Next Article in Journal
Mixed-Degree Cubature H Information Filter-Based Visual-Inertial Odometry
Previous Article in Journal
Sparse Data Recommendation by Fusing Continuous Imputation Denoising Autoencoder and Neural Matrix Factorization
Previous Article in Special Issue
Effect of Excitation Beam Divergence on the Goos–Hänchen Shift Enhanced by Bloch Surface Waves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave-Assisted Hydrothermal Synthesis of SrTiO3:Rh for Photocatalytic Z-scheme Overall Water Splitting

Department of Materials Science and Engineering, National Dong Hwa University, Hualien 97401, Taiwan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(1), 55; https://doi.org/10.3390/app9010055
Submission received: 2 December 2018 / Revised: 17 December 2018 / Accepted: 20 December 2018 / Published: 24 December 2018
(This article belongs to the Special Issue Nanomaterials for Solar Water Splitting)

Abstract

:
Developing a photocatalyst system for solar energy conversion to electric energy or chemical energy is a topic of great interest for fundamental and practical importance. In this study, hydrogen production by a new Z-scheme photocatalysis water-splitting system was examined over Rh-doped SrTiO3 (denoted as Rh:SrTiO3) with Ru nanoparticle as cocatalyst for H2 evolution and BiVO4 photocatalyst for O2 evolution under visible light irradiation, where Co(bpy)32+/3+ was used as electron mediator. The catalysts were characterized by powder X-ray diffraction (XRD), scanning electron microscope (SEM), transmission electron microscopy (TEM), and Ultraviolet–visible spectroscopy. We present a fast and efficient method to synthesize Rh-doped SrTiO3 photocatalyst via microwave-assisted hydrothermal method. Our results showed a significant effect of Ti precursor on morphology of Rh:SrTiO3 prepared by microwave-assisted hydrothermal synthesis. The Ru/Rh:SrTiO3 prepared by TiCl4 precursor showed a nanoporous structure and high photocatalytic activity. The combination of Ru/Rh:SrTiO3 with BiVO4 achieves a high H2 evolution rate (317 μmoL g−1 h−1) and O2 evolution rate (168 μmol g−1 h−1) in 0.5 mM Co(bpy)32+/3+ solution under visible light irradiation.

1. Introduction

Hydrogen is an environmentally clean chemical fuel with high energy density. Developing a photocatalyst system to generate hydrogen from water gives us a chance to produce hydrogen from an inexhaustible renewable source, i.e., sunlight and water. Much effort has been devoted to studies of the splitting of water into hydrogen and oxygen in the years following the discovery of photocatalytic water splitting using a semiconductor photoelectrode, TiO2, as was demonstrated by Fujishima and Honda in 1972 [1]. From the thermodynamic perspective, it requires 237 kJ to transform one mole of water into hydrogen which is an energetically unfavorable uphill reaction. For a semiconductor photocatalyst for photocatalytic water-splitting reaction, the conduction band edge of the photocatalyst should be more negative than the reduction potential of H2O to form H2, and band edge of valence band should be more positive than the oxidation potential of H2O is required to form O2. The band gap of the semiconductor photocatalyst must be larger than the theoretical dissociation energy of the water molecule (1.23 eV). Several studies have reported that transition metal oxides with d0 and d10 electron configurations have showed photocatalytic activity of water splitting under UV light such as NaTaO3 [2], K4Nb6O17 [3] and M2Sb2O7 (M = Ca, Sr) [4]. Since 1998, there have been several excellent reviews in photocatalysis water splitting [5,6,7]. The sun is the most important light source of our world. At sea level, there is only about 8% of solar ultraviolet radiation (200–400 nm), about 50% is visible light, and about 40% is infrared radiation. Hence, numerous studies have attempted to develop highly efficient photocatalysts for water-splitting reaction which can be activated under visible light irradiation [8,9,10]. This improved photocatalytic activity can be achieved by introducing impurities such as transition metal [11] and non-metal elements [12] for a smaller band gap by lowering condition band edge or a rising valence band edge for use of solar light. However, a more negative conduction bend potential and a more positive valance band potential are thermodynamically favorable for the photocatalysis reduction and oxidation reactions. It is clear that satisfying these two conflicting conditions is a challenge for the development of visible light-responsive photocatalysts.
In recent years, a semiconductor-based Z-scheme system for overall water splitting has drawn much attention [13,14,15]. In a Z-scheme system, photocatalytic H2 evolution and O2 evolution occurs on two different photocatalysts and a redox couple as electron mediator. As compared to a conventional photocatalytic system with one photocatalyst, a Z-scheme system combines two photocatalysts for water reduction and oxidation where the electronic structure of the photocatalysts only have to satisfy the thermodynamics requirements to the half reactions for reduction and oxidation and thereby expands the variety of potential candidates for overall photocatalytic water splitting under visible light irradiation. In particular, SrTiO3 photocatalyst has high potential for photocatalytic applications due to its high photocatalytic activity for water-splitting reaction. However, due to the wide band gap of SrTiO3 (3.2 eV), intensive research has been carried out on doping transition metal ions in SrTiO3 to reduce band gap and improve photocatalytic activity under visible light irradiation [16,17]. A visible-light-driven Z-scheme overall water-splitting system using Ru loaded Rh-doped SrTiO3 (Ru/SrTiO3:Rh) photocatalyst for H2 evolution, BiVO4 photocatalyst for O2 evolution photocatalyst and a Tris (2,2′-bipyridine) cobalt(II)/Tris (2,2′-bipyridine) cobalt(III) used as electron mediator was achieved by Kudo and co-workers [13].
It is well known that, crystallinity is an important factor to influence photocatalytic activity. To achieve high crystalline quality, solid-state reaction is widely employed to synthesize semiconductor photocatalyst. However, the high-temperature synthesis process yields large particles and low surface area which is another important factor to consider. Several routes were taken to prepare SrTiO3 nano particles via wet chemical route, including sol-gel synthesis and hydrothermal synthesis. It was reported that a variety of SrTiO3 nanoparticles with specific morphologies such as nanocubes, nanowires, and porous structure can be fabricated via hydrothermal method [18].
In this paper, we present the synthesis of Rh-doped SrTiO3 photocatalysts via a one-step microwave-assisted hydrothermal method. In the microwave-assisted hydrothermal process, heat was not only transferred from the outside to the inside of autoclave reactor as conventional heating, but also generated from the inside of SrTiO3 due to its dielectric properties which led to a faster, and more efficient synthesis process [19]. By using anatase TiO2 and TiCl4 as titanium precursor, high crystallinity Rh-doped SrTiO3 photocatalysts with different morphologies were synthesized quickly and efficiently, and no post treatment at high temperature was required. The influence of Ti precursors on the morphology of Rh-doped SrTiO3 photocatalysts, as well as the photocatalytic activity for the Z-scheme overall water-splitting reaction under visible light irradiation were investigated. The results were compared with Rh-doped SrTiO3 prepared by solid-state method and conventional hydrothermal method.

2. Materials and Methods

2.1. Synthesis of Rh-Doped SrTiO3 with Ru Cocatalyst for H2 Evolution

Two different titanium precursors, anatase TiO2 (Alfa aesar) and TiCl4 Alfa aesar), were used in the microwave-assisted hydrothermal synthesis of Rh-doped SrTiO3.
The Rh-doped SrTiO3 prepared by anatase TiO2, MW-STO2Rh-A, were synthesized by using anatase TiO2, Sr(OH)2 and Na3RhCl6 as starting materials, where the molar composition of Ti:Sr:Rh was 100:98:2. The starting materials was added to 30 mL of 5M NaOH aqueous solution in sequence under vigorous stirring. Then, the start material was transferred to a 100 cm3 Teflon autoclave and the microwave-assisted hydrothermal treatment was performed in a commercial microwave digestion system (StartD, Milestone, Taiwan, 2011). The microwave hydrothermal process was performed at 180 °C for 1 h. After synthesis reactions, the product was washed by distillated H2O and dried at 60 °C overnight.
The Rh-doped SrTiO3 prepared by TiCl4, referred as MW-STO2Rh-T, were synthesized by a similar procedure: TiCl4, 1 M solution in toluene (Sigma-Aldrich, used as purchased), Sr(OH)2 and Na3RhCl6 were used as starting materials where the molar composition of Ti:Sr:Rh was 100:98:2. The color of MW-STO2Rh-A and MW-STO2Rh-T varies from light yellow-green to gray-green. The color of samples became darker with increasing Rh amount.
In this study, conventional hydrothermal synthesis and solid-state synthesis of Rh-doped SrTiO3 photocatalyst, denoted as HT-STO2Rh, and SS-STO2Rh were also prepared by for comparison. The HT-STO2Rh were synthesized by using anatase TiO2, Sr(OH)2 and Na3RhCl6 where the molar ratio of Ti:Sr:Rh= 100:98:2. The starting materials was added to 30 mL of 5M NaOH aqueous solution in sequence under vigorous stirring. The precursor solution was transferred to a 100 cm3 Teflon lined autoclave. Then, the hydrothermal synthesis was carried out at 180 °C for 24 h. The HT-STO2Rh showed a similar color to that of MW-STO2Rh-A and MW-STO2Rh-T.
The SS-STO2Rh were prepared a two-step solid-state reaction using anatase TiO2, SrCO3, Rh2O3 (molar ratio 100: 98: 1). The mixed precursor was calcinated in air at 1273 K for 10h then at 1273 K for 10 h with an intermediate grinding process between the two calcinations; essentially similar of the procedure by Konta et al. [20]. The color of SS-STO2Rh was black.
Ru nanoparticles were used as a H2 evolution cocatalyst and were loaded on Rh-doped SrTiO3 by photo-deposition method. Rh-doped SrTiO3 and an aqueous solution RuCl3∙xH2O (1 wt % Ru loading for complete photoreduction) were added to a methanol solution (vol. ratio of MeOH:H2O = 1:10). The suspension was irradiated in a photoreactor with a 400 W medium-pressure metal halide lamp (Phillips HPA400) for 12 h under string at room temperature. The sample was then filtered and washed with distilled water. Finally, the photocatalyst was dried at 60 °C overnight.

2.2. Synthesis of BiVO4 for O2 Evolution

The BiVO4 photocatalyst for O2 evolution was synthesized by microwave-assisted hydrothermal method. Bi2O3 and V2O5 in a molar ratio of 1:1 was added to 30 mL of 0.5 M HNO3 solution under vigorous stirring for 20 min. Then, the starting materials was transferred to a 100 cm3 Teflon reaction vessel and the microwave hydrothermal process was performed at 180 °C for 1 h. After synthesis reactions, the product was washed by distillated H2O and dried at 60 °C overnight. The BiVO4 photocatalyst had a vivid orange-yellow color.

2.3. Synthesis of [Co(bpy)3]SO4 for Electron Mediator

[Co(bpy)3]SO4 was synthesized by mixing stoichiometric amounts of CoSO4·7H2O and 2,2′-bipyridine (Bpy) in aqueous ethanol solution (vol. ratio of EtOH:H2O = 2:1). The solution was heated under stirring to the ligand was totally dissolved. The mixture was cooled to room temperature, to precipitate the crude product. Then, the product was recrystallized to remove impurities [13].

2.4. Characterization and Photocatalytic Reactions

The characterization methods included Powder X-ray diffraction (XRD, Rigaku X-ray diffractometer, MAX-2500V, Cu-Kα radiation, λ= 1.54178 Ǻ), UV-vis diffuse reflectance spectra (Varian Lary 5E diode array spectrometer, Varian), and FE-SEM (Field emission scanning electron microscope, JSM-7000F, JEOL), and transmission electron microscopy (TEM, JEM-2000FX microscope, JEOL).
The photocatalytic reaction was carried out in a reactor equipped with an inner irradiation quartz cell with a cooling water jacket under stirring at 43 °C. A 150 W Xe lamp was mounted inside the quartz cell as shown in Figure 1. The photocatalytic Z-scheme water splitting was performed in 550 mL 0.5 mM aqueous Co(bpy)3SO4 solution containing 0.2 g of H2 evolution and O2 evolution photocatalysts, respectively. The photocatalysts were used as prepared without treatment. The gas product was analyzed by a gas chromatography (China Gas Chromatography 9800) with a packed column (MS-5A, 3.5 m in length) and thermal conductivity detector.

3. Results and Discussion

3.1. Characterization of the Hydrogen Evolution Catalyst: Rh-Doped SrTiO3

Figure 2 shows the XRD patterns Rh-doped SrTiO3 photocatalysts prepared by different methods: conventional solid-state reaction (SS), hydrothermal synthesis (HT) and microwave-assisted hydrothermal method (MW). The XRD patterns of all prepared materials show obvious diffraction peaks at 2θ = 32.4°, 40.0°, 46.5°, and 57.8° which corresponded to the (110), (111), (200) and (221) planes of SrTiO3 (JCPDS 35-0734). No reflection peak corresponded to rhodium or rhodium oxide species of doping elements suggesting that dopant ions were introduced into structure of SrTiO3 or were too small to be detected by XRD.
The SS-STO2Rh showed a much weaker diffraction peaks than samples prepared via hydrothermal route. As compared with conventional solid-state reaction method, a better crystalline structure could be formed for Rh-doped SrTiO3 prepared by hydrothermal synthesis. The small diffraction peaks at 2θ of 31.5° were considered as the presence of Sr2TiO4 (JCPDS-39-1471).
On the other hand, the XRD results also shows that samples prepared via hydrothermal route contained trace amount of SrCO3 (JCPDS 84-0418, diffraction peaks located in 2θ of 20.5° and 25.4° as the (110) and (111) planes SrCO3) for which was due to the reaction of environment CO2 with the precursors such as Sr(OH)2 and NaOH. The sharp diffraction peaks indicated that the samples were fully crystallized SrTiO3 with cubic space group Pm3m (221). The average crystallite sizes of Rh-doped SrTiO3 samples (HT-STO2Rh, MW-STO2Rh-A and MW-STO2Rh-T) prepared by hydrothermal method were about 3.6 nm, determined by Scherrer equation. It should be noted that while conventional hydrothermal synthesis of Rh-doped SrTiO3 required a long synthesis time (>12 h), high crystallinity Rh-doped SrTiO3 was formed in a very short time (about 1 h) during microwave-assisted hydrothermal synthesis [21]. These results indicated that the lattice deformation of Rh-doped SrTiO3 are associated with the synthesis route.
To obtain the effects of synthesis route on their size and morphology, the samples were subjected to TEM analysis. Figure 3 depicts the TEM images of the Rh-doped SrTiO3 prepared from different methods. The TEM results show that their morphology was significant different. The high-temperature solid-state synthesis caused the formation of large SS-STO2Rh particles with a diameter of 1–5 μm while the HT-STO2Rh catalyst showed smaller particles ranged from 50–500 nm. The TEM images also revealed that the MW-STO2Rh-A and MW-STO2Rh-T samples prepared by microwave-assisted hydrothermal method have a smallest particle sizes of about 100–200 nm. The TEM images also illustrated that the morphology Rh-doped SrTiO3 by microwave-assisted hydrothermal method was changed sensitively by the titanium precursor. The MW-STO2Rh-A showed uniform cubic nano particles whereas the MW-STO2Rh-T showed a porous nanocrystal structure.
Synthesis of anatase SrTiO3 through hydrothermal reaction in NaOH solution was based on a series of dissolution-precipitation steps. The possible growth mechanism of SrTiO3 nanoparticles form can be considered as following [18]:
TiO2 (s) + (n − 4) OH (aq)(l) → Ti(OH)(n−4)−n (aq) n = 4–6
Ti(OH)(n−4)−n (aq) + Sr2+ (aq) + (6 − b) OH (aq) → SrTiO3 (s) + 3H2O
The TiO2 can be dissolved in alkaline solution and converted to soluble Ti(OH)(n−4)−n, n = 4–6, and further form SrTiO3 precipitation with Sr2+ ion. Kalyani et al. [18] reported that the hydrothermal process of was limited by the dissolution rate of titanium precursor where cubic shape SrTiO3 particles was formed form anatase TiO2 precursor and porous SrTiO3 particles was formed by TiOCl2 due to its rapid hydrolysis of soluble titanium complexes at 200 °C for 13–24 h. In this study, we showed a porous nanosphere of Rh-doped SrTiO3 can be synthesized by 1 M TiCl4 in toluene as titanium precursor, where the rapid hydrolysis of TiCl4 formed Ti(OH)n nano aggregates with a porous nanostructure in alkaline solution, which can be described as:
TiCl4 (aq) +10 H2O → Ti(OH)5 + 5H3O++ 4Cl
And reacted with Sr2+ ions and form porous SrTiO3 particles. Moreira et al. reported that the TiCl4 formed Ti(OH)5 and Sr(OH)2 clusters, then strong and stable Ti-O-Ti and Sr-O-Sr bond was formed during the dehydration processes and self-assembled to form porous SrTiO3 nanoparticles [22].
On the other hand, while small irregular particles (about 10–100 nm) was found in TEM image of SS-STO2Rh which could be Rh2O3 form via high-temperature solid-state reaction, no particles can be observed on the particle surfaces of HT-STO2Rh, MW-STO2Rh-A and MW-STO2Rh-T samples. It suggests that the Rh3+ ions from Na3RhCl6 were substituted into the Sr site of SrTiO3 efficiently through liquid phase hydrothermal synthesis without forming of Rh or RhOx nanoparticles on the surface of SrTiO3.
These results also indicated that while conventional hydrothermal synthesis of SrTiO3 required a long synthesis time (>12 h), high crystallinity Rh-doped SrTiO3 was formed in a very short time (about 1 h) during microwave-assisted hydrothermal synthesis.
Figure 4 shows the TEM images of Ru/MW-STO2Rh-A and Ru/MW-STO2Rh-T. As shown in Figure 4a, large Ru nanoparticles (>20 nm) were observed on the surface of MW-STO2Rh-A cubic nanoparticle. This suggests that Ru species could migrate easily along the particle surface of MW-STO2Rh-A, and form large Ru nanoparticles. On the other hand, the TEM image of Ru/MW-STO2Rh-T catalyst (Figure 4b) revealed clearly that about 2–5 nm Ru nanoparticles were formed on the rough particle surface of MW-STO2Rh-T. Our results showed that the deposition of Ru nanoparticles on the Rh-doped SrTiO3 was affected by the surface morphology. The diffusion barrier of Rh species on a rough and porous surface was higher than on a well-crystallized SrTiO3 particle with smooth surface. The precursor not only plays a significant role on particle morphology of Rh-doped SrTiO3 photocatalysts but also the particle size distribution of Ru cocatalysts which were deposited to the surface of SrTiO3 by photoreduction reaction.
The light absorption capacity of photocatalysts influences the efficiency of any photocatalytic reaction. Figure 5 shows the UV-vis diffuse reflectance spectra of the SrTiO3 and Rh-doped SrTiO3 samples prepared by different method. The band gap of the pristine SrTiO3 (MW-STO-A) estimated form the absorption edge of the UV-vis spectrum was about 3.2 eV. As can be seen, the absorption edge of Rh-doped SrTiO3 samples were clearly shifted to long wave region (Figure 5b–e). In this study, all the metal loading of catalysts was maintained at about 2 mol% to investigate the effects of other variables. Two shoulder peaks at 580 nm and 420 nm were observed in UV-vis spectra of Rh-doped SrTiO3 samples which were attributed to the electron transition of the in-gap levels induced by Rh doping, corresponding to the existence of Rh4+ and Rh3+ species, respectively. As compared to SS-STO2Rh, and HT-STO2Rh prepared by conventional method, the MW-STO2Rh-A and MW-STO2Rh-T showed a lower absorption peak at 580 nm which suggested that the Rh-doped SrTiO3 prepared by microwave-assisted hydrothermal method consisted less Rh4+ species than SS-STO2Rh prepared by solid-state method [23]. These results indicated donor and acceptor levels were introduced in the band gap of SrTiO3 by the doping Rh.

