Next Article in Journal
GCN–Informer: A Novel Framework for Mid-Term Photovoltaic Power Forecasting
Previous Article in Journal
Special Issue: Artificial Intelligence Technology in Medical Image Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Excitonic Evolution in WS2/MoS2 van der Waals Heterostructures Turned by Out-of-Plane Localized Pressure

1
State Key Laboratory of Quantum Optics and Quantum Optics Devices, Institute of Laser Spectroscopy, School of Physics and Electronic Engineering, Shanxi University, Taiyuan 030006, China
2
Xinzhou Institute of Innovation Ecosystem, Shanxi University, Xinzhou 034000, China
3
Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan 030006, China
*
Authors to whom correspondence should be addressed.
Submission received: 29 January 2024 / Revised: 28 February 2024 / Accepted: 3 March 2024 / Published: 5 March 2024
(This article belongs to the Topic Optical and Optoelectronic Materials and Applications)

Abstract

:
In this study, we explore the exciton dynamics in a WS2/MoS2 van der Waals (vdW) heterostructure under varying pressures by integrating a laser-confocal photoluminescence (PL) spectroscope and an atomic force microscope (AFM). For the WS2/MoS2 heterostructure, the exciton emission belonging to MoS2 is too weak to be distinguished from the PL spectra. However, upon contact with a Si probe, the emission intensity of WS2 excitons significantly decreases from 34,234 to 6560, thereby matching the intensity level of MoS2. This alteration substantially facilitates the exploration of interlayer excitonic properties within the heterostructures using PL spectroscopy. Furthermore, the Si probe can apply out-of-plane localized pressure to the heterostructure. With increasing pressure, the emission intensity of the WS2 trions decreases at a rate twice that of other excitons, and the exciton energy increases at a rate of 0.1 meV nN−1. These results elucidate that the WS2 trions are particularly sensitive to the out-of-plane pressure within a WS2/MoS2 vdW heterostructure.

1. Introduction

Layered two-dimensional (2D) materials have attracted considerable attention due to their excellent properties: unique electronic structure, tunable bandgaps, atomic-level thickness, and van der Waals (vdW) integration capability. Compared to the direct synthesis method, the mechanical transfer method for manufacturing vdW heterostructures allows for free stacking of 2D layers at any angle without restrictions from covalent bonds and lattice matching [1,2,3]. Through the mechanical transfer method, different 2D materials can be combined into heterogeneous structures without any limitations in epitaxy, which greatly facilitates the assembly of high-quality vdW heterostructures and expands the exploration of their optical properties [4,5,6]. vdW heterostructures have promoted the development of condensed physics, such as confirming Hofstadter butterfly topologies [7,8,9], finding thermodynamic evidence for fractional Chern insulators [10] and gate-tunable Mott insulators [11,12], and achieving unconventional superconductivity in magic-angle superlattices [13,14]. Transition metal dichalcogenides (TMDCs) exhibit strong spin–orbit coupling (SOC) and high exciton binding energy, making them excellent candidates for building a vdW heterostructure platform to study novel physics. As representative TMDCs, monolayer tungsten disulfide (ML-WS2) and molybdenum disulfide (ML-MoS2) have emerged as promising candidates for constructing heterostructures due to their unconventional properties, including the conversion of indirect/direct bandgaps [15,16,17,18,19], valleytronics [20,21,22,23], and the large binding energies of excitons [18,24], positioning them as ideal platforms for exploring the quantum effect in condensed matter systems. The interlayer interaction is one of the most important properties of MoS2/WS2 heterostructures [25,26,27]. Wang et al. experimentally observed that the photoexcited holes transferring from MoS2 to WS2 take place within an ultra-short time: 50 fs [25]. Lei et al. reported on the interlayer hole-transferring process and the formation of spatially interlayer excitons with longer lifetimes in MoS2/WS2 heterostructures [26]. Yu et al. determined that Raman scattering and their underlying interlayer interactions depend on the twist angle in twisted MoS2/WS2 heterostructures [27]. Understanding the interlayer interactions in vdW heterostructures under different conditions is crucial for optimizing their performance.
The precise tuning of bandgaps, photoluminescence energy, and quantum yield at the nanoscale greatly expands the application prospects of two-dimensional materials in micro-optoelectronic devices. Various methods have been reported for tuning the electronic structure of 2D materials, including electric field modulation [28,29], chemical modification [30,31], magnetic field modulation [32], and mechanical strain [15,33,34]. It is noteworthy that the strong planar mechanical strength and small bending stiffness of 2D materials, coupled with their reversibility compared to other tuning methods, make them particularly suitable for strain engineering [35,36]. For example, Castellanos-Gomez analyzed the strain tunability of the exciton properties of monolayer and few-layer MoS2, MoSe2, WS2, and WSe2 by bending the substrate [37]. Steele et al. employed exciton imaging to investigate the band structure of a monolayer MoS2 under macroscopic strain and observed a strain-induced bandgap reduction and exciton funneling phenomenon [38]. Although strain engineering has been extendedly explored in TMDC monolayers, multilayers, and heterostructures, most studies primarily focus on macroscopic strain, while single-point pressure is rarely studied. Due to the anisotropy of monolayer TMDCs, the regulation of excitons by macroscopic in-plane strain and out-of-plane localized strain is completely different. Therefore, a quantitative investigation into the single-point tuning of vdW heterostructures is necessary to study the exciton effect.
In this work, we develop a quantitative analysis of the excitonic evolution features in WS2/MoS2 vdW heterostructures under controlled pressure conditions. ML-WS2 and ML-MoS2 samples are synthesized through chemical vapor transport (CVT) and the chemical vapor deposition (CVD) method, respectively, and the WS2/MoS2 heterostructure is fabricated using a wet transfer technique, enabling precise control of the heterostructure composition and layer stacking. We integrate a spectroscopy measurement apparatus with an atomic force microscope (AFM) to accurately control and monitor the applied pressure. The observations reveal a diminution in PL intensity concomitant with an applied pressure across both the monolayer and ensemble heterostructure. Notably, within the heterostructure, a blue shift in the excitonic peak position under pressure articulates a modifiable electronic structure, suggesting that the trions are more sensitive than the natural excitons to the dielectric environment. This quantitative pressure tuning of the excitonic evolution offers a novel approach for modulating the electronic structure of vdW heterostructures and expands their potential applications in nanoscale light sources, nanophotonic switches, and nano-integrated excitonic circuits.

2. Materials and Methods

2.1. Synthesis of ML-WS2 and ML-MoS2

ML-MoS2 and ML-WS2 were both synthesized in an open-type double-temperature-zone tube furnace (MTI OTF-1200X-II, HF-Kejing, Hefei, China). Prior to the synthetic procedure, a SiO2/Si wafer (SiO2 thickness is 300 nm) with dimensions of 20 mm × 20 mm was subjected to ultrasonic cleaning in ethanol, isopropanol, and deionized water for 15 min each in sequence. Afterward, the SiO2/Si wafer was dried with N2 gas flow and served as the substrate for the growth of the 2D materials. For the synthesis of ML-MoS2, 0.5 g of S powder and 10 mg of MoO3 powder were placed in Zone I and Zone II, respectively. The distances between the S and MoO3 were 30 cm, and the as-cleaned SiO2/Si wafer was placed upside down on the quartz boat with MoO3 powder. The system was evacuated to 10 Pa and then backfilled with N2 to atmospheric pressure at a flow rate of 70 sccm. Zones I and II were separately heated to 350 °C and 760 °C at a rate of 10 °C min−1 and maintained at stable temperatures for 20 min to facilitate growth. Then, the system underwent natural cooling to room temperature.
To synthesize ML-WS2, a growth source of 20 mg of WS2 powder (99.9% metallic base) was positioned near the central section of the tube. The pre-cleaned SiO2/Si wafer was then placed 2.2 cm behind the WS2 powder. The tube was sealed and first evacuated to a pressure of 10 Pa, followed by the introduction of Ar gas until atmospheric pressure was reached. This process was repeated three times to ensure the complete removal of air. Zone I was heated to 1000 °C at a rate of 10 °C min−1 and kept for 1 h under the protection of 100 sccm of Ar gas flow to grow large ML-WS2 triangles. Finally, the tube furnace was naturally cooled to room temperature to prevent the triangle from breaking.

2.2. Fabrication of WS2/MoS2 vdW Heterostructure

WS2/MoS2 HS was fabricated via the wet transfer method in a multifunctional 2D material transfer platform (E1-T, METATEST, Nanjing, China). Firstly, the SiO2/Si wafer grown with ML-WS2 triangles was spin-coated with a 6% PMMA-phenyl ether solution. The spin time and spin rate were set to first spin at 600 rpm for 10 s to completely cover the wafer and then spin at 3000 rpm for 60 s to form a PMMA thin film on the surface. The PMMA-coated wafer was then heated on a hot plate at 110 °C for 15 min to tighten the fitness between the ML-WS2 and PMMA film, followed by immersion in a 2 M NaOH solution at 120 °C for 30 min to etch the SiO2 layer. Secondly, the PMMA/WS2 film floating on the surface of the NaOH solutions was picked up with a clean slide and washed in deionized water several times to remove the residual SiO2. The cleaned PMMA/WS2 film was carefully transferred to the upper movable sample holder on the 2D material transfer platform with the WS2 side facing downward, while the MoS2-grown SiO2/Si wafer was fixed on the lower heating sample holder with the WS2 side facing upward. The five-axis positioning stage was adjusted to align the PMMA/WS2 parallel to the MoS2/SiO2/Si and the movable sample holder was pushed down vertically until the film was fully attached to the wafer, ensuring that there were no visible bubbles. The lower heating sample holder was heated at 100 °C for 30 min to reinforce the adherence between the WS2 and MoS2. Finally, the wafer was immersed in acetone, IPA, and deionized water for 10 min each to remove the PMMA and dried with N2 gas flow at room temperature to obtain WS2/MoS2 on the SiO2/Si wafer.

2.3. Measurement

Optical images, Raman spectra, and PL spectra of the samples were acquired using a homemade laser scanning confocal PL-Raman spectrometer (LSCPRS). A laser with a wavelength of 532 nm and a power of 6 mW was utilized as the excitation source. The spot diameter was focused to ~1 µm using a long work-distance M plan fluorite semi-apochromat infinity-corrected objective (LMPLFLN 50X, Olympus, Tokyo, Japan). The exposure times were 3 and 10 s for the PL and Raman spectra, respectively. The emission intensity was recorded with a thermoelectrically cooled charge coupled device detector at −60 °C, and the wavelength was distinguished by 300 and 1800 lines per mm diffraction gratings for the PL and Raman spectra, respectively. The morphological features and height of the samples were investigated through atomic force microscopy (Ntegra, NT-MDT, Moscow, Russia) equipped with a Si probe (VIT_P/TERS, 50 N/m, NT-MDT & TipsNano, Tallinn, Estonia). The actual force constant was measured to be 45.5582 N m−1. Pressure was applied to the nanoscale contact area on the sample stage through a Si tip loaded onto the AFM instrument. The pressure force (F) was calculated using F = setpoint × 0.064 nN. The focusing position of the microscope was adjusted with reference to the tip to enable optical property measurements under different pressure conditions in contact mode.

2.4. DFT Calculation

The band structure of the WS2/MOS2 HS was calculated with the DFT method using VASP 5.1. For the modeling of WS2/MOS2 HS, the AB-stack structure of ML-WS2 and ML-MoS2 with aligned metal atoms was chosen to construct a supercell including 6 atoms. The initial interlayer distance between the two monolayers was set to 3.3 Å and optimized to 3.087 Å. The electron–ion interaction was described using ultra-soft pseudopotential plane waves, and the interaction correlation energy was treated using a generalized gradient approximation (GGA) with the Periodic–Boundary–Embedding (PBE) functional form. To prevent the interaction between the supercell layers caused by the periodic boundary condition, a vacuum layer with a thickness of 15 Å was added in the c direction. During the optimization and electron calculation processes, the internal interatomic interaction force and pressure thresholds were set to 0.03 eV Å−1 and 0.05 GPa, respectively. The k-point grid of 3 × 3 × 1, the energy convergence criterion of 10−5 eV, and the cutoff energy of 650 eV were optimized for use.

3. Results and Discussion

Figure 1a,b show the detailed synthetic procedures of ML-MoS2 and ML-WS2, respectively. Unlike the CVD synthesis method for ML-MoS2, ML-WS2 was prepared using the CVT method for easier peeling from the substrate [39]. Then, the formation of the WS2/MoS2 heterostructure (WS2/MoS2 HS) was achieved by transferring ML-WS2 onto ML-MoS2 using a wet transfer technique, as illustrated in Figure 2. Typically, the formation of two-dimensional vdW heterostructures is confirmed by integrating the results from optical imaging, AFM imaging, Raman spectra, and PL spectroscopy [25,40,41].
Figure 2a is an optical image of the WS2/MoS2 HS on the SiO2/Si substrate, displaying twist angles for WS2 and MoS2 at approximately 60°. The AFM topography image is shown in Figure 2b, showing that numerous abnormal points with random heights are unevenly distributed throughout the stacking region. The abnormal points are caused by the formation of unavoidable nanobubbles between the two monolayers during the fabrication process, and the gap distances are approximately equal to the diameter of the bubbles. Figure 2c is the height profile along the straight line indicated in Figure 2b. The presence of two steps below 1 nm in the height curve indicates that the heterostructure is composed of two monolayers pointing to ML-WS2 and ML-MoS2. The height of the second step (0.68 or 0.71 nm) is slightly higher than that of the first (0.63 nm), which is in agreement with the gap of the bilayer. It is also observed that most of the bubbles are less than 2 nm in diameter, but a small number of bubbles can reach diameters of over 10 nm. The nanobubbles enlarge the interlayer distance between the two monolayers, weakening or even eliminating interlayer interactions [42,43]. Thus, nanobubbles should be avoided in subsequent measurements.
Figure 2d presents the Raman spectra of ML-WS2, ML-MoS2, and WS2/MoS2 HS (the selected points are marked with yellow, purple, and blue crosses in Figure 2b, respectively). The characteristic peaks of WS2 located at 424 and 361 cm−1 are indexed to the out-of-plane A1g mode and in-plane E12g mode of the monolayer WS2 [44]. The out-of-plane A1g mode and in-plane E12g mode of individual MoS2 are located at 410 and 392 cm−1, respectively. It should be noted that the Raman shifts are relatively higher than those in some previous studies [45,46,47,48,49], which is induced by the extensive oxidation reactions occurring between atmospheric oxygen and the defects in MoS2 during long exposure to air [50]. The distance between the two Raman peaks is ~18 cm−1, which is consistent with previous studies and confirms the monolayer nature of MoS2 [51]. The Raman spectrum of the WS2/MoS2 region appears to contain all four peaks of the two monolayers, indicating the coexistence of MoS2 and WS2. Upon stacking, the intensities of the peaks corresponding to the WS2 E12g, MoS2 E12g, and MoS2 A1g modes decrease to 1605, 250, and 270, respectively. In contrast, the peak intensity of the WS2 A1g model increases from 254 to 340. Previous studies have shown that the out-of-plane A1g mode of WS2 is extremely sensitive to the number of layers and its intensity increases with increasing layer number [52,53,54]. The formation of a bilayer heterostructure can generate interlayer vdW force, enhancing the out-of-plane vibration of WS2. Therefore, the enhancement of the A1g mode of WS2 indicates that there is more than one layer of material in the test area. Combined with the morphology and height images characterized by AFM, these results corroborate the formation of the WS2/MoS2 HS.
To investigate the interlayer electronic coupling in the WS2/MoS2 HS, the PL spectra of ML-WS2, ML-MoS2, and WS2/MoS2 HS are depicted in Figure 2e. Fitted by the Lorenz function, the PL spectrum of WS2 exhibits two peaks at 1.972 and 2.003 eV, which can be attributed to the A-excitons and negative trions of ML-WS2. The PL spectrum of MoS2 can also be divided into two peaks at 1.816 and 1.984 eV, corresponding to the A-excitons and B-excitons of ML-MoS2. After the formation of the heterostructure, only the exciton peak from WS2 can be observed, possibly due to the low emission intensity of MoS2, making it indistinguishable in the overlapping region from the PL spectrum. It is worth noting that the energy from trions slightly red-shifts from 1.972 to 1.935 eV while the PL intensity maximum of the heterostructure decreases dramatically from 34,234 (for ML-WS2) to 6560. This abrupt decrease in intensity indicates the formation of a heterojunction between WS2 and MoS2 accompanied by interlayer charge transfer, which can suppress exciton recombination and result in a decreased PL intensity. No distinct peak of interlayer exciton is observed, which could be attributed to the reduced intensity caused by non-parallel stacking of the monolayers [55]. To study the influence of stacking order on the exciton performance, the PL spectrum of MoS2/WS2 HS with MoS2 in the upper layer is also conducted, as shown in Figure 2f. Interestingly, MoS2/WS2 and WS2/MoS2 exhibit similar exciton behavior, revealing that stacking order does not significantly affect the exciton behavior of the heterostructure. Given the absence of the peaks related to the interlayer exciton and MoS2 exciton in the PL spectrum of WS2/MoS2 HS, it becomes extremely challenging to analyze the excitonic properties and charge transfer in the vdW heterostructures.
To gain clearer insights into the variation in the heterostructure’s excitonic characteristics, it is essential to enhance the proportion of exciton emission of MoS2 in the PL spectrum of the heterostructure region. A feasible approach is to reduce the PL intensity of WS2 without affecting the exciton emission of MoS2. Considering that WS2 is the upper layer of the heterostructure, we propose a method for the upper WS2 layer in the heterojunction to contact the tip of the AFM probe and divert the excited electrons in the conduction band of WS2 using the high work function of the Si probe, thus inhibiting the emission caused by exciton recombination in WS2. The experimental setup is illustrated in Figure 3a, which integrates a homemade LSCPRS and an AFM.
To study the influence of the probe contact on the exciton properties, the PL spectra of ML-WS2, ML-MoS2, and WS2/MoS2 HS are measured with and without Si probe contact (Figure 3b). To better compare the energy and shape of the emission peaks, the PL intensities of ML-WS2 and WS2/MoS2 HS without Si probe contact are multiplied by a shrinkage factor of 0.02. It is shown that upon contact with the Si probe, the PL shapes of ML-WS2 and ML-MoS2 remain basically unchanged. The PL intensities slightly decrease because the Si probe inhibits exciton recombination by guiding the escape pathway of the electrons. The peak positions exhibit a blue shift, which can be attributed to the local electric field generated by the contacted Si probe, thereby influencing the band structure and exciton binding energy of WS2 and MoS2 [56]. Excitingly, with Si probe contact, the PL spectrum of the WS2/MoS2 HS not only experiences a substantial decrease in intensity but also undergoes a significant change in shape. The peak of MoS2 A-excitons is highlighted, enabling the detailed study of the excitonic properties in a vdW heterostructure using PL spectroscopy.
Moreover, an AFM equipped with a Si probe can serve as not only an electrode for electron transport but also a means to exert out-of-plane controllable pressure on 2D materials [57]. Previous studies have shown that 2D materials exhibit good reversibility under a pressure of 12 GPa and a strain of 0.08%, and applying pressure to 2D material using an AFM probe does not cause damage [58,59]. To investigate the pressure-related exciton characteristics, the PL spectra of ML-WS2, ML-MoS2, and WS2/MoS2 HS are measured under different applied forces in the range of 0–120 nN at 20 nN intervals, as shown in Figure 4a–c, respectively.
The emission peak intensities of all three regions exhibit a decreasing trend with increasing pressure, which can be related to the out-of-plane pressure-induced transition from the direct to indirect bandgaps of the monolayers, leading to a reduction in quantum yield and a significant weakening of PL intensity [15,60,61]. The spectral behavior in the heterostructure region is more complex than that in the monolayer region, as the PL properties of the heterostructure are influenced not only by the electronic structure of the monolayers but also by interlayer coupling [62]. To visually demonstrate the relationship between the PL spectra and pressure in the heterojunction region, the 2D pressure-dependent wavelength–intensity image is shown in Figure 4d. Three peaks of the heterostructure exhibit a strong pressure dependency. As the pressure increases, the intensity of the three peaks declines with varying degrees, where the peak related to WS2 trions exhibits the most noticeable decrease.
To gain insights into the influence of pressure on the exciton characteristics of the vdW heterostructure, the pressure-dependent PL curves are fitted with the Lorentzian function. The evolution of the intensities and positions of the fitted exciton peaks of ML-WS2, ML-MoS2, and WS2/MoS2 HS with increasing pressure are shown in Figure 5. Figure 5a reveals that with increasing pressure, the intensity of A excitons and trions in ML-WS2 decreases at the same rate but the exciton energy remains almost constant (Figure 5b). For ML-MoS2, the intensity of A excitons decreases much faster than that of B excitons (Figure 5c), and the variation in exciton energy is consistent with that of ML-WS2 (Figure 5d). The decline in PL intensity in the monolayer TMDCs is associated with the transfer of electrons from the K valley to its neighboring valleys during conduction. ML-WS2 and ML-MoS2 are both semiconductors with direct bandgaps. With increasing pressure, the conduction band exhibits an upward displacement at the K valley and a downward displacement at the Γ valley, enhancing the K−K bandgap and reducing the K−Γ bandgap [61]. Consequently, the quantum yield diminishes, resulting in a reduction in PL intensity. Once the pressure surpasses a certain threshold, the optical bandgap undergoes a transition, shifting from direct to indirect.
In the vdW heterostructure, the WS2 A-excitons and the MoS2 A-excitons exhibit similar decreasing trends in peak intensity under increasing pressure (Figure 5e). However, the WS2 trions experience a rapid decline with a reduction rate approximately twice that of the WS2 A-excitons. When the pressure increases to 100 nN, the intensity of WS2 trions is already quenched to a level smaller than that of the WS2 A-excitons. This is because the WS2 trions, composed of two electrons and one hole, are more significantly affected as the electrons transitioning to the conduction band decrease due to pressure-induced band change. In contrast, the WS2 A-excitons, composed of one electron and one hole, are less affected. Furthermore, the energy of WS2 trions in the heterostructure also exhibits an anomalous behavior (Figure 5f). Compared to the stability of the A-excitons of WS2 and MoS2, the trion energy increases at a rate of 0.1 meV nN−1 with increasing pressure. This blue-shift is attributed to the anticrossing effect of the electronic states under pressure, a phenomenon observed in III–V and II–VI semiconductor families [63,64,65,66,67,68]. The more pronounced change in WS2 trions compared to A-excitons is understood as the combined modulation of the electronic structure in the heterostructure region by interlayer coupling and strain induced by pressure.
To investigate the reversibility of pressure modulation on the WS2/MoS2 HS, the applied pressure is gradually decreased from 140 to 0 nN, as shown in Figure 6. All the emission intensities of the three peaks increase with weakening pressure. The emission intensity of the WS2 A excitons exhibits the fastest increasing rate. The emission peak positions of the A excitons from both WS2 and MoS2 remain almost unchanged, while the emission energy of the WS2 trions gradually shifts towards lower energies. Once the pressure weakens to 0 N, the peak shape and intensity of the PL spectrum almost recover to their initial states. These variation trends are opposite to the case of increasing pressure, demonstrating the good reversibility of pressure modulation on the excitonic behavior of the WS2/MoS2 HS.
A possible mechanism is proposed in Figure 7. Figure 7a shows the band structure of WS2/MoS2 HS calculated using the DFT method. At the K-point, both the conduction band and valence band of WS2 are higher than those of MoS2, indicating the formation of a type-II heterostructure. It is worth noting that the energy gap of the K−Γ transition is smaller than that of the K−K transition, suggesting the presence of an indirect bandgap, which may lead to a significant enhancement of non-radiative transitions.
Figure 7b illustrates the charge transfer in the heterostructure after contact with the Si probe. Prior to contact, the excited electrons flow from the WS2 conduction band to the MoS2, while the holes flow in the opposite direction. The spatial separation of electrons and holes leads to the dissociation of excitons, suppressing exciton recombination emission. After contact with the probe, a large number of excited electrons flow from the WS2 conduction band to the Si probe, thereby reducing the fluorescence intensity of WS2 to a similar level as MoS2. This can be attributed to the fact that, when the Si probe contacts the upper WS2 in the heterostructure region, the excited electrons in WS2 will migrate to the Si probe due to the difference in the work function, resulting in a significant dissociation of the exciton species in WS2 and a decrease in PL intensity. However, the lower MoS2 layer is not in contact with the Si probe, and the influence on the PL intensity is minimal. When the excitonic recombination efficiency in WS2 is lowered to be comparable to that in MoS2, the contribution of MoS2 in the heterostructure’s PL spectrum becomes prominent. When pressure is applied to the heterostructure, it causes changes in the coupling strength of the heterostructure. With increasing tip pressure, the energy of the A-excitons remains stable while the energy of the trions blue-shifts, indicating that the trions are more sensitive to the interlayer coupling strength than the A-excitons.

4. Conclusions

In summary, we employed a novel method to quantitatively study the exciton characteristics of WS2/MoS2 vdW heterostructures through tip contact PL spectra. Upon Si probe contact, the emissions of the A-excitons and trions can be observed in the PL spectra at the heterostructure region, specifically highlighting the contribution of MoS2 A-excitons. The Si probe not only acts as an electrode to remove electrons but also exerts out-of-plane pressure on the micro-region of the heterostructure to quantitatively control the electronic structure and interlayer coupling. As pressure on the heterostructure increases, the PL intensity corresponding to WS2 trions significantly decreases while the PL energy exhibits an obvious blue-shift, revealing that the trions are more sensitive to pressure than A-excitons. This work provides a valuable approach for understanding the relationship between interlayer interactions and electronic/optical properties in 2D TMDC vdW heterostructures. In future studies, more work should be conducted to further combine the local pressure induced by an AFM probe with novel physics in the TMDC heterostructures, such as interlayer excitons, moiré superlattices, and exciton condensations.

Author Contributions

Conceptualization, Y.F. and J.M.; methodology, W.K. and Z.R.; software, Z.R. and P.L.; validation, W.K. and J.C.; formal analysis, W.K. and Y.L.; investigation, W.K. and P.C.; data curation, J.W. and W.L.; writing—original draft preparation, W.K. and Y.C.; writing—review and editing, Y.F. and J.M.; visualization, W.K.; supervision, Y.F. and J.M.; project administration, Y.F. and J.M.; funding acquisition, J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program of China (2017YFA0304203); the National Natural Science Foundation of China (62020106014, 62175140, 61901249, 92165106, and 12104276); the PCSIRT (IRT-17R70); the 111 project (D18001); the Program for the Outstanding Innovative Teams of Higher Learning Institutions of Shanxi (OIT); the Applied Basic Research Project of Shanxi Province, China (201901D211191 and 201901D211188); and the collaborative grant by the Russian Foundation for Basic Research and NSF of China (62011530047 and 20-53-53025 in the RFBR classification).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Castellanos-Gomez, A.; Buscema, M.; Molenaar, R.; Singh, V.; Janssen, L.; van der Zant, H.S.J.; Steele, G.A. Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping. 2D Mater. 2014, 1, 011002. [Google Scholar] [CrossRef]
  2. 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]
  3. Pizzocchero, F.; Gammelgaard, L.; Jessen, B.S.; Caridad, J.M.; Wang, L.; Hone, J.; Boggild, 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]
  4. Fan, S.D.; Cao, R.; Wang, L.D.; Gao, S.; Zhang, Y.P.; Yu, X.; Zhang, H. Quantum tunneling in two-dimensional van der Waals heterostructures and devices. Sci. Chin. Mater. 2021, 64, 2359–2387. [Google Scholar] [CrossRef]
  5. 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]
  6. Wu, Y.F.; Zhu, M.Y.; Zhao, R.J.; Liu, X.J.; Zhao, Y.C.; Wei, H.X.; Zhang, J.Y.; Zheng, X.Q.; Shen, J.X.; Huang, H.; et al. The fabrication and physical properties of two-dimensional van der Waals heterostructures. Acta Phys. Sin. 2022, 71, 048502. [Google Scholar] [CrossRef]
  7. Dean, C.R.; Wang, L.; Maher, P.; Forsythe, C.; Ghahari, F.; Gao, Y.; Katoch, J.; Ishigami, M.; Moon, P.; Koshino, M.; et al. Hofstadter’s butterfly and the fractal quantum Hall effect in moire superlattices. Nature 2013, 497, 598–602. [Google Scholar] [CrossRef]
  8. Hunt, B.; Sanchez-Yamagishi, J.D.; Young, A.F.; Yankowitz, M.; LeRoy, B.J.; Watanabe, K.; Taniguchi, T.; Moon, P.; Koshino, M.; Jarillo-Herrero, P.; et al. Massive Dirac Fermions and Hofstadter Butterfly in a van der Waals Heterostructure. Science 2013, 340, 1427–1430. [Google Scholar] [CrossRef]
  9. Ponomarenko, L.A.; Gorbachev, R.V.; Yu, G.L.; Elias, D.C.; Jalil, R.; Patel, A.A.; Mishchenko, A.; Mayorov, A.S.; Woods, C.R.; Wallbank, J.R.; et al. Cloning of Dirac fermions in graphene superlattices. Nature 2013, 497, 594–597. [Google Scholar] [CrossRef]
  10. Spanton, E.M.; Zibrov, A.A.; Zhou, H.X.; Taniguchi, T.; Watanabe, K.; Zaletel, M.P.; Young, A.F. Observation of fractional Chern insulators in a van der Waals heterostructure. Science 2018, 360, 62–66. [Google Scholar] [CrossRef]
  11. Cao, Y.; Fatemi, V.; Demir, A.; Fang, S.; Tomarken, S.L.; Luo, J.Y.; Sanchez-Yamagishi, J.D.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; et al. Correlated insulator behaviour at half-filling in magic-angle graphene superlattices. Nature 2018, 556, 80–84. [Google Scholar] [CrossRef]
  12. 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]
  13. Idzuchi, H.; Pientka, F.; Huang, K.F.; Harada, K.; Gül, Ö.; Shin, Y.J.; Nguyen, L.T.; Jo, N.H.; Shindo, D.; Cava, R.J.; et al. Unconventional supercurrent phase in Ising superconductor Josephson junction with atomically thin magnetic insulator. Nat. Commun. 2021, 12, 5332. [Google Scholar] [CrossRef]
  14. Zhou, Z.J.; Hou, F.C.; Huang, X.L.; Wang, G.; Fu, Z.H.; Liu, W.L.; Yuan, G.W.; Xi, X.X.; Xu, J.; Lin, J.H.; et al. Stack growth of wafer-scale van der Waals superconductor heterostructures. Nature 2023, 621, 499–505. [Google Scholar] [CrossRef]
  15. Conley, H.J.; Wang, B.; Ziegler, J.I.; Haglund, R.F.; Pantelides, S.T.; Bolotin, K.I. Bandgap Engineering of Strained Monolayer and Bilayer MoS2. Nano Lett. 2013, 13, 3626–3630. [Google Scholar] [CrossRef]
  16. Kang, S.; Eshete, Y.A.; Lee, S.; Won, D.; Im, S.; Lee, S.; Cho, S.; Yang, H. Bandgap modulation in the two-dimensional core-shell-structured monolayers of WS2. Iscience 2022, 25, 103563. [Google Scholar] [CrossRef]
  17. Kim, J.S.; Maity, N.; Kim, M.; Fu, S.Y.; Juneja, R.; Singh, A.; Akinwande, D.; Lin, J.F. Strain-Modulated Interlayer Charge and Energy Transfers in MoS2/WS2 Heterobilayer. ACS Appl. Mater. Interfaces 2022, 14, 46841–46849. [Google Scholar] [CrossRef]
  18. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef]
  19. Splendiani, A.; Sun, L.; Zhang, Y.B.; Li, T.S.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  20. Mak, K.F.; He, K.L.; Shan, J.; Heinz, T.F. Control of valley polarization in monolayer MoS2 by optical helicity. Nat. Nanotechnol. 2012, 7, 494–498. [Google Scholar] [CrossRef]
  21. Su, H.M.; Wei, C.R.; Deng, A.Y.; Deng, D.M.; Yang, C.L.; Dai, J.F. Anomalous enhancement of valley polarization in multilayer WS2 at room temperature. Nanoscale 2017, 9, 5148–5154. [Google Scholar] [CrossRef]
  22. Xiao, D.; Liu, G.-B.; Feng, W.; Xu, X.; Yao, W. Coupled Spin and Valley Physics in Monolayers of MoS2 and Other Group-VI Dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802. [Google Scholar] [CrossRef]
  23. Zeng, H.L.; Dai, J.F.; Yao, W.; Xiao, D.; Cui, X.D. Valley polarization in MoS2 monolayers by optical pumping. Nat. Nanotechnol. 2012, 7, 490–493. [Google Scholar] [CrossRef]
  24. Chernikov, A.; Berkelbach, T.C.; Hill, H.M.; Rigosi, A.; Li, Y.; Aslan, B.; Reichman, D.R.; Hybertsen, M.S.; Heinz, T.F. Exciton Binding Energy and Nonhydrogenic Rydberg Series in Monolayer WS2. Phys. Rev. Lett. 2014, 113, 076802. [Google Scholar] [CrossRef]
  25. 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]
  26. Qin, C.; Liu, W.; Liu, N.; Zhou, Z.; Song, J.; Ma, S.-H.; Jiao, Z.-Y.; Lei, S. Observation of hole transfer in MoS2/WS2 van der Waals heterostructures. ACS Photonics 2022, 9, 1709–1716. [Google Scholar] [CrossRef]
  27. Wu, L.; Cong, C.; Shang, J.; Yang, W.; Chen, Y.; Zhou, J.; Ai, W.; Wang, Y.; Feng, S.; Zhang, H.; et al. Raman scattering investigation of twisted WS2/MoS2 heterostructures: Interlayer mechanical coupling versus charge transfer. Nano Res. 2021, 14, 2215–2223. [Google Scholar] [CrossRef]
  28. Liu, X.G.; Li, Z.Y. Electric Field and Strain Effect on Graphene-MoS2 Hybrid Structure: Ab Initio Calculations. J. Phys. Chem. Lett. 2015, 6, 3269–3275. [Google Scholar] [CrossRef]
  29. Wang, Y.S.; Song, N.H.; Yang, X.H.; Zhang, J.; Xu, B.; Li, M.; Zheng, Y.F.; Yang, D.P. Tailoring the electronic properties of graphyne/blue phosphorene heterostructure via external electric field and vertical strain. Chem. Phys. Lett. 2019, 730, 277–282. [Google Scholar] [CrossRef]
  30. Jia, X.M.; Cao, J.; Lin, H.L.; Zhang, M.Y.; Guo, X.M.; Chen, S.F. Transforming type-I to type-II heterostructure photocatalyst via energy band engineering: A case study of I-BiOCl/I-BiOBr. Appl. Catal. B-Environ. 2017, 204, 505–514. [Google Scholar] [CrossRef]
  31. Liao, C.S.; Yu, Z.L.; He, P.B.; Zhao, Y.Q.; Liu, B.; Cai, M.Q. Effects of composition modulation on the type of band alignments for Pd2Se3/CsSnBr3 van der waals heterostructure: A transition from type I to type II. J. Power Sources 2020, 478, 229078. [Google Scholar] [CrossRef]
  32. Zhong, D.; Seyler, K.L.; Linpeng, X.Y.; Wilson, N.P.; Taniguchi, T.; Watanabe, K.; McGuire, M.A.; Fu, K.M.C.; Xiao, D.; Yao, W.; et al. Layer-resolved magnetic proximity effect in van der Waals heterostructures. Nat. Nanotechnol. 2020, 15, 187–191. [Google Scholar] [CrossRef]
  33. Cho, C.; Wong, J.; Taqieddin, A.; Biswas, S.; Aluru, N.R.; Nam, S.; Atwater, H.A. Highly Strain-Tunable Interlayer Excitons in MoS2/WSe2 Heterobilayers. Nano Lett. 2021, 21, 3956–3964. [Google Scholar] [CrossRef]
  34. Susarla, S.; Manimunda, P.; Jaques, Y.M.; Hachtel, J.A.; Idrobo, J.C.; Amnulla, S.A.S.; Galvao, D.S.; Tiwary, C.S.; Ajayan, P.M. Deformation Mechanisms of Vertically Stacked WS2/MoS2 Heterostructures: The Role of Interfaces. ACS Nano 2018, 12, 4036–4044. [Google Scholar] [CrossRef]
  35. Zhan, H.; Guo, D.; Xie, G.X. Two-Dimensional Layered Materials: From Mechanical and Coupling Properties towards Applications in Electronics. Nanoscale 2019, 11, 13181–13212. [Google Scholar] [CrossRef]
  36. Peng, Z.; Chen, X.; Fan, Y.; Srolovitz, D.J.; Lei, D. Strain Engineering of 2D Semiconductors and Graphene: From Strain Fields to Band-Structure Tuning and Photonic Applications. Light Sci. Appl. 2020, 9, 190. [Google Scholar] [CrossRef]
  37. Carrascoso, F.; Li, H.; Frisenda, R.; Castellanos-Gomez, A. Strain engineering in single-, bi-and tri-layer MoS2, MoSe2, WS2 and WSe2. Nano Res. 2021, 14, 1698–1703. [Google Scholar] [CrossRef]
  38. Castellanos-Gomez, A.; Roldán, R.; Cappelluti, E.; Buscema, M.; Guinea, F.; Van Der Zant, H.S.; Steele, G.A. Local strain engineering in atomically thin MoS2. Nano Lett. 2013, 13, 5361–5366. [Google Scholar] [CrossRef]
  39. Modtland, B.J.; Navarro-Moratalla, E.; Ji, X.; Baldo, M.; Kong, J. Monolayer tungsten disulfide (WS2) via chlorine-driven chemical vapor transport. Small 2017, 13, 1701232. [Google Scholar] [CrossRef]
  40. 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]
  41. Zhang, J.; Du, L.; Feng, S.; Zhang, R.-W.; Cao, B.; Zou, C.; Chen, Y.; Liao, M.; Zhang, B.; Yang, S.A.; et al. Enhancing and controlling valley magnetic response in MoS2/WS2 heterostructures by all-optical route. Nat. Commun. 2019, 10, 4226. [Google Scholar] [CrossRef]
  42. Tyurnina, A.V.; Bandurin, D.A.; Khestanova, E.; Kravets, V.G.; Koperski, M.; Guinea, F.; Grigorenko, A.N.; Geim, A.K.; Grigorieva, I.V. Strained bubbles in van der Waals heterostructures as local emitters of photoluminescence with adjustable wavelength. ACS Photonics 2019, 6, 516–524. [Google Scholar] [CrossRef]
  43. Khestanova, E.; Guinea, F.; Fumagalli, L.; Geim, A.K.; Grigorieva, I.V. Universal shape and pressure inside bubbles appearing in van der Waals heterostructures. Nat. Commun. 2016, 7, 12587. [Google Scholar] [CrossRef]
  44. Lan, C.Y.; Li, C.; Yin, Y.; Liu, Y. Large-area synthesis of monolayer WS2 and its ambient-sensitive photo-detecting performance. Nanoscale 2015, 7, 5974–5980. [Google Scholar] [CrossRef]
  45. Oh, H.M.; Han, G.H.; Kim, H.; Bae, J.J.; Jeong, M.S.; Lee, Y.H. Photochemical reaction in monolayer MoS2 via correlated photoluminescence, Raman spectroscopy, and atomic force microscopy. ACS Nano 2016, 10, 5230–5236. [Google Scholar] [CrossRef]
  46. Wang, S.; Rong, Y.; Fan, Y.; Pacios, M.; Bhaskaran, H.; He, K.; Warner, J.H. Shape evolution of monolayer MoS2 crystals grown by chemical vapor deposition. Chem. Mater. 2014, 26, 6371–6379. [Google Scholar] [CrossRef]
  47. Chen, W.; Zhao, J.; Zhang, J.; Gu, L.; Yang, Z.; Li, X.; Yu, H.; Zhu, X.; Yang, R.; Shi, D.; et al. Oxygen-assisted chemical vapor deposition growth of large single-crystal and high-quality monolayer MoS2. J. Am. Chem. Soc. 2015, 137, 15632–15635. [Google Scholar] [CrossRef]
  48. Dobusch, L.; Schuler, S.; Perebeinos, V.; Mueller, T. Thermal light emission from monolayer MoS2. Adv. Mater. 2017, 29, 1701304. [Google Scholar] [CrossRef]
  49. Hu, Z.; Bao, Y.; Li, Z.; Gong, Y.; Feng, R.; Xiao, Y.; Wu, X.; Zhang, Z.; Zhu, X.; Ajayan, P.M.; et al. Temperature dependent Raman and photoluminescence of vertical WS2/MoS2 monolayer heterostructures. Sci. Bull. 2017, 62, 16–21. [Google Scholar] [CrossRef]
  50. He, Z.; Guo, Z.; Zhong, X.; Chen, X.; Xue, J.; Wang, X.; Chen, Y. Spectroscopic investigation of defects mediated oxidization of single-layer MoS2. Sci. Chi. Technol. Sci. 2021, 64, 611–619. [Google Scholar] [CrossRef]
  51. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef]
  52. Berkdemir, A.; Gutiérrez, H.R.; Botello-Méndez, A.R.; Perea-López, N.; Elías, A.L.; Chia, C.-I.; Wang, B.; Crespi, V.H.; López-Urías, F.; Charlier, J.-C.; et al. Identification of individual and few layers of WS2 using Raman Spectroscopy. Sci. Rep. 2013, 3, 1755. [Google Scholar] [CrossRef]
  53. Pak, S.; Lee, J.; Lee, Y.-W.; Jang, A.-R.; Ahn, S.; Ma, K.Y.; Cho, Y.; Hong, J.; Lee, S.; Jeong, H.Y.; et al. Strain-mediated interlayer coupling effects on the excitonic behaviors in an epitaxially grown MoS2/WS2 van der Waals heterobilayer. Nano Lett. 2017, 17, 5634–5640. [Google Scholar] [CrossRef]
  54. Xu, X.; Wang, C.; Xiong, W.; Liu, Y.; Yang, D.; Zhang, X.; Xu, J.J. Strain regulated interlayer coupling in WSe2/WS2 heterobilayer. Nanotechnology 2021, 33, 085705. [Google Scholar] [CrossRef]
  55. 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]
  56. Yan, P.; Wang, J.; Yang, G.; Lu, N.; Chu, G.; Zhang, X.; Shen, X. Microstructures. Chemical vapor deposition of monolayer MoS2 on sapphire, Si and GaN substrates. Superlattices Microstruct. 2018, 120, 235–240. [Google Scholar] [CrossRef]
  57. Qi, J.J.; Lan, Y.W.; Stieg, A.Z.; Chen, J.H.; Zhong, Y.L.; Li, L.J.; Chen, C.D.; Zhang, Y.; Wang, K.L. Piezoelectric effect in chemical vapour deposition-grown atomic-monolayer triangular molybdenum disulfide piezotronics. Nat. Commun. 2015, 6, 7430. [Google Scholar] [CrossRef]
  58. Lee, H.; Koo, Y.; Choi, J.; Kumar, S.; Lee, H.-T.; Ji, G.; Choi, S.H.; Kang, M.; Kim, K.K.; Park, H.-R.; et al. Drift-dominant exciton funneling and trion conversion in 2D semiconductors on the nanogap. Sci. Adv. 2022, 8, eabm5236. [Google Scholar] [CrossRef]
  59. Koo, Y.; Lee, H.; Ivanova, T.; Kefayati, A.; Perebeinos, V.; Khestanova, E.; Kravtsov, V.; Park, K.-D. Tunable interlayer excitons and switchable interlayer trions via dynamic near-field cavity. Light Sci. Appl. 2023, 12, 59. [Google Scholar] [CrossRef]
  60. Dou, X.M.; Ding, K.; Jiang, D.S.; Sun, B.Q. Tuning and Identification of Interband Transitions in Monolayer and Bilayer Molybdenum Disulfide Using Hydrostatic Pressure. ACS Nano 2014, 8, 7458–7464. [Google Scholar] [CrossRef]
  61. Fu, L.; Wan, Y.; Tang, N.; Ding, Y.M.; Gao, J.; Yu, J.C.; Guan, H.M.; Zhang, K.; Wang, W.Y.; Zhang, C.F.; et al. K-Λ crossover transition in the conduction band of monolayer MoS2 under hydrostatic pressure. Sci. Adv. 2017, 3, e1700162. [Google Scholar] [CrossRef]
  62. Xia, J.; Yan, J.; Wang, Z.; He, Y.; Gong, Y.; Chen, W.; Sum, T.C.; Liu, Z.; Ajayan, P.M.; Shen, Z. Strong coupling and pressure engineering in WSe2–MoSe2 heterobilayers. Nat. Phys. 2021, 17, 92–98. [Google Scholar] [CrossRef]
  63. Li, G.H.; Goni, A.R.; Syassen, K.; Brandt, O.; Ploog, K. State Mixing In InAs/GaAs Quantum Dots At The Pressure-Induced Gamma-X Crossing. Phys. Rev. B 1994, 50, 18420–18425. [Google Scholar] [CrossRef]
  64. Mattila, T.; Wei, S.H.; Zunger, A. Localization and anticrossing of electron levels in GaAs1−xNx alloys. Phys. Rev. B 1999, 60, 11245–11248. [Google Scholar] [CrossRef]
  65. Shan, W.; Walukiewicz, W.; Ager, J.W.; Haller, E.E.; Geisz, J.F.; Friedman, D.J.; Olson, J.M.; Kurtz, S.R. Band anticrossing in GaInNAs alloys. Phys. Rev. Lett. 1999, 82, 1221–1224. [Google Scholar] [CrossRef]
  66. Walukiewicz, W.; Shan, W.; Yu, K.M.; Ager, J.W.; Haller, E.E.; Miotkowski, I.; Seong, M.J.; Alawadhi, H.; Ramdas, A.K. Interaction of localized electronic states with the conduction band: Band anticrossing in II-VI semiconductor ternaries. Phys. Rev. Lett. 2000, 85, 1552–1555. [Google Scholar] [CrossRef]
  67. Wu, J.; Shan, W.; Walukiewicz, W.; Yu, K.M.; Ager, J.W.; Haller, E.E.; Xin, H.P.; Tu, C.W. Effect of band anticrossing on the optical transitions in GaAs1−xNx/GaAs multiple quantum wells. Phys. Rev. B 2001, 64, 085320. [Google Scholar] [CrossRef]
  68. Wu, J.; Walukiewicz, W.; Yu, K.M.; Ager, J.W.; Haller, E.E.; Hong, Y.G.; Xin, H.P.; Tu, C.W. Band anticrossing in GaP1−xNx alloys. Phys. Rev. B 2002, 65, 241303. [Google Scholar] [CrossRef]
Figure 1. The synthesis procedures of the monolayers. (a) Monolayer MoS2 by CVD. (b) Monolayer WS2 by CVT.
Figure 1. The synthesis procedures of the monolayers. (a) Monolayer MoS2 by CVD. (b) Monolayer WS2 by CVT.
Applsci 14 02179 g001
Figure 2. Characterization of WS2/MoS2 heterostructure. (a) Optical image. (b) AFM topography image. (c) AFM height curve. (d) Raman spectra of the three samples. (e) PL spectra of the three samples. (f) PL spectra of MoS2/WS2 HS and WS2/MoS2 HS.
Figure 2. Characterization of WS2/MoS2 heterostructure. (a) Optical image. (b) AFM topography image. (c) AFM height curve. (d) Raman spectra of the three samples. (e) PL spectra of the three samples. (f) PL spectra of MoS2/WS2 HS and WS2/MoS2 HS.
Applsci 14 02179 g002
Figure 3. (a) Illustration of the measurement. (b) PL spectra with probe contact.
Figure 3. (a) Illustration of the measurement. (b) PL spectra with probe contact.
Applsci 14 02179 g003
Figure 4. (ac) PL spectra of (a) ML-WS2, (b) ML-MoS2, and (c) WS2/MoS2 HS under out-of-plane pressure ranging from 0 to 120 nN. (d) Two-dimensional mapping of the HS PL spectra.
Figure 4. (ac) PL spectra of (a) ML-WS2, (b) ML-MoS2, and (c) WS2/MoS2 HS under out-of-plane pressure ranging from 0 to 120 nN. (d) Two-dimensional mapping of the HS PL spectra.
Applsci 14 02179 g004
Figure 5. The peak intensity and photo energy fitted from PL spectra of (a,b) ML-WS2, (c,d) ML-MoS2, and (e,f) WS2/MoS2 HS.
Figure 5. The peak intensity and photo energy fitted from PL spectra of (a,b) ML-WS2, (c,d) ML-MoS2, and (e,f) WS2/MoS2 HS.
Applsci 14 02179 g005
Figure 6. The PL spectra of WS2/MoS2 HS under applied pressure gradually decreasing from 140 to 0 nN.
Figure 6. The PL spectra of WS2/MoS2 HS under applied pressure gradually decreasing from 140 to 0 nN.
Applsci 14 02179 g006
Figure 7. (a) DFT-calculated band structure of WS2/MoS2 HS. (b) Illustration of the charge transfer in the heterostructure with Si probe contact.
Figure 7. (a) DFT-calculated band structure of WS2/MoS2 HS. (b) Illustration of the charge transfer in the heterostructure with Si probe contact.
Applsci 14 02179 g007
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

Kong, W.; Ren, Z.; Chen, P.; Cui, J.; Chen, Y.; Wu, J.; Li, Y.; Liu, W.; Li, P.; Fu, Y.; et al. Excitonic Evolution in WS2/MoS2 van der Waals Heterostructures Turned by Out-of-Plane Localized Pressure. Appl. Sci. 2024, 14, 2179. https://0-doi-org.brum.beds.ac.uk/10.3390/app14052179

AMA Style

Kong W, Ren Z, Chen P, Cui J, Chen Y, Wu J, Li Y, Liu W, Li P, Fu Y, et al. Excitonic Evolution in WS2/MoS2 van der Waals Heterostructures Turned by Out-of-Plane Localized Pressure. Applied Sciences. 2024; 14(5):2179. https://0-doi-org.brum.beds.ac.uk/10.3390/app14052179

Chicago/Turabian Style

Kong, Weihu, Zeqian Ren, Peng Chen, Jinxiang Cui, Yili Chen, Jizhou Wu, Yuqing Li, Wenliang Liu, Peng Li, Yongming Fu, and et al. 2024. "Excitonic Evolution in WS2/MoS2 van der Waals Heterostructures Turned by Out-of-Plane Localized Pressure" Applied Sciences 14, no. 5: 2179. https://0-doi-org.brum.beds.ac.uk/10.3390/app14052179

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop