Next Article in Journal
Quantitative Risk Assessment Methodology for Hydrogen Tank Rupture in a Tunnel Fire
Next Article in Special Issue
Modeling of Liquid Hydrogen Tank Cooled with Para-Orthohydrogen Conversion
Previous Article in Journal
Simulation of Biogas Conversion Using Porous Solid Oxide Electrochemical Cells: Virtual Prototyping
Previous Article in Special Issue
First Hydrogenation of TiFe with Addition of 20 wt.% Ti
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aluminum Cation Doping in Ruddlesden-Popper Sr2TiO4 Enables High-Performance Photocatalytic Hydrogen Evolution

1
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, China
2
WA School of Mines: Minerals, Energy and Chemical Engineering, Curtin University, Perth, WA 6845, Australia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Hydrogen 2022, 3(4), 501-511; https://doi.org/10.3390/hydrogen3040032
Submission received: 6 November 2022 / Revised: 24 November 2022 / Accepted: 29 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Hydrogen Energy Technologies)

Abstract

:
Hydrogen (H2) is regarded as a promising and renewable energy carrier to achieve a sustainable future. Among the various H2 production routes, photocatalytic water splitting has received particular interest; it strongly relies on the optical and structural properties of photocatalysts such as their sunlight absorption capabilities, carrier transport properties, and amount of oxygen vacancy. Perovskite oxides have been widely investigated as photocatalysts for photocatalytic water splitting to produce H2 because of their distinct optical properties, tunable band gaps and excellent compositional/structural flexibility. Herein, an aluminum cation (Al3+) doping strategy is developed to enhance the photocatalytic performance of Ruddlesden-Popper (RP) Sr2TiO4 perovskite oxides for photocatalytic H2 production. After optimizing the Al3+ substitution concentration, Sr2Ti0.9Al0.1O4 exhibits a superior H2 evolution rate of 331 μmol h−1 g−1, which is ~3 times better than that of Sr2TiO4 under full-range light illumination, due to its enhanced light harvesting capabilities, facilitated charge transfer, and tailored band structure. This work presents a simple and useful Al3+ cation doping strategy to boost the photocatalytic performance of RP-phase perovskites for solar water splitting.

1. Introduction

The energy crisis, greenhouse effect, and environmental issues caused by the overuse of fossil fuels have received increased attention recently. Thus, it is crucial to develop renewable and sustainable energy resources (e.g., wind, solar) and relevant technologies [1,2,3]. Solar energy is regarded as the most attractive renewable energy due to its abundance and clean nature; it can also be utilized by various routes [4,5,6,7,8,9]. Among them, photovoltaics and photocatalysis for the conversion of solar energy to fuels and electric power have received considerable attention for their efficient utilization of sunlight energy [10,11,12,13,14,15]. Principally, semiconductors used in photovoltaics directly transform solar energy to electricity, while semiconductors in photocatalysis enable the conversion from sunlight energy to chemical energy by means of photoinduced charge carriers for wastewater treatment and photocatalytic water splitting [10,12,13,16,17,18,19,20]. More importantly, photocatalysis is crucial in wastewater treatment and water splitting for hydrogen (H2) production to address worldwide energy demands and environmental pollution [21,22,23,24]. H2 is regarded as a highly promising energy carrier because clean water is only combustion product, while photocatalytic water splitting is considered one of the most promising methods of H2 generation, and the photocatalytic efficiency is determined by the photocatalysts [25,26,27,28]. Consequently, it is crucial to develop cost-effective and earth-abundant photocatalysts with advanced water-splitting abilities [29,30,31].
At present, various types of photocatalysts have been designed and developed for photocatalytic water splitting, such as metal oxides, metal sulfides, carbon nitrides, etc. [32,33,34,35]. Simple metal oxides (e.g., TiO2 and ZnO) have been extensively investigated in relation to photocatalytic water splitting, showing relatively high photocatalytic activity [36,37]. Nevertheless, the performance of the reported simple oxide-based photocatalysts is significantly limited by their fixed atomic environment, wide band gap, and inferior stability, which limit their photocatalytic efficiency [38,39]. As a result, complex metal oxides are thought to be more applicable as high-performance photocatalysts because of their rich and tunable atomic environments, which exhibit easily tailored optical, chemical, and structural properties [40,41,42,43,44]. In particular, perovskite oxides, a significant class of complex oxides, have received increasing attention in relation to photocatalytic H2 evolution reaction (HER) due to their easily tailored band gaps/structures, cost effectiveness, compositional/structural flexibility, and superior stability [12,40,41,42]. Additionally, the physical, optical, and electronic properties of conventional perovskites oxides (ABO3) with three-dimensional structures can be effectively tuned by reducing the dimensions [12,40]. Notably, two-dimensional Ruddlesden-Popper (RP) perovskite oxides have gained increasing attention on account of their superior photocatalytic performance and unique chemical, structural, electronic, and optical features, as well as the fact that they play a promoting role in enhancing the separation and transportation of photo-induced charge carriers [45,46]. Sr2TiO4, a typical RP-phase perovskite, has been employed as a photocatalyst for water splitting and dye degradation, but exhibited inferior visible-light-driven photocatalytic performance because of the wide band gap of ~3.5 eV [46].
Functional doping is reported to be a useful and simple strategy to improve the photocatalytic performance of RP-phase Sr2TiO4 photocatalysts for solar water splitting by tuning the band structure/gap, charge transfer capability, and oxygen vacancy concentrations [47,48,49]. For instance, Xiao et al. reported a synergistic strategy by combining silver (Ag+) cation doping and reducing pretreatment to enhance the photocatalytic activity of Sr2TiO4 [47]. The optimized r-Ag0.05Sr1.95TiO4, with a suitable amount of Ag+ cation substitution and reduced pretreatment, displayed a superior H2 generation rate of 541 μmol h−1 g−1 under visible light illumination due to the enlarged specific surface area, the optimized band gap value, the enhanced light-harvesting capability, the presence of more surface oxygen vacancies, and reduced carrier recombination. In another work, Han et al. employed a fluorine (F) anion doping strategy to boost the photocatalytic performance of Sr2TiO4 for H2 production [48]. Sr2TiO3.97F0.03 with an optimized amount of F doping displayed outstanding H2 production performance (282 μmol h−1 g−1) under full-range light irradiation, 44% higher than that of pristine Sr2TiO4, because of the increased amount of surface defect, the more strongly negative position of the conduction band, the reduced carrier recombination and well-balanced band gap value, the specific surface area, and the grain size. Nevertheless, the above-mentioned Ag+ and F doped Sr2TiO4 suffer from low doping concentrations of foreign ions (≤5 mol %), which may limit the improvement of the photocatalytic performance of Sr2TiO4. Meanwhile, metal cation doping with higher oxidation states (e.g., Rh4+) into RP-phase Sr2TiO4 competitively consumed photo-generated electrons, and played a detrimental role in the photocatalytic activity [50,51]. To solve the above-mentioned problems, we report a simple and useful strategy to increase the photocatalytic HER performance of Sr2TiO4 using a doping aluminum (Al3+) cation with a lower oxidation state and a higher doping concentration in the B-site of RP-phase Sr2TiO4.
Herein, we report a simple and useful strategy to increase the photocatalytic HER performance of Sr2TiO4 by doping an Al3+ cation into the B-site with higher doping concentrations. After optimizing the Al3+ doping concentration, Sr2Ti0.9Al0.1O4 demonstrated an increased photocatalytic H2 production rate of 331 μmol h−1 g−1 under full-range light illumination (λ ≥ 250 nm), which was nearly three times better than that of Sr2TiO4, due to the reduced particle sizes, suppressed band gaps, stronger light harvesting capabilities, and improved separation and transfer capabilities of the photoinduced charge carriers. This study presents a new and efficient RP-type photocatalyst for high-efficiency water splitting, and provides insights that will enable the acceleration of large-scale applications of photocatalytic water splitting.

2. Materials and Methods

Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15) with various Al3+ doping concentrations was prepared by the typical solid-state reaction route, using SrCO3, TiO2, and Al2O3 with different molar ratios as the precursors; these were mixed well in ethanol by 60-min high-energy ball milling. After drying, the mixed precursors were further annealed in air at 1000 °C for 10 h to obtain the final photocatalysts. The phase structures, microstructures, and chemical states of various elements in the as-prepared photocatalysts were analyzed by X-ray diffraction (XRD, Bruker D8 Advance), field emission scanning electron microscopy (FESEM, Hitachi S4800), and with an XPS spectrometer (Thermo ESCALAB 250), respectively. The energy dispersive X-ray spectrometer (EDX) was used to investigate the elemental compositions of various samples. The Autosorb-iQ3 instrument was employed to measure the Brunauer–Emmett–Teller (BET) surface areas of various samples. The ultraviolet-visible (UV-vis) spectrophotometer (Lambda 750S) and FL 6500 photoluminescence (PL) spectrometer were employed to measure the diffuse reflectance spectra and fluorescence emission of the as-prepared photocatalysts, respectively. The photocatalytic performance of various photocatalysts was determined using a photocatalytic performance evaluation instrument (CEL-PAEMD8, China Education Au-light); detailed information regarding this process can be found in our previous work [48].
A three-electrode system was used to investigate the photoelectrochemical properties of photocatalysts in a 0.1 M Na3PO4/Na2HPO4 aqueous solution (pH = 7.9) under full-range light irradiation on a CHI760E workstation. The photocatalyst powder-based slurries were spin-coated on F doped tin oxide (FTO) glass to serve as the working electrode [48], while Ag/AgCl and Pt functioned as the reference and counter electrodes, respectively. Cyclic voltammetry of 50 cycles at a scan rate of 200 mV s−1 was used to activate the working electrode. After that, the linear sweep voltammetry (LSV) curves were obtained at a scan rate of 10 mV s−1 and a voltage range of −1.6–0 V. Nyquist plots of various photocatalysts were obtained using electrochemical impedance spectroscopy (EIS) at −1.2 V under light irradiation with frequencies in the range of 0.1–105 Hz, using the CHI760E workstation.

3. Results and Discussion

Generally, a high calcination temperature (≥1000 °C) is required for the synthesis of Sr2TiO4 using the solid-state reaction route, due to the high energy barrier of SrO layer intercalation into the SrTiO3 with an ABO3 structure [45]. Figure 1 shows the XRD patterns of various Sr2TiO4 samples with different Al3+ doping amounts calcined at 1000 °C; the main diffraction peaks of the four samples are consistent with the peaks of RP-type Sr2TiO4 (JCPDS No. 00-039-1471), suggesting that Al3+ doping had no obvious impact on the phase structure of Sr2TiO4. It was found that the main peaks (31° ≤ 2θ ≤ 33°) of the Sr2Ti1 − xAlxO4 samples exhibited shifts to higher 2θ values, as compared with the pristine Sr2TiO4 in Figure 1, implying that the Al3+ cation with a smaller ionic radius (0.535 Å) than that of Ti4+ (0.605 Å) and Ti3+ (0.67 Å) cations was successfully doped in the lattice of RP-phase Sr2TiO4 with a lattice contraction [52,53]. Nevertheless, when the doping ratio of the Al3+ cation was higher than 0.1, an Sr4Ti3O10 (JCPDS: No. 00-022-1444) impurity phase appeared in the Sr2Ti0.85Al0.15O4 sample, which may have a detrimental effect on photocatalytic activity [54].
The microstructures, particle sizes, and elemental distributions of Al3+-doped Sr2TiO4 were investigated using SEM and EDX elemental mapping, which showed a significant effect on the photocatalytic performance. SEM images and the corresponding grain size histograms are provided in Figure 2a–d. It seems that the particles of various Sr2TiO4-based photocatalysts exhibited irregular shapes, with particle size distributions in the range of 150–300 nm. Although the Al3+ doping efficiently decreased the particle sizes of the pristine Sr2TiO4 photocatalyst, the more serious particle agglomeration inevitably reduced the BET specific surface areas, as shown in Figures S1 and S2 and Table S1. Furthermore, the corresponding EDX elemental mapping image in Figure 3 illustrates the well-distributed Sr, Ti, and O elements in pristine Sr2TiO4 and Sr, Ti, Al, and O elements in Sr2Ti0.9Al0.1O4, further confirming that Al3+ cations were successfully doped in the lattice of Sr2TiO4.
The XPS technique was employed to investigate the influence of Al3+ doping on the chemical states of various elements in Sr2TiO4, with results shown in Figure 4 and Figure S2. As depicted in Figure 4a, all samples exhibited two main peaks at binding energy (BE) positions of around 458.0 and 464.0 eV in Ti 2p XPS spectra, corresponding to Ti 2p3/2 and Ti 2p1/2, respectively [55]. As compared with pristine Sr2TiO4, the Ti XPS peaks of the Al3+ doped samples shifted to the lower BE positions, demonstrating the formation of Ti3+ cations induced by the Al3+ doping. In addition, O 1s XPS spectra were mainly composed of four peaks at BEs of around 530.1, 531.1, 532.0, and 533.2 eV, representing the lattice oxygen, O 2 2 /O species, the surface-absorbed hydroxyl groups/oxygen (-OH/O2), and absorbed carbonates/water on the surface (CO32−/H2O), respectively [56]. In fact, the content of surface oxygen vacancy as a crucial factor in determining the photocatalytic performance was closely tied to the O 2 2 /O species [57]. It can be seen that the number of O 2 2 /O species for Sr2Ti0.9Al0.1O4 was much higher than those of the other four samples, implying the highest amount of surface oxygen vacancies. Based on the Al 2p XPS spectra in Figure 4c, the intensity of the typical peak for Al at a BE of around 74 eV was gradually increased with the increase in the Al3+ doping concentration from 0.05 to 0.10, suggesting that the Al3+ cations of suitable amounts were successfully doped into the lattice of Sr2TiO4.
Besides the phase structure and morphology, the light-harvesting capability of the photocatalyst is another decisive factor for photocatalytic HER activity [47]. UV-vis spectra were used to identify the impact of Al3+ substitution amounts on the sunlight-harvesting capabilities of Sr2TiO4. As depicted in Figure 5a, all samples exhibited strong light absorption in the ultraviolet (UV) region. Moreover, the increase in the Al3+ doping concentration from 0 to 0.10 in Sr2TiO4 gradually led to improved light absorption capabilities at higher wavelengths because of the reduced band gaps. In addition, Sr2Ti0.9Al0.1O4 exhibited a stronger absorbance intensity and light harvesting capability than Sr2Ti0.85Al0.15O4 at wavelengths larger than 375 nm. Although Sr2Ti0.85Al0.15O4 exhibited stronger absorbance intensity at wavelengths from 340–375 nm, the inferior photocatalytic performance of Sr2Ti0.85Al0.15O4, compared to that of Sr2Ti0.9Al0.1O4, was mainly attributed to the reduced specific surface area, wider band gap, larger charge transfer resistance and the existence of Sr4Ti3O10 impurity. The Kubelka–Munk transformation displayed in Figure 5b was employed to obtain the band gap energies (Eg) of various Sr2TiO4 photocatalysts. Sr2TiO4 displayed an Eg of 3.32 eV, agreeing well with the reported Eg for Sr2TiO4 in previous investigations [58]. For Al3+-doped Sr2TiO4, decreased Eg values of 3.20, 3.06, and 3.16 eV were observed for Sr2Ti0.95Al0.05O4, Sr2Ti0.9Al0.1O4, and Sr2Ti0.85Al0.15O4, respectively, as displayed in Figure 5b and Table S1; these results suggest that the Al3+ doping is effective in narrowing the band gap of Sr2TiO4. Notably, Sr2Ti0.9Al0.1O4 exhibited the lowest Eg with much-enhanced light-harvesting capability, which was beneficial for improving the photocatalytic activity.
It has been widely reported that the separation, transport, and recombination behaviors of photogenerated charge carriers play crucial roles in the photocatalytic activity of photocatalysts, which can be characterized using the PL technique [59]. The PL spectra of the Sr2TiO4 and various Al3+ doped Sr2TiO4 are displayed in Figure 5c. Under excitation at a wavelength of 380 nm, all of the investigated samples exhibited emission peaks at a wavelength range of 420–550 nm. As compared with Sr2TiO4, Sr2Ti0.95Al0.05O4 and Sr2Ti0.9Al0.1O4 showed lower PL emission peak intensities, indicating the facilitated separation and transport capabilities of the charge carriers. Nevertheless, Sr2Ti0.85Al0.15O4 with excessive Al3+ doping exhibited a higher PL emission peak intensity than that of pristine Sr2TiO4 because of the inhibited separation and transport of photoexcited carriers induced by the existence of the impurity phase in Sr2Ti0.85Al0.15O4.
The photocatalytic activities of the as-prepared Sr2Ti1 − xAlxO4 (x = 0, 0.05, 0.1, and 0.15) photocatalysts were evaluated by testing the H2 production amounts in a mixed Na2SO3/Na2S aqueous solution under full-range light irradiation, in which Na2SO3 and Na2S functioned as sacrificial agents to react with the photoinduced holes. Figure 6a displays steady and continuous H2 production during photocatalytic water splitting under full-range light illumination, and the H2 production rates of all samples are presented in Figure 6b. Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15) exhibited better photocatalytic HER activity than pristine Sr2TiO4, demonstrating the effectiveness of Al3+ doping in boosting the photocatalytic activity. Notably, Sr2Ti0.9Al0.1O4 exhibited the maximum H2 evolution rate (of 331 μmol h−1 g−1) among the three Al3+ doped samples, which was nearly three times higher than that of the undoped Sr2TiO4 (85 μmol h−1 g−1) due to the enhanced light-harvesting capabilities, suppressed carrier recombination, and reduced Eg value. In addition, Sr2Ti0.85Al0.15O4 with excessive amounts of Al3+ doping exhibited an inferior H2 generation rate to that of Sr2Ti0.9Al0.1O4 due to the inhibited separation and transport of photoinduced carriers and the large Eg value. As for the photocatalytic stability of HER, the Sr2Ti0.9Al0.1O4 photocatalyst exhibited steady and continuous H2 production for 15 h (Figure 7a). The amount of H2 produced with Sr2Ti0.9Al0.1O4 was measured to 82.6 μmol in the first 5 h, and the initial H2 production rate was 328 μmol h−1 g−1. In addition, the H2 evolution amount was increased to 219.8 μmol after 15 h of operation and the average H2 evolution rate was 292 μmol h−1 g−1, which was nearly 90% of the initial value after 15 h of continuous operation (Figure 7a). As compared with the sample before the photocatalytic reaction, the Sr2Ti0.9Al0.1O4 photocatalyst exhibited no obvious changes in the phase structure and particle sizes after the stability test under full-range light illumination, as depicted in Figure 7b,d,e. As shown in Figure 7c, the cycling performance of Sr2Ti0.9Al0.1O4 photocatalyst was determined by testing the HER rates in three cycles. The Sr2Ti0.9Al0.1O4 photocatalyst exhibited a high recyclability without any obvious degradation in the HER rates (~297 μmol h−1 g−1).
The LSV curves of various Sr2Ti1 − xAlxO4 (x = 0, 0.05, 0.1, and 0.15) photocatalysts were measured under full-range light illumination to further verify the enhanced photocatalytic activity of Sr2Ti0.9Al0.1O4. As depicted in Figure 8a, after the cyclic voltammetry activation, the photocatalysts showed similar onset potential for the cathodic current. In addition, Sr2Ti0.9Al0.1O4 exhibited larger current densities than those of Sr2TiO4, Sr2Ti0.95Al0.05O4, and Sr2Ti0.85Al0.15O4, demonstrating the superior photocatalytic HER activity of Sr2Ti0.9Al0.1O4. Based on the EIS results in Figure 8b, the Sr2Ti0.9Al0.1O4 photocatalysts exhibited the smallest charge transfer resistance among the four investigated samples due to the suppressed recombination of photogenerated carriers, as evidenced by the PL results, which were consistent with the photocatalytic HER activities of various photocatalysts under irradiation.
Mott–Schottky curves were further tested to measure the energy band positions, including the conduction bands (CB) and the valence bands (VB) of the Sr2TiO4 and Sr2Ti0.9Al0.1O4 photocatalysts. As shown in Figure 8c, the flat band potentials of the photocatalysts can be acquired by extrapolating the Mott–Schottky curves to the X-axis. It can be observed that the flat band positions of Sr2TiO4 and Sr2Ti0.9Al0.1O4 were measured to be −0.37 and −0.68 eV (vs. Ag/AgCl), respectively. Thus, the CB edges of Sr2TiO4 and Sr2Ti0.9Al0.1O4 were calculated to be −0.17 and −0.48 eV (vs. the normal hydrogen electrode, NHE), respectively. Based on the Eg values measured by the Kubelka–Munk transformation, the VB edges of Sr2TiO4 and Sr2Ti0.9Al0.1O4 were calculated to be 3.15 and 2.58 eV (vs. NHE), respectively. Based on the schematic diagrams of the band structures of Sr2TiO4 and Sr2Ti0.9Al0.1O4, as shown in Figure 8d, it seems that the Al3+ doping effectively regulated the band positions of Sr2TiO4. The CB edge of Sr2Ti0.9Al0.1O4, which was more strongly negative than that of the pristine Sr2TiO4, enabled a much higher driving force of the photoexcited electrons for water reduction, and led to a significant enhancement of the photocatalytic HER performance of Sr2TiO4.

4. Conclusions

In conclusion, we successfully fabricated several Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15) photocatalysts using a conventional solid-state reaction method. The Sr2Ti0.9Al0.1O4 photocatalyst exhibited the most outstanding photocatalytic HER activity (331 μmol h−1 g−1) among the three investigated Al3+-doped Sr2TiO4 photocatalysts under full-range light irradiation, which was nearly three times higher than that of Sr2TiO4. Such a remarkable improvement in the photocatalytic performance of Sr2TiO4 induced by Al3+ doping was attributed to the reduced particle sizes, enhanced light harvesting capability, facilitated charge transfer, and tailored band structure. Furthermore, Sr2Ti0.9Al0.1O4 also exhibited good stability and recyclability for photocatalytic HER. In sum, this study reports a highly promising catalyst (Sr2Ti0.9Al0.1O4) for photocatalytic water splitting, which may contribute to the further development of perovskite-based photocatalysts for efficient solar conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/hydrogen3040032/s1, Figure S1: N2 adsorption and desorption curves of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15); Figure S2. Sr 3d XPS spectra of Sr2Ti1-xAlxO4 (x = 0, 0.05, 0.1, and 0.15); Table S1. BET specific surface areas and band gap energies of various Sr2Ti1−xAlxO4 samples (x = 0, 0.05, 0.1, and 0.15).

Author Contributions

Investigation, formal analysis, visualization, writing—original draft, data curation, validation, J.H.; investigation, formal analysis, data curation, X.H.; data curation, H.X.; writing—review and editing, R.R.; supervision, conceptualization, writing—review and editing, funding acquisition, W.W.; project administration, funding acquisition, W.Z.; supervision, writing-review and editing, funding acquisition, Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 22279057) and the Postgraduate Research and Practice Innovation Program of Jiangsu Province (No. KYCX22_1352).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Østergaard, P.A.; Duic, N.; Noorollahi, Y.; Mikulcic, H.; Kalogirou, S. Sustainable development using renewable energy technology. Renew. Energy 2020, 146, 2430–2437. [Google Scholar] [CrossRef]
  2. Razmjoo, A.; Gakenia Kaigutha, L.; Vaziri Rad, M.A.; Marzband, M.; Davarpanah, A.; Denai, M. A technical analysis investigating energy sustainability utilizing reliable renewable energy sources to reduce CO2 emissions in a high potential area. Renew. Energy 2021, 164, 46–57. [Google Scholar] [CrossRef]
  3. Saavedra, M.M.R.; de O. Fontes, C.H.; Freires, M.F.G. Sustainable and renewable energy supply chain: A system dynamics overview. Renew. Sust. Energy Rev. 2018, 82, 247–259. [Google Scholar] [CrossRef]
  4. Kannan, N.; Vakeesan, D. Solar energy for future world: A review. Renew. Sust. Energy Rev. 2016, 62, 1092–1105. [Google Scholar] [CrossRef]
  5. Mekhilef, S.; Saidur, R.; Safari, A. A review on solar energy use in industries. Renew. Sust. Energy Rev. 2011, 15, 1777–1790. [Google Scholar] [CrossRef]
  6. Gong, J.; Li, C.; Wasielewski, M.R. Advances in solar energy conversion. Chem. Soc. Rev. 2019, 48, 1862–1864. [Google Scholar] [CrossRef]
  7. Rabaia, M.K.H.; Abdelkareem, M.A.; Sayed, E.T.; Elsaid, K.; Chae, K.J.; Wilberforce, T.; Olabi, A.G. Environmental impacts of solar energy systems: A review. Sci. Total Environ. 2021, 754, 141989. [Google Scholar] [CrossRef]
  8. Solangi, K.H.; Islam, M.R.; Saidur, R.; Rahim, N.A.; Fayaz, H. A review on global solar energy policy. Renew. Sust. Energy Rev. 2011, 15, 2149–2163. [Google Scholar] [CrossRef]
  9. Tosti, S.; Pozio, A.; Farina, L.; Santucci, A. Hydrogen and oxygen production via water splitting in a solar-powered membrane reactor-A conceptual study. Hydrogen 2021, 2, 18–32. [Google Scholar] [CrossRef]
  10. He, J.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Single-atom catalysts for high-efficiency photocatalytic and photoelectrochemical water splitting: Distinctive roles, unique fabrication methods and specific design strategies. J. Mater. Chem. A 2022, 10, 6835–6871. [Google Scholar] [CrossRef]
  11. Huang, D.; Xiang, H.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Recent advances in nanostructured inorganic hole-transporting materials for perovskite solar cells. Nanomaterials 2022, 12, 2592. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, W.; Xu, M.; Xu, X.; Zhou, W.; Shao, Z. Perovskite oxide based electrodes for high-performance photoelectrochemical water Splitting. Angew. Chem. Int. Ed. 2020, 59, 136–152. [Google Scholar] [CrossRef] [PubMed]
  13. Xiang, H.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Two-dimensional Dion-Jacobson halide perovskites as new-generation light absorbers for perovskite solar cells. Renew. Sust. Energy Rev. 2022, 166, 112614. [Google Scholar] [CrossRef]
  14. Yang, X.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Recent advances in Cs2AgBiBr6-based halide double perovskites as lead-free and inorganic light absorbers for perovskite solar cells. Energy Fuels 2020, 34, 10513–10528. [Google Scholar] [CrossRef]
  15. Takata, T.; Jiang, J.; Sakata, Y.; Nakabayashi, M.; Shibata, N.; Nandal, V.; Seki, K.; Hisatomi, T.; Domen, K. Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef] [PubMed]
  16. Ren, G.; Han, H.; Wang, Y.; Liu, S.; Zhao, J.; Meng, X.; Li, Z. Recent advances of photocatalytic application in water treatment: A review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef]
  17. Saeed, M.; Muneer, M.; Haq, A.; Akram, N. Photocatalysis: An effective tool for photodegradation of dyes—A review. Environ. Sci. Pollut. Res. 2022, 29, 293–311. [Google Scholar] [CrossRef]
  18. Wu, Y.; Zhong, L.; Yuan, J.; Xiang, W.; Xin, X.; Liu, H.; Luo, H.; Li, L.; Chen, M.; Zhong, D.; et al. Photocatalytic optical fibers for degradation of organic pollutants in wastewater: A review. Environ. Chem. Lett. 2021, 19, 1335–1346. [Google Scholar] [CrossRef]
  19. Xiang, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Sodium fluoride sacrificing layer concept enables high-efficiency and stable methylammonium lead iodide perovskite solar cells. J. Mater. Sci. Technol. 2022, 113, 138–146. [Google Scholar] [CrossRef]
  20. Yang, X.; Xiang, H.; Huang, J.; Zhou, C.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Thiourea with sulfur-donor as an effective additive for enhanced performance of lead-free double perovskite photovoltaic cells. J. Colloid Interface Sci. 2022, 628, 476–485. [Google Scholar] [CrossRef]
  21. Ahmad, K.; Ghatak, H.R.; Ahuja, S.M. A review on photocatalytic remediation of environmental pollutants and H2 production through water splitting: A sustainable approach. Environ. Technol. Innov. 2020, 19, 100893. [Google Scholar] [CrossRef]
  22. Nasir, A.M.; Awang, N.; Jaafar, J.; Ismail, A.F.; Othman, M.H.D.; Rahman, M.A.; Aziz, F.; Yajid, M.A.M. Recent progress on fabrication and application of electrospun nanofibrous photocatalytic membranes for wastewater treatment: A review. J. Water Process. Eng. 2021, 40, 101878. [Google Scholar] [CrossRef]
  23. Riaz, S.; Park, S.-J. An overview of TiO2-based photocatalytic membrane reactors for water and wastewater treatments. J. Ind. Eng. Chem. 2020, 84, 23–41. [Google Scholar] [CrossRef]
  24. Yao, S.; Yuan, X.; Jiang, L.; Xiong, T.; Zhang, J. Recent progress on fullerene-based materials: Synthesis, properties, modifications, and photocatalytic applications. Materials 2020, 13, 2924. [Google Scholar] [CrossRef]
  25. Fajrina, N.; Tahir, M. A critical review in strategies to improve photocatalytic water splitting towards hydrogen production. Int. J. Hydrogen Energy 2019, 44, 540–577. [Google Scholar] [CrossRef]
  26. Liu, X.; Chi, J.; Dong, B.; Sun, Y. Recent progress in decoupled H2 and O2 production from electrolytic water splitting. ChemElectroChem 2019, 6, 2157–2166. [Google Scholar] [CrossRef]
  27. Miseki, Y.; Sayama, K. Photocatalytic water splitting for solar hydrogen production using the carbonate effect and the Z-Scheme reaction. Adv. Energy Mater. 2019, 9, 1801294. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Guan, J. Recent progress in single-atom catalysts for photocatalytic water splitting. Sol. RRL 2020, 4, 2000283. [Google Scholar] [CrossRef]
  29. Hu, C.; Lin, Y.-R.; Yang, H.-C. Recent developments in graphitic carbon nitride based hydrogels as photocatalysts. ChemSusChem 2019, 12, 1794–1806. [Google Scholar] [CrossRef]
  30. Li, Y.; Song, X.; Zhang, G.; Wang, L.; Liu, Y.; Chen, W.; Chen, L. 2D covalent organic frameworks toward efficient photocatalytic hydrogen evolution. ChemSusChem 2022, 15, e202200901. [Google Scholar] [CrossRef]
  31. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef] [PubMed]
  32. Reddy, C.V.; Reddy, K.R.; Shetti, N.P.; Shim, J.; Aminabhavi, T.M.; Dionysiou, D.D. Hetero-nanostructured metal oxide-based hybrid photocatalysts for enhanced photoelectrochemical water splitting—A review. Int. J. Hydrogen Energy 2020, 45, 18331–18347. [Google Scholar] [CrossRef]
  33. Han, Y.; Chen, Y.; Fan, R.; Li, Z.; Zou, Z. Promotion effect of metal phosphides towards electrocatalytic and photocatalytic water splitting. EcoMat 2021, 3, e12097. [Google Scholar] [CrossRef]
  34. Lin, L.; Yu, Z.; Wang, X. Crystalline carbon nitride semiconductors for photocatalytic water splitting. Angew. Chem. Int. Ed. 2019, 58, 6164–6175. [Google Scholar] [CrossRef] [PubMed]
  35. Soto Morillo, E.; Mota Toledo, N.; García Fierro, J.L.; Navarro Yerga, R.M. Role of the sulphur source in the solvothermal synthesis of Ag-CdS photocatalysts: Effects on the structure and photoactivity for hydrogen production. Hydrogen 2020, 1, 64–89. [Google Scholar] [CrossRef]
  36. Singh, R.; Dutta, S. A review on H2 production through photocatalytic reactions using TiO2/TiO2-assisted catalysts. Fuel 2018, 220, 607–620. [Google Scholar] [CrossRef]
  37. Hamid, S.B.A.; Teh, S.J.; Lai, C.W. Photocatalytic water oxidation on ZnO: A review. Catalysts 2017, 7, 93. [Google Scholar] [CrossRef] [Green Version]
  38. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  39. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  40. Grabowska, E. Selected perovskite oxides: Characterization, preparation and photocatalytic properties-A review. Appl. Catal. B Environ. 2016, 186, 97–126. [Google Scholar] [CrossRef]
  41. Kumar, A.; Kumar, A.; Krishnan, V. Perovskite oxide based materials for energy and environment-oriented photocatalysis. ACS Catal. 2020, 10, 10253–10315. [Google Scholar] [CrossRef]
  42. Rodionov, I.A.; Zvereva, I.A. Photocatalytic activity of layered perovskite-like oxides in practically valuable chemical reactions. Russ. Chem. Rev. 2016, 85, 248–279. [Google Scholar] [CrossRef]
  43. Abdi, F.F.; Berglund, S.P. Recent developments in complex metal oxide photoelectrodes. J. Phys. D Appl. Phys. 2017, 50, 193002. [Google Scholar] [CrossRef]
  44. Liu, Y.; Yang, B.; He, H.; Yang, S.; Duan, X.; Wang, S. Bismuth-based complex oxides for photocatalytic applications in environmental remediation and water splitting: A review. Sci. Total Environ. 2022, 804, 150215. [Google Scholar] [CrossRef]
  45. Hu, Y.; Mao, L.; Guan, X.; Tucker, K.A.; Xie, H.; Wu, X.; Shi, J. Layered perovskite oxides and their derivative nanosheets adopting different modification strategies towards better photocatalytic performance of water splitting. Renew. Sust. Energy Rev. 2020, 119, 109527. [Google Scholar] [CrossRef]
  46. Xiao, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Ruddlesden-Popper perovskite oxides for photocatalysis-based water splitting and wastewater treatment. Energy Fuels 2020, 34, 9208–9221. [Google Scholar] [CrossRef]
  47. Xiao, H.; Liu, P.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z. Enhancing the photocatalytic activity of Ruddlesden-Popper Sr2TiO4 for hydrogen evolution through synergistic silver doping and moderate reducing pretreatment. Mater. Today Energy 2022, 23, 100899. [Google Scholar] [CrossRef]
  48. Han, X.; Liu, P.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Non-metal fluorine doping in Ruddlesden–Popper perovskite oxide enables high-efficiency photocatalytic water splitting for hydrogen production. Mater. Today Energy 2022, 23, 100896. [Google Scholar] [CrossRef]
  49. Hu, C.; Chen, T.-S.; Huang, H.-X. Heterojunction of n-type Sr2TiO4 with p-type Bi5O7I with enhanced photocatalytic activity under irradiation of simulated sunlight. Appl. Surf. Sci. 2017, 426, 536–544. [Google Scholar] [CrossRef]
  50. Sun, X.; Xu, X. Efficient photocatalytic hydrogen production over La/Rh co-doped Ruddlesden-Popper compound Sr2TiO4. Appl. Catal. B Environ. 2017, 210, 149–159. [Google Scholar] [CrossRef]
  51. 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]
  52. Mao, M.M.; Chen, X.M.; Liu, X.Q. Structure and microwave dielectric properties of solid solution in SrLaAlO4-Sr2TiO4 System. J. Am. Ceram. Soc. 2011, 94, 3948–3952. [Google Scholar] [CrossRef]
  53. Zhang, J.; Xie, K.; Li, Y.; Qi, W.; Ruan, C.; Wu, Y. Composite cathode based on Mn-doped perovskite niobate-titanate for efficient steam electrolysis. Chin. J. Chem. Phys. 2014, 27, 457–464. [Google Scholar] [CrossRef] [Green Version]
  54. Noguera, C. Theoretical investigation of the Ruddlesden-Popper compounds Srn+1TinO3n+1 (n = 1–3). Philos. Mag. Lett. 2000, 80, 173–180. [Google Scholar] [CrossRef]
  55. Sun, X.; Mi, Y.; Jiao, F.; Xu, X. Activating layered perovskite compound Sr2TiO4 via La/N codoping for visible light photocatalytic water splitting. ACS Catal. 2018, 8, 3209–3221. [Google Scholar] [CrossRef]
  56. Shang, P.; Ye, Z.; Ding, Y.; Zhu, Z.; Peng, X.; Ma, G.; Li, D. Nanosponge-like solid solution of NiMo with a high hydrogen evolution reaction performance over a wide range of current densities. ACS Sustain. Chem. Eng. 2020, 8, 10664–10672. [Google Scholar] [CrossRef]
  57. She, S.; Yu, J.; Tang, W.; Zhu, Y.; Chen, Y.; Sunarso, J.; Zhou, W.; Shao, Z. Systematic study of oxygen evolution activity and stability on La1-xSrxFeO3-δ perovskite electrocatalysts in alkaline media. ACS Appl. Mater. Interfaces 2018, 10, 11715–11721. [Google Scholar] [CrossRef]
  58. Kwak, B.S.; Do, J.Y.; Park, N.K.; Kang, M. Surface modification of layered perovskite Sr2TiO4 for improved CO2 photoreduction with H2O to CH4. Sci. Rep. 2017, 7, 16370. [Google Scholar] [CrossRef]
  59. Jing, L.; Qu, Y.; Wang, B.; Li, S.; Jiang, B.; Yang, L.; Fu, W.; Fu, H.; Sun, J. Review of photoluminescence performance of nano-sized semiconductor materials and its relationships with photocatalytic activity. Sol. Energy Mater. Sol. Cells 2006, 90, 1773–1787. [Google Scholar]
Figure 1. XRD patterns of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1 and 0.15).
Figure 1. XRD patterns of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1 and 0.15).
Hydrogen 03 00032 g001
Figure 2. SEM images and the corresponding grain-size histograms of (a) Sr2TiO4, (b) Sr2Ti0.95Al0.05O4, (c) Sr2Ti0.9Al0.1O4, and (d) Sr2Ti0.85Al0.15O4.
Figure 2. SEM images and the corresponding grain-size histograms of (a) Sr2TiO4, (b) Sr2Ti0.95Al0.05O4, (c) Sr2Ti0.9Al0.1O4, and (d) Sr2Ti0.85Al0.15O4.
Hydrogen 03 00032 g002
Figure 3. SEM-EDX images of (a) Sr2TiO4 and (b) Sr2Ti0.9Al0.1O4. This is a figure. Schemes follow the same formatting.
Figure 3. SEM-EDX images of (a) Sr2TiO4 and (b) Sr2Ti0.9Al0.1O4. This is a figure. Schemes follow the same formatting.
Hydrogen 03 00032 g003
Figure 4. XPS spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15): (a) Ti 2p, (b) O 1s, and (c) Al 2p.
Figure 4. XPS spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15): (a) Ti 2p, (b) O 1s, and (c) Al 2p.
Hydrogen 03 00032 g004
Figure 5. (a) UV-vis diffuse reflectance spectra, (b) Kubelka–Munk curves and (c) PL spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15).
Figure 5. (a) UV-vis diffuse reflectance spectra, (b) Kubelka–Munk curves and (c) PL spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15).
Hydrogen 03 00032 g005
Figure 6. (a) H2 evolution amounts and (b) average H2 evolution rates of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15) samples under full-range sunlight illumination.
Figure 6. (a) H2 evolution amounts and (b) average H2 evolution rates of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15) samples under full-range sunlight illumination.
Hydrogen 03 00032 g006
Figure 7. (a) H2 production amounts of Sr2Ti0.9Al0.1O4 in a continuous test period of 15 h. (b) XRD patterns of Sr2Ti0.9Al0.1O4 before and after the stability test. (c) Recycling tests of Sr2Ti0.9Al0.1O4 for photocatalytic HER. SEM images of Sr2Ti0.9Al0.1O4 (d) before and (e) after the stability test.
Figure 7. (a) H2 production amounts of Sr2Ti0.9Al0.1O4 in a continuous test period of 15 h. (b) XRD patterns of Sr2Ti0.9Al0.1O4 before and after the stability test. (c) Recycling tests of Sr2Ti0.9Al0.1O4 for photocatalytic HER. SEM images of Sr2Ti0.9Al0.1O4 (d) before and (e) after the stability test.
Hydrogen 03 00032 g007
Figure 8. (a) LSV curves and (b) EIS spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15). (c) MS curves and (d) schematically illustrated band structures of Sr2TiO4 and Sr2Ti0.9Al0.1O4.
Figure 8. (a) LSV curves and (b) EIS spectra of Sr2Ti1−xAlxO4 (x = 0, 0.05, 0.1, and 0.15). (c) MS curves and (d) schematically illustrated band structures of Sr2TiO4 and Sr2Ti0.9Al0.1O4.
Hydrogen 03 00032 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

He, J.; Han, X.; Xiang, H.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. Aluminum Cation Doping in Ruddlesden-Popper Sr2TiO4 Enables High-Performance Photocatalytic Hydrogen Evolution. Hydrogen 2022, 3, 501-511. https://doi.org/10.3390/hydrogen3040032

AMA Style

He J, Han X, Xiang H, Ran R, Wang W, Zhou W, Shao Z. Aluminum Cation Doping in Ruddlesden-Popper Sr2TiO4 Enables High-Performance Photocatalytic Hydrogen Evolution. Hydrogen. 2022; 3(4):501-511. https://doi.org/10.3390/hydrogen3040032

Chicago/Turabian Style

He, Jingsheng, Xiao Han, Huimin Xiang, Ran Ran, Wei Wang, Wei Zhou, and Zongping Shao. 2022. "Aluminum Cation Doping in Ruddlesden-Popper Sr2TiO4 Enables High-Performance Photocatalytic Hydrogen Evolution" Hydrogen 3, no. 4: 501-511. https://doi.org/10.3390/hydrogen3040032

Article Metrics

Back to TopTop