Alkyl group-decorated g-C3N4 for enhanced gas-phase CO2 photoreduction

Chao Yang a, Yanting Hou a, Guoqiang Luo b, Jiaguo Yu a and Shaowen Cao *a
aState Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China. E-mail: swcao@whut.edu.cn
bChaozhou Branch of Chemistry and Chemical Engineering Guangdong Laboratory, Chaozhou 521000, China

Received 10th May 2022 , Accepted 27th July 2022

First published on 27th July 2022


Abstract

With excellent physical/chemical stability and feasible synthesis, g-C3N4 has attracted much attention in the field of photocatalysis. However, its weak photoactivity limits its practical applications. Herein, by easily planting hydrophobic alkyl groups onto g-C3N4, the hydrophilicity of g-C3N4 can be well regulated and its specific surface area be enlarged simultaneously. Such a modification ensures enhanced CO2 capture and increased active sites. In addition, the introduction of alkyl groups endows g-C3N4 with abundant charge density and efficient separation of photoinduced excitons. All these advantages synergistically contribute to the enhanced photocatalytic CO2 reduction performance over the optimized catalyst (DCN90), and the total CO2 conversion is 7.4-fold that of pristine g-C3N4 (CN).


Introduction

Photocatalytic CO2 reduction into high value-added chemical fuels is able to simultaneously mitigate the increasingly severe energy crisis and environmental problems, which is considered as one of the photochemical reactions with the most potential.1–7 To apply photocatalytic CO2 reduction technology in practice, the main research aim is to seek the proper photocatalyst. In recent decades, polymeric g-C3N4 has received great attention as a promising photocatalyst because of its excellent physical and chemical stability, easy synthesis methods, low cost, abundant resources, and visible-light response, etc.8–10 Nevertheless, due to the fast recombination of photogenerated charge carriers, insufficient active sites, and weak CO2 adsorption, the CO2 reduction performance of g-C3N4 is commonly unsatisfactory.11–14

To promote the photocatalytic activity of g-C3N4, plentiful strategies have been developed so far. For example, crafting heterojunctions or loading cocatalysts is an effective approach to boost the separation of photogenerated charge carriers.15–17 To provide more active sites, designing unique nanostructures is obviously crucial, which also favors the process of mass transfer including CO2 capture and the desorption of reduction products.18 Recently, upon introducing electron donor or acceptor units into the framework of g-C3N4, the obtained donor–acceptor-based g-C3N4 has demonstrated enhanced light-harvesting capacity and efficient separation of charges due to the strong electron push–pull effect.19,20

Although these methods play a positive role in improving photocatalytic CO2 reduction performance, there still exists an imperative scientific issue that is usually neglected. As is known, during gas-phase CO2 photoreduction, the simultaneous adsorption of H2O and CO2 molecules on the surface of the photocatalyst jointly determines the reaction rate.21,22 Specifically, the oxidation of H2O molecules to provide protons is considered as a rate-determining step. A beneficial H2O adsorption process can accelerate the oxidation of H2O for the fast supplementation of protons. However, excessive H2O adsorption will occupy the CO2 adsorption sites, resulting in the weakening of the CO2 hydrogenation process.23,24 Therefore, a good adsorption balance between CO2 and H2O molecules is very significant, which is rarely investigated now.

Herein, we achieve the facile regulation of the hydrophilicity of g-C3N4 by planting hydrophobic alkyl groups. It is found that the decline in the hydrophilicity of g-C3N4 trades off against the enhanced CO2 adsorption. In addition, the specific surface area of g-C3N4 is also enlarged, thus providing more active sites. Most importantly, the introduction of alkyl groups increases the electron density of g-C3N4 and promotes the dissociation of photogenerated charge carriers. These enhancements synergistically improve the photocatalytic CO2 reduction performance, and the total CO2 conversion over the optimal sample (DCN90) is 7.4-fold that of pristine g-C3N4. This research offers a straightforward and effective strategy for the design of highly efficient g-C3N4-based photocatalysts for energy conversion.

Results and discussion

Fig. 1a shows the synthesis route of alkyl group-functionalized g-C3N4. Under illumination, a series of radical reactions occur on the surface of g-C3N4 (CN), which can trigger the occurrence of additional reactions between CN and 1-decene (Fig. S1),25 thus successfully obtaining hydrophobic alkyl group-grafted g-C3N4 (DCNT, T is the irradiation time, T = 30, 60, 90, and 120 min). The X-ray diffraction (XRD) patterns and Fourier transform infrared spectra (FTIR) were recorded to investigate the chemical structures of the samples (Fig. S2 and S3). Compared with CN, the XRD peak position and the FTIR spectra of DCNT show no obvious change, indicating that the chemical structure of CN is well retained after planting alkyl groups.26 Notably, the DCNT samples possess wider XRD peaks (∼27°) than those of CN, which might result from the insertion of alkyl groups into the interlayer of CN.27
image file: d2nr02551e-f1.tif
Fig. 1 (a) Synthesis route of alkyl group-functionalized g-C3N4 with controlled hydrophilicity. High-resolution XPS (b) C 1s and (c) O 1s spectra of CN and DCN90.

To further analyze the chemical structures of the samples, X-ray photoelectron spectra (XPS) were recorded. Fig. S4a shows that both CN and DCN90 mainly consist of the C, N, and O elements and the O content in the DCN90 sample is lower than that of CN. As shown in Fig. S4b, the high-resolution N 1s spectra of CN and DCN90 can be fitted into three peaks, respectively, attributed to sp2-hybridized nitrogen (C–N[double bond, length as m-dash]C), tri-coordinated nitrogen (N–C3), and the terminal amino groups (–NHx).28–30 From Fig. 1b, the high-resolution C 1s spectrum of CN consists of three sub-peaks, indexed to sp2-hybridized carbon (N–C[double bond, length as m-dash]N), C–OH from adsorbed H2O, and adventitious carbon impurities, respectively.31–33 As for the DCN90 sample, a new peak located at 283.9 eV appears, which is assigned to sp3-hybridized C–H from the alkyl groups.34 Meanwhile, the peak assigned to the adsorbed H2O disappears. These results suggest that the hydrophobic alkyl groups have been successfully grafted onto the surface of CN via photoinduced grafting, thus leading to a decrease in H2O adsorption, which can be further confirmed by the weaker high-resolution O 1s spectrum for DCN90 (Fig. 1c).35 In addition, the peak positions of N–C[double bond, length as m-dash]N and C–N[double bond, length as m-dash]C shift to lower binding energies for DCN90 compared with CN, which is attributed to the increased electron density of g-C3N4 after grafting alkyl groups.

The introduction of hydrophobic alkyl groups could exert an imperative effect on the physical and chemical properties of CN. Therefore, the hydrophilicity of the samples was first explored. As shown in Fig. 2a, with prolonged photoinduced grafting time, the water contact angles of the samples gradually increase from 60° (CN) to 83° (DCN90). This result indicates a slightly decreased H2O adsorption capacity. The field emission scanning electron microscopic (FESEM) images show that CN exhibits a nanosheet-like morphology (Fig. S5a). After grafting the alkyl groups, the morphology of CN is still retained (Fig. S5b–e), which can be further confirmed by the transmission electron microscopic (TEM) images (Fig. S6). Atomic force microscopy (AFM) was applied to measure the thickness of the samples.36Fig. 2b provides the AFM image of CN and the corresponding height profile shows a thickness of 61.7 nm (the inset of Fig. 2b). Compared with CN, the DCN90 sample is thinner with a thickness of around 10 nm (Fig. 2c and the inset of Fig. 2c). This can be explained by the insertion of alkyl groups into the interlayer of CN, resulting in the exfoliation of CN, which is consistent with the results of XRD.


image file: d2nr02551e-f2.tif
Fig. 2 (a) Water contact angles of the samples. AFM images and the corresponding height profiles (insets) of (b) CN and (c) DCN90. (d) N2 adsorption–desorption isotherms, the corresponding pore size distribution curves (inset of (d)), and (e) the CO2 uptake curves of CN and DCN90.

In general, the thinner thinness of g-C3N4 signifies the larger specific surface area (SBET).37 To confirm this point, the N2 adsorption–desorption isotherms of CN and DCN90 were measured. As shown in Fig. 2d, both CN and DCN90 exhibit type-IV isotherms with an H3-type hysteresis loop.38 The corresponding pore size distribution curves of CN and DCN90 mainly display the mesopores and macropores. Compared with CN, the DCN90 sample possesses a higher N2 adsorption capacity and pore volume (inset of Fig. 2d), thus showing a larger SBET for DCN90 (66 m2 g−1) than for CN (48 m2 g−1). As is known, the larger SBET favors providing more adsorbed and active sites.39,40 Accordingly, the CO2 adsorption performance of CN and DCN90 was further tested. Fig. 2e shows that the CO2 uptake amount of DCN90 is 0.28 mmol g−1 at a relative pressure (P/P0) of 1, obviously higher than that of CN (0.16 mmol g−1). The slightly decreased hydrophilicity and enhanced CO2 adsorption might boost the adsorption balance between the H2O and CO2 molecules,21 which is beneficial for the progress of the interfacial CO2 reduction reaction.

To analyze the electronic structures of the samples, the UV–vis diffuse reflectance spectra (DRS) were first recorded to investigate the light-harvesting ability of the samples. As shown in Fig. 3a, upon grafting the hydrophobic alkyl groups, all the samples show enhanced light absorption in the UV region. However, in the visible region, the light absorption performance of samples shows no obvious change. In addition, the absorption edges of samples even display a slight blue-shift, which can be assigned to the typical quantum confinement effect,41 resulting from the decreased thickness of the samples by inserting the alkyl groups into the interlayer of g-C3N4. The absorption edges of CN and DCN90 are respectively 450 and 439 nm, and the corresponding band gaps (Eg) are 2.75 and 2.82 eV. The Mott–Schottky curves were analysed to determine the conduction band potential (ECB) of the samples (Fig. S7). The values of ECB for CN and DCN90 are respectively −1.29 and −1.44 V (vs. Ag/AgCl, pH = 6.7), which can be transformed to −1.11 and −1.26 V (vs. NHE, pH = 7) using the conversion formulae S1 and S2.42 Based on their Eg, the valence band potential (EVB) of CN and DCN90 is 1.64 and 1.56 V, respectively. Fig. 3b illustrates the band structure of CN and DCN90. The more negative ECB for DCN90 allows for a stronger electron driving force for CO2 photoreduction.43


image file: d2nr02551e-f3.tif
Fig. 3 (a) UV–vis DRS spectra of various samples. (b) Electronic band structure, (c) EPR spectra, (d) time-resolved PL spectra, (e) steady-state PL spectra, and (f) transient photocurrent response of the CN and DCN90 samples.

As is known, the separation of photoinduced charge carriers plays a critical role in a photocatalytic reaction.44 Consequently, the electron paramagnetic resonance (EPR) spectra were first recorded to investigate the charge density of the CN and DCN90 samples. Both CN and DCN90 exhibit a single Lorentzian line (g = 2.009) signal (Fig. 3c), attributed to the unpaired electrons from the aromatic heptazine ring.45 Compared with CN, the EPR signal of DCN90 is obviously enhanced. This is due to the introduction of alkyl groups because the alkyl groups act as electron donors for supplying more delocalized electrons.46 Furthermore, the time-resolved photoluminescence (PL) spectra were recorded to investigate the PL lifetime of the samples (Fig. 3d). According to the tri-exponential decay fit (Table 1), the PL lifetimes of CN and DCN90 are calculated to be 28.2 and 22.1 ns, respectively. The decreased lifetime indicates a faster electron transfer for DCN90 than CN, which could be attributed to the establishment of a built-in electric field between alkyl groups and g-C3N4.47 The steady-state PL spectra show that the PL intensity of DCN90 is obviously weakened compared with CN (Fig. 3e), suggesting the more efficient separation of photogenerated charges for DCN90, which is attributed to the decreased thickness of DCN90 for quick electron transfer from the bulk to the surface.48–50 In addition, the PL peak of DCN90 shows a blue shift compared with CN, resulting from the quantum confinement effect due to the decreased thickness of DCN90. The photocurrent tests show that DCN90 possesses a stronger photocurrent signal than CN. When the photocurrent signal is steady, DCN90 shows a slower decay of the photocurrent signal compared with CN whether the light source is on or off (Fig. 3f). This result indicates that DCN90 possesses a slower recombination of photogenerated charges than CN. The decreased electrochemical impedance for DCN90 further confirms the improved transfer of photogenerated charges after grafting alkyl groups (Fig. S8).51 The more delocalized electrons and efficient separation of charges demonstrate the higher photocatalytic performance.

Table 1 Summary of the photoluminescence decay time (τ) and their pre-exponential factor (B) of the CN and DCN90 samples. The calculation formula of average lifetime is as follows: τave = (B1τ12 + B2τ22 + B3τ32)/(B1τ1 + B2τ2 + B3τ3)
Samples Decay time (ns) Pre-exponential factor (%) τ ave (ns)
τ 1 τ 2 τ 3 B 1 B 2 B 3
CN 1.59 7.44 35.56 20.87 47.70 31.43 28.2
DCN90 1.08 5.29 27.67 20.41 48.55 31.04 22.1


To evaluate the photocatalytic activity of the samples, CO2 reduction tests were carried out under full spectrum irradiation without any cocatalyst or sacrificial reagent. Fig. 4a shows that CO is the primary CO2 reduction product, along with a small amount of CH4. This is because the reduction of CO2 into CO follows the two-electron reduction process, which is easier than the eight-electron reduction of CO2 into CH4. Unmodified CN shows a very low photocatalytic performance, and the yields of CO and CH4 are 2.61 and 0.20 μmol h−1 g−1, respectively. After grafting hydrophobic alkyl groups onto g-C3N4, the photocatalytic activity of samples is greatly enhanced. In particular, DCN90 exhibits an optimal CO2 reduction performance with yields of 19.77 μmol h−1 g−1 for CO and 1.09 μmol h−1 g−1 for CH4, which are higher than many reported results (Table 2). The total CO2 conversion of DCN90 is 7.4-fold that of CN.


image file: d2nr02551e-f4.tif
Fig. 4 (a) Photocatalytic CO2 reduction performance of various samples. (b) Cycle tests and (c) control experiments of CO2 photoreduction over the DCN90 sample.
Table 2 Comparison of the photocatalytic activity and the reaction conditions with other g-C3N4-based photocatalysts for CO2 reduction (LS, CS, TOR, SR, AOP, TEOA, and MeOH represent the light source, the CO2 source, the type of reaction, the sacrificial reagent, the amount of photocatalyst, triethanolamine, and methanol, respectively)
Materials LS CS TOR SR AOP (mg) Product yield (μmol h−1 g−1) Ref.
DCN90 300 W Xe lamp (full spectrum) CO2 gas Gas–solid 50 CO: 19.77 This work
CH4: 1.09
g-CN-0.01Dbc 300 W Xe lamp CO2 gas Liquid–solid 20 CO: 2.4 19
g-C3N4/7Ag/m-CeO2 8 W UV lamp CO2 gas Liquid–solid TEOA 100 CO: 1.39 52
CH4: 0.74
OCCN0.25 4 W UV lamp (254 nm, 40 μW cm−2) CO2 gas Liquid–solid 20 CO: 8.74 53
CMN 300 W Xe lamp CO2 gas Gas–solid 30 CO: 18.8 54
CH4: 1.8
10TC 300 W Xe lamp (λ ≥ 420 nm) NaHCO3 + H2SO4 Gas–solid 20 CO: 5.19 16
CH4: 0.044
CFC-0.2 350–780 nm CO2 gas Gas–solid 50 CO: 8.182 55
CH4: 0.0805
2Au-CN 300 W Xe lamp (full spectrum) CO2 gas Gas–solid 50 CO: 6.585 56
CH4: 1.55
3% Ni/NiO/g-C3N4 300 W Xe lamp (full spectrum) CO2 gas Gas–solid 50 CO: 13.955 57
CH4: 2.08
Ni5-CN 300 W Xe lamp (λ ≥ 420 nm) CO2 gas Gas–solid 25 CO: 8.6 58
CH4: 0.5
CABB@C3N4-82% AM1.5G (150 mW cm−2) CO2 gas Liquid–solid MeOH 15 CO: <0.75 59
CH4: <1.35
5BSCN 300 W Xe lamp (full spectrum) NaHCO3 + H2SO4 Gas–solid 50 CO: 8.2 15
BiCN-0.6 AM1.5G CO2 gas Liquid–solid 50 CO: 3.78 60
CH4: 1.65
CZTS-CN 400 W Xe lamp (λ ≥ 420 nm) CO2 gas Gas–solid 100 CO: 2.402 61
CH4: 0.752


In addition, photocatalytic cycle tests were conducted to investigate the stability of DCN90. After three consecutive runs, 81.7% of the total CO2 conversion performance of DCN90 is still retained with no obvious change in morphology, hydrophilicity, and chemical structure (Fig. 4b and S9–S11). A certain degree of decreased activity might be caused by the occupation of active sites due to the adsorption of intermediates on the surface of DCN90. Finally, control experiments were carried out to determine the carbon source of the reduction products. From Fig. 4c, without a photocatalyst and under a N2 atmosphere, a little CO is observed and no CH4 is detected. Furthermore, with DCN90 irradiated under a N2 atmosphere, a very similar result is obtained. The small amounts of CO produced might originate from the pollution of the apparatus.62 However, when DCN90 is irradiated under a CO2 atmosphere, a rapidly enhanced CO yield is observed, and CH4 is also detected. These results strongly affirm that CO and CH4 are produced by CO2 photoreduction.

Based on the above results of CO2 photoreduction and the characterization of the physical and photoelectrochemical properties of the samples, a rational photocatalytic mechanism is proposed. First, under irradiation, the electrons in VB are excited and jump into CB. As for the hydrophobic alkyl group-decorated DCN90, a more negative ECB allows for a stronger electron reduction ability. Furthermore, the electron-rich alkyl groups can increase the charge density of g-C3N4 and boost the separation of photogenerated charge carriers. In addition, due to the insertion effect of the hydrophobic alkyl groups, g-C3N4 becomes thinner, which enlarges the specific surface area to provide more active sites. Meanwhile, the hydrophilicity of g-C3N4 is also regulated to optimize the simultaneous adsorption of H2O and CO2 molecules. All these factors synergistically endow DCN90 with a more favorable photoelectronic nature, thus leading to greatly enhanced photocatalytic activity compared with pristine g-C3N4.

Conclusions

In summary, hydrophobic alkyl group-decorated g-C3N4 is prepared by an easy photoinduced grafting method. The optimal photocatalyst (DCN90) demonstrates excellent photocatalytic CO2 reduction activity. The yields of CO and CH4 reach 19.77 and 1.09 μmol h−1 g−1, respectively, and the total CO2 conversion is 7.4-fold that of pristine g-C3N4. The promotion of CO2 photoreduction performance is mainly attributed to the grafting of hydrophobic alkyl groups. On the one hand, alkyl groups regulate the band structure, increase the electron density, and boost the separation of photoinduced charge carriers of g-C3N4. On the other hand, alkyl groups are inserted into the interlayer of g-C3N4, enlarging the specific surface area to provide more active sites and optimizing the adsorption of H2O and CO2 molecules. All these factors synergistically endow hydrophobic alkyl group-modified g-C3N4 with enhanced photocatalytic activity. This work offers a straightforward yet effective strategy to design highly efficient g-C3N4-based photocatalysts for photocatalytic energy conversion.

Author contributions

Chao Yang: conceptualization, experiments and data processing, formal analysis, and writing – original draft. Yanting Hou: experiments and formal analysis. Guoqiang Luo: resource and writing – review & editing. Jiaguo Yu: resource and writing – review & editing. Shaowen Cao: formal analysis, resource, writing – review & editing and supervision.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was financially supported by the Key-Area Research and Development Program of Guangdong Province (2021B0707050001), the National Natural Science Foundation of China (51922081, 51961135303, and 51932007), and the Self-Innovation Research Funding Project of Hanjiang Laboratory (HJL202202A001).

References

  1. Y. S. Zhou, Z. T. Wang, L. Huang, S. Zaman, K. Lei, T. Yue, Z. A. Li, B. You and B. Y. Xia, Engineering 2D photocatalysts toward carbon dioxide reduction, Adv. Energy Mater., 2021, 11, 2003159 CrossRef CAS.
  2. M. Sayed, F. Y. Xu, P. Y. Kuang, J. X. Low, S. Y. Wang, L. Y. Zhang and J. G. Yu, Sustained CO2-photoreduction activity and high selectivity over Mn, C-codoped ZnO core-triple shell hollow spheres, Nat. Commun., 2021, 12, 4936 CrossRef CAS PubMed.
  3. M. Sayed, J. G. Yu, G. Liu and M. Jaroniec, Non-noble plasmonic metal-based photocatalysts, Chem. Rev., 2022, 122, 10484–10537 CrossRef CAS PubMed.
  4. X. G. Fei, H. Y. Tan, B. Cheng, B. C. Zhu and L. Y. Zhang, 2D/2D black phosphorus/g-C3N4 S-scheme heterojunction photocatalysts for CO2 reduction investigated using DFT calculations, Acta Phys.-Chim. Sin., 2021, 37, 2010027 Search PubMed.
  5. S. Wageh, A. A. Al-Ghamdi and L. J. Liu, S-scheme heterojunction photocatalyst for CO2 photoreduction, Acta Phys.-Chim. Sin., 2021, 37, 2010024 Search PubMed.
  6. X. Y. Jiang, J. D. Huang, Z. H. Bi, W. J. Ni, G. Gurzadyan, Y. A. Zhu and Z. Y. Zhang, Plasmonic active “hot spots”-confined photocatalytic CO2 reduction with high selectivity for CH4 production, Adv. Mater., 2022, 34, 2109330 CrossRef CAS PubMed.
  7. N. Lu, M. Y. Zhang, X. D. Jing, P. Zhang, Y. A. Zhu and Z. Y. Zhang, Electrospun semiconductor–based nano-heterostructures for photocatalytic energy conversion and environmental remediation: Opportunities and challenges, Energy Environ. Mater., 2022 DOI:10.1002/eem2.12338.
  8. Z. F. Chen, S. C. Lu, Q. L. Wu, F. He, N. Q. Zhao, C. N. He and C. S. Shi, Salt-assisted synthesis of 3D open porous g-C3N4 decorated with cyano groups for photocatalytic hydrogen evolution, Nanoscale, 2018, 10, 3008–3013 RSC.
  9. C. L. Zhu, T. Wei, Y. Wei, L. Wang, M. Lu, Y. P. Yuan, L. S. Yin and L. Huang, Unravelling intramolecular charge transfer in donor–acceptor structured g-C3N4 for superior photocatalytic hydrogen evolution, J. Mater. Chem. A, 2021, 9, 1207–1212 RSC.
  10. X. Y. Jiang, Z. Y. Zhang, M. H. Sun, W. Z. Liu, J. D. Huang and H. Y. Xu, Self-assembly of highly-dispersed phosphotungstic acid clusters onto graphitic carbon nitride nanosheets as fascinating molecular-scale Z-scheme heterojunctions for photocatalytic solar-to-fuels conversion, Appl. Catal., B, 2021, 281, 119473 CrossRef CAS.
  11. X. Zhao, Y. Y. Fan, W. S. Zhang, X. J. Zhang, D. X. Han, L. Niu and A. Ivaska, Nanoengineering construction of Cu2O nanowire arrays encapsulated with g-C3N4 as 3D spatial reticulation all-solid-state direct Z-scheme photocatalysts for photocatalytic reduction of carbon dioxide, ACS Catal., 2020, 10, 6367–6376 CrossRef CAS.
  12. P. F. Xia, B. C. Zhu, J. G. Yu, S. W. Cao and M. Jaroniec, Ultra-thin nanosheet assemblies of graphitic carbon nitride for enhanced photocatalytic CO2 reduction, J. Mater. Chem. A, 2017, 5, 3230–3238 RSC.
  13. Y. Xia, Z. H. Tian, T. Heil, A. Y. Meng, B. Cheng, S. W. Cao, J. G. Yu and M. Antonietti, Highly selective CO2 capture and its direct photochemical conversion on ordered 2D/1D heterojunctions, Joule, 2019, 3, 2792–2805 CrossRef CAS.
  14. Q. L. Xu, Z. H. Xia, J. M. Zhang, Z. Y. Wei, Q. Guo, H. L. Jin, H. Tang, S. Z. Li, X. C. Pan, Z. Su and S. Wang, Recent advances in solar-driven CO2 reduction over g–C3N4-based photocatalysts, Carbon Energy, 2022 DOI:10.1002/cey2.205.
  15. Y. H. Huang, K. Wang, T. Guo, J. Li, X. Y. Wu and G. K. Zhang, Construction of 2D/2D Bi2Se3/g-C3N4 nanocomposite with high interfacial charge separation and photo-heat conversion efficiency for selective photocatalytic CO2 reduction, Appl. Catal., B, 2020, 277, 119232 CrossRef CAS.
  16. C. Yang, Q. Y. Tan, Q. Li, J. Zhou, J. J. Fan, B. Li, J. Sun and K. L. Lv, 2D/2D Ti3C2 MXene/g-C3N4 nanosheets heterojunction for high efficient CO2 reduction photocatalyst: Dual effects of urea, Appl. Catal., B, 2020, 268, 118738 CrossRef CAS.
  17. D. P. Dong, C. X. Yan, J. D. Huang, N. Lu, P. Y. Wu, J. Wang and Z. Y. Zhang, An electron-donating strategy to guide the construction of MOF photocatalysts toward co-catalyst-free highly efficient photocatalytic H2 evolution, J. Mater. Chem. A, 2019, 7, 24180–24185 RSC.
  18. J. Wang, S. W. Cao and J. G. Yu, Nanocages of polymeric carbon nitride from low-temperature supramolecular preorganization for photocatalytic CO2 reduction, Sol. RRL, 2020, 4, 1900469 CrossRef CAS.
  19. X. H. Song, X. Y. Zhang, M. Wang, X. Li, Z. Zhu, P. W. Huo and Y. S. Yan, Fabricating intramolecular donor–acceptor system via covalent bonding of carbazole to carbon nitride for excellent photocatalytic performance towards CO2 conversion, J. Colloid Interface Sci., 2021, 594, 550–560 CrossRef CAS PubMed.
  20. S. J. Wan, J. S. Xu, S. W. Cao and J. G. Yu, Promoting intramolecular charge transfer of graphitic carbon nitride by donor–acceptor modulation for visible-light photocatalytic H2 evolution, Interdiscip. Mater., 2022, 1, 294–308 CrossRef.
  21. Y. Xia, K. Xiao, B. Cheng, J. G. Yu, L. Jiang, M. Antonietti and S. W. Cao, Improving artificial photosynthesis over carbon nitride by gas–liquid–solid interface management for full light-induced CO2 reduction to C1 and C2 fuels and O2, ChemSusChem, 2020, 13, 1730–1734 CrossRef CAS PubMed.
  22. Q. H. Zhang, Y. Xia and S. W. Cao, “Environmental phosphorylation” boosting photocatalytic CO2 reduction over polymeric carbon nitride grown on carbon paper at air–liquid–solid joint interfaces, Chin. J. Catal., 2021, 42, 1667–1676 CrossRef CAS.
  23. D. Raciti, M. Mao, J. H. Park and C. Wang, Mass transfer effects in CO2 reduction on Cu nanowire electrocatalysts, Catal. Sci. Technol., 2018, 8, 2364–2369 RSC.
  24. A. Li, Q. Cao, G. Y. Zhou, B. V. K. J. Schmidt, W. J. Zhu, X. T. Yuan, H. L. Huo, J. L. Gong and M. Antonietti, Three-phase photocatalysis for the enhanced selectivity and activity of CO2 reduction on a hydrophobic surface, Angew. Chem., Int. Ed., 2019, 58, 14549–14555 CrossRef CAS PubMed.
  25. B. Kumru, M. Antonietti and B. V. K. J. Schmidt, Enhanced dispersibility of graphitic carbon nitride particles in aqueous and organic media via a one-pot grafting approach, Langmuir, 2017, 33, 9897–9906 CrossRef CAS PubMed.
  26. C. Yang, S. S. Zhang, Y. Huang, K. L. Lv, S. Fang, X. F. Wu, Q. Li and J. J. Fan, Sharply increasing the visible photoreactivity of g-C3N4 by breaking the intralayered hydrogen bonds, Appl. Surf. Sci., 2020, 505, 144654 CrossRef CAS.
  27. F. Yang, D. Z. Liu, Y. X. Li, L. J. Cheng and J. H. Ye, Salt-template-assisted construction of honeycomb-like structured g-C3N4 with tunable band structure for enhanced photocatalytic H2 production, Appl. Catal., B, 2019, 240, 64–71 CrossRef CAS.
  28. L.-Z. Qin, Y.-Z. Lin, Y.-C. Dou, Y.-J. Yang, K. Li, T. Li and F.-T. Liu, Toward enhanced photocatalytic activity of graphite carbon nitride through rational design of noble metal-free dual cocatalysts, Nanoscale, 2020, 12, 13829–13837 RSC.
  29. H. Li, F. Li, J. G. Yu and S. W. Cao, 2D/2D FeNi-LDH/g-C3N4 hybrid photocatalyst for enhanced CO2 photoreduction, Acta Phys.-Chim. Sin., 2020, 37, 2010073 Search PubMed.
  30. S. Y. Gao, S. J. Wan, J. G. Yu and S. W. Cao, Donor–acceptor modification of carbon nitride for enhanced photocatalytic hydrogen evolution, Adv. Sustainable Syst., 2022, 6, 2200130 CrossRef.
  31. M.-M. Liu, S.-M. Ying, B.-G. Chen, H.-X. Guo and X.-G. Huang, Ag@g-C3N4 nanocomposite: An efficient catalyst inducing the reduction of 4-nitrophenol, Chin. J. Struct. Chem., 2021, 40, 1372–1378 CAS.
  32. Q. Li, S. C. Wang, Z. X. Sun, Q. J. Tang, Y. Q. Liu, L. Z. Wang, H. Q. Wang and Z. B. Wu, Enhanced CH4 selectivity in CO2 photocatalytic reduction over carbon quantum dots decorated and oxygen doping g-C3N4, Nano Res., 2019, 12, 2749–2759 CrossRef CAS.
  33. S. R. Tao, S. J. Wan, Q. Y. Huang, C. M. Li, J. G. Yu and S. W. Cao, Molecular engineering of g-C3N4 with dibenzothiophene groups as electron donor for enhanced photocatalytic H2-production, Chin. J. Struct. Chem., 2022, 41, 2206048–2206054 Search PubMed.
  34. O. V. Kuznetsov, A. Cole, M. Pulikkathara and V. N. Khabashesku, Sidewall alkylcarboxylation of carbon nanotubes through reactions of fluoronanotubes with functional free radicals, Russ. Chem. Bull., Int. Ed., 2011, 60, 2212–2221 CrossRef CAS.
  35. F. Dong, S. Guo, H. Q. Wang, X. F. Li and Z. B. Wu, Enhancement of the visible light photocatalytic activity of C-doped TiO2 nanomaterials prepared by a green synthetic approach, J. Phys. Chem. C, 2011, 115, 13285–13292 CrossRef CAS.
  36. P. F. Xia, S. W. Cao, B. C. Zhu, M. J. Liu, M. S. Shi, J. G. Yu and Y. F. Zhang, Designing 0D/2D S-scheme heterojunction over polymeric carbon nitride for visible-light photocatalytic inactivation of bacteria, Angew. Chem., Int. Ed., 2020, 59, 5218–5225 CrossRef CAS PubMed.
  37. C. Yang, Y. J. Wang, J. G. Yu and S. W. Cao, Ultrathin 2D/2D graphdiyne/Bi2WO6 heterojunction for gas-phase CO2 photoreduction, ACS Appl. Energy Mater., 2021, 4, 8734–8738 CrossRef CAS.
  38. M. Thommes, K. Kaneko, A. V. Neimark, J. P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol and K. S. W. Sing, Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report), Pure Appl. Chem., 2015, 87, 1051–1069 CrossRef CAS.
  39. S. W. Cao, Y. J. Wang, B. C. Zhu, G. C. Xie, J. G. Yu and J. R. Gong, Enhanced photochemical CO2 reduction in the gas phase by graphdiyne, J. Mater. Chem. A, 2020, 8, 7671–7676 RSC.
  40. S. Wageh, O. A. Al-Hartomy, M. F. Alotaibi and L.-J. Liu, Ionized cocatalyst to promote CO2 photoreduction activity over core–triple-shell ZnO hollow spheres, Rare Met., 2022, 41, 1077–1079 CrossRef CAS.
  41. R. Shi, F. L. Liu, Z. Wang, Y. X. Weng and Y. Chen, Black/red phosphorus quantum dots for photocatalytic water splitting: From a type I heterostructure to a Z-scheme system, Chem. Commun., 2019, 55, 12531–12534 RSC.
  42. Q. Li, Y. Xia, C. Yang, K. L. Lv, M. Lei and M. Li, Building a direct Z-scheme heterojunction photocatalyst by ZnIn2S4 nanosheets and TiO2 hollowspheres for highly-efficient artificial photosynthesis, Chem. Eng. J., 2018, 349, 287–296 CrossRef CAS.
  43. M. Zhang, Y. F. Li, W. Chang, W. Zhu, L. H. Zhang, R. X. Jin and Y. Xing, Negative inductive effect enhances charge transfer driving in sulfonic acid functionalized graphitic carbon nitride with efficient visible-light photocatalytic performance, Chin. J. Catal., 2022, 43, 526–535 CrossRef CAS.
  44. C. Cheng, B. W. He, J. J. Fan, B. Cheng, S. W. Cao and J. G. Yu, An inorganic/organic S-scheme heterojunction H2-production photocatalyst and its charge transfer mechanism, Adv. Mater., 2021, 33, 2100317 CrossRef CAS PubMed.
  45. L. Shi, L. Q. Yang, W. Zhou, Y. Y. Liu, L. S. Yin, X. Hai, H. Song and J. H. Ye, Photoassisted construction of holey defective g-C3N4 photocatalysts for efficient visible-light-driven H2O2 production, Small, 2018, 14, 1703142 CrossRef PubMed.
  46. J. Echeverría, Alkyl groups as electron density donors in π-hole bonding, CrystEngComm, 2017, 19, 6289–6296 RSC.
  47. G. G. Zhang, G. S. Li, Z. A. Lan, L. H. Lin, A. Savateev, T. Heil, S. Zafeiratos, X. C. Wang and M. Antonietti, Optimizing optical absorption, exciton dissociation, and charge transfer of a polymeric carbon nitride with ultrahigh solar hydrogen production activity, Angew. Chem., Int. Ed., 2017, 56, 13445–13449 CrossRef CAS PubMed.
  48. W. L. Wang, W. L. Zhao, H. C. Zhang, X. C. Dou and H. F. Shi, 2D/2D step-scheme α-Fe2O3/Bi2WO6 photocatalyst with efficient charge transfer for enhanced photo-Fenton catalytic activity, Chin. J. Catal., 2021, 42, 97–106 CrossRef CAS.
  49. T. M. Su, Z. D. Hood, M. Naguib, L. Bai, S. Luo, C. M. Rouleau, I. N. Ivanov, H. B. Ji, Z. Z. Qin and Z. L. Wu, 2D/2D heterojunction of Ti3C2/g-C3N4 nanosheets for enhanced photocatalytic hydrogen evolution, Nanoscale, 2019, 11, 8138–8149 RSC.
  50. K. Wang, X. Z. Feng, Y. Z. Shangguan, X. Y. Wu and H. Chen, Selective CO2 photoreduction to CH4 mediated by dimension-matched 2D/2D Bi3NbO7/g-C3N4 S-scheme heterojunction, Chin. J. Catal., 2022, 43, 246–254 CrossRef CAS.
  51. L. Cheng, P. Zhang, Q. Y. Wen, J. J. Fan and Q. J. Xiang, Copper and platinum dual-single-atoms supported on crystalline graphitic carbon nitride for enhanced photocatalytic CO2 reduction, Chin. J. Catal., 2022, 43, 451–460 CrossRef CAS.
  52. H. Q. Wang, J. R. Guan, J. Z. Li, X. Li, C. C. Ma, P. W. Huo and Y. S. Yan, Fabricated g-C3N4/Ag/m-CeO2 composite photocatalyst for enhanced photoconversion of CO2, Appl. Surf. Sci., 2020, 506, 144931 CrossRef CAS.
  53. X. H. Song, X. Li, X. Y. Zhang, Y. F. Wu, C. C. Ma, P. W. Huo and Y. S. Yan, Fabricating C and O co-doped carbon nitride with intramolecular donor-acceptor systems for efficient photoreduction of CO2 to CO, Appl. Catal., B, 2020, 268, 118736 CrossRef CAS.
  54. F. Li, D. N. Zhang and Q. J. Xiang, Nanosheet-assembled hierarchical flower-like g-C3N4 for enhanced photocatalytic CO2 reduction activity, Chem. Commun., 2020, 56, 2443–2446 RSC.
  55. Z. M. Sun, W. Fang, L. Zhao, H. Chen, X. He, W. X. Li, P. Tian and Z. H. Huang, g-C3N4 foam/Cu2O QDs with excellent CO2 adsorption and synergistic catalytic effect for photocatalytic CO2 reduction, Environ. Int., 2019, 130, 104898 CrossRef CAS PubMed.
  56. H. L. Li, Y. Gao, Z. Xiong, C. Liao and K. M. Shih, Enhanced selective photocatalytic reduction of CO2 to CH4 over plasmonic Au modified g-C3N4 photocatalyst under UV–vis light irradiation, Appl. Surf. Sci., 2018, 439, 552–559 CrossRef CAS.
  57. C. Q. Han, R. M. Zhang, Y. H. Ye, L. Wang, Z. Y. Ma, F. Y. Su, H. Q. Xie, Y. Zhou, P. K. Wong and L. Q. Ye, Chainmail co-catalyst of NiO shell-encapsulated Ni for improving photocatalytic CO2 reduction over g-C3N4, J. Mater. Chem. A, 2019, 7, 9726–9735 RSC.
  58. L. Cheng, H. Yin, C. Cai, J. J. Fan and Q. J. Xiang, Single Ni atoms anchored on porous few-layer g-C3N4 for photocatalytic CO2 reduction: The role of edge confinement, Small, 2020, 16, 2002411 CrossRef CAS PubMed.
  59. Y. Y. Wang, H. L. Huang, Z. Z. Zhang, C. Wang, Y. Y. Yang, Q. Li and D. S. Xu, Lead-free perovskite Cs2AgBiBr6@g-C3N4 Z-scheme system for improving CH4 production in photocatalytic CO2 reduction, Appl. Catal., B, 2021, 282, 119570 CrossRef CAS.
  60. J.-W. Gu, R.-T. Guo, Y.-F. Miao, Y.-Z. Liu, G.-L. Wu, C.-P. Duan and W.-G. Pan, Noble-metal-free Bi/g-C3N4 nanohybrids for efficient photocatalytic CO2 reduction under simulated irradiation, Energy Fuels, 2021, 35, 10102–10112 CrossRef CAS.
  61. A. Raza, A. A. Haidry and J. Saddique, In situ synthesis of Cu2ZnSnS4/g-C3N4 heterojunction for superior visible light-driven CO2 reduction, J. Phys. Chem. Solids, 2022, 165, 110694 CrossRef CAS.
  62. K. F. Wang, L. Zhang, Y. Su, D. K. Shao, S. W. Zeng and W. Z. Wang, Photoreduction of carbon dioxide of atmospheric concentration to methane with water over CoAl-layered double hydroxide nanosheets, J. Mater. Chem. A, 2018, 6, 8366–8373 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2nr02551e

This journal is © The Royal Society of Chemistry 2022