Photocatalytic methane oxidation over a TiO2/SiNWs p–n junction catalyst at room temperature

  1. 1 ORCID Logo ,
  2. 1,2 ORCID Logo ,
  3. 1,2 ,
  4. 3 ORCID Logo and
  5. 1,2 ORCID Logo
1Institute of Chemical Technology, Vietnam Academy of Science and Technology, 1A TL29 Street, Thanh Loc Ward, District 12, Ho Chi Minh City 700000, Vietnam
2Graduate University of Science and Technology, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet Street, Cau Giay District, Hanoi 100000, Vietnam
3Department of Chemical and Biological Engineering, Gachon University, 1342 Seongnamdaero, Sujeong-gu, Seongnam-si, Gyeonggi-do 13120, Republic of Korea
  1. Author email
  2. Corresponding author email
Guest Editor: B. T. Huy
Beilstein J. Nanotechnol. 2024, 15, 1132–1141. https://doi.org/10.3762/bjnano.15.92
Received 18 May 2024, Accepted 13 Aug 2024, Published 02 Sep 2024
Full Research Paper
cc by logo

Abstract

Rapid recombination of charge carriers in semiconductors is a main drawback for photocatalytic oxidative coupling of methane (OCM) reactions. Herein, we propose a novel catalyst by developing a p–n junction titania–silicon nanowires (TiO2/SiNWs) heterostructure. The structure is fabricated by atomic layer deposition of TiO2 on p-type SiNWs. The TiO2/SiNWs heterostructure exhibited an outstanding OCM performance under simulated solar light irradiation compared to the single components. This enhanced efficiency was attributed to the intrinsic electrical field formed between n-type TiO2 and p-type SiNWs, which forces generated charge carriers to move in opposite directions and suppresses charge recombination. Besides, surface morphology and optical properties of the the p–n TiO2/SiNWs catalyst are also beneficial for the photocatalytic activity. It is expected that the results of this study will provide massive guidance in synthesizing an efficient photocatalyst for CH4 conversion under mild conditions.

Introduction

Methane (CH4), which can take the form of liquefied natural gas, is one of the crucial sources of industrial chemicals and energy. However, CH4 is a major precursor for tropospheric ozone, causing severe air pollution. Because of its rising atmospheric concentration, CH4 poses a global warming potential approximately thirty fold larger than that of carbon dioxide (CO2) [1-3]. Therefore, it is a challenging mission to eliminate CH4 from the atmosphere. Conventionally, CH4 activation is carried out at high temperatures (>650 °C) via thermal methane conversion to value-added products. However, combustion of CH4 for energy production usually generates great amounts of carbon dioxide as well as coke deposition on catalyst surfaces [4-7]. Therefore, sustainable strategies for both green conversion and atmospheric removal of CH4 are urgently necessary [8-11]. Semiconductor-based photocatalysis has been attracting scientists’ attention because of its environmental friendliness and easy handling [12-14]. Photocatalytic metal oxide semiconductor materials have been utilized for converting solar energy into valuable chemical energy in the field of CH4 conversion [15-17]. Methane oxidation presents a particularly promising strategy. The primary objective is to convert methane into valuable products such as formaldehyde (HCHO), methanol (CH3OH), and other value-added oxygenates, which serve as essential precursors in various manufacturing and production processes [18,19].

The n-type semiconductor titanium dioxide (TiO2) has been discovered as a potential photocatalyst material because of its high stability, good dispersibility, and narrow energy bandgap. However, pristine TiO2 shows only low photocatalytic efficiency because of the high recombination rate between holes and electrons and the low visible-light harvesting ability [20-22]. The rapid recombination of charge carriers prior to their participation in reactions significantly reduces the efficiency of methane oxidation reactions [23,24]. To address these issues of TiO2 nanomaterials, many scientists have developed TiO2-based nanostructure composites as advanced photocatalysts [25-30]. The recombination of charge carriers is mainly attributed to the anisotropic movement of generated electron–hole pairs in semiconductors. Therefore, the implementation of a driving force could remarkably accelerate the oriented motion of electrons and holes, which could suppress recombination and eventually improve photocatalytic efficiency. For years, doping of metal nanoparticles (NPs) into a semiconductor matrix has been extensively studied to enhance photocatalytic CH4 oxidation performance. Metal NPs in, for example, Au/TiO2, Au@Pd/ZnO, and Pt@Cu/TiO2 composites act as electron scavenger centers and own more free electrons for reactions [25-27]. However, the generated electron affinity of metal NPs is sometimes insufficient and cannot prevent recombination or maintain electrons for further reactions.

As an advanced solution for catalysis modification, p–n junction photocatalysts with an intrinsic electric field formed at the interface have emerged, which effectively force charge carriers to move in opposite directions and hinder recombination [31-33]. Very recently, Cu2O/BiVO4, Ag2O/Bi12O17Cl2 and CuFe2O4/Bi4Ti3O12 composite powders have shown improved efficiencies in water treatment based on p–n configuration advantages [34-36]. However, the wetness impregnation synthesis of those powder co-catalysts faces the issues of low surface area, low reproducibility, and difficult control of large-scale production. Therefore, the development of novel catalysts with unique morphologies by using precise tools is extremely essential and important [37-39].

Herein, we constructed a robust p–n junction catalyst by atomic layer deposition (ALD) of TiO2 thin films on a p-type SiNW substrate for enhancing the photocatalytic efficiency in CH4 oxidation. Pristine p-Si wafers have limited surface area and are highly susceptible to mechanical failure because of their brittle nature; in contrast, the etched SiNW arrays exhibit superior optical absorption and enhanced surface catalytic reaction properties. The intimate contact between 1D Si NWs and thin TiO2 layers reduces the recombination rate of electron–hole pairs. Additionally, TiO2/SiNWs offer flexibility, improved bandgap energy, and enhanced light harvesting across a broad spectrum, leading to higher photocatalytic efficiency. Combining SiNWs and TiO2 presents an opportunity to leverage the strengths of both materials while mitigating their respective limitations. This study offers new insights into the design of an efficient system for OCM.

Results and Discussion

Structural and morphological properties

For understanding the crystalline structure of TiO2 and SiNWs, X-ray diffraction patterns were recorded as displayed in Figure 1. The XRD pattern of a pure Si wafer and p-type SiNWs display a main peak at 2θ of 33.2°, which was attributed to the reflection from (200) planes. Despite being etched with concentrated acid and Ag+ ions, there was no significant change in the peak position of p-type SiNWs, which corresponds to the original Si phase (JCPDS No.27-1402) [40,41]. In the case of the as-synthesized composite, the favored growth of TiO2(101) on the surface of p-Si NWs has been noticed [42]. The minor (112), (200), (105), and (211) peaks at 2θ = 38.2°, 48.5°, 53.3°, and 55.1° indicate the formation of anatase TiO2 (JCPDS No.21-1272) [43,44]. As expected, the crystal orientation of the TiO2/SiNWs catalyst obviously led to the creation of a robust p–n junction photocatalyst.

[2190-4286-15-92-1]

Figure 1: XRD patterns of p-Si, p-type SiNWs, and the TiO2/SiNWs sample.

The SEM analysis further confirms the morphological evolution during etching and TiO2 ALD. As revealed in Supporting Information File 1, Figure S1, the morphology of SiNWs is characterized by individual nanowires that bunch together in a highly ordered manner, resulting in well-oriented Si NW arrays perpendicular to the Si bulk surface. Figure 2 shows cross-sectional- and top-view SEM images of the as-prepared TiO2/SiNWs sample. The TiO2/SiNWs arrays were well prepared with an average length of 4 µm. Moreover, the surface of the SiNWs was fully decorated by the TiO2 passivation layer and became blurry. The active pure 25 nm TiO2 layer exhibits flake-like morphology as displayed in Figure 2c,d.

[2190-4286-15-92-2]

Figure 2: Cross-sectional view (a, b) and top view (c, d) SEM images of the TiO2/SiNWs device with a higher-magnification view of the red rectangles on the right-hand side. A thin layer of TiO2 (estimated at 25 nm) was deposited on the SiNWs by using ALD with 500 deposition cycles.

Optical properties

Nanowire arrays offer a better optical absorption than planar Si wafers because of the specific morphology [45]. The optical absorption of the as-prepared catalyst is shown in Figure 3a. The UV–vis diffuse reflection spectrum of TiO2/SiNWs catalyst is drastically reduced in comparison to the pure SiNWs. The superior antireflection property of the TiO2/SiNWs catalyst may be attributed to the vertical wires, which enable strong light scattering leading to enhancement in light harvesting. The optical bandgap values of SiNWs and TiO2/SiNWs are estimated at around 3.8 and 3.3 eV, respectively. Figure 3b displays the current–voltage (IV) curves of the photocatalyst under dark and light conditions. The current of the sample under light conditions is higher than that under dark conditions. The slope of the IV characteristic starts to increase, showing that generated electrons strongly influence the electrical properties of the samples.

[2190-4286-15-92-3]

Figure 3: The UV–visible diffuse reflectance spectra (a) and IV curve under light/dark conditions (b) of the TiO2/SiNWs sample.

The surface interaction with gases during photocatalytic oxidative coupling can be analyzed using water contact angle analysis (as shown in Supporting Information File 1, Figure S2). The wettability of pure p-Si and the p-Si NW array are illustrated in Figure S3 (Supporting Information File 1). Pure p-Si had a water contact angle of 50.24°. Because of the nanowire array morphology, the p-Si NWs were more hydrophilic nature with a water contact angle of 3.36°, which manifests superior photocatalytic oxidative coupling.

Raman spectra were conducted to confirm the surface composition of the synthesized photocatalysts. As depicted in Figure 4a, the Raman spectrum of Si exhibits a single peak located at 519 cm−1, corresponding to the first-order transverse optical (TO) mode of Si [46]. For the TiO2/Si photocatalyst, two distinct peaks were observed, namely, (i) the characteristic Eg vibration of TiO2, located at 146 cm−1, and (ii) the TO phonon mode of Si (Figure 4b) [47-49]. Consequently, the combined surface-sensitive Raman and bulk-sensitive XRD results reveal that the n-type TiO2 coating layer on p-type SiNWs does not influence the crystalline structure.

[2190-4286-15-92-4]

Figure 4: Raman spectra of pure p-Si and Si NW (a), and the TiO2/SiNWs thin film (b).

Photocatalytic OCM

The photocatalytic OCM activity of TiO2/SiNWs sample under aerobic conditions is described in Figure 5. In a batch reactor, the photocatalytic CH4 oxidation progresses as a function of irradiation time. Besides ethane (C2H6), carbon dioxide (CO2) was detected as a by-product. Moreover, propane (C3H8) and H2 were observed as result of the oxidative cross-coupling of methane and ethane (Figure 5a). The conversion reaction of CH4 can be described as follows:

[Graphic 1]

To evaluate the reaction ratio-dependent photocatalytic OCM efficiency, we varied the gas pressure ratios between CH4 and air. As shown in Figure 5b, more CO2 was measured at lower CH4-to-air ratios because of the higher O2 content. Therefore, optimizing the CH4/air ratio is important for improving the OCM reaction. The highest coupling selectivity was around 90% at CH4/air ≈ 4.5:0.5, comparable to or higher than that of typical reported photocatalysts (Table 1). Although the TiO2/Si composite did not achieve the best results compared to other photocatalysts, it remains a viable option for methane oxidation processes.

[2190-4286-15-92-5]

Figure 5: Photocatalytic CH4 oxidation as a function of irradiation time (a) and different CH4/air composition (b) over TiO2/SiNWs. Reaction conditions: TiO2/SiNWs chip (1 × 2 cm2), total pressure of 5 bar, 20 °C, reaction time t = 6 h in (b).

Table 1: Comparison of the OCM performance of the TiO2/Si composite with previous photocatalysts.

Photocatalysts Amount of catalyst Reaction conditions Light Productivity (μmol·g−1·h−1) and selectivity Ref.
Au@Zn2Ti3O8 0.03 g CH4/O2 = 15:1, mild conditions 50 W LED C2H6: 1219
Sel.: 81%
[50]
Pt@TiO2 0.075 g CH4 and water UV lamp C2H6: 57
Sel.: 62%
[51]
AuZnO@TiO2 0.02 g CH4/air = 69:1, mild conditions 300 W Xe lamp C2H6: 5020
Sel.: 90%
[30]
Ag HPW@TiO2 0.1 g CH4 and air, 0.3 MPa 400 W Xe lamp C2H6: 21
Sel.: 90%
[52]
Au@ZnO 0.005 g CH4/O2 = 99:1, mild conditions 2 LEDs C2-C4:684
Sel.: 83%
[53]
TiO2/Si 1 × 2 cm2 CH4/air = 4.5:0.5, 0.5 MPa 300 W Xe lamp C2H6: 210 µmol/cm2 in 6 h
Sel.: 90%
this work

In order to investigate the individual effects of p-type SiNWs and TiO2 layer on the photocatalytic OCM efficiency, comparable samples were irradiated under the same conditions. The methane coupling was analyzed with different samples using pure p-type SiNW, TiO2/SiNW, TiO2/glass catalyst, as shown in Figure 6a. In fact, only negligible CH4 conversion was discovered over p-type SiNWs, and TiO2/glass under the same conditions. It can be explained that the superior wettability of TiO2/SiNWs (hydrophilic surfaces) compared to TiO2/glass (hydrophobic surfaces) enables partial adsorption of water molecules, which facilitates the generation of radicals necessary for photocatalytic reactions. Furthermore, the TiO2 nanostructure and p-type SiNWs are crucial in photogenerated charge separation and adsorption enhancement under UV–vis light. In other words, the ethane productivity of TiO2/Si NWs was five times higher (210 µmol/cm2) than that of pure p-Si NWs (20 µmol/cm2) and glass/TiO2 (30 µmol/cm2).

[2190-4286-15-92-6]

Figure 6: Photocatalytic CH4 oxidation over different types of catalyst (a), and under different types of irradiation (b). Note that only TiO2/SiNWs was used in panel (b). Reaction conditions: 5 bar, 20 °C, t = 6 h.

Photocatalytic OCM over TiO2/Si NWs catalysts was recorded in a batch reactor under different wavelengths of light. As shown in Figure 6b, no products were detected under the visible-light irradiation. The photocatalytic performance under UV illumination was significantly lower than that under full illumination.

Figure 7 shows the recyclability of the p–n TiO2/SiNWs photocatalyst. Note that after each cycle, the reactor was completely evacuated, and fresh gases were refilled for the following run. The photocatalytic activity of p–n TiO2/SiNWs remained almost unchanged after four consecutive reaction cycles, indicating a high recyclability in the batch reactor.

[2190-4286-15-92-7]

Figure 7: Stability of photocatalytic CH4 oxidation reaction over the TiO2/SiNWs sample.

Plausible charge transport mechanism

Figure 8 shows a schematic mechanism of the photocatalytic activity of the p–n TiO2/SiNWs hierarchical structures. According to previous studies the energy bandgaps of p-Si and n-type TiO2 were assumed to be 1.1 eV and 3.3 eV, respectively [54-56]. Because of the different electron affinities (Eea,TiO2 ≈ 4.10 eV, Eea,Si ≈ 4.05 eV) [57,58] the electrons are excited and moved from the valance band minimum (VBM, 0.74V vs NHE) of SiNWs to the conduction band maximum (CBM, 0.35V vs NHE) of TiO2 to enter the equilibrium state under irradiation following Aderson’s model [39]. The photogenerated electrons tend to produce C2H6, C3H8, and H2 from H+. The photogenerated holes at the VBM of TiO2 create oxidized intermediates and H+ [59-61]. The whole process can be expressed by the S-scheme mechanism, as follows:

[Graphic 2]

at the CB of n-type TiO2

[Graphic 3]

and at the VB of p-Si NW

[Graphic 4]

[2190-4286-15-92-8]

Figure 8: Proposed mechanism of photocatalytic CH4 oxidation reaction over p–n junction TiO2/SiNWs sample.

Conclusion

A p–n junction TiO2/SiNWs photocatalyst was synthesized via ALD and utilized for light-driven OCM. TiO2/SiNWs revealed excellent performance owing to the smooth transport of photogenerated electrons in the p–n junction, which lowers the e–h recombination rate. The nanowire array structure of the catalyst provides a surface that can massively increase light absorption, achieving an efficient C2H6 yield of 210 µmol/cm2 in 6 h with high selectivity under light illumination at room temperature. This research could offer new insights into composite photocatalysts for methane coupling.

Experimental

Chemicals and materials

Commercial p-type Si 3-inch wafers (⟨100⟩ orientation, boron-doped, resistivity = 0.01–1 Ω·cm) were purchased from Silicon Mitus Corporation, South Korea. Silver nitrate (AgNO3, 0.1 M), hydrofluoric acid HF (50 wt %), nitric acid (HNO3, 63%), acetone, and ethanol were provided from Sigma-Aldrich. Deionized (DI) water was used for cleaning steps.

Si NWs and TiO2/Si NWs preparation

First, a small piece (1 × 2 cm2) was cut from a commercial p-type Si wafer and washed several times using DI water, ethanol, and acetone in a sonication bath. Etching solution containing AgNO3 (0.1 M), HF (50 wt %) and H2O (2:1:2 vol %) was prepared and kept at 56 °C for 20 min. The clean Si substrate was rapidly immersed in the etching medium and etched by the Ag+ ions for 25 min to obtain 4 µm long SiNWs. Afterwards, remaining Ag on the Si surface was removed using HNO3 (63 wt %) for 10 min. The etched p-Si NWs substrate was eventually washed with DI water and dried under N2 flow, as shown in Figure 9.

[2190-4286-15-92-9]

Figure 9: Schematic illustration of TiO2/SiNWs chip fabrication starting from a p-type silicon wafer. A thin layer of TiO2 was deposited on the Si NWs via ALD.

Second, the as-prepared SiNWs sample was transferred to the chamber of an ALD system (R200 Advanced Picosun, 2013) for TiO2 thin film deposition. The TiO2 thin film deposition was carried out at 300 °C using TiCl4 (98% purity, 0.2 s pulse time) and H2O (0.1 s pulse time) as precursors. The vacuum level of the chamber was kept at 8 × 10−3 bar, and the deposition rate was 0.051 nm/cycle. After cooling down to room temperature, the sample was taken out from chamber for photocatalytic experiments.

Photocatalytic tests

The as-synthesized catalyst was placed in a custom-made batch reactor with a small transparent quartz window, which was directly connected to a gas chromatograph (GC) with thermal conductivity and flame ionization detectors. A 300 W Xenon lamp was utilized as a light source (HAL-320). First, the reactor containing the photocatalytic thin film samples was evacuated using a vacuum pump for 10 min and filled with a mixture of CH4/air (4.5/0.5 pressure ratio). The pressure ratio of the gas mixture was varied to study gas composition-dependent efficiency. The total pressure of gaseous reactants in batch reactor was established at 5 bar, and the photocatalytic system was kept in the dark (20 min) to reach equilibrium before exposure to simulated solar light. The temperature was kept at room temperature via a cooler.

Characterization

The crystalline structure and morphological properties of as-synthesized samples were analyzed using an X-ray diffraction system (XRD, Rigaku, SmartLab) with a 2θ range of 20–80° and a field-emission scanning electron microscope (FE-SEM, Hitachi, S-4700). The absorption properties of the thin films were analyzed using a diffuse reflectance UV–vis spectrometer (DRS-UV, Shimazu UV-2450). The chemical structure of the catalyst surface was analyzed using a Raman spectrometer (excitation of 532 nm, ANDOR Monora 500i). The surface wettability of the thin film sample was measured using a static contact angle system (Biosin Scientific), as shown in Supporting Information File 1, Figure S2. The contact angle between horizontal sample surface and the perimeter of the water drop was measured after 10 s of interaction. The in situ photocurrent measurements were carried out in the presence of gaseous reactants (CH4/air = 4.5/0.5) ranging from −2 V to +3 V under dark and light conditions. Each measurement was scanned with 0.05 V intervals. The bias supply and current signals were provided and recorded by a Keithley system adapted with an amplifier.

Supporting Information

Supporting Information File 1: Additional figures.
Format: PDF Size: 662.8 KB Download

Acknowledgements

We acknowledge Material projects for using the TiO2 and Si CIF files for the structures in the graphical abstract. The TiO2 structure was reproduced from https://next-gen.materialsproject.org/materials/mp-1215, and the Si structure was reproduced from https://next-gen.materialsproject.org/materials/mp-165 (both published by The Materials Project and distributed under the terms of the Creative Commons Attribution 4.0 International License, https://creativecommons.org/licenses/by/4.0). Both structures were combined using the Vesta software (Momma, K.; Izumi, F. J. Appl. Crystallogr. 2011, 44, 1272–1276).

Funding

Ta Thanh Hoai Qui was funded by the Postdoctoral Scholarship Programme of Vingroup Innovation Foundation (VINIF), code VINIF.2023.STS.62.

Competing 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.

Author Contributions

Qui Thanh Hoai Ta: supervision; validation; visualization; writing – original draft. Luan Minh Nguyen: data curation; formal analysis. Ngoc Hoi Nguyen: resources; software. Phan Khanh Thinh Nguyen: supervision; validation; visualization. Dai Hai Nguyen: writing – original draft; writing – review & editing.

Data Availability Statement

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

References

  1. Reay, D.; Smith, P.; van Amstel, A. Methane Sources and the Global Methane Budget. In Methane and Climate Change, 1st ed.; Reay, D.; Smith, P., Eds.; Routledge: Abingdon-on-Thames, United Kingdom, 2010; pp 1–13. doi:10.4324/9781849775090
    Return to citation in text: [1]
  2. Lai, D. Y. F. Pedosphere 2009, 19, 409–421. doi:10.1016/s1002-0160(09)00003-4
    Return to citation in text: [1]
  3. Van Amstel, A. J. Integr. Environ. Sci. 2012, 9, 5–30. doi:10.1080/1943815x.2012.694892
    Return to citation in text: [1]
  4. Zamri, A. A.; Ong, M. Y.; Nomanbhay, S.; Show, P. L. Environ. Res. 2021, 197, 111204. doi:10.1016/j.envres.2021.111204
    Return to citation in text: [1]
  5. Biernat, K.; Malinowski, A.; Gnat, M. The Possibility of Future Biofuels Production Using Waste Carbon Dioxide and Solar Energy; Biofuels - Economy, Environment and Sustainability; InTech: Rijeka, Croatia, 2013. doi:10.5772/53831
    Return to citation in text: [1]
  6. Soeder, D. J. Adv. Geo-Energy Res. 2021, 5, 274–285. doi:10.46690/ager.2021.03.04
    Return to citation in text: [1]
  7. Benson, S. M. In Carbon Dioxide Capture for Storage in Deep Geologic Formations – Results from the CO2 Capture Project: Vol 1-Capture and Separation of Carbon Dioxide from Combustion, Vol 2-Geologic Storage of Carbon Dioxide with Monitoring and Verification; Thomas, D. C., Ed.; Elsevier, 2005. doi:10.1016/b978-008044570-0/50126-4
    Return to citation in text: [1]
  8. Mar, K. A.; Unger, C.; Walderdorff, L.; Butler, T. Environ. Sci. Policy 2022, 134, 127–136. doi:10.1016/j.envsci.2022.03.027
    Return to citation in text: [1]
  9. Collins, W. J.; Webber, C. P.; Cox, P. M.; Huntingford, C.; Lowe, J.; Sitch, S.; Chadburn, S. E.; Comyn-Platt, E.; Harper, A. B.; Hayman, G.; Powell, T. Environ. Res. Lett. 2018, 13, 054003. doi:10.1088/1748-9326/aab89c
    Return to citation in text: [1]
  10. Song, H.; Meng, X.; Wang, S.; Zhou, W.; Wang, X.; Kako, T.; Ye, J. J. Am. Chem. Soc. 2019, 141, 20507–20515. doi:10.1021/jacs.9b11440
    Return to citation in text: [1]
  11. Jiang, Y.; Li, S.; Wang, S.; Zhang, Y.; Long, C.; Xie, J.; Fan, X.; Zhao, W.; Xu, P.; Fan, Y.; Cui, C.; Tang, Z. J. Am. Chem. Soc. 2023, 145, 2698–2707. doi:10.1021/jacs.2c13313
    Return to citation in text: [1]
  12. Zhu, J.; Liao, M.; Zhao, C.; Liu, M.; Han, A.; Zhu, C.; Sun, Y.; Zhao, M.; Ye, S.; Cao, H. Nano Res. 2023, 16, 6402–6443. doi:10.1007/s12274-023-5428-6
    Return to citation in text: [1]
  13. Kim, C.-M.; Jaffari, Z. H.; Abbas, A.; Chowdhury, M. F.; Cho, K. H. J. Hazard. Mater. 2024, 465, 132995. doi:10.1016/j.jhazmat.2023.132995
    Return to citation in text: [1]
  14. Mao, J.; Ta, Q. T. H.; Tri, N. N.; Shou, L.; Seo, S.; Xu, W. Appl. Mater. Today 2023, 35, 101939. doi:10.1016/j.apmt.2023.101939
    Return to citation in text: [1]
  15. Baharudin, L.; Rahmat, N.; Othman, N. H.; Shah, N.; Syed-Hassan, S. S. A. J. CO2 Util. 2022, 61, 102050. doi:10.1016/j.jcou.2022.102050
    Return to citation in text: [1]
  16. Zhou, Z.; Sun, Z.; Duan, L. Curr. Opin. Green Sustainable Chem. 2023, 39, 100721. doi:10.1016/j.cogsc.2022.100721
    Return to citation in text: [1]
  17. Taifan, W.; Baltrusaitis, J. Appl. Catal., B 2016, 198, 525–547. doi:10.1016/j.apcatb.2016.05.081
    Return to citation in text: [1]
  18. Tucci, F. J.; Rosenzweig, A. C. Chem. Rev. 2024, 124, 1288–1320. doi:10.1021/acs.chemrev.3c00727
    Return to citation in text: [1]
  19. Siang, T. J.; Jalil, A. A.; Liew, S. Y.; Owgi, A. H. K.; Rahman, A. F. A. Catal. Rev.: Sci. Eng. 2024, 66, 343–399. doi:10.1080/01614940.2022.2072450
    Return to citation in text: [1]
  20. Qamar, O. A.; Jamil, F.; Hussain, M.; Bae, S.; Inayat, A.; Shah, N. S.; Waris, A.; Akhter, P.; Kwon, E. E.; Park, Y.-K. Chem. Eng. J. 2023, 460, 141734. doi:10.1016/j.cej.2023.141734
    Return to citation in text: [1]
  21. Chen, S.; Hu, Y. H. Catal. Rev.: Sci. Eng. 2023, 1–41. doi:10.1080/01614940.2023.2169451
    Return to citation in text: [1]
  22. Rafique, M.; Hajra, S.; Irshad, M.; Usman, M.; Imran, M.; Assiri, M. A.; Ashraf, W. M. ACS Omega 2023, 8, 25640–25648. doi:10.1021/acsomega.3c00963
    Return to citation in text: [1]
  23. Li, X.; Li, C.; Xu, Y.; Liu, Q.; Bahri, M.; Zhang, L.; Browning, N. D.; Cowan, A. J.; Tang, J. Nat. Energy 2023, 8, 1013–1022. doi:10.1038/s41560-023-01317-5
    Return to citation in text: [1]
  24. Wang, Z.; Zhang, Z.; Wang, Z.; Lu, H.; Wang, L. Appl. Catal., A 2023, 654, 119082. doi:10.1016/j.apcata.2023.119082
    Return to citation in text: [1]
  25. Meng, L.; Chen, Z.; Ma, Z.; He, S.; Hou, Y.; Li, H.-H.; Yuan, R.; Huang, X.-H.; Wang, X.; Wang, X.; Long, J. Energy Environ. Sci. 2018, 11, 294–298. doi:10.1039/c7ee02951a
    Return to citation in text: [1] [2]
  26. Jiang, W.; Low, J.; Mao, K.; Duan, D.; Chen, S.; Liu, W.; Pao, C.-W.; Ma, J.; Sang, S.; Shu, C.; Zhan, X.; Qi, Z.; Zhang, H.; Liu, Z.; Wu, X.; Long, R.; Song, L.; Xiong, Y. J. Am. Chem. Soc. 2021, 143, 269–278. doi:10.1021/jacs.0c10369
    Return to citation in text: [1] [2]
  27. Lang, J.; Ma, Y.; Wu, X.; Jiang, Y.; Hu, Y. H. Green Chem. 2020, 22, 4669–4675. doi:10.1039/d0gc01608j
    Return to citation in text: [1] [2]
  28. Wu, X.; Zhang, H.; Xie, S.; Wang, Y. Chem Catal. 2023, 3, 100437. doi:10.1016/j.checat.2022.10.013
    Return to citation in text: [1]
  29. Li, X.; Xie, J.; Rao, H.; Wang, C.; Tang, J. Angew. Chem., Int. Ed. 2020, 59, 19702–19707. doi:10.1002/anie.202007557
    Return to citation in text: [1]
  30. Song, S.; Song, H.; Li, L.; Wang, S.; Chu, W.; Peng, K.; Meng, X.; Wang, Q.; Deng, B.; Liu, Q.; Wang, Z.; Weng, Y.; Hu, H.; Lin, H.; Kako, T.; Ye, J. Nat. Catal. 2021, 4, 1032–1042. doi:10.1038/s41929-021-00708-9
    Return to citation in text: [1] [2]
  31. Kim, M.; Lee, J.; Je, M.; Heo, B.; Yoo, H.; Choi, H.; Choi, J.; Lee, K. J. Mater. Chem. A 2021, 9, 2239–2247. doi:10.1039/d0ta10062e
    Return to citation in text: [1]
  32. Ding, Y.; Zhang, J.; Yang, Y.; Long, L.; Yang, L.; Yan, L.; Kong, W.; Liu, F.; Lv, F.; Liu, J. Int. J. Hydrogen Energy 2021, 46, 36069–36079. doi:10.1016/j.ijhydene.2021.08.175
    Return to citation in text: [1]
  33. Zhang, J.; Chen, X.; Bai, Y.; Li, C.; Gao, Y.; Li, R.; Li, C. J. Mater. Chem. A 2019, 7, 10264–10272. doi:10.1039/c8ta08199a
    Return to citation in text: [1]
  34. Wang, W.; Huang, X.; Wu, S.; Zhou, Y.; Wang, L.; Shi, H.; Liang, Y.; Zou, B. Appl. Catal., B 2013, 134–135, 293–301. doi:10.1016/j.apcatb.2013.01.013
    Return to citation in text: [1]
  35. Liu, T.; Shi, L.; Wang, Z.; Liu, D. Colloids Surf., A 2022, 632, 127811. doi:10.1016/j.colsurfa.2021.127811
    Return to citation in text: [1]
  36. Zhao, W.; Jin, Y.; Gao, C. H.; Gu, W.; Jin, Z. M.; Lei, Y. L.; Liao, L. S. Mater. Chem. Phys. 2014, 143, 952–962. doi:10.1016/j.matchemphys.2013.10.026
    Return to citation in text: [1]
  37. Kashyap, V.; Kumar, C.; Chaudhary, N.; Saxena, K. Silicon 2023, 15, 1203–1210. doi:10.1007/s12633-022-02068-w
    Return to citation in text: [1]
  38. Sawal, M. H.; Jalil, A. A.; Khusnun, N. F.; Hassan, N. S.; Bahari, M. B. Electrochim. Acta 2023, 467, 143142. doi:10.1016/j.electacta.2023.143142
    Return to citation in text: [1]
  39. Chandra, A.; Ghosh, S.; Sarkar, R.; Sarkar, S.; Chattopadhyay, K. K. Chemosphere 2024, 352, 141249. doi:10.1016/j.chemosphere.2024.141249
    Return to citation in text: [1] [2]
  40. Shao, M.-W.; Shan, Y.-Y.; Wong, N.-B.; Lee, S.-T. Adv. Funct. Mater. 2005, 15, 1478–1482. doi:10.1002/adfm.200500080
    Return to citation in text: [1]
  41. Ta, Q. T. H.; Tran, N. M.; Tri, N. N.; Sreedhar, A.; Noh, J.-S. Chem. Eng. J. 2021, 425, 131437. doi:10.1016/j.cej.2021.131437
    Return to citation in text: [1]
  42. Tesvara, C.; Yousuf, M. R.; Albrahim, M.; Troya, D.; Shrotri, A.; Stavitski, E.; Karim, A. M.; Sautet, P. ACS Catal. 2024, 14, 7562–7575. doi:10.1021/acscatal.4c01018
    Return to citation in text: [1]
  43. Nguyen, T. T. D.; Le, N. C. T.; Nguyen, D.; Huynh, V. T.; Nguyen, L. T. B.; Nguyen, P. T. Vietnam J. Chem. 2023, 61, 38–44. doi:10.1002/vjch.202300053
    Return to citation in text: [1]
  44. Masuda, Y.; Kato, K. J. Ceram. Soc. Jpn. 2009, 117, 373–376. doi:10.2109/jcersj2.117.373
    Return to citation in text: [1]
  45. Park, Y.-S.; Lee, J. S. ACS Photonics 2017, 4, 2587–2594. doi:10.1021/acsphotonics.7b01018
    Return to citation in text: [1]
  46. Elfiad, A.; Benredouane, S.; Naama, S.; Moulai, F.; Friha, S.-E.; Hadjersi, T. J. Photochem. Photobiol., A 2024, 451, 115503. doi:10.1016/j.jphotochem.2024.115503
    Return to citation in text: [1]
  47. Vemula, S. K.; Katta, V. S.; Ojha, S.; Singh, F.; Raavi, S. S. K. Opt. Mater. (Amsterdam, Neth.) 2024, 148, 114947. doi:10.1016/j.optmat.2024.114947
    Return to citation in text: [1]
  48. Chen, Q.; Wang, Z.; Liao, H.; Li, R.; Feng, Z.; Li, Z.; Lin, L.; Li, J.; Chen, G. Chem. Eng. J. 2024, 493, 152316. doi:10.1016/j.cej.2024.152316
    Return to citation in text: [1]
  49. Cai, J.; Wang, Z.; Jia, S.; Feng, Z.; Ren, Y.; Lin, L.; Chen, G.; Zheng, Z. ACS Appl. Mater. Interfaces 2022, 14, 13703–13712. doi:10.1021/acsami.1c23939
    Return to citation in text: [1]
  50. Huang, Q.; Cai, J.; Wei, F.; Fan, Y.; Liang, Z.; Liu, K.; Lu, X. F.; Ding, Z.; Wang, S. J. Mater. Chem. A 2024, 12, 21334–21340. doi:10.1039/d4ta02049a
    Return to citation in text: [1]
  51. Yu, L.; Shao, Y.; Li, D. Appl. Catal., B 2017, 204, 216–223. doi:10.1016/j.apcatb.2016.11.039
    Return to citation in text: [1]
  52. Yu, X.; Zholobenko, V. L.; Moldovan, S.; Hu, D.; Wu, D.; Ordomsky, V. V.; Khodakov, A. Y. Nat. Energy 2020, 5, 511–519. doi:10.1038/s41560-020-0616-7
    Return to citation in text: [1]
  53. Yang, X.; Chen, M.; Weng, C.; Zhuge, D.; Jin, F.; Xiao, Y.; Tian, D.; Yin, Q.; Li, L.; Zhang, X.; Shi, G.; Lu, X.; Yan, L.; Wang, L.; Wen, B.; Zhao, Y.; Lin, J.; Wang, F.; Zhang, W.; Chen, Y. Adv. Sci. 2024, 11, 2470173. doi:10.1002/advs.202470173
    Return to citation in text: [1]
  54. Somvanshi, D.; Jit, S. IEEE Trans. Nanotechnol. 2014, 13, 62–69. doi:10.1109/tnano.2013.2290553
    Return to citation in text: [1]
  55. Tran Huu, H.; Ngo Van, N.; Vo, V.; Nguyen Thi, L. Vietnam J. Catal. Adsorpt. 2021, 10, 56–62. doi:10.51316/jca.2021.090
    Return to citation in text: [1]
  56. Rawat, G.; Somvanshi, D.; Kumar, H.; Kumar, Y.; Kumar, C.; Jit, S. IEEE Trans. Nanotechnol. 2016, 15, 193–200. doi:10.1109/tnano.2015.2512565
    Return to citation in text: [1]
  57. Sze, S. M.; Li, Y.; Ng, K. K. Physics of Semiconductor Devices, 4th ed.; John Wiley & Sons, 2021. doi:10.1002/0470068329
    Return to citation in text: [1]
  58. Chen, J.-Z.; Chen, T.-H.; Lai, L.-W.; Li, P.-Y.; Liu, H.-W.; Hong, Y.-Y.; Liu, D.-S. Materials 2015, 8, 4273–4286. doi:10.3390/ma8074273
    Return to citation in text: [1]
  59. Liu, Y.; Ji, G.; Dastageer, M. A.; Zhu, L.; Wang, J.; Zhang, B.; Chang, X.; Gondal, M. A. RSC Adv. 2014, 4, 56961–56969. doi:10.1039/c4ra10670a
    Return to citation in text: [1]
  60. Gulati, S.; Mulvehill, M. C.; Thompson, T. C.; Spurgeon, J. M. Energy Fuels 2023, 37, 8573–8582. doi:10.1021/acs.energyfuels.3c00568
    Return to citation in text: [1]
  61. Yao, X.; Chen, L.; Liu, M.; Feng, D.; Wang, C.; Lu, F.; Wang, W.; Wang, X.; Cheng, Y.; Liu, H.; Chen, H.; Wang, W. Appl. Catal., B 2018, 221, 70–76. doi:10.1016/j.apcatb.2017.08.087
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities