Marlene
Lamers
,
Sebastian
Fiechter
,
Dennis
Friedrich
,
Fatwa F.
Abdi
* and
Roel
van de Krol
*
Institute 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; roel.vandekrol@helmholtz-berlin.de
First published on 17th September 2018
Metal oxide photoelectrodes typically suffer from poor carrier transport properties and extensive carrier recombination, which is caused by the presence of intrinsic or extrinsic defects in the material. Here, the influence of annealing temperature and atmosphere on the formation and suppression of defects in BiVO4—one of the best performing metal oxide photoanodes—is elucidated. Annealing in argon has little or no effect on the photoelectrochemical performance due to the competing effects of an increase in grain size (i.e., reduction of grain boundaries) and the unfavorable formation of oxygen vacancies. When annealing in air, the formation of oxygen vacancies is suppressed, resulting in up to ∼1.5-fold enhancement of the photocurrent and an order of magnitude increase of the charge carrier mobility. However, vanadium leaves the BiVO4 lattice above 500 °C, which leads to a decrease in carrier lifetime and photocurrent. This vanadium loss can be avoided by supplying excess vanadium in the gas phase during annealing. This leads to enhanced charge carrier mobility and lifetime, resulting in improved photocurrents. Overall, this strategy offers a general approach to prevent unfavorable changes of cation stoichiometry during high-temperature treatment of complex metal oxide photoelectrodes.
The above-mentioned limitations also apply to bismuth vanadate (BiVO4), which is one of the best performing metal oxide photoanodes for solar water splitting. The monoclinic scheelite phase, which is the most photoactive one, has a bandgap of 2.4–2.5 eV.12 Carrier transport in the material is slow, with reported carrier mobilities in the range of 10−2–10−1 cm2 V−1 s−1.10,13,14 In addition to having a small polaron conduction mechanism,10,11,15–17 the presence of defects in BiVO4 may lead to trap-mediated transport and/or bound polarons that reduce the mobility even further.13,14,18 Nevertheless, through various developments in the last 5–10 years, ∼90% of the theoretical maximum photocurrent of BiVO4 (7.5 mA cm−2, assuming all AM1.5 photons with energy higher than the 2.4 eV bandgap contribute to the photocurrent) has already been achieved.19 Pihosh et al. realized this record photocurrent by depositing a thin layer of BiVO4 onto WO3 nanorods to form a guest–host nanostructured photoelectrode. This orthogonalizes the direction of optical absorption and charge transport,20,21 which ensures that enough photons can be absorbed while the photogenerated carriers only need to travel short distances.
Despite these encouraging results, the fabrication of a complete solar water splitting device with a nanostructured BiVO4 photoanode is not straightforward. To generate the ∼1.5 V needed for water splitting, the BiVO4 is usually placed in front of one or two smaller-bandgap semiconductors to form a tandem device.22,23 Unfortunately, extensive optical scattering from nanostructured BiVO4 photoelectrodes prevents the low energy photons from reaching the bottom absorber. To circumvent this, Pihosh et al. used Pt as a reflective substrate and placed a double-junction GaAs/InGaAsP tandem solar cell in front of their BiVO4 at a 45° angle. The highly efficient III–V tandem cell absorbed enough of the diffusely back-reflected photons from the WO3/BiVO4 nanowires to reach an impressive 8.2% solar-to-hydrogen efficiency.19 In another report a beam splitter was used to direct the short wavelength part of the solar spectrum to the WO3/BiVO4 electrode and the long wavelength part to a perovskite solar cell, resulting in a device with a 7.7% solar-to-hydrogen efficiency.24 For practical applications, however, a simple stacked tandem configuration with a non-scattering top absorber (and without expensive III–V semiconductors) is preferred.25 To reduce the scattering of the WO3/BiVO4 nanowires one could revert to photolithographic techniques to create highly regular WO3 nanopillars followed by atomic layer deposition to deposit a BiVO4 top layer with a homogeneous thickness. In this paper, we aim for a potentially easier approach, which is to improve the carrier transport properties of the BiVO4. If successful, this would avoid the need for nanostructuring altogether.
One possible way to alleviate carrier transport limitations is to apply a high temperature treatment that may reduce the concentration of point defects and improve the crystallinity.26,27 While changes in the crystallinity can be readily observed by X-ray diffraction, changes in the nature and concentration of point defects are more difficult to track. To the best of our knowledge, no systematic studies have yet been carried out that correlate the influence of such heat treatments on the defect properties and the resulting PEC performance. In this paper, we systematically elucidate for the first time the formation of defects in BiVO4 thin films at high temperature in both oxidizing and reducing environments. We find that high temperature treatments of BiVO4 in air lead to an increase in grain size and improved carrier dynamics. At temperatures above 500 °C, however, vanadium is lost from the lattice. The concomitant decrease in photocurrent is presumably caused by the formation of vanadium vacancies. By introducing excess vanadium (V) in the gas phase during the heat treatment the formation of these vacancies can be avoided, leading to improved photocurrents.
The influence of vanadium loss on the crystal structure of BiVO4 thin films was studied by X-ray diffraction. Fig. 2 shows the grazing incidence X-ray diffractograms of as-prepared and annealed BiVO4 films deposited on fused silica (‘quartz’) substrates. All films show a pure monoclinic (scheelite-type) BiVO4 phase (clinobisvanite; space group: I2/b JCPDS card No. 14-0688) without any phase segregation. However, at temperatures of 600 °C and above, a re-orientation of the crystal lattice starts to occur and a strong preference for the (010) orientation is observed after annealing at 700 °C. This orientation is known to result in an enhanced photoelectrochemical performance.29,30 We attribute this re-orientation to changes in the amount of strain in the film, as caused by the vanadium loss and/or possible change in particle size.31 This is supported by the increased preference for the (010) orientation when uniaxially pressing BiVO4 powder into a pellet (Fig. S4†). The re-orientation of the crystal structure is also accompanied by a shift of the symmetric V–O stretching mode peak (∼826 cm−1) to a higher wavenumber, as shown in the Raman spectra (Fig. S5†). Since the wavenumber is inversely correlated with the interatomic distance,32 the observed shift can be attributed to a decrease in V–O bond length, which is consistent with the loss of vanadium that occurs under these conditions (Fig. 1).
To understand the influence of the annealing atmosphere, we performed the same high temperature MS measurement under argon flow. In contrast to the measurement in air (Fig. 1), neither V nor Bi loss was detected (Fig. S6†). Based on this, we conclude that the observed VO and VO2 species in the air-annealing experiment were formed by reaction of vanadium with oxygen from the environment and not oxygen from the lattice of BiVO4. The following defect-chemical reactions (in Kröger–Vink notation33) are proposed to describe the loss of vanadium:
(1) |
(2) |
It should be noted that a formal oxidation state of −5 for the vanadium vacancy is energetically highly unfavorable so that it is likely to spontaneously ionize to the −4 or −3 state:
(3) |
Since no elemental V loss was observed in the Ar-annealing experiment, the presence of V (g) when annealing in air is believed to be the product of the following disproportionation reaction:
2VO (g) ⇌ VO2 (g) + V (g) | (4) |
Indeed, the change of Gibbs free energy, ΔG, for reaction (4) is negative (−61.6 kJ mol−1 at 500 °C), suggesting that the reaction is spontaneous. The activation energy for the formation of in BiVO4 can be determined from an Arrhenius plot of the log of the ion current vs. the reciprocal temperature (Fig. S7†). The activation energies for reactions (2) and (3) are found to be 0.18 ± 0.02 eV and 0.23 ± 0.01 eV, respectively.
To investigate the influence of vanadium vacancies on the photoelectrochemical (PEC) performance, we fabricated thin film BiVO4 samples on FTO-coated glass substrates annealed under different temperatures and atmospheres as described in the Experimental section. The temperatures were limited between 450 and 550 °C, since scanning electron micrographs showed that agglomeration of the films starts to occur at temperatures above 550 °C (Fig. S8†). Within this temperature range, the crystal structure and orientation of the films do not change significantly (see Fig. 2). Since we are interested in the bulk charge separation properties of the films, the PEC measurements were performed in a 0.1 M phosphate buffer electrolyte (pH ∼ 7) with Na2SO3 added as a hole scavenger to remove any surface catalytic limitations. Fig. 3 shows the AM1.5 photocurrents at 1.2 V vs. RHE for BiVO4 films annealed at different temperatures up to 550 °C in air (black) and argon (purple) atmospheres. The heating atmosphere is found to have a profound influence on the resulting PEC performance. Annealing in Ar does not result in any change in the photocurrent of BiVO4. In contrast, the photocurrent is strongly affected by heat treatment in air. The photocurrent increases with increasing temperature until it reaches a maximum (∼1.5 fold increase) at about 500 °C. Above this temperature, the photocurrent decreases with increasing temperature.
Since the PEC measurements were performed with Na2SO3 as a hole scavenger, charge transfer limitations at the BiVO4/electrolyte interface do not play a role. Thus, the observed temperature dependence is unlikely to be caused by any changes in the surface chemistry. We further rule out this possibility by comparing the oxidation states of Bi and V in BiVO4 films annealed at 450 °C and 500 °C in air and Ar with XPS (Fig. S9†). The peak positions of all core levels remain the same, indicating no changes in the oxidation states of Bi and V despite the significant differences in the PEC performance. The lack of a peak shift is not surprising, considering that even a large amount of vanadium vacancies (1018–1019 cm−3) would still be much lower than the ∼1 at% detection limit of XPS. Furthermore, the photocurrent onset potential shows negligible variation with annealing temperature and atmosphere (see Fig. S10†), which further confirms that the heat treatments do not significantly affect the surface chemistry of the material.
Morphological changes with increasing temperature may also be the reason behind the observed PEC behavior. As shown in Fig. S8,† the grain size indeed increases with increasing annealing temperature. However, this is true for BiVO4 films annealed in air as well as those annealed in argon; thus, changes in morphology cannot explain the differences in photocurrent. Cross-section SEM images of the films show that the thickness of the films shows little change (Fig. S11†), and AFM measurements (Table S1†) reveal that the specific surface area of the films is also not significantly affected by the heat treatment.
We also rule out changes in optical absorption as a possible cause, since the monotonic increase of the long wavelength absorption with temperature (see Fig. S12†) is inconsistent with the observed photocurrent maximum at 500 °C in Fig. 3.
With changes in surface chemistry and film morphology being ruled out, we attribute the observed differences in PEC performance to the formation and reduction of defects in BiVO4. For samples annealed in air, the increase in grain size reduces the amount of grain boundary scattering, leading to an increase in photocurrent with temperature up to 500 °C (Fig. 3). Annealing in argon also results in larger grains, but, in contrast to air annealing, also leads to the formation of oxygen vacancies, . The formation of oxygen vacancies is evidenced by the loss of O2 without the accompanying loss of Bi or V, as shown in Fig. S6.† While the presence of can improve the conductivity, too many oxygen vacancies may reduce the space charge width and adversely affect charge separation. Moreover, it has been reported that can form deep trap states and act as recombination centers.34 This would cancel out the positive effect of the larger grain size and would explain the slight decrease in photocurrent for samples annealed in argon (Fig. 3). Anodic electrochemical treatment (i.e., applying positive bias) does not positively affect the photocurrent (Fig. S13†); thus, it seems not possible to re-fill the oxygen vacancies in our Ar-annealed BiVO4. At temperatures above 500° the photocurrent for samples annealed in air rapidly decreases (Fig. 3, black curve). Under these conditions the MS experiments showed a rapid loss of vanadium in the form of VO+, VO2+, and V+ (Fig. 1). This results in the formation of vanadium vacancies according to reactions (1) and (2). These have been reported to lie ∼0.3 eV above the valence band edge,35 which would be deep enough in the bandgap to form trap states and act as recombination centers. In contrast, no vanadium loss is observed when annealing in argon (Fig. S6†), nor does the photocurrent show any decrease under these conditions (Fig. 3, blue curve). Based on these observations, we conclude that the decrease in photocurrent for BiVO4 films annealed in air above 500 °C is due to the formation of vanadium vacancies.
We note that extending the annealing time of the BiVO4 films from 2 to 8 h at 500 °C in air also results in a lower PEC performance (Fig. S14†). We attribute this to the same V loss that causes the formation of V vacancies, which already starts at ∼450 °C (see Fig. 1).
The carrier transport properties of the air-annealed BiVO4 films are investigated using time-resolved microwave conductivity (TRMC). Typical TRMC curves (ϕΣμ vs. time) are shown in Fig. S15;† the carrier mobility can be obtained from the peak amplitude, and the carrier lifetime can be obtained from the decay of the transient curve.36Table 1 summarizes the TRMC results obtained at a laser pulse intensity of 3.5 × 1013 photons per pulse per cm2. Consistent with our previous report,13 the carrier mobility (μ) of BiVO4 annealed in air at 450 °C is rather low at ∼0.02 cm2 V−1 s−1, while its carrier lifetime (τ) of ∼70 ns is relatively long. The carrier diffusion length (LD) can be calculated using the following equation:
(5) |
(6) |
Temperature | Carrier mobility μ (10−2 cm2 V−1 s−1) | Carrier lifetime τ (ns) | Diffusion length LD (nm) |
---|---|---|---|
450 °C in air | 2 ± 0.4 | 68.8 ± 2.6 | 59.7 ± 0.3 |
500 °C in air | 21 ± 0.8 | 20.5 ± 1.4 | 105.5 ± 1.6 |
550 °C in air | 2 ± 0.1 | 19.7 ± 2.3 | 31.9 ± 0.1 |
550 °C in air + BiVO4 powder | 13.5 ± 2.5 | 34.7 ± 0.8 | 110.1 ± 0.9 |
It should be noted that diffusion of impurities from the FTO substrate (e.g. Sn from the F:SnO2 or Na from the glass) during high-temperature treatments may also affect the properties of the BiVO4. However, the transport properties measured by TRMC, which were obtained using Na- and Sn-free quartz substrates, show the same trends as the photocurrents that were measured using FTO substrates. Moreover, impurity diffusion cannot explain the pronounced influence of the annealing atmosphere that is observed in Fig. 3 and 4. Therefore, impurity diffusion is unlikely to play an important role in our experiments.
Our findings show that when BiVO4 is annealed at high temperature in air, the resulting loss of vanadium is detrimental to the PEC performance. As a strategy to overcome the vanadium loss, we annealed our BiVO4 film in the presence of BiVO4 powder (see Experimental section). By ensuring that the amount of BiVO4 powder exceeds the amount of BiVO4 in the film by several orders of magnitude, we create a VOx-rich atmosphere during the annealing. The high concentration of VOx species is then expected to shift the defect equilibrium of reactions (1) and (2) to the left and thus suppress the vanadium loss from the BiVO4 film. Under these conditions, no re-orientation of the crystal lattice is observed, even after annealing at temperatures as high as 700 °C (Fig. S16†). This suggests that the re-orientation of crystallites in the films is indeed induced by the loss of vanadium from the films. The exact mechanism of this re-orientation is still unclear and beyond the scope of the current study.
Fig. 4 shows that annealing in a VOx-rich atmosphere can partially compensate the loss of photocurrent observed in Fig. 3 when heating the films at 550 °C in air. This partial recovery is consistent with the improved carrier transport properties as measured by TRMC; annealing in VOx-rich atmosphere results in a 3.4-fold increase in the carrier diffusion length (Table 1).
Fig. 5 summarizes our findings on the influence of different annealing atmospheres on the properties of BiVO4 thin films. In all cases, the grain size of BiVO4 increases with increasing temperature, which results in a reduced number of grain boundaries and higher carrier mobilities. However, annealing in Ar introduces oxygen vacancies, which act as trap states in the film. This reduces the carrier lifetime and therefore nullifies the positive effect of larger grain sizes. The formation of oxygen vacancies can be prevented by annealing the samples in air. Unfortunately, the presence of air results in the loss of vanadium via volatile VOx species, especially above 500 °C. This is accompanied by a decrease in the photocurrent, which is presumably caused by the formation of vanadium vacancies that act as recombination centers. This can be avoided by annealing in the presence of BiVO4 powder in order to create a VOx-rich atmosphere. This prevents the formation of both oxygen vacancies and vanadium vacancies, and results in a BiVO4 film with higher carrier mobility and lifetime. This increases the carrier diffusion length and, therefore, the photocurrent of the films.
Fig. 5 Schematic illustration of the influence of annealing atmosphere on the morphology, defect formation, and carrier transport properties in BiVO4. |
Photoelectrochemical characterizations were performed in a three-electrode configuration. The measurement consisted of a Pt counter electrode, an Ag/AgCl reference electrode (XR300, saturated KCl and AgCl solution, Radiometer Analytical) and the BiVO4 films as the working electrode. Electrical contact to the films was provided by a copper wire as well as a tin-plated copper foil (CCK-18-101, Farnell) connected to the exposed FTO substrate. The electrolyte was 0.1 M potassium phosphate (KPi) buffer (pH ∼ 7) with added 0.5 M sodium sulfite (Na2SO3) as a hole scavenger. Potentials with respect to the reference electrode (VAg/AgCl) were applied by a potentiostat (EG&G PAR 273A) and converted to the reversible hydrogen electrode scale (VRHE) by using the Nernst equation:
VRHE (V) = VAg/AgCl (V) + 0.0591 × pH + VAg/AgCl0 | (7) |
Time-resolved microwave conductivity (TRMC) measurements were performed using a setup that has been described in detail elsewhere.14,37 In short, the BiVO4 films deposited on quartz were placed in a microwave cavity cell. A frequency-tripled Q-switched Nd:YAG laser at a wavelength of 355 nm was used as the excitation source with a 7 ns pulse. Microwaves in the X-band region (here: 8.4–8.7 GHz) were generated by a voltage controlled oscillator (SiversIMA VO3262X). For the carrier transport properties calculation, the dielectric constant of BiVO4 was taken as 68.43
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ta06269b |
This journal is © The Royal Society of Chemistry 2018 |