Home Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications
Article Open Access

Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications

  • Yumei Jing , Baoze Liu , Xukun Zhu , Fangping Ouyang , Jian Sun EMAIL logo and Yu Zhou ORCID logo EMAIL logo
Published/Copyright: April 2, 2020
Become an author with De Gruyter Brill

Abstract

Differing from its bulk counterparts, atomically thin two-dimensional transition metal dichalcogenides that show strong interaction with light are considered as new candidates for optoelectronic devices. Either physical or chemical strategies can be utilized to effectively tune the intrinsic electronic structures for adopting optoelectronic applications. This review will focus on the different tuning strategies that include its physics principles, in situ experimental techniques, and its application of various optoelectronic devices.

1 Introduction

Atomically thin two-dimensional (2D) transition metal dichalcogenides (TMDs) have been extensively studied in electronic, optoelectronic, and electrochemical devices because of its rich physical and chemical properties since graphene was discovered from 2004 [1]. Thanks to anisotropic layered structures, the electronic structures of 2D TMDs can be tuned by the multiple factors such as thickness thinning down, foreigner species interaction, external strain, and composition engineering or assembling heterostructures, which have been widely applied to optoelectronic devices such as electronic logic devices, photovoltaic devices, photodetection devices, nonlinear optical absorption devices, light emission devices, and so on [2]. Direct-bandgap monolayer MoS2 devices with improved carrier mobility have proven the ultrasensitive responsivity of 880 AW−1 at a wavelength of 561 nm and a response spectrum in the 400- to 680-nm range with equivalent noise level as the silicon devices [3]. A bilayer MoTe2-based light-emitting diode and photodetector for silicon photonic integrated circuits were demonstrated by Bie et al. [4]. Monolayer and bilayer MoS2 field-effect transistors with ionic liquid gate tuning show both holes and electrons accumulation in the ambipolar regime, thus emitting the visible light [5]. The photodetectors and emitting devices based on the 2D TMDs have been developed with large photoresponsivity, ultrahigh gain, tunable spectrum sensitivity, bandwidth, and considerable light-to-current or current- to-light conversion efficiency. At the same time with the rapid developments of large-scale preparation technology and direct patterning of specific position and geometry for the 2D films, such as chemical vapor deposition (CVD), liquid exfoliation, printing techniques, and so on, 2D TMD materials show the great potential for the optoelectronic integrated circuits, photodetection, imaging, and biomedical recording [6], [7], [8], [9], [10], [11].

In this review, considering 2D MoS2 as a star material, we intend to summarize the structure–optoelectronic properties relationships of 2D TMDs in Section 2. Different modulation strategies based on the distinctive crystal structure will be systematically introduced in Section 3. Finally, the development history, progress, and perspective of optoelectronic device application in 2D TMDs are summarized in Section 4.

2 Basic material properties

The TMD material family is rather rich owing to the combination of various transition metal elements and chalcogens and exhibits abundant charming physical properties. The material properties are strongly dependent on their chemical composition and atomic arrangement. Before reviewing the tuning of electronic structure for optoelectronics, some basic material properties related to the optoelectronics, including the crystal structure and electronic properties, are outlined in this section.

2.1 Crystal structure

Transition metal dichalcogenides can be generalized with the formula MX2, where M denotes a transition metal from group IIIB to group VIIIB of subgroup elements in the element periodic table (Mo, Ti, Zr, Nb, Ta, W, Pt, and so on), whereas X denotes a chalcogen (S, Se, or Te). A large number of TMDs exhibit graphite-like layered structure, which can be exfoliated into 2D layers. The layered structure is composed of 2D hexagonally packed X-M-X sandwich layers – metal atoms in the middle hexagonal plane covalently bonded to chalcogen atoms located in the top and bottom hexagonal planes, stacked layer-by-layer, as shown in Figure 1A. The adjacent X-M-X sandwich layers are held together to form bulk crystal by van der Waals interactions, which is much weaker than intralayer covalent bonds. The strong intralayer and weak interlayer interactions lead to the strong anisotropy in mechanical properties, responsible for a remarkably easy mechanical exfoliation to get thin layers or monolayer (a single X-M-X sandwich layer).

Figure 1: Crystal and electronic structures.(A) Three-dimensional schematic illustration of a typical TMD (MX2) structure with 2H phase. The pink ball represents the chalcogen atom X, and the gray ball represents the metal atom M, which are also the same in (B) and (C). (B) The schematic illustration of the two basic octahedral and trigonal prismatic coordination in TMDs. (C) Schematics of the three most commonly involved structural phases: 1T (tetragonal symmetry, one layer per repeat unit, octahedral coordination), 2H (hexagonal symmetry, two layers per repeat unit, trigonal prismatic coordination), and 3R (rhombohedral symmetry, three layers per repeat unit, trigonal prismatic coordination). The dot-dashed vertical lines indicate the alignments of interlayer M and X atoms. (D) Simplified band structure of bulk MoS2, showing the lowest conduction band c1 and the highest split valence bands v1 and v2. A and B are the direct-gap transitions, and I is the indirect-gap transition. E′g${E'_{\rm{g}}}$ is the indirect gap for the bulk, and Eg is the direct gap for the monolayer. Reproduced with permission from Mak et al. [12]. Copyright 2010 the American Physical Society.
Figure 1:

Crystal and electronic structures.

(A) Three-dimensional schematic illustration of a typical TMD (MX2) structure with 2H phase. The pink ball represents the chalcogen atom X, and the gray ball represents the metal atom M, which are also the same in (B) and (C). (B) The schematic illustration of the two basic octahedral and trigonal prismatic coordination in TMDs. (C) Schematics of the three most commonly involved structural phases: 1T (tetragonal symmetry, one layer per repeat unit, octahedral coordination), 2H (hexagonal symmetry, two layers per repeat unit, trigonal prismatic coordination), and 3R (rhombohedral symmetry, three layers per repeat unit, trigonal prismatic coordination). The dot-dashed vertical lines indicate the alignments of interlayer M and X atoms. (D) Simplified band structure of bulk MoS2, showing the lowest conduction band c1 and the highest split valence bands v1 and v2. A and B are the direct-gap transitions, and I is the indirect-gap transition. Eg is the indirect gap for the bulk, and Eg is the direct gap for the monolayer. Reproduced with permission from Mak et al. [12]. Copyright 2010 the American Physical Society.

The bulk TMDs exhibit a variety of polymorphs, which determine the unique structure and electronic properties, depending on stacking orders and metal atom coordination by the chalcogens. As illustrated in Figure 1B, within each X-M-X sandwich layer, the transition metal atom is coordinated by six chalcogens in two basic geometries: octahedral coordination or trigonal prismatic coordination, leading to different symmetries. The octahedral coordination with AbC (the letters denote different atom-arrange positions, whereas the uppercase and lowercase letters denote chalcogen and transition metal atoms, respectively) stacking sequence holds inversion symmetry for monolayer, whereas the trigonal prismatic coordination with AbA stacking sequence results in broken inversion symmetry in monolayer. There are three most commonly involved structural phases [1]: 1T (one sandwich layer per unit cell with AbC stacking order, octahedral coordination), 2H (two sandwich layers per unit cell with AbABaB stacking order, trigonal prismatic coordination), and 3R (three sandwich layers per unit cell with AbACaCBcB stacking order, trigonal prismatic coordination), as illustrated in Figure 1C. As an example, the most widely studied group VIB bulk TMDs (e.g. MoS2, WSe2, MoTe2) exhibit 2H or 3R phase [2], [13], [14], [15], [16], the group VB TMDs (e.g. TaSe2, TaS2) exhibit 2H or 1T phase [17], [18], [19], [20], [21], and all the group IVB TMDs (e.g. TiS2, ZrS2) are 1T phase [22], [23], [24], [25], [26], [27]. It has been found that phase changes can be induced by intercalation [28], [29], [30], [31], [32], [33], [34]. For example, 2H-MoS2 can transform to 1T phase by lithium intercalation [2], [31], [33], [34], [35]. These structural phases featured with different symmetries and stacking orders, relating to the d-electron count, primarily play important roles in affecting the electronic properties of TMDs.

2.2 Electronic properties

Most TMDs have common features in their band structures originating from d orbital on metal atoms, as demonstrated by theoretical calculations and spectroscopic experiments [1]. Different d-electron counts of transition metal elements from group IIIB to group VIII fill the nonbonding d bands to different degrees, resulting in the wide variety of electronic properties including insulator, semiconductor, semimetal, metal, and superconductor [2]. In general, the partially filled group VIB TMDs, such as MoX2 and WX2, are semiconducting [36], [37], whereas the fully filled group VB TMDs, such as NbX2 and TaX2, are metallic [38], [39].

Semiconducting TMDs such as MoX2 and WX2 possess bandgaps in the near-infrared to the visible region, promising interesting photonics and optoelectronics applications [3], [40], [41], [42], [43], [44], [45], [46], [47], [48], [49], [50], [51], [52]. Their bulk materials are indirect-bandgap semiconductors with the bottom of conduction band and the top of valence band maximum located at different high symmetry points of the Brillouin zone, respectively, whereas the monolayers change to larger direct-bandgap semiconductors [53], [54], [55], [56], [57], [58]. Compared to bulk TMDs, the electronic properties of 2D TMDs are considerably tuned as a result of quantum confined effect in the out-of-plane direction and the resulting reduced symmetries and alteration in hybridization between pz orbitals on chalcogen atoms and d orbitals on metal atoms near Γ point when thinning down thickness [55], [57], [59]. Except the indirect-to-direct bandgap transition and bandgap size enlargement, few-layer or monolayer 2D TMDs exhibit strong excitonic effects due to the weak charge carrier screening and subsequent strong Coulomb interactions caused by reducing thickness [57], [60], [61], [62], [63], [64], [65], [66], [67].

A simplified band diagram of MoS2, as a representative semiconducting TMD, is depicted pictorially in Figure 1D, showing the lowest conduction band c1 and the highest split valence bands v1 and v2. The indirect bandgap Eg for the bulk features conduction band minimum locates between Γ and K points and valence band maximum located at the Γ point, whereas the direct bandgap Eg for the monolayer locates at K point. The valence band around K point is split as a result of spin-orbit coupling [12], [68], [69], [70]. The direct-gap transitions between the maximum of the split valence bands and the minimum of conduction band, generally called the A and B excitons, will lead to two distinct low-energy resonance peaks in optical spectrum, while the indirect-gap transition I will give rise to an even lower-energy resonance peak. The indirect-to-direct bandgap transition and strong excitonic effect in monolayer TMDs have been widely demonstrated by absorption spectra and photoluminescence [12], [59], [70], [71], [72].

3 Tuning electronic structure for optoelectronics

Transition metal dichalcogenides with tunable bandgap available are desirable for many photonic and optoelectronic applications where the direct bandgap of TMDs would allow a strong absorption and efficient production of electron–hole pairs under photoexcitation. The electronics and optoelectronics engineering of 2D TMDs have been investigated extensively by theoretical calculations and experiments. It shows that the electronic structure and associated optical properties can be well tuned by a variety of physical parameters, such as thickness [12], [40], [55], [59], [70], [71], composition [73], [74], stacked heterostructures [41], [44], [46], [50], [75], [76], [77], [78], [79], [80], [81], [82], [83], strain [84], [85], [86], [87], [88], [89], [90], [91], [92], chemical modification [93], [94], and electric field [95], [96], [97], [98], [99], [100]. The progress in electronics and optoelectronics engineering and the physical mechanisms under various tuning methods will be reviewed in this section [101], [102].

3.1 Quantum confinement by thickness thinning down

The thickness-dependent band structures of TMDs, due to the resulting quantum confinement effect of thinning thickness, greatly affect the optical properties and optoelectronic performance. Benefiting from the easily mechanical cleavage [7] and bottom-up growth by CVD [103], [104], [105], [106], [107] or metal–organic chemical vapor deposition [15], [108], [109], atomic thin layers can be achieved to provide opportunities for studying thickness-dependent physical properties in experiments. As first reported by Splendiani et al. [59] in 2010, a strong photoluminescence, which is absent in bulk, emerges in the monolayer MoS2, indicating an indirect-to-direct bandgap transition in this d-electron material. Through characterization by absorption, photoluminescence, and photoconductivity spectroscopy, Mak et al. [12] traced the thickness-dependent electronic properties and the resulting optical properties of MoS2 from one to six layers (Figure 2C). It shows that the indirect bandgap increases with decreasing thickness and changes into direct bandgap at the monolayer, whereas the direct excitonic transition energy almost keeps unchanged. In addition, the monolayer MoS2 exhibits a considerable increase in luminescence quantum efficiency compared with bulk, as a result of the strong excitonic effect due to reducing thickness. The observation of thickness-dependent band structure and resulting optical properties evolution shows possibility of tuning electronic properties for optoelectronic applications by thinning thickness [3], [40], [42], [48], [110], [111].

