22
REVIEW 1700080 (1 of 21) www.small-methods.com © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim A Review of Direct Z-Scheme Photocatalysts Jingxiang Low, Chuanjia Jiang, Bei Cheng, Swelm Wageh, Ahmed A. Al-Ghamdi, and Jiaguo Yu* DOI: 10.1002/smtd.201700080 emerged as one of the most promising photoconversion technologies because of its ability to directly utilize solar energy for producing solar fuels such as hydrogen (H 2 ), methane (CH 4 ), methanol (CH 3 OH), formic acid (HCOOH), and formaldehyde (CH 2 O), and alleviating environmental pollution. [19–23] Despite the significant progress in photocatalysis that has been achieved in recent years, the photoconversion efficiency of photo- catalytic reactions is still low, and they are far from practical application, because of the rapid electron–hole recombination and poor light utilization of semiconduc- tors. [24–27] Therefore, it is of importance to search for advanced photocatalytic sys- tems with efficient light utilization and electron–hole separation. In the past several decades, diverse strategies have been studied to enhance the photoconversion efficiency of photo- catalysts, including doping, metal loading, heterojunction construction, etc. Among these proposed strategies, the com- bination of two semiconductors to form a type-II heterojunc- tion photocatalyst is one of the most facile ways for enhancing photocatalytic performance, due to its effectiveness for spatially separating the photogenerated electron–hole pairs through the band alignment between two semiconductors. [25,28] For a typical type-II heterojunction, both the conduction band (CB) and valence band (VB) of semiconductor A are higher than those of semiconductor B (Figure 1a). [29–31] Therefore, the photogen- erated electrons in the CB of semiconductor A will migrate to the CB of semiconductor B under light irradiation due to band alignment. Meanwhile, the photogenerated holes in the VB of semiconductor B will transfer to the VB of semiconductor A. Since photogenerated electrons and holes respectively accumu- late on semiconductor B and semiconductor A, spatial separa- tion of the electrons and holes can be achieved with a type-II heterojunction photocatalyst for enhancing the photocatalytic activity. [32–34] However, there are several obvious problems that restrict the wide application of type-II heterojunction photocata- lysts. In detail, the reduction and oxidation reactions of type-II heterojunction photocatalysts, respectively, occur for semicon- ductor B with a lower reduction potential and semiconductor A with a lower oxidation potential. [35,36] Therefore, the redox ability of type-II heterojunction photocatalysts will be greatly reduced. [37] In addition, it is difficult for electrons in semicon- ductor A and holes in semiconductor B to respectively migrate to the electron-rich CB of semiconductor B and the hole-rich Recently, great attention has been paid to fabricating direct Z-scheme photo- catalysts for solar-energy conversion due to their effectiveness for spatially separating photogenerated electron–hole pairs and optimizing the reduc- tion and oxidation ability of the photocatalytic system. Here, the historical development of the Z-scheme photocatalytic system is summarized, from its first generation (liquid-phase Z-scheme photocatalytic system) to its current third generation (direct Z-scheme photocatalyst). The advantages of direct Z-scheme photocatalysts are also discussed against their predecessors, including conventional heterojunction, liquid-phase Z-scheme, and all-solid- state (ASS) Z-scheme photocatalytic systems. Furthermore, characteriza- tion methods and applications of direct Z-scheme photocatalysts are also summarized. Finally, conclusions and perspectives on the challenges of this emerging research direction are presented. Insights and up-to-date informa- tion are provided to give the scientific community the ability to fully explore the potential of direct Z-scheme photocatalysts in renewable energy produc- tion and environmental remediation. Photocatalysts J. X. Low, Dr. C. J. Jiang, Prof. B. Cheng, Prof. J. G. Yu State Key Laboratory of Advanced Technology for Materials Synthesis and Processing Wuhan University of Technology 122 Luoshi Road, Wuhan 430070, P. R. China E-mail: [email protected], [email protected] Prof. S. Wageh, Prof. A. A. Al-Ghamdi, Prof. J. G. Yu Department of Physics Faculty of Science King Abdulaziz University Jeddah 21589, Saudi Arabia The ORCID identification number(s) for the author(s) of this article can be found under http://dx.doi.org/10.1002/smtd.201700080. 1. Introduction Due to rapid industrialization and population growth, the global energy crisis and environmental pollution have become two of the greatest challenges of human society in the 21st century. [1–5] The utilization of solar energy, which is a pow- erful, affordable, and renewable energy source, for energy production and pollutant elimination, is regarded as the best solution to these critical problems. [6–9] Therefore, much effort has been devoted toward converting solar energy into an appli- cable energy medium through various technologies, such as photocatalysis, [10,11] solar cells, [12,13] photoelectrochemical cells, [14–16] and so on. [17,18] Among them, photocatalysis has Small Methods 2017, 1, 1700080

A Review of Direct Z‐Scheme Photocatalysts

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: A Review of Direct Z‐Scheme Photocatalysts

REVIEW

1700080 (1 of 21)

www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

A Review of Direct Z-Scheme Photocatalysts

Jingxiang Low, Chuanjia Jiang, Bei Cheng, Swelm Wageh, Ahmed A. Al-Ghamdi, and Jiaguo Yu*

DOI: 10.1002/smtd.201700080

emerged as one of the most promising photoconversion technologies because of its ability to directly utilize solar energy for producing solar fuels such as hydrogen (H2), methane (CH4), methanol (CH3OH), formic acid (HCOOH), and formaldehyde (CH2O), and alleviating environmental pollution.[19–23] Despite the significant progress in photocatalysis that has been achieved in recent years, the photoconversion efficiency of photo­catalytic reactions is still low, and they are far from practical application, because of the rapid electron–hole recombination and poor light utilization of semiconduc­tors.[24–27] Therefore, it is of importance to search for advanced photocatalytic sys­tems with efficient light utilization and electron–hole separation.

In the past several decades, diverse strategies have been studied to enhance the photoconversion efficiency of photo­

catalysts, including doping, metal loading, heterojunction construction, etc. Among these proposed strategies, the com­bination of two semiconductors to form a type­II heterojunc­tion photocatalyst is one of the most facile ways for enhancing photocatalytic performance, due to its effectiveness for spatially separating the photogenerated electron–hole pairs through the band alignment between two semiconductors.[25,28] For a typical type­II heterojunction, both the conduction band (CB) and valence band (VB) of semiconductor A are higher than those of semiconductor B (Figure 1a).[29–31] Therefore, the photogen­erated electrons in the CB of semiconductor A will migrate to the CB of semiconductor B under light irradiation due to band alignment. Meanwhile, the photogenerated holes in the VB of semiconductor B will transfer to the VB of semiconductor A. Since photogenerated electrons and holes respectively accumu­late on semiconductor B and semiconductor A, spatial separa­tion of the electrons and holes can be achieved with a type­II heterojunction photocatalyst for enhancing the photocatalytic activity.[32–34] However, there are several obvious problems that restrict the wide application of type­II heterojunction photocata­lysts. In detail, the reduction and oxidation reactions of type­II heterojunction photocatalysts, respectively, occur for semicon­ductor B with a lower reduction potential and semiconductor A with a lower oxidation potential.[35,36] Therefore, the redox ability of type­II heterojunction photocatalysts will be greatly reduced.[37] In addition, it is difficult for electrons in semicon­ductor A and holes in semiconductor B to respectively migrate to the electron­rich CB of semiconductor B and the hole­rich

Recently, great attention has been paid to fabricating direct Z-scheme photo-catalysts for solar-energy conversion due to their effectiveness for spatially separating photogenerated electron–hole pairs and optimizing the reduc-tion and oxidation ability of the photocatalytic system. Here, the historical development of the Z-scheme photocatalytic system is summarized, from its first generation (liquid-phase Z-scheme photocatalytic system) to its current third generation (direct Z-scheme photocatalyst). The advantages of direct Z-scheme photocatalysts are also discussed against their predecessors, including conventional heterojunction, liquid-phase Z-scheme, and all-solid-state (ASS) Z-scheme photocatalytic systems. Furthermore, characteriza-tion methods and applications of direct Z-scheme photocatalysts are also summarized. Finally, conclusions and perspectives on the challenges of this emerging research direction are presented. Insights and up-to-date informa-tion are provided to give the scientific community the ability to fully explore the potential of direct Z-scheme photocatalysts in renewable energy produc-tion and environmental remediation.

Photocatalysts

J. X. Low, Dr. C. J. Jiang, Prof. B. Cheng, Prof. J. G. YuState Key Laboratory of Advanced Technology for Materials Synthesis and ProcessingWuhan University of Technology122 Luoshi Road, Wuhan 430070, P. R. ChinaE-mail: [email protected], [email protected]. S. Wageh, Prof. A. A. Al-Ghamdi, Prof. J. G. YuDepartment of PhysicsFaculty of ScienceKing Abdulaziz UniversityJeddah 21589, Saudi Arabia

The ORCID identification number(s) for the author(s) of this article can be found under http://dx.doi.org/10.1002/smtd.201700080.

1. Introduction

Due to rapid industrialization and population growth, the global energy crisis and environmental pollution have become two of the greatest challenges of human society in the 21st century.[1–5] The utilization of solar energy, which is a pow­erful, affordable, and renewable energy source, for energy production and pollutant elimination, is regarded as the best solution to these critical problems.[6–9] Therefore, much effort has been devoted toward converting solar energy into an appli­cable energy medium through various technologies, such as photocatalysis,[10,11] solar cells,[12,13] photoelectrochemical cells,[14–16] and so on.[17,18] Among them, photocatalysis has

Small Methods 2017, 1, 1700080

Page 2: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (2 of 21)

www.advancedsciencenews.com www.small-methods.com

VB of semiconductor A, due to electrostatic repulsion between electron–electron or hole–hole. Thus, it is highly desired to pre­pare new heterostructured photocatalytic systems to overcome these problems.

Biomimetic artificial photosynthesis through building of the direct Z­scheme photocatalyst represents a viable strategy for enhancing the photocatalytic performance.[38,39] In 2013, the concept of a direct Z­scheme photocatalyst was proposed by Yu et al. for explaining the high photocatalytic formalde­hyde (HCHO) degradation performance of the TiO2/g­C3N4 composite. Specifically, the structure of a direct Z­scheme pho­tocatalyst is similar to that of a type­II heterojunction photocat­alyst (Figure 1a,b), but its charge­carrier migration mechanism is different. Specifically, a typical direct Z­scheme system has a charge­carrier migration pathway that resembles the letter “Z” (Figure 1b).[40] During the photocatalytic reaction, the photo­generated electrons in semiconductor B, with lower reduction ability, recombine with the photogenerated holes in semicon­ductor A with a lower oxidation ability (Figure 1b).[41] Therefore, the photogenerated electrons in semiconductor A with high reduction ability and photogenerated holes in semiconductor B with a high oxidation ability can be maintained. As a result, the redox ability of the direct Z­scheme photocatalyst can be optimized. In addition, it should be noted that charge­carrier migration for the direct Z­scheme photocatalyst is physically more feasible than that for type­II heterojunction photocata­lysts, since the migration of photogenerated electrons from the CB of semiconductor B to the photogenerated hole­rich VB of semiconductor A is favorable due to the electrostatic attraction between the electron and hole.

Since 2013, huge accomplishments related to direct Z­scheme photocatalysts for photocatalytic applications have been achieved by numerous research groups. Therefore, it is of significance to summarize the recent discoveries and achieve­ments in the field of direct Z­scheme photocatalysts. Here, the development and basic principle of direct Z­scheme photo­catalysts are discussed. Then, the characterization methods for direct Z­scheme photocatalysts are summarized. The recent advances and trends in the study of direct Z­scheme photocata­lysts for various photocatalytic applications are also presented. Finally, the conclusions and perspectives for future research on the direct Z­scheme photocatalyst are provided.

2. Historical Development of Direct Z-Scheme Photocatalysts

In order to have a comprehensive understanding of direct Z­scheme photocatalysts, it is of significance to discuss devel­opment of the Z­scheme photocatalytic system from the 1st generation to the current 3rd generation (Figure 2). The con­cept of the Z­scheme photocatalytic system was originally proposed by Bard in 1979.[42] As shown in Figure 3, the liquid­phase Z­scheme photocatalytic system is built by combining two different semiconductors with a shuttle redox mediator (viz. an electron acceptor/donor (A/D) pair). Under light irra­diation, both semiconductor A and semiconductor B are photo­excited, thereby generating electrons and holes respectively in their CB and VB. Then, the photogenerated electrons from

semiconductor B will transfer to the VB of semiconductor A via a shuttle redox mediator according to Reaction (1) and (2), leaving photogenerated holes in the VB of semiconductor B.

Jingxiang Low obtained his B.Eng. (Hons) from Multimedia University, Malaysia in 2011 and his M.S. in materials science from Wuhan University of Technology. He is currently a Ph.D. candidate under the supervision of Prof. Jiaguo Yu at the State Key Laboratory of Advanced Technology for Materials Synthesis and

Processing, Wuhan University of Technology. His current research includes photocatalytic H2 production and CO2 reduction. See more details at: http://www.researcherid.com/rid/N-2381-2014.

Chuanjia Jiang received his B.S. and M.S. in environ-mental engineering from Tsinghua University, Beijing in 2009 and 2011, respectively, and his Ph.D. in civil and environmental engineering in 2016 from Duke University. In 2016, he joined Wuhan University of Technology as an Assistant Professor. His current research inter-

ests include materials for air and water pollution con-trol. See more details at: http://www.researcherid.com/rid/C-9398-2014.

Jiaguo Yu received his B.S. and M.S. in chemistry from Central China Normal University and Xi’an Jiaotong University, respectively, and his Ph.D. in materials sci-ence in 2000 from Wuhan University of Technology. In 2000, he became a Professor at Wuhan University of Technology. He was a post-doctoral fellow at the Chinese

University of Hong Kong from 2001 to 2004, a visiting sci-entist from 2005 to 2006 at the University of Bristol, and a visiting scholar from 2007 to 2008 at the University of Texas at Austin. His current research interests include semicon-ductor photocatalysis, photocatalytic hydrogen production, CO2 reduction to hydrocarbon fuels, and so on. See more details at: http://www.researcherid.com/rid/G-4317-2010.

Small Methods 2017, 1, 1700080

Page 3: A Review of Direct Z‐Scheme Photocatalysts

1700080 (3 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

( )+ →−A e @CB of semiconductor B D (1)

( )+ →+D h @VB of semiconductor A A (2)

Therefore, the photogenerated electrons remain on semicon­ductor A with the higher reduction potential, while the pho­togenerated holes remain on semiconductor B with the higher oxidation potential, thereby achieving the optimization of the redox potential of the photocatalytic system.

Although the 1st generation Z­scheme photocatalytic system is efficient for photocatalytic reactions, it suffers from several obvious problems. First, backward reactions exist due to the use of reversible redox mediators such as I−/IO3− and Fe2+/Fe3+. For example, the acceptor, such as Fe3+, and the donor, such as Fe2+, will compete with the reactants for reduction and oxidation reaction, respectively during the photocatalytic reac­tion. Therefore, the photoconversion efficiency of the Z­scheme photocatalytic system will be greatly reduced. Second, the 1st generation Z­scheme photocatalytic system can only be applied in the liquid phase. Therefore, its practical application in the

gas and solid phases is greatly limited. In order to solve these problems, Tada et al. proposed the concept of the 2nd genera­tion Z­scheme photocatalytic system, namely the all­solid­state (ASS) Z­scheme photocatalytic system, in 2006 (Figure 4).[43] An ASS Z­scheme photocatalytic system is composed of two dif­ferent semiconductors and a noble­metal nanoparticle (NP) as the electron mediator. Since the noble­metal NP is used as an electron mediator in the ASS Z­scheme photocatalytic system, the backward reaction of the first­generation Z­scheme pho­tocatalytic system can be inhibited. However, the use of noble metals, which are rare and expensive, greatly limits the wide application of the ASS Z­scheme photocatalytic system. More­over, noble­metal NPs are normally strong light absorbers.[44,45] Therefore, the light­absorption ability of the photocatalyst will be also greatly reduced by construction of the ASS Z­scheme photocatalyst.

Thereafter, Wang et al. prepared a mediator­free ASS Z­scheme photocatalytic system in 2009.[46] It was found that the Z­scheme electron–hole transfer mechanism can be applied to ZnO and CdS when they are in intimate contact. In 2013, our group proposed the concept of the third­generation Z­scheme photocatalytic system, namely the direct Z­scheme photocatalyst (Figure 1b).[47] The advantages of the previous two generations of Z­scheme photocatalytic systems are fully inher­ited by the direct Z­scheme photocatalyst, including improved electron–hole separation efficiency and optimized redox poten­tial. In detail, a direct Z­scheme photocatalyst consists of only two semiconductors that have direct contact at their interface. In comparison with the previous two generations, electron or hole mediators are not required in the direct Z­scheme photo­catalyst. Therefore, the construction cost of the Z­scheme photocatalytic system can be greatly reduced. Moreover, the light­shielding effect caused by the loading of the metal­based mediator can also be overcome by building the direct Z­scheme photocatalyst. Owing to these advantages, the direct Z­scheme photocatalyst has been widely studied for various photocatalytic applications (see Table 1).[40,48,49]

3. Characterization Methods for Direct Z-Scheme Photocatalysts

As discussed in Section 1, the structure of a direct Z­scheme photocatalyst is similar to that of a type­II heterojunction photocata­lyst. Therefore, it is imperative to investigate the charge­carrier migration mechanism for the direct Z­scheme photocatalyst through various characterization methods, in order to differentiate it from type­II heterojunc­tion photocatalysts. To date, various char­acterization methods have been proposed for this purpose, including photocatalytic­reduction testing, radical species trapping, metal loading, X­ray photoelectron spec­troscopy (XPS), effective mass calculation, and internal electric­field simulation. The following subsections elucidate these char­acterization methods. It should be kept

Small Methods 2017, 1, 1700080

Figure 1. Comparison of the charge-carrier separation mechanism on the type-II heterojunction (a) and the direct Z-scheme (b) photocatalyst built on two different semiconductors (SCs).

Figure 2. The roadmap of the evolution of Z-scheme photocatalytic system from the 1st gen-eration to the 3rd generation.

Page 4: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (4 of 21)

www.advancedsciencenews.com www.small-methods.com

in mind that utilizing only a single characterization method cannot give precise information on the charge­carrier migra­tion mechanism for the direct Z­scheme photocatalyst. Thus, a comprehensive investigation is always required to confirm the formation of direct Z­scheme photocatalysts through a combi­nation of different characterization methods.

3.1. Photocatalytic-Reduction Testing

It is well known that not all the photogenerated electrons reaching the surface of a photocatalyst can be utilized for the photocatalytic­reduction reaction. Specifically, only the photogen­erated electrons in a semiconductor with sufficient reduction potential can be applied for specific photocatalytic­reduction reactions. The standard redox potentials of various photocata­lytic reduction reaction are shown in Reaction (3) to (8):

+ →= − =

+ −2H 2e H ,0.41V vs NHE at pH 7

2

0E

(3)

+ + →= − =

+ −CO 2H 2e HCOOH,0.61V vs NHE at pH 7

2

0E

(4)

+ + → += − =

+ −CO 2H 2e CO H O,0.53V vs NHE at pH 7

2 2

0E

(5)

+ + → += − =

+ −CO 4H 4e HCHO H O,0.48V vs NHE at pH 7

2 2

0E

(6)

+ + → += − =

+ −CO 6H 6e CH OH H O,0.38V vs NHE at pH 7

2 3 2

0E

(7)

+ + → += − =

+ −CO 8H 8e CH 2H O,0.24V vs NHE at pH 7

2 4 2

0E

(8)

Therefore, it is feasible to determine the accumulation of photogenerated electrons in specific semiconductors of a direct Z­scheme photocatalyst through an investigation of the final product of the photocatalytic­reduction reaction. For example, our group confirmed the formation of the CdS/WO3 direct Z­scheme photocatalyst by determining the products of photocat­alytic CO2 reduction (Figure 5a,b).[61] It was found that the WO3 shows no photocatalytic CO2 activity for CH4 production due to its low reduction potential (0.5 V vs normal hydrogen electrode (NHE)) (Figure 5c). Meanwhile, the prepared CdS exhibited good photocatalytic CO2­reduction activity for CH4 production because of its sufficient reduction potential (−0.6 V vs NHE) (Figure 5c). Moreover, all the prepared CdS/WO3 composites exhibited better CH4 production efficiency. This result is because of the forma­tion of the direct Z­scheme heterojunction between CdS and WO3, which can enhance the electron–hole separation efficiency and optimize the reduction ability of the CdS/WO3. In detail, according to the direct Z­scheme mechanism, the photogen­erated electrons will accumulate on the CdS, which has suffi­cient reduction potential for reduction of CO2 into CH4. If the photogenerated electrons migrate according to the conventional type­II heterojunction, no CH4 can be produced by using CdS/WO3. This is because, according to the conventional heterojunc­tion mechanism, the photogenerated electrons will accumulate on the WO3, which does not have sufficient reduction potential. This result confirms the formation of a direct Z­scheme photo­catalytic system instead of a type­II heterojunction between the CdS and the WO3 (Figure 5c). Moreover, it should be noted that CH4, which require more electrons (8 electrons) and less reduc­tion potential (−0.24 V vs NHE) is the main reaction product during the photocatalytic CO2­reduction test using the CdS/WO3 composite (Figure 5a). This is due to the accumulation of photo­generated electrons on a specific semiconductor, which is ben­eficial for the multi­electron reaction of CH4 production. This study demonstrates that the photocatalytic reduction test can be a viable method to investigate the charge­carrier migration mechanism for the direct Z­scheme photocatalyst.

3.2. Radical Species Trapping Test

It is well established that the hydroxyl (·OH) radical can be generated on semiconductors that can produce photogenerated

Small Methods 2017, 1, 1700080

Figure 3. Schematic illustration of the 1st generation liquid-phase Z-scheme photocatalytic system, where A and D respectively represent the electron acceptor and donor.

Figure 4. Schematic illustration of the 2nd generation Z-scheme photo-catalytic system, all-solid-state (ASS) Z-scheme photocatalytic system.

Page 5: A Review of Direct Z‐Scheme Photocatalysts

1700080 (5 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimSmall Methods 2017, 1, 1700080

Table 1. Studies of direct Z-scheme heterojunction for various photocatalytic applications.

Photocatalyst Light source Application Performance Ref.

Ag2CrO4–graphene oxide 300 W xenon arc lamp

(400 nm cutoff filter)

Methylene blue degradation k = 2.8 × 10−1 min−1 [50]

g-C3N4/TiO2 3 W 365 nm UV lamp Brilliant red X3B degradation k = 5.1 × 10−2 min−1 [51]

CaIn2S4/TiO2 200 W mercury vapor lamp Metronidazole degradation k = 1.91 × 10−2 min−1 [52]

Bi2O3/g-C3N4 500 W xenon arc lamp (400 nm cutoff filter) Rhodamine B degradation k = 1.01 × 10−2 min−1 [53]

g-C3N4–TiO2 15 W 365 nm UV lamp Formaldehyde degradation k = 7.36 × 10−2 min−1 [47]

ZnIn2S4/Bi2WO6 500 W tungsten–halogen lamp Metronidazole degradation k = 1.66 × 10−2 min−1 [54]

Bi2O3/NaNbO3 375 W mercury lamp Rhodamine B degradation k = 5.86 × 10−2 min−1 [55]

g-C3N4–Ag3PO4 300 W xenon arc lamp C2H4 degradation k = 1.05 h−1 [56]

NaNbO3/WO3 375 W mercury lamp Rhodamine B and Methylene

blue degradationk = 3.7 × 10−2 min−1 and

9.6 × 10−2 min−1

[57]

g-C3N4/Ag2CO3 300 W xenon arc lamp (400 nm cutoff filter) Rhodamine B degradation k = 1.36 × 10−1 min−1 [58]

ZnO/CdS 400 W xenon arc lamp H2 production 1007 mmol h−1g−1 [59]

Anatase/rutile TiO2 350 W xenon arc lamp H2 production 324 μmol h−1g−1 [60]

C,N-TiO2/g-C3N4 300 W xenon arc lamp (400 nm cutoff filter) H2 production 39.18 mmol h−1g−1 [27]

CdS–WO3 300 W xenon arc lamp (420 nm cutoff filter) CO2 reduction 1.02 μmol h−1g−1 CH4 [61]

g-C3N4/ZnO 300 W xenon arc lamp CO2 reduction 0.6 μmol h−1g−1 CH3OH [62]

Cu2O/Fe2O3 300 W xenon arc lamp (400 nm cutoff filter) CO2 reduction 1.67 μmol h−1gcat−1 CO [63]

Figure 5. a) Gas-chromatography detection of the photocatalytic CO2 reduction products over 5 mol% CdS-loaded WO3 (C5). b) Performance of photo-catalytic CO2 reduction to CH4 over WO3 (C0), pure CdS (C100), and CdS-loaded WO3 samples with different CdS loadings: 1 mol% (C1), 2 mol% (C2), C5, 10 mol% (C10), 20 mol% (C20). c) Schematic illustration of the charge-carrier migration mechanism on the CdS/WO3 type-II heterojunction and direct Z-scheme photocatalyst. Reproduced with permission.[61] Copyright 2015, Wiley-VCH.

Page 6: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (6 of 21)

www.advancedsciencenews.com www.small-methods.com

holes with sufficient oxidation potential (2.4 V vs NHE), while the superoxide (·O2−) radical can be produced from photogen­erated electrons in a semiconductor with sufficient reduction potential (−0.33 V vs NHE) under light irradiation.[64–66] Since direct Z­scheme photocatalysts and type­II heterojunction photo catalysts can respectively maximize and minimize the redox potential of the photocatalyst, it is feasible to differentiate these two photocatalytic systems by investigating the produc­tion efficiency of the radical species on the photocatalyst.

3.2.1. ·OH-Radical Trapping Test

The hydroxyl­radical trapping test is one of the most com­monly applied methods to confirm the formation of a direct Z­scheme photocatalyst. Specifically, the ·OH radical can be produced by the reaction of OH−/H2O with the photogenerated hole on the semiconductor with an oxidation potential greater than 2.4 V vs NHE. Therefore, the ·OH radical trapping test is a simple yet effective method to confirm the accumula­tion of photogenerated holes on either semiconductor with a high oxidation potential or a low oxidation potential. Typi­cally, the production of ·OH radicals can be investigated by photoluminescence (PL) spectroscopy, using terephthalic acid (TA) as a probe molecule. The nonfluorescent TA molecules can react with an ·OH radical to produce highly fluorescent 2­hydroxyterephthalic acid (HTA), which is detectable by the PL spectrometer. The PL peak intensity of the HTA is in pro­portion to the number of produced ·OH radicals. For example, the PL spectral changes of the g­C3N4/TiO2 direct Z­scheme photo catalyst with different illumination times are shown in Figure 6a.[47] Obviously, the PL intensity of the HTA in the sample increases with increasing illumination time. This result indicates that the photogenerated holes on the g­C3N4/TiO2 composite has sufficient oxidation potential for gener­ating ·OH radicals, which can react with TA to produce HTA. Since the photogenerated holes on g­C3N4 do not have suffi­cient oxidation potential to react with OH−/H2O for generating ·OH radicals, it can be concluded that the photogenerated holes accumulate on the TiO2 (Figure 6b). The accumulation of photogenerated holes on TiO2 is in accordance with the charge­carrier migration mechanism of the direct Z­scheme

photocatalyst. Therefore, the formation of the direct Z­scheme system instead of the type­II heterojunction between TiO2 and g­C3N4 can be confirmed.

3.2.2. ·O2− Radical Trapping Test

Similarly, the electron­generation efficiency of the photocatalyst can be examined by investigating the generation of ·O2− radical ions, which are normally determined by electron spin resonance (ESR) spectroscopy. It should be noted that the ·O2− radical ions are not directly detectable by the ESR spectrometer. Therefore, 5,5­dimethyl­pyrroline N­oxide (DMPO) is used to react with ·O2− to form DMPO­·O2−, which has obvious ESR characteristic peaks. For instance, the ESR signals of DMPO­·O2− produced by g­C3N4, Ag3PO4/g­C3N4, and Ag3PO4 under 30 s light irra­diation in methanol are shown in Figure 7a.[56] Basically, six ESR characteristic peaks of DMPO­·O2− can be found for pure g­C3N4 and the Ag3PO4/g­C3N4 composite. Meanwhile, no peak can be observed in the ESR spectra of Ag3PO4 and the blank test, indicating that no ·O2− is produced for these samples. The results indicate that the photogenerated electrons mainly accumulate in the g­C3N4 of the Ag3PO4/g­C3N4 composite which has sufficiently high reduction potential to produce ·O2−. Therefore, the migration of photogenerated electrons in the Ag3PO4/g­C3N4 follows the direct Z­scheme instead of the type­II heterojunction transfer mechanism (Figure 7b).

3.3. Metal Loading

The photo­deposition of metal NPs is widely applied to directly confirm the reduction site of photocatalytic system, thus pro­viding information on the migration pathway of charge car­riers. In detail, metal precursors such as HAuCl4, H2PtCl6, and AgNO3 are firstly dissolved in water to create a metal­ion solu­tion. Under light irradiation, these metal ions will be reduced into metal NPs and in situ deposited at the electron­rich reduc­tion site of the photocatalytic system by reacting with the photo­generated electrons according to Reaction (9):

+ →+ − 0M ne Mn

(9)

Small Methods 2017, 1, 1700080

Figure 6. a) Photoluminescence (PL) spectra of the TiO2/g-C3N4 under different light irradiation periods in the NaOH solution (2 × 10−3 m) in the presence of terephthalic acid (5 × 10−4 m), and b) comparison of the redox potential of TiO2 and g-C3N4 for producing ·OH radicals and ·O2− radicals. Reproduced with permission.[47] Copyright 2013, The Royal Society of Chemistry.

Page 7: A Review of Direct Z‐Scheme Photocatalysts

1700080 (7 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Therefore, the accumulation of photogenerated electrons on a specific semiconductor of the direct Z­scheme photocata­lyst can be confirmed. For example, Xu et al. applied the metal photo­deposition method to confirm the electron­migration pathway for the rutile/anatase TiO2 direct Z­scheme photo­catalyst.[60] As shown in the transmission electron microscopy (TEM) and high­resolution TEM images (Figure 8), the Pt NPs are selectively loaded on rutile TiO2, indicating that the photo­generated electrons transfer to rutile TiO2 instead of anatase TiO2, which is consistent with the electron­migration mecha­nism of the direct Z­scheme photocatalyst.

3.4. XPS Characterization

XPS characterization has been extensively used to determine the surface chemistry of materials. Recently, the application of the XPS characterization has been extended to investigate the changes in the electronic density on the different surfaces of a photocatalyst through investigating the shift in the binding ener­gies. In detail, the introduction of a foreign material on a semi­conductor can cause a shift in the binding energy of a specific element of the semiconductor if the electron migration occurs on its surface. Specifically, a positive shift in the binding energy indicates a decrease of the electron density, whereas a negative shift indicates an increase in the electron density. Therefore, the shift of binding energy in the XPS spectra can be utilized for determining the electron­migration pathway for a direct Z­scheme photocatalyst. For instance, Peng and co­workers confirmed the formation of a direct Z­scheme heterojunction between g­C3N4 and ZnO through XPS characterization.[62] The high­resolution XPS spectra of N 1s of g­C3N4 and the ZnO/g­C3N4 composite are shown in Figure 9a, and those of O 1s of ZnO and the ZnO/g­C3N4 composite are shown in Figure 9b. Compared with pure g­C3N4, the N 1s states of the ZnO/g­C3N4 composite were found to shift toward the lower­energy region (Figure 9a). Meanwhile, the O 1s states of the ZnO/g­C3N4 com­posite shifted toward the higher­energy region in comparison with that of pure ZnO (Figure 9b). This result indicates the migration of electrons from ZnO to g­C3N4 on the ZnO/g­C3N4 composite. This migration mechanism of electrons is attributed to the formation of a weak internal electric field between the ZnO and the g­C3N4. Specifically, the Fermi level of g­C3N4 is higher than that of ZnO (Figure 9c). When ZnO is coupled with g­C3N4, electrons on the g­C3N4 will migrate to ZnO to obtain a Fermi level equilibrium, leaving a positive charge on the g­C3N4 (Figure 9d). Thus, the interface between the ZnO and the g­C3N4 is charged, thereby creating a weak internal electric field at the interface. During XPS characterization, both the ZnO and the g­C3N4 are photoexcited, and the photogenerated electrons will migrate from the ZnO to the g­C3N4 under the influence of the internal electric field (Figure 9e). This migration mecha­nism of the photogenerated electrons is in accordance with the direct Z­scheme mechanism. Therefore, the formation of the direct Z­scheme instead of a type­II heterojunction between the ZnO and the g­C3N4 can be confirmed.

Small Methods 2017, 1, 1700080

Figure 7. a) Electron spin resonance (ESR) spectra for DMPO-·O2− signal of different samples under 30 s light illumination, and b) schematic illus-tration for the charge-carrier transfer mechanism for a Ag3PO4/g-C3N4 type-II heterojunction and a direct Z-scheme photocatalyst. Reproduced with permission.[56] Copyright 2015, The Royal Society of Chemistry.

Figure 8. a,b) TEM (a) and high-resolution TEM (b) images of the rutile/anatase TiO2 direct Z-scheme photocatalyst with the loading of Pt NPs. Reproduced with permission.[60] Copyright 2014, Elsevier B.V.

Page 8: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (8 of 21)

www.advancedsciencenews.com www.small-methods.com

3.5. Effective-Mass Calculation

Supplementary to the above experimental methods, theoretical simulation is also a powerful method to evaluate the charge­carrier­migration mechanism in a photocatalytic system. Generally, first­principles simulation based on a density func­tional theory (DFT) calculation is carried out to analyze the charge­carrier generation and migration for a specific photocat­alyst by analyzing the effective masses of its charge carriers. For example, Yu et al. demonstrated that theoretical material simu­lation is an effective method for investigating the formation of the direct Z­scheme system for a ZnO/g­C3N4 composite.[62] In detail, the electronic­band structure of ZnO and g­C3N4 were firstly attained using first­principles simulation based on DFT calculation (Figure 10). Then, the effective masses of the charge carriers of the ZnO and the g­C3N4 can be respectively calcu­lated via the parabolic fitting of their CB minimum and VB maximum according to Equation (10) and (11):

= −(d /d )* 2 2 2 1m E k (10)

= / *v k m (11)

where ħ, d2E/dk2, k, m*, and v respectively represent the reduced Planck constant, the coefficient of the second­order term in a quadratic fit of E(k) curves for the band edge, the wave vector, the effective mass of a charge carrier, and the transfer rate of a charge carrier. Generally, the smaller the charge carrier’s effec­tive mass, the faster the charge­carrier transfer. In order to give a clearer image of the charge­carrier separation efficiency for the ZnO/g­C3N4 nanocomposite, the separation rate of the photo­generated holes to the electrons can be also calculated based on Equation (12):

= /* *D m mh e (12)

The calculated effective masses of electrons and holes and their separation rate in both the G­Z and the G­F directions in the reciprocal space are given in Table 2. It is obvious that the effective mass of the electron in the G­Z direction of ZnO (0.034) is much less than that of g­C3N4 (3.9), indicating that the photogenerated electrons of the ZnO in the G­Z direction show higher tendency to transfer than those of the g­C3N4 at the ZnO/g­C3N4 interface. These simulation results indicate that charge­carrier migration for ZnO and g­C3N4 is intrinsi­cally feasible and follows the Z­scheme mechanism instead of the type­II heterojunction mechanism.

3.6. Internal Electric-Field Simulation

Other than the effective masses of the photogenerated charge carriers, the surface energy and the interface formation energy, etc. are also found to have a great impact on the photocatalytic performance of a direct Z­scheme photocatalyst. For instance, Liu et al. found that the formation of an internal electric field between the g­C3N4 and the TiO2 is beneficial for facilitating the direct Z­scheme charge­carrier transfer efficiency.[67] In their study, the TiO2 (100) surface was chosen for constructing a TiO2/g­C3N4 heterostructure due to its high reactivity. The 4 × 2 unit of monolayer g­C3N4 and 4 × 1 unit of the TiO2 (100) surface were created for simulating the structure of the g­C3N4/TiO2 heterostructure (Figure 11a–c). The interface forma­tion energy of the TiO2/g­C3N4 was calculated to be −1.16 eV. The negative value of the interface formation energy reveals that the TiO2/g­C3N4 has a stable interface, indicating that

Small Methods 2017, 1, 1700080

Figure 9. a,b) High-resolution XPS spectra of the N 1s for g-C3N4 and ZnO/g-C3N4 (a) and of O 1s for ZnO and ZnO/g-C3N4 (b). c–e) Schematic illustration of the Fermi level of ZnO and g-C3N4 (c), formation of the weak internal electric field on ZnO/g-C3N4 (d), and charge-carrier separation on the ZnO/g-C3N4 (e). Reproduced with permission.[62] Copyright 2015, The Royal Society of Chemistry.

Page 9: A Review of Direct Z‐Scheme Photocatalysts

1700080 (9 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

charge­carrier migration can be easily achieved between TiO2 and g­C3N4. Finally, the three­dimensional charge­density dif­ference was calculated to give a clear picture of the charge­carrier migration for the TiO2/g­C3N4 (Figure 11d). It is obvious that the interface between the g­C3N4 and the TiO2 is the center of the charge­carrier redistribution. Meanwhile, a negligible change of the charge density can be observed in the interior of the TiO2 due to the weak interface formation energy between the g­C3N4 and the TiO2. The planar­averaged charge­density difference along the Z direction is also given (Figure 11e). The electrons on the g­C3N4 tend to migrate to TiO2 via the TiO2/g­C3N4 interface, leaving holes on the g­C3N4. The charge­carrier diffusion between the TiO2 and the g­C3N4 will continue until the equilibrium state of the system is obtained. As a result, the net charge­carrier accumu­lation causes the formation of an internal electric field at the

TiO2/g­C3N4 interface (Figure 11e). Upon light irradiation, the photogenerated electrons will migrate to the g­C3N4 under the influence of this internal electric field (see Figure 11f). The for­mation of the internal electric field is beneficial for accelerating the charge­carrier separation across the TiO2/g­C3N4 interface according to the direct Z­scheme mechanism.

4. Application of Direct Z-Scheme Photocatalysts

4.1. Photocatalytic Pollutant Degradation

Thousands of different organic pollutants enter the air, soil, and water everyday as a result of daily human activities.[68–70]

Small Methods 2017, 1, 1700080

Figure 10. a,b) The crystal structures of ZnO (a) and g-C3N4 (b), where the red, light gray, blue, and dark-gray balls represent oxygen, zinc, nitrogen, and carbon atoms, respectively. c,d) Electronic band structures of ZnO (c) and g-C3N4 (d). Reproduced with permission.[62] Copyright 2015, The Royal Society of Chemistry.

Table 2. The effective masses of the photogenerated electrons and holes of ZnO and g-C3N4 calculated through parabolic fitting of the CB minimum and VB maximum along a specific direction in the reciprocal space (see Figure 10c,d); mh

*, me*, and D represent the effective mass of

holes and electrons, and mh*/me

*, respectively.

Species Effective mass G-Z direction G-F direction

ZnO mh* 0.72 10.3

me* 0.034 0.9

D 20.9 11.5

g-C3N4 mh* 29 0.64

me* 3.9 0.41

D 7.4 1.6

Figure 11. a–c) Schematic illustration for the 4 × 2 unit of monolayer g-C3N4 (a), 4 × 1 unit of the TiO2 (100) surface (b), and structure of g-C3N4/TiO2 composite (c). d,e) The simulated three-dimensional elec-tron-density difference (d) and planar-averaged electron density differ-ence (e) for g-C3N4/TiO2, where the yellow and cyan areas respectively indicate electron accumulation and depletion. f) Schematic illustration of the charge-carrier migration mechanism on g-C3N4/TiO2 under the influence of the internal electric field. Reproduced with permission.[67] Copyright 2016, The Royal Society of Chemistry.

Page 10: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (10 of 21)

www.advancedsciencenews.com www.small-methods.com

These organic pollutants can greatly harm the environment and cause negative effects on human health. In order to eliminate these pollutants from the environment, various technologies have been proposed, including photocatalytic degradation,[71,72] biological degradation,[73] physical adsorption,[6,74,75] and fil­tration.[76,77] Among them, semiconductor­based photoca­talysis has captured substantial attention because of its ability to utilize sustainable solar energy for degradation of organic pollutants without causing any side effects to the environment. Although various semiconductors have been studied for photo­catalytic degradation of pollutants, their photocatalytic perfor­mance remains unsatisfactory.[35,78,79] Therefore, researchers have made enormous efforts toward developing novel photo­catalytic systems with high photocatalytic activities. Particularly, the development of direct Z­scheme photocatalysts has shown potential for use in enhancing the efficiency of photocatalytic pollutant degradation through the promotion of the separation of photogenerated electron–hole pairs and maximizing the redox potential of the photocatalytic system.

For example, our group reported a TiO2/g­C3N4 direct Z­scheme photocatalyst for photocatalytic HCHO decomposi­tion.[47] The TiO2/g­C3N4 direct Z­scheme photocatalyst was prepared by a simple calcination route using P25 and urea as the TiO2 and g­C3N4 sources, respectively. Upon calcination, urea was polycondensed into g­C3N4 and uniformly deposited onto the surface of the TiO2 (Figure 12a,b). The thermogravi­metric analysis was carried out to confirm the actual loading content of g­C3N4 in the TiO2/g­C3N4 composite. As shown in Figure 12c, the g­C3N4 can be fully decomposed at ca. 600 °C in

air. Meanwhile, the TiO2 exhibited negligible weight loss under similar conditions. Therefore, the actual g­C3N4 loading content can be simply estimated by calculating the weight differences of the samples before and after calcination at ca. 600 °C in air. According to this proposed method, the actual loading contents of g­C3N4 on U20, U100, U200, and U500 were respectively estimated to be 3, 12, 18, and 26 wt%. Then, the photocatalytic performance of the samples was investigated by the photo­catalytic decomposition of HCHO in air. It was found that the photocatalytic HCHO decomposition efficiency of the TiO2 increased as the content of g­C3N4 increased from 0 to 12 wt% (Figure 12d). According to the ·OH radical trapping test, this is attributed to the formation of a direct Z­scheme heterojunction between the g­C3N4 and the TiO2 (Figure 12e). However, further increase of the loading content of g­C3N4 caused a decrease in the photocatalytic performance of the samples. This is because overloading of the g­C3N4 caused a shielding effect on the TiO2 and inhibited the contact between the TiO2 and the reactant (Figure 12f). This work suggested that careful tuning of the content of an individual semiconductor in a direct Z­scheme photocatalyst is crucial to optimize its photocatalytic performance.

Furthermore, Zhu et al. reported the g­C3N4/Ag2WO4 direct Z­scheme photocatalyst for photocatalytic methyl orange deg­radation.[80] The g­C3N4/Ag2WO4 was prepared through the in situ precipitation method (Figure 13a). Specifically, g­C3N4 was firstly dispersed into the water to form negatively charged g­C3N4. Then, AgNO3 was added into the g­C3N4 suspen­sion, and the dissolved Ag+ ions could be easily anchored on

Small Methods 2017, 1, 1700080

Figure 12. a,b) Low-resolution (a) and high-resolution (b) TEM images of optimized g-C3N4/TiO2 (U100), in which the g-C3N4 loading content on the g-C3N4/TiO2 was simply tuned by changing the urea/P25 weight ratio from 0 (U0), 20 (U20), 100 (U100), and 200 (U200), to 500 (U500). c) Ther-mogravimetric analysis (TGA) spectra of the U0, U100 and g-C3N4. d) Comparison of photocatalytic formaldehyde degradation performance of U0, U20, U100, U200, U500, and g-C3N4. e,f) Schematic illustration for charge-carrier separation on the U100 (e) and U500 (f) samples. Reproduced with permission.[47] Copyright 2013, The Royal Society of Chemistry.

Page 11: A Review of Direct Z‐Scheme Photocatalysts

1700080 (11 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

g­C3N4 due to the electrostatic attraction between the Ag+ ions and the negatively charged g­C3N4. When Na2WO4 was intro­duced into the suspension, the Ag+ ions can react in situ with the WO4

2− ions to form Ag2WO4. Therefore, the Ag2WO4 NPs

can be uniformly anchored on the surface of the g­C3N4 (Figure 13b,c), providing a greater number of surface active sites for the photo­catalytic reaction. After loading of Ag2WO4, the PL fluorescence intensity of the g­C3N4 was greatly reduced (Figure 13d), indicating the high electron–hole separation rate for the g­C3N4/Ag2WO4. This result is attrib­uted to the formation of a direct Z­scheme heterojunction between the g­C3N4 and the Ag2WO4, which can spatially separate the photogenerated electrons and holes. As a result, the photocatalytic performance of the g­C3N4 for degradation of methyl orange was greatly enhanced by the loading of Ag2WO4 (Figure 13e).

Thereafter, Jo and Natarajan systemati­cally investigated the effect of the TiO2 mor­phology on the photocatalytic performance of the direct Z­scheme g­C3N4/TiO2 for the degradation of isoniazid.[81] The g­C3N4 was coupled with TiO2 NPs and TiO2 NTs via the wetness impregnation method to respectively form TiO2 NPs/g­C3N4 and TiO2 NTs/g­C3N4 composites (Figure 14). The photocatalytic performance of the TiO2 NPs/g­C3N4 and the TiO2 NTs/g­C3N4 is better than that of pure g­C3N4. According to the radical trap­ping test, this is due to the formation of a direct Z­scheme heterojunction between the g­C3N4 and the TiO2, which can reduce the electron–hole recombination rate and maxi­mize the redox potential of the photocatalyst. Moreover, the photocatalytic performance of the TiO2 NTs/g­C3N4 was higher than that of the TiO2 NPs/g­C3N4 because the TiO2 NTs/g­C3N4 has a larger specific surface area than TiO2 NPs/g­C3N4 to provide a greater

number of surface active sites for photocatalytic reaction. This work demonstrated that the rational tuning of the morphology of individual semiconductor in a direct Z­scheme photocatalyst is important for enhancing its photocatalytic performance.

Small Methods 2017, 1, 1700080

Figure 13. a) Schematic illustration for the growth mechanism of the g-C3N4/Ag2WO4. b,c) TEM images of the g-C3N4/Ag2WO4. d) PL spectra of the g-C3N4 and g-C3N4/Ag2WO4. e) Comparison of the photocatalytic performance of the g-C3N4, Ag2WO4 and g-C3N4/Ag2WO4. Reproduced with permission.[80] Copyright 2017, Elsevier B.V.

Figure 14. a,b) TEM images of TiO2 NPs/g-C3N4 (a) and TiO2 nanotubes (TiO2 NTs)/g-C3N4 (b) direct Z-scheme photocatalysts. Reproduced with permission.[81] Copyright 2015, Elsevier B.V.

Page 12: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (12 of 21)

www.advancedsciencenews.com www.small-methods.com

Additionally, exposed facets on the semiconductor can also have great influence on the photocatalytic pollutant degrada­tion performance of the direct Z­scheme photocatalyst. For instance, Huang et al. systematically investigated the effect of exposed {001} and {101} facets of TiO2 on the photocatalytic performance of the TiO2/g­C3N4 direct Z­scheme photocata­lyst.[51] They prepared TiO2 nanosheets (NSs) and TiO2 nano­boxes (NBs), which were coupled with g­C3N4 to respectively form TiO2 NS/g­C3N4 and TiO2 NB/g­C3N4 (Figure 15a,b). The TiO2 NS contacted with the g­C3N4 mainly through the {001} facets (Figure 15a,c), whereas the special structure of the TiO2 NBs caused TiO2 to mainly contact with g­C3N4 through the {101} facets (Figure 15b,d). Under light irradiation, photogen­erated electrons and holes on TiO2 accumulate at the {101} and {001} facets, respectively, due to the presence of the surface het­erojunction between the {101} and {001} facets of the TiO2.[82]

Then, the photogenerated electrons that are accumulated at the {101} facets of the TiO2 NBs can easily migrate to the g­C3N4 for a reduction reaction according to the direct Z­scheme migra­tion mechanism. Meanwhile, the photo generated holes remain on the {001} facets of the TiO2 NBs for an oxidation reaction. As a result, the electron–hole separation across the TiO2 NB/g­C3N4 is much faster than that of the TiO2 NS/g­C3N4, and TiO2 NB/g­C3N4 showed the highest photocatalytic performance among all the prepared samples for degradation of Brilliant Red X3B (Figure 15e). This work showed that appropriate inter­face engineering of the exposed facet of an individual semicon­ductor of a direct Z­scheme photocatalyst can greatly improve the electron–hole separation efficiency, thereby enhancing the photocatalytic performance.

Other than TiO2 and g­C3N4, other semiconductor­based direct Z­scheme photocatalysts have also been studied for photo catalytic pollutant degradation. For instance, Jo et al. reported a ZnIn2S4 marigold flower/Bi2WO6 direct Z­scheme photocatalyst for photocatalytic degradation of metronidazole (MTZ).[54] The ZnIn2S4 marigold flower was prepared by a simple hydrothermal method. During the hydrothermal reac­tion, ZnIn2S4 NPs were firstly formed (Figure 16a), which fur­ther grew into a nanopetal structure. Thereafter, the ZnIn2S4

Small Methods 2017, 1, 1700080

Figure 15. a,b) TEM images of TiO2 nanosheets (TiO2 NS)/g-C3N4 (a) and TiO2 hollow nanoboxes (TiO2 NB)/g-C3N4 (b). c,d) Schematic illustration of the contact interface between TiO2 and g-C3N4 on TiO2 NS/g-C3N4 (c) and TiO2 NB/g-C3N4 (d). e) Comparison of the photo-catalytic performance of the prepared samples for degradation of Brilliant Red X3B. Reproduced with permission.[51] Copyright 2015, Elsevier B.V.

Figure 16. a) Schematic illustration for the growth mechanism of ZnIn2S4 marigold flowers. b,c) Field-emission scanning electron microscopy (FESEM) images of the ZnIn2S4 marigold flower (b) and ZnIn2S4 marigold flower/Bi2WO6 direct Z-scheme photocatalyst (c). d) Light-absorption spectra of the prepared samples. Reproduced with permission.[54] Copy-right 2016, The Royal Society of Chemistry.

Page 13: A Review of Direct Z‐Scheme Photocatalysts

1700080 (13 of 21)

www.advancedsciencenews.com www.small-methods.com

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

nanopetals curled up under high temperature and pressure. Finally, these curled ZnIn2S4 nanopetals self­assembled and formed a hierarchical flower­like structure (Figure 16b). After obtaining the ZnIn2S4 marigold flowers, the ZnIn2S4 marigold flower/Bi2WO6 direct Z­scheme photocatalyst was obtained by wet­impregnation of the ZnIn2S4 marigold flower with a pre­cursor of Bi2WO6 (Figure 16c). After coupling with ZnIn2S4, it was found that the light­absorption spectrum of the Bi2WO6 was redshifted, indicating the intimate interfacial interaction between the Bi2WO6 and the ZnIn2S4 (Figure 16d). The photocatalytic MTZ degradation performance of the optimized ZnIn2S4 mari­gold flower/Bi2WO6 direct Z­scheme photocatalyst was higher than that of pure ZnIn2S4 and Bi2WO6. This enhanced photo­

catalytic activity is due to the enhanced charge­carrier separation efficiency and optimized redox potential of the direct Z­scheme ZnIn2S4 marigold flower/Bi2WO6 photocatalyst.

Recently, graphene­based materials have attracted wide attention from the scientific community due to their ultralarge specific surface area and high chemical and phys­ical stability.[83,84] Therefore, graphene­based materials such as graphene oxide (GO) have been employed, coupled with other semiconductors, forming direct Z­scheme photocatalysts. For example, our group prepared Ag2CrO4/GO direct Z­scheme photocatalysts through a simple self­assembly precipitation method for degradation of methylene blue (Figure 17a).[50] In detail, negatively charged GO was dispersed into a AgNO3

Small Methods 2017, 1, 1700080

Figure 17. a) Schematic illustration for the formation mechanism of the Ag2CrO4/graphene oxide (GO) direct Z-scheme photocatalyst. b,c) SEM (b) and TEM images of the optimized Ag2CrO4/GO direct Z-scheme photocatalysts (G1), in which Ag2CrO4 loaded with different GO contents: 0 (G0), 0.5 (G0.5), 0.75 (G0.75), 1 (G1), 2 (G2) and 3 wt% (G3) was prepared. d) Electrochemical impedance spectra of the prepared samples. e) Comparison of the photocatalytic performance of the prepared samples and nitrogen-doped TiO2 (N-TiO2) for degradation of methylene blue. Reproduced with permission.[50] Copyright 2015, Elsevier B.V.

Page 14: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (14 of 21)

www.advancedsciencenews.com www.small-methods.com

solution, and the Ag+ ions were adsorbed on the negatively charged GO due to electrostatic attraction. Then, a K2CrO4 solu­tion was added into the above suspension, and K2CrO4 in situ reacted with Ag+ ions on the negatively charged GO to create the Ag2CrO4/GO direct Z­scheme photocatalysts. As shown in the scanning electron microscopy (SEM) and TEM images (Figure 17b,c), Ag2CrO4 particles are successfully loaded on the GO surface, indicating the strong interaction between the Ag2CrO4 and the GO. Moreover, after coupling with the GO, the electron–hole separation efficiency of the Ag2CrO4 is greatly enhanced due to the formation of a direct Z­scheme heterojunction between the Ag2CrO4 and the GO (Figure 17d). As a result, the photocatalytic performance of the optimized Ag2CrO4/GO was 3.5 times higher than that of pure Ag2CrO4 for degradation of methylene blue due to the improved elec­tron–hole separation efficiency and the optimized redox potential of the Ag2CrO4/GO composite (Figure 17e). This work demonstrated that graphene­based material can be a potential candidate for constructing advanced direct Z­scheme photocatalysts.

4.2. Photocatalytic Hydrogen Production

Recently, hydrogen, a clean and high­heat­value energy carrier, has received great attention due to depletion of fossil­fuel energy sources and their adverse effects on the envi­ronment.[85–87] Photocatalytic water splitting for H2 produc­tion is an ideal strategy for the generation of H2 because of its simplicity and clean reaction,[88–90] in which only the photocata­lyst, sunlight, and water are required to produce H2 at room temperature.[29,91] Therefore, many semiconductor­based photo­catalysts have been studied and applied in photocatalytic H2 production.[92–95] Although this solar­energy conversion strategy exhibits many advantages, H2 production through photocata­lytic water splitting is still far from practical application. This is mainly due to the high electron–hole recombination rate on the photocatalysts.[96–99] Building a direct Z­scheme photocata­lyst can effectively suppress the electron–hole recombination and isolate the reduction and oxidation sites of the photocata­lytic reaction. Therefore, various direct Z­scheme photocatalysts have been studied for enhancing the efficiency of the photocata­lytic production of H2.

For example, Mukhopadhyay et al. reported hierarchical ZnO/CdS nanocomposites with enhanced photocatalytic water­splitting performance.[59] ZnO nanorods and mesoporous ZnO nanoflowers were firstly prepared through sequential sono­chemical and hydrothermal methods. Then, the prepared ZnO nanorods and mesoporous ZnO nanoflowers were coupled with CdS NPs to form ZC­1 and ZC­2 direct Z­scheme photo­catalysts, respectively (Figure 18a–d). It was found that the light­absorption range of both ZnO nanorods and mesoporous ZnO nanoflowers was extended into the visible­light range by the loading of CdS due to the relatively small bandgap value of CdS (Figure 18e). The extended light­absorption range of the samples is beneficial for harvesting both UV and visible wavelengths of the incident solar radiation. The PL intensi­ties of the ZC­1 and ZC­2 were lower than that of pure ZnO,

indicating that the electron–hole separation efficiencies of ZC­1 and ZC­2 are higher than that of pure ZnO. This is because the electron–hole separation efficiency of ZnO can be greatly enhanced by the CdS NPs through the fluorescence resonance energy transfer (FRET) mechanism (Figure 18f,g). In detail, the photogenerated electrons in the CB of the ZnO can transfer to the VB of the CdS, thereby reducing the fluorescence of the ZnO. More interestingly, the electron–hole recombination rate of ZC­1 is lower than that of ZC­2 because the absorption–emission interaction between the ZnO and the CdS of ZC­1 is higher than that of ZC­2 (Figure 18h). This is mainly attrib­uted to the large contact interface between the ZnO nanorods and CdS NPs, which can greatly accelerate the migration of the charge carriers across the interface between the ZnO and the CdS. As a result, the prepared ZC­1 exhibited the highest photo catalytic H2 production performance among all the pre­pared samples.

Recently, Chen et al. demonstrated that the combination of doping and the construction of the direct Z­scheme heterojunc­tion is a feasible way to enhance the photocatalytic performance of TiO2.[27] Specifically, a visible­light responsive TiO2/g­C3N4 direct Z­scheme photocatalyst was prepared through a simple one­pot solvothermal route with the assistance of concentrated nitric acid, for photocatalytic H2 production. Upon C and N doping, the C and N co­doped­TiO2/g­C3N4 (C–N TiO2/g­C3N4) exhibited a redshifted light­absorption band edge in comparison with the pure TiO2 NPs (Figure 19a), indicating the incorpora­tion of C and N atoms into TiO2. The extended light­absorption range of the samples is beneficial for enhancing their vis­ible­light reactivity. Meanwhile, the formation of the direct Z­scheme heterojunction between C–N TiO2 and g­C3N4 mark­edly improved their charge transfer and separation efficiency for photocatalytic reaction (Figure 19b,c). As a result, the photocata­lytic H2 production performance of the optimized C–N TiO2/ g­C3N4 (3 wt% C–N TiO2/g­C3N4) composite was 10.9 and 21.3 times higher than those of CN­TiO2 and pure g­C3N4, respec­tively (Figure 19d). Doping is deemed to be a simple method for improving the light absorption of the individual semicon­ductors in a direct Z­scheme photocatalyst to achieve its visible­light activity.

Thereafter, our group found that direct Z­scheme photocata­lysts can also be constructed from a single semiconductor with the same chemical composition but different phases.[60] In detail, anatase/rutile bi­phase nanofibers with different rutile contents were prepared by slow and rapid cooling of calcined electrospun TiO2 nanofibers (Figure 20a–d). Both anatase­ and rutile­phase TiO2 can be observed in the TEM images of the AR28 and AR45 samples (Figure 20b,d), indicating that the mixed­phase TiO2 nanofiber was successfully prepared with large a specific sur­face area (Table 3). Then, X­ray diffraction (XRD) characteriza­tion also further confirmed the presence of anatase and rutile phases in AR28 and AR45 (Figure 20e). The photocatalytic H2 production test was then performed to investigate the photocat­alytic performance of the prepared samples. It was found that the photocatalytic H2 production activity of the AR28 and AR45 was higher than that of the AR0 and AR100 (see Figure 20f). According to the Pt NPs photo­deposition test, this is due to the formation of a direct Z­scheme heterojunction between the

Small Methods 2017, 1, 1700080

Page 15: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (15 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

Figure 18. a,b) TEM images of ZnO nanorod/CdS NP (ZC-1) (a) and ZnO nanoflower/CdS NPs composites (ZC-2) (b). c,d) High-resolution TEM images of ZC-1 (c) and ZC-2 (d). e) Light-absorption spectra of ZnO rods, ZnO NPs, ZC-1 and ZC-2. f,g) PL emission spectra of ZC-1 (f) and ZC-2 (g) excited at 345 nm, in comparison with the emission spectra of ZnO and CdS. h) Emission spectra of ZC-1 and ZC-2 excited at 390 nm in comparison with the emission spectra of the CdS NPs. Reproduced with permission.[59] Copyright 2015, The Royal Society of Chemistry.

Page 16: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (16 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

anatase­ and rutile­phase TiO2 of the AR28 and AR45 samples (Figure 20g). Specifically, the Pt NPs were selectively deposited on the rutile­phase TiO2 during the photo­reduction process (see Figure 8), suggesting the accumulation of photogenerated electrons on the rutile­phase TiO2. This result confirms that the charge­carrier migration mechanism on the anatase/rutile bi­phase nanofibers obeys the Z­scheme mechanism instead of the type­II­heterojunction migration mechanism. Moreover, it was found that the photocatalytic performance of the AR45 was higher than that of the AR28. This is because the balance between the rutile and anatase phases is beneficial for the spa­tial separation of photogenerated electron and hole pairs.

4.3. Photocatalytic CO2 Reduction

The ever­rising demand for fossil fuels and a collateral increase in the concentration of atmospheric CO2 have urged the devel­opment of carbon­management technologies.[100–103] To this end, much research has been devoted to searching for new technologies for the reduction of CO2, for example biological conversion,[104] electrocatalysis,[105] photocatalysis, and photo­thermal catalysis.[106] Among them, photocatalytic CO2 reduc­tion into valuable chemical fuels such as CH4, CH3OH, HCOOH, and CH2O has received great attention because it can alleviate the current dependence on fossil fuels of human society, as well as reducing the CO2 concentration in the

atmosphere.[107–110] Notably, hydrocarbon fuels such as CH4 or CH3OH have a higher energy density than H2, making them easier for energy storage and more suitable for practical applications. Nevertheless, compared with other technologies, photo catalytic CO2 reduction for solar fuel production is still an enormous challenge because of its low solar­energy­con­version efficiency, which is primarily attributed to the rapid electron–hole recombination rate on the semiconductor.[111–113] Thus, it is necessary to develop advanced photocatalysts with good electron–hole separation efficiency.[114–116] Recent studies have indicated that building direct Z­scheme photocatalysts is one of the best strategies to prepare highly efficient photo­catalysts for CO2 reduction.

For example, Yu et al. prepared the ZnO/g­C3N4 direct Z­scheme photocatalyst for CO2 reduction. The efficiency of the photocatalytic CO2 reduction of the prepared ZnO/g­C3N4 was higher than that of commercial ZnO, lab­synthesized ZnO, and g­C3N4 for CH3OH production (Figure 21a).[62] According to the ·OH radical trapping test, this improved photocatalytic performance of the ZnO/g­C3N4 is due to the formation of a direct Z­scheme heterojunction between the g­C3N4 and the ZnO, which can promote electron–hole separation and opti­mize the redox potential of the photocatalytic system. Moreover, theoretical simulations were also carried out to further confirm the formation of the direct Z­scheme heterojunction between the g­C3N4 and the ZnO. Specifically, the effective masses of the photogenerated electrons and holes of the g­C3N4 and the ZnO

Figure 19. a) Light-absorption spectra of g-C3N4, 1 wt% C and N co-doped-TiO2/g-C3N4 (1 C–N TiO2/g-C3N4), 2 wt% C and N co-doped-TiO2/g-C3N4 (2 C–N TiO2/g-C3N4), 3 wt% C and N co-doped-TiO2/g-C3N4 (3 C–N TiO2/g-C3N4), 5 wt% C and N co-doped-TiO2/g-C3N4 (5 C–N TiO2/g-C3N4) and TiO2 NPs. b,c) Photocurrent response (b) and electrochemical impedance (c) spectra of the prepared samples. d) Comparison of the photocatalytic H2 production performance of the prepared samples. Reproduced with permission.[27] Copyright 2015, The Royal Society of Chemistry.

Page 17: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (17 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

were estimated through the parabolic fitting to the CB min­imum and VB maximum of the simulated band structure of the g­C3N4 and the ZnO. It was found that the calculated effective

masses of the photogenerated electrons in the G­Z direction of the ZnO (0.034) are much lower than those of the g­C3N4 (3.9), indicating that the photogenerated electrons in the G­Z direc­tion of the ZnO have a higher tendency to transfer than those of the g­C3N4. Therefore, the migration of electrons from the ZnO to the g­C3N4 is more favorable than that from the g­C3N4 to the ZnO. Both an ·OH radical trapping test and theoretical simulation results indicated that the enhanced photocatalytic performance of the ZnO/g­C3N4 is due to the formation of a direct Z­scheme heterojunction (Figure 21b,c).

Then, our group reported well­dispersed CdS NPs grown on hierarchical WO3 hollow spheres for photocatalytic CO2 reduction.[61] It was found that the CdS NPs were success­fully deposited on WO3 nanosheets (Figure 22a,b). The close contact between the CdS and the WO3 can accelerate the

Figure 20. a–d) SEM and TEM images of AR28 (a,b) and AR45 (c,d). e) XRD patterns of AR0 (I), AR28 (II), AR45 (III), and AR100 (IV). f) Comparison of the photocatalytic H2 production performance of the prepared samples. g) Schematic illustration of the charge-carrier migration for an anatase/rutile direct Z-scheme photocatalyst. Reproduced with permission.[60] Copyright 2014, Elsevier B.V.

Table 3. Physical properties of anatase/rutile bi-phase nanofibers with 0% (AR0), 28% (AR28), 45% (AR45), and 100% rutile (AR100) content.

Sample SBET [m2 g−1]

PVa) [m3 g−1]

APSa) [nm]

AR0 31 0.02 3.1

AR28 69 0.13 7.2

AR45 57 0.11 7.6

AR100 5 0.01 5.6

a)PV and APS respectively denote pore volume and average pore size.

Page 18: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (18 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

migration of charge carriers between the CdS and the WO3 during the photo catalytic reaction (Figure 22b). Moreover, the CO2­adsorption ability of the CdS/WO3 was higher than that of WO3 because the introduction of the CdS NPs can increase the specific surface area of the WO3. The improved CO2­adsorption ability of the sample is beneficial for enhancing the CO2 concen­tration on its surface, thereby improving the CO2­conversion efficiency (Figure 5b). Furthermore, the successful formation of the direct Z­scheme photocatalyst was evidenced from the ·OH radical trapping test using a terephthalic acid probe mole­cule. In detail, the PL signals of the HTA of the WO3 and CdS/WO3 gradually increase with irradiation time (see Figure 22c), indicating that WO3 and CdS/WO3 have sufficient oxidation potential to produce ·OH radicals (Figure 22d). Meanwhile, no PL signal can be observed for the CdS, suggesting that CdS does not have sufficient oxidation potential to produce ·OH radicals. These results demonstrate that the photogenerated holes accumulate on the WO3 instead of the CdS, indicating the formation of a direct Z­scheme heterojunction (Figure 5c). Moreover, it should be noted that PL signal of the CdS/WO3 is higher than that of WO3, indicating that a higher concen­tration of ·OH radicals was produced by the CdS/WO3. This is due to the faster electron–hole separation for the CdS/WO3. As a result, the photocatalytic performance of the CdS/WO3 was higher than that of pure CdS and WO3 for CO2 production (Figure 5b). This work demonstrates that the coupling of two different semiconductors can not only form a direct Z­scheme

photocatalyst for enhancing their electron–hole separation efficiency but also improves their physicochemical properties for dif­ferent functionalities.

5. Conclusion and Future Perspectives

Here, the historical development of the direct Z­scheme photo catalyst from its 1st genera­tion to the current 3rd generation has been outlined. Then, the recent advances of the direct Z­scheme photocatalyst including characterization methods and applications have been discussed. It is conspicuous that the construction of direct Z­scheme photocatalysts shows great potential in var­ious photocatalytic applications ranging from environmental remediation to solar fuel pro­duction. However, the development of direct Z­scheme photocatalysts is still at an early stage. Enormous challenges exist for fully exploring the potential of direct Z­scheme photocatalysts. Therefore, future develop­ment in this field is urgently needed and should be focused in following directions.

Firstly, the physical and chemical prop­erties of the individual semiconductor in a direct Z­scheme photocatalyst should be optimized. Much work is needed to optimize the physicochemical properties of each semi­

conductor in a direct Z­scheme photocatalyst to enhance the photoconversion efficiency. These research efforts may include the preparation of advanced semiconductors with high visible­light utilization for high solar­energy­conversion efficiency, low charge­carrier­transfer resistance for fast charge­carrier migra­tion and good physical and chemical stability for long­term application.

Secondly, the contact interface between two semiconduc­tors in the direct Z­scheme photocatalyst must be rationally controlled. The charge­carrier separation across the interface between two semiconductors is crucial for optimizing the photo catalytic performance of a direct Z­scheme photocatalyst. Therefore, maximizing and optimizing the contact interface between two semiconductors in a direct Z­scheme photocata­lyst can be an effective method to optimize their photocatalytic performance.

Thirdly, the photocatalytic performance and selectivity of a direct Z­scheme photocatalyst can be further improved by building ternary or multicomponent photocatalytic systems. One viable way is to deposit suitable oxidation and reduction co­catalysts respectively on oxidation and reduction sites of the direct Z­scheme photocatalyst in order to further enhance its electron–hole separation efficiency and photocatalytic performance.

Finally, the mechanisms of direct Z­scheme photocata­lysts remain largely unclear and must be extensively studied. Theoretical material simulation based on first­principles

Figure 21. a) Comparison of the photocatalytic CO2 reduction performance of commercial ZnO, ZnO, ZnO/g-C3N3, and g-C3N4 for CH3OH production. b,c) Schematic illustrations of the electron–hole pair separation mechanism on the type-II heterojunction (b) and the direct Z-scheme photocatalyst (c). Reproduced with permission.[62] Copyright 2015, The Royal Society of Chemistry.

Page 19: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (19 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

DFT combined with various characterization methods such as photo catalytic testing, radical­species trapping tests, XPS characterization, etc. can be an effective strategy for pro­viding a full image of the photocatalytic reaction for a direct Z­scheme photo catalyst. The in­depth understanding of the working mechanism of direct Z­scheme photocatalysts can provide researchers with various opportunities for opti­mizing the charge­carrier separation efficiency and tuning the selectivity of the photocatalytic reaction toward different products.

We believe that the further expansion of direct Z­scheme photocatalysts from these directions will greatly stimulate the application of photocatalysis in energy conversion and environ­mental remediation, thereby ensuring the sustainable develop­ment of human society.

AcknowledgementsThis review was partially supported by the 973 program (2013CB632402), NSFC (21433007, 51320105001, 21573170 and 51372190), the Fundamental Research Funds for the Central Universities (2015-III-034), Self-determined and Innovative Research Funds of SKLWUT (2015-ZD-1 and 2016-KF-17), and the Natural Science Foundation of Hubei Province of China (No. 2015CFA001).

KeywordsCO2 reduction, direct Z-scheme, hydrogen production, photocatalysts, pollutant degradation

Received: January 16, 2017Revised: February 15, 2017

Published online: April 18, 2017

[1] J. G. Yu, J. X. Low, W. Xiao, P. Zhou, M. Jaroniec, J. Am. Chem. Soc. 2014, 136, 8839.

[2] R. E. Smalley, MRS Bull. 2005, 30, 412.[3] Q. J. Xiang, B. Cheng, J. G. Yu, Angew. Chem., Int. Ed. 2015, 54,

11350.[4] J. Yang, D. Wang, H. Han, C. Li, Acc. Chem. Res. 2013, 46, 1900.[5] J. R. Ran, J. Zhang, J. G. Yu, M. Jaroniec, S. Z. Qiao, Chem. Soc.

Rev. 2014, 43, 7787.[6] X. Wang, W. Bi, P. Zhai, X. Wang, H. Li, G. Mailhot, W. Dong, Appl.

Surf. Sci. 2016, 360, 240.[7] G. Zhao, G. Liu, H. Pang, H. Liu, H. Zhang, K. Chang, X. Meng,

X. Wang, J. Ye, Small 2016, 12, 6160.[8] Y. Sohn, W. Huang, F. Taghipour, Appl. Surf. Sci. 2016, 396, 1696.[9] R. A. He, S. W. Cao, J. G. Yu, Acta Phys. Chim. Sin. 2016, 32, 2841.

[10] H. L. Tan, H. A. Tahini, X. Wen, R. J. Wong, X. Tan, A. Iwase, A. Kudo, R. Amal, S. C. Smith, Y. H. Ng, Small 2016, 12, 5295.

[11] X. Zhang, Y. Wang, B. Liu, Y. Sang, H. Liu, Appl. Catal., B 2017, 202, 620.

[12] M. B. Qadir, Y. Li, I. A. Sahito, A. A. Arbab, K. C. Sun, N. Mengal, A. A. Memon, S. H. Jeong, Small 2016, 12, 4508.

[13] O. Pachoumi, I. Nasrallah, H. Sirringhaus, Small Methods 2016, 1, 1600007.

[14] T. Hisatomi, J. Kubota, K. Domen, Chem. Soc. Rev. 2014, 43, 7520.[15] M. S. Faber, S. Jin, Energy Environ. Sci. 2014, 7, 3519.[16] W. D. Chemelewski, H.-C. Lee, J.-F. Lin, A. J. Bard, C. B. Mullins,

J. Am. Chem. Soc. 2014, 136, 2843.[17] J. Li, N. Wu, Catal. Sci. Technol. 2015, 5, 1360.

Figure 22. a) FESEM and high-resolution FESEM images (inset) of the WO3 hollow spheres. b) High-resolution TEM image of the CdS/WO3 direct Z-scheme photocatalyst. c) Comparison of the HTA PL intensity of the prepared samples at 456 nm against irradiation time in terephthalic acid solu-tion. d) Comparison of the redox potential of CdS and WO3 for producing ·OH and ·O2− radicals. Reproduced with permission.[61] Copyright 2015, Wiley-VCH.

Page 20: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (20 of 21)

www.advancedsciencenews.com www.small-methods.com

Small Methods 2017, 1, 1700080

[18] X. Li, J. G. Yu, M. Jaroniec, Chem. Soc. Rev. 2016, 45, 2603.[19] Q. Huang, J. G. Yu, S. W. Cao, C. Cui, B. Cheng, Appl. Surf. Sci.

2015, 358, 350.[20] J. Schneider, M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi,

M. Anpo, D. W. Bahnemann, Chem. Rev. 2014, 114, 9919.[21] Z. Ni, Y. Sun, Y. Zhang, F. Dong, Appl. Surf. Sci. 2016, 365, 314.[22] R. A. He, S. W. Cao, P. Zhou, J. G. Yu, Chin. J. Catal. 2014, 35, 989.[23] M. Marszewski, S. W. Cao, J. G. Yu, M. Jaroniec, Mater. Horiz.

2015, 2, 261.[24] J. X. Low, B. Cheng, J. G. Yu, Appl. Surf. Sci. 2017, 392, 658.[25] H. Wang, L. Zhang, Z. Chen, J. Hu, S. Li, Z. Wang, J. Liu, X. Wang,

Chem. Soc. Rev. 2014, 43, 5234.[26] S. W. Cao, J. X. Low, J. G. Yu, M. Jaroniec, Adv. Mater. 2015, 27,

2150.[27] W. Chen, T.-Y. Liu, T. Huang, X.-H. Liu, G.-R. Duan, X.-J. Yang,

S.-M. Chen, RSC Adv. 2015, 5, 101214.[28] S. J. Moniz, S. A. Shevlin, D. J. Martin, Z.-X. Guo, J. Tang, Energy

Environ. Sci. 2015, 8, 731.[29] K. Li, F.-Y. Su, W.-D. Zhang, Appl. Surf. Sci. 2016, 375, 110.[30] X. Li, J. G. Yu, J. X. Low, Y. P. Fang, J. Xiao, X. B. Chen, J. Mater.

Chem. A 2015, 3, 2485.[31] U. I. Gaya, A. H. Abdullah, J. Photochem. Photobiol., C 2008, 9, 1.[32] H. Li, Y. Zhou, W. Tu, J. Ye, Z. Zou, Adv. Funct. Mater. 2015, 25,

998.[33] J. L. White, M. F. Baruch, J. E. Pander III, Y. Hu, I. C. Fortmeyer,

J. E. Park, T. Zhang, K. Liao, J. Gu, Y. Yan, Chem. Rev. 2015, 115, 12888.

[34] M. R. Gholipour, C.-T. Dinh, F. Béland, T.-O. Do, Nanoscale 2015, 7, 8187.

[35] J. Xiao, Y. Xie, H. Cao, Chemosphere 2015, 121, 1.[36] S. Sun, Nanoscale 2015, 7, 10850.[37] J. X. Low, J. G. Yu, M. Jaroniec, S. Wageh, A. A. Al-Ghamdi, Adv.

Mater. 2017, DOI: 10.1002/adma.201601694.[38] P. Zhou, J. G. Yu, M. Jaroniec, Adv. Mater. 2014, 26, 4920.[39] Y. W. Su, W. H. Lin, Y. J. Hsu, K. H. Wei, Small 2014, 10, 4427.[40] H. Li, W. Tu, Y. Zhou, Z. Zou, Adv. Sci. 2016, 3, 1500389.[41] Y. Cao, Z. Wu, J. Ni, W. A. Bhutto, J. Li, S. Li, K. Huang, J. Kang,

Nano-Micro Lett. 2012, 4, 135.[42] A. J. Bard, J. Photochem. 1979, 10, 59.[43] H. Tada, T. Mitsui, T. Kiyonaga, T. Akita, K. Tanaka, Nat. Mater.

2006, 5, 782.[44] W. Tu, Y. Zhou, S. Feng, Q. Xu, P. Li, X. Wang, M. Xiao, Z. Zou,

Chem. Commun. 2015, 51, 13354.[45] C. Liu, J. Tang, H. M. Chen, B. Liu, P. Yang, Nano Lett. 2013, 13,

2989.[46] X. Wang, G. Liu, Z.-G. Chen, F. Li, L. Wang, G. Q. Lu, H.-M. Cheng,

Chem. Commun. 2009, 3452.[47] J. G. Yu, S. H. Wang, J. X. Low, W. Xiao, Phys. Chem. Chem. Phys.

2013, 15, 16883.[48] D. Zhou, Z. Chen, Q. Yang, C. Shen, G. Tang, S. Zhao, J. Zhang,

D. Chen, Q. Wei, X. Dong, ChemCatChem 2016, 8, 3064.[49] J. W. Fu, S. W. Cao, J. G. Yu, J. Materiomics 2015, 1, 124.[50] D. F. Xu, B. Cheng, S. W. Cao, J. G. Yu, Appl. Catal., B 2015, 164,

380.[51] Z. Huang, Q. Sun, K. Lv, Z. Zhang, M. Li, B. Li, Appl. Catal.,

B 2015, 164, 420.[52] W.-K. Jo, T. Sivakumar Natarajan, ACS Appl. Mater. Interfaces 2015,

7, 17138.[53] J. Zhang, Y. Hu, X. Jiang, S. Chen, S. Meng, X. Fu, J. Hazard. Mater.

2014, 280, 713.[54] W.-K. Jo, J. Y. Lee, T. S. Natarajan, Phys. Chem. Chem. Phys. 2016,

18, 1000.[55] S. Chen, Y. Hu, L. Ji, X. Jiang, X. Fu, Appl. Surf. Sci. 2014, 292,

357.

[56] S. Meng, X. Ning, T. Zhang, S.-F. Chen, X. Fu, Phys. Chem. Chem. Phys. 2015, 17, 11577.

[57] S. Chen, L. Ji, W. Tang, X. Fu, Dalton Trans. 2013, 42, 10759.[58] L. Shi, L. Liang, F. Wang, M. Liu, J. Sun, J. Mater. Sci. 2015, 50,

1718.[59] S. Mukhopadhyay, I. Mondal, U. Pal, P. S. Devi, Phys. Chem. Chem.

Phys. 2015, 17, 20407.[60] F. Y. Xu, W. Xiao, B. Cheng, J. G. Yu, Int. J. Hydrogen Energy 2014,

39, 15394.[61] J. Jin, J. G. Yu, D. P. Guo, C. Cui, W. K. Ho, Small 2015, 11,

5262.[62] W. L. Yu, D. F. Xu, T. Y. Peng, J. Mater. Chem. A 2015, 3, 19936.[63] J.-C. Wang, L. Zhang, W.-X. Fang, J. Ren, Y.-Y. Li, H.-C. Yao,

J.-S. Wang, Z.-J. Li, ACS Appl. Mater. Interfaces 2015, 7, 8631.[64] S. Shinde, C. Bhosale, K. Rajpure, Catal. Rev. 2013, 55, 79.[65] S. Gligorovski, R. Strekowski, S. Barbati, D. Vione, Chem. Rev.

2015, 115, 13051.[66] Q. J. Xiang, J. G. Yu, P. K. Wong, J. Colloid Interface Sci. 2011, 357,

163.[67] J. J. Liu, B. Cheng, J. G. Yu, Phys. Chem. Chem. Phys. 2016, 18,

31175.[68] R. Singh, S. Singh, P. Parihar, V. P. Singh, S. M. Prasad, Ecotoxicol.

Environ. Saf. 2015, 112, 247.[69] M. Mehrjouei, S. Müller, D. Möller, Chem. Eng. J. 2015, 263, 209.[70] V. Etacheri, C. Di Valentin, J. Schneider, D. Bahnemann, S. C. Pillai,

J. Photochem. Photobiol., C 2015, 25, 1.[71] Q. Zhai, L. Zhang, X. Zhao, H. Chen, D. Yin, J. Li, Appl. Surf. Sci.

2016, 377, 17.[72] D. Li, W. Shi, Chin. J. Catal. 2016, 37, 792.[73] A. Joss, S. Zabczynski, A. Göbel, B. Hoffmann, D. Löffler,

C. S. McArdell, T. A. Ternes, A. Thomsen, H. Siegrist, Water Res. 2006, 40, 1686.

[74] Y. Zhang, X. Lin, Q. Zhou, X. Luo, Appl. Surf. Sci. 2016, 389, 34.[75] X. Yao, L. Li, W. Zou, S. Yu, J. An, H. Li, F. Yang, L. Dong, Chin.

J. Catal. 2016, 37, 1369.[76] C. Li, M. Wei, D. G. Evans, X. Duan, Small 2014, 10, 4469.[77] K. Zhou, X.-Y. Hu, B.-Y. Chen, C.-C. Hsueh, Q. Zhang, J. Wang,

Y.-J. Lin, C.-T. Chang, Appl. Surf. Sci. 2016, 383, 300.[78] M. Gao, L. Zhu, W. L. Ong, J. Wang, G. W. Ho, Catal. Sci. Technol.

2015, 5, 4703.[79] M. Ge, C. Cao, J. Huang, S. Li, Z. Chen, K.-Q. Zhang, S. Al-Deyab,

Y. Lai, J. Mater. Chem. A 2016, 4, 6772.[80] B. C. Zhu, P. F. Xia, Y. Li, W. K. Ho, J. G. Yu, Appl. Surf. Sci. 2017,

391, 175.[81] W.-K. Jo, T. S. Natarajan, Chem. Eng. J. 2015, 281, 549.[82] C. P. Sajan, S. Wageh, A. A. Al-Ghamdi, J. G. Yu, S. W. Cao, Nano

Res. 2016, 9, 3.[83] X. Li, J. G. Yu, S. Wageh, A. A. Al-Ghamdi, J. Xie, Small 2016, 12,

6640.[84] Q. J. Xiang, J. G. Yu, M. Jaroniec, Chem. Soc. Rev. 2012, 41, 782.[85] T. Wang, Z. Luo, C. Li, J. Gong, Chem. Soc. Rev. 2014, 43,

7469.[86] T. Dushatinski, C. Huff, T. M. Abdel-Fattah, Appl. Surf. Sci. 2016,

385, 282.[87] T. Wu, X. Kang, M. W. Kadi, I. Ismail, G. Liu, H.-M. Cheng, Chin.

J. Catal. 2015, 36, 2103.[88] S. W. Cao, J. G. Yu, J. Phys. Chem. Lett. 2014, 5, 2101.[89] G. Zhang, M. Zhang, X. Ye, X. Qiu, S. Lin, X. Wang, Adv. Mater.

2014, 26, 805.[90] X. Zou, Y. Zhang, Chem. Soc. Rev. 2015, 44, 5148.[91] J. Jiang, H. Zhou, F. Zhang, T. Fan, D. Zhang, Appl. Surf. Sci. 2016,

368, 309.[92] D. J. Martin, K. Qiu, S. A. Shevlin, A. D. Handoko, X. Chen,

Z. Guo, J. Tang, Angew. Chem., Int. Ed. 2014, 53, 9240.

Page 21: A Review of Direct Z‐Scheme Photocatalysts

© 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim1700080 (21 of 21)

www.advancedsciencenews.com www.small-methods.com

[93] J. X. Low, S. W. Cao, J. G. Yu, S. Wageh, Chem. Comm. 2014, 50, 10768.

[94] N. Linares, A. M. Silvestre-Albero, E. Serrano, J. Silvestre-Albero, J. García-Martínez, Chem. Soc. Rev. 2014, 43, 7681.

[95] W. J. Ong, L. L. Tan, S. P. Chai, S. T. Yong, A. R. Mohamed, Chem-SusChem 2014, 7, 690.

[96] H. Ahmad, S. Kamarudin, L. Minggu, M. Kassim, Renewable Sustainable Energy Rev. 2015, 43, 599.

[97] S. Wang, X. Wang, Small 2015, 11, 3097.[98] P. Zhang, T. Wang, J. Gong, Adv. Mater. 2015, 27, 5328.[99] Y. Xu, R. Xu, Appl. Surf. Sci. 2015, 351, 779.

[100] J. X. Low, B. Cheng, J. G. Yu, M. Jaroniec, Energy Storage Mater. 2016, 3, 24.

[101] X. Li, J. Q. Wen, J. X. Low, Y. P. Fang, J. G. Yu, Sci. Chin. Mater. 2014, 57, 70.

[102] L. Yuan, Y.-J. Xu, Appl. Surf. Sci. 2015, 342, 154.[103] W. J. Lee, J. Lim, S. O. Kim, Small Methods 2016, 1, 1600014.[104] T. M. Mata, A. A. Martins, N. S. Caetano, Renewable Sustainable

Energy Rev. 2010, 14, 217.

[105] J. Liu, C. Guo, A. Vasileff, S. Qiao, Small Methods 2016, 1, 1600006.[106] W. C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S. M. Haile,

A. Steinfeld, Science 2010, 330, 1797.[107] Q. L. Xu, J. G. Yu, J. Zhang, J. F. Zhang, G. Liu, Chem. Comm.

2015, 51, 7950.[108] M. S. Akple, J. X. Low, Z. Y. Qin, S. Wageh, A. A. Al-Ghamdi,

J. G. Yu, S. W. Liu, Chin. J. Catal. 2015, 36, 2127.[109] X. Zheng, L. Zhang, Energy Environ. Sci. 2016, 9, 2511.[110] M. Tahir, B. Tahir, N. A. S. Amin, Appl. Surf. Sci. 2015, 356, 1289.[111] J. G. Yu, J. Jin, B. Cheng, M. Jaroniec, J. Mater. Chem. A 2014, 2,

3407.[112] L. Shen, R. Liang, L. Wu, Chin. J. Catal. 2015, 36, 2071.[113] X. Liu, S. Inagaki, J. Gong, Angew. Chem., Int. Ed. 2016, 55, 14924.[114] J. X. Low, J. G. Yu, W. K. Ho, J. Phys. Chem. Lett. 2015, 6,

4244.[115] S. Ye, R. Wang, M.-Z. Wu, Y.-P. Yuan, Appl. Surf. Sci. 2015,

358, 15.[116] W. Dai, H. Xu, J. Yu, X. Hu, X. Luo, X. Tu, L. Yang, Appl. Surf. Sci.

2015, 356, 173.

Small Methods 2017, 1, 1700080

Page 22: A Review of Direct Z‐Scheme Photocatalysts

本文献由“学霸图书馆-文献云下载”收集自网络,仅供学习交流使用。

学霸图书馆(www.xuebalib.com)是一个“整合众多图书馆数据库资源,

提供一站式文献检索和下载服务”的24 小时在线不限IP

图书馆。

图书馆致力于便利、促进学习与科研,提供最强文献下载服务。

图书馆导航:

图书馆首页 文献云下载 图书馆入口 外文数据库大全 疑难文献辅助工具