3.2. Characterization of the Oxygen Evolution Catalyst: BiVO4

The XRD (Figure 6) results show that single phase monoclinic BiVO4 (JCPDS 14-0688) was synthesized successfully by the microwave-assisted hydrothermal method. The optical absorption spectra of the BiVO4 are shown in Figure 7. The strong absorption lies in visible light region from 400 nm to 550 nm. The band gap of the BiVO4 sample estimated form the UV-vis spectrum was about 2.4 eV. Figure 8 depicts the SEM images of the BiVO4 catalyst which are dual-faceted microcrystals with {010} and {110} facets [24].

3.3. Photocatalytic Water-Splitting Activity

Figure 9 shows the H2 and O2 evolution time course on Ru/Rh:SrTiO3–BiVO4 photocatalysts taken from the 0.5 mM Co(bpy)32+ solution. The gas production stopped when the light was turned off and similar hydrogen and oxygen production rates were observed in 8 h. This indicates that the Z-scheme overall water splitting on these catalysts was photocatalytic and that the photocatalysts can be used without deactivation. Furthermore, the H2 and O2 ratio of products were constant during the reaction and close to the theoretical stoichiometric ratio (H2/O2 = 2).
The related reactions of the Z-scheme photocatalytic water splitting are written as following [13]:
On H2 evolution photocatalyst:
Co(bpy)32+ + h+ → Co(bpy)33+
2H+ + 2e → H2
On H2 evolution photocatalyst:
2 H2O + 4h+ → O2 + 4H+
Co(bpy)33+ + e → Co(bpy)32+
The average production rates of hydrogen and oxygen from water splitting over this period are listed in Table 1 which increased as the following order: Ru/HT-STO2Rh < Ru/SS-STO2Rh << Ru/MW-STO2Rh-A < Ru/SS-STO2Rh-T. These results show that the Rh-doped SrTiO3 photocatalyst prepared by microwave-assisted hydrothermal method exhibited a very high hydrogen evolution rate of water-splitting reaction under visible light irradiation. The activity of Ru/SS-STO2Rh-T drastically increased more than 5 times as compared to Ru/SS-STO2Rh, and 10 times than Ru/HT-STO prepared by convention hydrothermal method.
Conventional solid-state reaction is widely used to synthesize semiconductor photocatalyst. However, the high-temperature synthesis process leads to large particles and low surface area. It is known that preparation of photocatalyst nanoparticles via hydrothermal method. However, crystallinity is one of the most important factors of the photocatalytic activity of photocatalyst. Although HT-STO2Rh sample consisted smaller particles than SSR-STO2Rh sample, it showed a lower activity as compared to the SSR-STO2Rh sample.
Kato et al. [16] reported that the SrO defects in the Rh-doped SrTiO3 photocatalyst prepared by hydrothermal method are the recombination centers for the photogenerated electron and holes which is the main reason of low activity. Their results indicated that a post calcination at 1273 K with excess Sr was necessary to decrease SrO defects of in the Rh-doped SrTiO3 photocatalyst prepared by hydrothermal method. The low activity of Ru/HT-STO2Rh was attributed to the SrO defects formed in hydrothermal reaction.
The high activity of Ru/MW-STO2Rh-A and Ru/SS-STO2Rh-T samples indicated that the microwave-assisted hydrothermal method has distinct advantages in the synthesis of high crystallinity Rh-doped SrTiO3 nanoparticles with low SrO defects. Secondly, a lower Rh4+ cations as compared to that of Rh:SrTiO3 prepared by conventional solid-state reaction and hydrothermal method may also play a role in improving activity of Ru/MW-STO2Rh-A and Ru/SS-STO2Rh-T samples [11]. Furthermore, the Ru/SS-STO2Rh-T showed a much higher activity than Ru/SS-STO2Rh-A. This suggests that the photo-excited electron can transfer more efficiently from SS-STO2Rh-T to the well dispersed Ru cocatalysts and reduce electron-hole recombination. On the other hand, the larger and less uniform Ru particles are attributed to the lower activity of Ru/MW-STO2Rh-A (Figure 4). Our results also show that the photocatalytic steps on Ru/Rh:SrTiO3 is the rate limit step of the Z-scheme system of Ru/Rh:SrTiO3–BiVO4. The activity of Ru/Rh:SrTiO3 is determined by complex factors, such as preparation method of Rh:SrTiO3, titanium precursor, and the size and dispersion of Ru cocatalyst. It should be noticed that the photocatalysts were used as prepared without pretreatment. Therefore, it was believed that there were no effective heterojunctions between Ru/Rh:SrTiO3 and BiVO4 particles. The electron transformations were relied on Co[(bpy)]32+/2+, the electron mediator redox couple. On the other hand, reduced graphene oxide as a solid electron mediator for Z-scheme water splitting was reported Iwase et al. [25] which indicated that the surface properties of the solid-state electron mediator played an important role for the photocatalytic activity. Our study has demonstrated a simple method to synthesize highly efficient photocatalysts by microwave-assisted hydrothermal method. Further studies are needed for the applications in a Z-scheme system with solid-state electron mediator.

