FeVO4 and BiVO4 show zero IPCE signal at wavelengths higher than 525 nm, the extended IPCE response for Bi50 can be attributed to the absorption in the FeVO4 phase assisted by the presence of the BiVO4/FeVO4 heterojunction to improve charge separation as mentioned above. The equivalence of the optical onset and PEC onset wavelength for the Bi50 film therefore serves as evidence for a successful extension of the photoactivity in the mixed BiVO4/FeVO4 films.

Finally, Bi50 was measured in KPi at 1.5 V vs. RHE under illumination to investigate its stability. The photocurrent of the film continuously decreases in the 4000 s chronoamperometry measurement (Fig. S22a). Therefore, a thin layer of CoPi was deposited by a well-established method for surface protection and the thickness of CoPi is estimated to be ∼2 nm according to a report.41 After CoPi deposition, other than the initial transient, the photocurrent stability is improved. The film is able to retain a photocurrent of ∼0.6 mA cm−2 up to 4000 s (Fig. S22a), implying that a thin layer CoPi can help protect the surface and improve the stability. This CoPi deposited Bi50 film was later used for oxygen evolution measurements due to its improved stability. The amount of oxygen generated was quantitatively recorded by gas chromatography. The faradaic efficiency for oxygen evolution was calculated to be ∼88% (Fig. S23).

Conclusions

A series of BixFe1−xVO4 mixture mesoporous films was successfully fabricated by electrospinning. Instead of forming a solid solution, a mixed phase of BiVO4 and FeVO4 was detected from the X-ray diffraction patterns, and the indirect bandgap decreases with decreasing Bi/Fe atomic ratio. The mixed phase with an atomic ratio of Bi/Fe = 1, i.e., Bi50, was found to have the optimum charge carrier transfer and separation efficiencies as compared with BiVO4 and FeVO4. Improved charge transfer efficiency was achieved due to a higher catalytic activity of the Fe and Bi cation mixture, while the enhanced charge separation was attributed to the homogeneously distributed BiVO4/FeVO4 heterojunction as observed from TEM. In addition, IPCE measurements revealed that the Bi50 composition demonstrates photoactivity up to ∼560 nm, which is larger than that of pure BiVO4 and consistent with the indirect bandgap of ∼2.2 eV. Our work illustrates a simple yet successful strategy of forming a mixed phase heterojunction to expand light utilization as well as charge carrier transfer and separation efficiencies. Further improvement of the mixed phase films may be obtained through strategies well-reported in the literature, such as doping and nanostructuring.

Experimental

Preparation of BiFeVO4 films

All the films were fabricated by electrospinning in a similar procedure from our previous report.27 For one specific composition, BixFe1−xVO4 (0 ≤ x ≤ 1), 0.06x mmol bismuth nitrate pentahydrate (Bi(NO3)3·5H2O, 98%, Sigma Aldrich), 0.03(1 − x) mmol iron(III) acetylacetonate (Fe(acac)3, 99.9%, Sigma Aldrich) and 0.03 mmol vanadium acetylacetonate (VO(acac)2, 98%, Sigma Aldrich) were mixed so that the (Bi + Fe)/V ratio equals 1. The salt mixture was then dissolved in a mixed solvent containing 2 mL acetic acid (CH3COOH, 99.7%, Sigma Aldrich), 2.5 mL absolute ethanol (CH3CH2OH, 99.5%, Merck) and 2.5 mL N,N-dimethylformamide (HCON(CH3)2, 99.8%, Sigma Aldrich). After adding 1200 mg polyvinylpyrrolidone (PVP, (C6H9NO)n, Mw ∼ 1[thin space (1/6-em)]300[thin space (1/6-em)]000, Sigma Aldrich), the solution was then kept at room temperature with magnetic stirring overnight to obtain a viscous and homogeneous solution. The addition of PVP was done to obtain an appropriate solution viscosity for electrospinning.

The resulting solution was then used for electrospinning deposition. 4 mL prepared solution was sucked into a 16 cm diameter syringe. The syringe and the electrospinning needle (20G, 13 mm length) were connected with chemical resistant tubing (Tygon 2375, 1/16 × 3/16′′). The setup is shown in Scheme S1. Fluorine doped tin oxide (FTO) substrates (14 Ω cm−2, Nippon Sheet Glass) were subjected to a standard cleaning procedure (sonicated 20 min each in soap water, deionized water, acetone, deionized water and ethanol). After drying with nitrogen flow, the FTO substrates were placed at the center of the electrospinning stage (collection plate) partially wrapped with aluminum foil to allow contact between the FTO conductive side and the stage. During the electrospinning process, a voltage of 25 kV was applied between the needle and the collection plate at a distance of 12 cm. The deposition rate was controlled at 1.5 mL h−1 using the syringe pump. The deposition was carried out at room temperature at different time intervals, i.e., 15, 30 and 45 min. All the films were finally annealed at 550 °C in air for 3 h. The samples were denoted by the Bi content, for example Bi0.3Fe0.7VO4 can be denoted as Bi30.

Characterization

The morphology of the films was observed by field emission scanning electron microscopy (FESEM) using a JEOL-7600F microscope. Energy dispersive X-ray spectroscopy (EDX) analysis was done with an Oxford INCA detector. For each composition, measurements were done in at least five different locations on the films to obtain an average value. The crystal structure of the films was measured by X-ray diffraction (XRD) using a Shimadzu thin film X-ray diffractometer (Cu Kα source). Optical absorption spectroscopy was measured on a UV-vis-NIR spectrophotometer (PerkinElmer, Lambda 950). For structural characterization, Transmission Electron Microscopy (TEM) analysis was performed using a JEOL 2100F transmission electron microscope. Powders carefully scratched from the film deposited on an FTO substrate were used for the TEM analysis. The powders were loaded onto the copper grid after sonicating in ethanol for 20 min to allow better dispersion. Energy Dispersive X-ray Analysis (EDAX) was performed for elemental mapping.

Photoelectrochemical (PEC) measurements were carried out on a CHI 660D workstation (CH Instruments Inc.). The mixed vanadate photoanodes were irradiated using a 150 W xenon solar simulator (67005, Newport Corp.) with a solar filter (KG 3) to achieve a measured light intensity of 1 sun (100 mW cm−2). PEC measurements were performed in a PEC cell in a 3-electrode configuration, where the vanadate films were used as the anode, a Pt coil (3.6 cm2) as the counter electrode, and an Ag/AgCl electrode (KCl 3 M solution) as the reference electrode. The PEC performance of the photoanodes were measured in a potassium phosphate buffer electrolyte (0.1 M KPi, pH = 7). The illuminated area on the photoanodes is fixed at 0.125 cm2. Unless stated, all the PEC and IPCE measurements were performed under back-side illumination. Potential conversion to the reversible hydrogen electrode (RHE) scale was calculated using the equations below.

 
ERHE = EAg/AgCl + E0Ag/AgCl + 0.0591 × pH(1)
where E0Ag/AgCl = 0.21 V vs. NHE at 298 K.

For PEC measurements with a hole scavenger, 0.5 M hydrogen peroxide (H2O2, 30% w/w in H2O, Sigma Aldrich) was added to the KPi buffer solution. Incident photon-to-current conversion efficiency (IPCE) measurements were carried out using a combination of a xenon lamp (MAX-302, Asahi Spectra Co. Ltd.) and a monochromator (CMS-100, Asahi Spectra Co. Ltd.) from 350 to 620 nm wavelength. The IPCE values can be calculated using eqn (2).

 
image file: c9ta01791g-t1.tif(2)

Electrochemical active surface area (ECSA) measurements were performed using the same configuration as the PEC measurements, but without any irradiation (i.e., in the dark). The potential was swept between 0 V to 0.6 V vs. Ag/AgCl at different scan rates (20, 50, 100, 150, 200, and 300 mV s−1). The films were masked with tape to leave a fixed area (0.283 cm2) for electrical contact. More details of the ECSA measurements can be found in other reports.42,43 Electrical impedance spectroscopy (EIS) and Mott–Schottky measurements were conducted on an Autolab potentiostat (Metrohm-Autolab, AUT 83285) in a three electrode configuration using 0.1 M KPi solution. All the films were masked with tape to leave the same contact area (0.283 cm2).

X-ray photoelectron spectroscopy (XPS) and ultraviolet photoelectron spectroscopy (UPS) measurements were conducted on the same VG ESCALAB 2201-XL instrument (operation condition <5 × 10−10 mbar) equipped with a monochromatic Al Kα (1486.7 eV) X-ray source and a He I discharge lamp (21.2 eV) UV source.

A stability test was carried out in KPi at an applied potential of 1.5 V vs. RHE for both untreated and CoPi deposited Bi50 films. CoPi was deposited by electrodeposition at 1.5 V vs. RHE for 60 s using 0.1 M KPi electrolyte containing 0.5 mM Co(NO3)2. A two-compartment electrochemical cell divided by an ion exchange membrane was used for the gas evolution test. The CoPi-Bi50 film was subject to 1.5 V vs. RHE in 0.1 M KPi under AM1.5 illumination. The amount of O2 evolved was detected using a gas chromatograph (GC) (Agilent 7890A).

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Financial support by a Singapore Ministry of Education (MOE) Tier 2 grant (MOE2016T21030) is gratefully acknowledged. The authors thank Marlene Lamers for her assistance with the TRMC measurements.

References

  1. Y. Pihosh, I. Turkevych, K. Mawatari, J. Uemura, Y. Kazoe, S. Kosar, K. Makita, T. Sugaya, T. Matsui, D. Fujita, M. Tosa, M. Kondo and T. Kitamori, Sci. Rep., 2015, 5, 11141 CrossRef PubMed.
  2. F. A. Fatwa and P. B. Sean, J. Phys. D: Appl. Phys., 2017, 50, 193002 CrossRef.
  3. J. E. Yourey and B. M. Bartlett, J. Mater. Chem., 2011, 21, 7651–7660 RSC.
  4. I.-S. Cho, C. H. Kwak, D. W. Kim, S. Lee and K. S. Hong, J. Phys. Chem. C, 2009, 113, 10647–10653 CrossRef CAS.
  5. M. Kölbach, I. J. Pereira, K. Harbauer, P. Plate, K. Höflich, S. P. Berglund, D. Friedrich, R. van de Krol and F. F. Abdi, Chem. Mater., 2018, 30, 8322–8331 CrossRef.
  6. K. J. McDonald and K.-S. Choi, Chem. Mater., 2011, 23, 4863–4869 CrossRef CAS.
  7. N. Guijarro, P. Bornoz, M. Prévot, X. Yu, X. Zhu, M. Johnson, X. Jeanbourquin, F. Le Formal and K. Sivula, Sustainable Energy Fuels, 2018, 2, 103–117 RSC.
  8. L. Zhou, Q. Yan, A. Shinde, D. Guevarra, P. F. Newhouse, N. Becerra-Stasiewicz, S. M. Chatman, J. A. Haber, J. B. Neaton and J. M. Gregoire, Adv. Energy Mater., 2015, 5, 1500968 CrossRef.
  9. P. F. Newhouse, D. A. Boyd, A. Shinde, D. Guevarra, L. Zhou, E. Soedarmadji, G. Li, J. B. Neaton and J. M. Gregoire, J. Mater. Chem. A, 2016, 4, 7483–7494 RSC.
  10. C. D. Morton, I. J. Slipper, M. J. K. Thomas and B. D. Alexander, J. Photochem. Photobiol., A, 2010, 216, 209–214 CrossRef CAS.
  11. S. K. Biswas and J.-O. Baeg, Int. J. Hydrogen Energy, 2013, 38, 14451–14457 CrossRef CAS.
  12. M. Lamers, W. Li, M. Favaro, D. E. Starr, D. Friedrich, S. Lardhi, L. Cavallo, M. Harb, R. van de Krol, L. H. Wong and F. F. Abdi, Chem. Mater., 2018, 30, 8630–8638 CrossRef CAS.
  13. Y. Nakabayashi, M. Nishikawa and Y. Nosaka, J. Appl. Electrochem., 2016, 46, 9–16 CrossRef CAS.
  14. Z. Ding, Y. Fu, Z. Xie and Z. Li, Mater. Lett., 2011, 65, 460–463 CrossRef CAS.
  15. H. Liu, R. Nakamura and Y. Nakato, ChemPhysChem, 2005, 6, 2499–2502 CrossRef CAS PubMed.
  16. H. Liu, A. Imanishi, W. Yang and Y. Nakato, Electrochim. Acta, 2010, 55, 4130–4136 CrossRef CAS.
  17. H. Liu, R. Nakamura and Y. Nakato, Electrochem. Solid-State Lett., 2006, 9, G187–G190 CrossRef CAS.
  18. Y. Murakami, M. Ikarashi, M. Hashizume, A. Y. Nosaka and Y. Nosaka, Electrochem. Solid-State Lett., 2008, 11, H42–H46 CrossRef CAS.
  19. G. Bera, V. R. Reddy, P. Mal, P. Das and G. R. Turpu, AIP Conf. Proc., 2018, 1953, 080026 CrossRef.
  20. M. Zhang, Y. Ma, D. Friedrich, R. van de Krol, L. H. Wong and F. F. Abdi, J. Mater. Chem. A, 2018, 6, 548–555 RSC.
  21. M. Lamers, S. Fiechter, D. Friedrich, F. F. Abdi and R. van de Krol, J. Mater. Chem. A, 2018, 6, 18694–18700 RSC.
  22. A. P. Grosvenor, B. A. Kobe, M. C. Biesinger and N. S. McIntyre, Surf. Interface Anal., 2004, 36, 1564–1574 CrossRef CAS.
  23. M. Balamurugan, G. Yun, K.-S. Ahn and S. H. Kang, J. Phys. Chem. C, 2017, 121, 7625–7634 CrossRef.
  24. K. Lee, S. Shin, T. Degen, W. Lee and Y. S. Yoon, Nano Energy, 2017, 32, 397–407 CrossRef CAS.
  25. Z. Liu, Q. Lu, E. Guo and S. Liu, J. Nanopart. Res., 2016, 18, 236 CrossRef.
  26. C. D. Morton, PhD thesis, University of Greenwich, 2012.
  27. R. P. Antony, P. S. Bassi, F. F. Abdi, S. Y. Chiam, Y. Ren, J. Barber, J. S. C. Loo and L. H. Wong, Electrochim. Acta, 2016, 211, 173–182 CrossRef CAS.
  28. Y. Liu, Y. Guo, L. T. Schelhas, M. Li and J. W. Ager, J. Phys. Chem. C, 2016, 120, 23449–23457 CrossRef CAS.
  29. J. Yang, D. Wang, X. Zhou and C. Li, Chem.–Eur. J., 2013, 19, 1320–1326 CrossRef CAS PubMed.
  30. J. Li, W. Zhao, Y. Guo, Z. Wei, M. Han, H. He, S. Yang and C. Sun, Appl. Surf. Sci., 2015, 351, 270–279 CrossRef CAS.
  31. J. Bisquert, Phys. Chem. Chem. Phys., 2000, 2, 4185–4192 RSC.
  32. L. Zhang, E. Reisner and J. J. Baumberg, Energy Environ. Sci., 2014, 7, 1402–1408 RSC.
  33. X. Shi, I. Herraiz-Cardona, L. Bertoluzzi, P. Lopez-Varo, J. Bisquert, J. H. Park and S. Gimenez, Phys. Chem. Chem. Phys., 2016, 18, 9255–9261 RSC.
  34. H. Jung, S. Y. Chae, C. Shin, B. K. Min, O.-S. Joo and Y. J. Hwang, ACS Appl. Mater. Interfaces, 2015, 7, 5788–5796 CrossRef CAS PubMed.
  35. S. J. A. Moniz, J. Zhu and J. Tang, Adv. Energy Mater., 2014, 4, 1301590 CrossRef.
  36. S. Bai, H. Chu, X. Xiang, R. Luo, J. He and A. Chen, Chem. Eng. J., 2018, 350, 148–156 CrossRef CAS.
  37. S. Ho-Kimura, S. J. A. Moniz, A. D. Handoko and J. Tang, J. Mater. Chem. A, 2014, 2, 3948–3953 RSC.
  38. Q. Peng, J. Wang, Z. Feng, C. Du, Y. Wen, B. Shan and R. Chen, J. Phys. Chem. C, 2017, 121, 12991–12998 CrossRef CAS.
  39. J. Resasco, H. Zhang, N. Kornienko, N. Becknell, H. Lee, J. Guo, A. L. Briseno and P. Yang, ACS Cent. Sci., 2016, 2, 80–88 CrossRef CAS PubMed.
  40. J. Feng, Z. Wang, X. Zhao, G. Yang, B. Zhang, Z. Chen and Y. Huang, J. Phys. Chem. C, 2018, 122, 9773–9782 CrossRef CAS.
  41. C. Zachäus, F. F. Abdi, L. M. Peter and R. van de Krol, Chem. Sci., 2017, 8, 3712–3719 RSC.
  42. J. Kibsgaard and T. F. Jaramillo, Angew. Chem., Int. Ed., 2014, 53, 14433–14437 CrossRef CAS PubMed.
  43. B. K. Kang, G. S. Han, J. H. Baek, D. G. Lee, Y. H. Song, S. B. Kwon, I. S. Cho, H. S. Jung and D. H. Yoon, Adv. Mater. Interfaces, 2017, 4, 1700323 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ta01791g

This journal is © The Royal Society of Chemistry 2019
Click here to see how this site uses Cookies. View our privacy policy here.