Next Article in Journal
Improvement in Slag Resistance of No-Cement Refractory Castables by Matrix Design
Previous Article in Journal
Corrosion of MgO-C with Magnesium Aluminate Spinel Addition in A Steel Casting Simulator
Previous Article in Special Issue
Hydrothermal Corrosion Behaviors of Constituent Materials of SiC/SiC Composites for LWR Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Activity for Hydrogen Evolution of Heteroatom-Doped SrTiO3 Prepared Using a Graphitic-Carbon Nitride Nanosheet

Department of Applied Chemistry, Faculty of Engineering, Sanyo-Onoda City University, 1-1-1 Daigakudori, Sanyo-Onoda 756-0884, Yamaguchi, Japan
*
Author to whom correspondence should be addressed.
Ceramics 2020, 3(1), 22-30; https://doi.org/10.3390/ceramics3010003
Submission received: 19 October 2019 / Revised: 17 December 2019 / Accepted: 3 January 2020 / Published: 11 January 2020
(This article belongs to the Special Issue Physical Properties of Metals/Metal-Base Materials)

Abstract

:
We developed a novel method to synthesize a visible-light-responsible photocatalyst from a composite of SrTiO3 and a graphitic carbon nitride (g-C3N4) nanosheet. Heteroatoms were successfully doped into a lattice of SrTiO3 by mild calcination of a composite that the g-C3N4 nanosheet adsorbed on to the SrTiO3 surface. The absorption edge in the UV-Vis absorption spectrum of the doped sample was shifted to a longer wavelength region. The photocatalytic activity of the doped sample under UV light irradiation was higher than those of both pristine SrTiO3 and the g-C3N4 nanosheet, suggesting that the photocatalytic property of SrTiO3 was enhanced by doping. The doped sample showed photocatalytic activity under visible light irradiation (>420 nm), which was enhanced by Pt loading.

1. Introduction

Hydrogen production by photocatalytic solar water splitting has received much attention as the next generation clean energy source because CO2 emissions are minimal during electric power generation. Over the past few decades, a number of photocatalysts have been reported to achieve one-step overall water-splitting UV irradiation under visible light irradiation [1,2,3,4,5,6,7,8,9,10]. SrTiO3 and TiO2 have been widely used for water splitting photocatalysis under UV irradiation because of their non-toxicity and low cost [11,12,13,14,15]. However, these photocatalysts cannot operate under visible light irradiation because of their wide band gap. Many researchers have therefore developed visible-light-driven photocatalysts by band modification [16,17,18,19,20].
Doping, which includes replacing the host material with a foreign element at a crystal lattice point, is widely used to prepare visible-light-driven photocatalysts. Kudo et al. reported that co-doping of Cr3+-Ta5+ and Cr3+-Sb5+, and doping of Rh cations were effective, with doped SrTiO3 showing high photocatalytic activity for H2 evolution under visible light irradiation in the presence of a sacrificial agent [16,17]. The donor level was formed between the conduction band and the valence band. The transition from the donor level to the conduction bands of SrTiO3 and TiO2 responds to visible light absorption. For other band-modification methods in the oxide-based photocatalyst, sulfurization and nitridation have also been suggested. These processes were carried out by high-temperature calcination of an oxide-based photocatalyst in the flowing of H2S and NH3. The new bands based on S3p and/or N2p orbitals formed above O2p valence bands and led to a decrease in the band gap. Ohno et al. reported TiO1−xSx prepared by calcination of TiO2 in the flowing of H2S showed photocatalytic activity under visible light irradiation [18]. Domen et al., have reported on oxynitride- and oxysulfide-based photocatalysts for water splitting such as TaON, ATaO2N (A = Ca, Sr, Ba), (Ga1−xZnx)(N1−xOx) solid solution and Ln2Ti2S2O5 (Ln = Pr, Nd, Sm, Gd, Tb, Dy, Ho and Er) [21,22,23]. In particular, an Rh2−yCryO3-loaded (Ga1−xZnx)(N1−xOx) solid solution synthesized by nitriding a mixture Ga2O3 and ZnO has been shown to achieve overall water splitting under visible light irradiation without noticeable degradation. The quantum efficiency of this photocatalyst for overall water splitting reached ca. 2.5% at 420–440 nm. However, doping and sulfurization/nitridation require high-temperature calcination and the use of toxic reactants, respectively. From the perspectives of energy consumption suppression and reducing environmental load, the development of a more environmentally friendly process is required.
In the present study, we investigated facile doping under mild conditions in order to obtain a visible-light-driven photocatalyst. Graphitic C3N4 (g-C3N4) nanosheets were used as a doping agent. The bulk g-C3N4, an organic, metal-free polymeric, layered semiconductor has recently attracted attention as a visible-light-responsible photocatalyst [9,24,25,26,27,28], and its photocatalytic property was first reported by Wang et al. [29]. Moreover, the composite photocatalysts using g-C3N4 were reported and show high photocatalytic activity [30,31,32,33]. The layered g-C3N4 can be exfoliated into 2D thinner nanosheets. Although many investigations on the photocatalysis using bulk g-C3N4 have been carried out, there are few studies using g-C3N4 nanosheets. Therefore, we focused on the synthesis of a photocatalyst using g-C3N4 nanosheets. The 2D g-C3N4 nanosheet will strongly interact with other oxide-based semiconductor surfaces due to its high surface energy. By using a strong interaction of 2D g-C3N4, doping was carried out by the calcination of g-C3N4 nanosheet-adsorbed SrTiO3 nanoparticles in a nitrogen atmosphere. We optimized the amount of dopant by changing the volume of g-C3N4 nanosheet dispersion in the adsorption process. The photocatalytic hydrogen evolution in the presence of a sacrificial agent under UV and visible-light irradiation was studied. Moreover, we investigated the effect of co-catalyst loading on the photocatalytic activity under visible light irradiation.

2. Materials and Methods

2.1. Chemicals and Materials

Urea (99%), HNO3 (70%), and HCl (37%) were purchased from FUJIFILM Wako Pure Chemical Co. Ltd. (Osaka, Japan). These chemicals were used without purification. SrTiO3 nanoparticles were supplied by Toda Kogyo Co., Ltd., and calcined at 600 °C for 2 h before use.

2.2. Synthesis of Photocatalyst

A carbon nitride nanosheet was prepared following a method given in a previous study [34]. Urea was calcined at 600 °C in a crucible with a cover (heating rate: 5 °C/min) in order to obtain the bulk g-C3N4. The bulk C3N4 powder was washed with 0.1 M HNO3 and distilled water, followed by drying at 70 °C for 12 h. The powder was dispersed into a 15% HCl aqueous solution (150 mL). The resulting suspension was sonicated for 1 h and was then magnetically stirred for 24 h. The suspension was transferred to Teflon-lined stainless steel autoclaves and heated to 110 °C for 5 h. The obtained suspension was pump filtrated and washed at least five times to remove HCl and agglomerate the C3N4. To achieve g-C3N4 nanosheet dispersion, the wet product was dispersed into 100 mL distilled water.
The calcined SrTiO3 (0.3 g) was dispersed into 100 mL distilled water, followed by sonification for 10 min. The g-C3N4 nanosheet dispersion was dropped into the SrTiO3 dispersion under vigorous stirring. The amount of the dropped g-C3N4 nanosheet dispersion was varied in the range of 10 to 30 mL. The resulting suspension was stirred at room temperature for 1 h, and left without stirring for 3 h. The obtained sedimentation was centrifugated and dried at 110 °C overnight. The resulting powder was calcined at 600 °C for 5 h in flowing N2. The doped sample was abbreviated as SrTiO3-Dx (x is volume of added g-C3N4 nanosheet dispersion).

2.3. Characterization

The crystal structures of the samples were determined using an X-ray diffraction (XRD) diffractometer (Rigaku, SmartLab, Tokyo, Japan) with monochromated Cu Kα radiation (40 kV, 30 mA). X-ray fluorescence (XRF; Horiba, XGT-7200, Tokyo Japan) analysis was used to obtain the chemical composition of the samples. For the samples, the microstructures were observed and local elemental analysis was conducted using field emission scanning electron microscopy (FE-SEM) with energy-dispersive X-ray analysis (EDX, Hitachi S-4800, Tokyo, Japan). Diffuse reflectance absorption spectra were recorded using a UV-Vis spectrometer (Shimadzu, UV-3100PC, Kyoto, Japan) to determine the optical band gap energy.

2.4. Photocatalytic Reaction

The photocatalytic reaction was conducted in an external-irradiation quartz cell connected to a closed gas-circulating system. The photocatalyst (0.1 g) was suspended in 50% methanol aqueous solution (150 mL) in the cell. The rate of H2 evolution was determined via gas chromatography (Shimadzu GC-8A, TCD, Ar carrier, MS-5A) under irradiation from a 500 W Xe lamp (light density at 420 nm: 0.23 W/cm2, illuminated area: 78.5 cm2, distance between light source and cell: 10 cm). The light wavelength was controlled using a cut-off filter (Sigma Koki, Saitama, Japan). Additionally, the photocatalytic reaction was carried out after Pt photodeposition.

3. Results and Discussion

3.1. Characterization of Doped SrTiO3

A nanosheet will have an unstable surface compared to bulk material because of its high surface energy. Therefore, a g-C3N4 nanosheet will strongly adsorb SrTiO3 nanoparticles to form the precursor material for a doped sample. The doped sample can be synthesized by a solid state reaction between the g-C3N4 nanosheet and SrTiO3 owing to strong interaction. To optimize the amount of adsorbed g-C3N4 nanosheet in the precursor material, we synthesized the doped sample by changing the volume of g-C3N4 nanosheet dispersion. The crystal structures of the as-prepared samples were examined via XRD. Figure 1 shows the XRD patterns of pristine SrTiO3, g-C3N4 nanosheet, and the doped samples. The XRD pattern of the pure g-C3N4 nanosheet exhibited two diffraction peaks at 13.2° and 27.6°, which corresponded to the (100) and (002) diffraction planes, respectively [25,29,35]. These were attributed to the characteristic in-plane and inter-planar stacking peaks of the aromatic system in graphite-like carbon nitride, respectively [29]. The pristine SrTiO3 exhibits five distinct peaks at 22.66°, 32.27°, 39.79°, 46.28°, 52.11°, and 57.53°, which can be attributed to the (100), (110), (111), (200), (210), and (211) crystal planes, respectively. The doped samples were not observed as having the peaks owing to the g-C3N4; only the peaks ascribed to SrTiO3. It is suggested that this sample was not a simple composite of g-C3N4 and SrTiO3 nanoparticles. However, the diffraction peaks at 32.34° for the (110) planes of the doped samples (SrTiO3-D15) exhibited a slightly higher shift than that of pristine SrTiO3 (32.27°), which indicated that nitrogen or carbon was doped into SrTiO3, resulting in lattice distortion. The shift was also observed on other doped samples (Table S1).
The morphologies of the doped samples were examined via FE-SEM (Figure 2). As shown in Figure 2a, uniform spherical nanoparticles with grain sizes of approximately 30 nm could be observed. The g-C3N4 nanosheet particles (Figure 2b) had a thin lammellar shape and obvious particle aggregation was observed. Nanosheet aggregation may occur during the drying process. In the FE-SEM image (Figure 2c) of the calcination product from the SrTiO3/g-C3N4 composite, g-C3N4 nanosheet particles were not observed, which made it difficult to differentiate pristine SrTiO3 from the doped sample. From the energy dispersed X-ray spectroscopy (EDX) analysis results for pristine SrTiO3 and SrTiO3-D15 (Figure 3), it can be seen that the content of the C element in the doped sample is notably higher than that of pristine SrTiO3. In addition, the presence of the N element was observed in the doped sample, indicating that C and N were successfully doped into the lattice of SrTiO3. We attempted to perform an X-ray photoelectron spectroscopy (XPS) analysis of the doped samples to elucidate the chemical state of C and/or N. However, we could not sufficiently obtain the XPS spectra because of low concentration.
The diffuse reflectance absorption spectra were used to study the optical properties of the doped samples. Figure 4 shows the diffuse reflectance UV-Vis absorption spectra of the doped samples. The doped samples exhibited a longer light absorption edge compared to pristine SrTiO3, indicating that C or N doping was an efficient alternative for expanding light absorption by SrTiO3. Moreover, an obvious red shift was observed in the absorption edge of the doped samples. We also observed color change of samples from white to light yellow after doping (Figure S1). Figure 4b shows the Tauc plots obtained from the UV-vis absorption spectra by transformation based on the Kubelka–Munk function versus the energy of light. The transformation was conducted according to the formula: (αhν)2 = A(Eg), where α, ν, A, and Eg are the absorption coefficient, light frequency, proportionality constant, and band gap, respectively [36]. As shown in Table 1, the band gaps of the doped samples were smaller than that of pristine SrTiO3 and decreased with increasing g-C3N4 nanosheet dispersion during the preparation of the precursor to the doped samples. It is suggested that the g-C3N4 nanosheet can directly contribute to doping.

3.2. The Photocatalytic Property of Doped SrTiO3

At first, the photocatalytic activities of the doped samples were evaluated by considering the hydrogen evolution in the presence of methanol as a sacrificial agent. Figure 5 shows the hydrogen evolution rate for the doped samples synthesized using different amounts of g-C3N4 nanosheet dispersion. From the results of the UV-vis absorption analysis, it can be seen that the amount of dopant increased with increasing g-C3N4 nanosheet dispersion during the preparation of the precursor materials. The hydrogen evolution rate increased with increasing dopant concentration; however, excess increase of dopant makes the recombination center between the photogenerated electrons and holes increase [17]. SrTiO3-D15 prepared using 15 mL C3N4 nanosheet dispersion, therefore, showed maximum photocatalytic activity. The hydrogen evolution rate of SrTiO3-D15 was higher than those of pristine g-C3N4 and SrTiO3 nanoparticles (Figure 6), suggesting that the photocatalytic activity of SrTiO3-D15 increased due to a synergistic effect between SrTiO3 and the dopant. To enhance the photocatalytic activity, Pt was loaded onto the SrTiO3-D15 as a co-catalyst. The Pt loading sample showed photocatalytic activity under visible light irradiation (λ > 420 nm) (Figure S2). The 5-h photocatalytic test under UV irradiation using the Pt loaded sample could be repeated two times without noticeable deactivation (Figure S3). Figure 7 shows the effect of photocatalytic activity on Pt loading. The photocatalytic H2 evolution rate showed maximum value at 0.2 wt% loading. The photo-formed electron transfers to the Pt nanoparticle owing to a high work function, and proton reduction can efficiently take place over the Pt nanoparticles.

4. Conclusions

We demonstrate a new doping method using g-C3N4 nanosheets. C3N4 nanosheets reacted with SrTiO3 nanoparticles via a strong adsorption interaction to form the doped sample. The doped sample shows light absorption in a longer wavelength region (λ > 420 nm) compared to SrTiO3. The photocatalytic activity of the doped sample under UV irradiation was higher than those of both pristine g-C3N4 and SrTiO3 nanoparticles owing to their synergistic effect, and was enhanced by Pt loading. Moreover, the doped sample can operate under visible light irradiation. This doping method can applied to a broad range of oxide-based photocatalysts. We investigated the photocatalytic properties of various doped samples prepared by this method. However, the photocatalytic H2 evolution reaction in the doped sample cannot proceed without a sacrificial agent under visible light irradiation. Therefore, the overall water splitting under visible light irradiation should be investigated in the Z-scheme reaction system formed by combination with O2 evolution photocatalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/2571-6131/3/1/3/s1, Figure S1: The photograph of samples, Figure S2: Time course of H2 evolution reaction over 0.2 wt% Pt loaded SrTiO3-D15, Figure S3: The repeated photocatalytic test, Table S1: The peak position and d-spacing in XRD of doped samples.

Author Contributions

Conceptualization, K.I.; Methodology, K.I.; Formal Analysis, Y.Y. and M.S.; Investigation, Y.Y.; Resources, K.I.; Writing-Original Draft Preparation, K.I.; Writing-Review & Editing, K.I.; Supervision, K.I. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Regional Innovation Strategy Support Program 2014 of the Ministry of Education, Culture, Sports, Science and Technology, Japan (MEXT).

Acknowledgments

We thank Honmyo and Kawaguchi (Toda Kogyo Co. Ltd.) for supplying SrTiO3 powder and some advice.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2008, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef] [PubMed]
  3. Hisatomi, T.; Kubo, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef] [PubMed]
  4. Osterloh, F.E. Inorganic materials as catalysts for photochemical splitting of water. Chem. Mater. 2008, 20, 35–54. [Google Scholar] [CrossRef]
  5. 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]
  6. Abe, R. Recent progress on photocatalytic and photoelectrochemical water splitting under visible light irradiation. J. Photochem. Photobiol. C Photochem. Rev. 2010, 11, 179–209. [Google Scholar] [CrossRef]
  7. Maeda, K. Photocatalytic water splitting using semiconductor particles: History and recent developments. J. Photochem. Photobiol. C Photochem. Rev. 2011, 12, 237–268. [Google Scholar] [CrossRef]
  8. Maeda, K. Z-scheme water splitting using two different semiconductor photocatalysts. ACS Catal. 2013, 3, 1486–1503. [Google Scholar] [CrossRef]
  9. Wang, Y.; Wang, X.; Antonietti, M. Polymeric graphitic carbon nitride as a heterogeneous organocatalyst: From photochemistry to multipurpose catalysis to sustainable chemistry. Angew. Chem. Int. Ed. 2012, 51, 68–89. [Google Scholar] [CrossRef]
  10. Zhang, N.; Zhang, Y.; Xu, Y.-J. Recent progress on graphene-based photocatalysts: Current status and future perspectives. Nanoscale 2012, 4, 5792–5813. [Google Scholar] [CrossRef]
  11. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  12. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  13. Ni, M.; Leung, M.K.H.; Leung, D.Y.C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  14. Leung, D.Y.C.; Fu, X.L.; Wang, C.F.; Ni, M.; Leung, M.K.H.; Wang, X.X.; Fu, X.Z. Hydrogen production over titania-based photocatalysts. ChemSusChem 2010, 3, 681–694. [Google Scholar] [CrossRef]
  15. Qi, K.; Cheng, B.; Yu, J.; Ho, W. A review on TiO2-based Z-scheme photocatalysts. Chin. J. Catal. 2017, 38, 1936–1955. [Google Scholar] [CrossRef]
  16. Ishii, T.; Kato, H.; Kudo, A. H2 evolution from an aqueous methanol solution on SrTiO3 photocatalysts codoped with chromium and tantalum ions under visible light irradiation. J. Photochem. Photobiol. A Chem. 2004, 163, 181–186. [Google Scholar] [CrossRef]
  17. 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]
  18. Ohno, T.; Mitsui, T.; Matsumura, M. Photocatalytic activity of S-doped TiO2 photocatalyst under visible light. Chem. Lett. 2003, 32, 364–365. [Google Scholar] [CrossRef] [Green Version]
  19. 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]
  20. Irie, H.; Watanabe, Y.; Hashimoto, K. Carbon-doped anatase TiO2 powder as a visible-light sensitive photocatalyst. Chem. Lett. 2003, 32, 772–773. [Google Scholar] [CrossRef]
  21. Maeda, K.; Domen, K. New non-oxide photocatalysts designed for overall water splitting under visible light. J. Phys. Chem. C 2007, 111, 7851–7861. [Google Scholar] [CrossRef]
  22. Maeda, K.; Takata, T.; Hara, M.; Saito, N.; Inoue, Y.; Kobayashi, H.; Domen, K. GaN:ZnO solid solution as a photocatalyst for visible-light-driven overall water splitting. J. Am. Chem. Soc. 2005, 127, 8286–8287. [Google Scholar] [CrossRef]
  23. Maeda, K.; Teramura, K.; Takata, T.; Hara, M.; Saito, N.; Toda, K.; Inoue, Y.; Kobayashi, H.; Domen, K. Overall water splitting on (Ga1-xZnx)(N1-xOx) solid solution photocatalyst: Relationship between physical properties and photocatalytic activity. J. Phys. Chem. B 2005, 109, 20504–20510. [Google Scholar] [CrossRef] [PubMed]
  24. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic carbon nitride (g-C3N4)-based photocatalysts for artificial photosynthesis and environmental remediation: Are we a step closer to achieving sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  25. Sun, S.; Sun, M.; Fang, Y.; Wang, Y.; Wang, H. One-step in situ calcination synthesis of g-C3N4/N-TiO2 hybrids with enhanced photoactivity. RSC Adv. 2016, 6, 13063–13071. [Google Scholar] [CrossRef]
  26. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [CrossRef]
  27. Wang, X.; Maeda, K.; Chen, X.; Takanabe, K.; Domen, K.; Hou, Y.; Fu, X.; Antonietti, M. Polymer semiconductors for artificial photosynthesis: Hydrogen evolution by mesoporous graphitic carbon nitride with visible light. J. Am. Chem. Soc. 2009, 131, 1680–1681. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Liu, J.; Wu, G.; Chen, W. Porous graphitic carbon nitride synthesizedviadirect polymerization of ureafor efficient sunlight-driven photocatalytic hydrogen production. Nanoscale 2012, 4, 5300–5303. [Google Scholar] [CrossRef]
  29. Wang, X.C.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–82. [Google Scholar] [CrossRef]
  30. Xu, X.; Liu, G.; Randorn, C.; Irvine, J.T.S. g-C3N4 coated SrTiO3 as an efficient photocatalyst for H2 production in aqueous solution under visible light irradiation. Int. J. Hydrogen Energy 2011, 36, 13501–13507. [Google Scholar] [CrossRef]
  31. Li, J.; Zhang, M.; Li, Q.; Yang, J. Enhanced visible light activity on direct contact Z-scheme g-C3N4-TiO2 photocatalyst. Appl. Surf. Sci. 2017, 391, 184–193. [Google Scholar] [CrossRef]
  32. Chen, J.; Yang, Q.; Zhong, J.; Li, J.; Hu, C.; Deng, Z.; Duan, R. In-Situ construction of direct Z-scheme Bi2WO6/g-C3N4 composites with remarkably promoted solar-driven photocatalytic activity. Mater. Chem. Phys. 2018, 217, 207–215. [Google Scholar] [CrossRef]
  33. Opoku, F.; Govender, K.K.; Sittert, C.G.C.E.; Governder, P.P. Insights into the photocatalytic mechanism of mediator-free direct Z-scheme g-C3N4/Bi2MoO6(010) and g-C3N4/Bi2WO6(010) heterostructures: A hybrid density functional theory study. Appl. Surf. Sci. 2018, 427, 487–498. [Google Scholar] [CrossRef]
  34. Yang, Y.; Geng, L.; Guo, Y.; Meng, J.; Guo, Y. Easy dispersion and excellent visible-light photocatalytic activity of the ultrathin urea-derived g-C3N4 nanosheets. Appl. Surf. Sci. 2017, 425, 535–546. [Google Scholar] [CrossRef]
  35. He, Y.M.; Wang, Y.; Zhang, L.H.; Teng, B.T.; Fan, M.H. High-efficiency conversion of CO2 to fuel over ZnO/g-C3N4 photocatalyst. Appl. Catal. B Environ. 2015, 168, 1–8. [Google Scholar]
  36. Sun, M.X.; Fang, Y.L.; Wang, Y.; Sun, S.F.; He, J.; Yan, Z. Synthesis of Cu2O/graphene/rutile TiO2 nanorod ternary composites with enhanced photocatalytic activity. J. Alloys Compd. 2015, 650, 520–527. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of (a) pristine SrTiO3, (b) SrTiO3-D10, (c) SrTiO3-D15, (d) SrTiO3-D20, (e) SrTiO3-D25, (f) SrTiO3-D30, and (g) C3N4.
Figure 1. X-ray diffraction (XRD) patterns of (a) pristine SrTiO3, (b) SrTiO3-D10, (c) SrTiO3-D15, (d) SrTiO3-D20, (e) SrTiO3-D25, (f) SrTiO3-D30, and (g) C3N4.
Ceramics 03 00003 g001
Figure 2. Scanning electron microscopy (SEM) images of (a) pristine SrTiO3, (b) pristine g-C3N4, and (c) SrTiO3-D15.
Figure 2. Scanning electron microscopy (SEM) images of (a) pristine SrTiO3, (b) pristine g-C3N4, and (c) SrTiO3-D15.
Ceramics 03 00003 g002
Figure 3. Energy-dispersive X-ray (EDX) spectra of (a) pristine SrTiO3 and (b) SrTiO3-D15.
Figure 3. Energy-dispersive X-ray (EDX) spectra of (a) pristine SrTiO3 and (b) SrTiO3-D15.
Ceramics 03 00003 g003
Figure 4. Diffuse reflectance UV-Vis absorption spectra (a) of SrTiO3, C3N4, and the doped samples, and their Tauc plots (b).
Figure 4. Diffuse reflectance UV-Vis absorption spectra (a) of SrTiO3, C3N4, and the doped samples, and their Tauc plots (b).
Ceramics 03 00003 g004
Figure 5. Photocatalytic H2 evolution over the doped SrTiO3 synthesized using different amounts of g-C3N4 nanosheet dispersion under UV light irradiation. The photocatalytic reaction was performed in an external-irradiation quartz cell under irradiation from a 500 W Xe lamp, 150 mL of a 50% methanol solution, and 0.1 g of a photocatalyst.
Figure 5. Photocatalytic H2 evolution over the doped SrTiO3 synthesized using different amounts of g-C3N4 nanosheet dispersion under UV light irradiation. The photocatalytic reaction was performed in an external-irradiation quartz cell under irradiation from a 500 W Xe lamp, 150 mL of a 50% methanol solution, and 0.1 g of a photocatalyst.
Ceramics 03 00003 g005
Figure 6. Comparison of photocatalytic activity under UV light irradiation between SrTiO3-D15 and pristine SrTiO3 and C3N4.
Figure 6. Comparison of photocatalytic activity under UV light irradiation between SrTiO3-D15 and pristine SrTiO3 and C3N4.
Ceramics 03 00003 g006
Figure 7. The effect of Pt loading on the photocatalytic H2 evolution rate over SrTiO3-D15 under visible light irradiation. The photocatalytic reaction was performed in an external-irradiation quartz cell under irradiation from a 500 W Xe lamp with a cut-off filter (λ > 420 nm), 150 mL of 50% methanol solution, and 0.1 g of photocatalyst.
Figure 7. The effect of Pt loading on the photocatalytic H2 evolution rate over SrTiO3-D15 under visible light irradiation. The photocatalytic reaction was performed in an external-irradiation quartz cell under irradiation from a 500 W Xe lamp with a cut-off filter (λ > 420 nm), 150 mL of 50% methanol solution, and 0.1 g of photocatalyst.
Ceramics 03 00003 g007
Table 1. Band gap energy of the doped samples, SrTiO3, and C3N4 nanosheet.
Table 1. Band gap energy of the doped samples, SrTiO3, and C3N4 nanosheet.
SampleBand Gap/eV
SrTiO33.21
C3N4 nanosheet2.78
SrTiO3-D103.17
SrTiO3-D153.02
SrTiO3-D202.81
SrTiO3-D252.74
SrTiO3-D302.42

Share and Cite

MDPI and ACS Style

Ikeue, K.; Yamamoto, Y.; Suzuki, M. Photocatalytic Activity for Hydrogen Evolution of Heteroatom-Doped SrTiO3 Prepared Using a Graphitic-Carbon Nitride Nanosheet. Ceramics 2020, 3, 22-30. https://doi.org/10.3390/ceramics3010003

AMA Style

Ikeue K, Yamamoto Y, Suzuki M. Photocatalytic Activity for Hydrogen Evolution of Heteroatom-Doped SrTiO3 Prepared Using a Graphitic-Carbon Nitride Nanosheet. Ceramics. 2020; 3(1):22-30. https://doi.org/10.3390/ceramics3010003

Chicago/Turabian Style

Ikeue, Keita, Yuta Yamamoto, and Masashige Suzuki. 2020. "Photocatalytic Activity for Hydrogen Evolution of Heteroatom-Doped SrTiO3 Prepared Using a Graphitic-Carbon Nitride Nanosheet" Ceramics 3, no. 1: 22-30. https://doi.org/10.3390/ceramics3010003

Article Metrics

Back to TopTop