4. Conclusions

In this study, we demonstrated that microwave-assisted hydrothermal method has distinct advantages in synthesis Rh-doped SrTiO3 nanoparticles with high crystallinity. Our results showed that significant effect of Ti precursor on morphology of Rh:SrTiO3. The Rh:SrTiO3 nanoparticle in with uniform cubic structure and porous frameworks were successfully synthesized by anatase TiO2 and TiCl4 precursor, respectively. The Rh:SrTiO3 prepared by the microwave-assisted hydrothermal method worked as a hydrogen evolution catalyst for the Z-scheme overall water-splitting system and exhibited a high and stable photocatalytic activities. The Ru/MW-STO2Rh-T-BiVO4 Z-scheme system exhibited the highest photocatalytic activities with a rate H2 evolution rate (317 μmol g−1 h−1) and O2 evolution rate (168 μmol g−1 h−1) in 0.5 mM Co(Bpy)32+/3+ solution under visible light irradiation. The great enhancement of photocatalytic activity of Ru/MW-STO2Rh is attributed to the levels enhanced visible light absorption by Rh donor levels and the efficient charge separation achieved by the highly dispersed Ru cocatalysts on the porous structure of MW-STO2Rh.

Author Contributions

All authors work together on performed the experiments and analysis data.

Funding

This research was funded by the National Science Council, Taiwan, Republic of China (MOST 107-2221-E-259-009).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Kato, H.; Asakura, K.; Kudo, A. Highly Efficient water splitting into H2 and O2 over lanthanum-doped NaTaO3 photocatalysts with high crystallinity and surface nanostructure. J. Am. Chem. Soc. 2003, 125, 3082–3089. [Google Scholar] [CrossRef] [PubMed]
  3. Lin, H.Y.; Lee, T.H.; Sie, C.Y. Photocatalytic hydrogen production with nickel oxide intercalated K4Nb6O17 under visible light irradiation. Int. J. Hydrogen Energy 2008, 33, 4055–4063. [Google Scholar] [CrossRef]
  4. Sato, J.; Saito, N.; Nishiyama, H.; Inoue, Y. Photocatalytic water decomposition by RuO2-loaded antimonates, M2Sb2O7 (M = Ca, Sr), CaSb2O6 and NaSbO3, with d10 configuration. J. Photochem. Photobiol. A 2002, 148, 85. [Google Scholar] [CrossRef]
  5. Takata, T.; Tanaka, A.; Hara, M.; Kondo, J.N.; Domen, K. Recent progress of photocatalysts for overall water splitting. Catal. Today 1998, 44, 17–26. [Google Scholar] [CrossRef]
  6. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  7. Li, X.; Yu, J.; Low, J.; Fang, Y.; Xiao, J.; Chen, X. Engineering heterogeneous semiconductors for solar water splitting. J. Mater. Chem. A 2015, 3, 2485–2534. [Google Scholar] [CrossRef]
  8. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef]
  9. Suzuki, T.M.; Iwase, A.; Tanaka, H.; Sato, S.; Kudo, A.; Morikawa, T. Z-scheme water splitting under visible light irradiation over powdered metal-complex/semiconductor hybrid photocatalysts mediated by reduced graphene oxide. J. Mater. Chem. A 2015, 3, 13283–13290. [Google Scholar] [CrossRef]
  10. Lin, H.Y.; Shih, C.Y. Efficient one-pot microwave-assisted hydrothermal synthesis of M (M = Cr, Ni, Cu, Nb) and nitrogen co-doped TiO2 for hydrogen production by photocatalytic water splitting. J. Mol. Catal. A Chem. 2016, 411, 128–137. [Google Scholar] [CrossRef]
  11. Sasaki, Y.; Nemoto, H.; Saito, K.; Kudo, A. Solar Water Splitting Using Powdered Photocatalysts Driven by Z-Schematic Interparticle Electron Transfer without an Electron Mediator. J. Phys. Chem. C 2009, 113, 17536–17542. [Google Scholar] [CrossRef]
  12. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  13. Sasaki, Y.; Kato, H.; Kudo, A. [Co(bpy)3]3+/2+ and [Co(phen)3]3+/2+ Electron Mediators for Overall Water Splitting under Sunlight Irradiation Using Z-Scheme Photocatalyst System. J. Am. Chem. Soc. 2013, 135, 5441–5449. [Google Scholar] [CrossRef] [PubMed]
  14. Maeda, K.; Lu, D.; Domen, K. Solar-Driven Z-scheme Water Splitting Using Modified BaZrO3-BaTaO2N Solid Solutions as Photocatalysts. ACS Catal. 2013, 3, 1026–1033. [Google Scholar] [CrossRef]
  15. Wang, Q.; Li, Y.; Hisatomi, T.; Nakabayashi, M.; Shibata, N.; Kubota, J.; Domen, K. Z-scheme water splitting using particulate semiconductors immobilized onto metal layers for efficient electron relay. J. Catal. 2015, 328, 308–315. [Google Scholar] [CrossRef]
  16. Kato, H.; Sasaki, Y.; Shirakura, N.; Kudo, A. Synthesis of highly active rhodium-doped SrTiO3 powders in Z-scheme systems for visible-light-driven photocatalytic overall water splitting. J. Mater. Chem. A 2013, 1, 12327–12333. [Google Scholar] [CrossRef]
  17. Tonda, S.; Kumar, S.; Anjaneyulu, O.; Shanker, V. Synthesis of Cr and La-codoped SrTiO3 nanoparticles for enhanced photocatalytic performance under sunlight irradiation. Phys. Chem. Chem. Phys. 2014, 16, 23819–23828. [Google Scholar] [CrossRef] [PubMed]
  18. Kalyani, V.; Vasile, B.S.; Ianculescu, A.; Testino, A.; Carino, A.; Buscaglia, M.T.; Buscaglia, V.; Nanni, P. Hydrothermal Synthesis of SrTiO3: Role of Interfaces. Cryst. Growth Des. 2015, 15, 5712–5725. [Google Scholar] [CrossRef]
  19. Zheng, J.-Q.; Zhu, Y.-J.; Xu, J.-S.; Lu, B.-Q.; Qi, C.; Chen, F.; Wu, J. Microwave-assisted rapid synthesis and photocatalytic activity of mesoporous Nd-doped SrTiO3 nanospheres and nanoplates. Mater. Lett. 2013, 100, 62–65. [Google Scholar] [CrossRef]
  20. Konta, R.; Ishii, T.; Kato, H.; Kudo, A. Photocatalytic activities of noble metal ion doped SrTiO3 under visible light irradiation. J. Phys. Chem. B 2004, 108, 8992–8995. [Google Scholar] [CrossRef]
  21. Kiss, B.; Manning, T.D.; Hesp, D.; Didier, C.; Taylor, A.; Pickup, D.M.; Chadwick, A.V.; Allison, H.E.; Dhanak, V.R.; Claridge, J.B.; et al. Nano-structured rhodium doped SrTiO3-Visible light activated photocatalyst for water decontamination. Appl. Catal. B-Environ. 2017, 206, 547–555. [Google Scholar] [CrossRef]
  22. Moreira, M.L.; Longo, V.M.; Avansi, W., Jr.; Ferrer, M.M.; Andres, J.; Mastelaro, V.R.; Varela, J.A.; Longo, E. Quantum Mechanics Insight into the Microwave Nucleation of SrTiO3 Nanospheres. J. Phys. Chem. C 2012, 116, 24792–24808. [Google Scholar] [CrossRef]
  23. Kawasaki, S.; Akagi, K.; Nakatsuji, K.; Yamamoto, S.; Matsuda, I.; Harada, Y.; Yoshinobu, J.; Komori, F.; Takahashi, R.; Lippmaa, M.; et al. Elucidation of Rh-Induced In-Gap States of Rh:SrTiO3 Visible-Light-Driven Photocatalyst by Soft X-ray Spectroscopy and First-Principles Calculations. J. Phys. Chem. C 2012, 116, 24445–24448. [Google Scholar] [CrossRef]
  24. Tan, H.L.; Tahini, H.A.; Wen, X.; Wong, R.J.; Tan, X.; Iwase, A.; Kudo, A.; Amal, R.; Smith, S.C.; Ng, Y.H. Interfacing BiVO4 with Reduced Graphene Oxide for Enhanced Photoactivity: A Tale of Facet Dependence of Electron Shuttling. Small 2016, 12, 5295–5302. [Google Scholar] [CrossRef] [PubMed]
  25. Iwase, A.; Ng, Y.H.; Ishiguro, Y.; Kudo, A.; Amal, R. Reduced graphene oxide as a solid-state electron mediator in Z-scheme photocatalytic water splitting under visible light. J. Am. Chem. Soc. 2011, 133, 11054–11057. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Closed reaction system for photocatalysis water splitting (b) Emission spectrum of 150 W Xe lamp.
Figure 1. (a) Closed reaction system for photocatalysis water splitting (b) Emission spectrum of 150 W Xe lamp.
Applsci 09 00055 g001
Figure 2. XRD patterns of Rh-doped SrTiO3 catalysts (a) SS-STO2Rh (b) HT-STO2Rh (c) MW-STO2Rh-A (d) MW-STO2Rh-T.
Figure 2. XRD patterns of Rh-doped SrTiO3 catalysts (a) SS-STO2Rh (b) HT-STO2Rh (c) MW-STO2Rh-A (d) MW-STO2Rh-T.
Applsci 09 00055 g002
Figure 3. TEM images of Rh-doped SrTiO3 catalysts (a) SS-STO2Rh (b) HT-STO2Rh (c) MW-STO2Rh-A (d) MW-STO2Rh-T.
Figure 3. TEM images of Rh-doped SrTiO3 catalysts (a) SS-STO2Rh (b) HT-STO2Rh (c) MW-STO2Rh-A (d) MW-STO2Rh-T.
Applsci 09 00055 g003
Figure 4. TEM images of Rh-doped SrTiO3 catalysts with Ru cocatalysts (a) Ru/MW-STO2Rh-A (b) Ru/MW-STO2Rh-T.
Figure 4. TEM images of Rh-doped SrTiO3 catalysts with Ru cocatalysts (a) Ru/MW-STO2Rh-A (b) Ru/MW-STO2Rh-T.
Applsci 09 00055 g004
Figure 5. UV-vis spectra of Rh-doped SrTiO3 catalysts (a) MW-STO-A (b) SS-STO2Rh (c) HT-STO2Rh (d) MW-STO2Rh-A (e) MW-STO2Rh-T.
Figure 5. UV-vis spectra of Rh-doped SrTiO3 catalysts (a) MW-STO-A (b) SS-STO2Rh (c) HT-STO2Rh (d) MW-STO2Rh-A (e) MW-STO2Rh-T.
Applsci 09 00055 g005
Figure 6. XRD spectrum of BiVO4.
Figure 6. XRD spectrum of BiVO4.
Applsci 09 00055 g006
Figure 7. UV-vis spectrum of BiVO4.
Figure 7. UV-vis spectrum of BiVO4.
Applsci 09 00055 g007
Figure 8. FE-SEM image of BiVO4.
Figure 8. FE-SEM image of BiVO4.
Applsci 09 00055 g008
Figure 9. Time course of H2 evolution and O2 evolution for the Z-scheme photocatalysis water splitting on 1 wt % Ru/Rh:SrTiO3–BiVO4 where (a) Ru/SS-STO2Rh (b) Ru/HT-STO2Rh (c) Ru/MW-STO2Rh-A (d) Ru/MW-STO2Rh-T was used as H2 evolution catalyst.
Figure 9. Time course of H2 evolution and O2 evolution for the Z-scheme photocatalysis water splitting on 1 wt % Ru/Rh:SrTiO3–BiVO4 where (a) Ru/SS-STO2Rh (b) Ru/HT-STO2Rh (c) Ru/MW-STO2Rh-A (d) Ru/MW-STO2Rh-T was used as H2 evolution catalyst.
Applsci 09 00055 g009
Table 1. Photocatalytic water-splitting activity of Rh-doped SrTiO3–BiVO4 in 0.5 mM CoBpy2+/3+ solution.
Table 1. Photocatalytic water-splitting activity of Rh-doped SrTiO3–BiVO4 in 0.5 mM CoBpy2+/3+ solution.
CatalystH2 Evolution Rate (μmol eg−1 h−1)O2 Evolution Rate (μmol eg−1 h−1)
Ru/SS-STO2Rh5635
Ru/HT-STO2Rh2816
Ru/MW-STO2Rh-A14175
Ru/MW-STO2Rh-T317168

Share and Cite

MDPI and ACS Style

Lin, H.-y.; Cian, L.-T. Microwave-Assisted Hydrothermal Synthesis of SrTiO3:Rh for Photocatalytic Z-scheme Overall Water Splitting. Appl. Sci. 2019, 9, 55. https://doi.org/10.3390/app9010055

AMA Style

Lin H-y, Cian L-T. Microwave-Assisted Hydrothermal Synthesis of SrTiO3:Rh for Photocatalytic Z-scheme Overall Water Splitting. Applied Sciences. 2019; 9(1):55. https://doi.org/10.3390/app9010055

Chicago/Turabian Style

Lin, Hsin-yu, and Lyu-Ting Cian. 2019. "Microwave-Assisted Hydrothermal Synthesis of SrTiO3:Rh for Photocatalytic Z-scheme Overall Water Splitting" Applied Sciences 9, no. 1: 55. https://doi.org/10.3390/app9010055

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop