催化学报 ›› 2021, Vol. 42 ›› Issue (5): 710-730.DOI: 10.1016/S1872-2067(20)63698-1
王宗鹏a,b, 林志萍a,c, 申士杰a,c, 钟文武a,c,*(), 曹少文b,#(
)
收稿日期:
2020-04-29
接受日期:
2020-04-29
出版日期:
2021-05-18
发布日期:
2021-01-29
通讯作者:
钟文武,曹少文
基金资助:
Zongpeng Wanga,b, Zhiping Lina,c, Shijie Shena,c, Wenwu Zhonga,c,*(), Shaowen Caob,#(
)
Received:
2020-04-29
Accepted:
2020-04-29
Online:
2021-05-18
Published:
2021-01-29
Contact:
Wenwu Zhong,Shaowen Cao
About author:
# E-mail: swcao@whut.edu.cnSupported by:
摘要:
伴随着人类文明的快速发展, 一些危机慢慢显露出来, 例如能源危机、环境污染和全球变暖. 2019年5月11日, Mauna Loa天文台报告, 大气中的二氧化碳水平超过415 ppm, 达到人类历史上的最高记录, 欧盟随后于2019年11月宣布了气候紧急状态. 因此, 绿色能源技术已成为迫切需求, 以减少化石燃料的使用, 并减少污染物的产生.
光催化是直接利用太阳能的技术, 可以应用于水分解产氢、CO2还原、降解有机污染物、促进有机物合成等, 是解决能源和环境问题的最有前途的技术之一. 光催化剂是光催化技术的核心. 目前, 许多半导体材料可作为光催化剂, 并已被充分地研究, 例如TiO2、CdS、ZnO、BiVO4和C3N4等. 然而, 单一的半导体材料具有一些缺点, 阻碍了它们的实际应用. 其中, 限制这些半导体材料光催化活性的一个关键问题是, 光生电子-空穴对容易快速复合而不是参与光催化反应. 例如, ZnO中激子的寿命估计仅为数百皮秒, 大多数激子来不及参与到氧化还原反应中.
为了抑制电子-空穴对的复合, 需要应用特殊的策略. 构建异质结光催化材料已成为最有前途的方法之一. 通常, 可以根据相邻材料的能带结构, 将异质结分为以下几种类型: PN型异质结, II型异质结, Schottky型异质结和S型异质结. 以上异质结大都是由两种半导体材料复合而形成的. 除此之外, 还可以根据形成异质结的特殊材料, 补充两种特殊的异质结类型, 即晶面异质结和石墨烯基异质结. 晶面异质结是由同一材料, 由于暴露不同的晶面而形成的. 石墨烯具有独特的能带结构、极大的比表面积及优良的导电性, 可以与其他半导体形成各种类型的异质结. 这些异质结材料能有效抑制电子-空穴对的复合, 从而提高材料整体的光催化活性, 也已成为光催化剂家族的重要分支. 本文详细介绍了以上各种类型的异质结光催化剂的最新进展, 概述了实现高性能异质结光催化剂的基本策略, 并对异质结光催化剂未来发展方向进行了一些探讨.
王宗鹏, 林志萍, 申士杰, 钟文武, 曹少文. 异质结光催化材料的新进展[J]. 催化学报, 2021, 42(5): 710-730.
Zongpeng Wang, Zhiping Lin, Shijie Shen, Wenwu Zhong, Shaowen Cao. Advances in designing heterojunction photocatalytic materials[J]. Chinese Journal of Catalysis, 2021, 42(5): 710-730.
Fig. 5. (a) Band structure of Schottky heterojunction; (b) Illustration of the oscillation of electrons in a metal particle; (c) Harmonic model to describe the plasmon resonance.
Heterojunctions | Similarity | Difference |
---|---|---|
Type II and S-scheme heterojunctions | Similar band alignment | Recombination pathway |
Type II and PN heterojunctions | Similar charge separation pathway | The built-in field is against the charge separation for Type II heterojunction |
PN and Schottky heterojunctions | Strong internal field | Schottky barrier is formed for Schottky heterojunction |
Table 1 Similarity and difference of various heterojunctions.
Heterojunctions | Similarity | Difference |
---|---|---|
Type II and S-scheme heterojunctions | Similar band alignment | Recombination pathway |
Type II and PN heterojunctions | Similar charge separation pathway | The built-in field is against the charge separation for Type II heterojunction |
PN and Schottky heterojunctions | Strong internal field | Schottky barrier is formed for Schottky heterojunction |
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
WO3/TiO2 | Stearic acid degradation | 14 times higher than pure TiO2 | 2017 | [ |
NiO/TiO2 | H2 generation | 22 times greater than TiO2 | 2018 | [ |
CsPbBrCl2/g-C3N4 | Organic effluents degradation | Degrading 94% of Eosin B dye within 120 min | 2019 | [ |
CdS/ZnSe | Water splitting | Higher carrier mobility | 2018 | [ |
SnOx/Zn2SnO4 | Degradation of methyl orange | Complete decolorization in 30 min | 2017 | [ |
CuSbSe2/TiO2 | Degradation of organic dyes | 21.08% increase in the degradation efficiency than TiO2 | 2019 | [ |
WO3/BiOBr | Oxidation of amine to imine | 2 times higher than WO3 | 2020 | [ |
ZnO/Ag2S | Anti-bacteria | Significant antibacterial activity | 2020 | [ |
g-C3N4/ZnTe | Photosynthesis of ethanol | Generation rate of 17.1 μmol cm-2 h-1 (at -1.1 V vs. Ag/AgCl) | 2019 | [ |
I-BiOCl/I-BiOBr | Degradation of methyl orange | 7 times than BiOBr | 2017 | [ |
La2Ti2O7/In2S3 | H2 generation | 18 times higher than physical mixtures of La2Ti2O7 and In2S3 | 2019 | [ |
BiVO4/WO3 | Photoanode | 2.15 times enhancement than WO3 | 2016 | [ |
Pt/TiO2/CdS/Co3O4 | H2 generation | A generation rate of 2000 μmol g-1 h-1 | 2018 | [ |
VO2/CuWO4 | Degradation of azure II | Degrading 83.8% in 80 min | 2018 | [ |
ZnS/g-C3N4 | MB degradation | 2.6 times better compared to bare g-C3N4 | 2017 | [ |
Cu2O/g-C3N4 | H2 generation | 4 times higher than pure g-C3N4 | 2017 | [ |
CuS/BiFeO3 | Degradation of alachlor | 95% degradation within 1 h | 2018 | [ |
Cu2O/SnO2 | RhB degradation | 2-fold higher than individual SnO2 | 2016 | [ |
S/CdS | H2 generation | 8.14 mmol h-1 | 2017 | [ |
Mn3O4/MnO2 | MB degradation | 93.5% degradation within 1 h | 2017 | [ |
SnS2/H‐TiO2/Ti | H2 generation | 70 times higher than SnS2/TiO2/Ti | 2019 | [ |
TiO2/CdIn2S4 | H2 generation | 5.5 times higher than bare TiO2 | 2018 | [ |
IO/CdS | Degradation of xylenol blue | 97.6% degradation within 3 h | 2019 | [ |
g-C3N4/P25 | H2 generation | 24.5 times higher than pure g-C3N4 | 2019 | [ |
ZnIn2S4/BiPO4 | TC degradation | 84% degradation within 80 min | 2019 | [ |
Bi2SiO5/BiPO4 | Phenol degradation | 4.36 timed higher than Bi2SiO5 | 2018 | [ |
g-C3N4/Nb2O5 | RhB degradation | 9.4 times higher than Nb2O5 | 2017 | [ |
Table 2 Comparison of some recently developed type II heterojunction photocatalysts.
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
WO3/TiO2 | Stearic acid degradation | 14 times higher than pure TiO2 | 2017 | [ |
NiO/TiO2 | H2 generation | 22 times greater than TiO2 | 2018 | [ |
CsPbBrCl2/g-C3N4 | Organic effluents degradation | Degrading 94% of Eosin B dye within 120 min | 2019 | [ |
CdS/ZnSe | Water splitting | Higher carrier mobility | 2018 | [ |
SnOx/Zn2SnO4 | Degradation of methyl orange | Complete decolorization in 30 min | 2017 | [ |
CuSbSe2/TiO2 | Degradation of organic dyes | 21.08% increase in the degradation efficiency than TiO2 | 2019 | [ |
WO3/BiOBr | Oxidation of amine to imine | 2 times higher than WO3 | 2020 | [ |
ZnO/Ag2S | Anti-bacteria | Significant antibacterial activity | 2020 | [ |
g-C3N4/ZnTe | Photosynthesis of ethanol | Generation rate of 17.1 μmol cm-2 h-1 (at -1.1 V vs. Ag/AgCl) | 2019 | [ |
I-BiOCl/I-BiOBr | Degradation of methyl orange | 7 times than BiOBr | 2017 | [ |
La2Ti2O7/In2S3 | H2 generation | 18 times higher than physical mixtures of La2Ti2O7 and In2S3 | 2019 | [ |
BiVO4/WO3 | Photoanode | 2.15 times enhancement than WO3 | 2016 | [ |
Pt/TiO2/CdS/Co3O4 | H2 generation | A generation rate of 2000 μmol g-1 h-1 | 2018 | [ |
VO2/CuWO4 | Degradation of azure II | Degrading 83.8% in 80 min | 2018 | [ |
ZnS/g-C3N4 | MB degradation | 2.6 times better compared to bare g-C3N4 | 2017 | [ |
Cu2O/g-C3N4 | H2 generation | 4 times higher than pure g-C3N4 | 2017 | [ |
CuS/BiFeO3 | Degradation of alachlor | 95% degradation within 1 h | 2018 | [ |
Cu2O/SnO2 | RhB degradation | 2-fold higher than individual SnO2 | 2016 | [ |
S/CdS | H2 generation | 8.14 mmol h-1 | 2017 | [ |
Mn3O4/MnO2 | MB degradation | 93.5% degradation within 1 h | 2017 | [ |
SnS2/H‐TiO2/Ti | H2 generation | 70 times higher than SnS2/TiO2/Ti | 2019 | [ |
TiO2/CdIn2S4 | H2 generation | 5.5 times higher than bare TiO2 | 2018 | [ |
IO/CdS | Degradation of xylenol blue | 97.6% degradation within 3 h | 2019 | [ |
g-C3N4/P25 | H2 generation | 24.5 times higher than pure g-C3N4 | 2019 | [ |
ZnIn2S4/BiPO4 | TC degradation | 84% degradation within 80 min | 2019 | [ |
Bi2SiO5/BiPO4 | Phenol degradation | 4.36 timed higher than Bi2SiO5 | 2018 | [ |
g-C3N4/Nb2O5 | RhB degradation | 9.4 times higher than Nb2O5 | 2017 | [ |
Fig. 6. (a) HRTEM image of g-C3N4/ZnTe. (b) Products of g-C3N4/ZnTe-x-y (x and y represented the mass ratio of g-C3N4 and ZnTe) at -1.1 V vs. Ag/AgCl. (c) Products of g-C3N4/ZnTe-1-2 at various applied potentials. (d,e) The optimized geometry structures and binding energies for CO2 (d) and CO (e) adsorption on g-C3N4, ZnTe, and g-C3N4/ZnTe. (f) Schematic diagram of band configuration and the charge separation at the interface of g-C3N4/ZnTe heterojunction. (reproduced with permission [68], copyright 2019, Elsevier).
Fig. 7. (a) Band structure of type-II I-BiOCl/I-BiOBr composite. Photocatalytic activities (b,d) and corresponding kapp constant (c,e) of various samples for MO and phenol degradation. (reproduced with permission [69], copyright 2017, Elsevier).
Fig. 8. (a) Band alignment of the QD@LDH@BiVO4 (QD: CdTe quantum dots, LDH: Co-based layered double hydroxide) photoanode. (b-d) SEM images of BiVO4, LDH@BiVO4 and QD@LDH@BiVO4, respectively. (e) HRTEM images of QD@LDH@BiVO4. (f) Oxidation efficiency of the surface-reaching holes injected into the solution species. (reproduced with permission [87], copyright 2016, American Chemistry Society).
Fig. 10. (a) TEM image of the La2Ti2O7/In2S3 heterojunction. (b) HRTEM image of the La2Ti2O7/In2S3 heterojunction. (c) Schematic illustration of the band edge positions for both La2Ti2O7 and In2S3. (reproduced with permission [70], copyright 2019, Elsevier).
Fig. 9. (a,b) Top and side view of the C2N/WS2 heterojunction. Grey, green, purple and orange colors denote C, N, W and S atoms, respectively. (c) Band structure of the C2N/WS2 heterojunction. (d) Absolute energy positions of VBM (valence band maxima) and CBM (conduction band minima) for all three structures with respect to vacuum level. (reproduced with permission [88], copyright 2018, Elsevier).
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
Fe2O3/BiOI | RhB degradation | 4.24 times higher than single BiOI | 2016 | [ |
Cu3P/g-C3N4 | Hydrogen generation | 95 times higher activity than bare g-C3N4 | 2018 | [ |
MoSe2/Bi2WO6 | Degradation of gas-phase toluene | 7 times higher compared with pure MoSe2 | 2018 | [ |
ZnO/MnWO4 | Degradations of RhB, MB, MO | 22.5, 17.7, 26.8 times higher than that of the ZnO sample in degradations of RhB, MB, MO | 2018 | [ |
BiOI/BiPO4 | Degradation of tetracycline | 1.98 times and 2.46 times higher than BiOI and BiPO4 | 2019 | [ |
Ag2S/BiVO4 | Degradation of tetracycline | 90.2% degradation efficiency | 2019 | [ |
Ag2O/AgNbO3 | Degradation of RhB | Degradation efficiency of 95.4% within 90 min | 2019 | [ |
MoS2/MgIn2S4 | NH3 production | 4 and 7 times higher than bare MoS2 and MgIn2S4 | 2020 | [ |
Bi2O3/MoS2 | Hydrogen generation | 10 times higher than pure Bi2O3 and MoS2 | 2020 | [ |
ZnFe2O4/SnS2 | Degradation of MO | 1.7 times higher than individual ZnFe2O4 | 2020 | [ |
BiVO4@MoS2 | Reduction of Cr6+ | 76.5 % Cr6+ reduced within 90 min | 2016 | [ |
g-C3N4/Bi4Ti3O12 | Degradation of acid orange-II | 2.85 times higher than pure Bi4Ti3O12 | 2016 | [ |
WO6/BiOI | Antifouling | 99.99% P. aeruginosa, E. coli and S. aureus killed within 60 min | 2016 | [ |
BiPO4/BiOCl | MO degradation | A degradation efficiency of 98% within 14 min | 2015 | [ |
Ag2O/Bi5O7I | Degradation of Bisphenol A and phenol | A degradation efficiency of 99.9% within 40 min | 2017 | [ |
BiOI/CeO2 | MO degradation | A degradation efficiency of 93.75% within 50 min | 2017 | [ |
Ag2O/TiO2 | PNP degradation | 7.7 times higher than TiO2 | 2015 | [ |
CdWO4/BiOBr | Degradation of RhB | 2.8 times better than BiOBr | 2016 | [ |
BiPO4/BiOBr | o-Dichlorobenzene removal | A removal efficiency of 53.6% | 2017 | [ |
CuO/ZnO | Degradation of RhB and CR | Complete degradation within 60 min | 2017 | [ |
Ag2O/BiOCOOH | Degradation of RhB | 96.8% degradation within 60 min | 2017 | [ |
Bi2Sn2O7/Ag2CrO4 | Degradation of RhB, MO and MB | 97.5%, 90.4% and 99.8% degradation within 120 min, respectively | 2018 | [ |
Co3O4/Bi2O2CO3 | Degradation of MO | 90% degradation within 150 min | 2017 | [ |
BiOI/TiO2 | Selective hydroxylation of phenol | A selectivity of 92.1% | 2018 | [ |
Table 3 Comparison of some recently developed PN heterojunction photocatalysts.
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
Fe2O3/BiOI | RhB degradation | 4.24 times higher than single BiOI | 2016 | [ |
Cu3P/g-C3N4 | Hydrogen generation | 95 times higher activity than bare g-C3N4 | 2018 | [ |
MoSe2/Bi2WO6 | Degradation of gas-phase toluene | 7 times higher compared with pure MoSe2 | 2018 | [ |
ZnO/MnWO4 | Degradations of RhB, MB, MO | 22.5, 17.7, 26.8 times higher than that of the ZnO sample in degradations of RhB, MB, MO | 2018 | [ |
BiOI/BiPO4 | Degradation of tetracycline | 1.98 times and 2.46 times higher than BiOI and BiPO4 | 2019 | [ |
Ag2S/BiVO4 | Degradation of tetracycline | 90.2% degradation efficiency | 2019 | [ |
Ag2O/AgNbO3 | Degradation of RhB | Degradation efficiency of 95.4% within 90 min | 2019 | [ |
MoS2/MgIn2S4 | NH3 production | 4 and 7 times higher than bare MoS2 and MgIn2S4 | 2020 | [ |
Bi2O3/MoS2 | Hydrogen generation | 10 times higher than pure Bi2O3 and MoS2 | 2020 | [ |
ZnFe2O4/SnS2 | Degradation of MO | 1.7 times higher than individual ZnFe2O4 | 2020 | [ |
BiVO4@MoS2 | Reduction of Cr6+ | 76.5 % Cr6+ reduced within 90 min | 2016 | [ |
g-C3N4/Bi4Ti3O12 | Degradation of acid orange-II | 2.85 times higher than pure Bi4Ti3O12 | 2016 | [ |
WO6/BiOI | Antifouling | 99.99% P. aeruginosa, E. coli and S. aureus killed within 60 min | 2016 | [ |
BiPO4/BiOCl | MO degradation | A degradation efficiency of 98% within 14 min | 2015 | [ |
Ag2O/Bi5O7I | Degradation of Bisphenol A and phenol | A degradation efficiency of 99.9% within 40 min | 2017 | [ |
BiOI/CeO2 | MO degradation | A degradation efficiency of 93.75% within 50 min | 2017 | [ |
Ag2O/TiO2 | PNP degradation | 7.7 times higher than TiO2 | 2015 | [ |
CdWO4/BiOBr | Degradation of RhB | 2.8 times better than BiOBr | 2016 | [ |
BiPO4/BiOBr | o-Dichlorobenzene removal | A removal efficiency of 53.6% | 2017 | [ |
CuO/ZnO | Degradation of RhB and CR | Complete degradation within 60 min | 2017 | [ |
Ag2O/BiOCOOH | Degradation of RhB | 96.8% degradation within 60 min | 2017 | [ |
Bi2Sn2O7/Ag2CrO4 | Degradation of RhB, MO and MB | 97.5%, 90.4% and 99.8% degradation within 120 min, respectively | 2018 | [ |
Co3O4/Bi2O2CO3 | Degradation of MO | 90% degradation within 150 min | 2017 | [ |
BiOI/TiO2 | Selective hydroxylation of phenol | A selectivity of 92.1% | 2018 | [ |
Fig. 11. (a) The charge transfer pathway between Cu2-xSe and CdS under illumination. (b,c) SEM and TEM images of 10?wt% Cu2-xSe/CdS composite. (d) Photocatalytic performance of CdS and different ratios of Cu2-xSe/CdS composites. (reproduced with permission [118], copyright 2019, Elsevier).
Fig. 12. (a) The schematic transformation from α-Bi2O3 to the β-Bi2O3. (b,c) TEM and HRTEM images of Bi2O3/Bi2SiO5 heterojunction. (d) Photocurrent responses of α-Bi2O3 and Bi2O3/Bi2SiO5 heterojunctions. The products with the SiO2 mass ratio of 1%, 5%, 10%, 15% and 20% were denoted as BiSi-1, BiSi-2, BiSi-3, BiSi-4 and BiSi-5, respectively. (reproduced with permission [119], copyright 2018, Elsevier).
Fig. 13. (a,b) TEM and HRTEM images of the BiOBr/La2Ti2O7 heterojunction. (c) Photocatalytic activity of various samples for RhB degradation. (d) Schematic diagram of the possible photocatalytic mechanism over BiOBr/La2Ti2O7 under both UV and visible light irradiation. (reproduced with permission [120], copyright 2016, Elsevier).
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
Ru/TiO2 | Water splitting | H2 evolution rate of 40.2 µmol h-1 g-1 | 2016 | [ |
Au/Cu2O | MO degradation | 1.2 times higher than pure Cu2O microcubes | 2016 | [ |
TiO2/NC-Ru | Oxidation of benzyl alcohol | 99% selectivity | 2019 | [ |
TaON/Ni | Reduction of CO2 to CH4 | 5.1 times higher than pristine TaON | 2019 | [ |
Bi2MoO6/Ti3C2 | Molecular oxygen activation | 5.6 times higher than pristine samples | 2020 | [ |
CoSe2/CdS0.95Se0.05 | H2 evolution | 66 times higher than CdS | 2020 | [ |
Co1.4Ni0.6P | H2 evolution | H2 evolution rate of 405 µmol h-1 g-1 | 2020 | [ |
PdAg/g-C3N4 | Dehydrogenation of formic acid | TOF of 420 h-1 | 2019 | [ |
CdS/Ti3C2 | H2 evolution | 7 times higher than CdS | 2020 | [ |
CdS/Cd | H2 evolution | 40 times compared to pure CdS | 2019 | [ |
Pd/SiC | Coupling of iodobenezene and phenylboronic acid | A high turnover frequency of 1053 h-1 and a selectivity of nearly 100% | 2015 | [ |
Ag/Ag2O | MO degradation | 84%, 78% and 88% of MO are degraded under UV, visible and NIR irradiation | 2016 | [ |
CoP/g-C3N4 | H2 evolution | The 2D/2D structure performs 2.1 times better than 0D/2D structure | 2018 | [ |
Ru/TiO2 | Aerobic oxidation of benzyl alcohol | 84% of MO degraded for 25 min | 2016 | [ |
Ag/CeO2 | MB degradation | 95.33% of MB degraded within 30 min | 2015 | [ |
Pd/BiFeO3 | MO degradation | 10.7 times higher than BiFeO3 | 2016 | [ |
CQDs/H-γ-TaON | H2 evolution | H2 evolution rate of 496.5 µmol h-1 | 2015 | [ |
1T‐MoS2/CdS | H2 evolution | 39 times higher than pure CdS | 2016 | [ |
CdS/Cd | H2 evolution | H2 evolution rate of 1.68 mmol h-1 | 2015 | [ |
1T-MoS2/O-g-C3N4 2D | H2 evolution | H2 evolution rate of 1841.72 μmol h-1 g-1 | 2018 | [ |
Ni/NiTe2 | H2 evolution | A rate of 2214.2 μmol g-1 h-1 | 2020 | [ |
Ag/CN | H2 evolution | H2 evolution rate of 364.6 μmol g-1 | 2019 | [ |
Ag2WO4/Ti3C2 | Degradation of TC and SFE | Degradation rate of 62.9% and 88.6%, respectively | 2019 | [ |
Bi/Bi2O2CO3 | MB degradation | Complete degradation within 60 min | 2016 | [ |
Pt/PbS | H2 evolution | 14.38 times higher than PbS | 2017 | [ |
Table 4 Comparison of some recently developed Schottky heterojunction photocatalysts.
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
Ru/TiO2 | Water splitting | H2 evolution rate of 40.2 µmol h-1 g-1 | 2016 | [ |
Au/Cu2O | MO degradation | 1.2 times higher than pure Cu2O microcubes | 2016 | [ |
TiO2/NC-Ru | Oxidation of benzyl alcohol | 99% selectivity | 2019 | [ |
TaON/Ni | Reduction of CO2 to CH4 | 5.1 times higher than pristine TaON | 2019 | [ |
Bi2MoO6/Ti3C2 | Molecular oxygen activation | 5.6 times higher than pristine samples | 2020 | [ |
CoSe2/CdS0.95Se0.05 | H2 evolution | 66 times higher than CdS | 2020 | [ |
Co1.4Ni0.6P | H2 evolution | H2 evolution rate of 405 µmol h-1 g-1 | 2020 | [ |
PdAg/g-C3N4 | Dehydrogenation of formic acid | TOF of 420 h-1 | 2019 | [ |
CdS/Ti3C2 | H2 evolution | 7 times higher than CdS | 2020 | [ |
CdS/Cd | H2 evolution | 40 times compared to pure CdS | 2019 | [ |
Pd/SiC | Coupling of iodobenezene and phenylboronic acid | A high turnover frequency of 1053 h-1 and a selectivity of nearly 100% | 2015 | [ |
Ag/Ag2O | MO degradation | 84%, 78% and 88% of MO are degraded under UV, visible and NIR irradiation | 2016 | [ |
CoP/g-C3N4 | H2 evolution | The 2D/2D structure performs 2.1 times better than 0D/2D structure | 2018 | [ |
Ru/TiO2 | Aerobic oxidation of benzyl alcohol | 84% of MO degraded for 25 min | 2016 | [ |
Ag/CeO2 | MB degradation | 95.33% of MB degraded within 30 min | 2015 | [ |
Pd/BiFeO3 | MO degradation | 10.7 times higher than BiFeO3 | 2016 | [ |
CQDs/H-γ-TaON | H2 evolution | H2 evolution rate of 496.5 µmol h-1 | 2015 | [ |
1T‐MoS2/CdS | H2 evolution | 39 times higher than pure CdS | 2016 | [ |
CdS/Cd | H2 evolution | H2 evolution rate of 1.68 mmol h-1 | 2015 | [ |
1T-MoS2/O-g-C3N4 2D | H2 evolution | H2 evolution rate of 1841.72 μmol h-1 g-1 | 2018 | [ |
Ni/NiTe2 | H2 evolution | A rate of 2214.2 μmol g-1 h-1 | 2020 | [ |
Ag/CN | H2 evolution | H2 evolution rate of 364.6 μmol g-1 | 2019 | [ |
Ag2WO4/Ti3C2 | Degradation of TC and SFE | Degradation rate of 62.9% and 88.6%, respectively | 2019 | [ |
Bi/Bi2O2CO3 | MB degradation | Complete degradation within 60 min | 2016 | [ |
Pt/PbS | H2 evolution | 14.38 times higher than PbS | 2017 | [ |
Fig. 14. (a) A schematic of the generation of CdS/Cd Schottky junction. (b) Mechanism of the improved photocatalytic activity for CdS/Cd. (c) STEM image of CdS/Cd viewed from the [100] direction. (d,e) Enlarged view of the corresponding areas in (c). (f) Comparison of the photocatalytic hydrogen evolution over CdS and CdS/Cd. (g) Cycling hydrogen evolution measurements for CdS/Cd. (reproduced with permission [135], copyright 2019, Elsevier).
Fig. 15. (A) Illustration of the structure of Ru/TiO2/RuO2 heterojunctions. (B) Band structure of the Ru/TiO2/RuO2 heterojunctions. (C,D,E,F) TEM and HRTEM images of the Ru/TiO2/RuO2 heterojunctions. (G,H) Photocatalytic H2 and O2 evolution over Ru8.0/TiO2 NBs-T, where T indicates the annealing temperatures. (reproduced with permission [125], copyright 2016, Wiley).
Fig. 16. (a) Schematic illustration showing the synthesis of Pt@MIL-125/Au and the corresponding Pt/MIL-125/Au and MIL-125/Au analogues. (b) Photocatalytic H2 production rates of different catalysts. (c) Schematic illustration showing the electron migration at the two metal/MOF interfaces. (reproduced with permission [153], copyright 2018, Wiley).
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
V2O5/g-C3N4 | RhB degradation | 7.3 and 13.0 times higher than individual g-C3N4 and V2O5 | 2016 | [ |
Bi2Fe4O9/Bi2WO6 | RhB degradation | Complete degradation within 90 min | 2018 | [ |
Ag2CrO4/GO | MB degradation | 3.5 times higher than Ag2CrO4 | 2015 | [ |
TiO2/CdS | CO2 reduction | 3.5 and 5.4 times higher than CdS and TiO2 | 2019 | [ |
BP/Bi2WO6 | H2 generation | 9.15 times that of pristine | 2019 | [ |
ZnO/CeO2 | RhB degradation | 2.5 and 1.7 times than pristine ZnO and CeO2 | 2018 | [ |
CdS/α-Fe2O3 | H2 generation | H2 production rate of 45 mmol h-1 g-1 and high quantum efficiency of 46.9% | 2020 | [ |
CdS/g‐C3N4 | H2 generation | H2 production rate of 15.3 mmol g-1 h-1 | 2020 | [ |
WO3(H2O)0.333/Ag3PO4 | MB degradation | Complete degradation in 4 min | 2019 | [ |
g-C3N4/C-TiO2 | RhB degradation | 150 times higher than TiO2 | 2019 | [ |
WO3/g-C3N4 | H2 generation | H2 production rate of 982 μmol h-1 g-1 | 2019 | [ |
g-C3N4/Ag3VO4 | MB degradation | Degradation rate of 99.3% within 8 min | 2019 | [ |
SnFe2O4/ZnFe2O4 | Tetracycline degradation | 93.2% removal efficiency | 2020 | [ |
Ni2P-SNO/CdS-D | Hydrogen production | Hydrogen production rate of 7808 μmol g-1 h-1 | 2020 | [ |
BiOBr/Bi2MoO6 | Degradation of RhB | 66.7% total organic carbon could be removed | 2017 | [ |
Bi2S3/Pg-C3N4 | MB degradation | 29.74 times greater than Bi2S3 | 2019 | [ |
g-C3N4/SnS2 | CO2 reduction | CH3OH yield of 2.3 μmol g-1 and CH4 yield of 0.64 μmol g-1 | 2017 | [ |
WO3/g-C3N4 | TC degradation | Degradation rate of 90.54% within 120 min | 2020 | [ |
Pg-C3N4/Zn0.2Cd0.8S | H2 generation | hydrogen production rate of 6.69 mmol g-1 h-1 | 2020 | [ |
Ag3VO4/BiVO4 | Photocatalytic reduction of Cr6+ and oxidation of Bisphenol S | Photocatalytic reduction and oxidation efficiency of 74.9% and 94.8% | 2019 | [ |
LaCoO3/g-C3N4 | Phenol degradation | 7.5 times higher than g-C3N4 | 2019 | [ |
Cd0.5Zn0.5S/g-C3N4 | Dye degradation | 13 times higher than g-C3N4 | 2020 | [ |
NiO/BiOI | RhB degradation | 8 times higher than NiO | 2020 | [ |
Bi2S3/BiFeO3 | TC degradation | 6 times higher than Bi2S3 | 2020 | [ |
Table 5 Comparison of some recently developed S-Scheme heterojunction photocatalysts.
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
V2O5/g-C3N4 | RhB degradation | 7.3 and 13.0 times higher than individual g-C3N4 and V2O5 | 2016 | [ |
Bi2Fe4O9/Bi2WO6 | RhB degradation | Complete degradation within 90 min | 2018 | [ |
Ag2CrO4/GO | MB degradation | 3.5 times higher than Ag2CrO4 | 2015 | [ |
TiO2/CdS | CO2 reduction | 3.5 and 5.4 times higher than CdS and TiO2 | 2019 | [ |
BP/Bi2WO6 | H2 generation | 9.15 times that of pristine | 2019 | [ |
ZnO/CeO2 | RhB degradation | 2.5 and 1.7 times than pristine ZnO and CeO2 | 2018 | [ |
CdS/α-Fe2O3 | H2 generation | H2 production rate of 45 mmol h-1 g-1 and high quantum efficiency of 46.9% | 2020 | [ |
CdS/g‐C3N4 | H2 generation | H2 production rate of 15.3 mmol g-1 h-1 | 2020 | [ |
WO3(H2O)0.333/Ag3PO4 | MB degradation | Complete degradation in 4 min | 2019 | [ |
g-C3N4/C-TiO2 | RhB degradation | 150 times higher than TiO2 | 2019 | [ |
WO3/g-C3N4 | H2 generation | H2 production rate of 982 μmol h-1 g-1 | 2019 | [ |
g-C3N4/Ag3VO4 | MB degradation | Degradation rate of 99.3% within 8 min | 2019 | [ |
SnFe2O4/ZnFe2O4 | Tetracycline degradation | 93.2% removal efficiency | 2020 | [ |
Ni2P-SNO/CdS-D | Hydrogen production | Hydrogen production rate of 7808 μmol g-1 h-1 | 2020 | [ |
BiOBr/Bi2MoO6 | Degradation of RhB | 66.7% total organic carbon could be removed | 2017 | [ |
Bi2S3/Pg-C3N4 | MB degradation | 29.74 times greater than Bi2S3 | 2019 | [ |
g-C3N4/SnS2 | CO2 reduction | CH3OH yield of 2.3 μmol g-1 and CH4 yield of 0.64 μmol g-1 | 2017 | [ |
WO3/g-C3N4 | TC degradation | Degradation rate of 90.54% within 120 min | 2020 | [ |
Pg-C3N4/Zn0.2Cd0.8S | H2 generation | hydrogen production rate of 6.69 mmol g-1 h-1 | 2020 | [ |
Ag3VO4/BiVO4 | Photocatalytic reduction of Cr6+ and oxidation of Bisphenol S | Photocatalytic reduction and oxidation efficiency of 74.9% and 94.8% | 2019 | [ |
LaCoO3/g-C3N4 | Phenol degradation | 7.5 times higher than g-C3N4 | 2019 | [ |
Cd0.5Zn0.5S/g-C3N4 | Dye degradation | 13 times higher than g-C3N4 | 2020 | [ |
NiO/BiOI | RhB degradation | 8 times higher than NiO | 2020 | [ |
Bi2S3/BiFeO3 | TC degradation | 6 times higher than Bi2S3 | 2020 | [ |
Fig. 17. (a) TEM image of PCN nanosheets. (b) TEM image of the CeO2/PCN heterojunction photocatalyst. Inset: corresponding HRTEM image of (111) plane of CeO2. (c) HAADF-TEM image of the CeO2/PCN heterojunction photocatalyst. (d) Band structure of pure CeO2 and PCN. (e) The statistic counts of bacteria colonies treated with different samples, respectively. (f) The bacteria removal efficiency over the prepared samples. (reproduced with permission [177], copyright 2020, Wiley).
Fig. 18. (a,b) TEM and HRTEM images of the TiO2/CdS nanofiber. (c) Comparison of photocatalytic H2-production activity of T, TC5, TC10, TC20, and C samples. TCx: T and C represent TiO2 and CdS, respectively, x indicates the weight percentage of CdS with respect to TiO2. (d) Cycling H2-production stability for the TC10 sample. (e) S-scheme charge separation and pathway in TiO2/CdS nanofiber. (reproduced with permission [178], copyright 2019, Wiley).
Fig. 19. (a,b,c) SEM, TEM and HRTEM images of BP/BiOBr photocatalyst. (d) Photocatalytic degradation efficiency of tetracycline (TC) over various samples. (e) Photocatalytic oxygen evolution rate of various samples. (f) Schematic illustration for the charge transfer and reaction mechanism. BP: Black phosphorus. xBP/BiOBr: x indicates the weight ratio of BP respect to BiOBr. (reproduced with permission [179], copyright 2020, Elsevier).
Fig. 20. (a) Schematic illustration of the preparation of COF/SC photocatalysts. (b,c) Photocatalytic performance of the prepared COF/SC samples. (d) Schematic charge separation in COF/SC with COF-138/TiO2 as an example. (reproduced with permission [180], copyright 2020, Wiley).
Fig. 21. (a) Schematic illustration of the (101)/(001) surface heterojunction of TiO2. (b) Schematic illustration of the adjustment of the facet ratio through morphology control. (c) SEM images of samples with different facet ratios. The (101)/(001) facet ratios of T1, T2, T3, and T4 were 49, 11.5, 1.78, and 0.087, respectively. (d,e) Bacterial viability of E. coli and S. aureus exposed to different concentrations of TiO2 nanocrystals, respectively. (reproduced with permission [188], copyright 2017, American Chemical Society).
Fig. 22. (a,b) Schematic illustration of the formation of TiO2 surface heterojunction. (c) Enlarged view showing the exposed facets of TiO2. (d-f) SEM images of TiO2 nanorods and TiO2 nanorods loaded with nanosheets. (g) Photocatalytic performance of various samples. (h) Band alignment of the (111) and (101) facets of rutile TiO2. FH-TiO2: Facet Heterojunction TiO2, TiO2 nanorods loaded with nanosheets; P25: commercial TiO2; NRs-anatase: nanorods with anatase phase; NRs-TiO2: nanorods with rutile phase. (reproduced with permission [190], copyright 2019, Wiley).
Fig. 23. (a) SEM image of BiOI. (b,c) HRTEM images of BiOI viewed from the perpendicular and parallel directions of the BiOI plate, respectively. (d) Schematic illustration of the growth of BiOI nanoplate at different PH values. (e) Band alignment of (110) and (001) facet of BiOI. (f) Degradation curves of phenol on BiOI samples. BiOI-X: X means X mL distilled water was added in the hydrolysis process. (reproduced with permission [205], copyright 2016, Elsevier).
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
TiO2(001)/(101) (101)/(001) | H2 production | H2 evolution rate of 264 μmol h-1 | 2016 | [ |
TiO2(001)/(101) | Reduction CO2 to CH4 | CH4 production rate of 1.58 μmol h-1 g-1 | 2016 | [ |
TiO2(001)/(101) | Removing Hg0 | 87% Hg0 removal efficiency | 2017 | [ |
TiO2(001)/(101) | Anti-bacteria | 50% decrease in the bacterial viability in 6 h | 2017 | [ |
Rutile TiO2(111)/(101) | H2 production | 0.566 mmol g-1 h-1 H2 production rate in pure water | 2019 | [ |
18-facet SrTiO3(001)/(110) | Water splitting | Fivefold enhancement of AQE than 6-facet SrTiO3 | 2016 | [ |
ZnGa2O4(110)/(111) | H2 production | 12.5 mmol g-1 h-1 H2 production rate | 2017 | [ |
CaCu3Ti4O12(001)/(111) | TC degradation | 70 times higher than defect free CaCu3Ti4O12 | 2019 | [ |
m-BiVO4(010)/(110) | O2 evolution | Higher than plate-like BiVO4 | 2016 | [ |
BiOI(001)/(110) | MO degradation | 71.4% MO degradation efficiency | 2018 | [ |
anatase TiO2(001)/(101)/SnS2 | MB degradation | 1.5 times better than pure TiO2 | 2017 | [ |
anatase TiO2(001)/(101)/ZnS | MB degradation | 2 times better than pure TiO2 | 2017 | [ |
BiVO4(010)/(110) | RhB degradation | 53.8% degradation efficiency within 180 min | 2018 | [ |
In2O3 loaded BiVO4(010)/(110) | RhB degradation | 91% degradation efficiency within 180 min | 2017 | [ |
anatase TiO2(001)/(101) | H2 production | a generation rate of 670 μmol h-1 g-1 | 2016 | [ |
CdS loaded TiO2(001)/(101) | 4-CP degradation | 96.12% degradation efficiency | 2018 | [ |
anatase TiO2(001)/(101) | CO2 Reduction | 21 times increasement | 2014 | [ |
Table 6 Comparison of some recently developed surface heterojunction photocatalysts.
Heterojunction | Application | Performance | Year | Ref. |
---|---|---|---|---|
TiO2(001)/(101) (101)/(001) | H2 production | H2 evolution rate of 264 μmol h-1 | 2016 | [ |
TiO2(001)/(101) | Reduction CO2 to CH4 | CH4 production rate of 1.58 μmol h-1 g-1 | 2016 | [ |
TiO2(001)/(101) | Removing Hg0 | 87% Hg0 removal efficiency | 2017 | [ |
TiO2(001)/(101) | Anti-bacteria | 50% decrease in the bacterial viability in 6 h | 2017 | [ |
Rutile TiO2(111)/(101) | H2 production | 0.566 mmol g-1 h-1 H2 production rate in pure water | 2019 | [ |
18-facet SrTiO3(001)/(110) | Water splitting | Fivefold enhancement of AQE than 6-facet SrTiO3 | 2016 | [ |
ZnGa2O4(110)/(111) | H2 production | 12.5 mmol g-1 h-1 H2 production rate | 2017 | [ |
CaCu3Ti4O12(001)/(111) | TC degradation | 70 times higher than defect free CaCu3Ti4O12 | 2019 | [ |
m-BiVO4(010)/(110) | O2 evolution | Higher than plate-like BiVO4 | 2016 | [ |
BiOI(001)/(110) | MO degradation | 71.4% MO degradation efficiency | 2018 | [ |
anatase TiO2(001)/(101)/SnS2 | MB degradation | 1.5 times better than pure TiO2 | 2017 | [ |
anatase TiO2(001)/(101)/ZnS | MB degradation | 2 times better than pure TiO2 | 2017 | [ |
BiVO4(010)/(110) | RhB degradation | 53.8% degradation efficiency within 180 min | 2018 | [ |
In2O3 loaded BiVO4(010)/(110) | RhB degradation | 91% degradation efficiency within 180 min | 2017 | [ |
anatase TiO2(001)/(101) | H2 production | a generation rate of 670 μmol h-1 g-1 | 2016 | [ |
CdS loaded TiO2(001)/(101) | 4-CP degradation | 96.12% degradation efficiency | 2018 | [ |
anatase TiO2(001)/(101) | CO2 Reduction | 21 times increasement | 2014 | [ |
Heterojunction | Type | Application | Performance | Year | Ref. |
---|---|---|---|---|---|
BiVO4/G/ZnSe | Z-Scheme | Water splitting | 9 times higher than ZnSe | 2019 | [ |
CsPbBr3/GO | Schottky | CO2 reduction | A rate of 23.7 μmol/g h with a selectivity over 99.3% | 2017 | [ |
BiFeO3/rGO | Schottky | MB degradation | Complete degradation within 100 min | 2018 | [ |
Ag2CrO4/g-C3N4/GO | S-Scheme | CO2 reduction | 1.5 times higher than Ag2CrO4/g-C3N4 | 2018 | [ |
TiO2/rGO | Schottky | RhB degradation | A rate of 53.27% within 300 min | 2017 | [ |
rGO/ZnO | Schottky | RhB degradation | A rate of 99.6% within 180 min | 2017 | [ |
GO QD/TiO2 | Type II | H2 evolution | 100 μmol h-1 | 2017 | [ |
(BiO)2CO3/BiO2-x/graphene | Z-Scheme | Removal of NO | 2 times higher than (BiO)2CO3 | 2019 | [ |
ZnIn2S4/G | Schottky | H2 evolution | A rate of 40.85 μmol g-1 h-1 | 2020 | [ |
TiO2/rGO-EDA | Schottky | H2 evolution | A rate of 224.9 μmol g-1 h-1 | 2020 | [ |
MoS2/N-doped G | S-Scheme | Degradation of NH3 | A degradation efficiency of 99.5% in 8 h | 2020 | [ |
N-TiO2/N-doped G | Type II | MB degradation | A degradation efficiency of 87.9% within 180 min | 2016 | [ |
Cu2O/G/TNA | Type II | CO2 conversion | Methanol yield of 45 μmol cm-2 h-1 | 2016 | [ |
P25/Ag3PO4/GO | Schottky | RhB degradation | 75% degradation efficiency within 1 h | 2015 | [ |
TiO2/rGO/WO3 | Z-Scheme | Disinfection | 97.3 ± 3.8% of E. coli inactivation within 80 mins | 2017 | [ |
ZnO/g-C3N4/rGO | Type II | MB degradation | 99.5% degradation efficiency within 15 min | 2015 | [ |
ZnO/MoS2/RGO | S-Scheme | H2 evolution | H2 production rate of 28.616 mmol h-1 gcat-1 | 2017 | [ |
Cd0.5Zn0.5S/g-C3N4/RGO | S-Scheme | H2 evolution | H2 production rate of 39.24 mmol g-1 h-1 | 2018 | [ |
ZnO/GQD | Type II | MB degradation | 95% degradation within 70 min | 2018 | [ |
GQD/AgVO3 | Type II | IBP degradation | 5 times higher than AgVO3 | 2016 | [ |
NaTaO3/g-C3N4/G | Type II | MB degradation | 99% removal of RhB within 70 min | 2019 | [ |
CdS/WS2/graphene | Schottky | H2 evolution | H2 production rate of 1842 μmol g-1 h-1 | 2016 | [ |
TiO2-x/rGO | Schottky | H2 production | H2 production rate of 16 mmol g-1 h-1 | 2016 | [ |
Table 7 Comparison of some recently developed graphene-based heterojunction photocatalysts.
Heterojunction | Type | Application | Performance | Year | Ref. |
---|---|---|---|---|---|
BiVO4/G/ZnSe | Z-Scheme | Water splitting | 9 times higher than ZnSe | 2019 | [ |
CsPbBr3/GO | Schottky | CO2 reduction | A rate of 23.7 μmol/g h with a selectivity over 99.3% | 2017 | [ |
BiFeO3/rGO | Schottky | MB degradation | Complete degradation within 100 min | 2018 | [ |
Ag2CrO4/g-C3N4/GO | S-Scheme | CO2 reduction | 1.5 times higher than Ag2CrO4/g-C3N4 | 2018 | [ |
TiO2/rGO | Schottky | RhB degradation | A rate of 53.27% within 300 min | 2017 | [ |
rGO/ZnO | Schottky | RhB degradation | A rate of 99.6% within 180 min | 2017 | [ |
GO QD/TiO2 | Type II | H2 evolution | 100 μmol h-1 | 2017 | [ |
(BiO)2CO3/BiO2-x/graphene | Z-Scheme | Removal of NO | 2 times higher than (BiO)2CO3 | 2019 | [ |
ZnIn2S4/G | Schottky | H2 evolution | A rate of 40.85 μmol g-1 h-1 | 2020 | [ |
TiO2/rGO-EDA | Schottky | H2 evolution | A rate of 224.9 μmol g-1 h-1 | 2020 | [ |
MoS2/N-doped G | S-Scheme | Degradation of NH3 | A degradation efficiency of 99.5% in 8 h | 2020 | [ |
N-TiO2/N-doped G | Type II | MB degradation | A degradation efficiency of 87.9% within 180 min | 2016 | [ |
Cu2O/G/TNA | Type II | CO2 conversion | Methanol yield of 45 μmol cm-2 h-1 | 2016 | [ |
P25/Ag3PO4/GO | Schottky | RhB degradation | 75% degradation efficiency within 1 h | 2015 | [ |
TiO2/rGO/WO3 | Z-Scheme | Disinfection | 97.3 ± 3.8% of E. coli inactivation within 80 mins | 2017 | [ |
ZnO/g-C3N4/rGO | Type II | MB degradation | 99.5% degradation efficiency within 15 min | 2015 | [ |
ZnO/MoS2/RGO | S-Scheme | H2 evolution | H2 production rate of 28.616 mmol h-1 gcat-1 | 2017 | [ |
Cd0.5Zn0.5S/g-C3N4/RGO | S-Scheme | H2 evolution | H2 production rate of 39.24 mmol g-1 h-1 | 2018 | [ |
ZnO/GQD | Type II | MB degradation | 95% degradation within 70 min | 2018 | [ |
GQD/AgVO3 | Type II | IBP degradation | 5 times higher than AgVO3 | 2016 | [ |
NaTaO3/g-C3N4/G | Type II | MB degradation | 99% removal of RhB within 70 min | 2019 | [ |
CdS/WS2/graphene | Schottky | H2 evolution | H2 production rate of 1842 μmol g-1 h-1 | 2016 | [ |
TiO2-x/rGO | Schottky | H2 production | H2 production rate of 16 mmol g-1 h-1 | 2016 | [ |
Fig. 24. (a) The schematical formation of CNNA/rGO composite in the molten salt medium. CNNA, rGO represents crystalline carbon nitride nanorods and reduced graphene oxide, respectively. (b) FESEM image of CNNA/rGO. (c) Photocatalytic CO2-reduction property of bulk CN (bCN), CNNA and CNNA/rGO at various amount of rGO in water vapor-saturated wet CO2 gas. (reproduced with permission [236], copyright 2019, Elsevier).
Fig. 25. (a) Schematic illustration of NG/CdS HS preparation process. (b-e) Corresponding SEM images of materials sequentially appear from left to right in (a). (f) Photocatalytic CO2 reduction performance of CdS, CdG1, CdG2, CdG3, and CdG5. (g) Schematic illustration of the charge carrier migration mechanism. CdGx: x donates that the dosage of pyridine during fabrication was 10x μL. (reproduced with permission [237], copyright 2019, Wiley).
Fig. 26. TEM (a) and HRTEM (b) images of the WO3/TiO2/rGO (WTG) composite. (c) Photocatalytic activities of various samples. (d) Schematic illustration of the S-scheme heterojunction-based charge transfer mechanism in WO3/TiO2/rGO composite (reproduced with permission [238], copyright 2020, Elsevier).
|
[1] | 蔡铭洁, 刘艳萍, 董珂欣, 陈晓波, 李世杰. 漂浮型Bi2WO6/C3N4/碳布S型异质结光催化材料用于高效净化水体环境[J]. 催化学报, 2023, 52(9): 239-251. |
[2] | 孙丽娟, 于晓慧, 唐丽永, 王伟康, 刘芹芹. 构建K3PW12O40/CdS核壳S型异质结实现同步太阳能光催化分解水和选择性苯甲醇氧化反应[J]. 催化学报, 2023, 52(9): 164-175. |
[3] | 江梓聪, 程蓓, 张留洋, 张振翼, 别传彪. 氧化锌基梯型异质结光催化剂[J]. 催化学报, 2023, 52(9): 32-49. |
[4] | 刘博文, 蔡家杰, 张建军, 谭海燕, 程蓓, 许景三. MOF/CdS梯型光催化剂同时进行苯甲醇氧化和析氢反应及其机理研究[J]. 催化学报, 2023, 51(8): 204-215. |
[5] | 宋明明, 宋相海, 刘鑫, 周伟强, 霍鹏伟. ZnIn2S4/MOF-808微球结构S型异质结光催化剂的制备及其光还原CO2性能研究[J]. 催化学报, 2023, 51(8): 180-192. |
[6] | 邵秀丽, 李可, 李静萍, 程强, 王国宏, 王楷. 揭示NiS@Ta2O5纳米纤维中梯型电荷转移路径及光催化CO2转化性能[J]. 催化学报, 2023, 51(8): 193-203. |
[7] | 李世杰, 王春春, 董珂欣, 张鹏, 陈晓波, 李鑫. 新型MIL-101(Fe)/BiOBr S型异质催化剂用于高效光催化降解抗生素和还原六价铬: 光催化性能分析和光催化机理研究[J]. 催化学报, 2023, 51(8): 101-112. |
[8] | 孙利娟, 王伟康, 路平, 刘芹芹, 王乐乐, 唐华. 纳米高熵合金实现光催化剂肖特基势垒的调控用于光催化制氢与苯甲醇氧化耦合反应[J]. 催化学报, 2023, 51(8): 90-100. |
[9] | 袁鑫, 范海滨, 刘杰, 覃龙州, 王剑, 段秀, 邱江凯, 郭凯. 连续流技术在光氧化还原催化转化的最新进展[J]. 催化学报, 2023, 50(7): 175-194. |
[10] | 林敏, 罗美兰, 柳勇志, 沈锦妮, 龙金林, 张子重. 一维S型异质结W18O49-SiC海胆状复合催化剂增强光催化CO2还原[J]. 催化学报, 2023, 50(7): 239-248. |
[11] | 余治晗, 关晨, 岳晓阳, 向全军. 碳环渗入的结晶氮化碳S型同质结及其光催化析氢[J]. 催化学报, 2023, 50(7): 361-371. |
[12] | 朱君江, S. Wageh, Ahmed A. Al-Ghamdi. 利用飞秒技术研究电荷转移动力学[J]. 催化学报, 2023, 49(6): 5-7. |
[13] | 夏思源, 李奇远, 张仕楠, 许冬, 林秀, 孙露菡, 许景三, 陈接胜, 李国栋, 李新昊. Pd纳米立方体异质结尺寸依赖的电子界面效应对苯酚加氢反应的促进作用[J]. 催化学报, 2023, 49(6): 180-187. |
[14] | 张海波, 王中辽, 张金锋, 代凯. 金属硫化物基异质结光催化剂: 原理、影响、应用和原位表征[J]. 催化学报, 2023, 49(6): 42-67. |
[15] | Mengistu Tulu Gonfa, 申升, 陈浪, 胡彪, 周威, 白张君, 区泽堂, 尹双凤. 光催化苯制苯酚的研究进展[J]. 催化学报, 2023, 49(6): 16-41. |
阅读次数 | ||||||
全文 |
|
|||||
摘要 |
|
|||||