Next Article in Journal
Plasma–Liquid Synthesis of PLA/MXene Composite Films and Their Structural, Optical, and Photocatalytic Properties
Previous Article in Journal
Maximally Exploiting the Fe2+ at the Interface of Micro and Nano Bubbles in the Fenton-Coupled Micro and Nano Bubble System for Organic Pollutant Degradation
Previous Article in Special Issue
From Pigment to Photocatalyst: CdSe/CdS Solutions Mimicking Cadmium Red for Visible-Light Dye Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Defect Engineering via La Doping and Hydrogenation on Bi4Ti3O12 for Synergistically Enhancing Photocatalytic CO2 to CH3OH

1
School of Chemical and Material Engineering, Nanjing Polytechnic Institute, Nanjing 210048, China
2
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Jiangsu National Synergetic Innovation Center for Advanced Materials, Nanjing Tech University, Nanjing 211816, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(9), 889; https://doi.org/10.3390/catal15090889
Submission received: 30 August 2025 / Revised: 11 September 2025 / Accepted: 12 September 2025 / Published: 16 September 2025
(This article belongs to the Special Issue Catalysis Accelerating Energy and Environmental Sustainability)

Abstract

Developing highly efficient photocatalysts for CO2 reduction remains a great challenge. The large band gap and poor charge carrier dynamics are the major factors limiting the performance of Bi4Ti3O12 (BTO). Herein, a series of La-doped Bi4Ti3O12 (BLaxTO) nanosheets were synthesized and further modified by NaBH4 hydrogenation to create surface defect-rich H-BLaXTO nanosheets. Characterizations and theoretical calculations confirmed that the synergistic effect of La doping and hydrogenation significantly enhanced visible-light absorption, promoted charge separation, and improved the electron reduction capacity. When applied to photocatalytic CO2 reduction, the H-BLa0.2TO catalyst achieved a superior CH3OH production rate of 7.90 μmol·g−1·h−1, which is 5.6 times higher than that of pristine Bi4Ti3O12. Moreover, the H-BLa0.2TO catalyst maintained excellent stability over four consecutive cycles. This study offers an integrated strategy for constructing high-performance bismuth-based photocatalysts through elemental doping and defect engineering.

Graphical Abstract

1. Introduction

The extensive industrialization has resulted in massive consumption of fossil fuels and excessive CO2 emissions, culminating in energy and environmental issues [1,2,3]. In this context, CO2 conversion has emerged as a promising strategy to mitigate these issues. Among these, photocatalysis, which utilizes solar energy as a driving force, has garnered significant attention due to its green and sustainable characteristics [4,5,6]. However, the efficient and selective photocatalytic CO2 reduction into high-value-added products such as methanol (CH3OH) remains a considerable challenge. Accordingly, the rational design and development of highly efficient photocatalysts represents an urgent and critical need in the field of photocatalytic CO2 reduction [7,8,9,10].
In recent years, Bi4Ti3O12 (BTO), a typically Aurivillius-type perovskite, has attracted considerable attention in photocatalysis [11,12]. The crystal structure of single layer [Bi2O2]2+ and perovskite layer [Bi2Ti3O10]2− arranged alternately along c-axis is conducive to improving the separation rate of photogenerated electron–hole pairs. Moreover, the distinctive hybrid valence band structure, formed by Bi 6 s and O 2p orbitals, facilitates the generation of photogenerated carriers [13,14,15,16,17]. However, the pristine BTO suffers from the large band gap and poor charge carrier dynamics, which significantly inhibit its practical application in photocatalytic performance under visible light. Recent research proposed several strategies in modifying BTO to enhance its functionality. Yin et al. demonstrated that Fe doping combined with a ZnIn2S4 heterojunction was shown to narrow the bandgap and create a built-in electric field, significantly boosting CO2 reduction to CO and CH4 [18]. He et al. reported that decorating BTO with plasmonic metallic Bi and oxygen vacancies (OVs) via in situ reduction greatly improved visible-light absorption and charge separation, leading to highly efficient NO removal with suppressed toxic NO2 by-production [19]. Geng et al. reported that Ce4+ doping was demonstrated to induce lattice distortion, enhance ferroelectric polarization, and reduce the bandgap, resulting in superior piezo-photocatalytic RhB degradation under combined light and ultrasonic excitation [20]. Collectively, these studies confirm that tailoring electronic and surface properties through elemental doping and defect creation is a powerful route to improving its catalytic performance. Inspired by these approaches, this work proposes a synergistic strategy of A-site La3+ doping coupled with post-synthetic hydrogenation for advanced defect engineering in BTO. While the aforementioned studies utilized Fe/Ce doping or created OVs, the combination of La doping with a controlled hydrogenation process remains less explored. This approach holds promise for simultaneously achieving narrowing of the bandgap and suppression of recombination while introducing abundant surface OV sites and Ti3+ sites. We hypothesize that this dual approach will not only enhance visible-light absorption but also create a synergistic effect that drastically improves charge separation and provides more active sites, thereby developing a novel hydrogenated La-doped BTO (H-BLaxTO) catalyst for efficient photocatalytic CO2 reduction to methanol [21,22,23]. Although defect engineering in perovskite oxides has been explored, the specific effects of combining A-site La doping with NaBH4 hydrogenation on the photocatalytic behavior of BTO remain inadequately understood, warranting further systematic investigation.
Herein, we report La doping and NaBH4 hydrogenation co-modified BTO (H-BLa0.2TO) photocatalysts prepared via a glucose-assisted hydrothermal method. The performance of H-BLa0.2TO nanosheet was evaluated based on photocatalytic CO2 reduction into CH3OH. The result illustrated that the deliberate defect engineering played a crucial role in enhancing photocatalytic performance. Experiments and characterizations were applied to explore the inherent nature for the verification of enhanced photocatalytic activity. Furthermore, the density functional theory (DFT) was employed to further elucidate the electronic density of states.

