ArticlePDF Available

2D Transition Metal Dichalcogenides for Photocatalysis

Wiley
Angewandte Chemie International Edition
Authors:

Abstract and Figures

Two‐dimensional (2D) transition metal dichalcogenides (TMDs), a rising star in the post‐graphene era, are fundamentally and technologically intriguing for photocatalysis. Their extraordinary electronic, optical, and chemical properties endow them as promising materials for effectively harvesting light and catalyzing the redox reaction in photocatalysis. Here, we present a tutorial‐style review of the field of 2D TMDs for photocatalysis to educate researchers (especially the new‐comers), which begins with a brief introduction of the fundamentals of 2D TMDs and photocatalysis along with the synthesis of this type of material, then look deeply into the merits of 2D TMDs as co‐catalysts and active photocatalysts, followed by an overview of the challenges and corresponding strategies of 2D TMDs for photocatalysis, and finally look ahead this topic.
This content is subject to copyright. Terms and conditions apply.
Photocatalysis
2D Transition Metal Dichalcogenides for Photocatalysis
Ruijie Yang+, Yingying Fan+, Yuefeng Zhang, Liang Mei, Rongshu Zhu,* Jiaqian Qin,
Jinguang Hu, Zhangxing Chen, Yun Hau Ng, Damien Voiry, Shuang Li, Qingye Lu,
Qian Wang,* Jimmy C. Yu,* and Zhiyuan Zeng*
Angewandte
Chemie
Reviews www.angewandte.org
How to cite:
International Edition: doi.org/10.1002/anie.202218016
German Edition: doi.org/10.1002/ange.202218016
Angew. Chem. Int. Ed. 2023, e202218016 (1 of 29) © 2023 Wiley-VCH GmbH
Angewandte
Chemie
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Abstract: Two-dimensional (2D) transition metal dichalcogenides (TMDs), a rising star in the post-graphene era, are
fundamentally and technologically intriguing for photocatalysis. Their extraordinary electronic, optical, and chemical
properties endow them as promising materials for effectively harvesting light and catalyzing the redox reaction in
photocatalysis. Here, we present a tutorial-style review of the field of 2D TMDs for photocatalysis to educate researchers
(especially the new-comers), which begins with a brief introduction of the fundamentals of 2D TMDs and photocatalysis
along with the synthesis of this type of material, then look deeply into the merits of 2D TMDs as co-catalysts and active
photocatalysts, followed by an overview of the challenges and corresponding strategies of 2D TMDs for photocatalysis,
and finally look ahead this topic.
1. Introduction
Two-dimensional (2D) transition metal dichalcogenides
(TMDs) constitute a charming material library.[1] They are a
promising alternative to graphene (zero band gap), and have
the potential to go beyond it in some quarters,[2] due to their
versatile chemistry, ranging from true metals (e.g., NbS2and
VSe2), semimetals (e.g., WTe2and TiSe2), and semiconduc-
tors (e.g., MoS2and WS2), to insulators (e.g., HfS2).[2]
This class of materials has recently attracted wide-
ranging interest in electronics, photonics, optoelectronics,
energy conversion, electrocatalysis, environmental remedia-
tion, biosensing, and bioimaging, to name but a few,[1–5]
owing to their extraordinary properties, including atomically
thin nature, strong spin-orbit coupling, tunable band gap,
direct energy gaps of monolayer counterparts in the near-
infrared to a visible spectral region, strong excitonic effects,
and naturally abundance.[1–5]
The past decade has also witnessed a rise in curiosity of
2D TMDs for photocatalysis both fundamentally and
technologically.[6–9] An astounding increase in publications
regarding this topic has been seen in the last few years
(Figure 1), since the pioneering work[10] published in 2008
(using MoS2as co-catalyst for photocatalytic hydrogen
production). Despite this being a young field and challenges
remain, it indeed points to a productive present and a bright
future, both as co-catalysts[10–22] and active photocatalysts
(light-harvesting materials).[23–37]
In this Review, we first introduce the fundamentals of
2D TMDs and photocatalysis as well as the synthesis of this
type of material, before delving into the advantages of them
(as co-catalysts or active photocatalysts) compared to the
classical photocatalytic materials. Then, an essential over-
view is presented toward the challenges and corresponding
strategies (including the engineering of edge site, phase,
doping, vacancy, and interface) of 2D TMDs for photo-
catalysis. Finally, we outlook the future opportunities of
these materials in the photocatalysis field.
1.1. Fundamentals of 2D TMDs
Composition, crystal phases, and electronic band structure
of 2D TMDs are introduced here, aiming to give readers a
preliminary understanding of this material family before
introducing its photocatalytic applications.
[*] R. Yang,+Y. Fan,+Y. Zhang, L. Mei, Prof. Dr. Z. Zeng
Department of Materials Science and Engineering, and State Key
Laboratory of Marine Pollution, City University of Hong Kong
83 Tat Chee Avenue, Kowloon, Hong Kong 999077 (P. R. China)
E-mail: zhiyzeng@cityu.edu.hk
R. Yang,+Y. Fan,+Prof. Dr. J. Hu, Prof. Dr. Z. Chen, Prof. Dr. Q. Lu
Department of Chemical and Petroleum Engineering, University of
Calgary
2500 University Drive, NW, Calgary, Alberta, T2N 1N4 (Canada)
Prof. Dr. R. Zhu
State Key Lab of Urban Water Resource and Environment, School of
Civil and Environmental Engineering, Harbin Institute of Technol-
ogy Shenzhen
Shenzhen 518055 (P. R. China)
E-mail: rszhu@hit.edu.cn
Prof. Dr. J. Qin
Center of Excellence in Responsive Wearable Materials, Metallurgy
and Materials Science Research Institute, Chulalongkorn University
Bangkok 10330 (Thailand)
Prof. Dr. Y. Hau Ng
Low-Carbon and Climate Impact Research Centre, School of Energy
and Environment, City University of Hong Kong
83 Tat Chee Avenue, Kowloon, Hong Kong SAR (P. R. China)
Prof. Dr. D. Voiry
Institut Européen des Membranes, IEM, UMR 5635, Université
Montpellier, ENSCM, CNRS
Montpellier (France)
Prof. Dr. S. Li
College of Polymer Science and Engineering, State Key Laboratory
of Polymer Materials Engineering, Sichuan University
Chengdu (China)
Prof. Dr. Q. Wang
Graduate School of Engineering, Nagoya University
Furo-cho, Chikusa-ku, Nagoya 464-8603 (Japan)
and
Institute for Advanced Research, Nagoya University
Furo-cho, Chikusa-ku, Nagoya 464-8601 (Japan)
E-mail: wang.qian@material.nagoya-u.ac.jp
Prof. Dr. J. C. Yu
Department of Chemistry and Materials Science and Technology
Research Centre, The Chinese University of Hong Kong
Shatin, New Territories, Hong Kong 999077 (China)
E-mail: jimyu@cuhk.edu.hk
Prof. Dr. Z. Zeng
Shenzhen Research Institute, City University of Hong Kong
Shenzhen 518057 (China)
E-mail: zhiyzeng@cityu.edu.hk
[+] These authors contributed equally to this work.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (2 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1.1.1. Composition
TMDs, commonly marked as MX2(M: transition metal
elements from group IVB to group VIII; X: S, Se and Te
elements of group VIA) (Figure 2a), are typical 2D layered
nanomaterials. Each TMD monolayer contains three atom-
layers, forming an “X–M–X’’ sandwich structure.[1] The
adjacent monolayers, with a distance of 6–7 Å, are held
together by weak van der Waals forces, which makes them
readily isolated.[2]
1.1.2. Crystal Phases
TMDs exist in several structural phases (Figure 2b). Specif-
ically, there are three well-known polytypic structures
resulting from different coordination geometry of the
transition metal atoms, namely 2H (trigonal prismatic), 1T
(octahedral), and 1T(distorted octahedral).[38,39] Their
stacking orders of the three atomic planes (X–M–X) are also
different. 2H-TMDs exhibit a Bernal (ABA) stacking,
whereas the stacking order of atomic planes in 1T-TMDs is
rhombohedral ABC.[1] The diversity of structural phases
gives rise to various electronic properties. For example, 2H-
Ruijie Yang received his M.S. degree
from Harbin Institute of Technology,
China, in 2019. He is currently a Ph.D.
student at the University of Calgary,
Canada. Prior to this, he worked as a
research assistant in Dr. Zhiyuan Zeng’s
group at City university of Hong Kong.
His research interests focus on 2D
materials, catalysis, and in situ charac-
terizations.
Yingying Fan received his M.S. degree
from Harbin Institute of Technology,
China, in 2020. She is currently a Ph.D.
student at the University of Calgary,
Canada. Her research interests focus on
photocatalytic hydrogen production, 2D
transition metal dichalcogenides, and
biomass photo-refining.
Rongshu Zhu studied at the Harbin
University of Technology, receiving her
Ph.D. in 2007. Since then, she has been
engaged in teaching and scientific re-
search at Harbin Institute of Technology
(Shenzhen). In 2010, she was rated as
the local leading talent in Shenzhen. She
established the “Shenzhen Key Labora-
tory for organic pollution prevention
and control” in 2016 and “International
Joint Research Center for Persistent Toxic
Substances (IJRC-PTS) Harbin Institute of
Technology (Shenzhen) Sub Center” in
2017, where she is serving as the director
and executive deputy director, respectively. Her research interests
include catalysts for environmental application and energy conversion.
Qian Wang is currently an Associate
Professor at Nagoya University, Japan.
She obtained her Ph.D. in 2014 at the
University of Tokyo, Japan, where she
worked on the development of perov-
skite-type oxide photocatalysts for visi-
ble-light-driven water splitting. She then
worked as a postdoctoral researcher at
the Japan Technological Research Asso-
ciation of Artificial Photosynthetic Chem-
ical Processes (ARPChem) on the devel-
opment of standalone photocatalyst
devices for overall water splitting. In
2018, she became a Marie Sklodowska-
Curie Research Fellow at the University of Cambridge to develop
inorganic-organic hybrid photocatalysts. She joined Nagoya University
as an Associate Professor in May 2021 and established her research
group, which is currently developing new materials, approaches, and
technologies for solar energy storage in the form of renewable fuels via
artificial photosynthesis.
Jimmy C. Yu is an emeritus professor at
The Chinese University of Hong Kong.
Before retiring from teaching, he was
Choh-Ming Li Professor of Chemistry and
Head of United College. Professor Yu
received a B.S. degree from St. Martin’s
College and a Ph.D. from the University
of Idaho. He has been a highly cited
researcher for many years. His major
research interest is photocatalysis, and
he holds several patents on the fabrica-
tion and application of photocatalytic
nanomaterials. Prof. Yu has received
numerous honors including a State
Natural Science Award and Chang Jiang Scholar Chair Professorship.
Zhiyuan Zeng is an assistant Professor in
Department of Materials Science and
Engineering, City University of Hong
Kong. He obtained his B.E. and M.E. from
Central South University, Zhejiang Uni-
versity, in 2006 and 2008, respectively.
He completed his Ph.D. at Nanyang
Technological University, Singapore
(2013). He started his postdoctoral train-
ing at Lawrence Berkeley National Labo-
ratory (2013–2017), followed by Senior
Engineer at Applied Materials Inc. (2017–
2019, Silicon Valley); he joined CityU in
2019. His research interests include lith-
ium intercalation & exfoliation method for 2D TMDs, in situ liquid
phase TEM, environmental catalysis, water desalination, etc.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (3 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MoS2shows semiconducting attributes,[38] whereas 1T-MoS2
exhibits a metallic character.[1]
1.1.3. Electronic Band Structure
TMDs possess a tunable (thickness-dependent) electronic
band gap (Figure 3a), resulting from quantum confinement
effect. With 2H-MoS2as exemplification, their calculated
electronic band gap values range from 0.88 eV (bulk) to
1.71 eV (monolayer).[1, 40] Interestingly, when the thickness
of MoS2decreases to monolayer, it changes from an indirect
band gap semiconductor to a direct band gap semiconductor,
along with a dramatic jump in luminescence (Figure 3b).[41]
Additionally, 2.16 eV is the experimental value of the
electronic band gap for monolayer MoS2.[42] Naturally, the
band gap values between different TMD monolayers are
also different (Figure 3c),[23] which leads to their different
spectral responsivities (Figure 3d).[43]
1.2. Fundamentals of Photocatalysis
Rudimentary knowledge of photocatalysis is introduced
here, aiming to help readers have a primary impression of
this catalytic technology, before focusing it on 2D TMDs.
1.2.1. Advantages
Photocatalysis employs solar energy, which is an abundant,
non-polluting and renewable natural energy source, to drive
Figure 1. The pioneers and the annual number of publications of 2D transition metal dichalcogenides (TMDs) for photocatalysis. The number of
publications (in the past decade) were collected from Google Scholar at December 22, 2022 by searching the two topical keywords of “2D
Transition Metal Dichalcogenides” and “Photocatalysis”.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (4 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
chemical processes, representing a promising approach to
store the solar energy into high-valued chemicals such as
hydrogen, methane and carbon monoxide.[44–46] It shows
superiorities over traditional catalysis. Unlike traditional
thermal catalysis, which typically requires high temperatures
and a high-pressure operating environment, most photo-
catalysis is operated under ambient conditions.[44, 47,48] Be-
sides, overoxidation and catalyst deactivation are common
problem of thermal catalysis, but can be avoided in
photocatalysis.[44] All these factors contribute to the advant-
age of photocatalysis, which points to a bright future.
1.2.2. Basic process
In general, the basic photocatalytic process involves three
steps (Figure 4a):[49–53] (1) light absorption by the light-
absorbing material in photocatalyst; (2) separation and
migration of the photogenerated charge carriers; (3) surface
redox reactions with the assistance of cocatalysts on the
semiconductor surface. Of note, the carrier migration
process is always accompanied by carrier recombination,
including bulk recombination and surface recombination,
which is not conducive to photocatalysis and should be
avoided as much as possible.
Figure 2. Composition and crystal phase of TMDs. a) Composition of TMDs. b) Schematic models of different phases of TMDs, including
hexagonal (2H), octahedral (1T), and distorted octahedral (1T). Top view shows the typical atomic arrangement of a single layer (top panel). Side
view shows the typical packing sequences (bottom panel).
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (5 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 3. Electronic band structure of TMDs. a) The valence band (VB) and conduction band (CB) positions of TMDs as a function of the layer
number calculated by Perdew–Burke–Ernzerhof (PBE). The vacuum level of zero is used as reference. Data collected from ref. [40]. b) Calculated
band structures of bulk MoS2, and MoS2flakes with four layers, two layers, and one layer. The arrows in each part indicate the lowest energy
transitions. Reproduced with permission from ref. [41]. Copyright 2010, American Chemical Society. c) The valence band (VB) and conduction band
(CB) positions of TMDs monolayers calculated by many-body G0W0approximation method. The vacuum level is zero used as reference. Data
collected from ref. [23]. d) Spectral responses of 2D TMDs. NUV: near ultraviolet; VIS: visible; NIR: near infrared; and SIR: longwave infrared. Data
collected from ref. [43]
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (6 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1.2.3. Thermodynamics requirements
The occurrence of photocatalytic reaction needs to meet
thermodynamics requirements:[54–57] (1) the energy of inci-
dent photons should be equal to or greater than the optical
band gap (Eg) of the semiconductor; (2) the valance band
maximum (VBM) of semiconductor needs to be more
positive than the oxidation potential of the donor (E(D/D));
(3) while the conduction band minimum (CBM) needs to be
more negative than the reduction potential of the acceptor
(E(A+/A)).
1.2.4. Kinetic Challenges
In addition to the above-mentioned thermodynamics re-
quirements, photocatalysis also faces a lot of kinetic
challenges:[58] (1) the three basic steps span a huge timescale
from 1015 to 101s (1015–109s of light absorption, <1015 s
of charge separation and transport, and 103–101s of surface
reactions),[59] which poses a great challenge to maximize the
synergy between the three-step reaction; (2) an optimum
photocatalyst needs to simultaneously hold broad light-
absorption (needing narrow band gap) in the solar radiation
spectrum, and strong redox ability (needing wide band gap),
but their realization is inherently contradictory; (3) the
semiconductor photosensitizers typically lack active sites.
Therefore, cocatalysts are normally required to be loaded
Figure 4. Fundamentals of photocatalysis. a) Basic process of photocatalysis. b) Photocatalytic system configuration. c) Parameters for activity
evaluation of photocatalysis. STH: solar-to-hydrogen conversion efficiency. AQY: apparent quantum yield. TOF: turnover frequency.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (7 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
on the semiconductor surface to promote charge separation
and transfer and reduce the activation energy.[60]
1.2.5. Photocatalytic System Configuration
Reasonable construction of a photocatalytic system is vital
for photocatalysis. In general, there are two types of system
configurations in photocatalysis:[6,45] “suspension” system
and “fixed” system (Figure 4b). In the “suspension” system,
the particulate photocatalyst is directly suspended in the
reaction solution. As the name suggests, the “fixed” system
refers to photocatalyst fixed on a substrate for photo-
catalytic reaction. For detail information about the photo-
catalytic system configuration, the reader is referred to a
current review article.[6]
1.2.6. Parameters for Activity Evaluation
The ultimate pursuit of photocatalysis is the efficient
utilization of solar energy. Thus, how to evaluate efficiency
of the system is important. Herein, we outlined the common
parameters for the evaluation of photocatalytic activities
(Figure 4c): solar-to-chemical conversion efficiency (for
example, solar-to-hydrogen conversion efficiency, STH),
apparent quantum yield (AQY), and turnover frequency
(TOF). In addition to the above-mentioned parameters,
stability is another important index for the practical
application of photocatalysis, which is worthy of attention.
The direct evaluation strategy for stability is to conduct a
long-term photocatalytic reaction test. For more information
about the efficiency accreditation and testing protocols of
photocatalysts, the reader is referred to a current author-
itative article.[61]
1.3. Synthesis of 2D TMDs
Reliable synthesis methods of 2D TMDs are presented here
to provide readers a concise guide for the synthesis of this
type of material. Ideally, the synthesis methods of 2D TMDs
can be divided into two categories: top-down exfoliation
(from bulk to 2D nanosheets; exampled by mechanical
cleavage,[41,62, 63] direct liquid exfoliation,[64, 65] and intercala-
tion-based liquid exfoliation[66–70]) and bottom-up synthesis
(from small building block molecules to 2D nanosheets;
exampled by chemical vapour deposition growth[71–73] and
wet-chemical synthesis[74–77]). The pros and cons of each
method and their main applications are listed in Supporting
Information Table 1.
1.3.1. Mechanical Cleavage
Mechanical cleavage (Figure 5a), employing scotch tape to
cleave atom-layers from their bulk crystal, is a simple and
common method for the exfoliation of TMDs.[78] The
produced 2D TMDs have the characteristics of single crystal,
high purity, and cleanliness, which are suitable for device
applications.[1] Theoretically, this method can be used to
exfoliate all the layered bulk TMDs to generate the ultrathin
2D counterparts, but it is not scalable.[79] Additionally, it is
relatively difficult for the systematic controlling of the
thickness, size, and shape of the final products.
1.3.2. Direct Liquid Exfoliation
For obtaining exfoliated 2D TMDs on a large scale
(mechanical cleavage is not feasible), liquid exfoliation is
promising.[4] Direct liquid exfoliation (Figure 5b) refers to
the direct exfoliation of the bulk TMDs into their ultrathin
2D counterparts in solvents, by breaking the weak van der
Waals interaction via sonication or shear forces.[64, 65] N-
methyl-pyrrolidone (NMP) and dimethylformamide (DMF)
are the commonly used two solvents.[64] Simple operation
steps, low costs, solution-processability, and insensitivity to
air and water make it show great potential in commercial
applications. However, direct liquid exfoliation is plagued by
the low-yield of the monolayers, the small lateral size of the
exfoliated flakes and the toxicity of the used organic
solvents.[65]
1.3.3. Intercalation-based Liquid Exfoliation
Intercalation-based liquid exfoliation[66–70,80] (Figure 5c) is an
up-and-coming alternative for direct liquid exfoliation, and
have the potential to overcome its deficiency mentioned
above.[70] The underlying principle of intercalation-based
liquid exfoliation is to enlarge the interlayer spacing and
thereby weaken the interlayer adhesion before exfoliation
by intercalating foreign species (ions or molecules).[70] The
typical procedure involves chemical or electrochemical
intercalation of foreign species into the interlayer spaces of
layered TMDs, followed by a mild exfoliation process.[70]
Lithium-ion (Li+) is a commonly used intercalant, but
always introduces basal-plane defects and induces 2H-to-1T/
1Tphase transitions.[66,67, 70] Larger ions, such as tetraalky-
lammonium cation (R4N+), can avoid such phase transitions,
but the increased size inevitably leads to the increase in
intercalation barrier, and thus reduces the intercalation rate
of bulk TMDs crystals, which thereby decrease the exfolia-
tion rate and the yield of monolayers.[68, 69] Low energetic
cost, solution-processability, and yield of large-area TMD
monolayers are advantageous of intercalation-based liquid
exfoliation strategy.
1.3.4. CVD Growth
CVD growth (Figure 5d) is a classical bottom-up strategy
for the wafer-scale synthesis of 2D TMDs on substrates
(e.g., the traditional SiO2/Si).[71–73] In a typical process, the
vaporized reactive precursors (e.g., the sulphur powder and
MoO3powder[71]) pass through the surface of the target
substrate with the specific gas flow under a high temperature
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (8 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 5. Synthesis of 2D TMDs. a) Mechanical cleavage. b) Direct liquid exfoliation. c) Intercalation-based liquid exfoliation. d) CVD growth.
e) Wet-chemical synthesis. Left panel: hydro/solvothermal synthesis. Right panel: hot-injection synthesis.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (9 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
and high vacuum environment, during which the precursors
undergo reaction, decomposition, in situ nucleation and
growth on the substrate surface, finally forming the 2D
TMD flakes.[71–73] The produced 2D TMDs are large-size,
high-purity, and high-crystallinity. However, this method is
hampered by the harsh working environment (high temper-
ature and high vacuum).[1]
1.3.5. Wet-chemical Synthesis
2D TMD flacks can also be prepared by a wet-chemical
synthesis process (another famous bottom-up strategy), such
as hydro/solvothermal synthesis (left panel of Figure 5e)[74,75]
and hot-injection synthesis (right panel of Figure 5e).[76, 77]
This method is simpler and easier to operate than CVD
growth, but the product quality and crystallinity are not as
good as those obtained by CVD growth. Hydro/solvother-
mal synthesis of 2D TMDs occurs under high temperature
and high pressure in a sealed vessel and uses water/organic
solvent as the reaction medium, during which metal
precursor (such as molybdic acid) and chalcogen precursor
(such as thiourea) reacted with each other to form TMD
crystals. The resulting 2D TMD flacks hold rich defects and
active edge sites, showing outstanding catalytic
performance,[74,75] but their thickness is not conclusively
shown to be single layers. Such a method is simple and
scalable, but the product quality is sensitive to reaction
conditions, such as operating temperature, time, type and
concentration of precursors, solvents and surfactants. Hot-
injection method is an effective strategy for 2D TMDs
colloid synthesis, which induce chemical reactions by rapidly
injecting active reactant molecules (e.g., CS2and WCl6
precursors) into a hot solution containing special surfactants
(long chain oleylamine and/or oleic acid).[76,77] Such a hot-
injection method produces 2D TMDs with good dispersity,
high purity, and uniform size and shape, but limitations still.
The uses of hard-to-remove surfactants (long chain oleyl-
amine and oleic acid) are unfavorable for catalytic applica-
tions. Besides, the special operation of the precursor
injection leads to the impossibility of mass production.
1.3.6. Others
Beyond these commonly used methods mentioned above,
the past few years also saw the emergence of some novel
strategies for the production of 2D TMDs, such as molecular
beam epitaxy (MBE),[81] atomic layer deposition (ALD),[82]
pulsed laser deposition,[83] and atomic layer etching by
planar surface plasmon polaritons (SPPs).[84] The emergence
of these technologies enriches the toolbox of 2D TMDs
preparation.
1.3.7. Analysis
Rich methodologies have been established with their own
merits for the synthesis of 2D TMDs flakes (Supporting
Information Table 1). Still, none of them are yet perfect.
Which method is more suitable for researchers (especially
new comers)? Demands decide choices. For example, acting
as active photocatalyst needs 2D TMDs hold semiconduct-
ing phase; mechanical cleavage,[41,62, 63, 78] direct liquid
exfoliation,[64,65] organic ammonium ion intercalation-based
exfoliation,[68,69] CVD growth,[71–73] and hydrothermal
method[74,75] are feasible for the synthesis of semiconducting
phase 2D TMDs. By contrast, hydrothermal method is the
most-friendly one for novices, because it is simple, easy to
operate and has a high success rate. 2D TMDs as cocatalyst,
preferably metallic phase; mechanical cleavage,[78] lithium-
ion intercalation-based exfoliation,[66,67, 70] CVD growth,[85, 86]
and hydrothermal method[87,88] are workable for the fabrica-
tion of metallic phase 2D TMDs. Our recommend one is the
lithium-ion intercalation-based exfoliation method[66,67, 70] (of
particular the electrochemical lithium-ion intercalation-
based exfoliation method[67,70]). Because this method is more
capable of large-scale preparation of metallic phase 2D
TMDs than the mechanical cleavage, it is easier to operate
than the CVD growth, and the phase of the synthesized 2D
TMDs is purer than the hydrothermal method.[70] For the
detailed operation process of electrochemical lithium-ion
intercalation-based exfoliation method, the researchers (of
particular beginners) are recommended a recently published
protocol.[70] Of note, this method needs to be carried out in
an argon-filled glovebox to prevent lithium from being
oxidized by air. The second choice for synthesizing metallic
phase 2D TMDs is hydrothermal method, which is simple
and easy to operate, although the phase synthesized is not
too pure.[87,88]
2. Why 2D TMDs for Photocatalysis
Here gives the advantages of 2D TMDs in photocatalysis
compared to the classical photocatalytic materials (including
the classical cocatalysts, Pt, Pd, Rh, Ru, and Au, the classical
semiconductor photocatalysts, TiO2, WO3, and CdS, as well
as the classical 2D photocatalytic materials, graphene and
graphitic carbon nitride), to explain why there is a possibility
to replace the classical photocatalytic materials.
2.1. Merits of 2D TMDs as Cocatalysts
2D TMDs (of particular their metallic phase crystal, for
example 1T-MoS2) are promising cocatalysts in
photocatalysis[10–22] (some representative examples of 2D
TMDs as cocatalysts are listed in Supporting Information
Table 2). They are an ideal alternative to noble metals and
even have the potential to go beyond noble metals in
performance. For example, 2D MoS2[10] and WS2[11] as
cocatalysts showed a better promoting effect on CdS for
photocatalytic H2evolution from 10 vol % lactic solution
than noble metals, including Pt, Pd, Rh, Ru, and Au
(Figure 6a). Such remarkable performance of metallic phase
2D TMDs as cocatalysts in photocatalysis results from their
extraordinary features (listed below).
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (10 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 6. Merits and functions of 2D TMDs as cocatalyst in photocatalysis. a) The rate of H2evolution on CdS loaded with 0.2 wt % of MoS2, Pt,
Ru, Rh, Pd and Au. Date collected from ref. [10]. b) Merit of metallic phase 2D TMDs as cocatalyst: good conductivity. c) Function of metallic phase
2D TMDs as cocatalyst: providing trapping sites for photo-generated charges and promote charge separation, migration, and collection. d) The
electron occupations in Mo 4d orbital in 1T-MoS2and 2H-MoS2crystal fields. e) Merit of metallic phase 2D TMDs as cocatalyst: rich active sites on
both the edge and basal plane. f) Function of metallic phase 2D TMDs as cocatalyst: playing an active role as adsorption, activation, and reaction
site of reactant molecules to lower the activation energies and accelerate reaction kinetics.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (11 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2.1.1. Good Conductivity
Metallic phase 2D TMDs hold good conductivity (Fig-
ure 6b),[1,89] enabling them to provide trapping sites for
photo-generated charges and promote charge separation,
migration, and collection, thus improving quantum efficiency
(Figure 6c). For example, metallic 1T phase MoS2(with high
electrical conductivity of 10–100 S cm1) is almost 107times
more conductive than its semiconducting 2H phase and is
close to graphene (100 S cm1),[89] which endows it with
the role of ideal cocatalyst in photocatalysis like
graphene.[16,18] Generally, the high conductivity of materials
is related to their electronic structure and atomic
arrangement.[90] For metallic 1T phase MoS2, it has the
incomplete occupation of the lower-lying t2g orbital (Fig-
ure 6d),[90] which makes it easy for the outermost valence
electrons to break free from the atomic nucleus and become
free electrons. This property is the same as that of a metal
atom, which leads to its high conductivity like that of a
metal.
2.1.2. Rich Active Sites on Both the Edge and Basal Plane
In parallel, metallic phase 2D TMDs have rich active sites
on both the edge and basal plane (Figure 6e),[91] endowing
them to play an active role as adsorption, activation, and
reaction site of reactant molecules to lower the activation
energies and accelerate reaction kinetics (Figure 6f). This
point of view (rich active sites on both the edge and basal
plane) was confirmed by theoretical calculations and exper-
imental verification, which will be described in detail below
(with hydrogen evolution reaction as an exemplification).
In the field of photocatalysis, active sites are normally
located at the position with strong reactivity. The reactivity
of catalyst is related to adsorption energy.[7] Sabatier
principle shows that the ideal free energy of adsorption
should be zero.[92] Over free energy of adsorption always
leads to a poor binding between the catalyst and the reactant
(causing the low reaction rate).[92] Insufficient free energy of
adsorption will result in a slow desorption of atoms (leading
to the poisoning of the catalyst surface).[92] Both of these
situations lead to weak reactivity.
Pioneering theoretical calculations (based on density
functional theory) has shown a zero-nearing Gibbs free
energy of adsorption of hydrogen atom (ΔGH) for the
metallic phase TMDs edge[93, 94] and basal plane (Fig-
ure 7a,b, d),[95] reveling that metallic phase 2D TMDs have
strong reactivity (rich active sites) on both the edge and
basal plane. Numerous substantial studies have also corro-
borated this view experimentally.[91,96,97] However, for semi-
conducting phase 2D TMDs, although their edge sites have
zero-nearing ΔGH(Figure 7a,b), their basal plane’ ΔGHare
usually very high (Figure 7c).[95] This reveals that their
(semiconducting phase 2D TMDs) active sites for hydrogen
evolution reaction are only at the edges, and their basal
planes are catalytically inert.[95]
2.1.3. Dense Interface Junctions
Crucially, their (2D TMDs) atomically thin nature grants
them to form dense interface junctions with light-harvesting
semiconductors, which guarantees the rapid transfer of
photo-induced charges.[16] Such dense interface junctions are
constructed by chemical bond bridge or van der Waals
force.
Chemical bond bridges are easily formed between 2D
TMDs and light-harvesting semiconductors, such as BiS
bond bridges in Bi12O17Cl2
MoS2interface (Figure 8a),[16]
ZnS bond bridges in ZnIn2S4
MoS2interface (Fig-
ure 8b),[18] and ORe (or OMo) bond bridges in TiO2
ReS2
or TiO2
MoS2interface (Figure 8c).[20,21] Such bridges pro-
vide a high-speed channel for electron migration, thus
promoting the separation of photo-generated electron-hole
pairs, prolonging the carrier lifetime (II panel in Figure 8a)
(3446 ns of Bi12O17Cl2
MoS2; 136 ns of Bi12O17Cl2), and
improving photocatalytic efficiency (32.68 mmol·g1·h1for
H2evolution of Bi12O17Cl2
MoS2; 0.86 mmol g1h1for H2
evolution of Bi12O17Cl2). This is because the existence of
such bridges reduces the energy barrier and distance of
electron migration compared with bridge-free interface; for
example, the energy barrier and distance of electron
migration at BiS bond bridges is 5 eV and 2.3 Å, respec-
Figure 7. Theoretical calculated (based on density functional theory)
thermodynamics of hydrogen adsorption on 2D transition metal
dichalcogenides. a) The differential ΔGHas a function of the ΔGHX
(X =S) of TMD monolayers edge sites. b) The differential ΔGHas a
function of the ΔGHX (X =Se) of TMD monolayers edge sites. c) The
differential ΔGHas a function of the ΔGHX of semiconducting TMD
monolayers basal planes. e) The differential ΔGHas a function of the
ΔGHX of metallic TMD monolayers basal planes. ΔGH: Gibbs free
energy of adsorption of hydrogen atom. ΔGHX: Gibbs free energy of
adsorption of XH (X =S or Se). Reproduced with permission from
ref. [95]. Copyright 2015, Elsevier.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (12 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 8. Electron bridge formed between metallic phase 2D TMDs and light-harvesting semiconductors. a) BiS electron bridge formed in
Bi12O17Cl2
MoS2, showing an electron transfer from Bi to S along BiS bond bridge. Top left panel in part I: schematic illustration of the
morphology. Bottom left panel in part I: schematic illustration of the occurred photocatalytic redox reaction, showing that H2evolution reaction
(reductive reaction) occurred on MoS2layer; while, ascorbic acid conversion reaction (oxidation reaction) occurred on Bi12O17Cl2layer. Right panel
in part I: schematic illustration of the formed electron bridge. Part II: femtosecond-resolved transient absorption (TA) spectroscopy of
Bi12O17Cl2
MoS2with BiS bonds (red curve) and without BiS bonds (bule curve). Part III: the calculated electrostatic potentials of (Cl2)-(Bi12O17)-
(MoS2) with BiS bonds. Part IV: the calculated electrostatic potentials of (Cl2)-(Bi12O17)-(MoS2) without BiS bonds. Top left panel in part I and
part II-IV are reproduced with permission from ref. [16]. Copyright 2016, Nature Publishing Group. b) ZnS electron bridge formed in
ZnIn2S4
MoS2, showing an electron transfer from Zn to S along ZnS bond bridge. Top panel: schematic illustration of the morphology and
formed electron bridge. Bottom panel: schematic illustration of the occurred photocatalytic redox reaction, showing that H2evolution reaction
(reductive reaction) occurred on MoS2layer; while, lactic acid conversion reaction (oxidation reaction) occurred on ZnIn2S4layer. Top panel is
reproduced with permission from ref. [18]. Copyright 2018, American Chemistry Society. c) OMo electron bridge formed in TiO2
MoS2, showing
an electron transfer from O to Mo along OMo bond bridge. Top panel: schematic illustration of the morphology and formed electron bridge.
Bottom panel: schematic illustration of the occurred photocatalytic redox reaction, showing that H2evolution reaction (reductive reaction) occurred
on MoS2layer; while, ethanol conversion reaction (oxidation reaction) occurred on TiO2layer. Top panel is reproduced with permission from
ref. [20]. Copyright 2016, John Wiley & Sons.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (13 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
tively (III panel in Figure 8a); while, for BiS bond bridge
free interface, they are 17 eV and 4.5 Å, respectively (IV
panel in Figure 8a).
Alternatively, a plethora of dense interface junctions can
be formed by combine 2D TMDs with several other 2D
nanosheets in one vertical stack,[98–100] held together by van
der Waals forces[101–104] (the same forces that hold layered
crystals together). Such dense interface junctions have
ultrafast excited-state dynamics, allowing ultrafast electron
migration, long-lived spin and valley polarization in resident
carriers, and thereby promoting the separation of photo-
generated electron-hole pairs.[105]
2.1.4. Some Other Attractive Interfacial Surface Properties and
Behaviors
Metallic phase 2D TMDs also hold some other attractive
interfacial surface properties and behaviors, including active
site self-optimization,[22] rich sulfur vacancies as electron
enrichment domains,[106] induced photothermal effect,[107]
and beyond. These properties and behaviors stimulate more
potential for them as cocatalysts in photocatalysis. In the
following, we take the active site self-optimization as a
notable example to give a brief introduction.
Pioneering work by Wang et al.[22] has shown that 1T
phase MoS2(normally been considered as the main source
of active sites in hydrogen production reactions) can be
irreversibly converted to 1Tphase (a more catalytically
active phase) as true active sites as cocatalysts in photo-
catalytic hydrogen production reactions, named “active site
self-optimization”. They proposed that the adsorption of
hydrogen atoms will cause electron transfer, which is the
main driving force for this phase transition. Such an “active
site self-optimization” behaviors greatly boosted the photo-
catalytic performance of TiO2in the photochemical
environment.[22]
2.2. Merits of 2D TMDs as Active Photocatalysts
Semiconducting phase 2D TMDs (for instance 2H-MoS2)
are potential candidates of active photocatalysts (light-
harvesting materials)[24–26,28–37] (some representative exam-
ples of 2D TMDs as active photocatalysts are listed in
Supporting Information Table 3). Their advantages as active
photocatalysts are list below.
2.2.1 Narrow Band Gaps
They hold narrow band gaps (normally <2.4 eV, for
instance, 2.16 eV of MoS2monolayer[42]), endowing them
with wider spectral absorption (cut-off wavelength ranging
from near-infrared to visible spectral[43]) and more efficient
solar spectrum utilization (for example, 50% of solar
spectrum utilization of few-layer MoS2[25]) than traditional
semiconductors (e.g., TiO2–3.2 eV-4% of solar spectrum
utilization, WO3–2.8 eV-10% of solar spectrum utilization,
and CdS-2.4 eV-20 % of solar spectrum utilization) (Fig-
ure 9). Besides, their narrow band gaps resulting in their
Figure 9. Band gaps and light-absorption properties of semiconducting phase 2D TMDs and some representative photocatalysts.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (14 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
strong light-matter interactions,[108,109] ensuring their strong
electron-hole creation.
2.2.2. Well-placed Energy Levels
In parallel, their well-placed energy levels meet the
thermodynamics requirements of a variety of photocatalytic
reactions (Figure 10), including hydrogen (H2)
evolution,[23,29, 30, 35, 110] carbon dioxide (CO2) reduction,[32] and
nitrogen (N2) fixation,[27,111] sterilization,[25, 31] pollution
degradation,[24] and organic synthesis,[26,28] which have been
proven theoretically and experimentally.
2.2.3. Atomically Thin Nature
More to the point, their atomically thin nature grants them
more extraordinary properties as light-harvesting materials
in photocatalysis, including quantum confinement, short
transmission distance of carriers, large surface-area-to-
volume ratio, and so on.[6] Quantum confinement effect
endows them with tunable band structure, proven by
theoretical calculation[40] (Figure 3a) and experimental
detection[112] (Figure 11a), resulting in diversified optical
properties to meet a variety of photocatalytic reactions. The
short transmission distance of carriers from the inner to the
surface is good for the effective suppression of bulk
recombination of carriers, prolonging the lifetime of light
quanta, and thereby improving the quantum efficiency
(Figure 11b). Large surface-area-to-volume ratio is in favor
of light capturing, and makes a great contribution to rich
point contact between the reactants and catalytic sites, and
provides a good platform for the formation of doping, non-
killer defects and heterojunction, which are all beneficial for
the further enhancing of the photocatalytic performance
(Figure 11c).
2.2.4. As Ideal Platform for Fundamental Research
Additionally, semiconducting phase 2D TMDs, especially
monolayer, can act as an ideal platform to demonstrate
some fundamental principles in photocatalysis, resulting
from their unique properties mentioned above. For example,
they can be used to demonstrate the direction and time of
carrier transfer after light excitation at the heterostructure
interface.[108] A prior study has experimentally verified
this.[108] In this study, using 2D TMDs-based van der Waals
heterostructure (constructed with MoS2monolayer and WS2
monolayer) as a model, ultrafast transfer of carriers at this
van der Waals interface was demonstrated.[108] Photolumi-
nescence mapping and femtosecond pump-probe spectro-
scopy revealed that photoinduced holes migrate from the
MoS2layer to the WS2layer in a very short time ( 50 fs)
after light excitation (Figure 11d).
2.3. Unique Merits of 2D TMDs for Photocatalysis Compared
with Classical 2D Photocatalytic Materials
Why 2D TMDs can stand out in photocatalysis from many
classical 2D photocatalytic materials is explained here from
the current perspective.
2.3.1. Comparison with Graphene
Graphene, while being theoretically and practically interest-
ing for photocatalysis as classical 2D cocatalysts,[113–115] is
chemically inert and zero-band-gap.[2] This limits graphene
to acting as a cocatalyst in photocatalysis and always need to
be activated by functionalization with desired molecules.[116]
2D TMDs exhibit versatile chemistry, ranging from true
metals (zero-band-gap, like graphene), semimetals (near-
zero-band-gap), to semiconductors (narrow-band-gap).[2]
This offers 2D TMDs the opportunities for diverse funda-
mental and technological researches in a variety of photo-
catalytic reactions both acting as cocatalysts and active
photocatalysts. Such a diversity is the most obvious advant-
age of 2D TMDs over graphene in photocatalysis.
2.3.2. Comparison with Graphitic Carbon Nitride
The graphitic carbon nitride (g-C3N4),[117–119] a representative
2D semiconductor photocatalyst, has become a new research
hotspot in the field of photocatalysis. Its band gap is 2.7 eV,
theoretically resulting in 10%-20 % of solar spectrum
utilization.[117–119] As previously mentioned, 2D TMDs hold
narrower band gaps (normally <2.4 eV[42]), giving them with
wider spectral absorptions[43] and more efficient solar
spectrum utilizations (>20% of solar spectrum
utilization[25]). Such an efficient solar spectrum utilization is
the advantage of 2D TMDs over g-C3N4as active photo-
catalysts in photocatalysis.
Undeniably, from the current perspective, graphene (as
cocatalysts) and g-C3N4(as active photocatalysts) are more
popular and mature in photocatalysis. However, 2D TMDs
are promising photocatalytic materials and are expected to
become the next research hotspot, due to their diversity and
effective spectral utilization.
3. Challenges and Strategies
Although the photocatalysis of 2D TMDs is promising, it
still faces numerous challenges. The corresponding design
considerations and engineering strategies to address these
challenges are needed (Supporting Information Table 4).
3.1. Challenges
Semiconducting phase 2D TMDs as active photocatalysts
face the challenges of catalytically inert basal plane,[120]
sluggish carrier dynamics (originating from their high
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (15 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
electrical resistance which hinders charge transfer),[96] and
the mutual compromise of light spectrum harvesting range
versus redox potentials. Several engineering strategies,
including the engineering of edge site, phase, doping and
Figure 10. Well-placed energy levels of 2D TMDs and the redox potentials of some common photocatalytic reactions. Top panel: valence band (VB)
and conduction band (CB) positions of TMD monolayers, as well as the redox potentials of some common photocatalytic reactions at pH =0.
Bottom panel: a summary table, reflecting whether various TMD monolayers meet the thermodynamic requirements of different reactions. The
solid circle represents meeting the thermodynamic requirements of the specified reaction. The blank represents not meeting the thermodynamic
requirements of the specified reaction. SHE: standard hydrogen electrode. The data of the VB and CB positions in the top panel are collected from
ref. .[23].
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (16 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 11. Advantages of 2D TMDs as photosensitizers: quantum confinement, short transmission distance of carriers, large surface-area-to-
volume ratio, and an ideal platform for fundamental research. a) Angle-resolved photoemission spectroscopic (ARPES) spectra of MoSe2with one
layer, two layers, three layers, and eight layers along the Γ–K direction. Reproduced with permission from ref. [112]. Copyright 2014, Nature
Publishing Group. b) Schematic illustration of short transmission distance of carriers in 2D TMDs. c) Schematic illustration of large surface-area-
to-volume ratio of 2D TMDs. d) An example of 2D TMDs as ideal platform for fundamental research. Left panel: a schematic illustration of the
theoretically predicted band alignment and carrier transfer direction at the heterostructure composed of MoS2monolayer and WS2monolayer.
Right panel: a schematic illustration of the formed heterostructure composed of MoS2monolayer and WS2monolayer. Reproduced with permission
from ref. [108]. Copyright 2014, Nature Publishing Group.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (17 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
vacancy, as well as interface have been applied to deal with
these issues and are introduced below.
Additionally, 2D TMDs are prone to agglomeration and
chemical changes in the storage process due to their high
surface energy and rich active sites, lacking long-term
stability and durability. It is feasible to cope with this
challenge by storing 2D TMDs in the form of suspension
solution, along with stabilizing them via rheological
refinements,[121] or covalent functionalization.[122] Besides,
agglomeration of 2D TMDs also occurs frequently during
catalytic reactions. The effective way to solve this problem is
to grow 2D TMDs on the substrate.
3.2. Edge Site Engineering
Semiconducting phase 2D TMDs hold catalytically inert
basal planes,[120] and their catalytic active centers are often
located at the edges, where rich unsaturated atoms are
exposed.[29] Therefore, creating well-exposed active edge
sites (named edge site engineering) is a reasonable approach
to enhance the photocatalytic activity of 2D TMDs.
Using CVD method to construct vertically grown 2D
TMDs on the substrate has proved to be an effective
strategy for creating well-exposed active edge sites (Fig-
ure 12a).[25,29, 31] Such 2D MoS2is ~ 15 times as effective as
bulk MoS2in photocatalytic sterilization (log inactivation
efficiency of the indicator bacteria).[25] Besides, such 2D
ReS2also shows boosted photocatalytic performance for H2
production[29] and sterilization.[31] Alternatively, introducing
pores on the surface of 2D TMDs by a ball-milling method
is also a powerful tool for enriching boundary and edge sites
(Figure 12b).[120] Recent study shows that such MoS2with
rich pores (or said with fully activated basal planes) displays
excellent photocatalytic performance for H2evolution.[120] In
addition, there are also some potentially feasible strategies
for creating well-exposed active edge sites to boost their
photocatalytic performance, such as building a spiral
screwed structure 2D TMDs (Figure 12c).[123]
It is clear that creating well-exposed active edge sites
(edge site engineering) is a viable way to boost the photo-
catalytic activity of 2D TMDs.[25,29,31, 120] Still, this strategy is
in its infancy and its implementation often requires harsh
conditions (such as the high temperature and high vacuum
conditions[25,29, 31]). Continuous efforts should be devoted to
this field (edge site engineering of 2D TMDs for photo-
catalysis), with more emphasis on clever structural design
and easy edge site synthesis.
3.3. Phase Engineering
Previous theoretical[38,124] and experimental studies[125, 126]
demonstrated the diverse properties of 2D TMDs with
different phases. For example, 2H-MoS2monolayer is a
semiconductor with catalytically inert basal plane[120] and
sluggish carrier dynamics due to high electrical resistance;[96]
1T-MoS2monolayer holds zero band gap, showing metallic
nature, catalytically active basal plane, and high
conductivity;[97] 1T-MoS2possesses only a tiny band gap,
being a quasi-metallic phase and sharing properties similar
to those of the metallic 1T phase.[97]
Therefore, theoretically, introducing metallic/quasi-met-
allic phase 2D TMDs (as cocatalysts) into semiconducting
phase 2D TMDs (as photosensitizers), named phase engi-
neering, is expected to achieve complementary properties
and synergetic catalytic effects, thereby boosting the photo-
catalytic activity of semiconducting phase 2D TMDs.
Among them, metallic phase TMDs, acting as cocatalysts,
can accelerate the carrier dynamics and enrich the catalytic
sites. Besides, at the formed in-plane interface, photo-
induced electrons can quickly migrate from semiconducting
phase to metallic phase, promoting the separation of photo-
generated electron and hole and improving the quantum
efficiency (Figure 12d).
The feasibility of phase engineering to improve photo-
catalytic activity has been verified experimentally.[97,111, 126, 127]
For example, 2H-1T-1Tmultiphasic MoS2nanosheets[97]
and 2H/1T mixed-phase Mo1xWxS2nanosheets[111] show
boosting photocatalytic performance for H2production from
proton reduction, and ammonia (NH3) synthesis from nitro-
gen (N2) reduction, respectively.
The specific implementation methods of phase engineer-
ing include direct synthesis of 2D TMDs with multiphase by
wet chemical method (e.g., hydrothermal method),[111] or
controlled synthesis of 2D TMDs with multiphase by partial
phase transition.[97,126] The ways to realize partial phase
transition (from 2H to 1T) include lithium intercalation[97]
and supercritical CO2inducing.[126]
Undoubtedly, phase engineering is effective for boosting
the photocatalytic performance of 2D TMDs,[97,111, 126] but its
implementation faces two substantive challenges. First, the
metal phase 2D TMDs (for example 1T/1Tphase MoS2) are
normally metastable,[66,128] which are easy to occur phase
transition[66,128] when placed in the ambiance or at the
catalytic process. This will denature the catalysts and bring
uncertainty to their photocatalytic researches. Second, it is
not easy to precisely control the synthesis of multiphase 2D
TMDs with accurate and on-demand phase ratio, either
direct synthesis (via hydrothermal method[111]) or fabrication
through partial phase transition (via lithium intercalation[97]
or supercritical CO2inducing[126]). Therefore, efforts should
be paid to address these two main bottlenecks in future
researches. For example, improve the stability of metal
phase 2D TMDs via surface modifications to solve the
challenge of phase transition. In parallel, enhance the
controllability of phase transition to synthesize catalyst with
accurate phase ratio. Of note, electrochemical lithium-ion
intercalation-based exfoliation is a promising strategy, which
is expected to achieve precise regulation of phase transition
by controlling the current or cut-off voltage.[70]
3.4. Doping and Vacancy Engineering
Introducing heteroatom (named doping engineering, divided
into nonmetallic element doping and metal element doping,
or substitute doping and interstitial doping, Figure 13a) into
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (18 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 12. Edge site engineering and phase engineering of 2D TMDs for boosting photocatalytic performance. a) A effective strategy for creating
well-exposed active edge sites of 2D TMDs: constructing vertically grown 2D TMDs on the substrate. Left panel: a schematic illustration of the
constructed vertically grown 2D MoS2on the substrate and their photocatalytic sterilization application. Right panel: a schematic illustration of the
well-exposed active edge sites. Left panel is reproduced with permission from ref. [25]. Copyright 2016, Nature Publishing Group. b) A effective
strategy for creating well-exposed active edge sites of 2D TMDs: introducing pores on the surface of 2D TMDs. c) A effective strategy for creating
well-exposed active edge sites of 2D TMDs: building a spiral screwed structure of 2D TMDs. Reproduced with permission from ref. [123]. Copyright
2022, John Wiley & Sons. d) Phase engineering of 2D TMDs. Left panel: a schematic illustration of the constructed heterophase interface
composed of 1T/1TTMDs and 2H TMDs. Right panel: a schematic illustration of the electron transfer at the heterophase interface composed of
1TMoS2and 2H MoS2.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (19 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
semiconducting 2D TMDs could bring various attractive
advantages for photocatalytic application.[129–133] Theoretical
calculation shows that the generated defect levels caused by
the introduction of heteroatoms can adjust the band gaps of
2D TMDs, thereby tuning their redox ability.[129, 130] Subse-
quent experimental studies confirmed this.[131–133] For in-
stance, carbon-doped SnS2holds a larger band gap value of
2.54 eV than that of its pristine counterparts (2.43 eV)
(Figure 13b).[131] Besides, introducing heteroatom can also
effectively modify their optoelectronic properties (e.g.,
effectively strengthening the light absorption and prolonging
carrier lifetime, Figure 13c), thus boosting their photocata-
lytic performance (Figure 13d).
Creating vacancy (named vacancy engineering, divided
into metal vacancy and chalcogen vacancy, Figure 14a) is
also a prevalent strategy to unlock the large potential of 2D
TMDs for photocatalytic property improvement.[134–137] This
is because it can tune the electronic structure (Figure 14b)
and neighboring atomic arrangement of 2D TMDs to boost
the intrinsic activity.[134,135] Besides, acting as a carrier trap, it
is also beneficial for the separation of photogenerated
electrons and holes, prolonging their lifetimes, and thereby
improving the photocatalytic performance (Fig-
ure 14c).[136,137]
Notably, the degree of modification (e.g., the concen-
tration of doping and vacancy) needs to be moderate.
Because over modification may lead to the opposite results.
For example, excess vacancies may become the center of
carrier recombination and inhibit photocatalysis.[138]
Doping and vacancy engineering are indeed useful for
the enhancement of photocatalytic performance of 2D
TMDs. Still, the exploitation of such strategies in 2D TMDs-
based photocatalysis[131–133,136, 137] is just starting to emerge.
Although these strategies enjoy great popularity in the field
of catalysis.[139] Additional researches toward this topic
(doping and vacancy engineering in 2D TMDs-based photo-
catalysis) should be carried out in the future. This is a
research field full of opportunities. For example, modern
technical toolkits for doping and vacancy engineering are
constantly enriched and improved (such as plasma induced
doping,[140] laser-assisted doping,[141] neutron-transmutation
doping,[142] electron beam lithography inducing vacancy[143]),
Figure 13. Doping engineering of 2D TMDs for boosting photocatalytic performance. a) Doping categories. b) Band edge position of SnS2and
CSnS2. Data collected from ref. [131]. c) Optical properties of SnS2and CSnS2. Top panel: UV/Vis diffuse reflectance spectra of SnS2and CSnS2.
Bottom panel: normalized photoluminescence (PL) spectra of SnS2and CSnS2. d) Photocatalytic CO2reduction (the formation of acetaldehyde)
activity of SnS2and CSnS2. Panel c, d are reproduced with permission from ref. [131]. Copyright 2018, Nature Publishing Group.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (20 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 14. Vacancy engineering of 2D TMDs for boosting photocatalytic performance. a) Vacancy categories. Left and middle panel: top view and
side view of the atomic structure model of TMD monolayer with vacancy. Right panel: a scanning transmission electron microscope (STEM) image
of MoS2with S vacancy. Right panel is reproduced with permission from ref. [135]. Copyright 2016, Nature Publishing Group. b) The band
structures of perfect monolayer MoS2, monolayer MoS2with Mo-vacancy, and with S-vacancy. Reproduced with permission from ref. [134].
Copyright 2014, Elsevier. c) Band edge position (left panel), optical properties (time-resolved photoluminescence spectra, middle panel), and
photocatalytic performance (degradation of methylene blue, right panel) of MoS2and MoS2with S-vacancy. NHE: normal hydrogen electrode.
Reproduced with permission from ref. [136]. Copyright 2017, Elsevier.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (21 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
which provide more sophisticated techniques to introducing
heteroatoms and vacancies into photocatalysts. Applying
these state-of-the-art techniques in 2D TMDs to boost their
photocatalytic performance is a promising research topic
and is worth making efforts.
3.5. Interface Engineering
When the interface is formed, the carrier migration will
show strong directivity.[43,144, 145] Such strong directivity con-
tributes to the effective separation of photogenerated
carriers, thus prolonging their lifetime and improving optical
quantum efficiency.[146] Therefore, a plethora of opportuni-
ties for boosting photocatalytic performance appear when
we start constructing a 2D TMDs-based heterojunction
interface.[43]
A well-designed 2D TMDs-based heterojunction inter-
face should be energy band matched, dimension (geometry)
matched, and chemically compatible.[43]
Theoretically, several 2D TMDs-based heterojunction
interfaces can be formed, according to the band alignments
and directivity of electron transfer, including type I (strad-
dling gap), type II (alternate gap), Z-scheme (alternate gap),
type III (broken gap), and Schottky junction (or ohmic
junction) (Figure 15a).[43] Among them, type II,[100,147, 148] Z-
scheme,[32,36, 149–151] and Schottky junction[28, 33, 152–155] (or ohmic
junction[156]) are energy band well matched. At these
heterojunction interface (type II, Z-scheme, and Schottky
junction or ohmic junction), photoinduced electrons or holes
are transfer from one building block to another and
enriched in different building block (as shown in Fig-
ure 15a).[157] Such a feature is conducive to the separation of
carriers. Besides, it also benefits the occurrence of photo-
catalytic oxidation and reduction reactions in different
places, thus improving the reaction efficiency.[157]
2D TMDs-based type II and Z-scheme heterojunctions
are formed between two semiconductors (semiconducting
phase 2D TMDs and another semiconductor). For Z-scheme
heterojunction, the redox potential of photoexcited elec-
tron–hole pairs are increased, resulting in higher redox
capability, but for type II heterojunctions, the opposite is
true (as shown in Figure 15a). Therefore, constructing 2d
TMDs-based Z-scheme heterojunction is more conducive to
the improvement of photocatalytic performance.
2D TMDs-based Schottky junction is composed of semi-
conducting phase 2D TMDs and near-zero band-gap crystal
(e.g., metals[28,33, 152, 153] and carbon materials[154, 155]). Near-
zero band-gap crystal normally hold high electrical con-
ductivity and can act as electron traps as well as provide rich
active reaction sites, thereby enhancing the photocatalytic
performance of 2D TMDs. It is worth noting that when
stacking 2D TMDs and plasmonic metals (e.g., Au,[152]
Ag,[153] Pd,[28] Cu[33]) together, some other fascinating syner-
getic effects appear. On the one hand, plasmonic metals can
enhance the light-absorption capability of 2D TMDs
through localized surface plasmon resonance (LSPR).[152]
LSPR can induce a large electric dipole (left panel of
Figure 15b), and thus form an intense electric field nearby
the plasmonic nanoparticles, called near-field enhancement
phenomenon.[43] The near-field enhancement phenomenon
leads to a strong local optical field and thereby enhances the
light absorption of 2D TMDs.[152] On the other hand, owing
to the plasmon resonance effects, plasmonic nanoparticles
can create high energy (“hot”) electrons, which can migrate
to the conduction band of 2D TMDs via the formation of a
Schottky junction (right panel of Figure 15b).[33] Such hot
electrons can drive the photocatalytic reaction.[33]
Beyond the energy band matching (mentioned above),
dimension (geometry) matching is another key factor for
building an efficient heterojunction interface. A well-
matched dimension (geometry) is beneficial for the forma-
tion of dense interface junctions between the two materials,
providing a high-speed channel for carrier migration.[43]
Figure 15c–f display some representative examples of well-
designed 2D TMDs-based heterojunction with 0D–2D,[28]
1D–2D,[158] 2D–2D,[16] and 3D–2D[159] matched dimension.
Among them, the 2D–2D heterojunction interface is the
famous one,[160–162] and can be divided into in-plane hetero-
junction interface (exemplified by in-plane MoS2
WS2
heterojunction interface,[163] Figure 15g) and vertical hetero-
junction interface (exemplified by PtSe2/Pt vertical hetero-
structure interface,[164] Figure 15h).
Additionally, the chemical compatibility of the two
building blocks (2D TMDs and another material) should
also be emphasized. In terms of 2D TMDs, metal chloride
(such as CdS,[158,165–169] ZnIn2S4,[17, 36, 170] and TMD
themselves[163,171, 172]) would be most chemically compatible
with them to make functional hybrid photocatalysts. This is
because they have the same or familiar non-metallic
elements (for example, the S element in MoS2and CdS),
which makes it easy for them to build heterostructures with
good interfaces by epitaxial growth.[158, 159,165, 173] At such an
interface, chemical bonds (for example, MoSCu bond
formed between MoS2and CuS interface[158]) are often
formed, connecting the two building blocks, which will
facilitate the rapid transfer of carriers and improve optical
quantum efficiency.[158,165]
Despite a plethora of opportunities appear when stack-
ing 2D TMDs with other units into a heterostructure, we
have not yet obtained an ideal heterostructure for practical
photocatalytic applications. It has also been challenging to
directly characterize the specific path of electron migration
in situ or operando at such heterojunction interfaces. It is
not yet possible to control the charge transfers and built-in
electric fields with atomic precision at such heterojunction
interfaces. These are both challenges and opportunities, and
researchers in the future are worth studying.
3.6. Analysis
Rich engineering strategies have been established to deal
with the challenges of intrinsic 2D TMDs faced for photo-
catalysis. These strategies have their own merits and bottle-
necks (Supporting Information Table 4), but none of them
are yet perfect. Of the strategies reviewed here, the most
readily and widely applicable one is interface engineering
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (22 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 15. Interface engineering. a) Type of heterojunction interface, classified according to the band alignments and directivity of electron transfer,
including type I (straddling gap), type II (alternate gap), Z-scheme (alternate gap), type III (broken gap), Schottky junction, and ohmic junction.
Note that Schottky junction is formed by the contact between a semiconductor (n-Type) with a lower work function and a conductor with a higher
work function, and form upward band bending. The ohmic junction is formed by the contact between a semiconductor with a higher work function
and a conductor with a lower work function, resulting in the downward bending of the energy band. b) Localized surface plasmon resonance and
hot electrons transfer at interface. Left panel: schematic illustration of localized surface plasmon resonance for a plasmonic nanoparticle. Right
panel: schematic illustration of the formation of hot electrons and interfacial charge transfer at the interface of plasmonic nanoparticles and 2D
TMDs. Ef: Fermi level. φB: Schottky barrier. c) TEM image of 0D–2D PdWS2hybrid. Reproduced with permission from ref. [28]. Copyright 2017,
American Chemical Society. d) TEM image of 1D–2D CdSMoS2hybrid. Reproduced with permission from ref. [158]. Copyright 2020, American
Chemistry Society. e) TEM image of 2D–2D Bi12O17Cl2
MoS2hybrid. Reproduced with permission from ref. [16]. Copyright 2016, Nature Publishing
Group. f) TEM image of 3D–2D TEM image of Cu2χS0.8Se0.2
MoS2hybrid. Reproduced with permission from ref. [159]. Copyright 2021, John Wiley
& Sons. g) High resolution annular-dark-field (ADF) scanning transmission electron microscopy (STEM) image and atomic model of the interface
of WS2
MoS2in-plane heterostructure. Reproduced with permission from ref. [163]. Copyright 2021, Elsevier. h) Atomic-resolution STEM image
and atomic model of the interface of PtSe2/Pt(111) vertical heterostructure. Reproduced with permission from ref. [164]. Copyright 2015, American
Chemical Society.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (23 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
(or named hybrid engineering, including, building hetero-
junction or coating cocatalyst). Although this strategy also
faces numbers bottlenecks as mentioned above, it is indeed
the most-friendly one for novices, owing to its easy
operation and obvious effect.
Notably, it is sometimes impossible to achieve a
satisfactory improvement effect by using only one strategy.
At this time, two or more strategies need to be applied
cooperatively. For example, creating well-exposed active
edge sites (edge site engineering) to deal with the challenge
of catalytically inert basal plane of semiconducting 2D
TMDs, along with coating conductors (such as Pt and
graphene) on 2D TMDs to cope with the challenge of
sluggish carrier dynamics.
4. Outlook
Undoubtedly, 2D TMDs, whether as cocatalysts or active
photocatalysts (light-harvesting materials), point to a pro-
ductive past, a moving present, and a bright future. Yet,
challenges remain, but opportunities abound. Here, we look
ahead the appealing directions of 2D TMDs for photo-
catalysis (Figure 16).
4.1. Extending Members
The family members of 2D TMDs used for photocatalysis
are continuously growing, which started with MoS2[10, 25]
Figure 16. 2D TMDs for photocatalysis: looking ahead. Panel of ‘Extending members’ shows that the family members of 2D TMDs using for
photocatalysis are continuously growing. Panel of ‘Improving existing strategies’ shows an example of existing strategy (edge site engineering) that
should be improved in the future for boosting photocatalytic performance of 2D TMDs. Reproduced with permission from ref. [179]. Copyright
2022, Nature Publishing Group. Panel of ‘Developing new strategies’ shows an example of new strategy (metal intercalation) that can be developed
in the future for boosting photocatalytic performance of 2D TMDs. Panel of ‘Understanding mechanisms’ shows an available in situ
characterization (in situ liquid cell transmission electron microscope) technology for understanding the photocatalytic mechanisms of 2D TMDs.
Panel of ‘Developing device’ shows a notable device example: flexible 100 cm2perovskite-BiVO4device. Reproduced with permission from
ref. [180]. Copyright 2022, Nature Publishing Group. Panel of ‘lab to fab transition’ shows a factory for conversion of solar energy.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (24 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
(MoSe2[17,36]) and WS2[11, 24] (WSe2[34, 37]), then ReS2[29, 149] and
SnS2,[131] have now evolved to PtS2,[174] PtSe2,[164] and beyond.
It looks like this process is only beginning, more members of
2D TMDs (e.g., ZrS2,[175] HfS2,[176] PdS2[177]) may pour into
photocatalysis in the near future predicted by high through-
put calculations. In the future, these theoretically feasible
TMDs need to be verified experimentally. Of note, although
ReS2has exciting physicochemical properties for photo-
catalytic applications, its metal element (Re) is not earth-
abundant (one of the rarest elements in the Earth’s
crust).[178] Therefore, its practical application prospect will
be limited by economic feasibility.
4.2. Improving Existing Strategies
Big promotion of photocatalytic performance achieves when
modifying 2D TMDs by the engineering of edge
site,[25,29, 31, 120] phases,[97, 111, 126] doping,[131–133] vacancy,[136, 137]
and interface. Still, compared with electrocatalysis, most of
the engineering technologies for photocatalysis of 2D TMDs
are in their infancy. In the future, continuing to dig deeper
into these engineering categories is essential. For instance,
fabricating ladder 2D TMDs to enrich edge sites;[179]
performing local phase transition to build periodic hetero-
phase interface;[125] applying laser patterning, thermal etch-
ing, and endo-epitaxial growth process to construct mosaic
heterostructures.[181]
4.3. Developing New Strategies
In parallel, developing new engineering strategies for photo-
catalysis using 2D TMDs also requires passion and effort.
For example, introducing intercalated metals into the
interlayers of 2D TMDs to boost their photocatalytic
performance by synergistic catalysis and confined catalysis
effect (confined catalysis: a new catalysis concept, to provide
a constrained environment for the catalytic reaction system
at the nanoscale, so as to achieve precise regulation of
catalytic performance and make catalysis “fast and
good”[182]).
4.4. Understanding Mechanisms
Understanding the photocatalytic mechanisms of 2D TMDs
should not be ignored. Clarifying some natures of science is
also necessary, such as who is the real active site? Whether
the phase of TMDs changes during the catalytic reaction?
How do electrons migrate at the interface? State-of-the-art
in situ characterization technologies offer powerful tools for
exploring these mechanisms and natures deeply in real-time.
These available in situ techniques include but are not limited
to, in situ liquid cell transmission electron microscope
(TEM), in situ X-ray absorption spectroscopy (XAS), and in
situ Raman. For example, in situ liquid cell TEM can
provide real-time visual detection of catalyst surface at the
atomic scale,[183,184] may open the black box of real photo-
catalytic sites of 2D TMDs-based materials. Beyond
advanced in situ characterization technologies, ultrafast
carrier detection (e.g., femtosecond pump-probe
spectroscopy[108]) and high throughput first-principles calcu-
lations are also effective tools for photocatalytic mechanism
research.
4.5. Developing Device
Methods and protocols should be established to develop 2D
TMDs-based photocatalytic devices, realizing an integration
of photo(electric)catalytic reaction, product collection and
real-time monitoring. Recently developed perovskite-BiVO4
devices for scalable solar fuel production can provide a
reference.[180] The development of modern micro-nano-
manufacturing technology (e.g., photolithography, physical
vapor deposition), solution-based deposition technology
(e.g., inkjet-printing, industrial roll-to-roll coating), and the
continuous development of functional materials (e.g., flexi-
ble material, superconductor) provide guarantee for the
manufacturing of such devices.
4.6. Lab to Fab Transition
The ultimate goal of photocatalytic technology is to serve
production and life. Realizing the transition of photo-
catalytic application of 2D TMDs-based photocatalysts from
lab to fab is therefore essential. Yet, there is a long way to
make this transition. Steps to be completed in turn in the
future for achieving such a transition include: enlarging the
developed devices, and/or assembling them in series or in
parallel, followed by placing them in compatible industrial
systems. During these steps, the evaluation of security
factors, economic factors, and market factors should not be
ignored. The recently realized solar-to-hydrogen farm is
such a pioneer, which is worth learning from.[185]
4.7. Others
In addition to seizing the opportunities mentioned above, it
is also meaningful to devote efforts to the following issues.
Building a complete and environment-friendly system for
the production, storage and recovery of 2D TMDs-based
photocatalysts should be emphasized. Because, most of the
2D TMDs, whether prepared or discarded, pollute the
environment. Constructing efficiency accreditation and
standardized testing protocols for 2D TMDs-based photo-
catalysts toward various photocatalytic reaction is urgent.
Currently, the experimental data of each research group
lacks the credibility of horizontal comparisons.[61] Because
their experimental conditions (such as the light condition,
ambient temperature, and pH value of the reaction solution)
are often different. This normally leads to the accumulation
of unverifiable and often misleading data, hindering the
advance in the research field.[61] Therefore, the efficiency
accreditation and standardized testing protocol are required.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (25 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
In all, opportunities and challenges coexist in this
fascinating field. We anticipate that 2D TMDs becoming a
new material-central for photocatalysis in the coming
decades, and photo(electro)catalysis becoming one of the
technologies central for solving the energy and environ-
mental crisis in the coming era.
Acknowledgements
Z.Y. Zeng thanks ECS scheme (CityU9048163) from RGC
in Hong Kong and the Basic Research Project from
Shenzhen Science and Technology Innovation Committee in
Shenzhen, China (No. JCYJ20210324134012034). Q. Wang
thanks JSPS Leading Initiative for Excellent Young Re-
searchers program, the JST Fusion Oriented Research for
disruptive Science and Technology program, and the JSPS
Grant-in-Aid for Young Scientists (Start-up) (No.
21 K20485). Q.Y. Lu thanks Natural sciences and engineer-
ing research council of Canada (NSERC) Discovery Grant
and Alberta Innovates Advance Program-NSERC Alliance
Grant.
Conflict of Interest
The authors declare no conflict of interest.
Keywords: 2D TMDs ·Active Photocatalyst ·Cocatalyst ·
MoS2·Photocatalysis
[1] S. Manzeli, D. Ovchinnikov, D. Pasquier, O. V. Yazyev, A.
Kis, Nat. Rev. Mater. 2017,2, 17033.
[2] M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh, H.
Zhang, Nat. Chem. 2013,5, 263.
[3] K. F. Mak, J. Shan, Nat. Photonics 2016,10, 216.
[4] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, M. S.
Strano, Nat. Nanotechnol. 2012,7, 699.
[5] X. Xu, W. Yao, D. Xiao, T. F. Heinz, Nat. Phys. 2014,10, 343.
[6] D. Voiry, H. S. Shin, K. P. Loh, M. Chhowalla, Nat. Chem.
Rev. 2018,2, 0105.
[7] D. Voiry, J. Yang, M. Chhowalla, Adv. Mater. 2016,28, 6197.
[8] Q. Lu, Y. Yu, Q. Ma, B. Chen, H. Zhang, Adv. Mater. 2016,
28, 1917.
[9] Z. Liang, R. Shen, Y. H. Ng, P. Zhang, Q. Xiang, X. Li, J.
Mater. Sci. Technol. 2020,56, 89.
[10] X. Zong, H. Yan, G. Wu, G. Ma, F. Wen, L. Wang, C. Li, J.
Am. Chem. Soc. 2008,130, 7176.
[11] X. Zong, J. Han, G. Ma, H. Yan, G. Wu, C. Li, J. Phys. Chem.
C2011,115, 12202.
[12] Q. Xiang, J. Yu, M. Jaroniec, J. Am. Chem. Soc. 2012,134,
6575.
[13] W. Zhou, Z. Yin, Y. Du, X. Huang, Z. Zeng, Z. Fan, H. Liu,
J. Wang, H. Zhang, Small 2013,9, 140.
[14] K. Chang, Z. Mei, T. Wang, Q. Kang, S. Ouyang, J. Ye, ACS
Nano 2014,8, 7078.
[15] J. Chen, X.-J. Wu, L. Yin, B. Li, X. Hong, Z. Fan, B. Chen, C.
Xue, H. Zhang, Angew. Chem. Int. Ed. 2015,54, 1210; Angew.
Chem. 2015,127, 1226.
[16] J. Li, G. Zhan, Y. Yu, L. Zhang, Nat. Commun. 2016,7,
11480.
[17] M.-Q. Yang, Y.-J. Xu, W. Lu, K. Zeng, H. Zhu, Q.-H. Xu,
G. W. Ho, Nat. Commun. 2017,8, 14224.
[18] S. Zhang, X. Liu, C. Liu, S. Luo, L. Wang, T. Cai, Y. Zeng, J.
Yuan, W. Dong, Y. Pei, Y. Liu, ACS Nano 2018,12, 751.
[19] S. Xie, Z. Shen, J. Deng, P. Guo, Q. Zhang, H. Zhang, C. Ma,
Z. Jiang, J. Cheng, D. Deng, Y. Wang, Nat. Commun. 2018,9,
1181.
[20] H. He, J. Lin, W. Fu, X. Wang, H. Wang, Q. Zeng, Q. Gu, Y.
Li, C. Yan, B. K. Tay, C. Xue, X. Hu, S. T. Pantelides, W.
Zhou, Z. Liu, Adv. Energy Mater. 2016,6, 1600464.
[21] X. Wang, B. Chen, D. Yan, X. Zhao, C. Wang, E. Liu, N.
Zhao, F. He, ACS Appl. Mater. Interfaces 2019,11, 23144.
[22] L. Wang, X. Liu, J. Luo, X. Duan, J. Crittenden, C. Liu, S.
Zhang, Y. Pei, Y. Zeng, X. Duan, Angew. Chem. Int. Ed.
2017,56, 7610; Angew. Chem. 2017,129, 7718.
[23] H. L. Zhuang, R. G. Hennig, J. Phys. Chem. C 2013,117,
20440.
[24] Y. Sang, Z. Zhao, M. Zhao, P. Hao, Y. Leng, H. Liu, Adv.
Mater. 2015,27, 363.
[25] C. Liu, D. Kong, P.-C. Hsu, H. Yuan, H.-W. Lee, Y. Liu, H.
Wang, S. Wang, K. Yan, D. Lin, P. A. Maraccini, K. M.
Parker, A. B. Boehm, Y. Cui, Nat. Nanotechnol. 2016,11,
1098.
[26] F. Raza, J. H. Park, H.-R. Lee, H.-I. Kim, S.-J. Jeon, J.-H.
Kim, ACS Catal. 2016,6, 2754.
[27] S. Sun, X. Li, W. Wang, L. Zhang, X. Sun, Appl. Catal. B
2017,200, 323.
[28] F. Raza, D. Yim, J. H. Park, H.-I. Kim, S.-J. Jeon, J.-H. Kim,
J. Am. Chem. Soc. 2017,139, 14767.
[29] Q. Zhang, W. Wang, J. Zhang, X. Zhu, Q. Zhang, Y. Zhang,
Z. Ren, S. Song, J. Wang, Z. Ying, R. Wang, X. Qiu, T. Peng,
L. Fu, Adv. Mater. 2018,30, 1707123.
[30] J. R. Dunklin, H. Zhang, Y. Yang, J. Liu, J. van de Lagemaat,
ACS Energy Lett. 2018,3, 2223.
[31] D. Ghoshal, A. Yoshimura, T. Gupta, A. House, S. Basu, Y.
Chen, T. Wang, Y. Yang, W. Shou, J. A. Hachtel, J. C.
Idrobo, T.-M. Lu, S. Basuray, V. Meunier, S.-F. Shi, N.
Koratkar, Adv. Funct. Mater. 2018,28, 1801286.
[32] Y. Wang, Z. Zhang, L. Zhang, Z. Luo, J. Shen, H. Lin, J.
Long, J. C. S. Wu, X. Fu, X. Wang, C. Li, J. Am. Chem. Soc.
2018,140, 14595.
[33] X. Xu, F. Luo, W. Tang, J. Hu, H. Zeng, Y. Zhou, Adv. Funct.
Mater. 2018,28, 1804055.
[34] Y. Wang, S. Zhao, Y. Wang, D. A. Laleyan, Y. Wu, B.
Ouyang, P. Ou, J. Song, Z. Mi, Nano Energy 2018,51, 54.
[35] Y. Pang, M. N. Uddin, W. Chen, S. Javaid, E. Barker, Y. Li,
A. Suvorova, M. Saunders, Z. Yin, G. Jia, Adv. Mater. 2019,
31, 1905540.
[36] X. Wang, X. Wang, J. Huang, S. Li, A. Meng, Z. Li, Nat.
Commun. 2021,12, 4112.
[37] M. Qorbani, A. Sabbah, Y.-R. Lai, S. Kholimatussadiah, S.
Quadir, C.-Y. Huang, I. Shown, Y.-F. Huang, M. Hayashi, K.-
H. Chen, L.-C. Chen, Nat. Commun. 2022,13, 1256.
[38] Y. Chen, Z. Lai, X. Zhang, Z. Fan, Q. He, C. Tan, H. Zhang,
Nat. Chem. Rev. 2020,4, 243.
[39] X. Qian, J. Liu, L. Fu, J. Li, Science 2014,346, 1344.
[40] J. Kang, S. Tongay, J. Zhou, J. Li, J. Wu, Appl. Phys. Lett.
2013,102, 012111.
[41] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim,
G. Galli, F. Wang, Nano Lett. 2010,10, 1271.
[42] H. M. Hill, A. F. Rigosi, K. T. Rim, G. W. Flynn, T. F. Heinz,
Nano Lett. 2016,16, 4831.
[43] X. Gan, D. Lei, R. Ye, H. Zhao, K.-Y. Wong, Nano Res.
2021,14, 2003.
[44] X. Li, C. Wang, J. Tang, Nat. Rev. Mater. 2022,7, 617.
[45] Q. Wang, C. Pornrungroj, S. Linley, E. Reisner, Nat. Energy
2022,7, 13.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (26 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[46] E. Pastor, M. Sachs, S. Selim, J. R. Durrant, A. A. Bakulin, A.
Walsh, Nat. Rev. Mater. 2022,7, 503.
[47] R. Yang, Y. Fan, R. Ye, Y. Tang, X. Cao, Z. Yin, Z. Zeng,
Adv. Mater. 2021,33, 2004862.
[48] R. Yang, Z. Guo, L. Cai, R. Zhu, Y. Fan, Y. Zhang, P. Han,
W. Zhang, X. Zhu, Q. Zhao, Z. Zhu, C. K. Chan, Z. Zeng,
Small 2021,17, 2103052.
[49] S. Chen, T. Takata, K. Domen, Nat. Rev. Mater. 2017,2,
17050.
[50] X. Li, J. Yu, M. Jaroniec, X. Chen, Chem. Rev. 2019,119,
3962.
[51] R. Yang, Y. Zhang, Y. Fan, R. Wang, R. Zhu, Y. Tang, Z.
Yin, Z. Zeng, Chem. Eng. J. 2022,428, 131145.
[52] R. Yang, L. Mei, Y. Fan, Q. Zhang, R. Zhu, R. Amal, Z. Yin,
Z. Zeng, Small Methods 2021,5, 2100887.
[53] Z. Liang, R. Shen, Y. H. Ng, Y. Fu, T. Ma, P. Zhang, Y. Li, X.
Li, Chem Catal. 2022,2, 2157.
[54] A. Kudo, Y. Miseki, Chem. Soc. Rev. 2009,38, 253.
[55] X. Li, J. Yu, J. Low, Y. Fang, J. Xiao, X. Chen, J. Mater.
Chem. A 2015,3, 2485.
[56] X. Li, J. Yu, M. Jaroniec, Chem. Soc. Rev. 2016,45, 2603.
[57] X. Li, J. Wen, J. Low, Y. Fang, J. Yu, Sci. China Mater. 2014,
57, 70.
[58] M. Xiao, Z. Wang, M. Lyu, B. Luo, S. Wang, G. Liu, H.-M.
Cheng, L. Wang, Adv. Mater. 2019,31, 1801369.
[59] G. Liu, C. Zhen, Y. Kang, L. Wang, H.-M. Cheng, Chem. Soc.
Rev. 2018,47, 6410.
[60] Q. Wang, Z. Pan, Nano Res. 2022,15, 10090.
[61] Z. Wang, T. Hisatomi, R. Li, K. Sayama, G. Liu, K. Domen,
C. Li, L. Wang, Joule 2021,5, 344.
[62] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V.
Khotkevich, S. V. Morozov, A. K. Geim, Proc. Natl. Acad.
Sci. USA 2005,102, 10451.
[63] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A.
Kis, Nat. Nanotechnol. 2011,6, 147.
[64] J. N. Coleman, M. Lotya, A. O’Neill, S. D. Bergin, P. J. King,
U. Khan, K. Young, A. Gaucher, S. De, R. J. Smith, I. V.
Shvets, S. K. Arora, G. Stanton, H.-Y. Kim, K. Lee, G. T.
Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Wang, J. F.
Donegan, J. C. Grunlan, G. Moriarty, A. Shmeliov, R. J.
Nicholls, J. M. Perkins, E. M. Grieveson, K. Theuwissen,
D. W. McComb, P. D. Nellist, V. Nicolosi, Science 2011,331,
568.
[65] V. Nicolosi, M. Chhowalla, M. G. Kanatzidis, M. S. Strano,
J. N. Coleman, Science 2013,340, 1226419.
[66] G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen, M.
Chhowalla, Nano Lett. 2011,11, 5111.
[67] Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey, H.
Zhang, Angew. Chem. Int. Ed. 2011,50, 11093; Angew. Chem.
2011,123, 11289.
[68] Z. Lin, Y. Liu, U. Halim, M. Ding, Y. Liu, Y. Wang, C. Jia, P.
Chen, X. Duan, C. Wang, F. Song, M. Li, C. Wan, Y. Huang,
X. Duan, Nature 2018,562, 254.
[69] J. Li, P. Song, J. Zhao, K. Vaklinova, X. Zhao, Z. Li, Z. Qiu,
Z. Wang, L. Lin, M. Zhao, T. S. Herng, Y. Zuo, W. Jonhson,
W. Yu, X. Hai, P. Lyu, H. Xu, H. Yang, C. Chen, S. J.
Pennycook, J. Ding, J. Teng, A. H. Castro Neto, K. S.
Novoselov, J. Lu, Nat. Mater. 2021,20, 181.
[70] R. Yang, L. Mei, Q. Zhang, Y. Fan, H. S. Shin, D. Voiry, Z.
Zeng, Nat. Protoc. 2022,17, 358.
[71] J. Zhou, J. Lin, X. Huang, Y. Zhou, Y. Chen, J. Xia, H.
Wang, Y. Xie, H. Yu, J. Lei, D. Wu, F. Liu, Q. Fu, Q. Zeng,
C.-H. Hsu, C. Yang, L. Lu, T. Yu, Z. Shen, H. Lin, B. I.
Yakobson, Q. Liu, K. Suenaga, G. Liu, Z. Liu, Nature 2018,
556, 355.
[72] L. Sun, G. Yuan, L. Gao, J. Yang, M. Chhowalla, M. H.
Gharahcheshmeh, K. K. Gleason, Y. S. Choi, B. H. Hong, Z.
Liu, Nat. Rev. Methods Primers 2021,1, 5.
[73] J. Zhou, C. Zhu, Y. Zhou, J. Dong, P. Li, Z. Zhang, Z. Wang,
Y.-C. Lin, J. Shi, R. Zhang, Y. Zheng, H. Yu, B. Tang, F. Liu,
L. Wang, L. Liu, G.-B. Liu, W. Hu, Y. Gao, H. Yang, W.
Gao, L. Lu, Y. Wang, K. Suenaga, G. Liu, F. Ding, Y. Yao, Z.
Liu, Nat. Mater. 2022, https://doi.org/10.1038/s41563-022-
01291-5.
[74] J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou,
X. W. Lou, Y. Xie, Adv. Mater. 2013,25, 5807.
[75] J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R.
Wang, Y. Lei, B. Pan, Y. Xie, J. Am. Chem. Soc. 2013,135,
17881.
[76] S. Jeong, D. Yoo, J.-T. Jang, M. Kim, J. Cheon, J. Am. Chem.
Soc. 2012,134, 18233.
[77] B. Mahler, V. Hoepfner, K. Liao, G. A. Ozin, J. Am. Chem.
Soc. 2014,136, 14121.
[78] Y. Huang, Y.-H. Pan, R. Yang, L.-H. Bao, L. Meng, H.-L.
Luo, Y.-Q. Cai, G.-D. Liu, W.-J. Zhao, Z. Zhou, L.-M. Wu,
Z.-L. Zhu, M. Huang, L.-W. Liu, L. Liu, P. Cheng, K.-H. Wu,
S.-B. Tian, C.-Z. Gu, Y.-G. Shi, Y.-F. Guo, Z. G. Cheng, J.-P.
Hu, L. Zhao, G.-H. Yang, E. Sutter, P. Sutter, Y.-L. Wang,
W. Ji, X.-J. Zhou, H.-J. Gao, Nat. Commun. 2020,11, 2453.
[79] H. Zhang, ACS Nano 2015,9, 9451.
[80] L. Mei, Z. Cao, T. Ying, R. Yang, H. Peng, G. Wang, L.
Zheng, Y. Chen, C. Y. Tang, D. Voiry, H. Wang, A. B.
Farimani, Z. Zeng, Adv. Mater. 2022,34, 2201416.
[81] M. M. Ugeda, A. J. Bradley, S.-F. Shi, F. H. da Jornada, Y.
Zhang, D. Y. Qiu, W. Ruan, S.-K. Mo, Z. Hussain, Z.-X.
Shen, F. Wang, S. G. Louie, M. F. Crommie, Nat. Mater. 2014,
13, 1091.
[82] J.-G. Song, J. Park, W. Lee, T. Choi, H. Jung, C. W. Lee, S.-
H. Hwang, J. M. Myoung, J.-H. Jung, S.-H. Kim, C. Lansalot-
Matras, H. Kim, ACS Nano 2013,7, 11333.
[83] M. I. Serna, S. H. Yoo, S. Moreno, Y. Xi, J. P. Oviedo, H.
Choi, H. N. Alshareef, M. J. Kim, M. Minary-Jolandan, M. A.
Quevedo-Lopez, ACS Nano 2016,10, 6054.
[84] X. Zhou, H. Hao, Y.-J. Zhang, Q. Zheng, S. Tan, J. Zhao, H.-
B. Chen, J.-J. Chen, Y. Gu, H.-Q. Yu, X.-W. Liu, Chem 2021,
7, 1626.
[85] Z. Lai, Q. He, T. H. Tran, D. V. M. Repaka, D.-D. Zhou, Y.
Sun, S. Xi, Y. Li, A. Chaturvedi, C. Tan, B. Chen, G.-H. Nam,
B. Li, C. Ling, W. Zhai, Z. Shi, D. Hu, V. Sharma, Z. Hu, Y.
Chen, Z. Zhang, Y. Yu, X. Renshaw Wang, R. V. Ramanujan,
Y. Ma, K. Hippalgaonkar, H. Zhang, Nat. Mater. 2021,20,
1113.
[86] Y. Yu, G.-H. Nam, Q. He, X.-J. Wu, K. Zhang, Z. Yang, J.
Chen, Q. Ma, M. Zhao, Z. Liu, F.-R. Ran, X. Wang, H. Li, X.
Huang, B. Li, Q. Xiong, Q. Zhang, Z. Liu, L. Gu, Y. Du, W.
Huang, H. Zhang, Nat. Chem. 2018,10, 638.
[87] W. Ding, L. Hu, J. Dai, X. Tang, R. Wei, Z. Sheng, C. Liang,
D. Shao, W. Song, Q. Liu, M. Chen, X. Zhu, S. Chou, X. Zhu,
Q. Chen, Y. Sun, S. X. Dou, ACS Nano 2019,13, 1694.
[88] J. Strachan, A. F. Masters, T. Maschmeyer, J. Mater. Chem. A
2021,9, 9451.
[89] M. Acerce, D. Voiry, M. Chhowalla, Nat. Nanotechnol. 2015,
10, 313.
[90] Z. Lei, J. Zhan, L. Tang, Y. Zhang, Y. Wang, Adv. Energy
Mater. 2018,8, 1703482.
[91] M. A. Lukowski, A. S. Daniel, F. Meng, A. Forticaux, L. Li,
S. Jin, J. Am. Chem. Soc. 2013,135, 10274.
[92] J. K. Nørskov, T. Bligaard, J. Rossmeisl, C. H. Christensen,
Nat. Chem. 2009,1, 37.
[93] B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H.
Nielsen, S. Horch, I. Chorkendorff, J. K. Nørskov, J. Am.
Chem. Soc. 2005,127, 5308.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (27 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[94] T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S.
Horch, I. Chorkendorff, Science 2007,317, 100.
[95] C. Tsai, K. Chan, J. K. Nørskov, F. Abild-Pedersen, Surf. Sci.
2015,640, 133.
[96] D. Voiry, M. Salehi, R. Silva, T. Fujita, M. Chen, T. Asefa,
V. B. Shenoy, G. Eda, M. Chhowalla, Nano Lett. 2013,13,
6222.
[97] R. Peng, L. Liang, Z. D. Hood, A. Boulesbaa, A. Puretzky,
A. V. Ievlev, J. Come, O. S. Ovchinnikova, H. Wang, C. Ma,
M. Chi, B. G. Sumpter, Z. Wu, ACS Catal. 2016,6, 6723.
[98] Y. Luo, K. Ren, S. Wang, J.-P. Chou, J. Yu, Z. Sun, M. Sun, J.
Phys. Chem. C 2019,123, 22742.
[99] L. Zhu, Y.-F. Ding, W.-J. Yang, S.-F. Yin, M.-Q. Cai, Phys.
Chem. Chem. Phys. 2021,23, 18125.
[100] R. Kumar, D. Das, A. K. Singh, J. Catal. 2018,359, 143.
[101] A. K. Geim, I. V. Grigorieva, Nature 2013,499, 419.
[102] K. S. Novoselov, A. Mishchenko, A. Carvalho, A. H. Cas-
tro Neto, Science 2016,353, aac9439.
[103] Y. Liu, N. O. Weiss, X. Duan, H.-C. Cheng, Y. Huang, X.
Duan, Nat. Rev. Mater. 2016,1, 16042.
[104] A. Castellanos-Gomez, X. Duan, Z. Fei, H. R. Gutierrez, Y.
Huang, X. Huang, J. Quereda, Q. Qian, E. Sutter, P. Sutter,
Nat. Rev. Methods Primers 2022,2, 58.
[105] C. Jin, E. Y. Ma, O. Karni, E. C. Regan, F. Wang, T. F. Heinz,
Nat. Nanotechnol. 2018,13, 994.
[106] B. Sun, Z. Liang, Y. Qian, X. Xu, Y. Han, J. Tian, ACS Appl.
Mater. Interfaces 2020,12, 7257.
[107] Y. Tang, W. Zhou, Q. Shang, Y. Guo, H. Hu, Z. Li, Y. Zhang,
L. Liu, H. Wang, X. Tan, T. Yu, J. Ye, Appl. Catal. B 2022,
310, 121295.
[108] X. Hong, J. Kim, S.-F. Shi, Y. Zhang, C. Jin, Y. Sun, S.
Tongay, J. Wu, Y. Zhang, F. Wang, Nat. Nanotechnol. 2014,9,
682.
[109] L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle,
A. Mishchenko, Y. J. Kim, R. V. Gorbachev, T. Georgiou,
S. V. Morozov, A. N. Grigorenko, A. K. Geim, C. Casiraghi,
A. H. C. Neto, K. S. Novoselov, Science 2013,340, 1311.
[110] H. Liu, B. Xu, J. M. Liu, J. Yin, F. Miao, C.-G. Duan, X. G.
Wan, Phys. Chem. Chem. Phys. 2016,18, 14222.
[111] J. Qin, W. Zhao, X. Hu, J. Li, P. Ndokoye, B. Liu, ACS Appl.
Mater. Interfaces 2021,13, 7127.
[112] Y. Zhang, T.-R. Chang, B. Zhou, Y.-T. Cui, H. Yan, Z. Liu,
F. Schmitt, J. Lee, R. Moore, Y. Chen, H. Lin, H.-T. Jeng, S.-
K. Mo, Z. Hussain, A. Bansil, Z.-X. Shen, Nat. Nanotechnol.
2014,9, 111.
[113] X. Li, J. Yu, S. Wageh, A. A. Al-Ghamdi, J. Xie, Small 2016,
12, 6640.
[114] Q. Xiang, B. Cheng, J. Yu, Angew. Chem. Int. Ed. 2015,54,
11350; Angew. Chem. 2015,127, 11508.
[115] X. Li, R. Shen, S. Ma, X. Chen, J. Xie, Appl. Surf. Sci. 2018,
430, 53.
[116] K. P. Loh, Q. Bao, G. Eda, M. Chhowalla, Nat. Chem. 2010,2,
1015.
[117] W.-J. Ong, L.-L. Tan, Y. H. Ng, S.-T. Yong, S.-P. Chai, Chem.
Rev. 2016,116, 7159.
[118] S. Cao, J. Low, J. Yu, M. Jaroniec, Adv. Mater. 2015,27, 2150.
[119] J. Wen, J. Xie, X. Chen, X. Li, Appl. Surf. Sci. 2017,391, 72.
[120] Y. Li, K. Yin, L. Wang, X. Lu, Y. Zhang, Y. Liu, D. Yan, Y.
Song, S. Luo, Appl. Catal. B 2018,239, 537.
[121] S. Pinilla, J. Coelho, K. Li, J. Liu, V. Nicolosi, Nat. Rev.
Mater. 2022,7, 717.
[122] D. Voiry, A. Goswami, R. Kappera, C. D. C. C. E. Silva, D.
Kaplan, T. Fujita, M. Chen, T. Asefa, M. Chhowalla, Nat.
Chem. 2015,7, 45.
[123] M. Su, W. Zhou, L. Liu, M. Chen, Z. Jiang, X. Luo, Y. Yang,
T. Yu, W. Lei, C. Yuan, Adv. Funct. Mater. 2022,32, 2111067.
[124] D. Voiry, A. Mohite, M. Chhowalla, Chem. Soc. Rev. 2015,
44, 2702.
[125] R. Kappera, D. Voiry, S. E. Yalcin, B. Branch, G. Gupta,
A. D. Mohite, M. Chhowalla, Nat. Mater. 2014,13, 1128.
[126] Y. Qi, Q. Xu, Y. Wang, B. Yan, Y. Ren, Z. Chen, ACS Nano
2016,10, 2903.
[127] Y. R. Girish, R. Biswas, M. De, Chem. Eur. J. 2018,24, 13871.
[128] W. Zhao, R. M. Ribeiro, G. Eda, Acc. Chem. Res. 2015,48,
91.
[129] Y. Zhao, J. Tu, Y. Sun, X. Hu, J. Ning, W. Wang, F. Wang, Y.
Xu, L. He, J. Phys. Chem. C 2018,122, 26570.
[130] Y. Zhao, W. Wang, C. Li, Y. Sun, H. Xu, J. Tu, J. Ning, Y.
Xu, L. He, Appl. Surf. Sci. 2018,456, 133.
[131] I. Shown, S. Samireddi, Y.-C. Chang, R. Putikam, P.-H.
Chang, A. Sabbah, F.-Y. Fu, W.-F. Chen, C.-I. Wu, T.-Y. Yu,
P.-W. Chung, M. C. Lin, L.-C. Chen, K.-H. Chen, Nat.
Commun. 2018,9, 169.
[132] Y. Li, S. Wu, J. Zheng, Y.-K. Peng, D. Prabhakaran, R. A.
Taylor, S. C. E. Tsang, Mater. Today 2020,41, 34.
[133] A. R. Shelke, H.-T. Wang, J.-W. Chiou, I. Shown, A. Sabbah,
K.-H. Chen, S.-A. Teng, I. A. Lin, C.-C. Lee, H.-C. Hsueh,
Y.-H. Liang, C.-H. Du, P. L. Yadav, S. C. Ray, S.-H. Hsieh,
C.-W. Pao, H.-M. Tsai, C.-H. Chen, K.-H. Chen, L.-C. Chen,
W.-F. Pong, Small 2022,18, 2105076.
[134] L.-P. Feng, J. Su, Z.-T. Liu, J. Alloys Compd. 2014,613, 122.
[135] D. Voiry, R. Fullon, J. Yang, C. de Carvalho Castro e Silva,
R. Kappera, I. Bozkurt, D. Kaplan, M. J. Lagos, P. E. Batson,
G. Gupta, A. D. Mohite, L. Dong, D. Er, V. B. Shenoy, T.
Asefa, M. Chhowalla, Nat. Mater. 2016,15, 1003.
[136] L. Lin, N. Miao, J. Huang, S. Zhang, Y. Zhu, D. D. Horsell,
P. Ghosez, Z. Sun, D. A. Allwood, Nano Energy 2017,38,
544.
[137] N. Luo, C. Chen, D. Yang, W. Hu, F. Dong, Appl. Catal. B
2021,299, 120664.
[138] L. Li, E. A. Carter, J. Am. Chem. Soc. 2019,141, 10451.
[139] S. Bai, N. Zhang, C. Gao, Y. Xiong, Nano Energy 2018,53,
296.
[140] Y. Sun, Y. Zang, W. Tian, X. Yu, J. Qi, L. Chen, X. Liu, H.
Qiu, Energy Environ. Sci. 2022,15, 1201.
[141] Y. Rho, K. Lee, L. Wang, C. Ko, Y. Chen, P. Ci, J. Pei, A.
Zettl, J. Wu, C. P. Grigoropoulos, Nat. Electronics 2022,5,
505.
[142] Z. Guo, Y. Zeng, F. Meng, H. Qu, S. Zhang, S. Hu, S. Fan, H.
Zeng, R. Cao, P. N. Prasad, D. Fan, H. Zhang, eLight 2022,2,
9.
[143] H. Yang, Y. Zhao, Q. Wen, Y. Mi, Y. Liu, H. Li, T. Zhai,
Nano Res. 2021,14, 4814.
[144] S. Bai, W. Jiang, Z. Li, Y. Xiong, ChemNanoMat 2015,1, 223.
[145] R. Shen, C. Jiang, Q. Xiang, J. Xie, X. Li, Appl. Surf. Sci.
2019,471, 43.
[146] T. Su, Q. Shao, Z. Qin, Z. Guo, Z. Wu, ACS Catal. 2018,8,
2253.
[147] Q. Li, C. Pan, J. Wang, L.-L. Wang, X. Zhu, Appl. Surf. Sci.
2022,605, 154720.
[148] Y. Liu, Z. Jiang, J. Jia, J. Robertson, Y. Guo, Appl. Surf. Sci.
2023,611, 155674.
[149] J. Yu, S. Seo, Y. Luo, Y. Sun, S. Oh, C. T. K. Nguyen, C. Seo,
J.-H. Kim, J. Kim, H. Lee, ACS Nano 2020,14, 1715.
[150] Y. Fu, F. Ye, X. Zhang, Y. He, X. Li, Y. Tang, J. Wang, D.
Gao, ACS Nano 2022,16, 18376.
[151] C.-F. Fu, X. Li, J. Yang, Chem. Sci. 2021,12, 2863.
[152] Y.-S. Huang, Y.-C. Hsiao, S.-H. Tzeng, Y.-H. Wu, T.-P.
Perng, M.-Y. Lu, Y.-L. Chueh, L.-J. Chen, Nano Energy 2020,
77, 105267.
[153] H.-R. Lee, J. H. Park, F. Raza, D. Yim, S.-J. Jeon, H.-I. Kim,
K. W. Bong, J.-H. Kim, Chem. Commun. 2016,52, 6150.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (28 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[154] X. Li, S. Guo, W. Li, X. Ren, J. Su, Q. Song, A. J. Sobrido, B.
Wei, Nano Energy 2019,57, 388.
[155] P. Atkin, T. Daeneke, Y. Wang, B. J. Carey, K. J. Berean,
R. M. Clark, J. Z. Ou, A. Trinchi, I. S. Cole, K. Kalantar-
zadeh, J. Mater. Chem. A 2016,4, 13563.
[156] M. H. D. Guimarães, H. Gao, Y. Han, K. Kang, S. Xie, C.-J.
Kim, D. A. Muller, D. C. Ralph, J. Park, ACS Nano 2016,10,
6392.
[157] J. Low, J. Yu, M. Jaroniec, S. Wageh, A. A. Al-Ghamdi, Adv.
Mater. 2017,29, 1601694.
[158] G. Liu, C. Kolodziej, R. Jin, S. Qi, Y. Lou, J. Chen, D. Jiang,
Y. Zhao, C. Burda, ACS Nano 2020,14, 5468.
[159] J. Chen, X.-J. Wu, Q. Lu, M. Zhao, P.-F. Yin, Q. Ma, G.-H.
Nam, B. Li, B. Chen, H. Zhang, Small 2021,17, 2006135.
[160] Y.-J. Yuan, Z. Shen, S. Wu, Y. Su, L. Pei, Z. Ji, M. Ding, W.
Bai, Y. Chen, Z.-T. Yu, Z. Zou, Appl. Catal. B 2019,246, 120.
[161] X. Lu, J. Xie, X. Chen, X. Li, Appl. Catal. B 2019,252, 250.
[162] Y. Wu, Z. Liu, Y. Li, J. Chen, X. Zhu, P. Na, Chin. J. Catal.
2019,40, 60.
[163] G.-J. Lai, L.-M. Lyu, Y.-S. Huang, G.-C. Lee, M.-P. Lu, T.-P.
Perng, M.-Y. Lu, L.-J. Chen, Nano Energy 2021,81, 105608.
[164] Y. Wang, L. Li, W. Yao, S. Song, J. T. Sun, J. Pan, X. Ren, C.
Li, E. Okunishi, Y.-Q. Wang, E. Wang, Y. Shao, Y. Y. Zhang,
H.-T. Yang, E. F. Schwier, H. Iwasawa, K. Shimada, M.
Taniguchi, Z. Cheng, S. Zhou, S. Du, S. J. Pennycook, S. T.
Pantelides, H.-J. Gao, Nano Lett. 2015,15, 4013.
[165] J. Chen, X.-J. Wu, Y. Gong, Y. Zhu, Z. Yang, B. Li, Q. Lu, Y.
Yu, S. Han, Z. Zhang, Y. Zong, Y. Han, L. Gu, H. Zhang, J.
Am. Chem. Soc. 2017,139, 8653.
[166] S. Ma, J. Xie, J. Wen, K. He, X. Li, W. Liu, X. Zhang, Appl.
Surf. Sci. 2017,391, 580.
[167] C. Ding, C. Zhao, S. Cheng, X. Yang, Chin. J. Catal. 2022,43,
403.
[168] Z.-W. Zhang, Q.-H. Li, X.-Q. Qiao, D. Hou, D.-S. Li, Chin. J.
Catal. 2019,40, 371.
[169] M. Xiong, J. Yan, B. Chai, G. Fan, G. Song, J. Mater. Sci.
Technol. 2020,56, 179.
[170] S. Zhang, S. Duan, G. Chen, S. Meng, X. Zheng, Y. Fan, X.
Fu, S. Chen, Chin. J. Catal. 2021,42, 193.
[171] J. Y. Lee, S. Kang, D. Lee, S. Choi, S. Yang, K. Kim, Y. S.
Kim, K. C. Kwon, S. H. Choi, S. M. Kim, J. Kim, J. Park, H.
Park, W. Huh, H. S. Kang, S. W. Lee, H.-G. Park, M. J. Ko,
H. Cheng, S. Han, H. W. Jang, C.-H. Lee, Nano Energy 2019,
65, 104053.
[172] H. Mao, H. Yang, J. Liu, S. Zhang, D. Liu, Q. Wu, W. Sun,
X.-M. Song, T. Ma, Chin. J. Catal. 2022,43, 1341.
[173] H. Ou-Yang, H.-M. Xu, X.-L. Zhang, Y.-Q. Liu, Y.-Q. He, L.
Shi, C. Gu, S.-K. Han, Small 2022,18, 2202109.
[174] J. Liu, H. Xu, J. Yan, J. Huang, Y. Song, J. Deng, J. Wu, C.
Ding, X. Wu, S. Yuan, H. Li, J. Mater. Chem. A 2019,7,
18906.
[175] Y. Bai, H. Zhang, X. Wu, N. Xu, Q. Zhang, J. Phys. Chem. C
2022,126, 2587.
[176] D. Singh, S. K. Gupta, Y. Sonvane, A. Kumar, R. Ahuja,
Catal. Sci. Technol. 2016,6, 6605.
[177] D. Saraf, S. Chakraborty, A. Kshirsagar, R. Ahuja, Nano
Energy 2018,49, 283.
[178] M. Rahman, K. Davey, S.-Z. Qiao, Adv. Funct. Mater. 2017,
27, 1606129.
[179] P. Yang, D. Wang, X. Zhao, W. Quan, Q. Jiang, X. Li, B.
Tang, J. Hu, L. Zhu, S. Pan, Y. Shi, Y. Huan, F. Cui, S. Qiao,
Q. Chen, Z. Liu, X. Zou, Y. Zhang, Nat. Commun. 2022,13,
3238.
[180] V. Andrei, G. M. Ucoski, C. Pornrungroj, C. Uswachoke, Q.
Wang, D. S. Achilleos, H. Kasap, K. P. Sokol, R. A. Jagt, H.
Lu, T. Lawson, A. Wagner, S. D. Pike, D. S. Wright, R. L. Z.
Hoye, J. L. MacManus-Driscoll, H. J. Joyce, R. H. Friend, E.
Reisner, Nature 2022,608, 518.
[181] Z. Zhang, Z. Huang, J. Li, D. Wang, Y. Lin, X. Yang, H. Liu,
S. Liu, Y. Wang, B. Li, X. Duan, X. Duan, Nat. Nanotechnol.
2022,17, 493.
[182] H. Li, J. Xiao, Q. Fu, X. Bao, Proc. Natl. Acad. Sci. USA
2017,114, 5930.
[183] W. Yuan, B. Zhu, K. Fang, X.-Y. Li, T. W. Hansen, Y. Ou, H.
Yang, J. B. Wagner, Y. Gao, Y. Wang, Z. Zhang, Science
2021,371, 517.
[184] R. Yang, L. Mei, Y. Fan, Q. Zhang, H.-G. Liao, J. Yang, J. Li,
Z. Zeng, Nat. Protocols 2022, https://doi.org/10.1038/s41596-
022-00762-y.
[185] H. Nishiyama, T. Yamada, M. Nakabayashi, Y. Maehara, M.
Yamaguchi, Y. Kuromiya, Y. Nagatsuma, H. Tokudome, S.
Akiyama, T. Watanabe, R. Narushima, S. Okunaka, N.
Shibata, T. Takata, T. Hisatomi, K. Domen, Nature 2021,598,
304.
Manuscript received: December 7, 2022
Accepted manuscript online: January 2, 2023
Version of record online: ■■,
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 (29 of 29) © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Reviews
Photocatalysis
R. Yang, Y. Fan, Y. Zhang, L. Mei, R. Zhu,*
J. Qin, J. Hu, Z. Chen, Y. Hau Ng, D. Voiry,
S. Li, Q. Lu, Q. Wang,* J. C. Yu,*
Z. Zeng* e202218016
2D Transition Metal Dichalcogenides for
Photocatalysis
The fundamentals of 2D transition metal
dichalcogenides (TMDs), their synthe-
sis, their advantages in photocatalysis,
and the strategies for boosting their
photocatalytic performance are summar-
ized in this review. Currently problems
and their solutions are presented.
Angewandte
Chemie
Reviews
Angew. Chem. Int. Ed. 2023, e202218016 © 2023 Wiley-VCH GmbH
15213773, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/anie.202218016 by University Of Calgary Libraries And Cultural Resources, Wiley Online Library on [19/01/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... Transition metal dichalcogenides (TMDCs) are typically layered materials, and monoor few-layer two-dimensional (2D) TMDC nanomaterials can be obtained by various techniques [1,2], holding extraordinary and special characteristics [3][4][5][6], which have been widely studied in the field of electronics [7], optoelectronics [8,9], energy storage [10,11], catalysis et al. [12][13][14][15]. TMDCs have a general formula of MX 2 (M stands for transition metal, and X stands for chalcogenide). ...
Article
Full-text available
Atomically thin two-dimensional transition metal dichalcogenides (TMDCs) have been regarded as ideal and promising nanomaterials that bring broad application prospects in extensive fields due to their ultrathin layered structure, unique electronic band structure, and multiple spatial phase configurations. TMDCs with different phase structures exhibit great diversities in physical and chemical properties. By regulating the phase structure, their properties would be modified to broaden the application fields. In this mini review, focusing on the most widely concerned molybdenum dichalcogenides (MoX2: X = S, Se, Te), we summarized their phase structures and corresponding electronic properties. Particularly, the mechanisms of phase transformation are explained, and the common methods of phase regulation or phase stabilization strategies are systematically reviewed and discussed. We hope the review could provide guidance for the phase regulation of molybdenum dichalcogenides nanomaterials, and further promote their real industrial applications.
Article
Tumor‐associated macrophages (TAMs) play a crucial function in solid tumor antigen clearance and immune suppression. Notably, 2D transitional metal dichalcogenides (i.e., molybdenum disulfide (MoS 2 ) nanozymes) with enzyme‐like activity are demonstrated in animal models for cancer immunotherapy. However, in situ engineering of TAMs polarization through sufficient accumulation of free radical reactive oxygen species for immunotherapy in clinical samples remains a significant challenge. In this study, defect‐rich metastable MoS 2 nanozymes, i.e., 1T2H‐MoS 2 , are designed via reduction and phase transformation in molten sodium as a guided treatment for human breast cancer. The as‐prepared 1T2H‐MoS 2 exhibited enhanced peroxidase‐like activity (≈12‐fold enhancement) than that of commercial MoS 2 , which is attributed to the charge redistribution and electronic state induced by the abundance of S vacancies. The 1T2H‐MoS 2 nanozyme can function as an extracellular hydroxyl radical generator, efficiently repolarizing TAMs into the M1‐like phenotype and directly killing cancer cells. Moreover, the clinical feasibility of 1T2H‐MoS 2 is demonstrated via ex vivo therapeutic responses in human breast cancer samples. The apoptosis rate of cancer cells is 3.4 times greater than that of cells treated with chemotherapeutic drugs (i.e., doxorubicin).
Article
Full-text available
2D semiconductor nanomaterials have received significant attention as photocatalysts for solar‐to‐hydrogen conversion due to their robust light absorption capacity, large specific surface area, and superior electron transport characteristics. Nevertheless, it is challenging to develop 2D quaternary copper‐based sulfides (QCSs) with functional integration of conducive structural features for photocatalysis from multiple elements and intricate control conditions. Herein, ultrathin 2D alloyed Cu−Ga−Zn−S (CGZS) curved nanobelts (NBs) are fabricated by using a facile one‐pot colloidal method. Subsequently, the derived Cu31S16‐CGZS Janus heterostructures are designed by increasing Cu concentrations. The formation mechanism of 2D curved alloys and Janus heterostructures is systematically investigated. Without cocatalysts, curved alloyed CGZS NBs demonstrated superior photocatalytic activities of 1264.2 µmol g⁻¹ h⁻¹ compared to Janus heterostructured Cu31S16−CGZS (463.1 µmol g⁻¹ h⁻¹) under visible light (> 400 nm). Experimental and theoretical results unveiled that the improved photocatalytic activities of curved CGZS NBs can be attributed to enhanced charge–carrier separation efficiency resulting from high crystallinity and ultrathin 2D structure, abundant active sites from curved structure, high‐active (0001) crystal facet with the lowest reaction Gibbs energy, work function, and metal‐like properties. This work offers new insights into functional integration into ultrathin 2D QCSs with enhanced visible‐light photocatalytic performance.
Article
Full-text available
Perovskite solar cells (PSCs) offer low costs and high power conversion efficiency. However, the lack of long-term stability, primarily stemming from the interfacial defects and the susceptible metal electrodes, hinders their practical application. In the past few years, two-dimensional (2D) materials (e.g., graphene and its derivatives, transitional metal dichalcogenides, MXenes, and black phosphorus) have been identified as a promising solution to solving these problems because of their dangling bond-free surfaces, layer-dependent electronic band structures, tunable functional groups, and inherent compactness. Here, recent progress of 2D material toward efficient and stable PSCs is summarized, including its role as both interface materials and electrodes. We discuss their beneficial effects on perovskite growth, energy level alignment, defect passivation, as well as blocking external stimulus. In particular, the unique properties of 2D materials to form van der Waals heterojunction at the bottom interface are emphasized. Finally, perspectives on the further development of PSCs using 2D materials are provided, such as designing high-quality van der Waals heterojunction, enhancing the uniformity and coverage of 2D nanosheets, and developing new 2D materials-based electrodes.
Article
Full-text available
Fundamentally understanding the complex electrochemical reactions that are associated with energy devices (e.g., rechargeable batteries, fuel cells and electrolyzers) has attracted worldwide attention. In situ liquid cell transmission electron microscopy (TEM) offers opportunities to directly observe and analyze in-liquid specimens without the need for freezing or drying, which opens up a door for visualizing these complex electrochemical reactions at the nano scale in real time. The key to the success of this technique lies in the design and fabrication of electrochemical liquid cells with thin but strong imaging windows. This protocol describes the detailed procedures of our established technique for the fabrication of such electrochemical liquid cells (~110 h). In addition, the protocol for the in situ TEM observation of electrochemical reactions by using the nanofabricated electrochemical liquid cell is also presented (2 h). We also show and analyze experimental results relating to the electrochemical reactions captured. We believe that this protocol will shed light on strategies for fabricating high-quality TEM liquid cells for probing dynamic electrochemical reactions in high resolution, providing a powerful research tool. This protocol requires access to a clean room equipped with specialized nanofabrication setups as well as TEM characterization equipment. This protocol describes the procedure for the fabrication of electrochemical liquid cells for in situ liquid cell transmission electron microscopy. This allows direct visualization of complex electrochemical reactions at the nano scale in real time.
Article
Full-text available
Photoelectrochemical (PEC) artificial leaves hold the potential to lower the costs of sustainable solar fuel production by integrating light harvesting and catalysis within one compact device. However, current deposition techniques limit their scalability1, whereas fragile and heavy bulk materials can affect their transport and deployment. Here we demonstrate the fabrication of lightweight artificial leaves by employing thin, flexible substrates and carbonaceous protection layers. Lead halide perovskite photocathodes deposited onto indium tin oxide-coated polyethylene terephthalate achieved an activity of 4,266 µmol H2 g−1 h−1 using a platinum catalyst, whereas photocathodes with a molecular Co catalyst for CO2 reduction attained a high CO:H2 selectivity of 7.2 under lower (0.1 sun) irradiation. The corresponding lightweight perovskite-BiVO4 PEC devices showed unassisted solar-to-fuel efficiencies of 0.58% (H2) and 0.053% (CO), respectively. Their potential for scalability is demonstrated by 100 cm2 stand-alone artificial leaves, which sustained a comparable performance and stability (of approximately 24 h) to their 1.7 cm2 counterparts. Bubbles formed under operation further enabled 30–100 mg cm−2 devices to float, while lightweight reactors facilitated gas collection during outdoor testing on a river. This leaf-like PEC device bridges the gulf in weight between traditional solar fuel approaches, showcasing activities per gram comparable to those of photocatalytic suspensions and plant leaves. The presented lightweight, floating systems may enable open-water applications, thus avoiding competition with land use. This work introduces lightweight, leaf-like photoelectrochemical devices for unassisted water splitting and syngas production, which could be used in the fabrication of floating systems for solar fuel production.
Article
Full-text available
The design and synthesis of advanced semiconductors is crucial for the full utilization of solar energy. Herein, colloidal selective‐epitaxial hybrid of tripartite semiconducting sulfides CuInS2Cd(In)SMoS2 heteronanostructures (HNs) via lateral‐ and vertical‐epitaxial growths, followed by cation exchange reactions, are reported. The lateral‐epitaxial CuInS2 and Cd(In)S enable effective visible to near‐infrared (NIR) solar spectrum absorption, and the vertical‐epitaxial ultrathin MoS2 offer sufficient edge sulfur sites for the hydrogen evolution reaction (HER). Furthermore, the integrated structures exhibit unique epitaxial‐staggered type II band alignments for continuous charge separation. They achieve the H2 evolution rate up to 8 mmol h⁻¹ g⁻¹, which is ≈35 times higher than bare CdS and show no deactivation after long‐term cycling, representing one of the most efficient and robust noble‐metal‐free photocatalysts. This design principle and transformation protocol open a new way for creating all‐in‐one multifunctional catalysts in a predictable manner.
Article
Full-text available
Chemical doping can be used to control the charge-carrier polarity and concentration in two-dimensional van der Waals materials. However, conventional methods based on substitutional doping or surface functionalization result in the degradation of electrical mobility due to structural disorder, and the maximum doping density is set by the solubility limit of dopants. Here we show that a reversible laser-assisted chlorination process can be used to create high doping concentrations (above 3 × 1013 cm−2) in graphene monolayers with minimal drops in mobility. The approach uses two lasers—with distinct photon energies and geometric configurations—that are designed for chlorination and subsequent chlorine removal, allowing highly doped patterns to be written and erased without damaging the graphene. To illustrate the capabilities of our approach, we use it to create rewritable photoactive junctions for graphene-based photodetectors. Two laser beams with different energies and configurations can be used to reversibly dope graphene via chlorination and chlorine removal, allowing rewritable graphene photodetectors to be fabricated.
Article
The use of semiconductor photocatalysis for hydrogen production is an ideal approach for achieving solar energy and conversion. In this work, we systematically examine the photocatalytic properties of WSe2/MoSi2N4 van der Waals heterojunctions using first-principles calculation and nonadiabatic molecular dynamics. The results demonstrate that the WSe2/MoSi2N4 heterojunction possesses a 1.81 eV indirect band gap and type-II band alignment, which ensures that the photoexcited electron–hole (e–h) pairs are spatially separated. The carrier lifetime of the photoexcited e–h pairs is 278 ps significantly longer than the interlayer hole transfer time of 335 fs, implying that the heterojunction has the high quantum efficiency. In addition, this heterojunction possesses superior optical absorption (10⁵ cm⁻¹), low overpotential for OER (0.60 V), and outstanding light absorption properties. These findings indicate that this heterojunction can achieve highly efficient and spontaneous photocatalytic water splitting.
Article
Currently, one of the main reasons for the ineffectiveness of tumor treatment is that the abnormally high tumor interstitial pressure (TIP) hinders the delivery of drugs to the tumor center and promotes intratumoral cell survival and metastasis. Herein, we designed a "nanomotor" by in situ growth of Ag2S nanoparticles on the surface of ultrathin WS2 to fabricate Z-scheme photocatalytic drug AWS@M, which could rapidly enter tumors by splitting water in interstitial liquid to reduce TIP, along with O2 generation. Moreover, the O2 would be further converted to reactive oxygen species (ROS), accompanied by increased local temperature of tumors, and the combination of ROS with thermotherapy could eliminate the deep tumor cells. Therefore, the "nanomotor'' could effectively reduce the TIP levels of cervical cancer and pancreatic cancer (degradation rates of 40.2% and 36.1%, respectively) under 660 nm laser irradiation, further enhance intratumor drug delivery, and inhibit tumor growth (inhibition ratio 95.83% and 87.61%, respectively), and the related mechanism in vivo was explored. This work achieves efficiently photocatalytic water-splitting in tumor interstitial fluid to reduce TIP by the nanomotor, which addresses the bottleneck problem of blocking of intratumor drug delivery, and provides a general strategy for effectively inhibiting tumor growth.
Article
Constructing layered heterostructured photocatalysts based on two dimensional semiconductor materials, such as transition metal dichalcogenides (TMDs) and graphitic carbon nitride (CxNy), has become a research hotspot in photocatalysis with the remarkable improvement of photoexcited charge separation efficiency. Herein, the electronic properties and photocatalytic activities of WS2/C2N bilayer and trilayer heterojunctions have been comparatively investigated using first-principles calculations. The results demonstrate that the WS2/C2N, WS2/C2N/WS2, and C2N/WS2/C2N are type-II van der Waals heterostructures, with an indirect bandgap of 1.79 eV, 1.72 eV, and 1.79 eV, respectively. Compared to WS2/C2N bilayer, the two sandwiched heterojunctions demonstrate stronger interfacial interactions, charge transfer, and visible-light absorption. The free energy calculations for redox reactions of the sandwiched heterojunctions further demonstrate high catalytic activity for hydrogen evolution reaction. More interestingly, biaxial strains can adjust the bandgap width (semiconductor→metal) and induce the band type transition (indirect→direct) in WS2/C2N sandwiched heterojunctions. For example, a weak stress about 1% leads to direct band gap both in type-II WS2/C2N/WS2 and C2N/WS2/C2N heterojunctions, without degrading their excellent abilities for water decomposition. This work not only provides rational strategies to design TMDs/CxNy-based photocatalysts in view of interface and bandgap engineering, but also reveals their potential applications in electronic and optoelectronic devices.
Article
The global adoption of efficient sustainable energy sources is a crucial step toward meeting energy demands while achieving carbon emission reduction targets. Solar energy has become a clean and cost-competitive alternative to traditional fossil fuels, but the intermittent nature of sunlight results in challenges associated with energy storage and transport. Photocatalytic carbon dioxide reduction intends to mimic natural photosynthesis for utilizing sunlight to chemically convert water and CO2 into fuels. In this process, the solar energy is captured and stored in fuels, so-called solar fuels, for widespread on-demand use. Heterogeneous solar fuel production systems are multi-component, comprising light-harvesting (photosensitizer) and catalytic (cocatalyst) units. Cocatalysts are indispensable for photocatalytic CO2 reduction systems, which promote charge carrier separation and transport, reduce the reaction activation energy, and alter the reaction route, thereby enhancing the activity and selectivity of the photocatalytic reactions. This review presents a comprehensive summary of the recent advancements in cocatalysts for photocatalytic CO2 reduction reaction (CO2RR), with the purpose of providing new insights and guidance to the field with regard to research directions and best practices. We summarize how various cocatalysts including inorganic nanoparticles, metal complexes, enzymes, and bacteria can be combined with semiconductor photosensitizer for light-driven photocatalytic CO2RR. Side-by-side comparisons reveal the strengths and limitations of each kind of cocatalysts and how lessons extracted from studying natural photosynthetic systems can be applied to investigations of artificial photosynthesis, presenting an outlook discussing possible future concepts for a more effective photocatalytic CO2 reduction process. [Figure not available: see fulltext.]
Article
The integration of dissimilar materials into heterostructures has become a powerful tool for engineering interfaces and electronic structure. The advent of 2D materials has provided unprecedented opportunities for novel heterostructures in the form of van der Waals stacks, laterally stitched 2D layers and more complex layered and 3D architectures. This Primer provides an overview of state-of-the-art methodologies for producing such van der Waals heterostructures, focusing on the two fundamentally different strategies, top-down deterministic assembly and bottom-up synthesis. Successful techniques, advantages and limitations are discussed for both approaches. As important as the fabrication itself is the characterization of the resulting engineered materials, for which a range of analysis techniques covering structure, composition and emerging functionality are highlighted. Examples of the properties of artificial van der Waals structures include optoelectronics and plasmonics, twistronics and unique functionality arising from the generalization of van der Waals assembly from 2D to 3D crystalline components. Finally, current issues of reproducibility, limitations and opportunities for future breakthroughs in terms of enhanced homogeneity, interfacial purity, feature control and ultimately orders-of-magnitude increased complexity of van der Waals heterostructures are discussed. Van der Waals epitaxy provides numerous opportunities for materials integration in heterostructures. This Primer provides an overview of methodologies for producing van der Waals heterostructures, focusing on top-down assembly and bottom-up synthesis, and discusses future opportunities for their continued development.
Article
Covalent organic frameworks (COFs) as efficient photocatalysts have attracted extensive attention due to their high crystallinity and tunable optical and electronic properties. This review first summarizes the fundamental aspects of COF photocatalysts—including the basic physical, chemical, optical, and electronic properties—and key considerations in the synthesis of crystalline COFs. Subsequently, deep insights into the photocatalytic mechanisms of COF-based photocatalysts (i.e., adsorption and activation mechanisms of reagents, the identifications of exact active sites, and the surface reaction mechanisms, as well as the kinetics of charge-carrier separation and transport) and engineering-modification strategies of COF-based photocatalysts are thoroughly addressed. Finally, various applications of COF-based photocatalysts are discussed, including photocatalytic hydrogen production, CO2 reduction, pollutant degradation, and organic transformation. A perspective on the challenges and future exploration of COF-based photocatalysts is also proposed. It is expected that this review will inspire new thoughts and will trigger exciting progress in COF-based photocatalysis.