Advanced Materials - 2021 - Wang - Stacking Engineered Heterostructures in Transition Metal Dichalcogenides

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Review

www.advmat.de

Stacking-Engineered Heterostructures in Transition


Metal Dichalcogenides
Shixuan Wang, Xuehao Cui, Chang’e Jian, Haowei Cheng, Mengmeng Niu, Jia Yu,
Jiaxu Yan,* and Wei Huang*

twistronics field, extending the twist idea


The layer-by-layer assembly of 2D transition metal dichalcogenide monolayer to various graphene analogues. Notably,
blocks to form a 3D stack, with a precisely chosen sequence/angle, is the transition metal dichalcogenides (TMDs)
newest development for these materials. In this way, one can create “van der are graphene analogues and have shown
Waals heterostructures (HSs),” opening up a new realm of materials engi- excellent optical, electrical, mechanical,
and thermodynamic properties.[11–13]
neering and novel devices with designed functionalities. Herein, a detailed Accordingly, they have been employed
systematic review of transition metal dichalcogenide stacking-engineered in a diverse range of nanodevice applica-
heterostructures, from controllable fabrication to typical characterization, tions such as optoelectronics,[14,15] bat-
and stacking-correlated physical behaviors is presented. Furthermore, recent teries,[16,17] sensors,[18,19] solar cells,[20,21]
advances in stacking design, such as stacking sequence, twist angles, and and catalysis.[22–24] Vertically stacking
TMD monolayers to form 3D blocks is an
moiré superlattice heterojunctions, are also comprehensively summarized.
emerging topic, with wide-ranging novel
Finally, the remaining challenges and possible strategies for using stacking possibilities.
engineering to tune the properties of 2D materials are also outlined. Stacking-designed heterobilayers,
whereby the monolayers are weakly bound
by van der Waals forces, allow for research
1. Introduction ideas similar to the magic angle,[7] and controlling the stacking
structures is a promising approach to adjust electron–hole pair
The successful manual exfoliation of graphene from graphite interactions. The twisted stacking of adjacent monolayers allows
in 2004[1] highlighted the 2D materials family, which includes crystallographic alignments and interlayer interactions to be
hexagonal boron nitride (h-BN),[2] black phosphorous (BP),[3] altered (as shown in Figure 1), and hence provides an extremely
and other graphene analogues.[4–6] Recently, eye-catching useful parameter for tuning the electronic properties of nano­
research in the 2D field has been performed by Cao et al., who materials. For instance, when a tungsten disulfide (WS2) or tung-
demonstrated magic-angle twisted bilayer graphene.[7–9] The sten diselenide (WSe2) monolayer is stacked on a molybdenum
authors realized unconventional superconductivity in magic- disulfide (MoS2) monolayer, the bandgap changes from direct to
angle graphene superlattices by stacking two nanosheets of indirect or inverse, forming HSs that exhibit unprecedented phe-
graphene that are relatively twisted by ≈1.1°. The magic-angle nomena.[25–28] Furthermore, the valley transfer of an exciton can
idea was further extended to graphene-like h-BN systems; also be tuned by changing the stacking configurations of HSs,
the twisted angle was controlled to tune intra- and interlayer particularly attractive for flexible optoelectronics.[29–31] In addi-
interactions, causing controversy in the physics community.[10] tion to the aforementioned studies, research[32,33] has also dem-
These remarkable works have led to the emergence of the onstrated that the stacking sequence of TMD monolayers in HSs
significantly influenced the photocatalytic activities,[23] photo-
voltaic performances,[34] and their stacking configurations. As
S. Wang, X. Cui, C. Jian, H. Cheng, M. Niu, J. Yu, Prof. J. X. Yan,
Prof. W. Huang such, distinct TMD monolayers can be vertically stacked to form a
Key Laboratory of Flexible Electronics (KLOFE) & Institute special-purpose van der Waals HS. By changing the stacking con-
of Advanced Materials (IAM) figurations and sequences of adjacent layers in real space, novel
Nanjing Tech University (Nanjing Tech) nanomaterials with interesting physical properties can be created.
30 South Puzhu Road, Nanjing 211800, China
E-mail: iamjxyan@njtech.edu.cn; iamwhuang@nwpu.edu.cn
Hitherto, experimental observations and theoretical calcula-
Prof. W. Huang
tions of research groups primarily focus on HSs composed of
Frontiers Science Center for Flexible Electronics two different TMD monolayers; there are few in-depth system-
Xi’an Institute of Flexible Electronics (IFE) and Xi’an atic explorations on the intrinsic mechanism of exciton transfer
Institute of Biomedical Materials & Engineering in stacking-engineered HSs. Moreover, a unified conclusion
Northwestern Polytechnical University has not been reached on the dependence of the dynamics on
127 West Youyi Road, Xi’an 710072, China
the constituent TMD layers, their stacking configurations, and
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adma.202005735.
order, whether in theory or experiment. Some studies even
provide opposing views. For example, Hoseok et  al.[28] and
DOI: 10.1002/adma.202005735 Gogoi et  al.[30] showed that interlayer exciton behaviors are

Adv. Mater. 2021, 33, 2005735 2005735  (1 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 1.  Superlattices formed by parallel (R-type) and antiparallel (H-type) aligned TMD bilayers. Summary of the unified naming for high-symmetry
stacks in TMD bilayers, obtained by sliding the top layer with respect to the bottom layer. The full stacking-engineered TMD bilayers are built primarily
upon such crystallographic alignments, with additional variations.

particularly susceptible to the bilayer stacking angles, while 2.1. Assembly


Ji et  al.[35] reported that charge transfer in MoS2/WS2 bilayers
was robust and stacking-independent. Moreover, the majority In general, all assembly methods are derived from the procedure
of previous studies only comparatively explained their obtained for graphene that was successfully tape-exfoliated from graphite in
phenomena, with minimal interpretation, failing to provide 2004.[1] Thus, although there are some differences between these
essential and unified explanations. approaches, all techniques operate using a similar experimental
Here, our aim is to underscore the enormous potential of system; an optical detection system combined with microposi-
stacking-tunable 2D layers, which is a relatively young field. This tioning transfer tools. TMD monolayers are obtained by mechanical
review focuses on stacking-designed heterobilayers based on peeling and chemical vapor deposition (CVD) growth.[40–42] Subse-
TMDs. First, the controllable preparation methods for stacking quently, monolayers can be transferred to a desired location on a
TMD heterobilayers are briefly introduced and compared, and substrate using a transfer medium, assisted by an optical detection
can be summarized into two major categories; assembly and system. After precise positioning on a micrometer or even nanom-
growth.[36–38] Second, the routine characterization of stacking eter scale, designed stacking structures have been constructed. The
bilayers is reviewed, with discussions on the techniques such as conventional processing steps are as follows (Figure 3):
Raman spectroscopy, photoluminescence (PL), second harmonic
I) Tape exfoliation: The initial TMD nanosheet is exfoliated from
generation (SHG), and scanning transmission electron micros-
bulk 2D materials using a sacrificial polymer similar to those
copy (STEM). Finally, the recent progress in stacking designs is
used for tape peeling (for example, poly(methyl methacrylate)
discussed in-depth based on two aspects: the stacking sequence,
(PMMA)), and the nanosheet and polymer are then attached
and the stacking configurations including twist angles and moiré
superlattices. The remaining challenges and possible strategies
for the use of stacking engineering are then discussed. This inte-
gration of previous investigations on stacking-engineered TMD
heterobilayers provides inspiration and support for latecomers
to the field of stacking-tunable photoelectronics.

2. Fabrication of Stacking-Designed Bilayers


To date, two ideas have been adopted to fabricate stacking TMD
heterobilayers: assembly and growth approaches. In this sec-
tion, we present a systematic overview of the fabrication of
stacked HSs based on TMD monolayers to gain a better under-
standing of precisely controllable stacking structures, including
a comprehensive description and a critical comparison. The
complete classification of the various fabrication methods is Figure 2.  Overview of the fabrication techniques for stacking-engineered
presented in Figure 2. HSs.

Adv. Mater. 2021, 33, 2005735 2005735  (2 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 3.  Schematic illustration of the mechanical assembly process: I) tape exfoliation, thermal release tape (TRT,[39] silicone elastomer); II) solution
etching; III) assembly. The STEM images show the stacking HSs with twisted angles of a) 29° and b) 50°, respectively. a,b) Reproduced with permis-
sion.[30] Copyright 2019, American Chemical Society.

to a transparent thermal release tape (TRT) in preparation for some cases, thermal annealing treatment is an essential last
position assembly. step for assembly, effectively enhancing the interlayer coupling
II) Solution etching: Compared to tape exfoliation, solution treat- of bilayers. Recently, there has been considerable interest in
ment is primarily carried out in order to obtain the mono­ creating large-scale assembly methods to meet experimental
layer through chemical corrosion. Initially, a protective and needs. Remarkably, Kang et  al.[43] used the tape exfoliation
supportive polymer film is spin-coated on the surface of the method to prepare wafer-scale semiconductor layers with supe-
monolayer, and the substrate is immersed in an etching so- rior spatial uniformity and pristine interlayer interfaces, assem-
lution (such as potassium hydroxide or hydrogen fluoride bling large-scale vertical HSs with varied stacking orientations.
solution), recovering the monolayer and polymer from the It is worth noting that this method uses a vacuum chamber-
solution and attaching to a TRT for alignment and release. assisted transfer and release; vacuum stacking significantly
III) Assembly: The nanosheet on the polymer is manipulated improves the interfacial quality and flatness of the layers, thus
to a desired position and orientation using microposition- reducing the amount of amorphous carbon or other contami-
ing transfer tools with the aid of a microscope, and aligned nants remaining at the interface. Compared to tape exfoliation,
with the predetermined target underneath. Subsequently, the layer recovery in the liquid phase is widely used to produce
sample is slowly lowered until it touches the target to form 2D flakes,[44–46] and this is an uncomplicated, easy to imple-
designed HSs, and the excess polymer is removed, as shown ment, and powerful approach. As a result, solution- etching
in Figure 3a,b.[30] The TRT can also be heated during the re- methods are more suitable for large-scale production than tape-
lease process, lowering its viscosity, allowing for facile removal exfoliation methods. Nonetheless, numerous disadvantages
of the tape and polymer. Based on these processes, additional exist for both processes; the sample is easily contaminated and
configurations and HS ordering can be easily achieved. prone to wrinkling and residues, however, the most important
issue is that the stacking configurations are difficult to control.
Consequently, the HSs with designed configurations are However, advantages such as rapid, cost-effective, and universal
superimposed using this deterministic assembly method. In applicability make these processes extremely attractive.

Adv. Mater. 2021, 33, 2005735 2005735  (3 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

2.2. Chemical Vapor Deposition Growth to micromechanical assembly, CVD growth is an alternative
method to prepare vertical HSs. Recent studies have demon-
As mentioned above, vertically stacked HSs can be assembled strated that CVD technology (Figure 4) can controllably and
using polymer-assisted layer-by-layer transfer methods, which efficiently synthesize different 2D materials,[47–51] offering a reli-
inevitably generate problems such as interface contamina- able and practical means to prepare HSs with desired physical
tion, layer wrinkles, and performance degradation. Compared properties.[52–58] To date, numerous studies have focused on the

Figure 4.  Controllable CVD growth: a) schematic illustrations of the temperature-dependent CVD growth processes. b) Schematic of the CVD setup.
c) Schematic illustration of the growth of NbS2/MoS2, from the three corners to the center of triangular MoS2. d) The ratio of AA and AB to (AA +
AB) varies with deposition temperature. The hollow icons represent the ratio of AA stacking, and the solid icons represent the ratio of AB stacking.
e) Schematic of the CVD growth process for lateral/vertical MoS2/WSe2 HSs with different W/Se ratios. f) Schematic illustration of the preparation of
MoS2/WS2 using WO3−x/MoO3−x nanowires as precursors. g) Representation of vertical ReS2/WS2 HSs with twinned growth. Au is used as the growth
substrate and W–Re alloy foil is used as the support substrate. h) Theoretical calculations confirm the twinned growth process of the ReS2/WS2 HSs.
EadsRe,Au(111)  = 1.52  eV, EadsW,Au(111)  = 2.77  eV, and EadsRe,WS2(001)  = 3.04  eV. i) Schematic view of the growth of Mo2C crystals on liquid Au substrates.
a) Reproduced with permission.[23] Copyright 2016, Wiley-VCH. b,c) Reproduced with permission.[64] Copyright 2018, American Chemical Society.
d) Reproduced with permission.[65] Copyright 2019, Springer Nature. e) Reproduced with permission.[66] Copyright 2019, Wiley-VCH. f) Reproduced
with permission.[62] Copyright 2015, Wiley-VCH. g,h) Reproduced under the terms of the CC-BY 4.0 license.[63] Copyright 2016, The Authors, published
by Springer Nature. i) Reproduced with permission.[67] Copyright 2019, IOP Publishing Ltd.

Adv. Mater. 2021, 33, 2005735 2005735  (4 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

preparation of stacking heterobilayers by the CVD method.[59–61] 2.2.2. Precursor Design


In particular, three types of preparation strategies deserve sig-
nificant attention, namely, temperature triggering,[23] precursor Two-step CVD methods employing different preparation condi-
design,[62] and substrate engineering.[63] These techniques are tions and chemical supplies are commonly used to synthesize
highly suitable for preparing HSs and show great potential for HSs.[68,75,76] However, changing the conditions between steps
practical application in a highly controllable manner. In this during the preparation process negatively impacts the resultant
section, studies on these three CVD growth approaches that aid materials produced, for example, alloying of Mo, W, and S.[77,78]
comprehension of the controllable synthesis of diverse stacking In an effort to overcome these negative effects, numerous
HSs are discussed. research efforts have been invested to rationally design the
experimental precursor in advance, allowing the develop-
ment of one-step methods to generate HSs.[62,79–82] Recently, Li
2.2.1. Temperature Triggering et  al.[66] showed that differences in the growth kinetics of pre-
cursors can be utilized to generate vertical or lateral HSs by
Previously, two-step CVD growth methods were employed rationally controlling the metal/chalcogenide ratio in the pre-
to synthesize vertically stacked heterojunctions.[68,69] How- cursor vapors. As shown in Figure 4e, at low W/Se ratios (WO3
ever, precursors were exposed to the gas phase during the and Se powders), nucleation of W1Se3 clusters is favored at
growth process; as a result, cross-contamination between dif- the edge of MoS2, forming seamless in-plane HSs. For a high
ferent atoms was often observed during the growth of HSs. To W/Se ratio, W1Se2 was prevalent on the surface of the MoS2
address this problem, Li et al. proposed a two-step CVD method layers and resulted in vertical stacking growth. Furthermore,
to prepare WSe2/MoS2 HSs on sapphire substrates containing Zhang et  al.[62] developed a controllable one-step growth
an atomically sharp interface.[70] Following this, Zhang et  al. method using WO3−x/MoO3−x nanowires as the experimental
and Johns et  al. used the same two-step method to synthesize precursor to successfully synthesize stacking heterobilayers
stacking WS2/MoS2 HSs on SiO2/Si substrates[71] and sapphire based on TMD monolayers.
substrates.[72] Unfortunately, cross-contamination of W and Mo First, MoO3−x was deposited on WO3−x nanowires to prepare
atoms was nonetheless observed. In this regard, Shi et  al.[23] the precursors (WO3−x/MoO3−x nanowires) for further experi-
designed an optimal growth-temperature-mediated CVD ments. These precursors possessed a core–shell structure in
method and successfully prepared HSs (MoS2/WS2 and WS2/ which WO3−x served as the core and MoO3−x acted as the shell
MoS2) in which atomic cross-contamination was not observed. (as shown in Figure  4f). Subsequently, gas-phase sulfurization
The authors employed Au foil as a metal substrate for was conducted by injecting sulfur gas into a furnace containing
growth because complete layered growth was achieved for the core–shell structure. In this process, the critical tempera-
MoS2 and WS2 domains on the foil, even at considerably dif- ture for sublimation of WO3−x is significantly higher than that
ferent temperatures. Two layers of composite MX2 monolayers of MoO3−x, ensuring the sublimation of MoO3−x and WO3−x in
were continuously grown on Au foil using two typical CVD different stages, leading to the sequential growth of MoS2 and
growth routes, and vertically stacked WS2/MoS2 (MoS2/WS2) WS2 heterojunctions. A detailed explanation of the process and
HSs were prepared naturally (as illustrated in Figure  4a). the reactions involved were reported by Humberto et  al.[83,84]
According to previous studies by Zhang et  al.,[73,74] at Remarkably, the results showed that the stacking orienta-
≈680 and 880 °C, the MX2 monolayers (MoS2 or WS2, respec- tions and configurations of HSs can be accurately modulated
tively) were synthesized on Au foil substrates using high- by proper design of the precursors. In particular, this strategy
temperature chemical reactions. Thereafter, further WS2 or opens up new avenues for producing a variety of novel 2D HSs
MoS2 monolayers grew over the substrate-covered, single-layer by rationally designing hetero-nanomaterials as experimental
sample. Finally, vertically stacked HSs (WS2/MoS2 and MoS2/ precursors.
WS2) were synthesized using a growth-temperature-mediated
two-step CVD method with high controllability and selec-
tivity. Furthermore, Fu et  al.[64] successfully synthesized high- 2.2.3. Substrate Engineering
quality metal/semiconductor TMD HSs (NbS2/MoS2) with the
assistance of halides, through temperature control, as shown During the synthesis of 2D HSs through CVD growth, the
in Figure  4b, effectively negating the shortcomings of metal surface of the catalytic substrate plays an important role in the
oxide precursors. Interestingly, when the top NbS2 layer nucle- decomposition of molecules in the inert gas stream and in pro-
ated on the bottom MoS2 layer, it gradually advanced from the moting atomic nucleation.[85–87] However, once the surface of
three corners to the center, which may be attributed to the the catalytic substrate is completely covered by the first mono­
higher growth temperature of NbS2 (schematically shown in layer of material, the decomposition, and nucleation rate of the
Figure  4c). Recently, Wu et  al.[65] demonstrated that vertically raw materials decrease significantly, inhibiting the growth of
stacked HSs (WSe2/WS2) can be synthesized with controllable subsequent layers. To solve such problems, researchers have
commensurate crystallographic alignments (so-called 0° AA devoted extensive research efforts to substrate development. For
and 60° AB stacking, as shown in Figure  1),[71] by regulating example, Zhang et al.[63] succeeded in designing a new type of
the deposition temperature (Figure  4d). These results indi- alloy substrate by exploiting the difference in adsorption ener-
cate that two-step CVD methods can be effectively optimized gies of Re and W atoms on the surface of Au(111), and utilized
by controlling the temperature, which can effectively suppress this substrate to prepare vertically stacked ReS2/WS2 HSs in a
cross-contamination of elements in stacking HSs. highly controllable manner (as shown in Figure  4g). Twinned

Adv. Mater. 2021, 33, 2005735 2005735  (5 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

growth behavior is a powerful approach to prepare vertically growth) in the following seven ways: controllability, simplicity,
stacked HSs. industrial scalability, homogeneity, cost effectiveness, fabrica-
The adsorption energy of Re and W atoms on Au(111), tion rate, and universality. This figure clearly shows that the
and the adsorption energy of Re atoms on WS2 were cal- mechanical transfer method originating from tape stripping
culated (EadsRe, Au(111)  = 1.52  eV, EadsW, Au(111)  = 2.77  eV, and demonstrates superior controllability, faster assembly rate, and
E adsRe,WS2 (001) = 3.04 eV), as shown in Figure 4h). The adsorption greater general applicability compared to the other techniques.
energy of the W atoms on Au(111) is relatively large; whereas the However, stacking HSs grown by the CVD methods possess the
adsorption energy of the Re atoms is considerably lower, leading best homogeneity, implying its possible superior performance.
to two distinct situations. W atoms can easily nucleate on Au(111) Notably, the solution-based synthesis strategy, owing to its cost
to generate WS2, whereas the nucleation of Re is difficult. Nev- effectiveness, low operational difficulty, and large sample size,
ertheless, once WS2 is generated on Au(111), and as a result of is currently more widely used in laboratories (Figure 5b). This
the strong adsorption energy (3.04  eV) of Re on WS2(001), the method is convenient for large-scale production of TMD heter-
Re atoms can nucleate on the surface of WS2 to form ReS2. ojunctions; nonetheless, it also displays some disadvantages in
Therefore, a ReS2/WS2/Au heterojunction with a completely that the prepared samples are prone to wrinkling and contami-
stacked structure can be successfully synthesized using one-step nation. In summary, when one intends to assemble a multilay-
CVD growth. This ingenious use of the difference in absorption ered heterojunction in a highly controllable manner, containing
energies of atoms can be extended to the preparation of other designed stacks, the tape-exfoliation method is preferred to
stacking HSs, and can significantly improve the controllability CVD growth. If a higher-quality HS must be synthesized, the
and feasibility of the fabrication process. It is worth mentioning CVD growth method may be the best choice. Although a variety
that recently, using liquid metal substrates to prepare HSs has of methods are available for the preparation of stacking TMD
received increasing attention (as shown in Figure  4i),[67] which heterobilayers, these methods are not perfect, and there are
may be a primary development direction for the one-step CVD various shortcomings that require urgent improvement. One of
preparation of heterojunctions.[88–90] the most prominent is that to date, it is extremely challenging
to prepare stacking HSs in a well-controlled manner. One can
achieve high-precision control, but not complete control.
2.3. Comparison and Selection of Preparation Methods

This section provides comprehensive comparisons and related 3. Characterization of Stacking-Designed Bilayers
guidelines for different preparation methods for the produc-
tion of stacked heterojunctions mentioned above. The various A comprehensive and detailed understanding of the stacks
preparation technologies to produce stacking heterobilayers in and the resulting intra- and interlayer interactions is crucial
TMD materials each possess their own characteristics, based for adjusting the physical properties of the material. In this
on the resultant stacking structures and special applications. section, a brief introduction to several mainstream characteri-
Figure 5a compares the aforementioned three types of prepara- zation methods, such as Raman spectroscopy, PL, SHG, and
tion strategies (i.e., tape exfoliation, solution etching, and CVD STEM, is provided and typically observed features are analyzed.

Figure 5.  Comprehensive comparison of the various fabrication strategies: a) comparison of the characteristics of three types of preparation methods,
namely, tape-exfoliation, solution-etching, and CVD methods. The tape-exfoliation, solution-etching, and CVD growth are represented by the orange,
purple, and blue areas, respectively. The data in pie chart (b) were collected and compiled from the top 500 most cited TMD-preparation-related articles
in the Web of Science database.

Adv. Mater. 2021, 33, 2005735 2005735  (6 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

3.1. Raman Spectroscopy Raman peak intensity at a particular wavenumber. Figure  6c


shows that the Raman intensity of the peak at 309.4 cm−1 can
3.1.1. High-Frequency Raman Spectroscopy be obtained by scanning along the dotted white line. Noticeably,
multiple peak fluctuations occur for the WSe2/WS2 bilayer, but
As a rapid, sensitive, and characteristic-optical detection not for the WS2 monolayer, indicating interlayer coupling in
method, Raman spectroscopy has been extensively employed WSe2/WS2 HSs.
to confirm the layer number, stacking modes, and interlayer Moreover, different intralayer Raman vibrational modes
coupling of 2D materials. In recent years, the Raman vibra- can identify the stacking sequence and thickness of the A
tion behavior of high-frequency (HF) modes in stacking mate- and B layers in predesigned A/B and B/A heterojunctions.
rials has been systematically studied and widely discussed. For example, Shi et  al.[23] accurately identified the stacking
For example, Wang et al.[91] explored the interlayer coupling of sequences and thicknesses of the WS2 and MoS2 layers in
WSe2/WS2 bilayers, with different stacking orientations, using HSs by discriminating the differences in the Raman spectra of
optical spectroscopy. Figure 6a shows a clear optical contrast WS2/MoS2 and MoS2/WS2. The HF Raman spectra (Figure 6d)
between the junction regions, individual monolayer regions, acquired from the stacking bilayer (as shown in Figure  4a)
and the Si/SiO2 substrate, while Figure  6b shows the typical present four discrete peaks that correspond to the E12g and
HF Raman spectra of WSe2/WS2 bilayers and their constituent A1g vibration modes of the individual MoS2 and WS2 layers.
monolayers. Evidently, the Raman spectrum of the WS2/WSe2 Remarkably, the A1g peak for MoS2 (≈405.2 cm−1) is attributed
bilayer is a composite of the spectra for individual WS2 and to MoS2/WS2/Au and is blue-shifted relative to the A1g peak
WSe2 monolayers; however, the presence of additional features for MoS2 (≈404.2 cm−1), attributed to WS2/MoS2/Au. This blue-
indicates interlayer coupling. For example, the shoulder in the shift of the A1g peak in MoS2/WS2/Au directly indicates the
A1g peak is usually observed at 261 cm−1 in the WSe2 monolayer, weakened n-doping effect in MoS2 caused by Au foils.[94,95] In
whereas in the WSe2/WS2 bilayer, this shoulder is red-shifted short, the blue-shift of the A1g Raman peak demonstrates that
by ≈2 cm−1. This spectral red-shift usually occurs as the number the MoS2 monolayer does not grow directly on the Au foil, but
of WS2 layers increases,[92,93] indicating that interlayer coupling grows on WS2/Au, further indicating the stacking sequence
occurs between the WS2 and WSe2 layers. Moreover, a new peak in layered structures. Although HF Raman spectroscopy is a
was observed at 309.4 cm−1 for the heterobilayer, which usually powerful tool to study the heterojunction formation, its effec-
only appears in multilayer WSe2,[93] further verifying inter- tiveness is reduced when exploring the quality of layer con-
layer coupling in the WSe2/WS2 bilayer, analogous to the WSe2 tact in hybridized 2D materials, for example, twisted stacking.
bilayer. Additional information regarding intralayer modes Nayak et al.[96] investigated the Raman spectra of samples with
in a particular sample can also be obtained by measuring the different rotation angles, probing the relationship between

Figure 6.  HF Raman spectra of stacking-engineered HSs: a) optical microscopy images and b) Raman spectra for a WSe2 monolayer, WS2 monolayer,
and WSe2/WS2 bilayer on a SiO2/Si substrate. c) Line scan profile of the intensity of the Raman peak (at 309.4 cm−1) obtained from the WSe2/WS2
bilayer sample along the dotted white line. d) Raman spectra exhibiting the stacking sequences of WS2/MoS2/Au, MoS2/WS2/Au, MoS2/Au, and
WS2/Au. e) Scanning electron microscopy (SEM) images of MoSe2/WSe2 HSs, twist range 0° ≤ θ ≤ 60°. The lower contrast bottom layer corresponds
to the WSe2 layers, and the higher contrast upper layer corresponds to the MoSe2 layers. The dark area is the overlap of the two. f) HF Raman spectra
of heterobilayers with different twist angles and an individual monolayer. a–c) Reproduced with permission.[91] Copyright 2016, American Chemical
Society. d) Reproduced with permission.[23] Copyright 2016, Wiley-VCH. e,f) Reproduced with permission.[96] Copyright 2017, American Chemical Society.

Adv. Mater. 2021, 33, 2005735 2005735  (7 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

interlayer excitons and twisted angle (0–60°, Figure  6e) in the rotating system in order to investigate the sensitivity of
stacking HSs (MoSe2/WSe2). The authors found that the HF the LBM to the interfacial conditions. The spatial profile of the
Raman vibrational modes were twist-angle-independent (as LBM of MoS2/WSe2 HS was mapped, and as expected, the LBM
shown in Figure  6f). In vertically stacked MoSe2/WSe2 HSs, was only observed in the overlapping region of adjacent TMD
there was no obvious correlation between the Raman vibration monolayers (Figure  7d). Figure  7e displays the LBM frequen-
modes mentioned above and the rotation angles, and the fre- cies of 58 heterobilayer samples, with different stacking orien-
quencies constantly remained within ± 1 cm−1. tations, as a function of the layer rotational angle (θ). Although
the LBM frequencies varied somewhat irregularly, due to dif-
ferent sample conditions (defects, stresses, interfaces, etc.),
3.1.2. Low-Frequency Raman Spectroscopy the average value decreased from 33.5 to 30.7 cm−1 when the
rotational angle increased from 0° to 60°. A slight decrease in
Compared to HF Raman modes, low-frequency (LF) Raman LBM frequency with increasing stacking angle indicates that
modes are more sensitive and are used as fingerprints to probe the interlayer coupling of HSs becomes slightly weaker, con-
different stacking configurations, order, and contact quality of sistent with the previous theoretical results obtained by Kang
hybridized layered materials.[97–100] The layer-breathing modes et  al.[104] As the angle increases from 0° to 60°, the interlayer
(LBM) and shear modes (SM) at low frequencies (<50 cm−1) distance becomes slightly larger while the interlayer absorbed
reflect the out-of-plane and in-plane vibrations of adjacent energy becomes slightly smaller. In addition, the LBM phonon
layers (as shown in Figure 7a). Figure 7b shows the LF Raman mode can also characterize subtle differences in the interlayer
characterization of MoSe2/WSe2 HSs with various twisted coupling of adjacent monolayers. As shown in Figure  7f, the
angles and individual monolayers, while Figure  7c shows the LBM peaks are located at 31.6, 28.3, 31.2, and 26.7 cm−1 for
normalized Raman shifts of the SM and LBM. Notably, as the the as-grown MoS2 films on WS2, AA, AB stacking HSs, and
relative angle between the two layers varied, the position of transfer-twisted HSs, respectively.[71] Evidently, the LBM fre-
the SM modes remained at 18.1 cm−1, whereas the position of quency of AB stacking HSs increased by 2 cm−1 compared with
the LBM changed continuously between 21 and 25 cm−1, dem- AA stacks, attributed to the slightly stronger interlayer coupling
onstrating that the LBM may be related to the stacking angle in AB stacking HSs. Unusually, the rarely reported SM Raman
of the heterobilayers. In addition, the changes in peak intensi- peak can be observed at ≈18 cm−1, and shifts in this peak are
ties can be attributed to a change in the lattice symmetry, in apparent, which may be related to the almost perfect crystal-
line with previous studies.[101,102] Furthermore, Lui et  al.[103] lographic alignment in the sample. These results demonstrate
used LF Raman to perform a detailed statistical evaluation of that the LBM in LF Raman spectra can serve as an efficient and

Figure 7.  LF Raman spectra of twist-stacking HSs: a) schematic of the lattice structure and vibrational modes for bilayer HSs. b) LF Raman spectra of
twisted MoSe2/WSe2 HSs and individual monolayers, obtained with a 532 nm laser. The positions of the SM and LBM are marked with dashed lines.
c) Normalized Raman shifts of the SM and LBM in part (b). d) Optical image and corresponding Raman intensity map of the LBM at ≈32 cm−1 for
MoS2/WSe2 HSs. e) LBM frequencies of 58 MoSe2/MoS2 samples as a function of the layer rotational angle (θ). Inset: schematic illustration of the
twisted angles between MoSe2 and MoS2 monolayers. f) LF Raman spectra of LBM and SM for as-grown continuous MoS2 films on WS2, AA and AB
stacked, and transfer-twisted stacked MoS2/WS2 HSs, respectively. a,f) Reproduced with permission.[71] Copyright 2015, Wiley-VCH. b,c) Reproduced
with permission.[96] Copyright 2017, American Chemical Society. d,e) Reproduced with permission.[103] Copyright 2015, American Physical Society.

Adv. Mater. 2021, 33, 2005735 2005735  (8 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

sensitive probe to detect differences in the stacking configura- originates from charge separation between the monolayers as
tions of vertical TMD heterojunctions. a result of the formation of a type II heterobilayer, as presented
in Figure  8b.[105–108] Another possible source of abnormal
intralayer exciton activity is the quality of the sample itself. As
3.2. Photoluminescence shown in Figure 8c, defects and intrusions are common in the
heterojunctions formed during the stacking process (assembly
PL is a highly efficient means of characterization and has or growth), and thus the interlayer distance varies as a func-
been extensively used to study the intra/interlayer exciton tion of position, so that excitons behave abnormally. Such trap-
activities of stacking-designed heterobilayers. The brightness intervening and abnormal exciton transfer is evident in the
and color of PL maps can be used to discern the number of time- and energy-resolved PL spectra following pulsed excita-
stacking layers, stacking configurations, and the dependence tion (Figure  8d).[109] The initial high energy is attributed to a
of stacking-engineered TMD heterobilayers. As shown in combination of a density-dependent blue-shift and hot inter-
Figure 8a, the individual WS2, WSe2 monolayer, and overlapped layer exciton emission during diffusion. The observed red-shift
WS2/WSe2 bilayer regions on the substrate can be accurately and the narrowing of the PL tail are attributed to a dynamic
identified by comparing the colors and relative brightness of reduction in the local interlayer exciton density as well as spa-
the PL images.[91] Strong PL emissions can be observed for tial relaxation of excitons to the local minima resulting from
individual WSe2 monolayers and WS2 monolayer regions, traps or interfacial contaminants.[29,109]
while quenching of the PL occurs in the overlapped region. The Moreover, PL can also be used to investigate twist-angle-
quenching of both interlayer PL emissions in overlapped HSs dependent heterobilayers. Alexeev et al.[110] used PL to conduct

Figure 8.  Stacking-dependent PL: a) PL images of WSe2/WS2 heterobilayers. b) Schematic illustration of type II band structure alignment in MoS2/
WS2. c) Schematic of WSe2/MoSe2 HSs with adsorbates in the adjacent layers that modulate the interlayer distance and the corresponding exciton
binding and emission energies. d) False-color plot of the time- and energy-resolved PL spectrum for interlayer excitons in the MoSe2/WSe2 HS. e) PL
spectra of WS2 and MoSe2 monolayers, and MoSe2/WS2 HSs assembled with stacking angle θ = 2°, and the two PL peaks of the heterojunction: P1
and P2. f) Normalized PL spectra of the P1 shoulder captured for MoSe2/WS2 HSs with twist angle from 1° (red) to 59° (blue). A typical PL spectrum
of an individual MoSe2 monolayer is marked by the black dotted curve. g) Phonon energy as a function of stacking angles, summarized from part (f).
The blue curve represents the results of theoretical calculations for the samples, and the dashed line represents the position of XA in individual MoSe2
monolayer. h) WS2 and MoSe2 band structures and Brillouin zone alignment for twist angle θ. Spin-up (spin-down) bands are colored green or red
(gray). i) Twist-angle (θ) dependence of interlayer exciton PL intensities for MoSe2/WSe2 HSs. a) Reproduced with permission.[91] Copyright 2016,
American Chemical Society. b) Reproduced with permission.[107] Copyright 2019, American Chemical Society. c,d) Reproduced with permission.[109]
Copyright 2017, IOP Publishing Ltd. e–h) Reproduced with permission.[110] Copyright 2019, Springer Nature. i) Reproduced with permission.[96] Copyright
2017, American Chemical Society.

Adv. Mater. 2021, 33, 2005735 2005735  (9 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 9.  Principle of SHG for artificially stacked bilayers: a) schematic of the SHG signal through the Brillouin zone. b) Optical image of a MoS2 bilayer
with twist angle θ = 25°. Scale bar is 5 µm. c) Schematic illustration of the polarization directions of the incident laser and the measured SHG. The angle
between the x-axis laser polarization is defined as ϕ, parallel to the armchair orientation of the layers. Polar plots of the SHG signal captured from d)
the individual monolayer, region (1), e) monolayer, region (2), and f) overlapped stacking, region (3), respectively. a) Reproduced with permission.[117]
Copyright 2017, American Chemical Society. b–f) Reproduced with permission.[118] Copyright 2014, American Chemical Society.

an in-depth investigation into the correlation between exciton interlayer conduction-band tunneling, termed the interlayer
transfer and stacking angles by studying a series of vertical exciton hybridization effect.[27,111–113] Thus, the red-shift of P1
HSs. As shown in Figure  8e, two peaks were observed in the in Figure 8f can be assigned to resonantly hybridized excitons
PL spectra of MoSe2/WS2 HSs (stacking angle = 2°): P1 and in moiré superlattices. Moreover, the results of Nayak et al.,[96]
P2, which were located at energies close to the exciton peaks who explored the evolution of twist-angle-dependent inter-
of MoSe2 and WS2 monolayers, respectively. P2 was almost layer excitons in MoSe2/WSe2 HSs, are also approximately in
identical to that of the WS2 monolayer (≈1.96  eV), attributed line with the above (Figure  8i). The high PL intensity in the
to strong signals originating from the surroundings that are high-symmetry stacking configurations can be attributed to
covered with a large number of WS2 monolayers. In contrast, increased interlayer coupling strength, improving the charge
P1 was significantly red-shifted relative to that of the MoSe2 transfer efficiency.
monolayer, with a significant decrease in the peak intensity.
Therefore, a follow-up investigation on the P1 peak of HSs with
different twisted angles was conducted. From the normalized 3.3. Second Harmonic Generation
PL spectra (Figure  8f,g) of P1 (rotation angle from 1° to 59°),
the P1 phonon energy as a function of the stacking angle can SHG is a powerful characterization technique for non­
be plotted and is shown in Figure  8g. Two distinct phases are centrosymmetric layered materials owing to its noncontact,
clearly observed: one where there is a steep variation in the peak simple, and versatile operation.[114–116] In this section, SHG char-
energy (≈60 meV) as the twist angle is close to either 0° or 60°, acterization of the layer-dependent crystal symmetry in TMD
and another where a plateau can be observed at ≈1.56  eV for materials and their applications as well as artificial stacking
large misalignment angles (from 8° to 52°). These variations in configurations are discussed. SHG is essentially a second-order
the photon energy point to misalignment of the Brillouin zones nonlinear optical process that can only be performed in mate-
in monolayers arising from their rotation, and this misalign- rials with broken inversion symmetry. A beam of light with
ment introduces a momentum mismatch between the valence- frequency ω is passed through the noncentrosymmetric mate-
and conduction-band edges of the adjacent layers.[111] As shown rials (see Figure 9a), in which two photons with frequency ω
in Figure  8h, for perfect lattice alignment (antialignment), the are quenched during harmonic generation and simultaneously
K valley mismatch, ∆K = K WS2 − K MoSe2 ( ∆K ′ = K WS′ 2 − K MoSe2 ) , generate a new photon with frequency 2ω. Significant features
is minimized, enabling the excitons to hybridize through of TMD heterobilayers such as optical valley polarization and

Adv. Mater. 2021, 33, 2005735 2005735  (10 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 10.  SHG characterization of stacking TMDs heterobilayers: a) polar plots of the SHG signal measured for the WSe2 monolayer (green circles) and
WS2 (yellow circles) regions, from which the rotation angle between the two layers is determined to be 0.5 ± 0.3°. SHG maps for the WSe2/MoSe2 HSs with
well-defined b) AA and c) AB stacking configurations; schematics of the corresponding atomic alignments are shown above the images. d) SHG maps for
the AA- and AB-stacking WSe2/WS2 HSs. e) Polarization-resolved SHG for a WS2 monolayer, AA-, and AB-stacking WSe2/WS2 HSs shown in (d). f) SHG
phase-resolved spectra for the stacked WSe2/MoSe2 HSs with a twist angle of ≈60°. g) SHG intensities of twisted WSe2/MoS2 as a function of the rotational
angle; Δf represents the rotational angle between layers. a) Reproduced with permission.[121] Copyright 2019, Springer Nature. b,c) Reproduced under the
terms of the CC-BY 4.0 license.[119] Copyright 2018, The Authors, published by Springer Nature. d,e) Reproduced with permission.[65] Copyright 2019, Springer
Nature. f) Reproduced with permission.[122] Copyright 2019, Springer Nature. g) Reproduced with permission.[120] Copyright 2017, American Chemical Society.

nonlinear optical effects, obtained from the layer-dependent the SHG signal of the stacked region can be interpreted as a
crystal symmetry of the materials, can be easily revealed by coherent superposition of the second-harmonic fields obtained
SHG. from different individual monolayers.[118]
TMD heterobilayers exhibit a strong layer-stacked depend- In practical applications, characteristic petal-like SHG pat-
ence on the SHG response. Some bilayers possess inversion terns are clearly observed for WS2 and WSe2 layers (Figure 10a),
symmetry and some do not, producing different second-order from which the rotation angle between the two layers is deter-
optical nonlinearities. As a result of this layer-stacked depend- mined to be 0.5  ± 0.3°. Moreover, as shown in Figure  10b,c,
ence, the TMD samples can be clearly distinguished from one Hsu et  al.[119] presented the SHG intensity maps of WSe2/
another using SHG. For instance, Hsu et  al.[118] studied the MoSe2 HSs with well-defined atomic arrangements: AA and AB
SHG of an artificially stacked MoS2 bilayer with various twist stacking (3R-like and 2H-like stacking, as shown in Figure  1).
angles, in which the optical waves were a coherent super- Notably, the SHG signal from the stacking region with an AA
position of the second harmonic field from each individual configuration is much more intense than that of the AB con-
layer. The results showed significant phase contrasts due to figuration, attributed to the enhanced SHG effect of the AA
the different stacking configurations of distinct samples (see configurations; this SHG signal was strongly suppressed in
Figures 9b,c). By using a back-reflecting geometry, the laser was the WSe2/MoSe2 with AB configuration. This phenomenon is
incident on the specimen along the x-direction (Figure 9c), par- more pronounced in the SHG characterization of WSe2/WS2
allel to the armchair orientation of the monolayer. According to HSs.[65] As shown in Figure  10d, AA and AB stacking WSe2/
the polar plots (Figure  9d,e) of the polarization-resolved SHG WS2 HSs exhibit vastly different SHG intensities. However, the
intensity as a function of the angle ϕ, the single flake exhibits SHG signal for AB stacking does not disappear completely (see
six-lobed features, similar to a blooming flower, and the petal- Figure  10e). This is because the monolayer of the grown AB-
like patterns lie along the vertical bisector of individual flakes. stacking WSe2/WS2 HSs still possess different second-order
The SHG signal captured from the stacking area also exhibits nonlinear coefficients, resulting in lattice mismatch. Accord-
a flower-like pattern with six-petal features (see Figure  9f), ingly, the AA-stacking configurations of TMD heterostacks can
proving that the SHG signal produced from the overlapped be clearly distinguished from the AB-stacking configurations.
region is not contributed solely by a single layer. In short, Moreover, SHG signals obtained by collecting and transforming

Adv. Mater. 2021, 33, 2005735 2005735  (11 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 11.  Atomic configurations of stacking-engineered HSs revealed by microscopy: a) close-up ADF-STEM images of four different high-symmetry
regions in MoS2/WSe2 moiré superlattices, labeled as AA, ABSe, bridge (Br), and ABW, and the corresponding atomic models. Conductive atomic
force microscopy images of b) R-type and c) H-type MoSe2/WSe2 heterobilayers with small twist angle. The domains present in the R-type stacking
superlattice are alternating triangular in shape, whereas the domains in H-type stacking superlattice are hexagonal. d) Diagram of the acquisition of
EEL-STEM spectra. The insert is rotation-angle-dependent EEL spectra of MoS2/WSe2 heterobilayers, in which the peak width changes significantly
as the twist angle increases from 0° to 30°. a) Reproduced with permission.[125] Copyright 2017, American Association for the Advancement of Sci-
ence. b,c) Reproduced with permission.[123] Copyright 2020, American Chemical Society. d) Reproduced with permission.[30] Copyright 2019, American
Chemical Society.

data produced by phase-resolved SHG equipment can also shows the atomically resolved annular dark field (ADF)-STEM
be translated to form a spectrum, as shown in Figure  10f (for images of MoS2/WSe2 HSs containing four differently aligned
60° stacking angle WSe2/MoSe2 HSs). In addition, the SHG regions, denoted as AA, ABSe, bridge (Br), and ABW. Therefore,
intensity can be parsed as a function of the rotational angle these four types of stacking configurations can be clearly distin-
for normalized observation and comparison (Figure  10g).[120] guished by the specific moiré pattern produced. Microscopes
In summary, SHG is a sensitive, efficient, and nondestructive are particularly effective when exploring the atomic interac-
characterization tool to determine the stacking configurations, tions of heterojunctions containing different stacking configu-
domain boundaries, and crystal polarities of HSs constructed rations. Recently, Rosenberger et  al.[123] used conductive atomic
from noncentrosymmetric layered materials. force microscopy to show that the moiré superlattices of MoSe2/
WSe2 and MoS2/WS2, at twist angles of ≤1°, undergo significant
atomic-level reconstruction, resulting in the formation of dis-
3.4. Other Microscopy Characterizations crete domains divided by regular side walls.[123] The domains in
the R-type stacking superlattice are alternating and triangular in
Compared to Raman, PL, and SHG, STEM is a more powerful shape (Figure 11b), whereas the domains in the H-type stacking
and intuitive technique for characterizing the microstructure of superlattice are hexagonal (Figure  11c). This split domain phe-
nanomaterials. A focused electron beam is scanned across the nomenon can be attributed to a potential energy gradient in the
surface of the material, and a certain number of the electrons superlattice, arising from interlayer stacking, consistent with
are either transmitted or scattered. The transmitted or the elas- the theoretical predictions of Stephen et al.[123,124] With regard to
tically scattered electron beams are subsequently captured by rotating heterobilayer systems, the difference in atomic position
a prepositioned detector, and after processing and amplifica- caused by twist can be shown using electron energy loss (EEL)
tion, a bright field or dark field image is obtained. Figure 11a spectroscopy. As illustrated in Figure  11d, an incident electron

Adv. Mater. 2021, 33, 2005735 2005735  (12 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 12.  Stacking order effects on vertical heterobilayer devices: a) schematic of the photocatalytic hydrogen evolution reaction (HER) with TMD
materials as electrodes. b) Photocatalytic H2 evolution curves of MoS2/Au, WS2/Au, WS2/MoS2/Au, and MoS2/WS2/Au with increasing irradiation
times. c,d)  Schematic illustrations of electron transfer occurring in MoS2/WS2 and WS2/MoS2 vertical stacks, respectively, under light irradiation.
Compared to WS2/MoS2 bilayers, MoS2/WS2 stacks exhibit a higher electron–hole separation efficiency and electron transfer rate, responsible for its
superior photocatalytic activity. e,f) Short-circuit currents (ISC, black) and open-circuit voltages (VOC, red) of e) GrB/WS2/MoS2/GrT and f) GrB/MoS2/
WS2/GrT vertical devices as a function of the gate voltage. Illustrations of the stacking order in the devices are shown above the graphs. g,h) Schematic
illustrations of type-II band alignments in GrB/MoS2/WS2/GrT and GrB/WS2/MoS2/GrT under positive source–drain bias. a) Reproduced with permis-
sion.[24] Copyright 2017, Wiley-VCH. b–d) Reproduced with permission.[23] Copyright 2016, Wiley-VCH. e–h) Reproduced with permission.[34] Copyright
2019, American Chemical Society.

beam illuminates the surface of the MoS2/WSe2 heterobilayers fascinating properties and providing a convenient way to arbi-
and is elastically scattered. The value of the partial energy loss trarily tune various properties for use in different types of
of electrons is the characteristic value of an element in the HSs, nano­ devices. The stacking sequence of heterobilayers intro-
thereby allowing acquisition of the EEL spectrum. Notably, Gogoi duces a new degree of freedom that can be useful for tuning
et  al.[30] used EEL spectroscopy to explore layer rotation-angle- the properties of optoelectronic devices as well as stacking
dependent excitonic activities, where the EEL spectral peak configurations.[32,33,126] For example, in photoelectrocatalysis,
widths are a function of the twist angles (Figure  11d), revealing Shi et al. demonstrated that WS2/MoS2 and MoS2/WS2 vertical
that the interlayer exciton transfer rate in coherent stacks is one stacks are ideal candidates for the hydrogen evolution reaction
order of magnitude higher than that in misaligned cases. (HER).[23] Figure 12a shows a schematic illustration of the pho-
tocatalytic HER using TMD heterobilayers as electrodes. The
H2 evolution rates of the WS2/MoS2 and MoS2/WS2 vertical
4. Stacking Sequences and Configurations stacks upon light irradiation are shown in Figure  12b, where
the MoS2/WS2/Au sample exhibited a higher rate, compared
Independent TMD monolayers, with diverse physical proper- to the reverse stacking order in WS2/MoS2/Au and other
ties, can be artificially stacked together to form HSs in a specifi- samples. In brief, MoS2/WS2 or WS2/MoS2, with opposite
cally designed sequence or configuration. As a result, the newly stacking orders, exhibited significantly different photocatalytic
formed heterojunction allows integration of the physical proper- activities upon light irradiation, and such differences can be
ties of each individual separated layer, generating new function- explained in terms of electron transfer in type-II band align-
alities. In this section, we review the dependence of dynamics ments. Figure  12c,d shows diagrams of the charge transfer
on the stacking sequences and configurations, including twist mechanisms in vertically stacked heterobilayers under light
angles, moiré superlattices, and potential applications. irradiation. The conduction band minimum position in MoS2
is lower than that of WS2 in the WS2/MoS2 stacks, generating
energy barriers that block electron transfer from MoS2 to WS2,
4.1. Stacking Sequences weakening H2 evolution. However, in the inverse stacking
structure (MoS2/WS2), the type-II band alignment of MoS2/
TMD heterobilayers are constructed with different indi- WS2 facilitated efficient electron transfer from WS2 to MoS2,
vidual monolayers placed on top of one another, leading to thereby enhancing H2 generation.

Adv. Mater. 2021, 33, 2005735 2005735  (13 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 13.  Twists in TMD heterobilayers: a) atomic models of the TMD heterobilayers with a twist angle of 0°, 13.2°, 21.8°, 27.8°, 32.2°, 38.2°, 46.8°,
and 60°, respectively. The interlayer distance in HSs is indicated by (d). b) Calculated bandgaps (top) and interlayer distances (bottom) of CrSe2/MoS2.
c) Illustration of the different stacking angles in devices (0°, 15°, and 30°), and the variation of the external quantum efficiencies with twist angles under
infrared illumination. d) Interlayer spin-valley physics of WSe2/MoSe2, showing spin-valley polarized resonant excitons and the valley-dependent optical
selection rules. a,b) Reproduced with permission.[127] Copyright 2017, Royal Society of Chemistry. c) Reproduced with permission.[102] Copyright 2018,
American Chemical Society. d) Reproduced under the terms of the CC-BY 4.0 license.[128] Copyright 2016, The Authors, published by Springer Nature.

TMD heterobilayers with inverse stacking sequences also 4.2. Stacking Configurations
show promise in optoelectronics, particularly photoelectronic
devices. Recently, Zhou et  al.[34] demonstrated that the stack 4.2.1. Twists in Heterobilayers
sequence played an important role in tuning the photocurrent
direction and photovoltaic effects under light irradiation, The exciton dynamics in twisted heterobilayers have been
according to experiments performed with WS2/MoS2 and studied by density functional theory (DFT) computa-
MoS2/WS2 samples. The absolute values of the short-circuit tions,[28,91,102] and various sample interlayer orientations can
current (ISC) and open-circuit voltage (VOC) as a function of be achieved with advanced preparation processes.[43,66] By arti-
VG are shown in Figure 12e,f. For the WS2/MoS2 nanodevice, ficially blocking TMD monolayers, heterobilayers with various
when VG was increased from −60 to 60 V, ISC and −VOC mono- twist angles can be obtained. Therefore, it is important to under-
tonically decreased from 120 nA and 2.4  mV to 30 nA and stand the intrinsic mechanism of interlayer rotation in order
0.7  mV, respectively. However, for the MoS2/WS2 structure, to tune the optoelectronic properties of 2D-nanodevices. The
the maximum photovoltaic effect occurred at VG = −60 V and periodic unit cells of twisted HSs exhibit clear differences as a
gradually decreased with increasing VG. This difference can result of various real space stacking orientations (Figure 13a).
also be attributed to the variety of electron transfer with dif- According to these models, Lu et  al.[127] performed first prin-
ferent band alignments. Figure 12g,h demonstrates that in the ciple calculations on CrSe2/MoS2 HSs. Figure  13b shows that
GrB/MoS2/WS2/GrT device, the direction of the type-II band an obvious transition from an indirect gap (Γ–K) to a direct gap
offset between MoS2 and WS2 changed under positive bias, (K–K) occurs as the rotation angle increases from 0° to 13.2°.
in contrast to that under zero source–drain bias, thus, the However, when the twist angles increased from 13.8° to 46.8°,
direction of the generated photocurrent in the device changed the bandgap remained constant at ≈0.76 eV. Moreover, both the
accordingly. However, although the direction of photocurrent 0° and 60° stacks exhibited the smallest bandgaps and layer
generated in GrB/WS2/MoS2/GrT coincided with that gener- distances, suggesting that they were the most stable stacking
ated in GrB/MoS2/WS2/GrT, a higher photocurrent was gen- configurations among the various rotations. Interestingly,
erated as a result of the larger band shift between WS2 and CrSe2/MoS2 HSs display type-I band alignment and are thus
MoS2. Therefore, TMD monolayers can be assembled with promising candidates for light-emitting diodes. Heterojunction
controllable sequences, allowing facile tuning of the charge devices fabricated with stacking-twisted structures show dif-
activities in heterobilayers. ferent photoelectric properties; this difference is closely related

Adv. Mater. 2021, 33, 2005735 2005735  (14 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 14.  The dependence of exciton transfer on twist angles: a) schematics of pump-induced exciton transient and valley alignment in MoS2/WS2 HSs. Upon
optical pump excitation, the electrons are rapidly transferred to MoS2, while holes are transferred to WS2, causing charge separation. Lastly, the interlayer
recombination is measured by probe light, and the transient absorption spectra b) of MoS2/WS2 HSs are obtained for coherent stacks and random stacks.
c) Lifetime, τ, of the excitons as a function of the rotation angles. d) Charge transfer (red shading) and recombination times (blue) in MoS2/WSe2 HSs with dif-
ferent twist angles. The insert is a schematic diagram of Brillouin zone alignment in twisted MoS2/WSe2. e) Transient absorption spectra of a MoS2 monolayer
and MoS2/WS2 HSs with three different stacking angles: 0°, 38°, and 60°. f) Theoretical concepts explaining the robust and ultrafast charge transfer in TMD
heterobilayers. Schematic of phonon-mediated charge transfer. Longitudinal acoustic (LA) phonons; Aʹ1, a specific optical phonon; transverse (TA) optical
phonons. The electron transfer between the K points is mediated by phonon coupling to the Q point, and the same for hole transfer, mediated by coupling
to the Γ point. a,b) Reproduced under the terms of the CC-BY 4.0 license.[28] Copyright 2015, The Authors, published by Springer Nature. c) Reproduced with
permission.[30] Copyright 2019, American Chemical Society. d) Reproduced with permission.[120] Copyright 2017, American Chemical Society. e) Reproduced
with permission.[35] Copyright 2017, American Chemical Society. f) Reproduced with permission.[132] Copyright 2017, American Physical Society.

to the stacking angle. As shown in Figure 13c, the I–V curve of same method, Heo et  al.[28] obtained the transient absorption
the sample rotated to 0° exhibited evident diode-like behaviors spectra (Figure  14b) of MoS2/WS2 HSs with two types of con-
and was sensitive to infrared illumination, while the 15° and figurations: coherent and random stacks. Evidently, the initial
30° samples were inactive.[102] This can be attributed to the fact decay was faster in the random stacks (τ1 = 3.3 ps) than in the
that the 0° HSs possessed excellent lattice matching, and as a coherent stacks (τ1 = 5.3 ps), which can be attributed to defect-
result the interlayer coupling was strong enough to form p–n assisted recombination in the random stacks.[130] Furthermore,
junctions. For the 15° and 30° samples, a van der Waals gap the slow decay component was reduced in the random stacks
occurred due to the formation of a potential barrier resulting (τ2 = 1.5 ns) in comparison to the coherent stacks (τ2 = 39 ps).
from larger lattice mismatch.[102] This was attributed to the direct bandgap of the coherent stacks,
Numerous groups also examined the nature of the spin- whereas the bandgap of random stacks was indirect, consistent
valley charge transfer (Figure  13d) for different types of TMD with other findings.[108,131] Gogoi et  al.[30] investigated the
heterobilayers, with a focus on the dependence on the twist rotation-angle-dependent exciton transfer in MoS2/WSe2 HSs
angles. Spin initialization is a crucial step for spin/valleytronic (Figure  14c), and found that aligned and antialigned configu-
devices that require net spin polarizations for reading, writing, rations, or so-called high-symmetry stacks, possessed a more
and transferring information.[129] Using pump–probe meas- transient lifetime compared to larger rotational mismatches.
urements, as shown in Figure 14a, immediately after optical The authors also demonstrated that momentum conservation
pump excitation, electrons were instantly transferred to MoS2, significantly enhances the charge transfer dynamics by approxi-
while holes were simultaneously transferred to WS2, causing mately one order of magnitude.
efficient charge separation. By gathering differential transmis- In sharp contrast to the above studies, the findings of Ji
sion data from the interlayer recombination using probe light, et  al.[35] indicated that the lifetime of charge transfer was
the transient absorption spectrum of MoS2/WS2 HSs was stacking-independent. As shown in Figure  14e, the signal
obtained, illustrating the electron transfer kinetics. Using the strength of MoS2/WS2 bilayers was approximately four times

Adv. Mater. 2021, 33, 2005735 2005735  (15 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 15.  Moiré-trapped interlayer excitons in heterobilayers: high-symmetry stacking configurations appear in moiré supercells with slight twists or
differences in the lattice constant: a) Hhh , Hhx , and HhM , and b) Rhh , Rhx , and RhM , respectively. c) Top and side views of three types of stacking configura-
tions, Rhh , Rhx , and RhM , and the corresponding exciton emission in the K valley after excitation by polarized light. Emission of exciton in the K valley at
the Rhh (Rhx ) site is left-circularly (right-circularly) polarized, however, at the RhM site, the emission is dipole-forbidden. d) Transient absorption signal
captured from 0° and 60° heterobilayers. Evidently, the transient absorption signal appeared to be weaker for the 60° sample. The energy landscape
of moiré potentials for 0° and 60° supercells allow trapping of the interlayer excitons (black and red spheres) in the local minima. 1, 2, and 3 indicate
e) Hhh , Hhx , and HhM , and f) Rhh , Rhx , and RhM sites. The position-dependent moiré potential governs the interlayer exciton dynamics. a,d–f) Reproduced
with permission.[137] Copyright 2020, Springer Nature. b) Reproduced with permission.[135] Copyright 2017, American Association for the Advancement
of Science. c) Reproduced with permission.[122] Copyright 2019, Springer Nature.

that of the MoS2 monolayer, however, when the MoS2/WS2 superconductivity in magic-angle graphene superlattices and
bilayers with different stacking angles were compared, the rise the extension of the h-BN system have recently caused con-
times varied slightly, indicating that the exciton transfer lifetime troversy in the physics community, leading to the emergence
was stacking-independent and robust.[28,30] The robust transfer of the twistronics field.[7,10] By artificially introducing small
lifetime resulted from the interlayer stretching/sliding of twists or layer mismatch in the heterojunctions, a long-period
stacked HSs, forming additional channels for exciton transfer. moiré pattern with various stacking configurations can be
The conclusions from Wang et  al.[91] insisted that interlayer formed (as shown in Figure 15a,b (Hhh , Hhx and HhM , Rhh , Rhx
charge transfer takes place within 450 fs, close to the pump and RhM )), dramatically changing the physical properties of the
pulse duration in their experiments, and therefore does not materials. A moiré supercell normally has a periodic size of
possess a measurable sensitivity to the stacking angle. Further 10  nm or greater, which varies as the twist angle or layer dis-
reinforcing this point, Zhu et  al.[120] explored twisted MoS2/ tance changes.[135,136] The different stacking configurations in
WSe2 HSs and found that the charge transfer signal appeared the moiré superlattice smoothly and periodically change with
within 40 fs, regardless of stacking angle (Figure 14d), and the position. The interlayer interaction is sensitive to the local envi-
decay timescale varied with the twist angle, however, no clear ronment. For instance, regarding the optical selection charac-
trend was apparent. The robustness of the charge transfer may teristics in the moiré pattern (Figure  15b), the exciton in the
be explained by phonon-mediated charge transfer. As shown Rhh (Rhx ) configurations only couples to σ+ (σ−) polarized light,
in Figure  14f, the proposed mechanism involves scattering by whereas the dipole in the RhM configuration is perpendicular to
phonons from the K valley to the Γ valley (for holes) or the Q the plane and cannot efficiently couple to the incident light, as
valley (for electrons), regions in which the interlayer coupling is shown in Figure 15c. Theoretical calculations showed that this
so strong that the charge transfer is undisturbed and rapid.[132] optical selectivity was derived from the difference in moiré
potentials at different sites. When σ+ and σ− polarized light is
alternated, the deeper moiré potential at the Rhh site can accom-
4.2.2. Moiré Superlattices modate additional and stronger interlayer exciton resonances.
These optical selections continuously change in the moiré
It is generally accepted that a moiré superlattice can effectively superlattice and are generally elliptically polarized.[122] Such
modulate the electronic band structure of layered materials to position-dependent excitons in the heterobilayers also display
form exclusive transmission characteristics, such as the fractal clear dependences on different types of stacking. Recently, Yuan
quantum Hall effect, topological-insulator behaviors, and Hof- et  al.[137] employed CVD-grown WS2/WSe2 HSs to investigate
stadter’s butterfly.[7,133,134] The observation of unconventional twist-angle-dependent interlayer exciton diffusion, and found

Adv. Mater. 2021, 33, 2005735 2005735  (16 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Figure 16.  Topology in moiré superlattice: a) schematic of the device, where the TMD monolayers possess a near-zero twist angle. b) Real-space
distribution of the exciton center-of-mass wavefunction. The three maps present the appearance of three different moiré minibands (states I–III) in
the strong-coupling regime (middle and bottom row). c) Left: local atomic registries are shown at three sites (R1, R2, and R3) in a moiré supercell,
which have negligible differences to TMD bilayers, as a result of interlayer translation. Different local sites are present in either the topological (red)
or normal (blue) insulating phase. The spin up (gray) and down (green) helical modes appeared at the boundaries of the topological/normal insu-
lating phases. Right: electrically switchable concentrated arrays of protected helical channels can be applied in field effect transistors. d) Energy map
of strain-engineered TMD bilayers with corresponding rotation between monolayers. The enlarged energy map shows five different energy states; an
imposed strain path is represented by the dashed white arrow and the relaxation path is indicated by the dashed gray arrow. The imposed strain path
drives state #1 to state #2 and state #3 to state #4. The system is spontaneously relaxed from state #2 to state #3 and from state #4 to state #5. a,b)
Reproduced with permission.[121] Copyright 2019, Springer Nature. c) Reproduced with permission.[139] Copyright 2016, Springer Nature. d) Reproduced
with permission.[140] Copyright 2019, American Chemical Society.

that the transient absorption signal appeared to be weaker for monolayer on an unstrained monolayer (Figure  16c).[139] The
the 60°-rotated HSs, owing to the faster charge transfer, which authors applied strain-regulated techniques to the band energy
arose from the difference in moiré potentials for 0° and 60° of the local moiré pattern. Strain engineering is a powerful tool
stacking (Figure 15d). The spatial variations in 0°-rotated sam- to tune the electronic signatures of TMD heterobilayers.[140]
ples (Figure 15f) are much stronger (deep potential) than those As shown in Figure 16d, controlling the interlayer rotation, by
for 60° samples (Figure 15e) (shallow potential).[137] Such moiré applying strain, led to the control of local energy in the moiré
potential-trapped valley excitons have also been demonstrated superlattice, offering another possibility for tuning moiré
in MoSe2/WSe2[138] and MoS2/WSe2[125] HSs. The presence of bands, which is topologically nontrivial.[141]
moiré potentials in heterobilayers significantly alters the inter-
layer exciton transport and dynamics.
The influence of topology on the physical properties is also 5. Summary and Outlook
common in all types of TMD moiré superlattices. Jin et  al.[121]
fabricated h-BN-encapsulated WSe2/WS2 HSs, with a measured Remarkable advancements have been achieved in the investi-
twist angle of 0.5  ± 0.3° (Figure 16a). In the strong-coupling gation and exploration of stacking-engineered heterobilayers,
regime of the large moiré potential of the superlattice, three such as the controllable preparation of vertical heterobilayers,
moiré exciton minibands appeared, as shown in Figure  16b. their outstanding photoelectric properties, and effective modu-
Thus, the moiré potential is stronger than the exciton kinetic lation of interlayer coupling through stacking configurations.
energy under this regime. Ultimately, the flat low-energy Although TMD-based HSs have shown extraordinary potential
exciton band exhibits a highly localized exciton density, forming and considerable progress has been achieved to date, many out-
a topological exciton band.[121] Similar topological effects have standing questions remain regarding controllable preparation
also been reported by Tong et al.[139] The authors demonstrated and the dependence of spin-valley dynamics on the stacking
topological mosaic and topologically protected helical modes at order and crystallographic alignments. As such, the following
the phase boundaries of topological or normal insulators in 1D is a summary of the main future research directions for TMD
moiré superlattices formed by the assembly of a strained TMD stacking.

Adv. Mater. 2021, 33, 2005735 2005735  (17 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

1) Precise control of the stacking configurations of samples Keywords


remains a major obstacle for future advancements; high-
2D materials, heterostructures, stacking, transition metal
precision control can be achieved, but not complete control.
dichalcogenides, twist angle
On the one hand, deterministic assembly methods require
optimization. Annealing is an essential step for artificial Received: August 24, 2020
assembly, which causes immense damage to the interface, Revised: October 30, 2020
therefore, innovative types of transfer carriers should be Published online: March 14, 2021
developed. On the other hand, scalable CVD growth for the
preparation of nanosheets should be enhanced to improve
the controllability of the stacking orientations, interfacial [1] K. S.  Novoselov, A. K.  Geim, S. V.  Morozov, D.  Jiang, Y.  Zhang,
contacts, and sample size. To advance the fabrication steps S. V. Dubonos, I. V. Grigorieva, A. A. Firsov, Science 2004, 306, 666.
further, it is essential to integrate stacking-engineered layers [2] K. Watanabe, T. Taniguchi, H. Kanda, Nat. Mater. 2004, 3, 404.
[3] V. Tran, R. Soklaski, Y. Liang, L. Yang, Phys. Rev. B 2014, 89, 235319.
into functional nanodevices, however, most studies are cur-
[4] D.  Xiao, G.-B.  Liu, W.  Feng, X.  Xu, W.  Yao, Phys. Rev. Lett. 2012,
rently in the experimental and theoretical stages.
108, 196802.
2) Although understanding of the valley exciton dynamics is [5] W.  Zhao, Z.  Ghorannevis, L.  Chu, M.  Toh, C.  Kloc, P.-H.  Tan,
developing, a quantitative description of the intrinsic mecha- G. Eda, ACS Nano 2013, 7, 791.
nism for charge transfer in heterobilayers remains elusive, [6] S.  Haldar, H.  Vovusha, M. K.  Yadav, O.  Eriksson, B.  Sanyal, Phys.
and the current theories cannot fully account for the Cou- Rev. B 2015, 92, 235408.
lombic interactions between electron–hole pairs. Numerous [7] Y.  Cao, V.  Fatemi, S.  Fang, K.  Watanabe, T.  Taniguchi, E.  Kaxiras,
groups have explored the dependence of the dynamics on the P. Jarillo-Herrero, Nature 2018, 556, 43.
constituent TMD layers, their stacking configurations, and [8] Y.  Cao, V.  Fatemi, A.  Demir, S.  Fang, S. L.  Tomarken, J. Y.  Luo,
stacking order in bilayers.[28,134,137,142] However, whether theo- J. D.  Sanchez-Yamagishi, K.  Watanabe, T.  Taniguchi, E.  Kaxiras,
R. C. Ashoori, P. Jarillo-Herrero, Nature 2018, 556, 80.
ry or experiment, a sufficient unified conclusion has not been
[9] Y.  Cao, J. Y.  Luo, V.  Fatemi, S.  Fang, J. D.  Sanchez-Yamagishi,
reached. Although one can argue that the excitonic states in
K. Watanabe, T. Taniguchi, E. Kaxiras, P. Jarillo-Herrero, Phys. Rev.
stacking bilayers are superpositions of the quasiparticle band Lett. 2016, 117, 116804.
states, a quantitative description of various excitonic correla- [10] A. Uri, S. Grover, Y. Cao, J. A. Crosse, K. Bagani, D. Rodan-Legrain,
tions remains a theoretical challenge. Y.  Myasoedov, K.  Watanabe, T.  Taniguchi, P.  Moon, M.  Koshino,
3) Another challenge for stacking bilayers is that the heavy com- P. Jarillo-Herrero, E. Zeldov, Nature 2020, 581, 47.
puting power required for incommensurate supercells or ir- [11] W.  Liao, S.  Zhao, F.  Li, C.  Wang, Y.  Ge, H.  Wang, S.  Wang,
regular bodies exceeds the capacity of traditional theoretical H. Zhang, Nanoscale Horiz. 2020, 5, 787.
calculations. Moiré superlattices with small twist angles are [12] C.  Jin, E. Y.  Ma, O.  Karni, E. C.  Regan, F.  Wang, T. F.  Heinz, Nat.
constructed using thousands of atoms, vacancies, impurities, Nanotechnol. 2018, 13, 994.
[13] M.  Chhowalla, H. S.  Shin, G.  Eda, L.-J.  Li, K. P.  Loh, H.  Zhang,
or combined systems. Thus, a description of the interactions
Nat. Chem. 2013, 5, 263.
between electrons and holes in stacking bilayers has not been
[14] Q. H.  Wang, K.  Kalantar-Zadeh, A.  Kis, J. N.  Coleman,
clearly or accurately presented to date. Therefore, new meth- M. S. Strano, Nat. Nanotechnol. 2012, 7, 699.
ods need to be optimized to cope with these issues. One exam- [15] H.  Chen, J.  Yin, J.  Yang, X.  Zhang, M.  Liu, Z.  Jiang, J.  Wang,
ple of such a case is the unfolding band calculation method, Z. Sun, T. Guo, W. Liu, P. Yan, Opt. Lett. 2017, 42, 4279.
in which the dense energy bands of the supercell are unfolded [16] J.  Zhou, J.  Qin, L.  Guo, N.  Zhao, C.  Shi, E.-z.  Liu, F.  He, L.  Ma,
onto the larger Brillouin zone of primitive cells with proper J. Li, C. He, J. Mater. Chem. A 2016, 4, 17370.
spectral weighting using explicit Kohn–Sham orbits.[143,144] [17] E. Yang, H. Ji, Y. Jung, J. Phys. Chem. C 2015, 119, 26374.
[18] K.  Zhang, T.  Zhang, G.  Cheng, T.  Li, S.  Wang, W.  Wei, X.  Zhou,
Overcoming these challenges would lead to greater under- W.  Yu, Y.  Sun, P.  Wang, D.  Zhang, C.  Zeng, X.  Wang, W.  Hu,
H. J. Fan, G. Shen, X. Chen, X. Duan, K. Chang, N. Dai, ACS Nano
standing and significant progress in the field of stacking 2D
2016, 10, 3852.
materials. At a fundamental level, stacking-engineered materials
[19] H.  Tan, W.  Xu, Y.  Sheng, C. S.  Lau, Y.  Fan, Q.  Chen, M.  Tweedie,
have exhibited novel electronic, excitonic, magnetic, and topo- X. Wang, Y. Zhou, J. H. Warner, Adv. Mater. 2017, 29, 1702917.
logical properties. The field of stacking design is young, and a [20] T.  Akama, W.  Okita, R.  Nagai, C.  Li, T.  Kaneko, T.  Kato, Sci. Rep.
plethora of unique physical phenomena is expected in the future. 2017, 7, 11967.
[21] A. Thilagam, J. Math. Chem. 2017, 55, 50.
[22] X. Chia, M. Pumera, Nat. Catal. 2018, 1, 909.
Acknowledgements [23] J. Shi, R. Tong, X. Zhou, Y. Gong, Z. Zhang, Q. Ji, Y. Zhang, Q. Fang,
This work was financially supported by the National Natural Science L. Gu, X. Wang, Z. Liu, Y. Zhang, Adv. Mater. 2016, 28, 10664.
Foundation of China (Grant Nos. 11704185 and 61935017), Natural [24] Y.  Ren, Q.  Xu, C.  Wang, X.  Zheng, Y.  Jia, Y.  Qi, Y.  Zhou, X.  Yang,
Science Foundation of Jiangsu Province, China (Grant No. BK20171021), Z. Zhang, ChemNanoMat 2017, 3, 632.
“High-level talents in six industries” in the Jiangsu Province (Grant No. [25] T. Heine, Acc. Chem. Res. 2015, 48, 65.
XCL-020), and the Projects of International Cooperation and Exchanges [26] J. He, K. Hummer, C. Franchini, Phys. Rev. B 2014, 89, 075409.
NSFC (51811530018). [27] J. Kang, S. Tongay, J. Zhou, J. Li, J. Wu, Appl. Phys. Lett. 2013, 102,
012111.
[28] H.  Heo, J. H.  Sung, S.  Cha, B.-G.  Jang, J.-Y.  Kim, G.  Jin, D.  Lee,
Conflict of Interest J.-H. Ahn, M.-J. Lee, J. H. Shim, H. Choi, M.-H. Jo, Nat. Commun. 2015,
The authors declare no conflict of interest. 6, 7372.

Adv. Mater. 2021, 33, 2005735 2005735  (18 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

[29] P.  Rivera, H.  Yu, K. L.  Seyler, N. P.  Wilson, W.  Yao, X.  Xu, Nat. [57] R. D.  Nikam, P. A.  Sonawane, R.  Sankar, Y.-T.  Chen, Nano Energy
Nanotechnol. 2018, 13, 1004. 2017, 32, 454.
[30] P. K. Gogoi, Y.-C. Lin, R. Senga, H.-P. Komsa, S. L. Wong, D. Chi, [58] K. Chen, X. Wan, J. Xu, Adv. Funct. Mater. 2017, 27, 1603884.
A. V.  Krasheninnikov, L.-J.  Li, M. B. H.  Breese, S. J.  Pennycook, [59] J. A. Robinson, ACS Nano 2016, 10, 42.
A. T. S. Wee, K. Suenaga, ACS Nano 2019, 13, 9541. [60] X. Zhang, F. Meng, J. R. Christianson, C. Arroyo-Torres, M. A. Lukowski,
[31] B.  Miller, A.  Steinhoff, B.  Pano, J.  Klein, F.  Jahnke, A.  Holleitner, D. Liang, J. R. Schmidt, S. Jin, Nano Lett. 2014, 14, 3047.
U. Wurstbauer, Nano Lett. 2017, 17, 5229. [61] L.  Tang, T.  Li, Y.  Luo, S.  Feng, Z.  Cai, H.  Zhang, B.  Liu,
[32] X. Luo, X. Lu, C. Cong, T. Yu, Q. Xiong, S. Y. Quek, Sci. Rep. 2015, H.-M. Cheng, ACS Nano 2020, 14, 4646.
5, 14565. [62] Q.  Zhang, X.  Xiao, R.  Zhao, D.  Lv, G.  Xu, Z.  Lu, L.  Sun, S.  Lin,
[33] S. Yi, J.-H. Choi, K. Lee, S. W. Kim, C. H. Park, J.-H. Cho, Phys. Rev. X.  Gao, J.  Zhou, C.  Jin, F.  Ding, L.  Jiao, Angew. Chem., Int. Ed.
B 2016, 94, 235428. 2015, 54, 8957.
[34] Y.  Zhou, W.  Xu, Y.  Sheng, H.  Huang, Q.  Zhang, L.  Hou, [63] T. Zhang, B. Jiang, Z. Xu, R. G. Mendes, Y. Xiao, L. Chen, L. Fang,
V.  Shautsova, J. H.  Warner, ACS Appl. Mater. Interfaces 2019, 11, T. Gemming, S. Chen, M. H. Rümmeli, L. Fu, Nat. Commun. 2016,
2234. 7, 13911.
[35] Z.  Ji, H.  Hong, J.  Zhang, Q.  Zhang, W.  Huang, T.  Cao, R.  Qiao, [64] Q. Fu, X. Wang, J. Zhou, J. Xia, Q. Zeng, D. Lv, C. Zhu, X. Wang,
C. Liu, J. Liang, C. Jin, L. Jiao, K. Shi, S. Meng, K. Liu, ACS Nano Y. Shen, X. Li, Y. Hua, F. Liu, Z. Shen, C. Jin, Z. Liu, Chem. Mater.
2017, 11, 12020. 2018, 30, 4001.
[36] A.  Castellanos-Gomez, M.  Buscema, R.  Molenaar, V.  Singh, [65] X. Wu, X. Wang, H. Li, Z. Zeng, B. Zheng, D. Zhang, F. Li, X. Zhu,
L.  Janssen, H. S. J.  van der  Zant, G. A.  Steele, 2D Mater. 2014, 1, Y. Jiang, A. Pan, Nano Res. 2019, 12, 3123.
011002. [66] F. Li, Y. Feng, Z. Li, C. Ma, J. Qu, X. Wu, D. Li, X. Zhang, T. Yang,
[37] C. R.  Dean, A. F.  Young, I.  Meric, C.  Lee, L.  Wang, S.  Sorgenfrei, Y.  He, H.  Li, X.  Hu, P.  Fan, Y.  Chen, B.  Zheng, X.  Zhu, X.  Wang,
K.  Watanabe, T.  Taniguchi, P.  Kim, K. L.  Shepard, J.  Hone, Nat. X. Duan, A. Pan, Adv. Mater. 2019, 31, 1901351.
Nanotechnol. 2010, 5, 722. [67] W.  Sun, X.  Wang, J.  Feng, T.  Li, Y.  Huan, J.  Qiao, L.  He, D.  Ma,
[38] Z. Cai, B. Liu, X. Zou, H.-M. Cheng, Chem. Rev. 2018, 118, 6091. Nanotechnology 2019, 30, 385601.
[39] Y.  Deng, Y.  Yu, Y.  Song, J.  Zhang, N. Z.  Wang, Z.  Sun, Y.  Yi, [68] Y. Gong, S. Lei, G. Ye, B. Li, Y. He, K. Keyshar, X. Zhang, Q. Wang,
Y. Z.  Wu, S.  Wu, J.  Zhu, J.  Wang, X. H.  Chen, Y.  Zhang, Nature J. Lou, Z. Liu, R. Vajtai, W. Zhou, P. M. Ajayan, Nano Lett. 2015, 15,
2018, 563, 94. 6135.
[40] M.  Samadi, N.  Sarikhani, M.  Zirak, H.  Zhang, H.-L.  Zhang, [69] R. M. Yunus, H. Endo, M. Tsuji, H. Ago, Phys. Chem. Chem. Phys.
A. Z. Moshfegh, Nanoscale Horiz. 2018, 3, 90. 2015, 17, 25210.
[41] H. Li, Y. Li, A. Aljarb, Y. Shi, L.-J. Li, Chem. Rev. 2018, 118, 6134. [70] M.-Y.  Li, Y.  Shi, C.-C.  Cheng, L.-S.  Lu, Y.-C.  Lin, H.-L.  Tang,
[42] P. K. Sahoo, S. Memaran, Y. Xin, L. Balicas, H. R. Gutiérrez, Nature M.-L.  Tsai, C.-W.  Chu, K.-H.  Wei, J.-H.  He, W.-H.  Chang,
2018, 553, 63. K. Suenaga, L.-J. Li, Science 2015, 349, 524.
[43] K.  Kang, K.-H.  Lee, Y.  Han, H.  Gao, S.  Xie, D. A.  Muller, J.  Park, [71] J.  Zhang, J.  Wang, P.  Chen, Y.  Sun, S.  Wu, Z.  Jia, X.  Lu, H.  Yu,
Nature 2017, 550, 229. W.  Chen, J.  Zhu, G.  Xie, R.  Yang, D.  Shi, X.  Xu, J.  Xiang, K.  Liu,
[44] K.  Wang, J.  Wang, J.  Fan, M.  Lotya, A.  O’Neill, D.  Fox, Y.  Feng, G. Zhang, Adv. Mater. 2016, 28, 1950.
X.  Zhang, B.  Jiang, Q.  Zhao, H.  Zhang, J. N.  Coleman, L.  Zhang, [72] Y. Yoo, Z. P. Degregorio, J. E. Johns, J. Am. Chem. Soc. 2015, 137,
W. J. Blau, ACS Nano 2013, 7, 9260. 14281.
[45] A. Ciesielski, P. Samorì, Chem. Soc. Rev. 2014, 43, 381. [73] J.  Shi, D.  Ma, G.-F.  Han, Y.  Zhang, Q.  Ji, T.  Gao, J.  Sun, X.  Song,
[46] G. S. Bang, K. W. Nam, J. Y. Kim, J. Shin, J. W. Choi, S.-Y. Choi, ACS C.  Li, Y.  Zhang, X.-Y.  Lang, Y.  Zhang, Z.  Liu, ACS Nano 2014, 8,
Appl. Mater. Interfaces 2014, 6, 7084. 10196.
[47] J. C.  Shaw, H.  Zhou, Y.  Chen, N. O.  Weiss, Y.  Liu, Y.  Huang, [74] Y.  Zhang, J.  Shi, G.  Han, M.  Li, Q.  Ji, D.  Ma, Y.  Zhang, C.  Li,
X. Duan, Nano Res. 2014, 7, 511. X. Lang, Y. Zhang, Z. Liu, Nano Res. 2015, 8, 2881.
[48] S. M.  Eichfeld, L.  Hossain, Y.-C.  Lin, A. F.  Piasecki, B.  Kupp, [75] S. Wang, X. Wang, J. H. Warner, ACS Nano 2015, 9, 5246.
A. G.  Birdwell, R. A.  Burke, N.  Lu, X.  Peng, J.  Li, A.  Azcatl, [76] H. Wang, F. Liu, W. Fu, Z. Fang, W. Zhou, Z. Liu, Nanoscale 2014,
S. McDonnell, R. M. Wallace, M. J. Kim, T. S. Mayer, J. M. Redwing, 6, 12250.
J. A. Robinson, ACS Nano 2015, 9, 2080. [77] Z. Zhang, P. Chen, X. Duan, K. Zang, J. Luo, X. Duan, Science 2017,
[49] A. Govind Rajan, J. H. Warner, D. Blankschtein, M. S. Strano, ACS 357, 788.
Nano 2016, 10, 4330. [78] Y.  Yu, S.  Hu, L.  Su, L.  Huang, Y.  Liu, Z.  Jin, A. A.  Purezky,
[50] V. Senthilkumar, L. C. Tam, Y. S. Kim, Y. Sim, M.-J. Seong, J. I. Jang, D. B. Geohegan, K. W. Kim, Y. Zhang, L. Cao, Nano Lett. 2015, 15,
Nano Res. 2014, 7, 1759. 486.
[51] Y.  Zhang, Y.  Yao, M. G.  Sendeku, L.  Yin, X.  Zhan, F.  Wang, [79] Y. Jung, J. Shen, Y. Sun, J. J. Cha, ACS Nano 2014, 8, 9550.
Z. Wang, J. He, Adv. Mater. 2019, 31, 1901694. [80] N.  Choudhary, J.  Park, J. Y.  Hwang, H.-S.  Chung, K. H.  Dumas,
[52] X.-Q. Zhang, C.-H. Lin, Y.-W. Tseng, K.-H. Huang, Y.-H. Lee, Nano S. I. Khondaker, W. Choi, Y. Jung, Sci. Rep. 2016, 6, 25456.
Lett. 2015, 15, 410. [81] J. M. Woods, Y. Jung, Y. Xie, W. Liu, Y. Liu, H. Wang, J. J. Cha, ACS
[53] Y.  Xue, Y.  Zhang, Y.  Liu, H.  Liu, J.  Song, J.  Sophia, J.  Liu, Z.  Xu, Nano 2016, 10, 2004.
Q. Xu, Z. Wang, J. Zheng, Y. Liu, S. Li, Q. Bao, ACS Nano 2016, 10, [82] F. Chen, L. Wang, X. Ji, J. Phys. Chem. C 2018, 122, 28337.
573. [83] H. R. Gutiérrez, N. Perea-López, A. L. Elías, A. Berkdemir, B. Wang,
[54] P.  Solís-Fernández, M.  Bissett, H.  Ago, Chem. Soc. Rev. 2017, 46, R. Lv, F. López-Urías, V. H. Crespi, H. Terrones, M. Terrones, Nano
4572. Lett. 2013, 13, 3447.
[55] A.  Behranginia, P.  Yasaei, A. K.  Majee, V. K.  Sangwan, F.  Long, [84] S. Najmaei, Z. Liu, W. Zhou, X. Zou, G. Shi, S. Lei, B. I. Yakobson,
C. J.  Foss, T.  Foroozan, S.  Fuladi, M. R.  Hantehzadeh, J.-C. Idrobo, P. M. Ajayan, J. Lou, Nat. Mater. 2013, 12, 754.
R.  Shahbazian-Yassar, M. C.  Hersam, Z.  Aksamija, [85] J. Wintterlin, M. L. Bocquet, Surf. Sci. 2009, 603, 1841.
A. Salehi-Khojin, Small 2017, 13, 1604301. [86] C. Oshima, A. Nagashima, J. Phys.: Condens. Matter 1997, 9, 1.
[56] R. Ai, X. Guan, J. Li, K. Yao, P. Chen, Z. Zhang, X. Duan, X. Duan, [87] C.  Zhang, S.  Zhao, C.  Jin, A. L.  Koh, Y.  Zhou, W.  Xu, Q.  Li,
ACS Nano 2017, 11, 3413. Q. Xiong, H. Peng, Z. Liu, Nat. Commun. 2015, 6, 6519.

Adv. Mater. 2021, 33, 2005735 2005735  (19 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

[88] J. Liu, L. Fu, Adv. Mater. 2019, 31, 1800690. [115] Y. R. Shen, Nature 1989, 337, 519.
[89] M. Zeng, L. Fu, Acc. Chem. Res. 2018, 51, 2839. [116] S. A.  Denev, T. T. A.  Lummen, E.  Barnes, A.  Kumar, V.  Gopalan,
[90] H. Liu, G. Qi, C. Tang, M. Chen, Y. Chen, Z. Shu, H. Xiang, Y. Jin, J. Am. Ceram. Soc. 2011, 94, 2699.
S.  Wang, H.  Li, M.  Ouzounian, T. S.  Hu, H.  Duan, S.  Li, Z.  Han, [117] H. Wang, X. Qian, Nano Lett. 2017, 17, 5027.
S. Liu, ACS Appl. Mater. Interfaces 2020, 12, 13174. [118] W.-T. Hsu, Z.-A. Zhao, L.-J. Li, C.-H. Chen, M.-H. Chiu, P.-S. Chang,
[91] K.  Wang, B.  Huang, M.  Tian, F.  Ceballos, M.-W.  Lin, Y.-C. Chou, W.-H. Chang, ACS Nano 2014, 8, 2951.
M. Mahjouri-Samani, A. Boulesbaa, A. A. Puretzky, C. M. Rouleau, [119] W.-T.  Hsu, L.-S.  Lu, P.-H.  Wu, M.-H.  Lee, P.-J.  Chen, P.-Y.  Wu,
M.  Yoon, H.  Zhao, K.  Xiao, G.  Duscher, D. B.  Geohegan, ACS Y.-C.  Chou, H.-T.  Jeng, L.-J.  Li, M.-W.  Chu, W.-H.  Chang, Nat.
Nano 2016, 10, 6612. Commun. 2018, 9, 1356.
[92] H.  Li, G.  Lu, Y.  Wang, Z.  Yin, C.  Cong, Q.  He, L.  Wang, F.  Ding, [120] H. Zhu, J. Wang, Z. Gong, Y. D. Kim, J. Hone, X. Y. Zhu, Nano Lett.
T. Yu, H. Zhang, Small 2013, 9, 1974. 2017, 17, 3591.
[93] W.  Zhao, Z.  Ghorannevis, K. K.  Amara, J. R.  Pang, M.  Toh, [121] C.  Jin, E. C.  Regan, A.  Yan, M.  Iqbal Bakti Utama, D.  Wang,
X. Zhang, C. Kloc, P. H. Tan, G. Eda, Nanoscale 2013, 5, 9677. S.  Zhao, Y.  Qin, S.  Yang, Z.  Zheng, S.  Shi, K.  Watanabe,
[94] Y. Shi, J.-K. Huang, L. Jin, Y.-T. Hsu, S. F. Yu, L.-J. Li, H. Y. Yang, Sci. T. Taniguchi, S. Tongay, A. Zettl, F. Wang, Nature 2019, 567, 76.
Rep. 2013, 3, 1839. [122] K.  Tran, G.  Moody, F.  Wu, X.  Lu, J.  Choi, K.  Kim, A.  Rai,
[95] J.  Shi, M.  Liu, J.  Wen, X.  Ren, X.  Zhou, Q.  Ji, D.  Ma, Y.  Zhang, D. A.  Sanchez, J.  Quan, A.  Singh, J.  Embley, A.  Zepeda,
C. Jin, H. Chen, S. Deng, N. Xu, Z. Liu, Y. Zhang, Adv. Mater. 2015, M.  Campbell, T.  Autry, T.  Taniguchi, K.  Watanabe, N.  Lu,
27, 7086. S. K.  Banerjee, K. L.  Silverman, S.  Kim, E.  Tutuc, L.  Yang,
[96] P. K.  Nayak, Y.  Horbatenko, S.  Ahn, G.  Kim, J.-U.  Lee, K. Y.  Ma, A. H. MacDonald, X. Li, Nature 2019, 567, 71.
A. R. Jang, H. Lim, D. Kim, S. Ryu, H. Cheong, N. Park, H. S. Shin, [123] M. R.  Rosenberger, H.-J.  Chuang, M.  Phillips, V. P.  Oleshko,
ACS Nano 2017, 11, 4041. K. M.  McCreary, S. V.  Sivaram, C. S.  Hellberg, B. T.  Jonker, ACS
[97] S.  Huang, L.  Liang, X.  Ling, A. A.  Puretzky, D. B.  Geohegan, Nano 2020, 14, 4550.
B. G. Sumpter, J. Kong, V. Meunier, M. S. Dresselhaus, Nano Lett. [124] S. Carr, D. Massatt, S. B. Torrisi, P. Cazeaux, M. Luskin, E. Kaxiras,
2016, 16, 1435. Phys. Rev. B 2018, 98, 224102.
[98] J. Ji, S. Dong, A. Zhang, Q. Zhang, Phys. E 2016, 80, 130. [125] C. Zhang, C.-P. Chuu, X. Ren, M.-Y. Li, L.-J. Li, C. Jin, M.-Y. Chou,
[99] A. A.  Puretzky, L.  Liang, X.  Li, K.  Xiao, K.  Wang, C.-K. Shih, Sci. Adv. 2017, 3, e1601459.
M.  Mahjouri-Samani, L.  Basile, J. C.  Idrobo, B. G.  Sumpter, [126] X. Hu, L. Kou, L. Sun, Sci. Rep. 2016, 6, 31122.
V. Meunier, D. B. Geohegan, ACS Nano 2015, 9, 6333. [127] N.  Lu, H.  Guo, Z.  Zhuo, L.  Wang, X.  Wu, X. C.  Zeng, Nanoscale
[100] X.  Ling, L.  Liang, S.  Huang, A. A.  Puretzky, D. B.  Geohegan, 2017, 9, 19131.
B. G. Sumpter, J. Kong, V. Meunier, M. S. Dresselhaus, Nano Lett. [128] J. R. Schaibley, P. Rivera, H. Yu, K. L. Seyler, J. Yan, D. G. Mandrus,
2015, 15, 4080. T.  Taniguchi, K.  Watanabe, W.  Yao, X.  Xu, Nat. Commun. 2016, 7,
[101] A. A. Puretzky, L. Liang, X. Li, K. Xiao, B. G. Sumpter, V. Meunier, 13747.
D. B. Geohegan, ACS Nano 2016, 10, 2736. [129] S. A.  Wolf, D. D.  Awschalom, R. A.  Buhrman, J. M.  Daughton,
[102] W.  Choi, I.  Akhtar, M. A.  Rehman, M.  Kim, D.  Kang, J.  Jung, S.  von  Molnár, M. L.  Roukes, A. Y.  Chtchelkanova, D. M.  Treger,
Y.  Myung, J.  Kim, H.  Cheong, Y.  Seo, ACS Appl. Mater. Interfaces Science 2001, 294, 1488.
2019, 11, 2470. [130] H. Wang, C. Zhang, F. Rana, Nano Lett. 2015, 15, 339.
[103] C. H.  Lui, Z.  Ye, C.  Ji, K.-C.  Chiu, C.-T.  Chou, T. I.  Andersen, [131] J.  Kim, C.  Jin, B.  Chen, H.  Cai, T.  Zhao, P.  Lee, S.  Kahn,
C. Means-Shively, H. Anderson, J.-M. Wu, T. Kidd, Y.-H. Lee, R. He, K.  Watanabe, T.  Taniguchi, S.  Tongay, M. F.  Crommie, F.  Wang,
Phys. Rev. B 2015, 91, 165403. Sci. Adv. 2017, 3, e1700518.
[104] J.  Kang, J.  Li, S.-S.  Li, J.-B.  Xia, L.-W.  Wang, Nano Lett. 2013, 13, [132] Y. Wang, Z. Wang, W. Yao, G.-B. Liu, H. Yu, Phys. Rev. B 2017, 95,
5485. 115429.
[105] X.  Zhu, N. R.  Monahan, Z.  Gong, H.  Zhu, K. W.  Williams, [133] K.  Kim, A.  DaSilva, S.  Huang, B.  Fallahazad, S.  Larentis,
C. A. Nelson, J. Am. Chem. Soc. 2015, 137, 8313. T.  Taniguchi, K.  Watanabe, B. J.  LeRoy, A. H.  MacDonald,
[106] L. Li, R. Long, O. V. Prezhdo, Chem. Mater. 2017, 29, 2466. Proc. Natl. Acad. Sci. USA 2017, 114, 3364.
[107] F. Chen, Y. Wang, W. Su, S. Ding, L. Fu, J. Phys. Chem. C 2019, 123, [134] Y. K.  Ryu, R.  Frisenda, A.  Castellanos-Gomez, Chem. Commun.
30519. 2019, 55, 11498.
[108] H.  Chen, X.  Wen, J.  Zhang, T.  Wu, Y.  Gong, X.  Zhang, J.  Yuan, [135] H. Yu, G.-B. Liu, J. Tang, X. Xu, W. Yao, Sci. Adv. 2017, 3, e1701696.
C.  Yi, J.  Lou, P. M.  Ajayan, W.  Zhuang, G.  Zhang, J.  Zheng, Nat. [136] S. Zhu, H. T. Johnson, Nanoscale 2018, 10, 20689.
Commun. 2016, 7, 12512. [137] L.  Yuan, B.  Zheng, J.  Kunstmann, T.  Brumme, A. B.  Kuc,
[109] P.  Nagler, G.  Plechinger, M. V.  Ballottin, A.  Mitioglu, S.  Meier, C. Ma, S. Deng, D. Blach, A. Pan, L. Huang, Nat. Mater. 2020, 19,
N.  Paradiso, C.  Strunk, A.  Chernikov, P. C. M.  Christianen, 617.
C. Schüller, T. Korn, 2D Mater. 2017, 4, 025112. [138] K. L. Seyler, P. Rivera, H. Yu, N. P. Wilson, E. L. Ray, D. G. Mandrus,
[110] E. M.  Alexeev, D. A.  Ruiz-Tijerina, M.  Danovich, M. J.  Hamer, J. Yan, W. Yao, X. Xu, Nature 2019, 567, 66.
D. J. Terry, P. K. Nayak, S. Ahn, S. Pak, J. Lee, J. I. Sohn, M. R. Molas, [139] Q.  Tong, H.  Yu, Q.  Zhu, Y.  Wang, X.  Xu, W.  Yao, Nat. Phys. 2017,
M.  Koperski, K.  Watanabe, T.  Taniguchi, K. S.  Novoselov, 13, 356.
R. V.  Gorbachev, H. S.  Shin, V. I.  Fal’ko, A. I.  Tartakovskii, Nature [140] S. Zhu, P. Pochet, H. T. Johnson, ACS Nano 2019, 13, 6925.
2019, 567, 81. [141] F. Wu, T. Lovorn, E. Tutuc, I. Martin, A. H. MacDonald, Phys. Rev.
[111] D. A. Ruiz-Tijerina, V. I. Fal’ko, Phys. Rev. B 2019, 99, 125424. Lett. 2019, 122, 086402.
[112] T. Deilmann, K. S. Thygesen, Nano Lett. 2018, 18, 2984. [142] Y.  Liu, N. O.  Weiss, X.  Duan, H.-C.  Cheng, Y.  Huang, X.  Duan,
[113] A. O. Slobodeniuk, Ł. Bala, M. Koperski, M. R. Molas, P. Kossacki, Nat. Rev. Mater. 2016, 1, 16042.
K. Nogajewski, M. Bartos, K. Watanabe, T. Taniguchi, C. Faugeras, [143] G. C.  Constantinescu, N. D. M.  Hine, Phys. Rev. B 2015, 91,
M. Potemski, 2D Mater. 2019, 6, 025026. 195416.
[114] P. B. Petersen, R. J. Saykally, J. Phys. Chem. B 2006, 110, 14060. [144] W. Ku, T. Berlijn, C.-C. Lee, Phys. Rev. Lett. 2010, 104, 216401.

Adv. Mater. 2021, 33, 2005735 2005735  (20 of 21) © 2021 Wiley-VCH GmbH
15214095, 2021, 16, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.202005735 by National Taiwan University, Wiley Online Library on [24/04/2023]. 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
www.advancedsciencenews.com www.advmat.de

Shixuan Wang received his B.E. degree from Jiangsu Normal University in 2018. He is currently
pursuing a M.E. degree under the supervision of Prof. Jiaxu Yan at the Institute of Advanced
Materials (IAM), Nanjing Tech University. His research interests include first principle calculation,
optoelectronic properties of layered 2D van der Waals nanomaterials and their potential device
applications.

Jiaxu Yan received his B.E. degree in Materials Physics from University of Science & Technology,
Beijing (2007) and received his Ph.D. degree in Physics from Nanyang Technological University
(NTU), Singapore (2015). After two years of postdoctoral research at NTU, he is currently a full
professor in Nanjing Tech University. His research interests are the first-principles study and
design of functional materials and optoelectronic properties of 2D materials/organic–inorganic
halide perovskites.

Wei Huang received his Ph.D. degree from Peking University (1992). In 2001, he became a
chair professor at Fudan University, where he founded the Institute of Advanced Materials
(IAM). In 2006, he was appointed as the Deputy President of Nanjing University of Posts
and Telecommunications (NUPT, 2006). He assumed his duty as the President of Nanjing
Tech University (Nanjing Tech, 2012) and was appointed as Deputy President and Provost
of Northwestern Polytechnical University (NPU, 2017). His research interests include
organic/plastic materials and devices, nanomaterials, nanotechnology, among others.

Adv. Mater. 2021, 33, 2005735 2005735  (21 of 21) © 2021 Wiley-VCH GmbH

You might also like