2. Results and Discussion

2.1. Structure and Morphology

The crystal structures of the as-synthesized BTO, BLaXTO, and H-BLa0.2TO nanosheets were characterized by XRD. As shown in Figure 1a, the distinct diffraction peaks observed at 2θ = 10.8°, 16.3°, 21.7°, 23.3°, 30.1°, 32.8°, and 33.1° could be assigned to the (004), (006), (008), (111), (117), (200), and (020) of Bi4Ti3O12 (JCPDS Card No. 72-2181) [24]. No additional peaks were observed, illustrating that the introduction of La3+ and further NaBH4 hydrogenation did not change the crystal structure of BTO. The crystallite sizes of samples in Table S1 were calculated using the Scherrer equation. To further investigate the effect of A-site La doping on the catalyst structure, Raman spectra of BTO, BLa0.2TO, and H-BLa0.2TO were measured, as shown in Figure 1b. BTO exhibited strong and sharp phonon modes at 117, 148, 189, 228, 266, 328, 356, 534, 562, 614, and 848 cm−1, illustrating well-crystallized nature of the catalyst [25]. The low-frequency modes (<200 cm−1) are associated with the vibrations of the [Bi2O2]2+ layers, while the high-frequency modes (>200 cm−1) are generally classified as those related to the motion of cations in the [TiO6] octahedra. Compared with BTO, BLa0.2TO showed decreased phonon peak frequencies, peak broadening, and enhanced diffuseness. A slight red shift was observed in the high-frequency phonon modes, while the low-frequency modes remained largely unchanged. These changes suggest an altered coordination environment and modified structure of the [TiO6] octahedra, resulting from the substitution of Bi sites in the [Bi2Ti3O10]2− layers by La [26].
The microstructures of the synthesized BTO, BLaxTO, and H-BLa0.2TO photocatalysts were characterized by SEM. As depicted in Figure 2a–c and Figure S2, all samples exhibited a nanosheet-like shape with varying sizes. Notably, as the concentration of the La3+ doping increased, the dimensions of the nanosheets gradually decreased. This increased the number of active sites and light capture sites, which was beneficial for photocatalytic reactions. Figure 2d exhibited the EDS elemental mapping of the H-BLa0.2TO, where Bi, La, Ti, and O elements were evenly distributed throughout the nanosheets, demonstrating the successful preparation [27].
The TEM and HR-TEM images (Figure 3) provided more details of BTO, BLa0.2TO, and H-BLa0.2TO. All catalysts exhibited nanosheets of varying sizes, indicating that neither La incorporation nor the NaBH4 hydrogenation process disrupted the catalyst structure. In the HRTEM image of BTO, lattice fringes with an interplanar spacing of approximately 0.404 nm were observed, corresponding to the (008) crystallographic plane of Bi4Ti3O12. Similarly, lattice spacings of about 0.400 nm and 0.405 nm were identified in the HRTEM images of BLa0.2TO and H-BLa0.2TO, respectively, both of which can also be indexed to the (008) plane of Bi4Ti3O12 according to the standard reference [28].
The chemical composition and elemental states of the BTO, BLa0.2TO, and H-BLa0.2TO photocatalysts were analyzed by XPS, and the results were presented in Figure 4. As shown in the Bi 4f spectra (Figure 4a), the characteristic peaks of Bi 4f5/2 and Bi 4f7/2 were observed at binding energies of 158.9 eV and 164.3 eV, respectively. In BLa0.2TO, the Bi 4f5/2 peak exhibited a slight positive shift of 0.1 eV. After hydrogenation, both the Bi 4f5/2 and Bi 4f7/2 peaks of H-BLa0.2TO shifted positively by 0.3 eV and 0.2 eV, respectively. The La 3d spectra (Figure 4b) displayed characteristic doublets corresponding to La 3d5/2 and La 3d3/2 in both BLa0.2TO and H-BLa0.2TO. Peaks located at 834.0 eV and 838.0 eV were attributed to La 3d5/2, while those at 851.4 eV and 855.0 eV were assigned to La 3d3/2. The consistent La 3d peak positions in both samples indicated that the incorporated La3+ ions remained stable within the crystal lattice even after the NaBH4 hydrogenation treatment. In the Ti 2p spectra (Figure 4c), the characteristic peaks of Ti4+ 2p3/2 and Ti4+ 2p1/2 were located at 457.8 eV and 463.5 eV, respectively. An additional peak observed at 466.0 eV is associated with Bi 4p3/2. For BLa0.2TO, the Ti4+ 2p3/2 and 2p1/2 peaks remained at similar positions, while a peak at 466.3 eV was assigned to Bi 4s. Notably, in H-BLa0.2TO, a new peak appeared at 458.4 eV, which was attributed to Ti3+, indicating the partial reduction of Ti4+ to Ti3+ due to the NaBH4 treatment. Furthermore, compared to BLa0.2TO, the Ti4+ 2p3/2 and 2p1/2 peaks in H-BLa0.2TO exhibited negative and positive shifts of 0.6 eV and 0.1 eV, respectively, likely resulting from charge compensation induced by Ti3+ formation. The O 1s spectra (Figure 4d) revealed two main contributions at 529.6 eV and 530.7 eV, corresponding to lattice oxygen (OL) and oxygen vacancies (OV), respectively. The introduction of La at the A-site did not significantly alter the concentration of surface oxygen vacancies in BLa0.2TO. In contrast, the hydrogenation process with NaBH4 notably increased the amount of OVs on the surface of H-BLa0.2TO [29,30,31]. The concentration of oxygen vacancies on the catalyst surface was further determined by O2-TPD, as shown in Figure S3. The peak observed between 200 °C and 500 °C can be corresponded to the oxidation of OVs. The introduction of La at the A-site had little influence on the oxygen vacancy concentration, whereas the high-temperature hydrogenation with NaBH4 significantly increased it. These results are consistent with the XPS O 1s analysis [32].

2.2. Photoelectrochemical Properties

The optical absorption properties of the catalysts were analyzed using UV–vis DRS. As shown in Figure 5a, all samples exhibited strong and steep absorption edges in the visible-light region. The incorporation of La3+ into the Bi4Ti3O12 lattice introduces metal impurity levels near the bottom of the conduction band (CB), reducing the bandgap and thereby enhancing light absorption. Consequently, BLa0.2TO demonstrated improved visible-light absorption compared to BTO. The subsequent NaBH4 hydrogenation further enhanced the visible-light absorption capability of H-BLa0.2TO, leading to higher light utilization efficiency. Furthermore, the valence band (VB) positions of the photocatalysts were determined via XPS valence band spectra. As depicted in Figure 5b, the VB positions of BLa0.2TO and H-BLa0.2TO were measured to be 2.05 V and 1.75 V, respectively. The CB positions of BLa0.2TO and H-BLa0.2TO were estimated based on Mott–Schottky measurements. The flat-band potentials, obtained from the Mott–Schottky plots (Figure S4), were −0.95 V and −0.98 V (vs. Ag/AgCl) for BLa0.2TO and H-BLa0.2TO, respectively. After conversion, the corresponding CB positions were calculated to be −0.85 V for BLa0.2TO and −0.88 V for H-BLa0.2TO. In comparison, the conduction band of pristine BTO has been previously reported to be −0.8 V. The A-site La doping resulted in a more negative conduction band position of BLa0.2TO than that of BTO, indicating that the electrons in the conduction band of BLa0.2TO possess stronger reducing ability. The hydrogenation process with NaBH4 further optimized the conduction band position of H-BLa0.2TO, endowing it with even greater electron reduction power, which is highly beneficial for the photocatalytic CO2 conversion to CH3OH [33,34].
The migration and separation efficiency of photoinduced charge carriers in the samples were further investigated through I-t and EIS measurements. As shown in the Nyquist plots (Figure 6a), H-BLa0.2TO exhibited the smallest arc radius, indicating the lowest charge transfer resistance and the fastest photoinduced electron migration rate among the samples. These results suggest that H-BLa0.2TO possesses enhanced charge separation efficiency and reduced recombination, contributing to its superior photocatalytic activity. The I-t curve (Figure 6b) further confirmed that H-BLa0.2TO generated the highest photocurrent density, demonstrating its excellent ability to separate photoinduced electron–hole pairs [35].
To gain further insight into the charge separation behavior, steady-state and time-resolved photoluminescence spectroscopy were conducted. The steady-state PL spectra (Figure 6c) showed that La doping at the A-site resulted in a significant quenching of the PL intensity, indicating that the introduction of La effectively suppressed the recombination of photogenerated carriers. The TRPL decay curves (Figure 6d) revealed a shorter fluorescence lifetime for the La-doped sample, suggesting an enhanced charge transfer capability due to A-site La substitution [36].

2.3. Photocatalysis Performance

The photocatalytic performance of the as-prepared samples was evaluated by the photocatalytic reduction of CO2 to CH3OH. Figure 7a shows the production rates of CO, CH3OH, and CH4 during the reaction. The CH3OH production rates over BTO, BLa0.1TO, BLa0.2TO, BLa0.3TO, BLa0.4TO, and H-BLa0.2TO were 1.40, 3.04, 4.70, 4.37, 3.85, and 7.90 μmol·g−1·h−1, respectively. Among the non-hydrogenated samples, BLa0.2TO exhibited the highest CH3OH yield. After hydrogenation, the CH3OH production rate was further enhanced. This improvement can be attributed to the introduction of an optimal amount of La into the A-site, which forms impurity energy levels that facilitate the separation of photogenerated charge carriers. Additionally, the hydrogenation process promotes the formation of surface defects, modifying the band structure and thereby enhancing the photocatalytic activity. Control experiments were carried out over H-BLa0.2TO under different conditions (Figure 7b). No CH3OH was detected in the absence of the catalyst, light irradiation, or CO2 flow, confirming that the CH3OH production indeed originated from the photocatalytic reduction of CO2. The stability of the H-BLa0.2TO catalyst was examined through four consecutive recycling experiments (Figure 7c) and by comparing the XRD patterns before and after the reaction (Figure 7d). No significant loss of photocatalytic activity was observed after four cycles. Moreover, the XRD patterns of H-BLa0.2TO remained virtually unchanged after the reaction, indicating that the crystal structure was well maintained and demonstrating the excellent cycling stability of the catalyst.

2.4. Photocatalytic Mechanism

Density functional theory (DFT) calculations were performed to investigate the electronic band structure, total density of states (TDOS), and projected density of states (PDOS) of H-BLa0.2TO, providing insight into the contributions of atomic orbitals to the valence and conduction bands. As shown in Figure 8a, the calculated band gap of H-BLa0.2TO is 0.953 eV, which is significantly narrower than that of pristine BTO (1.449 eV), confirming its enhanced visible-light absorption capability. Moreover, the valence band maximum and conduction band minimum are located at different high-symmetry points in the Brillouin zone, indicating an indirect band gap nature. The TDOS of H-BLa0.2TO (Figure 8b) reveals that the electronic states near the band edges are predominantly contributed by p-orbitals. Further analysis of the PDOS (Figure 8c–f) for Bi, La, Ti, and O atoms indicates that the valence band is mainly formed by O 2p orbitals, which exhibit higher energy states near this region. This suggests that the O 2p orbitals play a critical role in determining the electronic properties of the valence band in H-BLa0.2TO [37,38].
Based on the above discussion, a possible reaction mechanism for the photocatalytic CO2 reduction over the H-BLa0.2TO catalyst is proposed (Figure 9). La doping at the A-site of Bi4Ti3O12 via the sol–gel-assisted molten salt method introduces metal impurity energy levels, which not only narrow the bandgap of the catalyst but also effectively suppress the recombination of photogenerated charge carriers. Subsequent hydrogenation with NaBH4 constructs abundant surface defects, further enhancing the reduction ability of the electrons in the conduction band and thereby improving the overall photocatalytic performance. Under light irradiation, H-BLa0.2TO absorbs photons and generates electron–hole pairs. The photogenerated electrons in the conduction band then facilitate the reduction of CO2 and H2O to produce CH3OH.

3. Experimental Section

3.1. Materials

Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O, ≥99.9%), Lanthanum nitrate hexahydrate (La(NO3)3·6H2O, ≥99.9%), Tetrabutyl titanate (C16H36O4Ti, ≥99.9%), Potassium chloride (KCl, AR), Sodium chloride (NaCl, AR), Sodium borohydride (NaBH4, AR), Acetic anhydride (C4H6O3, AR), Ethylene glycol(C2H6O2, AR), and Glacial acetic acid (CH3COOH, AR) were obtained from Aladdin Reagent.

3.2. Fabrication of H-BLa0.2TO Nanosheets

The hydrogenated La-doped Bi4Ti3O12 (H-BLa0.2TO) nanosheets were fabricated via a sequential process involving sol–gel synthesis, molten salt annealing, and a subsequent hydrogenation treatment.
Synthesis of Bi3.8La0.2Ti3O12 (BLa0.2TO). Typically, 9.5 mmol of Bi(NO3)3·5H2O and 5 mmol of La(NO3)3·6H2O were dissolved in a mixed solvent of 7.5 mL acetic anhydride and 3 mL glacial acetic acid (solution A). Simultaneously, 7.5 mmol of C16H36O4Ti was dissolved in a mixture of 12.5 mL ethylene glycol and 2 mL glacial acetic acid (solution B). Solution B was then added dropwise into solution A under vigorous stirring. The resulting mixture was continuously stirred for 30 min and aged for 10 h at room temperature. The obtained gel was dried at 110 °C for 48 h to form a xerogel. The dry xerogel was thoroughly ground with a eutectic mixture of NaCl/KCl (0.125 mol each) and subsequently calcined at 800 °C for 2 h in a muffle furnace. The product was washed repeatedly with deionized water to remove soluble salts and dried at 100 °C for 6 h, yielding Bi3.8La0.2Ti3O12 nanosheets, designated as BLa0.2TO. Using the same procedure but adjusting the molar ratios of the precursors, Bi4-xLaₓTi3O12 samples with different doping levels (x = 0.1, 0.3, 0.4) were also prepared.
Hydrogenation Treatment. The as-prepared BLa0.2TO powder (1.0 g) was uniformly mixed with 0.2 g of sodium borohydride (NaBH4) by grinding in an agate mortar for 30 min. The mixture was then transferred into a ceramic boat and placed in a tube furnace. Under a continuous Ar flow, the temperature was raised to 350 °C at a ramp rate of 5 °C·min−1 and maintained for 1 h. After cooling down to room temperature naturally, the final product was collected and washed with deionized water until neutral, followed by drying at 80 °C for 8 h. The resulting hydrogenated catalyst was denoted as H-BLa0.2TO. The preparation process is displayed in Figure 10.

4. Conclusions

In summary, a series of La-doped Bi4Ti3O12 nanosheets were successfully synthesized via a sol–gel-assisted molten salt method, followed by a NaBH4 hydrogenation treatment to obtain the hydrogenated H-BLa0.2TO catalyst. The synergistic effect of A-site La doping and hydrogenation engineering was systematically demonstrated to significantly enhance the photocatalytic performance for CO2 reduction to CH3OH. The incorporation of La3+ into the Bi4Ti3O12 lattice effectively narrowed the bandgap by introducing impurity energy levels, thereby extending the visible-light absorption range. More importantly, the hydrogenation process with NaBH4 created abundant surface oxygen vacancies and partially reduced Ti4+ to Ti3+, which served as active sites and further optimized the electronic structure. A combination of experimental characterizations and DFT calculations confirmed that these modifications collectively promoted the separation and migration of photogenerated charge carriers while endowing the catalyst with a more negative conduction band potential, thereby enhancing its reducing power. Consequently, the optimized H-BLa0.2TO catalyst achieved a remarkable CH3OH production rate of 7.90 μmol·g−1·h−1, which was 5.6 times higher than that of pristine BTO. Furthermore, the catalyst exhibited excellent structural and catalytic stability over multiple reaction cycles. This work highlights defect engineering via elemental doping and post-synthesis treatment as a highly effective strategy for designing efficient bismuth-based photocatalysts for solar-driven CO2 conversion to value-added fuels.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15090889/s1, Figure S1: XRD spectrum of BLaXTO; Figure S2: SEM images of (a) BLa0.1TO, (b) BLa0.3TO, and (c) BLa0.4TO nanosheets; Figure S3: O2-TPD of different samples; Figure S4: Mott-Schottky diagram of (a) BLa0.2TO and (b) H-BLa0.2TO; Table S1: The crystallite sizes of the samples.

Author Contributions

Writing—original draft preparation, conceptualization, and investigation, L.X. and Y.W.; methodology, C.Q.; supervision and funding acquisition, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

The research received no external funding.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Yergaziyeva, G.; Kuspanov, Z.; Mambetova, M.; Khudaibergenov, N.; Makayeva, N.; Daulbayev, C. Advancements in Catalytic, Photocatalytic, and Electrocatalytic CO2 Conversion Processes: Current Trends and Future Outlook. J. CO2 Util. 2024, 80, 102682. [Google Scholar] [CrossRef]
  2. Garba, M.D.; Usman, M.; Khan, S.; Shehzad, F.; Galadima, A.; Ehsan, M.F.; Ghanem, A.S.; Humayun, M. CO2 towards Fuels: A Review of Catalytic Conversion of Carbon Dioxide to Hydrocarbons. J. Environ. Chem. Eng. 2021, 9, 104756. [Google Scholar] [CrossRef]
  3. Francis, K.; Kim, H.; Hyeak, C.; Seop, H. Covalent Organic Framework-Based Catalysts for Efficient CO2 Utilization Reactions. Coord. Chem. Rev. 2022, 473, 214835. [Google Scholar] [CrossRef]
  4. Li, S.; Yang, Y.; Wan, S.; Wang, R.; Yu, M.; Song, F.; Zhong, Q. Supramolecular Self-Assembled Deficient Carbon Nitride Nanotubes for Efficient Photocatalytic CO2 Reduction. J. Colloid Interface Sci. 2023, 651, 726–733. [Google Scholar] [CrossRef]
  5. Wang, L.; Wang, L.; Zhang, J.; Liu, X.; Wang, H.; Zhang, W.; Yang, Q.; Ma, J.; Dong, X.; Yoo, S.J.; et al. Selective Hydrogenation of CO2 to Ethanol over Cobalt Catalysts. Angew. Chem. Int. Ed. 2018, 57, 6104–6108. [Google Scholar] [CrossRef]
  6. Wang, S.; Wang, J.; Wang, Y.; Sui, X.; Wu, S.; Dai, W.; Zhang, Z.; Ding, Z.; Long, J. Insight into the Selectivity-Determining Step of Various Photocatalytic CO2 Reduction Products by Inorganic Semiconductors. ACS Catal. 2024, 14, 10760–10788. [Google Scholar] [CrossRef]
  7. Stolarczyk, J.K.; Bhattacharyya, S.; Polavarapu, L.; Feldmann, J. Challenges and Prospects in Solar Water Splitting and CO2 Reduction with Inorganic and Hybrid Nanostructures. ACS Catal. 2018, 8, 3602–3635. [Google Scholar] [CrossRef]
  8. Ding, J.; He, D.; Du, P.; Wu, J.; Hu, Q.; Chen, Q.; Jiao, X. Design Photocatalysts to Poost Parrier Dynamics in Plastics Photoconversion into Fuels. ACS Appl. Mater. Interfaces 2024, 16, 35865–35873. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, S.; Zhou, Q.; Wen, D.; Wu, C.; Pan, Y.; Liu, X.; Huang, Z.; Li, N. Particle Size-Dependent Charge Transfer Dynamics for Boosting CO2 Photoreduction Over Ag/TiO2 Heterojunction. ACS Catal. 2024, 14, 8105–8115. [Google Scholar] [CrossRef]
  10. Zhang, L.; Yang, Y.; Tian, J.; Li, J.; Chen, G.; Zhou, L.; Sun, Y.; Qiu, Y. Selective Photocatalytic Reduction of CO2 to Syngas over Tunable Metal-Perovskite Interface. ChemSusChem 2022, 15, e202102729. [Google Scholar] [CrossRef] [PubMed]
  11. Jia, G.; Sun, F.; Zhou, T.; Wang, Y.; Cui, X.; Guo, Z.; Fan, F.; Yu, J.C. Charge Redistribution of a Spatially Differentiated Ferroelectric Bi4Ti3O12 Single Crystal for Photocatalytic Overall Water Splitting. Nat. Commun. 2024, 15, 4746. [Google Scholar] [CrossRef]
  12. Shen, M.; Shi, Y.; Wang, Z.; Wu, T.; Hu, L.; Wu, L. Enhanced Photocatalytic Benzyl Alcohol Oxidation over Bi4Ti3O12 Ultrathin Nanosheets. J. Colloid Interface Sci. 2022, 608, 2529–2538. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Y.; Zheng, D.; Zheng, X.; Han, W. Oxygen Vacancies and Electrical Properties of Different Valence Ions (W6+, Nb5+, Zr4+, Cr3+) Doped Bi4Ti3O12 Ceramics. Ceram. Int. 2024, 50, 2778–2787. [Google Scholar] [CrossRef]
  14. Liu, L.; Huang, H.; Chen, F.; Yu, H.; Tian, N.; Zhang, Y.; Zhang, T. Cooperation of Oxygen Vacancies and 2D Ultrathin Structure Promoting CO2 Photoreduction Performance of Bi4Ti3O12. Sci. Bull. 2020, 65, 934–943. [Google Scholar] [CrossRef] [PubMed]
  15. Gao, F.; Li, W.; Duan, W.; Liao, G.; Wang, C. Improved Solar-Powered Water-Splitting Performance of Bi4Ti3O12/TiO2 Composite with Synergistically Interacted Heterointerfaces Under Platinum Cocatalysis. Small 2025, 21, 2503677. [Google Scholar] [CrossRef]
  16. Cheng, X.; Cao, T.; Li, X.; Luo, G.; Cao, S.; Wang, C.; Guo, D. Enhanced Visible-Light Photocatalytic Performance of Hollow Bi/Bi4Ti3O12 Microspheres Assembled with Bi4Ti3O12 Nanosheets and Bi Nanodots Synthesized by Hydrothermal Method. J. Alloys Compd. 2024, 1010, 177483. [Google Scholar] [CrossRef]
  17. Liu, L.; Hu, J.; Sheng, Y.; Akhoundzadeh, H.; Tu, W.; Siow, W.J.S.; Ong, J.H.; Huang, H.; Xu, R. Ru Single Atom Dispersed Cu Nanoparticle with Dual Sites Enables Outstanding Photocatalytic CO2 Reduction. ACS Nano 2024, 18, 26271–26280. [Google Scholar] [CrossRef]
  18. Yin, Y.; Wang, K.; Peng, R.; Zhou, J.; Li, N. Interface Charge Separation in Fe-Doped Bi4Ti3O12/ZnIn2S4 Heterostructure for Enhancing Photocatalytic CO2 Reduction. J. Alloys Compd. 2025, 1022, 180167. [Google Scholar] [CrossRef]
  19. He, W.; Jing, L.; Zhang, W.; Dong, X.; Zou, Y.; Liu, X.; Lv, X.; Chen, P.; Liao, J.; Zhang, X.; et al. Plasmonic Metallic Bi-Modified Defective Bi4Ti3O12 Nanosheets with Upward Migrating Electrons for Efficient Photocatalytic NO Removal and NO2 Inhibition. Chin. Chem. Lett. 2025, 111357. [Google Scholar] [CrossRef]
  20. Geng, S.; Sui, X.; Li, Y.; Wang, H.; Zhao, X.; Chang, L.; Duan, X. Reinforcing Piezo-Photocatalytic Properties and Enhancing RhB Degradation Efficiency of Ce-Doped Bi4Ti3O12: Mechanistic Insights. J. Mater. Chem. A 2025, 13, 23838–23852. [Google Scholar] [CrossRef]
  21. Wang, Y.; Qiu, C.; Xie, Y.; Wang, L.; Ding, J.; Zhang, J.; Wan, H.; Guan, G. Intentionally Introducing Oxygen Vacancies and Ti3+ Defects on the Surface of Bi4Ti3O12 Nanosheets for Promoting the Photoreduction of CO2 to CH3OH. ACS Appl. Nano Mater. 2024, 7, 3012–3023. [Google Scholar] [CrossRef]
  22. Qiu, C.; Wang, L.; Chen, R.; Zhang, J.; Ding, J.; Zhang, J.; Wan, H.; Guan, G. Insight of the state for deliberately introduced A-site defect in nanofibrous LaFeO3 for boosting artificial photosynthesis of CH3OH. ACS Appl. Mater. Interfaces 2023, 15, 56945–56956. [Google Scholar] [CrossRef]
  23. Arif, M.; Zhang, M.; Mao, Y.; Bu, Q.; Ali, A.; Qin, Z.; Muhmood, T.; Shahnoor; Liu, X.; Zhou, B.; et al. Oxygen Vacancy Mediated Single Unit Cell Bi2WO6 by Ti Doping for Ameliorated Photocatalytic Performance. J. Colloid Interface Sci. 2021, 581, 276–291. [Google Scholar] [CrossRef]
  24. Jia, P.; Li, Y.; Zheng, Z.; Wang, Y.; Liu, T. Ferroelectric Polarization Promotes the Excellent CO2 Photoreduction Performance of Bi4Ti3O12 Synthesized by Molten Salt Method. J. Alloys Compd. 2022, 920, 165880. [Google Scholar] [CrossRef]
  25. Tech, S.J.; Choong, C.E.; Choi, E.H.; Yoon, Y.; Jang, M. Ru-doped Aurivillius phase perovskites Bi4Ti3O12 for boosting photo-redox of N2 in photocatalytic nitrogen fixation. Appl. Surf. Sci. 2025, 684, 161890. [Google Scholar]
  26. Anwar, K.; Naqvi, F.K.; Beg, S. Facile Hydrothermal Synthesis of Yb Doped Bi4V2O11 Nanoparticles for the Improvement of Photocatalytic and AC Impedance Performance: Investigation of Polymorphism, Adsorption Isotherm, Optical Properties and Possible Mechanism with the Help of GCMS. J. Alloys Compd. 2023, 938, 168483. [Google Scholar] [CrossRef]
  27. Wang, J.; Liu, W.; Zhong, D.; Ma, Y.; Ma, Q.; Wang, Z.; Pan, J. Fabrication of Bismuth Titanate Nanosheets with Tunable Crystal Facets for Photocatalytic Degradation of Antibiotic. J. Mater. Sci. 2019, 54, 13740–13752. [Google Scholar] [CrossRef]
  28. Sun, S.; Dong, Q.; Chen, Z. Macroscopic Polarization Enhancement Boosting Piezo-Photocatalytic Performance of Bi4Ti3O12 via La-doping. Ceram. Int. 2025, 51, 3846–3856. [Google Scholar] [CrossRef]
  29. Jiang, H.; Wang, L.; Kaneko, H.; Gu, R.; Su, G.; Li, L.; Zhang, J.; Song, H.; Zhu, F.; Yamaguchi, A.; et al. Light-driven CO2 methanation over Au-grafted Ce0.95Ru0.05O2 solid-solution catalysts with activities approaching the thermodynamic limit. Nat. Catal. 2013, 6, 519–530. [Google Scholar] [CrossRef]
  30. Fang, H.; Kang, Y.; Yuan, S.; Zhang, M.; Rui, Z. One step synthesis and interfacial properties of black Ag/TiO2-x for enhancing sunlight absorption with application to photothermocatalytic VOCs degradation. Appl. Surf. Sci. 2024, 655, 159519. [Google Scholar] [CrossRef]
  31. Tang, Q.; Wu, J.; Chen, X.-Z.; Sanchis-Gual, R.; Veciana, A.; Franco, C.; Kim, D.; Surin, I.; Pérez-Ramírez, J.; Mattera, M.; et al. Tuning Oxygen Vacancies in Bi4Ti3O12 Nanosheets to Boost Piezo-Photocatalytic Activity. Nano Energy 2023, 108, 108202. [Google Scholar] [CrossRef]
  32. Mao, C.; Cheng, H.; Tian, H.; Li, H.; Xiao, W.-J.; Xu, H.; Zhao, J.; Zhang, L. Visible Light Driven Selective Oxidation of Amines to Imines with BiOCl: Does Oxygen Vacancy Concentration Matter. Appl. Catal. B-Environ. 2018, 228, 87–96. [Google Scholar] [CrossRef]
  33. Fang, F.; Zhang, J.; Xiang, L.; Chen, C.; Feng, N.; Lv, Y.; Chang, K.; Huang, J. Ordering bimetallic Cu-Pd catalysts onto orderly mesoporous SrTiO3-crystal nanotubular networks for efficient carbon dioxide photoreduction. Angew. Chem. Int. Ed. 2024, 63, e202405807. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, R.; Zhang, Z.; Wu, J.; Chen, X.; Wang, L.; Yin, H.; Li, H.; Ding, J.; Wan, H.; Guan, G. “Carbon diffusion” engineered carbon nitride nanosheets for high-efficiency photocatalytic solar-to-fuels conversion. Renew. Energy 2022, 197, 943–952. [Google Scholar] [CrossRef]
  35. Feng, X.; Sun, T.; Feng, X.; Chen, L.; Yang, Y.; Zhang, F. Engineering the Near-Surface Structure of WO3 by an Amorphous Layer with Trivalent Ni and Self-Adapting Oxygen Vacancies for Efficient Photocatalytic and Photoelectrochemical Acidic Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2022, 14, 54769–54780. [Google Scholar] [CrossRef]
  36. Huang, M.; Zhang, S.; Wu, B.; Wei, Y.; Yu, X.; Gan, Y.; Lin, T.; Yu, F.; Sun, F.; Jiang, Z.; et al. Selective Photocatalytic Oxidation of Methane to Oxygenates over Cu–W–TiO2 with Significant Carrier Traps. ACS Catal. 2022, 12, 9515–9525. [Google Scholar] [CrossRef]
  37. Li, G.; Lian, Z.; Wan, Z.; Liu, Z.; Qian, J.; Deng, Y.; Zhang, S.; Zhong, Q. Efficient Photothermal-Assisted Photocatalytic NO Removal on Molecular Cobalt Phthalocyanine/Bi2WO6 Z-Scheme Heterojunctions by Promoting Charge Transfer and Oxygen Activation. Appl. Catal. B 2022, 317, 121787. [Google Scholar] [CrossRef]
  38. Xia, P.; Cao, S.; Zhu, B.; Liu, M.; Shi, M.; Yu, J.; Zhang, Y. Designing a 0D/2D S-Scheme Heterojunction over Polymeric Carbon Nitride for Visible-Light Photocatalytic Inactivation of Bacteria. Angew. Chem. Int. Ed. 2020, 59, 5218–5225. [Google Scholar] [CrossRef]
Figure 1. (a) XRD and (b) Raman spectra of BTO, BLa0.2TO, and H-BLa0.2TO.
Figure 1. (a) XRD and (b) Raman spectra of BTO, BLa0.2TO, and H-BLa0.2TO.
Catalysts 15 00889 g001
Figure 2. SEM images of (a) BTO, (b) BLa0.2TO, and (c) H-BLa0.2TO nanosheets, and the (d) EDS elemental mappings of H-BLa0.2TO.
Figure 2. SEM images of (a) BTO, (b) BLa0.2TO, and (c) H-BLa0.2TO nanosheets, and the (d) EDS elemental mappings of H-BLa0.2TO.
Catalysts 15 00889 g002
Figure 3. TEM and HRTEM images of (a,d) BTO, (b,e) BLa0.2TO, (c,f) H-BLa0.2TO.
Figure 3. TEM and HRTEM images of (a,d) BTO, (b,e) BLa0.2TO, (c,f) H-BLa0.2TO.
Catalysts 15 00889 g003
Figure 4. XPS spectra of (a) Bi 4f, (b) La 3d, (c) Ti 2p, (d) O 1s.
Figure 4. XPS spectra of (a) Bi 4f, (b) La 3d, (c) Ti 2p, (d) O 1s.
Catalysts 15 00889 g004
Figure 5. (a) UV–vis DRS spectra, (b) VB-XPS plot, and (c) band gap structure diagram of samples.
Figure 5. (a) UV–vis DRS spectra, (b) VB-XPS plot, and (c) band gap structure diagram of samples.
Catalysts 15 00889 g005
Figure 6. (a) EIS, (b) I-t, (c) PL, and (d) TRPL of samples.
Figure 6. (a) EIS, (b) I-t, (c) PL, and (d) TRPL of samples.
Catalysts 15 00889 g006
Figure 7. (a) Performance diagram for photocatalytic CO2 reduction, (b) controlled experimental diagram of H-BLa0.2TO under different conditions, (c) photocatalytic cycle stability of H-BLa0.2TO, and (d) XRD pattern of H-BLa0.2TO before and after use.
Figure 7. (a) Performance diagram for photocatalytic CO2 reduction, (b) controlled experimental diagram of H-BLa0.2TO under different conditions, (c) photocatalytic cycle stability of H-BLa0.2TO, and (d) XRD pattern of H-BLa0.2TO before and after use.
Catalysts 15 00889 g007
Figure 8. Theoretical calculation diagram of (a) band structure and (b) TDOS of H-BLa0.2TO, PDOS of (c) Bi atom, (d) La atom, (e) Ti atom, and (f) O atom.
Figure 8. Theoretical calculation diagram of (a) band structure and (b) TDOS of H-BLa0.2TO, PDOS of (c) Bi atom, (d) La atom, (e) Ti atom, and (f) O atom.
Catalysts 15 00889 g008
Figure 9. Possible photocatalytic mechanism of CO2.
Figure 9. Possible photocatalytic mechanism of CO2.
Catalysts 15 00889 g009
Figure 10. Schematic diagram of H-BLa0.2TO nanosheet synthesis route.
Figure 10. Schematic diagram of H-BLa0.2TO nanosheet synthesis route.
Catalysts 15 00889 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xue, L.; Wang, Y.; Qiu, C.; Wan, H. Defect Engineering via La Doping and Hydrogenation on Bi4Ti3O12 for Synergistically Enhancing Photocatalytic CO2 to CH3OH. Catalysts 2025, 15, 889. https://doi.org/10.3390/catal15090889

AMA Style

Xue L, Wang Y, Qiu C, Wan H. Defect Engineering via La Doping and Hydrogenation on Bi4Ti3O12 for Synergistically Enhancing Photocatalytic CO2 to CH3OH. Catalysts. 2025; 15(9):889. https://doi.org/10.3390/catal15090889

Chicago/Turabian Style

Xue, Lijun, Yuxuan Wang, Chenhui Qiu, and Hui Wan. 2025. "Defect Engineering via La Doping and Hydrogenation on Bi4Ti3O12 for Synergistically Enhancing Photocatalytic CO2 to CH3OH" Catalysts 15, no. 9: 889. https://doi.org/10.3390/catal15090889

APA Style

Xue, L., Wang, Y., Qiu, C., & Wan, H. (2025). Defect Engineering via La Doping and Hydrogenation on Bi4Ti3O12 for Synergistically Enhancing Photocatalytic CO2 to CH3OH. Catalysts, 15(9), 889. https://doi.org/10.3390/catal15090889

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