Figure 2: Quantum confinement with tuning thickness.(A) Band structures of bulk MoS2, its monolayer, and few layers calculated at the DFT/PBE level. The horizontal dashed lines indicate the Fermi level. The arrows indicate the fundamental bandgap (direct or indirect) for a given system. The top of valence band (blue/dark gray) and bottom of conduction band (green/light gray) are highlighted. Reproduced with permission from Kuc et al. [55]. Copyright 2011 American Physical Society. (B) ARPES spectra of monolayer, bilayer, trilayer, and 8 ML MoSe2 thin films along the G–K direction. White and green dotted lines indicate the energy positions of the apices of valence bands at the G and K points, respectively, with energy values written in the same colors. Reproduced with permission from Zhang et al. [70]. Copyright 2014 Springer Nature Limited. (C) Left: PL spectra for monolayer and bilayer MoS2 samples in the photon energy range from 1.3 to 2.2 eV. Inset of left: PL QY of thin layers for N=1–6. Middle: Normalized PL spectra by the intensity of peak A of thin layers of MoS2 for N=1–6. Feature I for N=4–6 is magnified, and the spectra are displaced for clarity. Right: Bandgap energy of thin layers of MoS2, inferred from the energy of the PL feature I for N=2–6 and from the energy of the PL peak A for N=1. The dashed line represents the (indirect) bandgap energy of bulk MoS2. Reproduced with permission from Mak et al. [12]. Copyright 2010 the American Physical Society. (D) Left: Schematic 3D view of single-layer transistor with hexagonal structure MoS2 nanosheet, 50-nm-thick Al2O3 dielectric and ITO top-gate under monochromatic light. Middle: The schematic band diagrams of ITO (gate)/Al2O3 (dielectric)/single-layer (1 L), double-layer (2 L), and triple-layer (3 L) MoS2 (n-channel) under the light (Elight=hν) illustrate the photoelectric effects for bandgap measurements. Right: Respective photocurrent dynamics of single-layer MoS2-based top-gate transistors under monochromatic red and green lights. Reproduced with permission from Lee et al. [40]. Copyright 2012 American Chemical Society.
Figure 2:

Quantum confinement with tuning thickness.

(A) Band structures of bulk MoS2, its monolayer, and few layers calculated at the DFT/PBE level. The horizontal dashed lines indicate the Fermi level. The arrows indicate the fundamental bandgap (direct or indirect) for a given system. The top of valence band (blue/dark gray) and bottom of conduction band (green/light gray) are highlighted. Reproduced with permission from Kuc et al. [55]. Copyright 2011 American Physical Society. (B) ARPES spectra of monolayer, bilayer, trilayer, and 8 ML MoSe2 thin films along the G–K direction. White and green dotted lines indicate the energy positions of the apices of valence bands at the G and K points, respectively, with energy values written in the same colors. Reproduced with permission from Zhang et al. [70]. Copyright 2014 Springer Nature Limited. (C) Left: PL spectra for monolayer and bilayer MoS2 samples in the photon energy range from 1.3 to 2.2 eV. Inset of left: PL QY of thin layers for N=1–6. Middle: Normalized PL spectra by the intensity of peak A of thin layers of MoS2 for N=1–6. Feature I for N=4–6 is magnified, and the spectra are displaced for clarity. Right: Bandgap energy of thin layers of MoS2, inferred from the energy of the PL feature I for N=2–6 and from the energy of the PL peak A for N=1. The dashed line represents the (indirect) bandgap energy of bulk MoS2. Reproduced with permission from Mak et al. [12]. Copyright 2010 the American Physical Society. (D) Left: Schematic 3D view of single-layer transistor with hexagonal structure MoS2 nanosheet, 50-nm-thick Al2O3 dielectric and ITO top-gate under monochromatic light. Middle: The schematic band diagrams of ITO (gate)/Al2O3 (dielectric)/single-layer (1 L), double-layer (2 L), and triple-layer (3 L) MoS2 (n-channel) under the light (Elight=) illustrate the photoelectric effects for bandgap measurements. Right: Respective photocurrent dynamics of single-layer MoS2-based top-gate transistors under monochromatic red and green lights. Reproduced with permission from Lee et al. [40]. Copyright 2012 American Chemical Society.

Band structure calculations through the first principles show qualitative agreement with experimental researches. The thickness-dependent band structure of various TMDs has been studied by lots of theoretical works including ab initio calculations [36], [55], [58], [59], [112]. As an example shown in Figure 2A, the valence band maximum of MoS2 shifts from Γ to K point when the thickness is thinned to monolayer, resulting in indirect-to-direct bandgap transition [55]. The unusual thickness-dependent band structure of TMDs is attribute to quantum confinement effect and the characters of d-electron orbitals that comprise the conduction and valence bands [59], [70], [71]. Theoretical calculations for MoS2 [59] show that conduction band states at the K point are mainly composed of localized d orbitals on Mo atoms, located in the middle of the S-Mo-S sandwich layers and having relatively weak interlayer coupling. However, states near the Γ point originate from combination of d orbitals on Mo atoms and antibonding pz orbitals on S atoms, having strong interlayer coupling and being more sensitive to thickness. Therefore, as thickness decreases, the conduction band states near K point are almost unchanged, while the conduction bands and valence bands near the Γ point are changed considerably [113]. The direct observation of the thickness-dependent electronic band structure evolution is achieved by angle-resolved photoemission spectroscopy (ARPES) measurements on mechanically exfoliated and CVD-grown MoS2 and molecular beam epitaxy–grown MoSe2 thin films [70], [71]. The significant step-by-step evolution of the valence band (Figure 2B) provides direct evidence of the valence band maximum shifts from Γ to K point (indirect-to-direct bandgap transition), when the thickness is thinned to monolayer [70]. A lot of TMDs including MoX2 and WX2 are expected to possess a similar indirect-to-direct bandgap transition with decreasing thickness to monolayer, covering the bandgap energy range of ~1 to 2 eV [12], [40], [55], [70], [71], [114], [115], [116].

The controllability of band structure with tuning thickness paves the way for new optoelectronics. The electronic band structures of TMDs described earlier directly influence their optical properties. A variety of photodetectors based on thickness-modulated TMDs have been demonstrated. Lee et al. [40] have realized photodetection of different wavelengths using MoS2 layers with different thickness. As shown in Figure 2D, devices with single- and double-layer MoS2, exhibiting significant energy bandgaps of 1.82 and 1.65 eV, respectively, are demonstrated effective for green light detection, whereas devices made of triple-layer MoS2 with a bandgap of 1.35 eV are effective for red light detection. Yin et al. [110] have fabricated a single-layer MoS2 phototransistor showing good photoresponsibility.

3.2 Defect engineering

Previous studies have shown 2D material–fabricated electronic devices could emit photons out of the materials, detect incident photons, or control the properties of incident light [117]. Photoluminescence transition of bulk to monolayer MoS2 is directly related to their different electronic structures [43], [59]. Briefly speaking, photoluminescence consists of photoexcitation (i.e. photons that excite electrons to a higher energy level), relaxation processes, and emitting electromagnetic radiation, whose emitting intensity is proportional to the decrease in the concentration of defects. The defects of 2D materials are experimentally revealed: (1) point defects including foreigner dopants, vacancies, and ad adsorbed atoms; (2) line defects such as grain boundaries, edges, and so on. Hong et al. [118] have experimentally revealed monolayer MoS2 of physical vapor deposition often showed antisite defects with one Mo atom replacing one or two S atoms (MoS or MoS2). However, mechanical-exfoliated and chemical vapor–deposited MoS2 samples are dominantly shown with S vacancies with one (VS) or two (VS2) S atoms absent [118], [119]. The researchers exceptionally found that antisite defects (MoS or MoS2) reduced the mobility of monolayer MoS2 by three times than by the presence of vacancies (VS or VS2), whereas the phonon-limited mobility of holes carried by the intrinsic valence band is much more sensitive. The measured carrier motilities of monolayer MoS2 are mostly controlled by the primary type of defects, regardless of carrier density and contact resistance.

Influence of defects could not only change electronic structure and charge-carrier mobility of TMDs but also create midgap states in the bandgap and introduce the in-gap defect levels, which can significantly modify and improve the performance of optoelectronic properties [120]. As shown in Figure 3G, defect engineering and strong oxygen bonding on the defect sites of monolayer MoS2 could tune the photoluminescence intensity up to at least thousands of times. The physical-adsorbed O2/H2O molecules at defect sites induced heavy p-type doping and switched the trions recombination to exciton recombination process. The reduced nonradioactive recombination from excitons at defect sites could guarantee the enhancement of photoluminescence intensity under the vacuum (Figure 3A). The strong bonding energy of 2.395 eV for an oxygen molecule adsorbed on an S vacancy of MoS2 was used to controllably introduce defects and oxygen bonding in MoS2 by oxygen plasma, resulting in a controllable manipulation of photoluminescence [120].

Figure 3: Defect engineering.Photoluminescence intensity mapping images of exfoliated monolayer MoS2: (A) as-exfoliated, (B) annealed in vacuum for 1 h at 350°C, (C) pumping down to 0.1 Pa. The maps have the same color bar; photoluminescence intensity images of another monolayer MoS2: (D) as-prepared, (E) annealed in vacuum for 1 h at 500°C, (F) pumping down to 0.1 Pa. The images have the same color bar; (G, H) PL spectra taken from locations A to D in the images and change of normalized PL intensities (as compared to the original values) of locations B–D throughout the annealing and pumping process. Reproduced with permission from Nan et al. [120]. Copyright 2014 American Chemical Society.
Figure 3:

Defect engineering.

Photoluminescence intensity mapping images of exfoliated monolayer MoS2: (A) as-exfoliated, (B) annealed in vacuum for 1 h at 350°C, (C) pumping down to 0.1 Pa. The maps have the same color bar; photoluminescence intensity images of another monolayer MoS2: (D) as-prepared, (E) annealed in vacuum for 1 h at 500°C, (F) pumping down to 0.1 Pa. The images have the same color bar; (G, H) PL spectra taken from locations A to D in the images and change of normalized PL intensities (as compared to the original values) of locations B–D throughout the annealing and pumping process. Reproduced with permission from Nan et al. [120]. Copyright 2014 American Chemical Society.

3.3 Chemical composition modification

Ternary TMD alloys composed of two MX2 compounds with controllable concentrations show composition-dependent band structure, providing an effective strategy for tuning optical properties [66], [74], [121], [122], [123], [124], [125], [126], [127]. As mentioned earlier, the electronic properties of TMDs range from insulating, semiconducting, to metallic, due to the d-electron counts related to chemical compositions of TMDs. A necessary bandgap for optoelectronic applications requires at least one of the alloyed material is semiconducting. Thus, alloys are investigated either between metals and semiconductors, or semiconductors to semiconductors. For successfully alloying different materials, lattice match and bond distance are crucial [124]. Figure 4A shows how these TMDs meet the requirements by plotting theoretical bandgap differences with lattice mismatch [124]. The shaded area in the upper left corner corresponds to the largest difference in bandgaps (metal–semiconductor alloys), while the smallest lattice mismatch, potentially indicating the most suitable compound pairs for bandgap-tunable alloying. In this metal–semiconductor–type alloys, the represented metal is VX2, whereas the represented semiconductor is MoX2 or WX2. The shaded area in the lower left corner corresponds to the semiconductor–semiconductor alloys with small lattice mismatch and moderate bandgap variation with concentration. The represented semiconductor–semiconductor–type alloys are Mo–W dichalcogenides. In general, alloying TMDs within the same transition metal group or different chalcogenides are expected because they have the same lattice symmetry, as well as only small lattice mismatches. Figure 4B shows the calculated bandgaps for two representative types of TMD alloys (Mo1−xWxS2 represents alloys of the same transition metal group, and MoSe2(1−x)S2x represents alloys of different chalcogenides) at different concentrations [124]. It shows significant modulation of bandgaps by composition concentrations, expected to be observed in optical absorption or emission experiments.

Figure 4: Chemical composition tuning.(A) Lattice constant matching for alloying pairs of 2H TMDs based on LSDA lattice constants. Triangles designate oxides; squares, sulfides; circles, selenides; and diamonds, tellurides. (B) Calculated bandgaps in Mo1−xWxS2 (downward triangles, red) and MoSe2(1−x) 2x (upward triangles, blue) alloys as a function of concentration x. The bandgaps in the thermodynamic ground states are shown with larger symbols. (A) and (B) reproduced with permission from Kutana et al. [124]. Copyright 2014 The Royal Society of Chemistry. (C) PL spectrum of the complete composition MoSe2(1−x)2x nanosheets and a typical PL mapping of a single ternary nanosheet (the inset, scale bar, 7 um) excited with a 488-argon ion laser. Reproduced with permission from Li et al. [127]. Copyright 2014 American Chemical Society. (D) Left: Current–voltage characteristics at different laser powers (λ=633 nm) for five devices ranging in composition from MoS2 (Eg=1.85 eV) to MoSe2 (Eg=1.55 eV). The plots are enlarged to the same scale. Right: Photocurrent measured at a source-drain bias of 2 V as a function of laser power for single-layer MoS1.2Se0.4 (Eg=1.74 eV). The superlinear behavior is observed across compositions. Reproduced with permission from Klee et al. [128]. Copyright 2015 American Chemical Society.
Figure 4:

Chemical composition tuning.

(A) Lattice constant matching for alloying pairs of 2H TMDs based on LSDA lattice constants. Triangles designate oxides; squares, sulfides; circles, selenides; and diamonds, tellurides. (B) Calculated bandgaps in Mo1−xWxS2 (downward triangles, red) and MoSe2(1−x) 2x (upward triangles, blue) alloys as a function of concentration x. The bandgaps in the thermodynamic ground states are shown with larger symbols. (A) and (B) reproduced with permission from Kutana et al. [124]. Copyright 2014 The Royal Society of Chemistry. (C) PL spectrum of the complete composition MoSe2(1−x)2x nanosheets and a typical PL mapping of a single ternary nanosheet (the inset, scale bar, 7 um) excited with a 488-argon ion laser. Reproduced with permission from Li et al. [127]. Copyright 2014 American Chemical Society. (D) Left: Current–voltage characteristics at different laser powers (λ=633 nm) for five devices ranging in composition from MoS2 (Eg=1.85 eV) to MoSe2 (Eg=1.55 eV). The plots are enlarged to the same scale. Right: Photocurrent measured at a source-drain bias of 2 V as a function of laser power for single-layer MoS1.2Se0.4 (Eg=1.74 eV). The superlinear behavior is observed across compositions. Reproduced with permission from Klee et al. [128]. Copyright 2015 American Chemical Society.

Two-dimensional TMD alloys can be obtained by mechanical exfoliation of bulk alloys [122] and synthesis through physical vapor deposition [129], [130] or CVD [74], [131], [132], stimulating experimental exploration of bandgap engineering by chemical composition control. It is characterized by the combination of Raman and photoluminescence spectrum that 2D TMD alloys show tunable optical properties modified by the composition of alloys [66], [74], [122], [126], [127], [129], [131]. Figure 4C shows the photoluminescence spectra of the MoSe2(1−x)S2x nanosheet synthesized by CVD, with the spectral peaks continuously shifted from 668 nm (for pure MoS2) to 795 nm (for pure MoSe2) [127]. It indicates the bandgap of 2D TMD alloys can be continuously tuned through fully control of compositions. The 2D TMD alloys with continuously tunable bandgaps can have application in photodetection application, such as modulating cutoff detection wavelengths. An experimental study on photodetector based on MoSe2(1−x)S2x shows significant decrease of the photocurrent at fixed wavelength for Se-rich alloys compared to S-rich ones along with decreased diffusion length of photogenerated carriers (Figure 4D) [128]. An unusual superliner dependent of photocurrent on light intensity has also been observed. Both theoretical and experimental investigations show 2D TMD alloys possess the potential for tunable optoelectronics [133].

3.4 Foreigner species intercalation

Intercalation of metal atoms, molecules, and ions into 2D-layered TMDs has been widely studied and demonstrated to be an effective method for tuning electronic and optical properties of the host materials [21], [29], [30], [31], [32], [33], [34], [134], [135], [136]. Intercalation is a chemical process to reversibly insert various foreign species, such as zero-valent metal atoms, ions, and organic molecules, into the interlayer gaps of a host material (Figure 5A). Various intercalation methods, such as ion exchange intercalation, electrochemical intercalation, and zero-valent intercalation, have been developed for fundamental researches and technical applications [32]. The weak interlayer van der Waals interaction in TMDs leads to relatively large interlayer gap, showing that the intercalation into TMDs is easily achieved and affords a convenient way to modify and tune the electronic and optical properties for future optoelectronics applications.

Figure 5: Foreigner species intercalation.(A) Schematic for intercalation in the layered TMD materials, where guest species can be inserted into the interlayer gap by diverse methods. Reproduced with permission from Jung et al. [32]. Copyright 2016 the Partner Organisations. (B) Left: Schematic of the band structure of 2H-MoS2 at the K-point in the Brillouin zone without Li intercalation. Optical transitions due to the A and B excitons are depicted. These transitions possibly cause absorption peaks in 2H-MoS2. Right: Schematic of the band structure of 1T-Lix=1MoS2 at the K-point in the Brillouin zone after Li intercalations. 1T-Lix=1MoS2 shows a metallic behavior as the conduction band, and the valence band becomes overlapped, and the Fermi level lies in an incompletely filled band. This metallic band structure with no bandgap substantially reduces the light absorption. (C) Left: Experimental setup for electrochemical tuning of Li concentration in MoS2. Middle: Optical transmission spectra of MoS2 flakes with thicknesses of 9 nm before (red) and after (blue) Li intercalations. Right: Thickness dependence of MoS2 optical transmissions at 500 nm. The results showed enhancement in transmission after Li intercalation. (B) and (C) reproduced with permission from Xiong et al. [31]. Copyright 2015 American Chemical Society. (D) Top: Active control of PL in the 2D MoS2 nanoflakes. PL images of the nanostructured MoS2 film at the intercalating voltages of 0 V, −2 V, −4 V, −6 V, −8V, −10 V, and −12 V. Bottom: In situ PL intensity change of a selected area “1” in the MoS2 film under different intercalating/deintercalating voltages of ±2, ±4, ±6, ±8, ±10, and ±12 V in the 0.1 M LiClO4 solution (blue line) at the step duration of 60 s, whereas a selected uncoated area “2” within the substrate area is chosen as the reference (red line). Reproduced with permission from Wang et al. [134]. Copyright 2013 American Chemical Society. (E) Left: PL images of films made of drop-casted 2D MoS2 in initial and at the intercalating voltages of −4, −6, −8, and −10 V (at a broadband excitation light source covering wavelengths ranging from 400 to 500 nm). The changes in PL of 2D nanoflakes at different applied voltages are presented. Yellow light is emitted from the 2D MoS2 nanoflakes at the initial stage. When a voltage of −4 V is applied to the sample, the PL level is only slightly reduced. When the applied voltage is decreased from −6 to −10 V, PL intensity drops dramatically, and there is a near-total photoquenching at the voltage of −10 V. Right: UV-vis absorbance spectra of the 2D nanoflakes under different electrochemical forces, in which the pristine 2D MoS2 is used as the differential reference. Main absorption peaks in the experimental are observed at 332, 380, 430, 498, and 714 nm. The wavelengths for theoretical estimates of plasmonic peaks as a function of intercalated Li+ ions are listed along dotted lines. An intercalated fluoride-doped tin oxide (FTO) peak is observed at 325 nm, showing a minimal interference. Reproduced with permission from Wang et al. [135]. Copyright 2015 American Chemical Society.
Figure 5:

Foreigner species intercalation.

(A) Schematic for intercalation in the layered TMD materials, where guest species can be inserted into the interlayer gap by diverse methods. Reproduced with permission from Jung et al. [32]. Copyright 2016 the Partner Organisations. (B) Left: Schematic of the band structure of 2H-MoS2 at the K-point in the Brillouin zone without Li intercalation. Optical transitions due to the A and B excitons are depicted. These transitions possibly cause absorption peaks in 2H-MoS2. Right: Schematic of the band structure of 1T-Lix=1MoS2 at the K-point in the Brillouin zone after Li intercalations. 1T-Lix=1MoS2 shows a metallic behavior as the conduction band, and the valence band becomes overlapped, and the Fermi level lies in an incompletely filled band. This metallic band structure with no bandgap substantially reduces the light absorption. (C) Left: Experimental setup for electrochemical tuning of Li concentration in MoS2. Middle: Optical transmission spectra of MoS2 flakes with thicknesses of 9 nm before (red) and after (blue) Li intercalations. Right: Thickness dependence of MoS2 optical transmissions at 500 nm. The results showed enhancement in transmission after Li intercalation. (B) and (C) reproduced with permission from Xiong et al. [31]. Copyright 2015 American Chemical Society. (D) Top: Active control of PL in the 2D MoS2 nanoflakes. PL images of the nanostructured MoS2 film at the intercalating voltages of 0 V, −2 V, −4 V, −6 V, −8V, −10 V, and −12 V. Bottom: In situ PL intensity change of a selected area “1” in the MoS2 film under different intercalating/deintercalating voltages of ±2, ±4, ±6, ±8, ±10, and ±12 V in the 0.1 M LiClO4 solution (blue line) at the step duration of 60 s, whereas a selected uncoated area “2” within the substrate area is chosen as the reference (red line). Reproduced with permission from Wang et al. [134]. Copyright 2013 American Chemical Society. (E) Left: PL images of films made of drop-casted 2D MoS2 in initial and at the intercalating voltages of −4, −6, −8, and −10 V (at a broadband excitation light source covering wavelengths ranging from 400 to 500 nm). The changes in PL of 2D nanoflakes at different applied voltages are presented. Yellow light is emitted from the 2D MoS2 nanoflakes at the initial stage. When a voltage of −4 V is applied to the sample, the PL level is only slightly reduced. When the applied voltage is decreased from −6 to −10 V, PL intensity drops dramatically, and there is a near-total photoquenching at the voltage of −10 V. Right: UV-vis absorbance spectra of the 2D nanoflakes under different electrochemical forces, in which the pristine 2D MoS2 is used as the differential reference. Main absorption peaks in the experimental are observed at 332, 380, 430, 498, and 714 nm. The wavelengths for theoretical estimates of plasmonic peaks as a function of intercalated Li+ ions are listed along dotted lines. An intercalated fluoride-doped tin oxide (FTO) peak is observed at 325 nm, showing a minimal interference. Reproduced with permission from Wang et al. [135]. Copyright 2015 American Chemical Society.

It has been reported that intercalation of alkali metal ions (Li+, Na+, K+) into 2D-layered TMDs induces structural phase change and alters electronic properties, which result in modified optical properties [29], [30], [31], [33], [134], [135]. A commonly studied example is the 2H-to-1T phase transition in Li-intercalated 2D-layered MoS2, which consequently results in alteration of electronic band structure. It is predicted by calculations that the 2H phase of the original 2D MoS2 is semiconducting with a sizeable bandgap, whereas the 1T phase of Li-intercalated MoS2 is metallic with overlapped conduction and valence bands (see schematic in Figure 5B) [31]. This consequently affects many optical properties and opens up new opportunities for optoelectronics. Xiong et al. [31] observe that without bandgap in 1T phase the light transmission in ultrathin MoS2 layers after Li intercalation is substantially enhanced due to the light absorption reduction (Figure 5C). Wang et al. [134] report large modulation of photoluminescence of liquid phase–exfoliated MoS2 nanoflakes by electrochemical controlled ion intercalation, showing significant application prospect in bio-optical sensors, as well as optical modulators or switches (Figure 5D). The observed modulation of photoluminescence is attribute to the lattice expansion, as well as the transition from originally semiconducting 2H into metallic 1T crystal phase after ion intercalation. In addition, they successfully achieved plasmon resonances of 2D MoS2 nanoflakes in the visible and near-UV regimes by the electrochemical intercalating Li+ ions into 2D MoS2 nanoflakes, providing a great application potential for future plasmonic biosensing and optical systems (Figure 5E) [135]. The emergence of plasmon resonances is ascribed to the formation of semimetallic states controlled by doping level of intercalated Li+. The phase change induced by alkali metal intercalation may be attributed to the d-electron count alteration as a result of charge transfer from s orbital of the alkali metal to the d orbital of the transition metal.

Phase change from 1T to 2H has been observed in TaS2 after Li intercalation [137]. Except alkali metals, molecular intercalation in 2D MoS2 also induces phase transition from semiconducting 2H to metallic 1T phase [34]. Local phase transition has been proposed to achieve 2H–1T hybrid structures [138]. The influence of intercalation on optical properties is not limited to 2D TMDs. Many nontransition metal chalcogenides are demonstrated to show evident modulated optical properties after intercalation. Bi2Se3 nanoplates intercalated by zero-valent Cu atoms have been studied to show dramatic enhancement of optical transmission, due to the increase in the effective bandgap caused by the increased free electron density introduced via intercalation of Cu atoms [139]. Organic molecules–intercalated Bi2Se3 nanoribbon has been reported to show a wide tunability of the photonic and plasmonic properties of the host material [140], [141]. In short, the ability to modify the electronic structures and optical properties of 2D TMDs by intercalation provides a new effective control manner for future tunable optoelectronics and attracts a new round of attention.

3.5 Strain engineering

Strain engineering has both theoretically and experimentally proved to be an effective approach to continuously tune the bandgap of TMDs, which subsequently modulates the electronic, optical, and photonic properties of TMDs [84], [85], [86], [87], [88], [89], [90], [91], [142], [143], [144]. The ability to continuously tune the band structure of a TMD is most desirable for optic and photoelectronic applications, motivating a vast array of theoretical works that study the strain effect on band structure. Applying strain on a crystal will cause lattice stretching or compressing in different directions, resulting in the change of lattice constant. This causes the change of overlapping and hybridization of electron orbital and can be responsible for the modified electronic and optical properties. In order to study the influence of mechanical strains on the monolayer group VIB TMDs MoX2 and WX2, Johari and Shenoy [84] performed first-principles density functional theory based calculations and revealed that both tensile and shear strain continuously reduce the bandgap, whereas tensile strain decreases the bandgap more rapidly (Figure 6A). Specially, they found that application of biaxial tensile strain causes a direct-to-indirect bandgap and semiconductor-to-metal transition of the monolayer TMDs. Another theoretical work by Yun et al. [85] found, for the monolayer of MoX2 and WX2, the tensile strain reduces the bandgap, whereas the compressive strain enhances bandgap. For those beyond group VIB TMDs, density functional theory calculations suggest different strain effects. For example, the bandgap of group IVB TMDs (TiX2, ZrX2, HfX2) is predicted to increase with the tensile strain. All these calculations suggest a reversible and flexible manner to engineer the electronic and optic properties of TMDs.

Figure 6: Strain engineering on tuning electronic structures.(A) Bandgap of monolayer TMDs with respect to strain, ε, which varies from 0 to 10%. Strain is applied to the optimized structures (ε=0) through various approaches, such as uniaxial expansion in x-direction (xx), y-direction (yy), homogeneous expansion in both x- and y-directions (xx+yy), expansion in x-direction and compression in y-direction (xx−yy), and compression in x-direction and expansion in y-direction (yy−xx) with the same magnitude of strain. The first three strain profiles correspond to tensile strain, whereas the latter two represent pure shear strain. Left: MoS2. Right: WS2. Reproduced with permission from Johari and Shenoy [84]. Copyright 2012 American Chemical Society. (B) Left: Schematic diagram of a cantilever used in the experiment. It is made of a square-shaped PMMA substrate of length L=9.5 cm and thickness t=0.245 cm. MoS2 samples were deposited near a corner of the substrate. Strain was applied on the samples by clamping one edge (gray) and bending the opposite edge of the substrate. Strain on the sample was calibrated by the sample location x and the cantilever deflection δ as described in the text. Insets: Atomic structure of monolayer MoS2 and optical image of one sample with both monolayer (1L) and bilayer (2L) regions. The scale bar is 5 μm. Right: Absorption (left panel) and PL (right panel) spectrum of a monolayer MoS2 sample under tensile strains up to 0.52% along the zigzag direction. The dashed blue and red lines are guide to the eye of the red shift of the peaks. Reproduced with permission from He et al. [87]. Copyright 2013 American Chemical Society. (C) Top left: Schematic of an inhomogeneously strained membrane ribbon with varying width λ (dashed line). The two electrodes mechanically impose a vertical displacement on the central region of the membrane. Without the two electrodes, the equilibrium geometry of the membrane is flat. Top right: Schematic device setup for elastic strain–engineered artificial atom (not drawn to scale). Notice that although the bulk of the electrodes is metallic (e.g. transparent conducting oxide), they are coated with semiconducting buffers to facilitate selective quasi-particle collection. Bottom left–right: Three broad-spectrum solar energy funneling mechanisms arising from a different band bending and exciton binding profile in the strain-engineered semiconducting membrane. Reproduced with permission from Feng et al. [145]. Copyright 2012 Macmillan Publishers Limited.
Figure 6:

Strain engineering on tuning electronic structures.

(A) Bandgap of monolayer TMDs with respect to strain, ε, which varies from 0 to 10%. Strain is applied to the optimized structures (ε=0) through various approaches, such as uniaxial expansion in x-direction (xx), y-direction (yy), homogeneous expansion in both x- and y-directions (xx+yy), expansion in x-direction and compression in y-direction (xxyy), and compression in x-direction and expansion in y-direction (yyxx) with the same magnitude of strain. The first three strain profiles correspond to tensile strain, whereas the latter two represent pure shear strain. Left: MoS2. Right: WS2. Reproduced with permission from Johari and Shenoy [84]. Copyright 2012 American Chemical Society. (B) Left: Schematic diagram of a cantilever used in the experiment. It is made of a square-shaped PMMA substrate of length L=9.5 cm and thickness t=0.245 cm. MoS2 samples were deposited near a corner of the substrate. Strain was applied on the samples by clamping one edge (gray) and bending the opposite edge of the substrate. Strain on the sample was calibrated by the sample location x and the cantilever deflection δ as described in the text. Insets: Atomic structure of monolayer MoS2 and optical image of one sample with both monolayer (1L) and bilayer (2L) regions. The scale bar is 5 μm. Right: Absorption (left panel) and PL (right panel) spectrum of a monolayer MoS2 sample under tensile strains up to 0.52% along the zigzag direction. The dashed blue and red lines are guide to the eye of the red shift of the peaks. Reproduced with permission from He et al. [87]. Copyright 2013 American Chemical Society. (C) Top left: Schematic of an inhomogeneously strained membrane ribbon with varying width λ (dashed line). The two electrodes mechanically impose a vertical displacement on the central region of the membrane. Without the two electrodes, the equilibrium geometry of the membrane is flat. Top right: Schematic device setup for elastic strain–engineered artificial atom (not drawn to scale). Notice that although the bulk of the electrodes is metallic (e.g. transparent conducting oxide), they are coated with semiconducting buffers to facilitate selective quasi-particle collection. Bottom left–right: Three broad-spectrum solar energy funneling mechanisms arising from a different band bending and exciton binding profile in the strain-engineered semiconducting membrane. Reproduced with permission from Feng et al. [145]. Copyright 2012 Macmillan Publishers Limited.

In experiments, mechanical strain can be applied through bending of samples on flexible substrates or via lattice mismatch between substrates and epitaxy layers. Using a cantilever device to apply a uniaxial tensile strain, it has been observed that the A and B resonance peaks in both absorption and photoluminescence spectrum, corresponding to direct bandgap transitions, show a similar red-shift rate with increasing the tensile strain in both monolayer and bilayer MoS2 (Figure 6B) [87], whereas the I peak related to the indirect bandgap transition in bilayer MoS2 has a larger red-shift rate. This experimental result indicates applying a uniaxial tensile strain can continuously reduce the bandgap of atomic thin MoS2 layer, which agrees well with the first-principles calculations. On the contrary, through a piezoelectric substrate, the compressive strain was applied to trilayer MoS2, and it has been demonstrated to increase the bandgap, manifested as both the photoluminescence peak and Raman modes are blue-shifted [88], [146].

The effective strain tunability of the electronic structure of TMDs promises a wide range of applications in optoelectronics. Applying local tensile strain, TMDs can exhibit spatially varying bandgap [86], and subsequent confinement potentials for photoinduced excitons can be generated to trap excitons for quantum optics, photodetection, and photovoltaics because of the exciton funnel effect that excitons will move to lower bandgap regions induced by locally applied tensile strain before recombining. This funnel effect has been utilized to design a photovoltaic device by introducing inhomogeneous elastic strain [145]. An “artificial atom” made of a nanoindented MoS2 monolayer (Figure 6C) has been proposed and predicted to be able to absorb a broad range of the solar spectrum along the elastic strain gradient and concentrate photoexcited charge carriers. Introducing compressive strain, photoresponsibility is controllably modulated by the piezo-phototronic effect in a flexible MoS2/WSe2 van der Waals photodiode, leading to excellent strain-tunable photodetection performance [52]. Benefiting from the various manners of applying mechanical strains and the richness of TMDs, the development of novel optoelectronics is on the way.

3.6 Heterostructure constructing

Forming heterostructures of materials is a common strategy to modify the electronic band structures of the materials [147]. The heterostructures composed of different 2D TMDs stacked along the out-of-plane direction (vertical heterostructure) or in-plane direction (lateral structure) [148] can exhibit significant different electronic and optical properties compared with each component material, having great potential for atomically thin optoelectronic and photovoltaic applications [75], [149], [150], [151], [152]. For a heterostructure, many aspects, such as the stacking order, lattice mismatch, component layers, and so on, can affect its electronic structure. For example, it is theoretically found that heterobilayers composed of particularly stacked different monolayer TMDs exhibit a direct bandgap, where electrons and holes are physically separated and localized in different layers (Figure 7A) [153]. The predicted direct bandgaps of the heterobilayer TMDs ranging from 0.79 to 1.15 eV are much smaller than the direct bandgaps of monolayer WS2 (2.1 eV) and MoS2 (1.8 eV), which extends the optical applications in the infrared range. Additionally, the band offset calculations show both the conduction band minimum and valence band maximum of WX2 are higher than those of MoX2 due to the higher energy of the 5d orbital of W than that of the 4d orbital of Mo (Figure 7B) [76]. As a result, MoX2/WX2 heterostructures will form a type II band alignment with the conduction band minimum and valence band maximum located in MoX2 and WX2, separately. A type II band–aligned heterostructure, in which free electrons and holes prefer to spontaneously stay at separate positions (Figure 7B), is suitable for optoelectronics and light harvesting. Accordingly, TMD heterostructures enable bandgap engineering, which could further offer more flexible choices for the construction of novel high-performance optoelectronic devices.

Figure 7: Optoelectronics of 2D TMD heterostructures.(A) Left: Bilayer formed by overlapping WSe2 on WS2 with stacking type B denoted as bilayer (WS2, WSe2, B) and bilayer formed by placing WSe2 on MoS2 with stacking type A denoted as bilayer (MoS2, WSe2, A). In staking type B, the transition metal atoms are on top of the chalcogen atoms. In stacking type A, the chalcogen atoms are on top of each other. The dotted lines indicate the alignment of atoms between layers, and the arrows correspond to the distance dS-Se between chalcogen atoms of different layers. Right: Type 2 family band structure showing the direct bandgap of bilayer (WS2, WSe2, B), bilayer (WSe2, MoS2, A), and the infinite number of layers case: Crystal (WS2, WSe2, A). The red line indicates the top of the valence band. Reproduced with permission from Terrones et al. [153]. Copyright 2013 Springer Nature Limited. (B) Left: Calculated band alignment for MX2 monolayers. Solid lines are obtained by PBE, and dashed lines are obtained by HSE06. The dotted lines indicate the water reduction (H+/H2) and oxidation (H2O/O2) potentials. The vacuum level is taken as zero reference. Middle and right: Charge densities of VBM (middle) and CBM (right) states for monolayer WX2-MoX2 lateral heterostructures with common X. Reproduced with permission from Kang et al. [76]. Copyright 2013 American Institute of Physics. (C): (i) Bottom left: Schematic diagram of a van der Waals–stacked MoS2/WSe2 heterojunction device with lateral metal contacts. Top: Enlarged crystal structure, with purple, red, yellow, and green spheres representing Mo, S, W, and Se atoms, respectively. Bottom right: Optical image of the fabricated device, where D1 and D2 (S1 and S2) indicate the metal contacts for WSe2 (MoS2). Scale bar, 3 μm. (ii) Schematic illustrations of exciton dissociation (top) and interlayer recombination (bottom) processes. Top: Horizontal and vertical arrows represent charge transfer and intralayer recombination processes, respectively. Bottom: Red and blue arrows indicate Shockley–Read–Hall (SRH) and Langevin recombination processes, respectively. (iii) Photoresponse characteristics at various gate voltages under white-light illumination. Inset: Color plot of photocurrent as a function of voltages Vds (x axis) and Vg (y axis). The dashed line represents the profile of short-circuit current density Jsc at Vds=0 V. (iv) Top: Photocurrent map of the device presented in (i) for Vds=0 V and 532-nm laser excitation. The junction area and metal electrodes are indicated by dashed and solid lines, respectively. Scale bar, 3 μm. Bottom: Photoluminescence spectra measured from the isolated monolayers (blue curve for MoS2; red curve for WSe2) and the stacked junction region (brown curve). (v) Photoluminescence spatial maps for emission at 1.66 eV (top) and 1.88 eV (bottom), corresponding to direct gap transitions of monolayer WSe2 and MoS2, respectively. The junction area is indicated by dashed lines. Scale bars, 3 μm. Reproduced with permission from Lee et al. [46]. Copyright 2014 Macmillan Publishers Limited.
Figure 7:

Optoelectronics of 2D TMD heterostructures.

(A) Left: Bilayer formed by overlapping WSe2 on WS2 with stacking type B denoted as bilayer (WS2, WSe2, B) and bilayer formed by placing WSe2 on MoS2 with stacking type A denoted as bilayer (MoS2, WSe2, A). In staking type B, the transition metal atoms are on top of the chalcogen atoms. In stacking type A, the chalcogen atoms are on top of each other. The dotted lines indicate the alignment of atoms between layers, and the arrows correspond to the distance dS-Se between chalcogen atoms of different layers. Right: Type 2 family band structure showing the direct bandgap of bilayer (WS2, WSe2, B), bilayer (WSe2, MoS2, A), and the infinite number of layers case: Crystal (WS2, WSe2, A). The red line indicates the top of the valence band. Reproduced with permission from Terrones et al. [153]. Copyright 2013 Springer Nature Limited. (B) Left: Calculated band alignment for MX2 monolayers. Solid lines are obtained by PBE, and dashed lines are obtained by HSE06. The dotted lines indicate the water reduction (H+/H2) and oxidation (H2O/O2) potentials. The vacuum level is taken as zero reference. Middle and right: Charge densities of VBM (middle) and CBM (right) states for monolayer WX2-MoX2 lateral heterostructures with common X. Reproduced with permission from Kang et al. [76]. Copyright 2013 American Institute of Physics. (C): (i) Bottom left: Schematic diagram of a van der Waals–stacked MoS2/WSe2 heterojunction device with lateral metal contacts. Top: Enlarged crystal structure, with purple, red, yellow, and green spheres representing Mo, S, W, and Se atoms, respectively. Bottom right: Optical image of the fabricated device, where D1 and D2 (S1 and S2) indicate the metal contacts for WSe2 (MoS2). Scale bar, 3 μm. (ii) Schematic illustrations of exciton dissociation (top) and interlayer recombination (bottom) processes. Top: Horizontal and vertical arrows represent charge transfer and intralayer recombination processes, respectively. Bottom: Red and blue arrows indicate Shockley–Read–Hall (SRH) and Langevin recombination processes, respectively. (iii) Photoresponse characteristics at various gate voltages under white-light illumination. Inset: Color plot of photocurrent as a function of voltages Vds (x axis) and Vg (y axis). The dashed line represents the profile of short-circuit current density Jsc at Vds=0 V. (iv) Top: Photocurrent map of the device presented in (i) for Vds=0 V and 532-nm laser excitation. The junction area and metal electrodes are indicated by dashed and solid lines, respectively. Scale bar, 3 μm. Bottom: Photoluminescence spectra measured from the isolated monolayers (blue curve for MoS2; red curve for WSe2) and the stacked junction region (brown curve). (v) Photoluminescence spatial maps for emission at 1.66 eV (top) and 1.88 eV (bottom), corresponding to direct gap transitions of monolayer WSe2 and MoS2, respectively. The junction area is indicated by dashed lines. Scale bars, 3 μm. Reproduced with permission from Lee et al. [46]. Copyright 2014 Macmillan Publishers Limited.

Experimentally, TMD heterostructures are widely available through sequential transfer or various growth methods [149]. Stimulated by the graphene stacked on h-BN showing ultrahigh mobility, concerted efforts are expected to explore the modified physical properties of heterostructures formed by different 2D layers. In case of TMDs, high-quality heterostructures have been demonstrated not only to improve electrical transport properties, but also to enhance the optoelectronic performance [41], [44], [46], [50], [78], [80], [81], [154]. Researchers observed significant photoluminescence quenching and gate-tunable efficient photocurrent generation in a vertical heterostructure with the n-type monolayer MoS2 transferred onto the top of a p-type monolayer WSe2 (Figure 7C) [46]. The observed gate-tunable photovoltaic response originates from spontaneous dissociation of photogenerated excitons into free electrons and holes in different layers and subsequent tunneling-assisted interlayer recombination driven by large band offsets in a heterostructure with type II band alignment. The charge transfer processes in photoexcited MoS2/WS2 heterostructures have been proved to be ultrafast through photoluminescence mapping and femtosecond pump–probe spectroscopy [78]. It is found that hole transfers from the MoS2 layer to the WS2 layer within 50 fs after optical excitation.

Besides the heterostructures composed of two different TMDs, there are a lot of heterostructures formed by TMDs with other 2D materials, showing tunable optoelectronic properties [49]. The first to mention is the TMDs/graphene heterostructures [41], [44] because their distinct properties could be utilized to realize novel functionalities. The high mobility of graphene means fast response rate, whereas the large direct bandgap of TMDs results in strong light absorption as mentioned earlier. It is observed the strong light–matter interactions in TMDs/graphene heterostructures lead to enhanced photon absorption and photocurrent generation, allowing development of extremely efficient flexible photonic and optoelectronic devices [41]. Besides, ultrathin p-GaTe/n-MoS2 heterostructure, which poses the type II band alignment, is constructed and proved to have high photovoltaic and photodetecting performance [50]. The light-emitting diodes based on n-type monolayer MoS2 and p-type Si stacked vertically are fabricated to realize rectification and light emission from the entire surface of the heterojunction [155], [156].

It is worth to mention that geometrical alignments between different layers in heterostructures also play an important role in tuning the electronic structure. Heterostructures, including TMD bilayers, with well-defined interlayer twist angle can be obtained via transfer one monolayer onto the top of another. Experimental and theoretical study found that the interlayer twist angle has a strong influence on the indirect optical transition energy and second-harmonic generation [81]. As the interlayer twist affects the interlayer distance, which determines the interlayer interactions, band structures engineering by tuning twist angle is available for optoelectronic applications.

4 Conclusions and outlook

We have summarized several strategies on the electronic structures modulation of 2D TMDs, including thinning thickness down, defect engineering, chemical composition modification, foreigner species intercalation, strain engineering, and heterostructure construction, to tune its optical absorption, charge carrier’s mobility, and band structure alignments, thus adapting for the different optoelectronic devices. As the great development of nanofabrication techniques, it is more convenient to realize electronic and optical properties modification and stay compatible with optoelectronic devices, through tuning strategies such as thinning thickness down, chemical composition modification, and heterostructure constructing. Foreigner species intercalation has been identified as an effective modulation method, but still is challenged with ordering stability of intercalated atoms. Strain engineering is proved to be an effective approach to continuously tune the bandgap of TMDs but is challenged with how to quantitatively apply strain in devices. At the same time, there are much more fruitful tuning freedoms for 2D TMDs or the whole family of 2D materials, whose electronic structures and optical properties can be coupled with magnetic properties, biological compatibility, surface plasmonic structures, piezoelectricity/ferroelectricity, electrochemical activities, and so on, for designing the multifunctional device application [157], [158], [159], [160], [161], [162], [163], [164]. On the other hand, considering the preparation methods, we can expect the combination of diverse tuning methods or large-scale preparation with retaining precise modulated properties [165]; thus, we can implement the tunable optoelectronic devices not only in the fundamental research but also for the industrial-level application [166]. With the rapid development of controlled synthesis and physical and chemical approaches, the multifunctional magnetic and mechanical response, thermal transport devices [167], and electrochemical storage and conversion devices [168], [169] will be continuously springing up [170], [171], [172], [173], [174], [175], [176], [177], [178], [179].

References

[1] Chhowalla M, Shin HS, Eda G, Li LJ, Loh KP, Zhang H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nature Chem. 2013;5:263–75.10.1038/nchem.1589Search in Google Scholar PubMed

[2] Wilson JA, Yoffe AD. Transition metal dichalcogenides discussion and interpretation of observed optical, electrical and structural properties. Adv Phys 1969;18:193–335.10.1080/00018736900101307Search in Google Scholar

[3] Lopez-Sanchez O, Lembke D, Kayci M, Radenovic A, Kis A. Ultrasensitive photodetectors based on monolayer MoS2. Nat Nanotechnol 2013;8:497–501.10.1038/nnano.2013.100Search in Google Scholar PubMed

[4] Bie Y-Q, Grosso G, Heuck M, et al. A MoTe2-based light-emitting diode and photodetector for silicon photonic integrated circuits. Nat Nanotechnol 2017;12:1124–9.10.1038/nnano.2017.209Search in Google Scholar PubMed

[5] Jo S, Ubrig N, Berger H, Kuzmenko AB, Morpurgo AF. Mono- and bilayer WS2 light-emitting transistors. Nano Lett 2014;14:2019–25.10.1021/nl500171vSearch in Google Scholar PubMed

[6] Wang H, Li C, Fang P, Zhang Z, Zhang JZ. Synthesis, properties, and optoelectronic applications of two-dimensional MoS2 and MoS2-based heterostructures. Chem Soc Rev 2018;47:6101–27.10.1039/C8CS00314ASearch in Google Scholar PubMed

[7] Li H, Wu JMT, Yin ZY, Zhang H. Preparation and applications of mechanically exfoliated single-layer and multi-layer MoS2 and WSe2 nanosheets. Acc Chem Res 2014;47:1067–75.10.1021/ar4002312Search in Google Scholar PubMed

[8] Butler SZ, Hollen SM, Cao L, et al. Progress, challenges, and opportunities in two-dimensional materials beyond graphene. ACS Nano 2013;7:2898–926.10.1021/nn400280cSearch in Google Scholar PubMed

[9] Wang H, Yuan H, Sae Hong S, Li Y, Cui Y. Physical and chemical tuning of two-dimensional transition metal dichalcogenides. Chem Soc Rev 2015;44:2664–80.10.1039/C4CS00287CSearch in Google Scholar

[10] Guo Z, Zhang H, Lu S, et al. From black phosphorus to phosphorene: basic solvent exfoliation, evolution of Raman scattering, and applications to ultrafast photonics. Adv Funct Mater 2015;25:6996–7002.10.1002/adfm.201502902Search in Google Scholar

[11] Sang KD, Wang H, Qiu M, et al. Two dimensional β-InSe with layer-dependent properties: band alignment, work function and optical properties. Nanomaterials 2019;9:82.10.3390/nano9010082Search in Google Scholar PubMed PubMed Central

[12] Mak KF, Lee C, Hone J, Shan J, Heinz TF. Atomically thin MoS2: a new direct-gap semiconductor. Phys Rev Lett 2010;105:136805.10.1103/PhysRevLett.105.136805Search in Google Scholar PubMed

[13] Keum DH, Cho S, Kim JH, et al. Bandgap opening in few-layered monoclinic MoTe2. Nat Phys 2015;11:482–6.10.1038/nphys3314Search in Google Scholar

[14] Barrera D, Wang QX, Lee YJ, et al. Solution synthesis of few-layer 2H-MX2 (M=Mo, W; X=S, Se). J Mater Chem C 2017;5:2859–64.10.1039/C6TC05097BSearch in Google Scholar

[15] Kalanyan B, Kimes WA, Beams R, et al. Rapid wafer-scale growth of polycrystalline 2H-MoS2 by pulsed metal–organic chemical vapor deposition. Chem Mater 2017;29:6279–88.10.1021/acs.chemmater.7b01367Search in Google Scholar PubMed PubMed Central

[16] Zeng ZXS, Sun XX, Zhang DL, et al. Controlled vapor growth and nonlinear optical applications of large-area 3R phase WS2 and WSe2 atomic layers. Adv Funct Mater 2019;29:1806874.10.1002/adfm.201806874Search in Google Scholar

[17] Horiba K, Ono K, Oh JH, et al. Charge-density wave and three-dimensional Fermi surface in 1T-TaSe2 studied by photoemission spectroscopy. Phys Rev B 2002;66:073106.10.1103/PhysRevB.66.073106Search in Google Scholar

[18] Aiura Y, Bando H, Kitagawa R, et al. Electronic structure of layered 1T-TaSe2 in commensurate charge-density-wave phase studied by angle-resolved photoemission spectroscopy. Phys Rev B 2003;68:073408.10.1103/PhysRevB.68.073408Search in Google Scholar

[19] Bovet M, Popovic D, Clerc F, et al. Pseudogapped Fermi surfaces of 1T-TaS2 and 1T-TaSe2: a charge density wave effect. Phys Rev B 2004;69:125117.10.1103/PhysRevB.69.125117Search in Google Scholar

[20] Suzuki M-T, Harima H. Electronic band structures and charge density wave of 2H-MX2(M=Nb,Ta,X=S,Se). Phys B 2005;359–361:1180–2.10.1016/j.physb.2005.01.336Search in Google Scholar

[21] Wu Y, Xing H, Lian C-S, et al. Ion intercalation engineering of electronic properties of two-dimensional crystals of 2H-TaSe2. Phys Rev Mater 2019;3:104003.10.1103/PhysRevMaterials.3.104003Search in Google Scholar

[22] Martinez H, Auriel C, Gonbeau D, Loudet M, PfisterGuillouzo G. Studies of 1T TiS2 by STM, AFM and XPS: the mechanism of hydrolysis in air. Appl Surf Sci 1996;93:231–5.10.1016/0169-4332(95)00339-8Search in Google Scholar

[23] Fang CM, deGroot RA, Haas C. Bulk and surface electronic structure of 1T-TiS2 and 1T-TiSe2. Phys Rev B 1997;56: 4455–63.10.1103/PhysRevB.56.4455Search in Google Scholar

[24] Slough CG, Giambattista B, Johnson A, McNairy WW, Wang C, Coleman RV. Scanning tunneling microscopy of 1T-TiSe2 and 1T-TiS2 at 77 and 4.2 K. Phys Rev B 1988;37:6571–4.10.1103/PhysRevB.37.6571Search in Google Scholar PubMed

[25] Sharma S, Nautiyal T, Singh GS, Auluck S, Blaha P, Ambrosch-Draxl C. Electronic structure of 1T-TiS2. Phys Rev B 1999;59:14833–6.10.1103/PhysRevB.59.14833Search in Google Scholar

[26] Kidd TE, Miller T, Chou MY, Chiang TC. Electron–hole coupling and the charge density wave transition in TiSe2. Phys Rev Lett 2002;88:226402.10.1103/PhysRevLett.88.226402Search in Google Scholar PubMed

[27] Cercellier H, Monney C, Clerc F, et al. Evidence for an excitonic insulator phase in 1T-TiSe2. Phys Rev Lett 2007;99:146403.10.1103/PhysRevLett.99.146403Search in Google Scholar PubMed

[28] Benavente E, Santa Ana MA, Mendizabal F, Gonzalez G. Intercalation chemistry of molybdenum disulfide. Coord Chem Rev 2002;224:87–109.10.1016/S0010-8545(01)00392-7Search in Google Scholar

[29] Eda G, Yamaguchi H, Voiry D, Fujita T, Chen MW, Chhowalla M. Photoluminescence from chemically exfoliated MoS2. Nano Lett 2011;11:5111–6.10.1021/nl201874wSearch in Google Scholar PubMed

[30] Guo YS, Sun DZ, Ouyang B, et al. Probing the dynamics of the metallic-to-semiconducting structural phase transformation in MoS2 crystals. Nano Lett 2015;15:5081–8.10.1021/acs.nanolett.5b01196Search in Google Scholar PubMed

[31] Xiong F, Wang H, Liu X, et al. Li intercalation in MoS2: In situ observation of its dynamics and tuning optical and electrical properties. Nano Lett 2015;15:6777–84.10.1021/acs.nanolett.5b02619Search in Google Scholar PubMed

[32] Jung Y, Zhou Y, Cha JJ. Intercalation in two-dimensional transition metal chalcogenides. Inorg Chem Front 2016;3:452–63.10.1039/C5QI00242GSearch in Google Scholar

[33] Xia J, Wang J, Chao DL, et al. Phase evolution of lithium intercalation dynamics in 2H-MoS2. Nanoscale 2017;9:7533–40.10.1039/C7NR02028GSearch in Google Scholar

[34] He QY, Lin ZY, Ding MN, et al. In situ probing molecular intercalation in two-dimensional layered semiconductors. Nano Lett 2019;19:6819–26.10.1021/acs.nanolett.9b01898Search in Google Scholar PubMed

[35] Bissessur R, Kanatzidis MG, Schindler JL, Kannewurf CR. Encapsulation of polymers into MoS2 and metal to insulator transition in metastable MoS2. J Chem Soc Chem Commun 1993;1582–5.10.1039/c39930001582Search in Google Scholar

[36] Kobayashi K, Yamauchi J. Electronic-structure and scanning-tunneling-microscopy image of molybdenum dichalcogenide surfaces. Phys Rev B 1995;51:17085–95.10.1103/PhysRevB.51.17085Search in Google Scholar

[37] Radisavljevic B, Radenovic A, Brivio J, Giacometti V, Kis A. Single-layer MoS2 transistors. Nat Nanotechnol 2011;6: 147–50.10.1038/nnano.2010.279Search in Google Scholar PubMed

[38] Terrones H, Terrones M. Electronic and vibrational properties of defective transition metal dichalcogenide Haeckelites: new 2D semi-metallic systems. 2D Mater 2014;1:011003.10.1088/2053-1583/1/1/011003Search in Google Scholar

[39] Hinsche NF, Thygesen KS. Electron-phonon interaction and transport properties of metallic bulk and monolayer transition metal dichalcogenide TaS2. 2D Mater 2018;5:015009.10.1088/2053-1583/aa8e6cSearch in Google Scholar

[40] Lee HS, Min SW, Chang YG, et al. MoS2 nanosheet phototransistors with thickness-modulated optical energy gap. Nano Lett 2012;12:3695–700.10.1021/nl301485qSearch in Google Scholar PubMed

[41] Britnell L, Ribeiro RM, Eckmann A, et al. Strong light–matter interactions in heterostructures of atomically thin films. Science 2013;340:1311–4.10.1126/science.1235547Search in Google Scholar PubMed

[42] Tsai DS, Liu KK, Lien DH, et al. Few-layer MoS2 with high broadband photogain and fast optical switching for use in harsh environments. ACS Nano 2013;7:3905–11.10.1021/nn305301bSearch in Google Scholar PubMed

[43] Sundaram RS, Engel M, Lombardo A, et al. Electroluminescence in single layer MoS2. Nano Lett 2013;13:1416–21.10.1021/nl400516aSearch in Google Scholar PubMed

[44] Yu WJ, Liu Y, Zhou HL, et al. Highly efficient gate-tunable photocurrent generation in vertical heterostructures of layered materials. Nat Nanotechnol 2013;8:952–8.10.1038/nnano.2013.219Search in Google Scholar PubMed PubMed Central

[45] Ross JS, Klement P, Jones AM, et al. Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions. Nat Nanotechnol 2014;9:268–72.10.1038/nnano.2014.26Search in Google Scholar PubMed

[46] Lee CH, Lee GH, van der Zande AM, et al. Atomically thin p-n junctions with van der Waals heterointerfaces. Nat Nanotechnol 2014;9:676–81.10.1038/nnano.2014.150Search in Google Scholar PubMed

[47] Baugher BWH, Churchill HOH, Yang Y, Jarillo-Herrero P. Optoelectronic devices based on electrically tunable p-n diodes in a monolayer dichalcogenide. Nat Nanotechnol 2014;9:262–7.10.1038/nnano.2014.25Search in Google Scholar PubMed

[48] Pospischil A, Furchi MM, Mueller T. Solar-energy conversion and light emission in an atomic monolayer p-n diode. Nat Nanotechnol 2014;9:257–61.10.1038/nnano.2014.14Search in Google Scholar PubMed

[49] Withers F, Del Pozo-Zamudio O, Mishchenko A, et al. Light-emitting diodes by band-structure engineering in van der Waals heterostructures. Nat Mater 2015;14:301–6.10.1038/nmat4205Search in Google Scholar PubMed

[50] Wang F, Wang ZX, Xu K, et al. Tunable GaTe-MoS2 van der Waals p-n Junctions with novel optoelectronic performance. Nano Lett 2015;15:7558–66.10.1021/acs.nanolett.5b03291Search in Google Scholar PubMed

[51] Massicotte M, Schmidt P, Vialla F, et al. Picosecond photoresponse in van der Waals heterostructures. Nat Nanotechnol 2016;11:42–6.10.1038/nnano.2015.227Search in Google Scholar PubMed

[52] Lin P, Zhu L, Li D, Xu L, Pan C, Wang Z. Piezo-phototronic effect for enhanced flexible MoS2/WSe2 van der Waals photodiodes. Adv Funct Mater 2018;28:1802849.10.1002/adfm.201802849Search in Google Scholar

[53] Klein A, Tiefenbacher S, Eyert V, Pettenkofer C, Jaegermann W. Electronic band structure of single-crystal and single-layer WS2: influence of interlayer van der Waals interactions. Phys Rev B 2001;64:205416.10.1103/PhysRevB.64.205416Search in Google Scholar

[54] Ellis JK, Lucero MJ, Scuseria GE. The indirect to direct band gap transition in multilayered MoS2 as predicted by screened hybrid density functional theory. Appl Phys Lett 2011;99:261908.10.1063/1.3672219Search in Google Scholar

[55] Kuc A, Zibouche N, Heine T. Influence of quantum confinement on the electronic structure of the transition metal sulfide TS2. Phys Rev B 2011;83:245213.10.1103/PhysRevB.83.245213Search in Google Scholar

[56] Peelaers H, Van de Walle CG. Effects of strain on band structure and effective masses in MoS2. Phys Rev B 2012;86:241401.10.1103/PhysRevB.86.241401Search in Google Scholar

[57] Komsa HP, Krasheninnikov AV. Effects of confinement and environment on the electronic structure and exciton binding energy of MoS2 from first principles. Phys Rev B 2012;86:241201.10.1103/PhysRevB.86.241201Search in Google Scholar

[58] Cheiwchanchamnangij T, Lambrecht WRL. Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2. Phys Rev B 2012;85:205302.10.1103/PhysRevB.85.205302Search in Google Scholar

[59] Splendiani A, Sun L, Zhang YB, et al. Emerging photoluminescence in monolayer MoS2. Nano Lett 2010;10:1271–5.10.1021/nl903868wSearch in Google Scholar PubMed

[60] Ramasubramaniam A. Large excitonic effects in monolayers of molybdenum and tungsten dichalcogenides. Phys Rev B 2012;86:115409.10.1103/PhysRevB.86.115409Search in Google Scholar

[61] Mak KF, He KL, Lee C, et al. Tightly bound trions in monolayer MoS2. Nat Mater 2013;12:207–11.10.1038/nmat3505Search in Google Scholar PubMed

[62] Berkelbach TC, Hybertsen MS, Reichman DR. Theory of neutral and charged excitons in monolayer transition metal dichalcogenides. Phys Rev B 2013;88:045318.10.1103/PhysRevB.88.045318Search in Google Scholar

[63] Qiu DY, da Jornada FH, Louie SG. Optical spectrum of MoS2: many-body effects and diversity of exciton states. Phys Rev Lett 2013;111:216805.10.1103/PhysRevLett.111.216805Search in Google Scholar PubMed

[64] Chernikov A, Berkelbach TC, Hill HM, et al. Exciton binding energy and nonhydrogenic Rydberg series in monolayer WS2. Phys Rev Lett 2014;113:076802.10.1103/PhysRevLett.113.076802Search in Google Scholar PubMed

[65] Ye ZL, Cao T, O’Brien K, et al. Probing excitonic dark states in single-layer tungsten disulphide. Nature 2014;513:214–8.10.1038/nature13734Search in Google Scholar PubMed

[66] Mann J, Ma Q, Odenthal PM, et al. 2-dimensional transition metal dichalcogenides with tunable direct band gaps: MoS2(1−x)Se2x monolayers. Adv Mater 2014;26: 1399–404.10.1002/adma.201304389Search in Google Scholar PubMed

[67] Wang G, Marie X, Gerber I, et al. Giant enhancement of the optical second-harmonic emission of WSe2 monolayers by laser excitation at exciton resonances. Phys Rev Lett 2015;114:097403.10.1103/PhysRevLett.114.097403Search in Google Scholar PubMed

[68] Zhu ZY, Cheng YC, Schwingenschlogl U. Giant spin-orbit-induced spin splitting in two-dimensional transition-metal dichalcogenide semiconductors. Phys Rev B 2011;84:153402.10.1103/PhysRevB.84.153402Search in Google Scholar

[69] Sun LF, Yan JX, Zhan D, et al. Spin-orbit splitting in single-layer MoS2 revealed by triply resonant Raman scattering. Phys Rev Lett 2013;111:126801.10.1103/PhysRevLett.111.126801Search in Google Scholar PubMed

[70] Zhang Y, Chang TR, Zhou B, et al. Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2. Nat Nanotechnol 2014;9:111–5.10.1038/nnano.2013.277Search in Google Scholar PubMed

[71] Jin W, Yeh P-C, Zaki N, et al. Direct measurement of the thickness-dependent electronic band structure of MoS2 using angle-resolved photoemission spectroscopy. Phys Rev Lett 2013;111:106801.10.1103/PhysRevLett.111.106801Search in Google Scholar PubMed

[72] Li JB, Xiao S, Liang S, et al. Switching freely between superluminal and subluminal light propagation in a monolayer MoS2 nanoresonator. Opt Express 2017;25:13567–76.10.1364/OE.25.013567Search in Google Scholar PubMed

[73] Mouri S, Miyauchi Y, Matsuda K. Tunable photoluminescence of monolayer MoS2 via chemical doping. Nano Lett 2013;13: 5944–8.10.1021/nl403036hSearch in Google Scholar PubMed

[74] Duan XD, Wang C, Fan Z, et al. Synthesis of WS2xSe2-2x alloy nanosheets with composition-tunable electronic properties. Nano Lett 2016;16:264–9.10.1021/acs.nanolett.5b03662Search in Google Scholar PubMed

[75] Geim AK, Grigorieva IV. Van der Waals heterostructures. Nature 2013;499:419–25.10.1038/nature12385Search in Google Scholar PubMed

[76] Kang J, Tongay S, Zhou J, Li JB, Wu JQ. Band offsets and heterostructures of two-dimensional semiconductors. Appl Phys Lett 2013;102:012111.10.1063/1.4774090Search in Google Scholar

[77] Duan XD, Wang C, Shaw JC, et al. Lateral epitaxial growth of two-dimensional layered semiconductor heterojunctions. Nat Nanotechnol 2014;9:1024–30.10.1038/nnano.2014.222Search in Google Scholar PubMed

[78] Hong XP, Kim J, Shi SF, et al. Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures. Nat Nanotechnol 2014;9:682–6.10.1038/nnano.2014.167Search in Google Scholar PubMed

[79] Lu N, Guo HY, Wang L, Wu XJ, Zeng XC. Van der Waals trilayers and superlattices: modification of electronic structures of MoS2 by intercalation. Nanoscale 2014;6:4566–71.10.1039/C4NR00783BSearch in Google Scholar PubMed

[80] Tongay S, Fan W, Kang J, et al. Tuning interlayer coupling in large-area heterostructures with CVD-grown MoS2 and WS2 monolayers. Nano Lett 2014;14:3185–90.10.1021/nl500515qSearch in Google Scholar PubMed

[81] van der Zande AM, Kunstmann J, Chernikov A, et al. Tailoring the electronic structure in bilayer molybdenum disulfide via interlayer twist. Nano Lett 2014;14:3869–75.10.1021/nl501077mSearch in Google Scholar PubMed

[82] Howell SL, Jariwala D, Wu CC, et al. Investigation of band-offsets at monolayer-multilayer MoS2 junctions by scanning photocurrent microscopy. Nano Lett 2015;15:2278–84.10.1021/nl504311pSearch in Google Scholar PubMed

[83] Xie DD, Hu WN, Jiang J. Bidirectionally-trigged 2D MoS2 synapse through coplanar-gate electric-double-layer polymer coupling for neuromorphic complementary spatiotemporal learning. Org Electron 2018;63:120–8.10.1016/j.orgel.2018.09.007Search in Google Scholar

[84] Johari P, Shenoy VB. Tuning the electronic properties of semiconducting transition metal dichalcogenides by applying mechanical strains. ACS Nano 2012;6:5449–56.10.1021/nn301320rSearch in Google Scholar PubMed

[85] Yun WS, Han SW, Hong SC, Kim IG, Lee JD. Thickness and strain effects on electronic structures of transition metal dichalcogenides: 2H-MX2 semiconductors (M=Mo, W; X=S, Se, Te). Phys Rev B 2012;85:033305.10.1103/PhysRevB.85.033305Search in Google Scholar

[86] Castellanos-Gomez A, Roldan R, Cappelluti E, et al. Local strain engineering in atomically thin MoS2. Nano Lett 2013;13:5361–6.10.1021/nl402875mSearch in Google Scholar PubMed

[87] He K, Poole C, Mak KF, Shan J. Experimental demonstration of continuous electronic structure tuning via strain in atomically thin MoS2. Nano Lett 2013;13:2931–6.10.1021/nl4013166Search in Google Scholar PubMed

[88] Hui YY, Liu XF, Jie WJ, et al. Exceptional tunability of band energy in a compressively strained trilayer MoS2 sheet. ACS Nano 2013;7:7126–31.10.1021/nn4024834Search in Google Scholar PubMed

[89] Guo HY, Lu N, Wang L, Wu XJ, Zeng XC. Tuning electronic and magnetic properties of early transition-metal dichalcogenides via tensile strain. J Phys Chem C 2014;118:7242–9.10.1021/jp501734sSearch in Google Scholar

[90] Zhao Z, Zhang H, Yuan H, et al. Pressure induced metallization with absence of structural transition in layered molybdenum diselenide. Nat Commun 2015;6:7312.10.1038/ncomms8312Search in Google Scholar PubMed PubMed Central

[91] Li PF, Li L, Zeng XC. Tuning the electronic properties of monolayer and bilayer PtSe2 via strain engineering. J Mater Chem C 2016;4:3106–12.10.1039/C6TC00130KSearch in Google Scholar

[92] Peng K, Fu LJ, Yang HM, Ouyang J, Tang AD. Hierarchical MoS2 intercalated clay hybrid nanosheets with enhanced catalytic activity. Nano Res 2017;10:570–83.10.1007/s12274-016-1315-3Search in Google Scholar

[93] Zou H, Zeng QY, Peng MQ, Zhou WZ, Dai XY, Ouyang FP. Electronic structures and optical properties of P and Cl atoms adsorbed/substitutionally doped monolayer MoS2. Solid State Commun 2018;280:6–12.10.1016/j.ssc.2018.05.005Search in Google Scholar

[94] Guo JJ, Xie DD, Yang BC, Jiang J. Low-power logic computing realized in a single electric-double-layer MoS2 transistor gated with polymer electrolyte. Solid-State Electron 2018;144:1–6.10.1016/j.sse.2018.02.007Search in Google Scholar

[95] Ramasubramaniam A, Naveh D, Towe E. Tunable band gaps in bilayer transition-metal dichalcogenides. Phys Rev B 2011;84:205325.10.1103/PhysRevB.84.205325Search in Google Scholar

[96] Newaz AKM, Prasai D, Ziegler JI, et al. Electrical control of optical properties of monolayer MoS2. Solid State Commun 2013;155:49–52.10.1016/j.ssc.2012.11.010Search in Google Scholar

[97] Xiao J, Long MQ, Li XM, Zhang QT, Xu H, Chan KS. Effects of van der Waals interaction and electric field on the electronic structure of bilayer MoS2. J Phys Condens Mat 2014;26:405302.10.1088/0953-8984/26/40/405302Search in Google Scholar PubMed

[98] Chu T, Ilatikhameneh H, Klimeck G, Rahman R, Chen ZH. Electrically tunable bandgaps inbilayer MoS2. Nano Lett 2015;15:8000–7.10.1021/acs.nanolett.5b03218Search in Google Scholar PubMed

[99] Zhang H, Zhou WZ, Liu Q, et al. Transport properties and device-design of Z-shaped MoS2 nanoribbon planar junctions. Phys E 2017;93:143–7.10.1016/j.physe.2017.06.004Search in Google Scholar

[100] Guo JJ, Jiang J, Yang BC. Low-voltage electric-double-layer MoS2 transistor gated via water solution. Solid-State Electron 2018;150:8–15.10.1016/j.sse.2018.10.001Search in Google Scholar

[101] Sang DK, Wang H, Guo Z, Xie N, Zhang H. Recent developments in stability and passivation techniques of phosphorene toward next-generation device applications. Adv Funct Mater 2019;29:1903419.10.1002/adfm.201903419Search in Google Scholar

[102] Guo Z, Chen S, Wang Z, et al. Metal-ion–modified black phosphorus with enhanced stability and transistor performance. Adv Mater 2017;29:1703811.10.1002/adma.201703811Search in Google Scholar PubMed

[103] Kang K, Xie SE, Huang LJ, et al. High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity. Nature 2015;520:656–60.10.1038/nature14417Search in Google Scholar PubMed

[104] Lehtinen O, Komsa HP, Pulkin A, et al. Atomic scale microstructure and properties of Se-deficient two-dimensional MoSe2. ACS Nano 2015;9:3274–83.10.1021/acsnano.5b00410Search in Google Scholar PubMed

[105] Wang SS, Wang XC, Warner JH. All chemical vapor deposition growth of MoS2:h-BN vertical van der Waals heterostructures. ACS Nano 2015;9:5246–54.10.1021/acsnano.5b00655Search in Google Scholar PubMed

[106] Naylor CH, Parkin WM, Ping JL, et al. Monolayer single-crystal 1T′-MoTe2 grown by chemical vapor deposition exhibits weak antilocalization effect. Nano Lett 2016;16:4297–304.10.1021/acs.nanolett.6b01342Search in Google Scholar PubMed PubMed Central

[107] Hao S, Yang BC, Gao YL. Orientation-specific transgranular fracture behavior of CVD-grown monolayer MoS2 single crystal. Appl Phys Lett 2017;110:153105.10.1063/1.4979974Search in Google Scholar

[108] Eichfeld SM, Hossain L, Lin YC, et al. Highly scalable, atomically thin WSe2 grown via metal–organic chemical vapor deposition. ACS Nano 2015;9:2080–7.10.1021/nn5073286Search in Google Scholar PubMed

[109] Eichfeld SM, Colon VO, Nie YF, Cho K, Robinson JA. Controlling nucleation of monolayer WSe2 during metal–organic chemical vapor deposition growth. 2D Mater 2016;3:025015.10.1088/2053-1583/3/2/025015Search in Google Scholar

[110] Yin Z, Li H, Li H, et al. Single-layer MoS2 phototransistors. ACS Nano 2012;6:74–80.10.1021/nn2024557Search in Google Scholar PubMed

[111] Huang H, Ren X, Li Z, et al. Two-dimensional bismuth nanosheets as prospective photo-detector with tunable optoelectronic performance. Nanotechnology 2018;29:235201.10.1088/1361-6528/aab6eeSearch in Google Scholar PubMed

[112] Lebègue S, Eriksson O. Electronic structure of two-dimensional crystals from ab initio theory. Phys Rev B 2009;79:115409.10.1103/PhysRevB.79.115409Search in Google Scholar

[113] Dong YL, Zeng BW, Xiao J, et al. Effect of sulphur vacancy and interlayer interaction on the electronic structure and spin splitting of bilayer MoS2. J. Phys.-Condens. Mat. 2018;30:125302.10.1088/1361-648X/aaad3bSearch in Google Scholar PubMed

[114] Boker T, Severin R, Muller A, et al. Band structure of MoS2, MoSe2, and alpha-MoTe2: angle-resolved photoelectron spectroscopy and ab initio calculations. Phys Rev B 2001;64:235305.10.1103/PhysRevB.64.235305Search in Google Scholar

[115] Ding Y, Wang YL, Ni J, Shi L, Shi SQ, Tang WH. First principles study of structural, vibrational and electronic properties of graphene-like MX2 (M=Mo, Nb, W, Ta; X=S, Se, Te) monolayers. Phys B 2011;406:2254–60.10.1016/j.physb.2011.03.044Search in Google Scholar

[116] Ataca C, Sahin H, Ciraci S. Stable, single-layer MX2 transition-metal oxides and dichalcogenides in a honeycomb-like structure. J Phys Chem C 2012;116:8983–99.10.1021/jp212558pSearch in Google Scholar

[117] Xiao P, Mao J, Ding K, et al. Solution-processed 3D RGO–MoS2/pyramid Si heterojunction for ultrahigh detectivity and ultra-broadband photodetection. Adv Mater 2018;30:1801729.10.1002/adma.201801729Search in Google Scholar PubMed

[118] Hong J, Hu Z, Probert M, et al. Exploring atomic defects in molybdenum disulphide monolayers. Nat Commun 2015;6:6293.10.1038/ncomms7293Search in Google Scholar PubMed PubMed Central

[119] Ye G, Gong Y, Lin J, et al. Defects engineered monolayer MoS2 for improved hydrogen evolution reaction. Nano Lett 2016;16:1097–103.10.1021/acs.nanolett.5b04331Search in Google Scholar PubMed

[120] Nan H, Wang Z, Wang W, et al. Strong photoluminescence enhancement of MoS2 through defect engineering and oxygen bonding. ACS Nano 2014;8:5738–45.10.1021/nn500532fSearch in Google Scholar PubMed

[121] Komsa HP, Krasheninnikov AV. Two-dimensional transition metal dichalcogenide alloys: stability and electronic properties. J Phys Chem Lett 2012;3:3652–6.10.1021/jz301673xSearch in Google Scholar PubMed

[122] Chen YF, Xi JY, Dumcenco DO, et al. Tunable band gap photoluminescence from atomically thin transition-metal dichalcogenide alloys. ACS Nano 2013;7:4610–6.10.1021/nn401420hSearch in Google Scholar PubMed

[123] Kim IS, Sangwan VK, Jariwala D, et al. Influence of stoichiometry on the optical and electrical properties of chemical vapor deposition derived MoS2. ACS Nano 2014;8:10551–8.10.1021/nn503988xSearch in Google Scholar PubMed PubMed Central

[124] Kutana A, Penev ES, Yakobson BI. Engineering electronic properties of layered transition-metal dichalcogenide compounds through alloying. Nanoscale 2014;6:5820–5.10.1039/C4NR00177JSearch in Google Scholar

[125] Tan TL, Ng M-F, Eda G. Stable monolayer transition metal dichalcogenide ordered alloys with tunable electronic properties. J Phys Chem C 2016;120:2501–8.10.1021/acs.jpcc.5b10739Search in Google Scholar

[126] Park J, Kim MS, Park B, et al. Composition-tunable synthesis of large-scale Mo1-xWxS2 alloys with enhanced photoluminescence. ACS Nano 2018;12:6301–9.10.1021/acsnano.8b03408Search in Google Scholar PubMed

[127] Li H, Duan X, Wu X, et al. Growth of alloy MoS2xSe2(1−x) nanosheets with fully tunable chemical compositions and optical properties. J Am Chem Soc 2014;136:3756–9.10.1021/ja500069bSearch in Google Scholar PubMed

[128] Klee V, Preciado E, Barroso D, et al. Superlinear composition-dependent photocurrent in CVD-grown monolayer MoS2(1-x)Se2x alloy devices. Nano Lett 2015;15:2612–9.10.1021/acs.nanolett.5b00190Search in Google Scholar PubMed

[129] Feng Q, Zhu Y, Hong J, et al. Growth of large-area 2D MoS2(1-x) Se2x semiconductor alloys. Adv Mater 2014;26:2648–53.10.1002/adma.201306095Search in Google Scholar PubMed

[130] Feng Q, Mao N, Wu J, et al. Growth of MoS2(1−x)Se2x (x=0.41−1.00) monolayer alloys with controlled morphology by physical vapor deposition. ACS Nano 2015;9:7450–5.10.1021/acsnano.5b02506Search in Google Scholar PubMed

[131] Li H, Zhang Q, Duan X, et al. Lateral growth of composition graded atomic layer MoS2(1−x)Se2x nanosheets. J Am Chem Soc 2015;137:5284–7.10.1021/jacs.5b01594Search in Google Scholar PubMed

[132] Wu D, Shi J, Zheng XM, et al. CVD grown MoS2 nanoribbons on MoS2 covered sapphire(0001) without catalysts. Phys Status Solidi-R 2019;13:1900063.10.1002/pssr.201900063Search in Google Scholar

[133] Liu XC, Yuan YH, Qu DS, Sun J. Ambipolar MoS2 field-effect transistor by spatially controlled chemical doping. Phys Status Solidi-R 2019;13:1900208.10.1002/pssr.201900208Search in Google Scholar

[134] Wang YC, Ou JZ, Balendhran S, et al. Electrochemical control of photoluminescence in two-dimensional MoS2 nanoflakes. ACS Nano 2013;7:10083–93.10.1021/nn4041987Search in Google Scholar PubMed

[135] Wang YC, Ou JZ, Chrimes AF, et al. Plasmon resonances of highly doped two-dimensional MoS2. Nano Lett 2015;15:883–90.10.1021/nl503563gSearch in Google Scholar PubMed

[136] Zhang J, Yang A, Wu X, et al. Reversible and selective ion intercalation through the top surface of few-layer MoS2. Nat Commun 2018;9:5289.10.1038/s41467-018-07710-zSearch in Google Scholar PubMed PubMed Central

[137] Ganal P, Olberding W, Butz T, Ouvrard G. Soft chemistry induced host metal coordination change from octahedral to trigonal prismatic in 1T-TaS2. Solid State Ionics 1993;59:313–9.10.1016/0167-2738(93)90067-DSearch in Google Scholar

[138] Lorenz T, Teich D, Joswig JO, Seifert G. Theoretical study of the mechanical behavior of individual TiS2 and MoS2 nanotubes. J Phys Chem C 2012;116:11714–21.10.1021/jp300709wSearch in Google Scholar

[139] Yao J, Koski KJ, Luo W, et al. Optical transmission enhancement through chemically tuned two-dimensional bismuth chalcogenide nanoplates. Nat Commun 2014;5:5670.10.1038/ncomms6670Search in Google Scholar PubMed

[140] Cha JJ, Koski KJ, Huang KC, et al. Two-dimensional chalcogenide nanoplates as tunable metamaterials via chemical intercalation. Nano Lett 2013;13:5913–8.10.1021/nl402937gSearch in Google Scholar PubMed

[141] Qi YH, Shu CC, Dong DY, Petersen IR, Jacobs K, Gong SQ. Fast quantum state transfer in hybrid quantum dot–metal nanoparticle systems by shaping ultrafast laser pulses. J Phys D Appl Phys 2019;52:425101.10.1088/1361-6463/ab33ebSearch in Google Scholar

[142] Lu P, Wu XJ, Guo WL, Zeng XC. Strain-dependent electronic and magnetic properties of MoS2 monolayer, bilayer, nanoribbons and nanotubes. Phys Chem Chem Phys 2012;14:13035–40.10.1039/c2cp42181jSearch in Google Scholar PubMed

[143] Rice C, Young RJ, Zan R, et al. Raman-scattering measurements and first-principles calculations of strain-induced phonon shifts in monolayer MoS2. Phys Rev B 2013;87:081307.10.1103/PhysRevB.87.081307Search in Google Scholar

[144] Li AL, Pan JL, Yang ZX, Zhou L, Xiong X, Ouyang FP. Charge and strain induced magnetism in monolayer MoS2 with S vacancy. J Magn Magn Mater 2018;451:520–5.10.1016/j.jmmm.2017.11.074Search in Google Scholar

[145] Feng J, Qian X, Huang C-W, Li J. Strain-engineered artificial atom as a broad-spectrum solar energy funnel. Nat Photon 2012;6:866–72.10.1038/nphoton.2012.285Search in Google Scholar

[146] Yu DQ, Li SQ, Qi WH, Wang MP. Temperature-dependent Raman spectra and thermal conductivity of multi-walled MoS2 nanotubes. Appl Phys Lett 2017;111:123102.10.1063/1.5003111Search in Google Scholar

[147] Shi J, Wu D, Zheng XM, et al. From MoO2@MoS2 core-shell nanorods to MoS2 nanobelts. Phys Status Solidi B-Basic Solid State Physics 2018;255:1800254.10.1002/pssb.201800254Search in Google Scholar

[148] Wu NA, Yang ZX, Zhou WZ, et al. Collective electronic behaviors of laterally heterostructured armchair MoS2-NbS2 nanoribbons. J. Appl Phys 2015;118:084306.10.1063/1.4929759Search in Google Scholar

[149] Wang H, Liu FC, Fu W, Fang ZY, Zhou W, Liu Z. Two-dimensional heterostructures: fabrication, characterization, and application. Nanoscale 2014;6:12250–72.10.1039/C4NR03435JSearch in Google Scholar PubMed

[150] Li QZ, Tang LP, Zhang CX, et al. Seeking the Dirac cones in the MoS2/WSe2 van der Waals heterostructure. Appl Phys Lett 2017;111:171602.10.1063/1.4998305Search in Google Scholar

[151] Liao W, Huang Y, Wang H, Zhang H. Van der Waals heterostructures for optoelectronics: progress and prospects. Appl Mater Today 2019;16:435–55.10.1016/j.apmt.2019.07.004Search in Google Scholar

[152] Cao R, Wang H-D, Guo Z-N, et al. Black phosphorous photodetectors: black phosphorous/indium selenide photoconductive detector for visible and near-infrared light with high sensitivity (Advanced Optical Materials 12/2019). Adv Opt Mater 2019;7:1970047.10.1002/adom.201970047Search in Google Scholar

[153] Terrones H, Lopez-Urias F, Terrones M. Novel hetero-layered materials with tunable direct band gaps by sandwiching different metal disulfides and diselenides. Sci Rep 2013;3:1549.10.1038/srep01549Search in Google Scholar PubMed PubMed Central

[154] Chen HL, Wen XW, Zhang J, et al. Ultrafast formation of interlayer hot excitons in atomically thin MoS2/WS2 heterostructures. Nat Commun 2016;7:12512.10.1038/ncomms12512Search in Google Scholar PubMed PubMed Central

[155] Lopez-Sanchez O, Alarcon Llado E, Koman V, Morral AFI, Radenovic A, Kis A. Light generation and harvesting in a van der Waals heterostructure. ACS Nano 2014;8:3042–8.10.1021/nn500480uSearch in Google Scholar PubMed PubMed Central

[156] Hao S, Yang BC, Yuan JY, Gao YL. Substrate induced anomalous electrostatic and photoluminescence properties of monolayer MoS2 edges. Solid State Commun 2017;249:1–6.10.1016/j.ssc.2016.10.007Search in Google Scholar

[157] Wu D, Guo J, Du J, et al. Highly polarization-sensitive, broadband, self-powered photodetector based on graphene/PdSe2/germanium heterojunction. ACS Nano 2019;13:9907–17.10.1021/acsnano.9b03994Search in Google Scholar PubMed

[158] Wang W, Klots A, Prasai D, Yang Y, Bolotin KI, Valentine J. Hot electron-based near-infrared photodetection using bilayer MoS2. Nano Lett 2015;15:7440–44.10.1021/acs.nanolett.5b02866Search in Google Scholar PubMed

[159] Mao J, Yu Y, Wang L, et al. Ultrafast, broadband photodetector based on MoSe2/silicon heterojunction with vertically standing layered structure using graphene as transparent electrode. Adv Sci 2016;3:1600018.10.1002/advs.201600018Search in Google Scholar PubMed PubMed Central

[160] Zhou Y, Wu D, Zhu Y, et al. Out-of-plane piezoelectricity and ferroelectricity in layered α-In2Se3 nanoflakes. Nano Lett 2017;17:5508–13.10.1021/acs.nanolett.7b02198Search in Google Scholar PubMed

[161] Song Y, Shi X, Wu C, Tang D, Zhang H. Recent progress of study on optical solitons in fiber lasers. Appl Phys Rev 2019;6:021313.10.1063/1.5091811Search in Google Scholar

[162] Guo S, Zhang Y, Ge Y, Zhang S, Zeng H, Zhang H. 2D V-V binary materials: status and challenges. Adv Mater 2019;31:1902352.10.1002/adma.201902352Search in Google Scholar PubMed

[163] Tao W, Kong N, Ji X, et al. Emerging two-dimensional monoelemental materials (Xenes) for biomedical applications. Chem Soc Rev 2019;48:2891–912.10.1039/C8CS00823JSearch in Google Scholar

[164] Xing C, Huang W, Xie Z, et al. Ultrasmall bismuth quantum dots: facile liquid-phase exfoliation, characterization, and application in high-performance UV-vis photodetector. ACS Photonics 2018;5:621–9.10.1021/acsphotonics.7b01211Search in Google Scholar

[165] Xie Z, Xing C, Huang W, et al. Ultrathin 2D nonlayered tellurium nanosheets: facile liquid-phase exfoliation, characterization, and photoresponse with high performance and enhanced stability. Adv Funct Mater 2018;28:1705833.10.1002/adfm.201705833Search in Google Scholar

[166] Wang B, Zhong SP, Zhang ZB, Zheng ZQ, Zhang YP, Zhang H. Broadband photodetectors based on 2D group IVA metal chalcogenides semiconductors. Appl Mater Today 2019;15:115–38.10.1016/j.apmt.2018.12.010Search in Google Scholar

[167] Li SQ, Qi WH, Xiong SY, Yu DQ. Thermal conductivity of single-wall MoS2 nanotubes. Appl Phys A-Mater 2018;124:218.10.1007/s00339-018-1640-3Search in Google Scholar

[168] Zhang P, Qin FR, Zou L, et al. Few-layered MoS2/C with expanding d-spacing as a high-performance anode for sodium-ion batteries. Nanoscale 2017;9:12189–95.10.1039/C7NR03690FSearch in Google Scholar

[169] Xie JR, Zhu KJ, Min J, et al. In-situ grown ultrathin MoS2 nanosheets on MoO2 hollow nanospheres to synthesize hierarchical nanostructures and its application in lithium-ion batteries. Ionics 2019;25:1487–94.10.1007/s11581-019-02863-3Search in Google Scholar

[170] Zhou Y, Silva JL, Woods JM, et al. Revealing the contribution of individual factors to hydrogen evolution reaction catalytic activity. Adv Mater 2018;30:1706076.10.1002/adma.201706076Search in Google Scholar PubMed

[171] Jing Y, Huang S, Wu J, et al. A Single-electron transistor made of a 3D topological insulator nanoplate. Adv Mater 2019;31:1903686.10.1002/adma.201903686Search in Google Scholar PubMed

[172] Zhou Y, Pondick JV, Silva JL, et al. Unveiling the interfacial effects for enhanced hydrogen evolution reaction on MoS2/WTe2 hybrid structures. Small 2019;15:1900078.10.1002/smll.201900078Search in Google Scholar PubMed

[173] Zhu X, Li A, Wu D, et al. Tunable large-area phase reversion in chemical vapor deposited few-layer MoTe2 films. J Mater Chem C 2019;7:10598–604.10.1039/C9TC03216ASearch in Google Scholar

[174] Zhou Y, Jang H, Woods JM, et al. Direct synthesis of large-scale WTe2 thin films with low thermal conductivity. Adv Funct Mater 2017;27:1605928.10.1002/adfm.201605928Search in Google Scholar

[175] Yuan Z, Bak S-M, Li P, et al. Activating layered double hydroxide with multivacancies by memory effect for energy-efficient hydrogen production at neutral pH. ACS Energy Lett 2019;4:1412–8.10.1021/acsenergylett.9b00867Search in Google Scholar

[176] Zheng W, Xie T, Zhou Y, et al. Patterning two-dimensional chalcogenide crystals of Bi2Se3 and In2Se3 and efficient photodetectors. Nat Commun 2015;6:6972.10.1038/ncomms7972Search in Google Scholar PubMed PubMed Central

[177] Ding S-S, Huang W-Q, Yang Y-C, et al. Dual role of monolayer MoS2 in enhanced photocatalytic performance of hybrid MoS2/SnO2 nanocomposite. J. Appl Phys 2016;119:205704.10.1063/1.4952377Search in Google Scholar

[178] Li A, Pan J, Yang Z, Zhou L, Xiong X, Ouyang F. Charge and strain induced magnetism in monolayer MoS2 with S vacancy. J Magn Magn Mater 2018;451:520–5.10.1016/j.jmmm.2017.11.074Search in Google Scholar

[179] Xie D, Hu W, Jiang J. Bidirectionally-trigged 2D MoS2 synapse through coplanar-gate electric-double-layer polymer coupling for neuromorphic complementary spatiotemporal learning. Org Electron 2018;63:120–8.10.1016/j.orgel.2018.09.007Search in Google Scholar

Received: 2019-12-31
Revised: 2020-02-20
Accepted: 2020-02-20
Published Online: 2020-04-02

© 2020 Jian Sun, Yu Zhou et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Editorial
  2. 2D Xenes: from fundamentals to applications
  3. Reviews
  4. Monolayer MoS2 for nanoscale photonics
  5. 2D photonic memristor beyond graphene: progress and prospects
  6. MXenes: focus on optical and electronic properties and corresponding applications
  7. Advances in photonics of recently developed Xenes
  8. Nonlinear optical properties of anisotropic two-dimensional layered materials for ultrafast photonics
  9. Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications
  10. Recent advances in graphene and black phosphorus nonlinear plasmonics
  11. Fabrication, optical properties, and applications of twisted two-dimensional materials
  12. Novel layered 2D materials for ultrafast photonics
  13. 2D organic-inorganic hybrid perovskite materials for nonlinear optics
  14. Fine structures of valley-polarized excitonic states in monolayer transitional metal dichalcogenides
  15. MXenes for future nanophotonic device applications
  16. Two-dimensional nanomaterials for Förster resonance energy transfer–based sensing applications
  17. 2D materials integrated with metallic nanostructures: fundamentals and optoelectronic applications
  18. Graphene plasmonic devices for terahertz optoelectronics
  19. Research Articles
  20. Real-time dynamics of soliton collision in a bound-state soliton fiber laser
  21. Ultra-strong anisotropic photo-responsivity of bilayer tellurene: a quantum transport and time-domain first principle study
  22. Topological insulator overlayer to enhance the sensitivity and detection limit of surface plasmon resonance sensor
  23. Magnons scattering induced photonic chaos in the optomagnonic resonators
  24. Quantum confinement-induced enhanced nonlinearity and carrier lifetime modulation in two-dimensional tin sulfide
  25. Phosphorene-assisted silicon photonic modulator with fast response time
  26. High-performance monolayer MoS2 photodetector enabled by oxide stress liner using scalable chemical vapor growth method
  27. Enhancing the generating and collecting efficiency of single particle upconverting luminescence at low power excitation
  28. Biexcitons in 2D (iso-BA)2PbI4 perovskite crystals
  29. Broadband nonlinear optical response in GeSe nanoplates and its applications in all-optical diode
  30. Plasmonic nanocavity enhanced vibration of graphene by a radially polarized optical field
  31. Facile synthesis of sulfur@titanium carbide Mxene as high performance cathode for lithium-sulfur batteries
  32. The pump fluence and wavelength-dependent ultrafast carrier dynamics and optical nonlinear absorption in black phosphorus nanosheets
  33. Indium selenide film: a promising saturable absorber in 3- to 4-μm band for mid-infrared pulsed laser
  34. Temperature-stable black phosphorus field-effect transistors through effective phonon scattering suppression on atomic layer deposited aluminum nitride
  35. Real-time and noninvasive tracking of injectable hydrogel degradation using functionalized AIE nanoparticles
  36. MXene-Ti3C2 assisted one-step synthesis of carbon-supported TiO2/Bi4NbO8Cl heterostructures for enhanced photocatalytic water decontamination
  37. Nanofocusing of acoustic graphene plasmon polaritons for enhancing mid-infrared molecular fingerprints
  38. Effects of gap thickness and emitter location on the photoluminescence enhancement of monolayer MoS2 in a plasmonic nanoparticle-film coupled system
Downloaded on 12.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2019-0574/html?lang=en&srsltid=AfmBOorD5rfY2bw4thC8j4UzwOKaaNiaDHMYXP4EDWjy5RG_fSIvY--s
Scroll to top button