Next Article in Journal
Phytochemicals, Antioxidant and Antimicrobial Potentials and LC-MS Analysis of Centaurea parviflora Desf. Extracts
Previous Article in Journal
Yeast Particles for Encapsulation of Terpenes and Essential Oils
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Two Dimensional Heterostructures for Optoelectronics: Current Status and Future Perspective

by
Zaheer Ud Din Babar
1,2,†,
Ali Raza
2,†,
Antonio Cassinese
2,3 and
Vincenzo Iannotti
2,3,*
1
Scuola Superiore Meridionale (SSM), University of Naples Federico II, Largo S. Marcellino 10, 80138 Naples, Italy
2
Department of Physics “Ettore Pancini”, University of Naples Federico II, Piazzale Tecchio 80, 80125 Naples, Italy
3
CNR–SPIN (Institute for Superconductors, Oxides and Other Innovative Materials and Devices), Piazzale V. Tecchio 80, 80125 Naples, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(5), 2275; https://doi.org/10.3390/molecules28052275
Submission received: 31 December 2022 / Revised: 5 February 2023 / Accepted: 16 February 2023 / Published: 28 February 2023

Abstract

:
Researchers have found various families of two-dimensional (2D) materials and associated heterostructures through detailed theoretical work and experimental efforts. Such primitive studies provide a framework to investigate novel physical/chemical characteristics and technological aspects from micro to nano and pico scale. Two-dimensional van der Waals (vdW) materials and their heterostructures can be obtained to enable high-frequency broadband through a sophisticated combination of stacking order, orientation, and interlayer interactions. These heterostructures have been the focus of much recent research due to their potential applications in optoelectronics. Growing the layers of one kind of 2D material over the other, controlling absorption spectra via external bias, and external doping proposes an additional degree of freedom to modulate the properties of such materials. This mini review focuses on current state-of-the-art material design, manufacturing techniques, and strategies to design novel heterostructures. In addition to a discussion of fabrication techniques, it includes a comprehensive analysis of the electrical and optical properties of vdW heterostructures (vdWHs), particularly emphasizing the energy-band alignment. In the following sections, we discuss specific optoelectronic devices, such as light-emitting diodes (LEDs), photovoltaics, acoustic cavities, and biomedical photodetectors. Furthermore, this also includes a discussion of four different 2D-based photodetector configurations according to their stacking order. Moreover, we discuss the challenges that remain to be addressed in order to realize the full potential of these materials for optoelectronics applications. Finally, as future perspectives, we present some key directions and express our subjective assessment of upcoming trends in the field.

1. Introduction

Since its mechanical exfoliation in 2004, graphene has drawn much consideration and earned a colossal reputation as the “wonder material”. Various features, such as ultra-high-carrier mobility [1], a large active surface [2], superior in-plane thermal conductivity, and comparatively better mechanical properties (a Young’s modulus of 1 TPa and intrinsic strength of 130 GPa) [3], have made graphene a viable material for numerous applications. For instance, this includes applications such as ultrafast high-frequency photodetectors and transparent electrodes [4,5]. However, the reduced energy bandgap in graphene limits research as graphene-based transistors often have a low on/off current ratio [6]. In the meantime, several other 2D materials, e.g., WTe2 [7,8], Pb1−xSnxTe, Bi2Te3 [9,10], black phosphorous (BP), MoS2 [11], WS2 [12], WSe2 [13], and boron nitride (BN) [14], possess varying bandgaps and have been studied extensively. In addition to graphene’s zero-band gap, 2D materials have different bandgaps that can be further tuned by varying the thickness of layers. This has drawn significant interest in the scientific community and refocused the attention toward the fabrication of novel structures. The thickness-dependent bandgaps of such 2D materials lead to their applications in several devices, namely, photodetectors [15,16,17], field-effect transistors [18,19], and flexible electronics [20,21,22], to name a few. Furthermore, such 2D material facilitates atomic-scale integration, enabling advanced heterostructure devices with new physics and better technological aspects [23,24]. Generally, 2D material-based vdWHs can be produced via chemical-vapor deposition (CVD) or mechanical-transfer growth strategies [25,26]. The typical semiconductor-based heterostructures need comparable lattice structures among corresponding component semiconductors. In comparison, vdWHs have weak interaction between layers, so they have less stringent lattice-mismatching requirements [27,28].
It is imperative to develop and implement non-destructive methods to probe and explore the physicochemical and structural properties of 2D layered materials. Considering their small size, 2D nanostructured materials are fascinating in terms of their vibrational and molecular structure. Since the electrons have a short de Broglie wavelength, it allows for better spatial resolution. Similarly, electron microscopy limits the resolution of a light microscope to about 1 μm. In that respect, SEM and TEM are two popular microscopy techniques [29]. These approaches can offer detailed information on the crystalline nature and corresponding relationships between layer thickness, layer spacing, and element arrangement [30,31,32,33]. For example, selected-area electron diffraction (SAED) can discriminate a monolayer or a multilayer of graphene depending on the diffraction-intensity ratio from 110 to 100 [34]. Furthermore, the unique diffraction pattern can be identified when the interlayer spacing changes within a monolayer or multilayer of a 2D material.
On the other hand, Raman spectroscopy involves the inelastic scattering of monochromatic light, which is mainly categorized into Stokes and anti-Stokes scattering (emitted energy is larger than incoming energy). It involves the scattering of incoming light at different wavelengths than the initial wavelength when the sample is irradiated [35]. By comparing the incoming and outgoing light, vibrational modes can be determined, which are then used to create a fingerprint of each material. Photoluminescence (PL) measurements are the most commonly employed tool to characterize the electronic structure [29]. The radiative recombination of the electron–hole pair may be preceded by a non-radiative relaxation mechanism, leading to phonon emission. Comparing the non-radiative and radiative processes is vital to determine the characteristics of most 2D nanostructured materials. Multiple microscopy methods are applied to examine the optical characteristics of 2D materials, which are greatly influenced by their physical structure. It is commonly used to study morphology and topography using AFM, SEM, ultrasonic force microscopy (UFM), and TEM. In many cases, AFM can provide a layer-thickness measurement with preciseness as low as 5% [31,36]. In contrast, anomalies occur when the tip interacts differently with the substrate and the stacked layers [37]. However, the distance between the tip and the surface from any hysteretic parts is crucial to controlling these discrepancies and obtaining the ideal height profile of a single layer [37]. To determine the thickness of a single layer precisely; it is preferable to compare the heights of the first two layers instead of recording the difference in height between the substrate and the first layer.
When it comes to evaluating mechanical properties, such as mechanical interaction with substrates, ultrasonic force microscopy (UFM) is an effective method of assessing stiffness at the nanoscale [38,39]. Precisely, the sample is vibrated using a low vibrational amplitude (0.5–2 nm) and high vibrational frequencies (0.10–10 MHz) that are higher than the cantilever resonance frequencies of the AFM. It is possible to identify subsurface structures, including cavities, subsurface interfaces, and sample-substrate interfaces, by analyzing the material’s stiffness. It is possible to monitor ultrasonic vibration at the tip–sample junction on a separate channel and quantify it simultaneously with AFM using nonlinear rectification [29].
The present review focuses on the existing state-of-the-art concerning the material design, fabrication routes, and strategies to design novel heterostructures. Distinct fabrication approaches, e.g., mechanical transfer and chemical synthesis, are discussed extensively. In the meantime, critical factors and controlling parameters in each fabrication approach are also highlighted. Electric and optical features of vdWHs are thoroughly reviewed, with a particular emphasis on energy-band alignments. Later on, several optoelectronic applications of 2D vdWHs, such as photodetectors in biomedical fields, LEDs, photovoltaics, and acoustic cavities, are also reviewed. Notably, we explain four different configurations of 2D material-based photodetectors according to the stacking fashion of the layers. As future perspectives, we present some key directions and express our subjective assessment of upcoming trends.

2. Current Status

Two-dimensional stacked crystals, also called 2D heterostructures, have become increasingly popular over the past few decades. The diversity of their chemistry and physics has fascinated researchers and led to developing new perspectives. Similar efforts have facilitated the growth of 2D heterostructures with interesting thermal, optical, and electrical properties. Two-dimensional materials with weak vdW interactions among layers can be exfoliated into isolated atomic layers. Such isolated layers can further be rearranged into horizontally and vertically stacked heterostructures. In addition to conventional indirect growth of 2D crystals, current developments in vapor-phase deposition provide more opportunities for direct growth to realize such heterostructures (e.g., in-plane and vertically stacked). Therefore, 2D heterostructures produced by stacking semiconductors engender further opportunities in the semiconductor industry. This leads to applications such as highly efficient LEDs, photodetectors, neuromorphic devices, solar cells, lasers, and many others. Two-dimensional heterostructures are typically fabricated by combining two or more semiconductors, providing distinct electronic-band structures at the interface. Generally, a class of materials with identical elements, such as carbon materials (g-C3N4 and graphene), and/or transition-metal dichalcogenides (TMDs) such as MoS2 and WS2, etc., or similar attributes of individual elements, can be grouped into a particular classification. As guidance, we considered a clear taxonomy based on the latest advances that could prove helpful for the scientific society to foresee and delve into the following research fields (Figure 1).
Recently, the discovery of numerous 2D materials has radically changed the dynamics of semiconductor junctions. We expect this advancement to be an appropriate approach to representing several 2D materials discovered so far. The 2D-material family contains the majority of members, e.g., 2D carbon families [24,41]; magnetic materials, particularly their derivatives [42,43]; superconductors [44,45]; perovskites [46]; semiconducting dichalcogenides [47,48]; complex metallic dichalcogenides [49,50]; chrome-based dichalcogenides [51,52,53]; semimetals with polymorphs [54,55], in particular, polymorphism-based TMDs [56,57,58]; and mono-elemental materials, also known as “homonuclear materials” [59,60]. Moreover, it includes 2D ferroelectric and multiferroic materials [61,62,63], oxides [64,65], hydroxides [66,67], and other 2D materials (Figure 1) [68,69]. In future electronics and optoelectronics applications, developing these 2D materials will make it easier to build novel heterostructures with improved efficiency. Generally, different 2D materials are prone to certain limitations in particular devices. However, it has been observed that synergistic effects become visible when combined with other 2D materials or reassembled in stacked layers, resulting in a complete, robust material with superior performance. Aligned transfer, mechanical exfoliation, liquid-phase exfoliation, CVD, layer-by-layer (LbL), and electrostatic self-assembly are some of several approaches used to fabricate 2D heterostructures with unalike junctions/interfaces. In addition, sophisticated characterization techniques have allowed the designing of new 2D heterostructures with distinct heterointerfaces/junctions (Figure 2). Primarily, the fabrication of heterostructures from materials with different dimensions (2D/3D, 2D/2D, and 2D/1D) ensures substantial potential to improve the functions of futuristic optoelectronics. Still, scalable production of low-dimensional materials is critically important to reach the theoretical extent of performance. The requirement of advanced and controllable scalable transfer approaches to grow distinct integrated vdWHs with high quality impedes their true potential. Therefore, it is necessary to design a reliable procedure to transfer low-dimensional materials on a broader scale and ensure exceptional performance.

3. Electronic and Optical Features in vdW Heterostructures

3.1. Electronic Properties

The weak vdW interaction among two types of stacked 2D materials significantly changes the electrical properties of adjacent layers. It has been shown that ultra-flat and impurities-free (i.e., chemically pure) h-BN can substantially minimize the undesirable substrate effects on graphene [70,71]. Compared with conventional SiO2 substrates, graphene on the h-BN substrate has a field-effect mobility of about one-fold-larger in magnitude (∼1.4 × 105 cm2 V−1 s−1). In addition, the efficiency can be enhanced even further by encapsulating the graphene, thereby protecting it from surface contamination that could adversely affect its performance [72]. Such engineering approaches and interface variations are prevalent and have proven successful for 2D materials [71,73], especially for air-sensitive BP [74], NbSe2 [75], and CrI3 [76]. Cui et al. created a MoS2 transistor primarily based on vdWH in which graphene electrodes were employed as a source/drain while MoS2 layers were contained within h-BN, as visualized in Figure 3a,b. The proposed heterostructure attained a superior Hall mobility of ~34,000 cm2 V−1 s−1 at low temperatures [77].
Furthermore, h-BN is an optically thin dielectric material capable of withstanding significant electric-field strengths (≥0.5 V per layer) [23]. The main advantage of h-BN was to achieve a sophisticated interface. The graphene/h-BN heterostructures display unique moiré patterns due to mismatched lattices and misaligned orientations [78]. Lattice constants of h-BN and graphene exhibit a discrepancy of 1.8%, sufficient to observe the surface construction. The rotation-dependent moiré pattern can be observed in graphene engineered on h-BN according to theoretical analysis and experimental demonstration (Figure 3c–e). Such rotation-dependent moiré patterns function as periodic potentials and trigger the evolution of distinct Dirac points (Figure 3f,g) [79]. A commensurate–incommensurate transformation of graphene/h-BN heterostructures was discovered after the substantial alignment of crystallographic orientations of two crystals [80]. It is crucial to note that identical lattice constants separate moiré patterns in regions with relaxed graphene lattices. This effect is diminished whenever elastic energy becomes enormous to be recompensated via vdW in an incommensurate condition. Contrarily, graphene layers twisted at certain angles demonstrate unique physics [81]. Cao et al. reported that double-layer graphene twisted at an angle of 1.1° shows flat bands near zero Fermi energy, thus describing the evolution of insulating states [82]. As a result of electrostatic doping, magic-angle double-layer graphene also exhibits unusual superconductivity with a critical temperature as high as 1.7 K. Analogous to graphene and h-BN, precise twisting of one layer over the other, such as the stacked layers of monolayer TMDs (e.g., the layers of MoS2/WSe2) [83], and several other heterostructures (e.g., MoS2/WSe2 [83,84], WSe2/WS2 [85], and MoSe2/WS2 [86], express intriguing electrical, optoelectronic, and magnetic properties. It shows a considerable promise towards re-engineering the surface in TMD heterostructures.
Figure 3. (a) Schematic depiction of MoS2-based multi-terminal device encapsulated with h-BN. (b) False-color cross-sectional STEM images of the fabricated device. Adapted with permission from Ref. [77]. Copyright 2015, Nature Publishing Group. STM topography images with moiré patterns at (c) 2.4 nm, (d) 6.0 nm, and (e) 11.5 nm. The scale bar is 5 nm. (f) A theoretical low density of state behavior at three distinct rotation angles between graphene and h-BN, where green, blue, and red represent 6.3 nm, 10.0 nm, and 12.5 nm, respectively. (g) Experimental dI/dV behavior at distinct moiré wavelengths of 13.4 nm (red) and 9.0 nm (black) and the arrows indicate dips in the dI/dV curves. Adapted with permission from Ref. [79]. Copyright 2012, Nature Publishing Group.
Figure 3. (a) Schematic depiction of MoS2-based multi-terminal device encapsulated with h-BN. (b) False-color cross-sectional STEM images of the fabricated device. Adapted with permission from Ref. [77]. Copyright 2015, Nature Publishing Group. STM topography images with moiré patterns at (c) 2.4 nm, (d) 6.0 nm, and (e) 11.5 nm. The scale bar is 5 nm. (f) A theoretical low density of state behavior at three distinct rotation angles between graphene and h-BN, where green, blue, and red represent 6.3 nm, 10.0 nm, and 12.5 nm, respectively. (g) Experimental dI/dV behavior at distinct moiré wavelengths of 13.4 nm (red) and 9.0 nm (black) and the arrows indicate dips in the dI/dV curves. Adapted with permission from Ref. [79]. Copyright 2012, Nature Publishing Group.
Molecules 28 02275 g003

3.2. Optoelectronic Properties

In recent decades, extensive research efforts have been devoted to layer-dependent band structures and many other characteristics of several 2D materials [23,87]. Furthermore, vdW stacking can further revamp the optoelectronic features of 2D materials [88,89]. The research on vdW interactions among the layers and dynamic excitons in 2D materials is made possible by rapidly evolving routes such as PL, ultrafast optical, and Raman spectroscopy. Novel modes of lattice vibration could be developed via interfacial interactions at an interface in 2D semimetals and semiconductors-based vdWHs. Li et al. used ultralow-frequency Raman spectroscopy to identify numerous novel characteristics of layer-breathing modes in MoS2/graphene-based vdWH [90]. Contrary to high-frequency modes (E12g and A1g), these modes are governed via flake thickness and susceptible to interfacial coupling. Moreover, the interfacial force constant for layer-breathing was reported to be α0 (I) = 60 × 1018 Nm−3, as determined through the linear chain model. Fang et al. employed a micromachining transfer approach to create a vdW hetero-bilayer heterostructure based on monolayer WSe2 and MoS2 (Figure 4a) and studied the interlayer carrier coupling using PL and absorbance spectra [91]. The PL spectra shown in Figure 4b correspond to a Stokes-like shift of ~100 meV. This shifting follows the type-II band alignment, which holds spatially indirect emission and absorption (Figure 4c). Upon optical or electrical stimulation, abundant emissive and non-emissive dark excitons are present in 2D vdW optoelectronic structures. The ultra-fast interlayer energy transfer (approx. 200 femtoseconds) studied by Wu et al. in WSe2/MoTe2 vdWHs exhibited a near-unity PL output. This shows that, in addition to emissive excitons, non-emissive excitons such as spin-forbidden dark excitons and momentum-forbidden indirect excitons also exhibit similar behavior. Furthermore, in vdW layers, the transmission of both an electron and a hole denotes a dominating Dexter-type energy transfer (Figure 4d–f). The primary illustration for ultrafast charge transfer was described by Hong et al. in MoS2/WS2-based vdWHs. They discovered a rapid (in under 50 fs) hole transfer from MoS2 to WS2, which has excellent potential for various applications, including optoelectronics, light harvesting, and many more [92]. The limit of indirect excitons is generally restricted to low temperatures despite extensive study on indirect excitons in semiconductor vdWHs based on group III–V and II–VI elements, e.g., AlA/GaA-linked quantum wells [93]. Moreover, Figure 4h demonstrates the formation of a coupled quantum well in the vdW MoS2/h-BN heterostructures illustrated in Figure 4g. Indirect excitons were discovered by Calman et al. at room temperature [94]. Strikingly, the lifespan of indirect excitons in MoS2/h-BN vdWHs was longer than direct excitons in monolayer MoS2.

3.3. Energy-Band Alignments

The electronic orbital extends to each other in vdWHs despite the faint interactions among layers at the interface that affect the band structure in every layer [96,97]. A 2D stacked vdW layered structure can exhibit significant interaction between interlayers, producing unique physical characteristics. Because of Dirac electrons’ linear dispersion, graphene has a high carrier mobility of 104 cm2/Vs [98]. Still, zero bandgaps of graphene limit its applications in transistors [99,100]. In contrast, graphene-based vdWHs have pioneered several growing research topics, which may compensate for the graphene’s zero bandgaps. For example, controlled growth of graphene layers over h-BN can induce arbitrary rotational orientation between the lattices of individual components of the device. Due to the higher lattice constant of h-BN and this rotation of lattices, topographic moiré patterns are created as shown in Figure 5a–c [79]. In addition, as displayed in Figure 5d, the moiré pattern functions as a faint periodic potential and thereby producing a new set of Dirac points [79]. This result implies the possibility of controlling graphene’s electronic structures by stacking it over h-BN layers. In the case of a new Dirac point, the lattice mismatch among h-BN and neighboring graphene layers can tune the density of states in the graphene layer.
Similarly, it suggests that stacking various TMDs can change the electronic structure. Several research groups have theoretically predicted and demonstrated experimentally that vdWHs based on monolayer MoA2/WA2 (A = S, Se, and/or Te) exhibit type-II band alignments (Figure 5e) [101,102]. Optically active conduction-band (CB) minimum and valence-band (VB) maximum sites are limited to opposing layers, and the lowest-energy excitons can split spatially (Figure 5f–g) [102]. This is promising in some applications, such as optoelectronics and solar-energy conversion. The type-II band alignment in 2D MoS2/WSe2-based heterostructure was illustrated by Li et al. with 0.76 eV as the conduction offset and a 0.83 eV valence-band offset [103]. In contrast, according to Cho et al., monolayer n-type MX2 (where X = Se/Te; M = Mo/W)-based and p-type MX2 (where X = S/Se; M = Zr/Hf)-based vdWHs have synergistic coupling that offer broken gap junctions with promising tunneling efficiencies [104]. Such features are of great significance in low-power logic devices. Although 2D MX2 (M = Mo/W; X = S/Se) monolayer TMDs are direct-bandgap semiconductors, they change to indirect-bandgap semiconductors as the layers increase. It is due to the non-negligible transformation of the Γ-point to an intermediate state (Γ-Q) [105]. Furthermore, electric and electronic characteristics can be modified for stacked bilayer homostructures by varying the interlayer spacing or twisting the layers. Wang et al. developed MoS2 bilayers and found that the size of indirect bandgap changed noticeably according to twist angles, as shown in Figure 5h [105]. Furthermore, it shows that the most significant red shift was observed for AB and AA stacking of layers. In contrast, the red shift was found to be drastically smaller or negligible in the case of other angles (Figure 5i). Here, “AB” suggests the configuration when the S atoms of the top layer are on the S atoms of the bottom layers, whereas “AA” is when the S atoms are on top of the Mo atoms of the bottom layer. It is worth-mentioning that exciton dynamics, Raman spectroscopy, absorption spectra, and PL are optimal approaches for probing optical characteristics of vdWHs because of their efficiency, accuracy, and nondestructive observations [106,107].
Figure 5. STM topographical photographs at (a) 2.4 nm, (b) 6.0 nm, and (c) 11.5 nm, (d) The arrows in the experimental dI/dV curves stipulate the dips at two different moiré wavelengths, 9.0 nm (black) and 13.4 nm (red). The black curve shows a dip energy of 0.28 eV, whereas the red curve shows a dip energy of 0.22 eV from the Dirac point respectively. Further, the valence band exhibits much deeper dips than the conduction band, which indicates different relative strengths. Adapted with permission from Ref. [79]. Copyright 2012, Nature Publishing Group. (e) Visualization of band alignment for monolayer semiconducting TMDs as well as monolayer SnS2. Adapted with permission from Ref. [104]. Copyright 2013, American Institute of Physics. (f) VB maximum and (g) CB minimum charge densities in monolayer WX2/MoX2 (X  =  S, Se, and Te) lateral heterostructures with similar X sites. Adapted with permission from Ref. [102]. Copyright 2013, American Institute of Physics. (h) Optical photographs of mono- and bilayer MoS2 at certain twisting angles. (i) Display of energy values evaluated from Kohn–Sham K-valley indirect bandgaps (green) and direct bandgaps (orange). Adapted with permission from Ref. [105]. Copyright 2014, Nature Publishing Group.
Figure 5. STM topographical photographs at (a) 2.4 nm, (b) 6.0 nm, and (c) 11.5 nm, (d) The arrows in the experimental dI/dV curves stipulate the dips at two different moiré wavelengths, 9.0 nm (black) and 13.4 nm (red). The black curve shows a dip energy of 0.28 eV, whereas the red curve shows a dip energy of 0.22 eV from the Dirac point respectively. Further, the valence band exhibits much deeper dips than the conduction band, which indicates different relative strengths. Adapted with permission from Ref. [79]. Copyright 2012, Nature Publishing Group. (e) Visualization of band alignment for monolayer semiconducting TMDs as well as monolayer SnS2. Adapted with permission from Ref. [104]. Copyright 2013, American Institute of Physics. (f) VB maximum and (g) CB minimum charge densities in monolayer WX2/MoX2 (X  =  S, Se, and Te) lateral heterostructures with similar X sites. Adapted with permission from Ref. [102]. Copyright 2013, American Institute of Physics. (h) Optical photographs of mono- and bilayer MoS2 at certain twisting angles. (i) Display of energy values evaluated from Kohn–Sham K-valley indirect bandgaps (green) and direct bandgaps (orange). Adapted with permission from Ref. [105]. Copyright 2014, Nature Publishing Group.
Molecules 28 02275 g005
The production of monolayer WSe2 and MoS2-based vdWH was explained by Javey et al., who demonstrated a distinct excitonic PL peak at 1.55 eV lower than that of WSe2 (1.64 eV) and MoS2 (1.87 eV) monolayers [91]. Moreover, they discovered that adding dielectric h-BN layers to the vdW gap can alter interlayer coupling. Shin et al. also examined the interlayer coupling between MX2-based vdWHs at distinct twist angles [108]. At 0°, the upper layer’s Se-atoms were above the Se-atoms of the underneath layer (Figure 6a). At 60°, the metal atoms of the underneath layer were just below the Se-atoms of the upper layer. Further, PL excitonic peaks were observed at one order of magnitude smaller and slightly red shifted compared to other materials. The red shift of vdWH may have been due to the variations in band structure, and the PL quenching in hetero-bilayer could have been driven by a reduction in PL quantum yield in bilayer systems. Based on band alignment, an observed peak at 1.35 eV in Figure 6b can be ascribed to interlayer excitons. In particular, d-orbitals of W and Mo dominated the energy levels of WSe2 and MoSe2. It can also be observed that the VB and CB of MoSe2 were at lower energies than those of WS2 because the energy of the 4d orbital of Mo is lower than that of the 5d orbital of W. This causes the type-II band to align between hetero-bilayer systems. When exposed to light, individual WSe2 and MoSe2 observe exciton formation. As the excitons are located between the CB and VB in each layer, they have lower energy levels than unbound electrons and holes. The photoinduced electrons and holes are subsequently separated and transferred to CB of MoSe2 and VB of WSe2. Hence, the holes in WSe2 VB and electrons in MoSe2 CB recombine to generate excitons, consequently leading to interlayer exciton emission. Furthermore, they discovered that PL was maximum at 0° and 60° twist angles during the interlayer exciton emission and diminished at other twist angles (Figure 6c). This shows that the hetero-bilayer system has a highly symmetrical stacking architecture with robust coupling among the layers at twist angles of 0° and 60°. This leads to excellent charge-transfer efficiency due to the short distance between layers, resulting in increased PL intensity. This finding opens ways for modulating the optical characteristics of vdWHs in intricate aspects. Wang et al. employed femtosecond pump-probe and PL mapping spectroscopy in a photoinduced monolayer MoS2/WS2 heterojunction and demonstrated ultrafast charge transport (Figure 6d) [92]. They identified a significant quenching action of PL spectra at heterojunction compared to the single-layered material (Figure 6e), indicating the high interlayer charge transfer. Following that, they examined the transient absorption spectra of the proposed device in the range of 2.0 to 2.5 eV (Figure 6f) and discovered a hole-transfer time (from MoS2 to WS2 layer) of 50 fs. Such rapid charge transfer in the proposed vdWHs can be used for efficient solar-energy conversion and high-performance photodetectors.

4. Fabrication Routes

4.1. Mechanical Transfer

For various 2D materials, wet- and dry-transfer procedures are commonly employed as the mechanical fabrication routes. For example, Dean et al. explained the mechanical wet-transfer method to develop mono- and bilayer graphene over h-BN substrates [70]. As indicated in Figure 7a–e, a polymer film comprising a water-soluble layer and a handling layer made of poly(methyl methacrylate) (PMMA) was initially used to translate the graphene. The water-soluble layers get dissolved when immersed in deionized water, which leaves graphene/PMMA layers on top. A micro-manipulator attached to an optical microscope was utilized to position the graphene/PMMA membrane against the target h-BN layer after being translated on a glass slide (Figure 7a(ii)). The substrate was rinsed with acetone in subsequent stages to remove the remaining PMMA and heat treated at 110 °C to remove the adsorbed water (Figure 7a(iii–iv)), thus improving the contact between the layers. Optical images of graphene/h-BN created by using this mechanical wet-transfer approach are shown in Figure 7b. Furthermore, simply repeating the previously mentioned steps enables the fabrication of complex heterostructures. However, the residues on the surface could impede the transport procedure. Therefore, anticipating the robust vdW interaction among 2D materials, improved pick-and-lift approaches have been established to facilitate the complicated multilayer stacking with impurity-free interfaces [109,110]. A polypropylene carbonate (PPC)/polydimethylsiloxane (PDMS) stamp was utilized to place a desired h-BN flake on a SiO2 substrate at 40 °C (Figure 7c(i)). Subsequently, h-BN was translated to a different substrate at 110 °C (Figure 7c(ii,iii)). Moreover, a subsequent stack can be selected and successively placed onto an underlying h-BN substrate, as shown in Figure 7c(iv–vii). Compared to earlier pick-and-drop strategies, the advantage of the vdW pick-and-drop approach is that there is no interaction between the interface and any polymer during transfer. h-BN/Graphene/h-BN vdWHs assembled through this kind of growth are presented in Figure 7d,e. In addition, AFM analysis, shown in Figure 7e, demonstrated that the graphene layer is completely enclosed among neighboring h-BN layers with no prominent blisters. Moreover, several studies have been performed on comparable transfer methods, even with minor alterations in stamp preparation [111]. For instance, as an alternative print, Xu et al. produced a graphene-sandwiched heterostructure (h-BN/graphene/h-BN) through polycarbonate (PC)/PDMS [111]. It is critical to note that some essential principles, such as the purity of stack, time required for transfer, and simplicity of operation, should be considered while evaluating alternative transfer approaches [112].

4.2. Chemical Synthesis

Even though challenging, mechanical transfer procedures show enormous potential in fabricating unique vdWHs using 2D material that incites astonishing physical phenomena and imparts distinct features. Still, such a procedure has scalability constraints. Some alternative strategies, such as the bottom-up approach, have been established to overcome limitations and ensure scalability. The atomically flat surface and the absence of dangling bonds make the graphene an excellent template for other 2D materials and have been used to build several vdWHs [113,114,115]. Compared with manual stacking, direct stacking of 2D vdWHs allows simultaneously automated (machine-driven) manufacturing of several layers. Direct-growth 2D vdWHs may be produced using several methods, including single-step in situ fabrication, physical epitaxy (epitaxial growth), and CVD.
Furthermore, graphene-based vdWHs are helpful in place of various electrical applications owing to fundamental aspects such as minor-resistivity connection originating from semimetal/semiconducting interfaces. For example, graphene film was initially developed using a Cu-foil (graphene/Cu) via the liquid precursor CVD technique. Subsequently, ammonia borane (NH3-BH3) was utilized to create h-BN layers on graphene/Cu film produced in the previous step. Furthermore, a two-step CVD route was applied by Liu et al. [116] for the growth of h-BN over graphene. In addition, graphene can also be adapted as an epitaxial substrate for TMDs. By using CVD at lower temperature conditions (400 °C), Shi et al., for instance, reported a modest method for producing MoS2/graphene vdWHs [115].
Moreover, Chen et al. fabricated continuous monolayer MoS2 films over graphene as a substrate that provides sufficient information about the quality of MoS2 films, which entirely depends on nucleation density and growth time and can be controlled by precisely controlling these parameters [117]. Optics-electronic integration on a single chip has much potential owing to the wafer-scale development of vdWHs. Furthermore, it should be noted that the characteristics of resultant vdWHs were influenced by the quality of the substrate onto which the graphene flake was transferred [118]. Because MoS2 and WS2 grow sequentially due to nucleation variations and growth rates, vertical heterostructures (WS2/MoS2) commonly grow at 850 °C compared to in-plane structures (Figure 8a–c). An optical photograph of an in-plane WS2/MoS2 heterostructure on an SiO2 substrate is presented in Figure 8d, whereas Figure 8e shows a Z-contrast image of the step edge of a WS2/MoS2 bilayer on SiO2 substrate. In addition, the direct vdW epitaxial growth of the WSe2/SnS2-based vertical bilayer p–n junction with lateral diameters close to the millimeter scale has also been reported [119]. Using WSe2 powder as a precursor, the underlying WSe2 monolayer was produced at 1100 °C, whereas above this temperature, an SnS2 layer was produced with SnO2 with S-powders as precursors and the Ar as the carrier gas (Figure 8f). Furthermore, the AFM images presented in Figure 8h–i substantiate that the bilayer vdWH, which is shown in Figure 8g, has one layer of WSe2 and one layer of SnS2. Various vdWHs, such as MoTe2/Wse2 [120], GaSe/MoSe2 [121], GaSe/MoS2 [122], Sb2Te3/MoS2 [123], and many others, have been fabricated using a direct CVD method. Aside from the above-mentioned CVD procedures, several other approaches, such as metal-organic (MO)-CVD [124] and molecular beam epitaxy (MBE) [125,126], can be employed to build vdW 2D heterostructures directly. MOCVD uses organometallic compounds because they vaporize at a higher pressure than metallic elements [127]. After vaporization, these compounds are transferred to the growth chamber via a carrier gas. Since MOCVD uses organometallic precursors that degrade below 400 °C, it is possible to perform the synthesis at low temperatures. It also indicates that organometallic compounds can be fine-tuned in a bubbler modifier that controls the flow rate of the carrier gas (typically H2 or N2). However, a thorough study of the purity and decomposition pathway of the precursors is essential to produce carbon-free films under the required growth conditions [128,129]. MOCVD can be performed in cold wall reactors and horizontal hot wall reactors, unlike other techniques, such as powder-based chemical-vapor deposition (P-CVD). Cold wall reactors minimize unwanted gas-phase reactions by limiting the region of thermal heating near the substrate. It is worth mentioning that the MOCVD technique is much more universally applied to 2D heterostructures, including group III oxides [130] and group III nitrides [34]. The importance of defects was described by Azizi et al. via MOCVD to fabricate freestanding WSe2/graphene vdWHs [124]. They employed CVD to create a thin graphene film on a Cu-substrate, which was then transferred to a gold TEM grid. To create a monolayer WSe2, an Au grid and graphene sheet were mounted inside an MOCVD device. The growth was carried out in a hydrogen atmosphere using dimethyl selenide as selenium and tungsten hexacarbonyl (W(CO)6) as a tungsten precursor.
A further potential method is an ultrahigh vacuum (UHV) –MBE, which is a common technique employed for II–VI and III–V semiconductors. TaSe2/MoSe2 and TaSe2/HfSe2 vdWHs were produced by Tsoutsou et al. by utilizing UHV–MBE, and the resulting structures displayed good structural stability [131]. The direct-growth approach is generally considered the most promising strategy to realize scalable production of various vdW 2D heterostructures. Furthermore, it provides absolute control over the thickness and stoichiometry of layers and prevents exposure to ambient contaminants that harm the device’s performance. Additionally, controlled development of vdW 2D heterostructures continues to be a significant challenge despite significant efforts and cutting-edge technologies over the past 10 years. Specific factors that rely on growth conditions, such as crystallinity, homogeneity, and thickness, should be considered. Nevertheless, each 2D material has its own requirements and has a significant impact on interface properties. Additionally, ideal epitaxial growth can only be achieved between materials with perfectly matched lattices.
Figure 8. Depiction of synthesis mechanism of vdWHs: (a) demonstration of WS2/MoS2 heterostructure procedure. (b,c) In-plane and vertical stacking of WS2 (650 °C) and MoS2 (850 °C) monolayers to grow (in-plane/vertical) heterostructures. (d) Display of optical photograph for an in-plane WS2/MoS2 heterostructure on an SiO2 substrate. (e) Z-contrast image of the step edge of WS2/MoS2 bilayer on SiO2 substrate. The green dashed boundary implies the edge of the step, while the two triangles show the orientation of the MoS2 (top) and WS2 (bottom) layers. Adapted with permission from Ref. [132]. Copyright 2014, Nature Publishing Group. (f) Pictorial demonstration of dual-step vapor epitaxy fabrication of WSe2/SnS2 vdWHs. (g) Optical photograph of a stacked triangular flake of 1L SnS2/1L WSe2. (h) AFM photograph of the section in (g) indicated by the blue rectangle. (i) AFM photograph of the section in (g) indicated by the red rectangle. Adapted with permission from Ref. [119]. Copyright 2017, Nature Publishing Group.
Figure 8. Depiction of synthesis mechanism of vdWHs: (a) demonstration of WS2/MoS2 heterostructure procedure. (b,c) In-plane and vertical stacking of WS2 (650 °C) and MoS2 (850 °C) monolayers to grow (in-plane/vertical) heterostructures. (d) Display of optical photograph for an in-plane WS2/MoS2 heterostructure on an SiO2 substrate. (e) Z-contrast image of the step edge of WS2/MoS2 bilayer on SiO2 substrate. The green dashed boundary implies the edge of the step, while the two triangles show the orientation of the MoS2 (top) and WS2 (bottom) layers. Adapted with permission from Ref. [132]. Copyright 2014, Nature Publishing Group. (f) Pictorial demonstration of dual-step vapor epitaxy fabrication of WSe2/SnS2 vdWHs. (g) Optical photograph of a stacked triangular flake of 1L SnS2/1L WSe2. (h) AFM photograph of the section in (g) indicated by the blue rectangle. (i) AFM photograph of the section in (g) indicated by the red rectangle. Adapted with permission from Ref. [119]. Copyright 2017, Nature Publishing Group.
Molecules 28 02275 g008

5. 2D Heterostructure-Based Optoelectronic Devices

5.1. Photodetectors

The development of highly effective photodetectors with a broad optical response is made possible by the rapid growth of 2D materials with specific optoelectronic and electrical properties [5,133]. As described earlier, the dangling bond-free nature of 2D materials allows for the simple fabrication of multidimensional heterostructures [134,135]. The fabrication of 2D vdWHs is a potential strategy for realizing flexible photodetector devices with diverse topologies [136,137]. Based on varied material stacking, four primary forms of 2D vdWHs for photodetectors are established. The first one is based on graphene/2D-semiconductor/graphene (G/2D-SC/G) heterostructures in which graphene serves as the contact layer and 2D semiconductors serve as the photoactive layer. The spectral detection range in this 2D vdWH is primarily dictated by various bandgap energies of 2D semiconductors. These patterns are deemed advantageous for minimizing contact resistances and accelerating the split-up of photogenic carrier pairs to yield highly effective photocurrent and increased photo-reaction speed.
Additionally, by tuning the gate voltage, the work function of graphene can readily be changed to change the photocurrent of the proposed device. A 2D vdWH composed of h-BN/G/WSe2/G/h-BN with an ultrashort photo-response-time constant has been demonstrated. Such heterostructures exhibit an exceptional photodetection performance, which acts as a function of bias voltage and can also be controlled by adjusting the thickness of WSe2 flakes [138]. In this case, the encapsulating layer is h-BN, the source and drain electrodes are monolayer graphene flakes, and the photoactive material is WSe2. Furthermore, graphene improves the light response by increasing the charge-transfer rate (Figure 9a). In the meantime, measurements of the temporal response of h-BN/G/WSe2/G/h-BN photodetector were made using time-resolved photocurrent measurements (Figure 9b). The photocurrent response-time constant of the G/WSe2/G photodetector was estimated to be 5.5 ps under sub-picosecond laser stimulation (Figure 9c). In principle, the development of efficient and quick photodetection response depends on intrinsic features, e.g., the atomically narrow charge-extraction channel of the device.
The structural design of the second device is founded on the photogate of vertical vdWHs, where different 2D materials with distinct properties are commonly coupled to provide higher carrier mobility (as the charge-transport layer) and outstanding light absorption (as the light-absorption layer). Graphene has higher carrier mobility and offers superior interface properties in the heterostructure so it can function as a charge-transport layer. On the other end, MoTe2 has better light-absorption characteristics and serves as a light-absorption layer. A MoTe2/graphene vdWHs-based photodetector was recently reported (Figure 9d) [138]. Upon illumination, the photoexcited carriers in the MoTe2 layer were transported easily and rapidly to the graphene layer because of the matching energy-band structure (Figure 9e). By unifying the properties of both materials, vdW heterojunction devices exhibited an exceptional performance under 1064 nm laser stimulation, thus providing a high photo responsivity, an ultrahigh photoconductive gain, and a detectivity (D*) of ~970.82 A/W, 4.69 × 108, and 1011 Jones, respectively (Figure 9e,f). This shows that the high mobility and high light absorption of two different 2D materials can successfully be combined in vdWHs to achieve high-performance photodetection.
The third kind of vdWHs, i.e., vdW p–n heterojunctions, has also been studied significantly and is considered the most fundamental and extensively used building block for industrial semiconductor devices. p–n vdWHs may effectively segregate photo-excited carriers and provide photocurrent regardless of the momentum mismatch among the layers. There are two significant benefits to fabricating an atomic vdW p–n junction in this situation: (i) the photoexcited electron–hole pairs created by individual material can be effectively divided by a built-in electric field close to the p–n junction interface, enabling rapid photodetection, and (ii) complementary p–n junction-based photodetectors typically explain the benefits of using two materials in p–n junctions that give rise to different behaviors, such as detection-band complementarity and polarized light-response interaction. Despite the interlayer momentum mismatch, p–n vdW heterojunctions efficiently split up photo-excited carriers and lead to a photocurrent [139]. Chen et al. use ultrafast visible/infrared micro-spectroscopy to investigate the interlayer charge-transport kinetics in MoS2/WS2 vdWHs [140]. The charge transfers between the MoS2 and WS2 layers produced an intermediate state of electron–hole pairs with excess energy before creating firmly bound interlayer excitons. The surplus energy in the intermediates makes it easier for them to dissociate, which helps generate the photocurrent. A recently reported BP/MoS2 heterostructure composed of mid-wave infrared (MWIR) polarized detectors with “Au” as the contact with the hole and back reflector and “MoS2” as the MWIR windows and contact source with the electron. The proposed device displayed a clear room-temperature photo response with excellent performance [141]. Because of BP’s thickness-dependent light-absorption properties, a thick flake of 150 nm was chosen as the light-absorption layer, with an 80% x-polarized incoming laser absorption capacity. At ambient temperature, a 150 nm BP/15 nm MoS2 vdWH-based photodetector demonstrated an EQE (external quantum efficiency) of 35% at 2.5–3.5 μm and a specific D* of up to 1010 Jones at 3.8 μm, with an R = 0.9 A/W·m. Furthermore, a photodetector based on MoS2 sandwiched between two BP layers (BPt/MoS2/BPb) was fabricated for polarization-resolved and bias-dependent photodetection (Figure 9g,h). The photo response of MoS2/BPt and MoS2/BPb as a function of the polarizer angle under a linearly polarized laser at λ = 3.5 μm is presented in Figure 9i. It illustrates that the proposed device can simultaneously detect two linear polarization components under unpolarized illumination. Following that, BPt crystal orientation of the top BP layer (BPt) is perpendicular (out-of-the-axis) to the BP layer at the bottom (BPb), and the adjacent MoOx/Pd layer functions as a hole collector from the BPt layer.
Figure 9. (a) Atomistic representation of photoexcited carrier dynamics in an h-BN/G/WSe2/G/h-BN heterostructure. Excitons are produced, separated, and delivered to graphene electrodes under pulsed-laser excitation. (b) Demonstration of time-resolved photocurrent-analysis method. (c) Photocurrent time-delay behavior demonstrates a light response-time constant of 5.5 ps. Adapted with permission from Ref. [138]. Copyright 2016, Nature Publishing Group. (d) A schematic representation of photodetector based on MoTe2/graphene. (e) Charge-transfer behavior of the device (with and without illumination). (f) The behavior of photocurrent and photoresponsivity as a function of incident light power at 980 nm. Adapted with permission from Ref. [142]. Copyright 2017 John Wiley & Sons, Ltd. (g) Schematic of BPt/MoS2/BPb vdW heterojunction-based devices. (h) Cross-sectional TEM images of polarization-resolved BP/MoS2 heterojunction photodetector. (i) Photo response of MoS2/BPt and MoS2/BPb under a linearly polarized laser at λ = 3.5 μm. Adapted with permission from Ref. [141]. Copyright 2018, Nature Publishing Group.
Figure 9. (a) Atomistic representation of photoexcited carrier dynamics in an h-BN/G/WSe2/G/h-BN heterostructure. Excitons are produced, separated, and delivered to graphene electrodes under pulsed-laser excitation. (b) Demonstration of time-resolved photocurrent-analysis method. (c) Photocurrent time-delay behavior demonstrates a light response-time constant of 5.5 ps. Adapted with permission from Ref. [138]. Copyright 2016, Nature Publishing Group. (d) A schematic representation of photodetector based on MoTe2/graphene. (e) Charge-transfer behavior of the device (with and without illumination). (f) The behavior of photocurrent and photoresponsivity as a function of incident light power at 980 nm. Adapted with permission from Ref. [142]. Copyright 2017 John Wiley & Sons, Ltd. (g) Schematic of BPt/MoS2/BPb vdW heterojunction-based devices. (h) Cross-sectional TEM images of polarization-resolved BP/MoS2 heterojunction photodetector. (i) Photo response of MoS2/BPt and MoS2/BPb under a linearly polarized laser at λ = 3.5 μm. Adapted with permission from Ref. [141]. Copyright 2018, Nature Publishing Group.
Molecules 28 02275 g009

Photodetectors in Biomedical Field

Past years have seen extensive use of TMDs in numerous biomedical applications such as UV-radiation monitors, retinal prostheses, and pulse oximeters [143,144]. In addition to significantly reducing medical costs, wearable photodetectors are essential for health monitoring. In this regard, advanced strategies are required to fabricate devices with greater flexibility, ease of integration, and environmentally stable materials. Two-dimensional materials are an obvious choice for wearable electronics due to their inherent advantages of size-dependent properties, flexibility, and intriguing optoelectronic characteristics. For example, tattoo-like graphene devices record skin hydration, body temperature, and heart rate (HR). Recently, a wearable device was created by Polat et al. by placing light-sensitive lead-sulfide (PbS) quantum dots (QDs) on graphene [145]. Here, graphene functions as a susceptible substrate in connection with QDs, which means that their incorporation in the biomedical sector has excellent potential for sensing over a broad spectrum of wavelengths [146]. Electron–hole pairs in PbS QDs are produced upon illumination, where the holes are transmitted to graphene, and electrons remain confined inside the QDs. Holes transmitted in graphene improve conductivity, thus producing a controlled electrical response. The high sensitivity can be attributed to the long-term electron trapping in QDs leading to an increase in photoconductive gain through different charge carrier productions for an absorbed photon. This inherent gain reduces the necessity for an external amplifier, lowers energy usage, and creates a small device for wearable applications. It can distinguish specific light wavelengths that hemoglobin in blood absorbs and assess volumetric variations in blood vessels connected to HR. Additionally, the wearable electronic device enables the forecast of saturated oxygen levels (SpO2) in the blood because light absorption differs from oxygen levels in blood cells. This can also function as HR and respiration rate (RR) indicators on human skin. Surprisingly, the sensor is sensitive enough to use ambient light, which is essential for useful applications in health monitoring. As illustrated in Figure 10a–c, Polat et al. recently reported a novel class of transparent and flexible wearables fabricated of graphene-coupled semiconducting quantum dots (GQD) [145]. They reported numerous prototype wearable devices that noninvasively monitor essential health signals such as HR, RR, and SpO2. It demonstrates the utilization of ambient light with minimal power usage. Additionally, near-field communication (NFC) circuit boards with flexible UV-sensitive photodetectors enable wireless data and power transmission between photodetectors and cellphones. This technology opens the door for entirely connected wearables through wireless UV-index probing and empowers the user (Figure 10d–j).

5.2. Light-Emitting Diodes

Having the possibility of direct bandgap to the near-infrared (NIR), single-layer TMDs present a promising opportunity for optoelectronic applications and offer promise in the fabrication of LEDs [147,148]. Several studies have demonstrated electroluminescence (EL) in a monolayer MoS2-based transistor, possibly due to a similar excited state (i.e., exciton-A) [149]. Since EL emission is restricted to metal contacts in the case of single-layer MoS2, QE was limited to 10−5. Fabricating a LED based on a p–n diode proved to be a practical approach to increasing EL efficiency. Recently, a dual-gating method efficiently determined the in-plane p–n junctions in monolayer TMDs in which the active area was centered at the depletion region [150,151].
On the contrary, vertical vdWHs have emerged in place of effective carrier addition in LEDs because of the wide active area that spans overlapping junctions [152,153]. For example, Duan et al. [152] studied EL characteristics of p–WSe2/n–MoS2 diodes (Figure 11a). However, after a 3V forward bias, photocurrent demonstrated that the EL signal primarily originated from the overlapped region close to metal electrodes. Only a few electrons could move from MoS2 to WSe2, but WSe2 holes are injected into MoS2 when a forward bias was minimal and below a certain threshold. In MoS2, radiative recombination was minimal because of the indirect bandgap. Excitons are added into p- and n-type materials when the bias voltage is raised over the threshold. In addition, they could flow through the junction because of the increasing migration of MoS2 CB. Whereas radiative recombination in WSe2 dominated the EL, it increases linearly with injected current. Remarkably, the fitting of EL spectra using multiple Gaussian functions caused prominent peaks at 546 and 483 nm that corresponded to electron luminescence; these peaks can be used to investigate the interaction among electron-orbital in the lattice of WSe2. Notably, the EL peaks of the electrons were found at 546 and 483 nm after data fitting using Gaussian functions. These peaks are significant for interpreting the electron-orbital interaction in WSe2 (specifically, the interaction between the electron and its orbital). To create functional stacking of structures for controlled light emission, it is necessary to incorporate tunneling layers into electrode contacts and p–n junctions. This reduces leakage current in stacked arrangements and increases the lifespan of electrons and holes in TMD quantum wells [154,155]. A metal-insulator semiconductor-based tunnel diode made of monolayer WS2, multi-layer graphene, and h-BN exhibited an excellent QE (~1%) [156]. As shown in Figure 11b (quantum wells), the graphene-based effective LEDs created by Novoselov et al. consisted of conductive layers in which h-BN served as tunneling barriers and a variety of TMDs as quantum wells [157]. Advanced current densities and bridged contact resistance enabled luminous LEDs by pumping excitons from graphene electrodes into the TMD layer (Figure 11c). Nevertheless, there was an abrupt shift in the PL spectrum at a specific gate voltage, showing a different peak at 1.90 eV. This shift was produced via the Fermi level of the bottom graphene moving over the CB of MoS2, enabling electrons to flow into the quantum well. They also included numerous quantum wells stacked in sequence to enhance the possibility of radiatively recombining the injected carriers. The acquired QE could approach 10%, almost 10-fold greater than planar p–n diodes [150] and, interestingly, 100 times better than Schottky-barrier diodes [149]. In addition, monolayer TMDs had direct bandgap semiconductors consistent with tremendous tunability, making them perfect for atomically thin LEDs. White LEDs offer promising potential in lighting and display because of their elevated brightness at low power consumption and an extended lifespan for efficient operations. Chen et al. engineered a white LED that comprises n-MoS2/p-MoS2/p-GaN as blue, green, and orange emitters [158]. These heterostructures exposed the capacity to create light sources (atomically thin) with white LEDs, with PL spectra at 481 nm (blue, p-GaN), 525 nm (green, p-MoS2), and 642 nm (orange, n-MoS2) [158]. Since type-II band alignment is primarily acquired in vdWHs, the excitons pairs can be effectively separated. Furthermore, thermal light emission from graphene [159] and MoS2 [159] was reported with improved bright-light emission through the suspended device. However, because of limited interlayer separation, temporally isolated electrons and holes endure significant Coulomb interactions, leading to a tightly bound interlayer exciton (XI) [160]. To investigate the EL characteristics, Xu et al. [161] electrostatically developed a MoSe2/WSe2 hetero-bilayer-based in-plane p–n diode (Figure 11d). Several peaks in MoSe2 and WSe2 PL spectra from 1.56 to 1.74 eV indicated intralayer A-excitons and neutral, charged, and localized excitons (Figure 11e). Furthermore, Xu et al. [161] created a heterostructure consisting of stacked graphene layers as electrode materials separated by a thin layer of WSe2 with BN barriers, as visualized in Figure 11f [162]. When the external bias was zero, the Fermi level for graphene layers fell within the bandgap limits of WSe2. Moreover, the addition of spin from a ferromagnetic electrode was introduced by Kis et al. into a WSe2/MoS2 monolayer heterostructure [163]. This caused circularly polarized light emission that was controllable by an external magnetic field (Figure 11g).

5.3. Photovoltaics

Attempts have been made to simultaneously explore the conversion of light energy by means of advancing vdWH-based photodetectors and LEDs. As reported by Jariwala et al., 2D vdWHs have a power-conversion efficiency (PCE) of greater than 25% [163], which is equivalent to standard solar cells. Using vdW 2D heterojunctions that consist of higher-quality hetero interfaces allows for excellent charge separation when exposed to photo-excitation; this could result in more significant potential for ultrathin and lightweight photovoltaic applications. Furthermore, combining monolayer 2D materials, conventional semiconductors, and direct bandgap materials can produce effective single junctions. Following this, the fabrication of a photovoltaic device described by Sanchez et al. consisting of a MoS2/p-Si monolayer-based hybrid p–n heterojunction demonstrated a broad-spectrum response and an EQE of more than 4% [165]. Moreover, studies have been performed on chemical and/or electrostatic doping in the same 2D flakes to design new homojunctions [166,167]. At 0.14 V open-circuit voltage (Voc), Li et al. demonstrated a NIR photovoltaic outcome in a p–n BP homogenous junction doped with Al when irradiated with a 1550 nm laser. The proposed photovoltaic device showed a PEC of 0.66% [168]. Pospischil et al. also investigated an in-plane p–n homojunction based on a WSe2 monolayer using electrostatic doping under local gate control and demonstrated a PCE of 0.5% in addition to a fill factor of 0.5 [166]. A substitute method for building efficient and adaptable solar devices is to use vdWHs fabricated by stacked 2D crystals. The evaluation of photovoltaic impact was provided by Brignell et al. for graphene/WS2 or MoS2/graphene stacks realized via the dry-transfer technique [26]. A significant photocurrent of 1 A has been detected in vertical graphene/WS2/graphene heterostructures by providing a doping effect to two distinct layers of graphene and tuning the Fermi levels appropriately. Furthermore, placing Au nanospheres over graphene/TMDs/vdWH causes maximized light absorption of the photoactive layer, leading to a 10-fold increase in photocurrent. Moreover, Yu et al. reported a better photovoltaic effect in vertical stacks of graphene/MoS2/metal and graphene/MoS2/graphene, attaining 0.3 V of Voc, 50 s of rapid temporal response, and 2 μA short-circuit current (Figure 12a–c) [169]. As a dual-gate device, a graphene/MoS2/graphene vertical heterostructure can also be constructed, which allows the external electric field to regulate the amplitude of the output photocurrent. They show significant promise for solar-energy applications due to their ability to segregate photo-excited carriers effectively. The photocurrent is produced upon diffusion of relaxed electron holes to connections. Furchi et al. used monolayers of MoS2 (n-type) and WSe2 (p-type) to build a vdW p–n heterojunction. Upon illumination under white light, they observed a substantial photovoltaic effect with a 0.5 fill factor and 0.2% PCE [170]. The lowest-energy CB states can be found in MoS2 layers, whereas the highest-energy VB states can be found in WSe2 layers, forming a type-II heterojunction. In addition, WSe2 and MoS2 produce exciton pairs when they absorb photons from incoming light. Moreover, charge transfer happens through the heterojunction because of type-II offsets of CB and VB after the relaxation of photogenerated carriers. By introducing graphene electrodes directly to the top and bottom of a vertical MoS2/WSe2 junction, Lee et al. (Figure 12d) developed a graphene-sandwiched vdWH. Instead of diffusion, this method enables more effective vertical charge transfer to collect the carriers. The photocurrent can be acquired where the graphene electrode and p–n junction overlap, according to photocurrent mapping (Figure 12e). Figure 12f shows 70 nA ISC under a 532 nm laser (920 Wcm−2), demonstrating the lateral device’s superior carrier -collecting performance. Further structural engineering of vertical vdW p–n heterojunctions can result in a broader spectrum of absorption spectra. Long et al. [171] employed MoS2/graphene/WSe2 (p–g–n) vdWH to produce a broadband photovoltaic detector (Figure 12g). The optical image at a scale bar of 5 μm is presented in Figure 12h, where yellow, light gray, and green dashed lines highlight MoS2, graphene, and WSe2, respectively. This indicates that the photocurrent originates from the narrow p–g–n junction, as presented in the photocurrent mapping (Figure 12i), at Vds = 0 V when irradiated with an 830 nm laser (with a power of ~20.5 W). Hence, a noticeable photo response occurs at the touched area of WSe2, graphene, and MoS2.

5.4. Acoustic Cavities

The use of 2D materials in acoustic cavities is one of the most promising research areas [172,173]. At the same time, the heterostructures of such materials exhibit superior physical properties and provide robust responses [173,174]. Zalalutdinov et al. demonstrated a room-temperature longitudinal acoustic phonon lifetime in h-BN- and MoS2-based vdWHs (Figure 13) [173]. It exhibited a frequency range of 50–600 GHz, phonon lifetime of 2 ns at 100 GHz in the case of MoS2, quality-factor index (f × Q) > 1014, and coupling power (i.e., 47 GHz) for acoustic cavities (Γ). Compared to the h-BN phonon lifetime (0.2 ns), these findings describe significant promise for applying 2D materials at a broad spectrum of frequencies from GHz to THz. These results stimulate advanced acoustic cavity-design engineering with phonon-mediated signals and provide a further degree of freedom by manipulating the quantum nature of phonons using 2D vdWH.

6. Challenges and Prospects

Combining different materials and controlling their properties opens up new possibilities for exploring their potential in various applications. However, more research is needed to improve their performance and overcome the challenges associated with implementing these structures. The investigation of innovative 2D materials, for example, graphene, TMDs, BP, MXenes, and others, has led to advances in functional information devices. From theoretical design to device configuration, as well as in the preparation of materials and integration techniques, there has been significant progress in developing 2D material-based optoelectronic devices. This includes sensors, frequency converters, plasmon-generator modulators, ultrafast lasers, and photodetectors. Due to their vast potential, 2D materials can effectively improve the performance of optoelectronic devices and have gained substantial consideration in this field. This article summarizes the properties of current 2D materials and their applications in the field of optoelectronics.
In recent years, 2D vdW structures have been the focus of increased research, but only a few structures have been synthesized experimentally. Derivatives of two naturally occurring vdW materials, molybdenite and graphite, produced through exfoliation and synthesis, have fascinated the scientific community. In addition, to improve the performance of 2D materials, modifications in the fabrication and pre-/post-processing approaches are essential. Techniques such as doping, chemical modification, and electrostatic control can also provide the maximum potential of a material while overcoming possible limitations. In particular, paying attention to the complexities associated with producing high-quality 2D heterostructures with significant phase purity and precise controllability is important. In this respect, it is essential to highlight synthesis methods. Two-dimensional materials can be obtained in several ways, including liquid-phase exfoliation (LPE), mechanical exfoliation (ME), MOCVD, and CVD. However, there are limitations to most of these approaches that prevent further research and application of 2D materials. For instance, many fundamental studies and characterization tools have widely been used to study 2D materials produced by dry mechanical exfoliation. Nonetheless, the lateral dimensions of the layers are usually in micrometers, and it is tricky to cope with the thickness. In contrast, it is possible to produce wafer-scale monolayers using CVD, but the polycrystallinity of the material leads to higher defect density and produces low-quality sheets. To overcome this, CVD systems can be modified to have a better vacuum, more uniform gas flow/pulse, and other reactors in parallel to controlling chemical interactions between precursors in the vapor phase. In addition, the uppermost surface of metal substrates (such as Cu, Ge, W, etc.) must be fabricated first from single-crystal substrates. Even though LPE is scalable for numerous applications, it often lacks high-quality flakes and produces small flakes without precise control. Concerning morphological characterizations, SEM, TEM, STM, and spectroscopic techniques (e.g., Raman, FTIR, etc.) are typically used to analyze 2D materials but specific imaging or characterization techniques are also crucial. However, the effect of lateral thickness, number of layers, doping, modulation of charge carriers, and optimization of protocols to tune electronic and optoelectronic properties needed to be thoroughly investigated to obtain controlled performance and quality. Additionally, it is vital to examine defects and doping at the atomic scale and understand how they affect the band structures of 2D materials.
Importantly, there is also a lack of detailed underlying physics that can be used to analyze carrier transport in vdWHs, which impedes the development of high-performance vdWH devices. Equivalently, band alignment is a fundamental concept and pivotal to understand when describing device physics. Depending on the material, band alignment can be type I, type II, or type III, and it is possible to assess band alignment analytically (e.g., via PL measurements) or theoretically. Significantly, due to the large and expanding family of 2D materials and the infinite number of possible heterostructures, a systematic method for estimating the band alignment of vdWHs is needed. Understanding the physics and adjusting/tuning the coupling between layers and interlayer alignment is also tricky. These variables drastically affect the band alignment and, therefore, the applications of vdWHs. Such technological constraints make it difficult for researchers to overcome the current limitations and use the physical properties of vdW structures for practical purposes. Besides the issues mentioned above, the environmental stability of 2D heterostructures is another critical concern that needs to be addressed to achieve outstanding potential for real-world applications. This includes thermal decomposition, chemical stability, oxidative/environmental stability, and mechanical stability (especially for flexible optoelectronics). Significant experimental/theoretical studies have been performed [175]; however, it needs to be addressed meticulously. In conclusion, and as far as practical aspects are concerned, there are still many obstacles to overcome, but these obstacles could open up many doors. In addition, based on the current state of development, we present here some key directions and our subjective assessment of the development trend:
  • Functional optoelectronic devices based on vdWHs are designed for various optoelectronic applications, including ultrafast lasers, high-speed modulators, ultrasensitive sensors, ultrahigh-responsivity photodetectors, and ultralow damping plasmonics. Most research has focused on graphene, but TMDs and black phosphorus have interesting band structures and spectral responses. Moreover, newly discovered family of 2D materials i.e., MXenes can prove to a better replacement.
  • The possibility of using new metamaterials in optoelectronics, such as perovskites and topological insulators, should be investigated.
  • Develop innovative methods to prepare 2D materials, such as developing functional inks.
  • Research into new optical arrangements and systems for 2D optoelectronic devices, including phonon lasers and extraordinary spots, should be conducted.
  • Incorporating artificial intelligence (AI) into 2D material research to create brain-like devices, and synergistic integration with the Internet of Things (IoT) can lead to the possibility of self-sustaining futuristic smart devices.
  • Several 2D materials can be implanted into Programmable Interface Controllers (PICs), resulting in highly integrated, multifunctional optical devices thanks to recent advances in 2D material production.

Author Contributions

Writing—original draft preparation, Z.U.D.B. and A.R.; writing—review and editing, Z.U.D.B., A.R. and V.I.; data curation, A.C. and V.I.; supervision, V.I.; project administration, V.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank all individuals and organizations permitted to republish their figures and other relevant information.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novoselov, K.S.; Jiang, Z.; Zhang, Y.; Morozov, S.V.; Stormer, H.L.; Zeitler, U.; Maan, J.C.; Boebinger, G.S.; Kim, P.; Geim, A.K. Room-Temperature Quantum Hall Effect in Graphene. Science 2007, 315, 1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Novoselov, K.S.; Fal′ko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef] [PubMed]
  4. Xia, F.; Mueller, T.; Lin, Y.-m.; Valdes-Garcia, A.; Avouris, P. Ultrafast graphene photodetector. Nat. Nanotechnol. 2009, 4, 839–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Koppens, F.H.L.; Mueller, T.; Avouris, P.; Ferrari, A.C.; Vitiello, M.S.; Polini, M. Photodetectors based on graphene, other two-dimensional materials and hybrid systems. Nat. Nanotechnol. 2014, 9, 780–793. [Google Scholar] [CrossRef]
  6. Oostinga, J.B.; Heersche, H.B.; Liu, X.; Morpurgo, A.F.; Vandersypen, L.M.K. Gate-induced insulating state in bilayer graphene devices. Nat Mater 2008, 7, 151–157. [Google Scholar] [CrossRef] [Green Version]
  7. MacNeill, D.; Stiehl, G.M.; Guimaraes, M.H.D.; Buhrman, R.A.; Park, J.; Ralph, D.C. Control of spin–orbit torques through crystal symmetry in WTe2/ferromagnet bilayers. Nat. Phys. 2017, 13, 300–305. [Google Scholar] [CrossRef]
  8. Tang, S.; Zhang, C.; Wong, D.; Pedramrazi, Z.; Tsai, H.-Z.; Jia, C.; Moritz, B.; Claassen, M.; Ryu, H.; Kahn, S.; et al. Quantum spin Hall state in monolayer 1T’-WTe2. Nat. Phys. 2017, 13, 683–687. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, Q.; Cai, K.; Li, J.; Huang, Y.; Wang, Z.; Xu, K.; Wang, F.; Zhan, X.; Wang, F.; Wang, K.; et al. Rational Design of Ultralarge Pb1−xSnxTe Nanoplates for Exploring Crystalline Symmetry-Protected Topological Transport. Adv. Mater. 2016, 28, 617–623. [Google Scholar] [CrossRef]
  10. Li, H.; Cao, J.; Zheng, W.; Chen, Y.; Wu, D.; Dang, W.; Wang, K.; Peng, H.; Liu, Z. Controlled Synthesis of Topological Insulator Nanoplate Arrays on Mica. J. Am. Chem. Soc. 2012, 134, 6132–6135. [Google Scholar] [CrossRef]
  11. Viti, L.; Hu, J.; Coquillat, D.; Knap, W.; Tredicucci, A.; Politano, A.; Vitiello, M.S. Black Phosphorus Terahertz Photodetectors. Adv. Mater. 2015, 27, 5567–5572. [Google Scholar] [CrossRef] [Green Version]
  12. Chernikov, A.; Ruppert, C.; Hill, H.M.; Rigosi, A.F.; Heinz, T.F. Population inversion and giant bandgap renormalization in atomically thin WS2 layers. Nat. Photonics 2015, 9, 466–470. [Google Scholar] [CrossRef]
  13. Srivastava, A.; Sidler, M.; Allain, A.V.; Lembke, D.S.; Kis, A.; Imamoğlu, A. Optically active quantum dots in monolayer WSe2. Nat. Nanotechnol. 2015, 10, 491–496. [Google Scholar] [CrossRef]
  14. Cui, F.; Feng, Q.; Hong, J.; Wang, R.; Bai, Y.; Li, X.; Liu, D.; Zhou, Y.; Liang, X.; He, X.; et al. Synthesis of Large-Size 1T′ ReS2xSe2(1−x) Alloy Monolayer with Tunable Bandgap and Carrier Type. Adv. Mater. 2017, 29, 1705015. [Google Scholar] [CrossRef]
  15. Choi, W.; Cho, M.Y.; Konar, A.; Lee, J.H.; Cha, G.-B.; Hong, S.C.; Kim, S.; Kim, J.; Jena, D.; Joo, J.; et al. High-Detectivity Multilayer MoS2 Phototransistors with Spectral Response from Ultraviolet to Infrared. Adv. Mater. 2012, 24, 5832–5836. [Google Scholar] [CrossRef]
  16. Xie, Y.; Zhang, B.; Wang, S.; Wang, D.; Wang, A.; Wang, Z.; Yu, H.; Zhang, H.; Chen, Y.; Zhao, M.; et al. Ultrabroadband MoS2 Photodetector with Spectral Response from 445 to 2717 nm. Adv. Mater. 2017, 29, 1605972. [Google Scholar] [CrossRef]
  17. Xu, Y.; Ali, A.; Shehzad, K.; Meng, N.; Xu, M.; Zhang, Y.; Wang, X.; Jin, C.; Wang, H.; Guo, Y.; et al. Solvent-Based Soft-Patterning of Graphene Lateral Heterostructures for Broadband High-Speed Metal–Semiconductor–Metal Photodetectors. Adv. Mater. Technol. 2017, 2, 1600241. [Google Scholar] [CrossRef] [Green Version]
  18. Kim, S.; Konar, A.; Hwang, W.-S.; Lee, J.H.; Lee, J.; Yang, J.; Jung, C.; Kim, H.; Yoo, J.-B.; Choi, J.-Y.; et al. High-mobility and low-power thin-film transistors based on multilayer MoS2 crystals. Nat. Commun. 2012, 3, 1011. [Google Scholar] [CrossRef] [Green Version]
  19. Huang, J.-K.; Pu, J.; Hsu, C.-L.; Chiu, M.-H.; Juang, Z.-Y.; Chang, Y.-H.; Chang, W.-H.; Iwasa, Y.; Takenobu, T.; Li, L.-J. Large-Area Synthesis of Highly Crystalline WSe2 Monolayers and Device Applications. ACS Nano 2014, 8, 923–930. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, M.; Wu, J.; Lin, L.; Liu, Y.; Deng, B.; Guo, Y.; Lin, Y.; Xie, T.; Dang, W.; Zhou, Y.; et al. Chemically Engineered Substrates for Patternable Growth of Two-Dimensional Chalcogenide Crystals. ACS Nano 2016, 10, 10317–10323. [Google Scholar] [CrossRef]
  21. Zhou, Y.; Nie, Y.; Liu, Y.; Yan, K.; Hong, J.; Jin, C.; Zhou, Y.; Yin, J.; Liu, Z.; Peng, H. Epitaxy and Photoresponse of Two-Dimensional GaSe Crystals on Flexible Transparent Mica Sheets. ACS Nano 2014, 8, 1485–1490. [Google Scholar] [CrossRef] [PubMed]
  22. Chen, X.; Shehzad, K.; Gao, L.; Long, M.; Guo, H.; Qin, S.; Wang, X.; Wang, F.; Shi, Y.; Hu, W.; et al. Graphene Hybrid Structures for Integrated and Flexible Optoelectronics. Adv. Mater. 2020, 32, 1902039. [Google Scholar] [CrossRef] [PubMed]
  23. Novoselov, K.S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A.H. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Geim, A.K.; Grigorieva, I.V. Van der Waals heterostructures. Nature 2013, 499, 419–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, M.-Y.; Shi, Y.; Cheng, C.-C.; Lu, L.-S.; Lin, Y.-C.; Tang, H.-L.; Tsai, M.-L.; Chu, C.-W.; Wei, K.-H.; He, J.-H.; et al. Epitaxial growth of a monolayer WSe2-MoS2 lateral p-n junction with an atomically sharp interface. Science 2015, 349, 524–528. [Google Scholar] [CrossRef] [Green Version]
  26. Britnell, L.; Ribeiro, R.M.; Eckmann, A.; Jalil, R.; Belle, B.D.; Mishchenko, A.; Kim, Y.J.; Gorbachev, R.V.; Georgiou, T.; Morozov, S.V.; et al. Strong Light-Matter Interactions in Heterostructures of Atomically Thin Films. Science 2013, 340, 1311–1314. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, W.; Wang, Q.; Chen, Y.; Wang, Z.; Wee, A.T.S. Van der Waals stacked 2D layered materials for optoelectronics. 2D Mater. 2016, 3, 022001. [Google Scholar] [CrossRef]
  28. Liu, Y.; Weiss, N.O.; Duan, X.; Cheng, H.-C.; Huang, Y.; Duan, X. Van der Waals heterostructures and devices. Nat. Rev. Mater. 2016, 1, 16042. [Google Scholar] [CrossRef]
  29. Gupta, A.; Sakthivel, T.; Seal, S. Recent development in 2D materials beyond graphene. Prog. Mater. Sci. 2015, 73, 44–126. [Google Scholar] [CrossRef]
  30. Coleman, J.N.; Lotya, M.; O’Neill, A.; Bergin, S.D.; King, P.J.; Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R.J.; et al. Two-Dimensional Nanosheets Produced by Liquid Exfoliation of Layered Materials. Science 2011, 331, 568–571. [Google Scholar] [CrossRef] [Green Version]
  31. Fukuda, K.; Akatsuka, K.; Ebina, Y.; Ma, R.; Takada, K.; Nakai, I.; Sasaki, T. Exfoliated Nanosheet Crystallite of Cesium Tungstate with 2D Pyrochlore Structure: Synthesis, Characterization, and Photochromic Properties. ACS Nano 2008, 2, 1689–1695. [Google Scholar] [CrossRef]
  32. Aksit, M.; Hoselton, B.C.; Kim, H.J.; Ha, D.-H.; Robinson, R.D. Synthesis and Properties of Electrically Conductive, Ductile, Extremely Long (∼50 μm) Nanosheets of KxCoO2·yH2O. ACS Appl. Mater. Interfaces 2013, 5, 8998–9007. [Google Scholar] [CrossRef]
  33. Gutiérrez, H.R.; Perea-López, N.; Elías, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; López-Urías, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary Room-Temperature Photoluminescence in Triangular WS2 Monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef] [Green Version]
  34. Kakanakova-Georgieva, A.; Gueorguiev, G.K.; Sangiovanni, D.G.; Suwannaharn, N.; Ivanov, I.G.; Cora, I.; Pécz, B.; Nicotra, G.; Giannazzo, F. Nanoscale phenomena ruling deposition and intercalation of AlN at the graphene/SiC interface. Nanoscale 2020, 12, 19470–19476. [Google Scholar] [CrossRef]
  35. Verble, J.L.; Wieting, T.J. Lattice Mode Degeneracy in MoS2 and Other Layer Compounds. Phys. Rev. Lett. 1970, 25, 362–365. [Google Scholar] [CrossRef]
  36. Kim, J.; Cote, L.J.; Kim, F.; Huang, J. Visualizing Graphene Based Sheets by Fluorescence Quenching Microscopy. J. Am. Chem. Soc. 2010, 132, 260–267. [Google Scholar] [CrossRef]
  37. Nemes-Incze, P.; Osváth, Z.; Kamarás, K.; Biró, L.P. Anomalies in thickness measurements of graphene and few layer graphite crystals by tapping mode atomic force microscopy. Carbon 2008, 46, 1435–1442. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, X.; Song, J.; Liu, J.; Wang, Z.L. Direct-Current Nanogenerator Driven by Ultrasonic Waves. Science 2007, 316, 102–105. [Google Scholar] [CrossRef] [Green Version]
  39. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  40. Pham, P.V.; Bodepudi, S.C.; Shehzad, K.; Liu, Y.; Xu, Y.; Yu, B.; Duan, X. 2D Heterostructures for Ubiquitous Electronics and Optoelectronics: Principles, Opportunities, and Challenges. Chem. Rev. 2022, 122, 6514–6613. [Google Scholar] [CrossRef]
  41. Feng, L.; Zhang, W.X. The structure and magnetism of graphone. AIP Adv. 2012, 2, 042138. [Google Scholar] [CrossRef]
  42. Zhou, J.; Xie, Z.; Liu, R.; Gao, X.; Li, J.; Xiong, Y.; Tong, L.; Zhang, J.; Liu, Z. Synthesis of Ultrathin Graphdiyne Film Using a Surface Template. ACS Appl. Mater. Interfaces 2019, 11, 2632–2637. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, M.-C.; Huang, C.-C.; Cheung, C.-H.; Chen, C.-Y.; Tan, S.G.; Huang, T.-W.; Zhao, Y.; Zhao, Y.; Wu, G.; Feng, Y.-P.; et al. Prospects and Opportunities of 2D van der Waals Magnetic Systems. Ann. Der Phys. 2020, 532, 1900452. [Google Scholar] [CrossRef] [Green Version]
  44. Liu, X.-Y.; Chen, H.; Wang, R.; Shang, Y.; Zhang, Q.; Li, W.; Zhang, G.; Su, J.; Dinh, C.T.; de Arquer, F.P.G.; et al. 0D–2D Quantum Dot: Metal Dichalcogenide Nanocomposite Photocatalyst Achieves Efficient Hydrogen Generation. Adv. Mater. 2017, 29, 1605646. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, E.; Xu, X.; Zou, Y.-C.; Ai, L.; Dong, X.; Huang, C.; Leng, P.; Liu, S.; Zhang, Y.; Jia, Z.; et al. Nonreciprocal superconducting NbSe2 antenna. Nat. Commun. 2020, 11, 5634. [Google Scholar] [CrossRef]
  46. Osada, M.; Sasaki, T. The rise of 2D dielectrics/ferroelectrics. APL Mater. 2019, 7, 120902. [Google Scholar] [CrossRef] [Green Version]
  47. Shi, J.; Huan, Y.; Xiao, M.; Hong, M.; Zhao, X.; Gao, Y.; Cui, F.; Yang, P.; Pennycook, S.J.; Zhao, J.; et al. Two-Dimensional Metallic NiTe2 with Ultrahigh Environmental Stability, Conductivity, and Electrocatalytic Activity. ACS Nano 2020, 14, 9011–9020. [Google Scholar] [CrossRef]
  48. Wang, H.; Li, C.; Fang, P.; Zhang, Z.; Zhang, J.Z. Synthesis, properties, and optoelectronic applications of two-dimensional MoS2 and MoS2-based heterostructures. Chem. Soc. Rev. 2018, 47, 6101–6127. [Google Scholar] [CrossRef]
  49. Zhao, S.; Hotta, T.; Koretsune, T.; Watanabe, K.; Taniguchi, T.; Sugawara, K.; Takahashi, T.; Shinohara, H.; Kitaura, R. Two-dimensional metallic NbS2: Growth, optical identification and transport properties. 2d Mater. 2016, 3, 025027. [Google Scholar] [CrossRef]
  50. Najafi, L.; Bellani, S.; Oropesa-Nuñez, R.; Martín-García, B.; Prato, M.; Mazánek, V.; Debellis, D.; Lauciello, S.; Brescia, R.; Sofer, Z.; et al. Niobium disulphide (NbS2)-based (heterogeneous) electrocatalysts for an efficient hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 25593–25608. [Google Scholar] [CrossRef] [Green Version]
  51. Chen, C.; Chen, X.; Wu, C.; Wang, X.; Ping, Y.; Wei, X.; Zhou, X.; Lu, J.; Zhu, L.; Zhou, J.; et al. Air-Stable 2D Cr5Te8 Nanosheets with Thickness-Tunable Ferromagnetism. Adv. Mater. 2022, 34, 2107512. [Google Scholar] [CrossRef]
  52. Rahman, S.; Torres, J.F.; Khan, A.R.; Lu, Y. Recent Developments in van der Waals Antiferromagnetic 2D Materials: Synthesis, Characterization, and Device Implementation. ACS Nano 2021, 15, 17175–17213. [Google Scholar] [CrossRef]
  53. Wang, Y.D.; Yao, W.L.; Xin, Z.M.; Han, T.T.; Wang, Z.G.; Chen, L.; Cai, C.; Li, Y.; Zhang, Y. Band insulator to Mott insulator transition in 1T-TaS2. Nat. Commun. 2020, 11, 4215. [Google Scholar] [CrossRef]
  54. Sun, Y.; Li, Y.; Li, T.; Biswas, K.; Patanè, A.; Zhang, L. New Polymorphs of 2D Indium Selenide with Enhanced Electronic Properties. Adv. Funct. Mater. 2020, 30, 2001920. [Google Scholar] [CrossRef]
  55. Wang, H.; Chen, Y.; Zhu, C.; Wang, X.; Zhang, H.; Tsang, S.H.; Li, H.; Lin, J.; Yu, T.; Liu, Z.; et al. Synthesis of Atomically Thin 1T-TaSe2 with a Strongly Enhanced Charge-Density-Wave Order. Adv. Funct. Mater. 2020, 30, 2001903. [Google Scholar] [CrossRef]
  56. Zhang, M.; He, Y.; Yan, D.; Xu, H.; Wang, A.; Chen, Z.; Wang, S.; Luo, H.; Yan, K. Multifunctional 2H-TaS2 nanoflakes for efficient supercapacitors and electrocatalytic evolution of hydrogen and oxygen. Nanoscale 2019, 11, 22255–22260. [Google Scholar] [CrossRef]
  57. Chen, Y.; Lai, Z.; Zhang, X.; Fan, Z.; He, Q.; Tan, C.; Zhang, H. Phase engineering of nanomaterials. Nat. Rev. Chem. 2020, 4, 243–256. [Google Scholar] [CrossRef]
  58. Yin, X.; Tang, C.S.; Zheng, Y.; Gao, J.; Wu, J.; Zhang, H.; Chhowalla, M.; Chen, W.; Wee, A.T.S. Recent developments in 2D transition metal dichalcogenides: Phase transition and applications of the (quasi-)metallic phases. Chem. Soc. Rev. 2021, 50, 10087–10115. [Google Scholar] [CrossRef]
  59. Hu, Z.; Niu, T.; Guo, R.; Zhang, J.; Lai, M.; He, J.; Wang, L.; Chen, W. Two-dimensional black phosphorus: Its fabrication, functionalization and applications. Nanoscale 2018, 10, 21575–21603. [Google Scholar] [CrossRef]
  60. Xu, Y.; Shi, Z.; Shi, X.; Zhang, K.; Zhang, H. Recent progress in black phosphorus and black-phosphorus-analogue materials: Properties, synthesis and applications. Nanoscale 2019, 11, 14491–14527. [Google Scholar] [CrossRef]
  61. Ghosh, T.; Samanta, M.; Vasdev, A.; Dolui, K.; Ghatak, J.; Das, T.; Sheet, G.; Biswas, K. Ultrathin Free-Standing Nanosheets of Bi2O2Se: Room Temperature Ferroelectricity in Self-Assembled Charged Layered Heterostructure. Nano Lett. 2019, 19, 5703–5709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Liu, F.; You, L.; Seyler, K.L.; Li, X.; Yu, P.; Lin, J.; Wang, X.; Zhou, J.; Wang, H.; He, H.; et al. Room-temperature ferroelectricity in CuInP2S6 ultrathin flakes. Nat. Commun. 2016, 7, 12357. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, H.; Pan, L.; Xiao, M.; Fang, J.; Cui, Y.; Wei, Z. Iron-doping induced multiferroic in two-dimensional In2Se3. Sci. China Mater. 2020, 63, 421–428. [Google Scholar] [CrossRef] [Green Version]
  64. Behera, B.; Sutar, B.C.; Pradhan, N.R. Recent progress on 2D ferroelectric and multiferroic materials, challenges, and opportunity. Emergent Mater. 2021, 4, 847–863. [Google Scholar] [CrossRef]
  65. Zhao, Y.; Zhao, J.; Li, Y.; Ma, D.; Hou, S.; Li, L.; Hao, X.; Wang, Z. Room temperature synthesis of 2D CuO nanoleaves in aqueous solution. Nanotechnology 2011, 22, 115604. [Google Scholar] [CrossRef]
  66. Wen, X.; Zhang, W.; Yang, S. Synthesis of Cu(OH)2 and CuO Nanoribbon Arrays on a Copper Surface. Langmuir 2003, 19, 5898–5903. [Google Scholar] [CrossRef]
  67. Pereira, C.C.L.; Lima, J.C.; Moro, A.J.; Monteiro, B. Layered europium hydroxide system for phosphorous sensing and remediation. Appl. Clay Sci. 2017, 146, 216–222. [Google Scholar] [CrossRef]
  68. Ares, P.; Wang, Y.B.; Woods, C.R.; Dougherty, J.; Fumagalli, L.; Guinea, F.; Davidovitch, B.; Novoselov, K.S. Van der Waals interaction affects wrinkle formation in two-dimensional materials. Proc. Natl. Acad. Sci. USA 2021, 118, e2025870118. [Google Scholar] [CrossRef]
  69. Papadopoulou, K.A.; Chroneos, A.; Parfitt, D.; Christopoulos, S.-R.G. A perspective on MXenes: Their synthesis, properties, and recent applications. J. Appl. Phys. 2020, 128, 170902. [Google Scholar] [CrossRef]
  70. Dean, C.R.; Young, A.F.; Meric, I.; Lee, C.; Wang, L.; Sorgenfrei, S.; Watanabe, K.; Taniguchi, T.; Kim, P.; Shepard, K.L.; et al. Boron nitride substrates for high-quality graphene electronics. Nat. Nanotechnol. 2010, 5, 722–726. [Google Scholar] [CrossRef]
  71. Vu, Q.A.; Fan, S.; Lee, S.H.; Joo, M.-K.; Yu, W.J.; Lee, Y.H. Near-zero hysteresis and near-ideal subthreshold swing in h-BN encapsulated single-layer MoS2 field-effect transistors. 2D Mater. 2018, 5, 031001. [Google Scholar] [CrossRef]
  72. Mayorov, A.S.; Gorbachev, R.V.; Morozov, S.V.; Britnell, L.; Jalil, R.; Ponomarenko, L.A.; Blake, P.; Novoselov, K.S.; Watanabe, K.; Taniguchi, T.; et al. Micrometer-Scale Ballistic Transport in Encapsulated Graphene at Room Temperature. Nano Lett. 2011, 11, 2396–2399. [Google Scholar] [CrossRef] [Green Version]
  73. Das, S.; Gulotty, R.; Sumant, A.V.; Roelofs, A. All Two-Dimensional, Flexible, Transparent, and Thinnest Thin Film Transistor. Nano Lett. 2014, 14, 2861–2866. [Google Scholar] [CrossRef]
  74. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black phosphorus field-effect transistors. Nat. Nanotechnol. 2014, 9, 372–377. [Google Scholar] [CrossRef] [Green Version]
  75. Xi, X.; Wang, Z.; Zhao, W.; Park, J.-H.; Law, K.T.; Berger, H.; Forró, L.; Shan, J.; Mak, K.F. Ising pairing in superconducting NbSe2 atomic layers. Nat. Phys. 2016, 12, 139–143. [Google Scholar] [CrossRef] [Green Version]
  76. Huang, B.; Clark, G.; Klein, D.R.; MacNeill, D.; Navarro-Moratalla, E.; Seyler, K.L.; Wilson, N.; McGuire, M.A.; Cobden, D.H.; Xiao, D.; et al. Electrical control of 2D magnetism in bilayer CrI3. Nat. Nanotechnol. 2018, 13, 544–548. [Google Scholar] [CrossRef]
  77. Cui, X.; Lee, G.-H.; Kim, Y.D.; Arefe, G.; Huang, P.Y.; Lee, C.-H.; Chenet, D.A.; Zhang, X.; Wang, L.; Ye, F.; et al. Multi-terminal transport measurements of MoS2 using a van der Waals heterostructure device platform. Nat. Nanotechnol. 2015, 10, 534–540. [Google Scholar] [CrossRef]
  78. Yu, G.L.; Gorbachev, R.V.; Tu, J.S.; Kretinin, A.V.; Cao, Y.; Jalil, R.; Withers, F.; Ponomarenko, L.A.; Piot, B.A.; Potemski, M.; et al. Hierarchy of Hofstadter states and replica quantum Hall ferromagnetism in graphene superlattices. Nat. Phys. 2014, 10, 525–529. [Google Scholar] [CrossRef] [Green Version]
  79. Yankowitz, M.; Xue, J.; Cormode, D.; Sanchez-Yamagishi, J.D.; Watanabe, K.; Taniguchi, T.; Jarillo-Herrero, P.; Jacquod, P.; LeRoy, B.J. Emergence of superlattice Dirac points in graphene on hexagonal boron nitride. Nat. Phys. 2012, 8, 382–386. [Google Scholar] [CrossRef] [Green Version]
  80. Woods, C.R.; Britnell, L.; Eckmann, A.; Ma, R.S.; Lu, J.C.; Guo, H.M.; Lin, X.; Yu, G.L.; Cao, Y.; Gorbachev, R.V.; et al. Commensurate–incommensurate transition in graphene on hexagonal boron nitride. Nat. Phys. 2014, 10, 451–456. [Google Scholar] [CrossRef]
  81. Chittari, B.L.; Chen, G.; Zhang, Y.; Wang, F.; Jung, J. Gate-Tunable Topological Flat Bands in Trilayer Graphene Boron-Nitride Moir\’e Superlattices. Phys. Rev. Lett. 2019, 122, 016401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Cao, Y.; Fatemi, V.; Fang, S.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Jarillo-Herrero, P. Unconventional superconductivity in magic-angle graphene superlattices. Nature 2018, 556, 43–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Tran, K.; Moody, G.; Wu, F.; Lu, X.; Choi, J.; Kim, K.; Rai, A.; Sanchez, D.A.; Quan, J.; Singh, A.; et al. Evidence for moiré excitons in van der Waals heterostructures. Nature 2019, 567, 71–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Seyler, K.L.; Rivera, P.; Yu, H.; Wilson, N.P.; Ray, E.L.; Mandrus, D.G.; Yan, J.; Yao, W.; Xu, X. Signatures of moiré-trapped valley excitons in MoSe2/WSe2 heterobilayers. Nature 2019, 567, 66–70. [Google Scholar] [CrossRef] [Green Version]
  85. Jin, C.; Regan, E.C.; Yan, A.; Iqbal Bakti Utama, M.; Wang, D.; Zhao, S.; Qin, Y.; Yang, S.; Zheng, Z.; Shi, S.; et al. Observation of moiré excitons in WSe2/WS2 heterostructure superlattices. Nature 2019, 567, 76–80. [Google Scholar] [CrossRef] [Green Version]
  86. Alexeev, E.M.; Ruiz-Tijerina, D.A.; Danovich, M.; Hamer, M.J.; Terry, D.J.; Nayak, P.K.; Ahn, S.; Pak, S.; Lee, J.; Sohn, J.I.; et al. Resonantly hybridized excitons in moiré superlattices in van der Waals heterostructures. Nature 2019, 567, 81–86. [Google Scholar] [CrossRef] [Green Version]
  87. Mak, K.F.; Shan, J. Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides. Nat. Photonics 2016, 10, 216–226. [Google Scholar] [CrossRef]
  88. Pradhan, N.R.; Talapatra, S.; Terrones, M.; Ajayan, P.M.; Balicas, L. Optoelectronic Properties of Heterostructures: The Most Recent Developments Based on Graphene and Transition-Metal Dichalcogenides. IEEE Nanotechnol. Mag. 2017, 11, 18–32. [Google Scholar] [CrossRef]
  89. Lemme, M.C.; Akinwande, D.; Huyghebaert, C.; Stampfer, C. 2D materials for future heterogeneous electronics. Nat. Commun. 2022, 13, 1392. [Google Scholar] [CrossRef]
  90. Li, H.; Wu, J.-B.; Ran, F.; Lin, M.-L.; Liu, X.-L.; Zhao, Y.; Lu, X.; Xiong, Q.; Zhang, J.; Huang, W.; et al. Interfacial Interactions in van der Waals Heterostructures of MoS2 and Graphene. ACS Nano 2017, 11, 11714–11723. [Google Scholar] [CrossRef]
  91. Fang, H.; Battaglia, C.; Carraro, C.; Nemsak, S.; Ozdol, B.; Kang, J.S.; Bechtel, H.A.; Desai, S.B.; Kronast, F.; Unal, A.A.; et al. Strong interlayer coupling in van der Waals heterostructures built from single-layer chalcogenides. Proc. Natl. Acad. Sci. USA 2014, 111, 6198–6202. [Google Scholar] [CrossRef] [Green Version]
  92. Hong, X.; Kim, J.; Shi, S.-F.; Zhang, Y.; Jin, C.; Sun, Y.; Tongay, S.; Wu, J.; Zhang, Y.; Wang, F. Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures. Nat. Nanotechnol. 2014, 9, 682–686. [Google Scholar] [CrossRef] [Green Version]
  93. Grosso, G.; Graves, J.; Hammack, A.T.; High, A.A.; Butov, L.V.; Hanson, M.; Gossard, A.C. Excitonic switches operating at around 100 K. Nat. Photonics 2009, 3, 577–580. [Google Scholar] [CrossRef]
  94. Calman, E.V.; Fogler, M.M.; Butov, L.V.; Hu, S.; Mishchenko, A.; Geim, A.K. Indirect excitons in van der Waals heterostructures at room temperature. Nat. Commun. 2018, 9, 1895. [Google Scholar] [CrossRef] [Green Version]
  95. Wu, L.; Chen, Y.; Zhou, H.; Zhu, H. Ultrafast Energy Transfer of Both Bright and Dark Excitons in 2D van der Waals Heterostructures Beyond Dipolar Coupling. ACS Nano 2019, 13, 2341–2348. [Google Scholar] [CrossRef]
  96. Tan, C.; Zhang, H. Epitaxial Growth of Hetero-Nanostructures Based on Ultrathin Two-Dimensional Nanosheets. J. Am. Chem. Soc. 2015, 137, 12162–12174. [Google Scholar] [CrossRef]
  97. Lau, S.P.; Li, L.-J.; Chai, Y. Advances in Two-Dimensional Layered Materials. Adv. Funct. Mater. 2017, 27, 1701403. [Google Scholar] [CrossRef] [Green Version]
  98. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, W.; Huang, J.-K.; Chen, C.-H.; Chang, Y.-H.; Cheng, Y.-J.; Li, L.-J. High-Gain Phototransistors Based on a CVD MoS2 Monolayer. Adv. Mater. 2013, 25, 3456–3461. [Google Scholar] [CrossRef]
  100. Zhou, X.; Zhang, Q.; Gan, L.; Li, H.; Xiong, J.; Zhai, T. Booming Development of Group IV–VI Semiconductors: Fresh Blood of 2D Family. Adv. Sci. 2016, 3, 1600177. [Google Scholar] [CrossRef] [Green Version]
  101. Zhang, C.; Gong, C.; Nie, Y.; Min, K.-A.; Liang, C.; Oh, Y.J.; Zhang, H.; Wang, W.; Hong, S.; Colombo, L.; et al. Systematic study of electronic structure and band alignment of monolayer transition metal dichalcogenides in Van der Waals heterostructures. 2d Mater. 2017, 4, 015026. [Google Scholar] [CrossRef]
  102. Kang, J.; Tongay, S.; Zhou, J.; Li, J.; Wu, J. Band offsets and heterostructures of two-dimensional semiconductors. Appl. Phys. Lett. 2013, 102, 012111. [Google Scholar] [CrossRef] [Green Version]
  103. Chiu, M.-H.; Zhang, C.; Shiu, H.-W.; Chuu, C.-P.; Chen, C.-H.; Chang, C.-Y.S.; Chen, C.-H.; Chou, M.-Y.; Shih, C.-K.; Li, L.-J. Determination of band alignment in the single-layer MoS2/WSe2 heterojunction. Nat. Commun. 2015, 6, 7666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Gong, C.; Zhang, H.; Wang, W.; Colombo, L.; Wallace, R.M.; Cho, K. Band alignment of two-dimensional transition metal dichalcogenides: Application in tunnel field effect transistors. Appl. Phys. Lett. 2013, 103, 053513. [Google Scholar] [CrossRef]
  105. Liu, K.; Zhang, L.; Cao, T.; Jin, C.; Qiu, D.; Zhou, Q.; Zettl, A.; Yang, P.; Louie, S.G.; Wang, F. Evolution of interlayer coupling in twisted molybdenum disulfide bilayers. Nat. Commun. 2014, 5, 4966. [Google Scholar] [CrossRef] [Green Version]
  106. Lu, X.; Luo, X.; Zhang, J.; Quek, S.Y.; Xiong, Q. Lattice vibrations and Raman scattering in two-dimensional layered materials beyond graphene. Nano Res. 2016, 9, 3559–3597. [Google Scholar] [CrossRef]
  107. Pawbake, A.S.; Pawar, M.S.; Jadkar, S.R.; Late, D.J. Large area chemical vapor deposition of monolayer transition metal dichalcogenides and their temperature dependent Raman spectroscopy studies. Nanoscale 2016, 8, 3008–3018. [Google Scholar] [CrossRef]
  108. Nayak, P.K.; Horbatenko, Y.; Ahn, S.; Kim, G.; Lee, J.-U.; Ma, K.Y.; Jang, A.R.; Lim, H.; Kim, D.; Ryu, S.; et al. Probing Evolution of Twist-Angle-Dependent Interlayer Excitons in MoSe2/WSe2 van der Waals Heterostructures. ACS Nano 2017, 11, 4041–4050. [Google Scholar] [CrossRef]
  109. Wang, L.; Meric, I.; Huang, P.Y.; Gao, Q.; Gao, Y.; Tran, H.; Taniguchi, T.; Watanabe, K.; Campos, L.M.; Muller, D.A.; et al. One-Dimensional Electrical Contact to a Two-Dimensional Material. Science 2013, 342, 614. [Google Scholar] [CrossRef] [Green Version]
  110. Pizzocchero, F.; Gammelgaard, L.; Jessen, B.S.; Caridad, J.M.; Wang, L.; Hone, J.; Bøggild, P.; Booth, T.J. The hot pick-up technique for batch assembly of van der Waals heterostructures. Nat. Commun. 2016, 7, 11894. [Google Scholar] [CrossRef] [Green Version]
  111. Xu, J.; Singh, S.; Katoch, J.; Wu, G.; Zhu, T.; Žutić, I.; Kawakami, R.K. Spin inversion in graphene spin valves by gate-tunable magnetic proximity effect at one-dimensional contacts. Nat. Commun. 2018, 9, 2869. [Google Scholar] [CrossRef] [Green Version]
  112. Frisenda, R.; Navarro-Moratalla, E.; Gant, P.; Pérez De Lara, D.; Jarillo-Herrero, P.; Gorbachev, R.V.; Castellanos-Gomez, A. Recent progress in the assembly of nanodevices and van der Waals heterostructures by deterministic placement of 2D materials. Chem. Soc. Rev. 2018, 47, 53–68. [Google Scholar] [CrossRef] [Green Version]
  113. Pan, Y.; Fölsch, S.; Nie, Y.; Waters, D.; Lin, Y.-C.; Jariwala, B.; Zhang, K.; Cho, K.; Robinson, J.A.; Feenstra, R.M. Quantum-Confined Electronic States Arising from the Moiré Pattern of MoS2–WSe2 Heterobilayers. Nano Lett. 2018, 18, 1849–1855. [Google Scholar] [CrossRef] [Green Version]
  114. Seo, J.; Lee, J.; Jeong, G.; Park, H. Site-Selective and van der Waals Epitaxial Growth of Rhenium Disulfide on Graphene. Small 2019, 15, 1804133. [Google Scholar] [CrossRef]
  115. Shi, Y.; Zhou, W.; Lu, A.-Y.; Fang, W.; Lee, Y.-H.; Hsu, A.L.; Kim, S.M.; Kim, K.K.; Yang, H.Y.; Li, L.-J.; et al. van der Waals Epitaxy of MoS2 Layers Using Graphene As Growth Templates. Nano Lett. 2012, 12, 2784–2791. [Google Scholar] [CrossRef]
  116. Liu, Z.; Song, L.; Zhao, S.; Huang, J.; Ma, L.; Zhang, J.; Lou, J.; Ajayan, P.M. Direct Growth of Graphene/Hexagonal Boron Nitride Stacked Layers. Nano Lett. 2011, 11, 2032–2037. [Google Scholar] [CrossRef]
  117. Chen, T.; Zhou, Y.; Sheng, Y.; Wang, X.; Zhou, S.; Warner, J.H. Hydrogen-Assisted Growth of Large-Area Continuous Films of MoS2 on Monolayer Graphene. ACS Appl. Mater. Interfaces 2018, 10, 7304–7314. [Google Scholar] [CrossRef]
  118. Oyedele, A.D.; Rouleau, C.M.; Geohegan, D.B.; Xiao, K. The growth and assembly of organic molecules and inorganic 2D materials on graphene for van der Waals heterostructures. Carbon 2018, 131, 246–257. [Google Scholar] [CrossRef]
  119. Yang, T.; Zheng, B.; Wang, Z.; Xu, T.; Pan, C.; Zou, J.; Zhang, X.; Qi, Z.; Liu, H.; Feng, Y.; et al. Van der Waals epitaxial growth and optoelectronics of large-scale WSe2/SnS2 vertical bilayer p–n junctions. Nat. Commun. 2017, 8, 1906. [Google Scholar] [CrossRef] [Green Version]
  120. Wu, R.; Tao, Q.; Dang, W.; Liu, Y.; Li, B.; Li, J.; Zhao, B.; Zhang, Z.; Ma, H.; Sun, G.; et al. van der Waals Epitaxial Growth of Atomically Thin 2D Metals on Dangling-Bond-Free WSe2 and WS2. Adv. Funct. Mater. 2019, 29, 1806611. [Google Scholar] [CrossRef]
  121. Li, X.; Lin, M.-W.; Lin, J.; Huang, B.; Puretzky, A.A.; Ma, C.; Wang, K.; Zhou, W.; Pantelides, S.T.; Chi, M.; et al. Two-dimensional GaSe/MoSe2 misfit bilayer heterojunctions by van der Waals epitaxy. Sci. Adv. 2016, 2, e1501882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Zhou, N.; Wang, R.; Zhou, X.; Song, H.; Xiong, X.; Ding, Y.; Lü, J.; Gan, L.; Zhai, T. P-GaSe/N-MoS2 Vertical Heterostructures Synthesized by van der Waals Epitaxy for Photoresponse Modulation. Small 2018, 14, 1702731. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, H.; Li, D.; Ma, C.; Zhang, X.; Sun, X.; Zhu, C.; Zheng, B.; Zou, Z.; Luo, Z.; Zhu, X.; et al. Van der Waals epitaxial growth of vertically stacked Sb2Te3/MoS2 p–n heterojunctions for high performance optoelectronics. Nano Energy 2019, 59, 66–74. [Google Scholar] [CrossRef]
  124. Azizi, A.; Eichfeld, S.; Geschwind, G.; Zhang, K.; Jiang, B.; Mukherjee, D.; Hossain, L.; Piasecki, A.F.; Kabius, B.; Robinson, J.A.; et al. Freestanding van der Waals Heterostructures of Graphene and Transition Metal Dichalcogenides. ACS Nano 2015, 9, 4882–4890. [Google Scholar] [CrossRef]
  125. Xu, Z.; Khanaki, A.; Tian, H.; Zheng, R.; Suja, M.; Zheng, J.-G.; Liu, J. Direct growth of hexagonal boron nitride/graphene heterostructures on cobalt foil substrates by plasma-assisted molecular beam epitaxy. Appl. Phys. Lett. 2016, 109, 043110. [Google Scholar] [CrossRef] [Green Version]
  126. Aretouli, K.E.; Tsoutsou, D.; Tsipas, P.; Marquez-Velasco, J.; Aminalragia Giamini, S.; Kelaidis, N.; Psycharis, V.; Dimoulas, A. Epitaxial 2D SnSe2/2D WSe2 van der Waals Heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 23222–23229. [Google Scholar] [CrossRef]
  127. Kozhakhmetov, A.; Torsi, R.; Chen, C.Y.; Robinson, J.A. Scalable low-temperature synthesis of two-dimensional materials beyond graphene. J. Phys. Mater. 2020, 4, 012001. [Google Scholar] [CrossRef]
  128. Marx, M.; Nordmann, S.; Knoch, J.; Franzen, C.; Stampfer, C.; Andrzejewski, D.; Kümmell, T.; Bacher, G.; Heuken, M.; Kalisch, H.; et al. Large-area MoS2 deposition via MOVPE. J. Cryst. Growth 2017, 464, 100–104. [Google Scholar] [CrossRef]
  129. Kranthi Kumar, V.; Dhar, S.; Choudhury, T.H.; Shivashankar, S.A.; Raghavan, S. A predictive approach to CVD of crystalline layers of TMDs: The case of MoS2. Nanoscale 2015, 7, 7802–7810. [Google Scholar] [CrossRef]
  130. Kakanakova-Georgieva, A.; Giannazzo, F.; Nicotra, G.; Cora, I.; Gueorguiev, G.K.; Persson, P.O.Å.; Pécz, B. Material proposal for 2D indium oxide. Appl. Surf. Sci. 2021, 548, 149275. [Google Scholar] [CrossRef]
  131. Tsoutsou, D.; Aretouli, K.E.; Tsipas, P.; Marquez-Velasco, J.; Xenogiannopoulou, E.; Kelaidis, N.; Aminalragia Giamini, S.; Dimoulas, A. Epitaxial 2D MoSe2 (HfSe2) Semiconductor/2D TaSe2 Metal van der Waals Heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 1836–1841. [Google Scholar] [CrossRef]
  132. Gong, Y.; Lin, J.; Wang, X.; Shi, G.; Lei, S.; Lin, Z.; Zou, X.; Ye, G.; Vajtai, R.; Yakobson, B.I.; et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat Mater 2014, 13, 1135–1142. [Google Scholar] [CrossRef] [Green Version]
  133. Guo, Q.; Pospischil, A.; Bhuiyan, M.; Jiang, H.; Tian, H.; Farmer, D.; Deng, B.; Li, C.; Han, S.-J.; Wang, H.; et al. Black Phosphorus Mid-Infrared Photodetectors with High Gain. Nano Lett. 2016, 16, 4648–4655. [Google Scholar] [CrossRef] [Green Version]
  134. Jariwala, D.; Marks, T.J.; Hersam, M.C. Mixed-dimensional van der Waals heterostructures. Nat Mater 2017, 16, 170–181. [Google Scholar] [CrossRef] [Green Version]
  135. Deng, J.; Guo, Z.; Zhang, Y.; Cao, X.; Zhang, S.; Sheng, Y.; Xu, H.; Bao, W.; Wan, J. MoS2/Silicon-on-Insulator Heterojunction Field-Effect-Transistor for High-Performance Photodetection. IEEE Electron Device Lett. 2019, 40, 423–426. [Google Scholar] [CrossRef]
  136. Yao, J.; Yang, G. Flexible and High-Performance All-2D Photodetector for Wearable Devices. Small 2018, 14, 1704524. [Google Scholar] [CrossRef]
  137. Zhang, K.; Fang, X.; Wang, Y.; Wan, Y.; Song, Q.; Zhai, W.; Li, Y.; Ran, G.; Ye, Y.; Dai, L. Ultrasensitive Near-Infrared Photodetectors Based on a Graphene–MoTe2–Graphene Vertical van der Waals Heterostructure. ACS Appl. Mater. Interfaces 2017, 9, 5392–5398. [Google Scholar] [CrossRef]
  138. Massicotte, M.; Schmidt, P.; Vialla, F.; Schädler, K.G.; Reserbat-Plantey, A.; Watanabe, K.; Taniguchi, T.; Tielrooij, K.J.; Koppens, F.H.L. Picosecond photoresponse in van der Waals heterostructures. Nat. Nanotechnol. 2016, 11, 42–46. [Google Scholar] [CrossRef]
  139. Shehzad, K.; Shi, T.; Qadir, A.; Wan, X.; Guo, H.; Ali, A.; Xuan, W.; Xu, H.; Gu, Z.; Peng, X.; et al. Designing an Efficient Multimode Environmental Sensor Based on Graphene–Silicon Heterojunction. Adv. Mater. Technol. 2017, 2, 1600262. [Google Scholar] [CrossRef] [Green Version]
  140. Chen, H.; Wen, X.; Zhang, J.; Wu, T.; Gong, Y.; Zhang, X.; Yuan, J.; Yi, C.; Lou, J.; Ajayan, P.M.; et al. Ultrafast formation of interlayer hot excitons in atomically thin MoS2/WS2 heterostructures. Nat. Commun. 2016, 7, 12512. [Google Scholar] [CrossRef]
  141. Bullock, J.; Amani, M.; Cho, J.; Chen, Y.-Z.; Ahn, G.H.; Adinolfi, V.; Shrestha, V.R.; Gao, Y.; Crozier, K.B.; Chueh, Y.-L.; et al. Polarization-resolved black phosphorus/molybdenum disulfide mid-wave infrared photodiodes with high detectivity at room temperature. Nat. Photonics 2018, 12, 601–607. [Google Scholar] [CrossRef] [Green Version]
  142. Yu, W.; Li, S.; Zhang, Y.; Ma, W.; Sun, T.; Yuan, J.; Fu, K.; Bao, Q. Near-Infrared Photodetectors Based on MoTe2/Graphene Heterostructure with High Responsivity and Flexibility. Small 2017, 13, 1700268. [Google Scholar] [CrossRef] [PubMed]
  143. Guo, F.; Song, M.; Wong, M.-C.; Ding, R.; Io, W.F.; Pang, S.-Y.; Jie, W.; Hao, J. Multifunctional Optoelectronic Synapse Based on Ferroelectric Van der Waals Heterostructure for Emulating the Entire Human Visual System. Adv. Funct. Mater. 2022, 32, 2108014. [Google Scholar] [CrossRef]
  144. Kim, D.-H.; Lu, N.; Ma, R.; Kim, Y.-S.; Kim, R.-H.; Wang, S.; Wu, J.; Won, S.M.; Tao, H.; Islam, A.; et al. Epidermal Electronics. Science 2011, 333, 838–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Polat, E.O.; Mercier, G.; Nikitskiy, I.; Puma, E.; Galan, T.; Gupta, S.; Montagut, M.; Piqueras, J.J.; Bouwens, M.; Durduran, T.; et al. Flexible graphene photodetectors for wearable fitness monitoring. Sci. Adv. 2019, 5, eaaw7846. [Google Scholar] [CrossRef] [Green Version]
  146. Konstantatos, G.; Badioli, M.; Gaudreau, L.; Osmond, J.; Bernechea, M.; de Arquer, F.P.G.; Gatti, F.; Koppens, F.H.L. Hybrid graphene–quantum dot phototransistors with ultrahigh gain. Nat. Nanotechnol. 2012, 7, 363–368. [Google Scholar] [CrossRef]
  147. Lien, D.-H.; Kang, J.S.; Amani, M.; Chen, K.; Tosun, M.; Wang, H.-P.; Roy, T.; Eggleston, M.S.; Wu, M.C.; Dubey, M.; et al. Engineering Light Outcoupling in 2D Materials. Nano Lett. 2015, 15, 1356–1361. [Google Scholar] [CrossRef]
  148. Jo, S.; Ubrig, N.; Berger, H.; Kuzmenko, A.B.; Morpurgo, A.F. Mono- and Bilayer WS2 Light-Emitting Transistors. Nano Lett. 2014, 14, 2019–2025. [Google Scholar] [CrossRef] [Green Version]
  149. Sundaram, R.S.; Engel, M.; Lombardo, A.; Krupke, R.; Ferrari, A.C.; Avouris, P.; Steiner, M. Electroluminescence in Single Layer MoS2. Nano Lett. 2013, 13, 1416–1421. [Google Scholar] [CrossRef] [Green Version]
  150. Ross, J.S.; Klement, P.; Jones, A.M.; Ghimire, N.J.; Yan, J.; Mandrus, D.G.; Taniguchi, T.; Watanabe, K.; Kitamura, K.; Yao, W.; et al. Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p–n junctions. Nat. Nanotechnol. 2014, 9, 268–272. [Google Scholar] [CrossRef]
  151. Zhang, Y.J.; Oka, T.; Suzuki, R.; Ye, J.T.; Iwasa, Y. Electrically Switchable Chiral Light-Emitting Transistor. Science 2014, 344, 725–728. [Google Scholar] [CrossRef]
  152. Cheng, R.; Li, D.; Zhou, H.; Wang, C.; Yin, A.; Jiang, S.; Liu, Y.; Chen, Y.; Huang, Y.; Duan, X. Electroluminescence and Photocurrent Generation from Atomically Sharp WSe2/MoS2 Heterojunction p–n Diodes. Nano Lett. 2014, 14, 5590–5597. [Google Scholar] [CrossRef] [Green Version]
  153. Lee, C.-H.; Lee, G.-H.; van der Zande, A.M.; Chen, W.; Li, Y.; Han, M.; Cui, X.; Arefe, G.; Nuckolls, C.; Heinz, T.F.; et al. Atomically thin p–n junctions with van der Waals heterointerfaces. Nat. Nanotechnol. 2014, 9, 676–681. [Google Scholar] [CrossRef] [Green Version]
  154. Binder, J.; Withers, F.; Molas, M.R.; Faugeras, C.; Nogajewski, K.; Watanabe, K.; Taniguchi, T.; Kozikov, A.; Geim, A.K.; Novoselov, K.S.; et al. Sub-bandgap Voltage Electroluminescence and Magneto-oscillations in a WSe2 Light-Emitting van der Waals Heterostructure. Nano Lett. 2017, 17, 1425–1430. [Google Scholar] [CrossRef] [Green Version]
  155. Liu, C.-H.; Clark, G.; Fryett, T.; Wu, S.; Zheng, J.; Hatami, F.; Xu, X.; Majumdar, A. Nanocavity Integrated van der Waals Heterostructure Light-Emitting Tunneling Diode. Nano Lett. 2017, 17, 200–205. [Google Scholar] [CrossRef]
  156. Wang, S.; Wang, J.; Zhao, W.; Giustiniano, F.; Chu, L.; Verzhbitskiy, I.; Zhou Yong, J.; Eda, G. Efficient Carrier-to-Exciton Conversion in Field Emission Tunnel Diodes Based on MIS-Type van der Waals Heterostack. Nano Lett. 2017, 17, 5156–5162. [Google Scholar] [CrossRef]
  157. Withers, F.; Del Pozo-Zamudio, O.; Mishchenko, A.; Rooney, A.P.; Gholinia, A.; Watanabe, K.; Taniguchi, T.; Haigh, S.J.; Geim, A.K.; Tartakovskii, A.I.; et al. Light-emitting diodes by band-structure engineering in van der Waals heterostructures. Nat. Mater. 2015, 14, 301–306. [Google Scholar] [CrossRef]
  158. Nikam, R.D.; Sonawane, P.A.; Sankar, R.; Chen, Y.-T. Epitaxial growth of vertically stacked p-MoS2/n-MoS2 heterostructures by chemical vapor deposition for light emitting devices. Nano Energy 2017, 32, 454–462. [Google Scholar] [CrossRef]
  159. Dobusch, L.; Schuler, S.; Perebeinos, V.; Mueller, T. Thermal Light Emission from Monolayer MoS2. Adv. Mater. 2017, 29, 1701304. [Google Scholar] [CrossRef] [Green Version]
  160. Ceballos, F.; Bellus, M.Z.; Chiu, H.-Y.; Zhao, H. Ultrafast Charge Separation and Indirect Exciton Formation in a MoS2–MoSe2 van der Waals Heterostructure. ACS Nano 2014, 8, 12717–12724. [Google Scholar] [CrossRef]
  161. Ross, J.S.; Rivera, P.; Schaibley, J.; Lee-Wong, E.; Yu, H.; Taniguchi, T.; Watanabe, K.; Yan, J.; Mandrus, D.; Cobden, D.; et al. Interlayer Exciton Optoelectronics in a 2D Heterostructure p–n Junction. Nano Lett. 2017, 17, 638–643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Clark, G.; Schaibley, J.R.; Ross, J.; Taniguchi, T.; Watanabe, K.; Hendrickson, J.R.; Mou, S.; Yao, W.; Xu, X. Single Defect Light-Emitting Diode in a van der Waals Heterostructure. Nano Lett. 2016, 16, 3944–3948. [Google Scholar] [CrossRef] [PubMed]
  163. Jariwala, D.; Davoyan, A.R.; Wong, J.; Atwater, H.A. Van der Waals Materials for Atomically-Thin Photovoltaics: Promise and Outlook. ACS Photonics 2017, 4, 2962–2970. [Google Scholar] [CrossRef] [Green Version]
  164. Sanchez, O.L.; Ovchinnikov, D.; Misra, S.; Allain, A.; Kis, A. Valley Polarization by Spin Injection in a Light-Emitting van der Waals Heterojunction. Nano Lett. 2016, 16, 5792–5797. [Google Scholar] [CrossRef] [PubMed]
  165. Lopez-Sanchez, O.; Alarcon Llado, E.; Koman, V.; Fontcuberta i Morral, A.; Radenovic, A.; Kis, A. Light Generation and Harvesting in a van der Waals Heterostructure. ACS Nano 2014, 8, 3042–3048. [Google Scholar] [CrossRef]
  166. Pospischil, A.; Furchi, M.M.; Mueller, T. Solar-energy conversion and light emission in an atomic monolayer p–n diode. Nat. Nanotechnol. 2014, 9, 257–261. [Google Scholar] [CrossRef]
  167. Baugher, B.W.H.; Churchill, H.O.H.; Yang, Y.; Jarillo-Herrero, P. Optoelectronic devices based on electrically tunable p–n diodes in a monolayer dichalcogenide. Nat. Nanotechnol. 2014, 9, 262–267. [Google Scholar] [CrossRef] [Green Version]
  168. Liu, Y.; Cai, Y.; Zhang, G.; Zhang, Y.-W.; Ang, K.-W. Al-Doped Black Phosphorus p–n Homojunction Diode for High Performance Photovoltaic. Adv. Funct. Mater. 2017, 27, 1604638. [Google Scholar] [CrossRef]
  169. Yu, W.J.; Liu, Y.; Zhou, H.; Yin, A.; Li, Z.; Huang, Y.; Duan, X. Highly efficient gate-tunable photocurrent generation in vertical heterostructures of layered materials. Nat. Nanotechnol. 2013, 8, 952–958. [Google Scholar] [CrossRef] [Green Version]
  170. Furchi, M.M.; Pospischil, A.; Libisch, F.; Burgdörfer, J.; Mueller, T. Photovoltaic Effect in an Electrically Tunable van der Waals Heterojunction. Nano Lett. 2014, 14, 4785–4791. [Google Scholar] [CrossRef]
  171. Long, M.; Liu, E.; Wang, P.; Gao, A.; Xia, H.; Luo, W.; Wang, B.; Zeng, J.; Fu, Y.; Xu, K.; et al. Broadband Photovoltaic Detectors Based on an Atomically Thin Heterostructure. Nano Lett. 2016, 16, 2254–2259. [Google Scholar] [CrossRef] [Green Version]
  172. Zalalutdinov, M.K.; Robinson, J.T.; Fonseca, J.J.; LaGasse, S.W.; Pandey, T.; Lindsay, L.R.; Reinecke, T.L.; Photiadis, D.M.; Culbertson, J.C.; Cress, C.D.; et al. Acoustic cavities in 2D heterostructures. Nat. Commun. 2021, 12, 3267. [Google Scholar] [CrossRef]
  173. Soubelet, P.; Reynoso, A.A.; Fainstein, A.; Nogajewski, K.; Potemski, M.; Faugeras, C.; Bruchhausen, A.E. The lifetime of interlayer breathing modes of few-layer 2H-MoSe2 membranes. Nanoscale 2019, 11, 10446–10453. [Google Scholar] [CrossRef] [Green Version]
  174. Greener, J.D.G.; Akimov, A.V.; Gusev, V.E.; Kudrynskyi, Z.R.; Beton, P.H.; Kovalyuk, Z.D.; Taniguchi, T.; Watanabe, K.; Kent, A.J.; Patanè, A. Coherent acoustic phonons in van der Waals nanolayers and heterostructures. Phys. Rev. B 2018, 98, 075408. [Google Scholar] [CrossRef] [Green Version]
  175. Chakraborty, S.K.; Kundu, B.; Nayak, B.; Dash, S.P.; Sahoo, P.K. Challenges and opportunities in 2D heterostructures for electronic and optoelectronic devices. iScience 2022, 25, 103942. [Google Scholar] [CrossRef]
Figure 1. Summary of the current status of 2D materials presented to date. Adapted with permission from Ref. [40]. Copyright 2022, American Chemical Society.
Figure 1. Summary of the current status of 2D materials presented to date. Adapted with permission from Ref. [40]. Copyright 2022, American Chemical Society.
Molecules 28 02275 g001
Figure 2. Characteristic approaches executed for research on 2D heterostructures. Adapted with permission from Ref. [40]. Copyright 2022, American Chemical Society.
Figure 2. Characteristic approaches executed for research on 2D heterostructures. Adapted with permission from Ref. [40]. Copyright 2022, American Chemical Society.
Molecules 28 02275 g002
Figure 4. (a) Optical microscope images for the hetero-bilayer (WSe2/MoS2) over the Si/SiO2 substrate. (b) Normalized PL. (c) Energy-band diagram of photoexcited hetero-bilayer WSe2/MoS2. Adapted with permission from Ref. [91]. Copyright 2014, National Academy of Sciences of the United States of America. (d) Characteristic WSe2/MoTe2 samples with monolayer MoTe2 and bilayer WSe2, monolayer WSe2, and with their heterostructures for monolayer MoTe2. (e) Transient reflectance (TR) spectra of WSe2 and exciton bleach-recovery kinetics in WSe2/MoTe2 heterostructures at several temperatures. (f) Displays the near-unity transfer of WSe2 excitons (dark and bright) into free electrons and holes in the MoTe2 monolayer. Adapted with permission from Ref. [95]. Copyright 2019, American Chemical Society. (g) Microscopic images of MoS2/h-BN heterostructure. (h) Energy-band mapping of MoS2/h-BN heterostructure (where IX = indirect exciton; DX = direct exciton). Adapted with permission from Ref. [94]. Copyright 2015, Nature Publishing Group.
Figure 4. (a) Optical microscope images for the hetero-bilayer (WSe2/MoS2) over the Si/SiO2 substrate. (b) Normalized PL. (c) Energy-band diagram of photoexcited hetero-bilayer WSe2/MoS2. Adapted with permission from Ref. [91]. Copyright 2014, National Academy of Sciences of the United States of America. (d) Characteristic WSe2/MoTe2 samples with monolayer MoTe2 and bilayer WSe2, monolayer WSe2, and with their heterostructures for monolayer MoTe2. (e) Transient reflectance (TR) spectra of WSe2 and exciton bleach-recovery kinetics in WSe2/MoTe2 heterostructures at several temperatures. (f) Displays the near-unity transfer of WSe2 excitons (dark and bright) into free electrons and holes in the MoTe2 monolayer. Adapted with permission from Ref. [95]. Copyright 2019, American Chemical Society. (g) Microscopic images of MoS2/h-BN heterostructure. (h) Energy-band mapping of MoS2/h-BN heterostructure (where IX = indirect exciton; DX = direct exciton). Adapted with permission from Ref. [94]. Copyright 2015, Nature Publishing Group.
Molecules 28 02275 g004
Figure 6. (a) Representation of front and lateral views of MoS2/WSe2 heterostructures at varying twist angles. (b) Photoexcited excitonic band alignment of MoSe2/WSe2 heterostructures. (c) Interlayer exciton peak intensity as a function of twist angle. Adapted with permission from Ref. [108]. Copyright 2017, American Chemical Society. (d) Atomistic representation of heterostructures based on MoS2/WS2. (e) PL spectrum of isolated MoS2, WS2, and heterostructures based on MoS2/WS2. (f) 2D plots of transient absorption spectra from a MoS2/WS2 heterostructure at 77 K(top); and an isolated MoS2 monolayer (down) upon excitation of the MoS2 A-exciton transitions. Probe photon energy, probe-pump time delay, and transient absorption signal are shown by a horizontal axis, a vertical axis, and a color scale, respectively. Whereas the pump-induced decrease in absorption is indicated by positive signals. Adapted with permission from Ref. [92]. Copyright 2014, Nature Publishing Group.
Figure 6. (a) Representation of front and lateral views of MoS2/WSe2 heterostructures at varying twist angles. (b) Photoexcited excitonic band alignment of MoSe2/WSe2 heterostructures. (c) Interlayer exciton peak intensity as a function of twist angle. Adapted with permission from Ref. [108]. Copyright 2017, American Chemical Society. (d) Atomistic representation of heterostructures based on MoS2/WS2. (e) PL spectrum of isolated MoS2, WS2, and heterostructures based on MoS2/WS2. (f) 2D plots of transient absorption spectra from a MoS2/WS2 heterostructure at 77 K(top); and an isolated MoS2 monolayer (down) upon excitation of the MoS2 A-exciton transitions. Probe photon energy, probe-pump time delay, and transient absorption signal are shown by a horizontal axis, a vertical axis, and a color scale, respectively. Whereas the pump-induced decrease in absorption is indicated by positive signals. Adapted with permission from Ref. [92]. Copyright 2014, Nature Publishing Group.
Molecules 28 02275 g006
Figure 7. (a) Graphic representation of transfer approach utilized to produce devices using graphene on h-BN. (b) Optical photograph for graphene on h-BN (after performing the transfer). The inset shows electrical connections. Adapted with permission from Ref. [70]. Copyright 2010, Nature Publishing Group. (c) Schematics of pick-up and drop-down flow process for assembly of 2D heterostructures showing that flakes can be dropped-down and picked-up at the preferable positions. (d) Optical microscopic photograph of a staked graphene sandwiched among h-BN flakes after the procedure shown in (c). (e) The AFM of the stacks show no optically visible blisters of contaminations. Adapted with permission from Ref. [110]. Copyright 2016, Nature Publishing Group.
Figure 7. (a) Graphic representation of transfer approach utilized to produce devices using graphene on h-BN. (b) Optical photograph for graphene on h-BN (after performing the transfer). The inset shows electrical connections. Adapted with permission from Ref. [70]. Copyright 2010, Nature Publishing Group. (c) Schematics of pick-up and drop-down flow process for assembly of 2D heterostructures showing that flakes can be dropped-down and picked-up at the preferable positions. (d) Optical microscopic photograph of a staked graphene sandwiched among h-BN flakes after the procedure shown in (c). (e) The AFM of the stacks show no optically visible blisters of contaminations. Adapted with permission from Ref. [110]. Copyright 2016, Nature Publishing Group.
Molecules 28 02275 g007
Figure 10. (a) Illustration of transparent and flexible GQD photodetector implanted in an HR tracking wristband. (b) A zoomed-in view of stretchable photodetector on polyethylene naphthalate substrate with a 1 mm2 graphene channel entirely encased in a thin 30 nm-thick coating of PbS QDs. (c) Schematic representation of graphene and QD setup on a flexible substrate assembly. Photodetection in graphene/PbS QD heterostructure-based health patches for wearable fitness via HR and RR monitoring: (d,e) a mobile-phone screen with an adaptable health patch. (f) Signals from photoplethysmography (PPG) at 630 and 940 nm. (g) Two simultaneous readings of a health patch and a cutting-edge PPG sensor. (h) Bland–Altman assessment of the patch (i) Fourier transforms of PPG’s. (j) Bland–Altman evaluation of the obtained RR. Adapted with permission from Ref. [145]. Copyright 2019, AAAS.
Figure 10. (a) Illustration of transparent and flexible GQD photodetector implanted in an HR tracking wristband. (b) A zoomed-in view of stretchable photodetector on polyethylene naphthalate substrate with a 1 mm2 graphene channel entirely encased in a thin 30 nm-thick coating of PbS QDs. (c) Schematic representation of graphene and QD setup on a flexible substrate assembly. Photodetection in graphene/PbS QD heterostructure-based health patches for wearable fitness via HR and RR monitoring: (d,e) a mobile-phone screen with an adaptable health patch. (f) Signals from photoplethysmography (PPG) at 630 and 940 nm. (g) Two simultaneous readings of a health patch and a cutting-edge PPG sensor. (h) Bland–Altman assessment of the patch (i) Fourier transforms of PPG’s. (j) Bland–Altman evaluation of the obtained RR. Adapted with permission from Ref. [145]. Copyright 2019, AAAS.
Molecules 28 02275 g010
Figure 11. (a) Graphical illustration of heterostructure based on WSe2/MoS2. (b) EL spectra (false color) of device using 100 μA injection current. Adapted with permission from Ref. [152]. Copyright 2014, American Chemical Society. (c) Atomistic representation of h-BN/graphene/2h-BN/WS2/2h-BN/graphene/h-BN single-quantum wall heterostructure and optical view of EL using the alike device. Adapted with permission from Ref. [157]. Copyright 2015, Nature Publishing Group. (d) Representation of assembly for heterostructure. (e) Visualization of PL and EL spectra of heterostructure (inset: display of carrier transportation). Adapted with permission from Ref. [162]. Copyright 2017, American Chemical Society. (f) Demonstration of working of LED composed of a vertical heterostructure. Adapted with permission from Ref. [162]. Copyright 2016, American Chemical Society. (g) Schematics depiction of the proposed device. Adapted with permission from Ref. [164]. Copyright 2016, American Chemical Society.
Figure 11. (a) Graphical illustration of heterostructure based on WSe2/MoS2. (b) EL spectra (false color) of device using 100 μA injection current. Adapted with permission from Ref. [152]. Copyright 2014, American Chemical Society. (c) Atomistic representation of h-BN/graphene/2h-BN/WS2/2h-BN/graphene/h-BN single-quantum wall heterostructure and optical view of EL using the alike device. Adapted with permission from Ref. [157]. Copyright 2015, Nature Publishing Group. (d) Representation of assembly for heterostructure. (e) Visualization of PL and EL spectra of heterostructure (inset: display of carrier transportation). Adapted with permission from Ref. [162]. Copyright 2017, American Chemical Society. (f) Demonstration of working of LED composed of a vertical heterostructure. Adapted with permission from Ref. [162]. Copyright 2016, American Chemical Society. (g) Schematics depiction of the proposed device. Adapted with permission from Ref. [164]. Copyright 2016, American Chemical Society.
Molecules 28 02275 g011
Figure 12. (a) Multilayer MoS2 sandwiched among graphene electrodes (edge view). (b) The graphene layer at the bottom graphene (red), middle (blue), and top MoS2 layer (yellow) are each falsely colored in the SEM photograph of the device. (c) I–V behavior of the vertical device. Adapted with permission from Ref. [169]. Copyright 2021, Nature Publishing Group. (d) Representation of graphene sandwiched between an MoS2/WSe2 junction. (e) Photocurrent map for the corresponding junction (Vds = 0 V) and an optical photograph of a single-layer p–n junction device (1L–1L) sandwiched between a graphene electrode. (f) I–V curves of the device publicized in the figure restrained in the dark (black) and under 532 nm laser excitation (red). Adapted with permission from Ref. [153]. Copyright 2014, Nature Publishing Group. (g) Illustration of photodetector based on p–g–n heterostructure. (h) Optical microscopic view of the device. (i) Photocurrent mapping from the narrow p–g–n junction at Vds = 0 V. Adapted with permission from Ref. [171]. Copyright 2016, American Chemical Society.
Figure 12. (a) Multilayer MoS2 sandwiched among graphene electrodes (edge view). (b) The graphene layer at the bottom graphene (red), middle (blue), and top MoS2 layer (yellow) are each falsely colored in the SEM photograph of the device. (c) I–V behavior of the vertical device. Adapted with permission from Ref. [169]. Copyright 2021, Nature Publishing Group. (d) Representation of graphene sandwiched between an MoS2/WSe2 junction. (e) Photocurrent map for the corresponding junction (Vds = 0 V) and an optical photograph of a single-layer p–n junction device (1L–1L) sandwiched between a graphene electrode. (f) I–V curves of the device publicized in the figure restrained in the dark (black) and under 532 nm laser excitation (red). Adapted with permission from Ref. [153]. Copyright 2014, Nature Publishing Group. (g) Illustration of photodetector based on p–g–n heterostructure. (h) Optical microscopic view of the device. (i) Photocurrent mapping from the narrow p–g–n junction at Vds = 0 V. Adapted with permission from Ref. [171]. Copyright 2016, American Chemical Society.
Molecules 28 02275 g012
Figure 13. 2D heterostructure-based acoustic cavities: (a) Illustration of MoS2/h-BN heterostructure-based acoustic cavities. (b) Finite-element modeling simulation of the normalized mechanical strain of suspended MoS2. (c) Time-varying reflectance of suspended MoS2. (d) Time-dependent reflectivity spectra from various MoS2 cavities calculated via FFT. (e) The quantity of MoS2 determines the frequency of various MoS2 acoustic cavities. Adapted with permission from Ref. [172]. Copyright 2021, Nature Publishing Group.
Figure 13. 2D heterostructure-based acoustic cavities: (a) Illustration of MoS2/h-BN heterostructure-based acoustic cavities. (b) Finite-element modeling simulation of the normalized mechanical strain of suspended MoS2. (c) Time-varying reflectance of suspended MoS2. (d) Time-dependent reflectivity spectra from various MoS2 cavities calculated via FFT. (e) The quantity of MoS2 determines the frequency of various MoS2 acoustic cavities. Adapted with permission from Ref. [172]. Copyright 2021, Nature Publishing Group.
Molecules 28 02275 g013
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Babar, Z.U.D.; Raza, A.; Cassinese, A.; Iannotti, V. Two Dimensional Heterostructures for Optoelectronics: Current Status and Future Perspective. Molecules 2023, 28, 2275. https://doi.org/10.3390/molecules28052275

AMA Style

Babar ZUD, Raza A, Cassinese A, Iannotti V. Two Dimensional Heterostructures for Optoelectronics: Current Status and Future Perspective. Molecules. 2023; 28(5):2275. https://doi.org/10.3390/molecules28052275

Chicago/Turabian Style

Babar, Zaheer Ud Din, Ali Raza, Antonio Cassinese, and Vincenzo Iannotti. 2023. "Two Dimensional Heterostructures for Optoelectronics: Current Status and Future Perspective" Molecules 28, no. 5: 2275. https://doi.org/10.3390/molecules28052275

Article Metrics

Back to TopTop