Next Article in Journal
Band Dynamics of Multimode Resonant Nanophotonic Lattices with Adjustable Liquid Interfaces
Next Article in Special Issue
Enhanced Sunlight-Powered Photocatalysis and Methanol Oxidation Activities of Co3O4-Embedded Polymeric Carbon Nitride Nanostructures
Previous Article in Journal
Polycrystalline Transparent Al-Doped ZnO Thin Films for Photosensitivity and Optoelectronic Applications
Previous Article in Special Issue
A Review on Montmorillonite-Based Nanoantimicrobials: State of the Art
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature-Dependent Phonon Scattering and Photoluminescence in Vertical MoS2/WSe2 Heterostructures

1
Hunan Institute of Optoelectronic Integration, College of Materials Science and Engineering, Hunan University, Changsha 410082, China
2
Wuhan National Laboratory for Optoelectronics, School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(16), 2349; https://doi.org/10.3390/nano13162349
Submission received: 22 July 2023 / Revised: 8 August 2023 / Accepted: 9 August 2023 / Published: 16 August 2023
(This article belongs to the Special Issue Nanomaterials and 2D Materials Based on Semiconductors and Metals)

Abstract

:
Transition metal dichalcogenide (TMD) monolayers and their heterostructures have attracted considerable attention due to their distinct properties. In this work, we performed a systematic investigation of MoS2/WSe2 heterostructures, focusing on their temperature-dependent Raman and photoluminescence (PL) characteristics in the range of 79 to 473 K. Our Raman analysis revealed that both the longitudinal and transverse modes of the heterostructure exhibit linear shifts towards low frequencies with increasing temperatures. The peak position and intensity of PL spectra also showed pronounced temperature dependency. The activation energy of thermal-quenching-induced PL emissions was estimated as 61.5 meV and 82.6 meV for WSe2 and MoS2, respectively. Additionally, we observed that the spectral full width at half maximum (FWHM) of Raman and PL peaks increases as the temperature increases, and these broadenings can be attributed to the phonon interaction and the expansion of the heterostructure’s thermal coefficients. This work provides valuable insights into the interlayer coupling of van der Waals heterostructures, which is essential for understanding their potential applications in extreme temperatures.

1. Introduction

Transition metal dichalcogenides have recently garnered significant interest due to their scalability and thickness-dependent electrical and optical properties [1]. The evolution of bulk TMDs to atomically thin 2D layered structures is characterized by their transition from indirect to direct bandgap semiconductors [2]. Such a transition of bandgaps has significant implications for their optoelectronic properties and potential applications [3,4,5,6]. These 2D semiconducting materials, MX2 (M = Mo, W; X = S, Se), are exemplified by compounds such as molybdenum disulfide (MoS2) and tungsten diselenide (WSe2) with direct bandgaps of 1.9 eV and 1.6 eV, respectively [7]. Beyond monolayers, vertically-stacked heterostructures based on the van der Waals force provide a fascinating platform for investigating novel physical phenomena [8]. With the assistance of an artificial stacking arrangement, electronic band engineering of heterostructures can be achieved, resulting in the modulation of optical properties via interlayer coupling effects [9].
Heterostructures composed of two-dimensional materials have been shown to exhibit various intriguing phenomena, including ultrafast charge transfer [10], high PL quantum yield [11], and interlayer valley physics [12]. These properties make them potential candidates for applications in field-effect transistors [13], electrocatalysis [14], and photodetectors [15]. The thermodynamic stability of these heterostructures plays a critical role in their optoelectronic properties [16]. A comprehensive understanding of the correlation between phonon scattering and heterostructure interfacial coupling is imperative to achieve the optimal performance of such structures [17,18]. Raman spectroscopy is an essential tool to investigate the interaction between electrons and phonons in 2D materials and their heterostructures [19]. Research on temperature-dependent Raman spectroscopy of few-layered WS2 provides valuable insights into their intrinsic phonon scattering processes and thermal properties [20]. Several early experiments investigated the impact of temperature on the phononic properties of few-layered TMDs, including MoS2, WS2, MoSe2, and others [21]. These studies have demonstrated that temperature-dependent behaviors of phononic modes in these nanosheets are attributed to the effects of thermal expansion and anharmonic resonance [22]. Moreover, some studies investigated the impact of interlayer coupling on the Raman spectra of TMDs heterostructures, which demonstrated a strong dependence on the stacking orientation of the constituent monolayers [23,24].
This work provides a systematic investigation into the temperature-dependent Raman characteristics of a heterostructure composed of MoS2 and WSe2 monolayers. Raman spectroscopy measurements were conducted over the controlled temperature range of 79 to 473 K to investigate the in-plane ( E 2 g 1 ) and out-of-plane ( A 1 g ) vibrational modes of the heterostructure. Experimental observations indicate that both Raman modes display a nearly linear dependence across the entire temperature range, alongside a commensurate variation in their full width at half maximum. The primary focus of this study is on attempting to understand the variation of interlayer interactions, particularly temperature-dependent phonon interactions, and the thermal expansion coefficients of the heterostructure under high vacuum conditions. Furthermore, temperature-dependent photoluminescence was also observed and utilized to determine the activation energy of the PL emissions caused by thermal quenching. This work provides a platform to investigate the photodynamics of low-dimensional materials and also helps to understand the electronic and photonic properties of van der Waals heterostructures.

2. Materials and Methods

The synthesis of 2D transition metal dichalcogenides has revolutionized modern techniques, such as physical vapor deposition (PVD) [25], chemical vapor deposition (CVD) [26], electrodeposition [27], and thermal synthesis [28], facilitating their advancement and enhancing their capabilities for diverse applications. In this work, a MoS2 monolayer was grown by a chemical vapor deposition method using sulfur (S) and molybdenum oxide (MoO3) powders. MoO3 and S powder were put in separate boats at the center of a fused quartz tube located in a furnace. Subsequently, a SiO2/Si wafer was suspended on one of the boats, and the temperature was ramped up to 810 °C for 30 min and maintained for 10 min. When it cooled down to room temperature, triangular-shaped MoS2 monolayers were successfully prepared. Further details can be found in our previous work [29]. The synthesis procedure was conducted at atmospheric pressure using Ar gas as a carrier agent with a flow rate of 70 sccm. The same CVD procedure was employed to grow WSe2 utilizing tungsten oxide (WO3) and selenium (Se) powder as source materials, with the furnace temperature kept constant at 950 °C and naturally allowed to cool. A comprehensive description of the synthesis procedures and a schematic depiction of the CVD setup can be found in Supplementary Section S1. A simple wet transfer technique was employed to prepare the MoS2/WSe2 heterostructures [30]. Further details about the transfer process and the schematic illustration can be found in Supplementary Section S2. The heterostructure underwent annealing in a vacuum furnace at 70 °C for 12 h to eliminate water molecules and promote better crystallinity.
For spectral measurements of materials, PL and Raman spectra were analyzed using an iHR550 Raman spectrometer from Horiba with a laser excitation of 532 nm. An objective lens providing 50 times magnification was utilized to enable a more detailed analysis of the sample’s features. The diameter of the laser spot was about 1 µm. The spectrometer was equipped with gratings of 300 g mm−1 and 1200 g mm−1 to facilitate accurate analysis of the spectral data. Furthermore, a temperature-controlled cryostat (THMS-600, Linkam Scientific Instruments Ltd., Redhill, UK) was utilized to investigate the temperature-dependent behavior of the heterostructure (Supplementary Section S3). The cryostat utilized liquid nitrogen to attain the lowest temperature of the boiling point of liquid nitrogen. Samples were loaded into the cryostat full of nitrogen to avoid any impact from the surface adsorption of gas molecules and to enable precise temperature control during measurements.

3. Results and Discussion

Figure 1a shows the schematic view of the MoS2/WSe2 heterostructure. The MoS2 and WSe2 monolayers were in a 2H phase, behaving as semiconductors with direct band gaps. Figure 1b shows the scanning electron microscopy (SEM) image of the as-prepared MoS2/WSe2 heterostructure, where the top WSe2 monolayer has been transferred onto the bottom MoS2 monolayer. The scale bar is 10 µm. Figure 1c shows the optical image of heterostructures, where monolayer and heterostructure regions can be observed. The monolayer thicknesses of MoS2 (region 1) and WSe2 (region 2) were determined to be 0.8 nm and 0.76 nm, respectively, using atomic force microscopy (AFM) as described in Supplementary Section S1. The height profiles of both monolayer regions are plotted in Figure 1d. Additional AFM results of the heterostructure (region 3) with a 0.39 nm interlayer distance can be found in Supplementary Section S1.
Raman spectroscopy was employed to characterize the structure and spectral properties of the MoS2/WSe2 heterostructure. The sample was put into a vacuum annealing furnace to anneal the heterostructure sample and mitigate the influence of defects and gas adsorption (Supplementary Section S2). Moreover, to avoid any local heating of the sample, which can alter peak positions in the Raman spectra, the excitation power was carefully controlled at 1.2 mW. The Raman spectra in Figure 2a display two distinct peaks corresponding to the longitudinal (E) and transverse (A) modes of monolayer regions in the MoS2/WSe2 heterostructure. The Raman modes of E 2 g 1 and A 1 g for MoS2 (red curve) show a separation of about 18 cm−1, which is consistent with previously reported works on monolayer MoS2 [31].
The transverse and longitudinal modes were detected at 252 cm−1 and 260 cm−1, respectively, in the monolayer WSe2 region of the heterostructure, consistent with previous studies [32]. The Raman spectra acquired from the heterostructure region (black curve) shown in Figure 2a exhibit identical vibrational modes, as observed in both monolayers. Nevertheless, a marginal reduction is observed in the intensity of these modes, which is attributable to enhanced excitation light scattering caused by the presence of the heterostructure. The MoS2/WSe2 heterostructure manifests a type II band alignment, which facilitates effective charge transfer, yet concurrently reduces the PL intensity by spatially separating the valence and conduction bands. The photoluminescence peaks exhibited by the heterostructure in Figure 2b are attributed to the “A” excitons, corresponding to the direct bandgap transitions at the K-point of the Brillouin zone in both MoS2 and WSe2. For the MoS2 spectrum, an additional peak at higher energy near 2.02 eV was observed, corresponding to the “B” exciton formed due to transitions between the spin-orbit split valence band and the conduction band [33]. However, the emission strength of the B exciton is relatively weaker than the A exciton due to its lower oscillator strength, which resulted from an indirect transition.
PL mapping was performed at specific wavelengths to assess the photoemissions of MoS2/WSe2 heterostructures. As illustrated in Figure 2c,d, the resulting PL emission maps at 758 nm and 655 nm provide valuable insight into the distribution of PL intensity from MoS2 and WSe2, respectively, in the heterostructure. Remarkably, the photoluminescence maps obtained in this study demonstrate a stronger emission signal emanating from the MoS2 monolayer, which can be ascribed to its elevated radiative recombination rate and binding energy. Importantly, these findings align with previous investigations of similar systems [34]. Moreover, photoluminescence measurements were conducted for both MoS2 and WSe2 in the monolayer position as well as the heterostructure region. Notably, the peak intensities at 655 nm and 758 nm in the heterostructure’s region 3 exhibit significant quenching compared to the monolayer regions (1 and 2). This behavior can be attributed to the ultrafast charge transfer phenomenon observed in atomically thin MoS2/WSe2 heterostructures. Detailed discussions can be found in Supplementary Section S4 for a more in-depth analysis of these findings. All Raman and PL spectra in Figure 2 were measured at room temperature using a 532 nm laser of 1.2 mW.
To better comprehend the vibrational modes and interlayer coupling in MoS2/WSe2 heterostructures, temperature-dependent Raman measurements ranging from 79 to 473 K, as illustrated in Figure 3a,b, enable insights into their thermal behavior. The Raman findings indicate strong scattering intensities of the E 2 g 1 and A 1 g modes for MoS2/WSe2 heterostructures with increasing temperatures. Thermal expansion of the lattice is known to trigger a decline in the vibrational frequency with increasing temperature, thereby inducing a downward shift of the Raman modes towards lower frequencies, as corroborated by the outcomes shown in Figure 3a,b. Remarkably, the uniform redshift across all Raman modes with increasing temperature implies a systematic alteration in the vibrational characteristics of the heterostructure. At high temperatures, it looks as though split peaks appear apart from the main peaks, but actually, we are observing the temperature-dependent enhancement and widening of inconspicuous peaks at low temperatures. To visualize these changes, Figure 3c illustrates the temperature-dependent peak positions of the Raman modes. Our experimental data exhibit a linear decline with temperature and are well-fitted by a linear function.
Elevated temperatures lead to increased atomic vibrations in the lattices of WSe2 and MoS2, which causes an increase in the scattering rate and a reduction in the lifetime of vibrational modes. These effects contribute to the broadening of the Raman spectra, as demonstrated by the increase in full width at half maximum (FWHM) observed in Figure 3d. The FWHM of the WSe2 region in the heterostructure is particularly significant, perhaps due to its lower binding energy, which leads to the higher scattering of lattice modes at higher temperatures (indicated in Figure 3d). When the temperature rises, the thermal energy of the lattice vibrations increases, resulting in an increased number of phonon scattering events. In addition, the thermal energy of the lattice vibrations in 2D materials also increases, resulting in a greater number of phonon scattering events. This, in turn, leads to an increase in the FWHM of the Raman peak associated with the A 1 g   mode. In MoS2, this effect is more pronounced for the A 1 g mode compared to the E 2 g 1 mode due to the latter’s less sensitive in-plane vibrations of Mo and S atoms, as illustrated in Figure 3d. It is noteworthy that slight deviations observed at certain temperatures may be attributed to slight instrumentation instability in our Raman setup. The differences in Raman peak broadening between the E 2 g 1 and A 1 g modes of WSe2 and MoS2 provide valuable information about the behavior of these materials under varying thermal conditions.
In addition to its effects on the Raman spectra, temperature can also influence the photon emissions of the MoS2/WSe2 heterostructures. The aspect of temperature-dependent photoluminescence in heterostructures is crucial in device physics to improve their operation performance. The combined effect of thermal and optical energies that contribute to the de-trapping of carriers and desorption of adsorbates (such as O2 and H2O) could result in changes in the conductivity of the material, as well as the optical emission properties, as discussed elsewhere [35]. In the case of heterostructures, temperature-dependent PL spectroscopy has revealed new and interesting effects, including an anomalous redshift, which could be related to thermal quenching, defect engineering, and strain. Several studies have explored and linked the temperature-dependent PL spectra to the underlying band structure, electronic properties, and potential device applications of heterostructures. To investigate thermal quenching and the underlying band structure, photoluminescence measurements of the heterostructure were conducted over a range of temperatures (79 K to 473 K), as illustrated in Figure 4. The results of the photoluminescence measurements reveal a significant temperature dependence in both the peak position and intensity of the spectrum. As the temperature increases, both the WSe2 and MoS2 peaks experience redshifts towards longer wavelengths, as demonstrated in Figure 4a,b. The reduction in photoluminescence intensity with increasing temperature is a well-known phenomenon observed in both two-dimensional transition metal dichalcogenides and bulk semiconductors, attributable to the amplification of non-radiative recombination, thermal activation of electrons and holes, and phonon-assisted recombination.
Here, the investigation of the temperature dependence of photoluminescence intensity in MoS2/WSe2 heterostructures revealed that the Arrhenius formula (Equation (1)) provides a good fit for the experimental data. This formula correlates the photoluminescence quenching rate with the thermal activation energy for non-radiative recombination, as demonstrated in Figure 4c,d. These findings offer insights into the fundamental processes involved in optoelectronic devices and emphasize the significance of carefully controlling temperature effects in such systems [36].
I T = A 1 + C   e ( E / k B   T )  
Here, in Equation (1), the temperature-dependent PL intensity (I) is the main variable, along with the density of quenching centers (A) and the strength of electron–phonon coupling (C). The thermal quenching process is characterized by the thermal activation energy (E), and the Boltzmann constant (kB) is utilized in the formula to relate temperature and energy. The thermal activation energies of MoS2 (82.6 meV) and WSe2 (61.5 meV) were determined via Arrhenius fitting of the experimental data presented in Figure 4c,d. The two materials exhibit different thermal activation energies, E, due to differences in the bandgap, effective mass, and mechanical strain. The distinct crystal structures and chemical bonding may also contribute to different effective masses in the MoS2/WSe2 heterostructure, resulting in varying carrier mobilities and activation energies. The observed differences in the thermal activation energies between MoS2 and WSe2 can ultimately be attributed to material-specific factors, including band structures, carrier dynamics, and mechanical forces. As shown in Figure 4e,f, the photoluminescence peak position, corresponding to the direct excitonic transition energy in MoS2/WSe2, is affected by temperature variations. An increasing temperature causes a redshift in the PL peak, indicating a reduction in the energy gap of the heterostructure [37]. This energy gap refers to the energy of the lowest allowable exciton state, which corresponds to the energy required to create an electron–hole pair in the system. The decreased energy gap shifts the lowest allowable exciton state to lower energy. This effect is commonly observed in various materials, including semiconductors and insulators. Density functional theory (DFT) calculations are performed in Supplementary Section S4. The relationship between temperature and the PL peak position is typically described by empirical equations, such as the one proposed by Varshni [38], given in Equation (2):
E g T = E g 0 + α T 2 T + β
In this equation, E g 0 represents the band gap at a temperature of absolute zero, where T = 0 K; β is a constant approximating the Debye temperature of the material; and α is the coefficient representing the change in bandgap energy with temperature. The Varshni fitting shown in Figure 4e,f provides a good agreement with the experimental data, yielding bandgap values of E0 = 1.94 eV and 1.64 eV for MoS2 and WSe2, respectively. In the same way, the O’Donnell and Chen formula, as given in Equation (3), is another model to fit the temperature-dependent PL and obtain the bandgap of the material [39].
E g = E g 0 S < h 2 π > cot h ω 4 π k B T 1
Here, the parameter E g 0   gives the bandgap at zero temperature, while S gives information on electron–phonon coupling, and ω is the photon frequency. The fitting curves in Figure 4e,f for both MoS2 and WSe2 peak position shifting versus temperature are well-matched to the experimental data. The band gap values obtained using the O’Donnell and Chen formula for MoS2 and WSe2 are 1.93 eV and 1.65 eV, respectively, which are consistent with the results from the Varshni fitting, as well as with previously reported observations [40]. These temperature-dependent observations of phonon scattering and the interlayer coupling of heterostructures provide a significant influence on the performance of heterostructure devices; for example, the thermal noise is serious in photodetectors, and the carrier mobility of transistors is highly influenced by the environmental temperature.

4. Conclusions

Our study provides important insights into the temperature-dependent photoluminescence and bandgap characteristics of the MoS2/WSe2 heterostructure. These findings expand our understanding of this material system and have potential implications for various technological applications. We observed that both the longitudinal and transverse modes of MoS2 and WSe2 exhibit a linear variation in peak position and full width at half maximum with temperature. The quenching of the photoluminescence in the heterostructure is attributed to non-radiative recombination mechanisms. The behavior of the photoluminescence with temperature is similar to that observed in bulk MoS2 and WSe2. Furthermore, we found that differences in the bandgap, effective mass, and mechanical strain of MoS2 and WSe2 lead to different thermal activation energies, with values of 82.6 meV and 61.5 meV, respectively. Overall, our study contributes to the existing knowledge based on MoS2/WSe2 heterostructures and provides important insights into their temperature-dependent optoelectronic properties, which can enlighten the design and optimization of various devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13162349/s1: Section S1. CVD preparation of MoS2 and WSe2 and their corresponding AFM analysis. Section S2. Preparation of MoS2/WSe2 heterostructure. Section S3. Experimental setup of Raman and PL spectroscopy. Section S4. PL Measurement and Details of the DFT Calculation of MoS2/WSe2 Hetrostructure. References [41,42,43,44,45,46,47,48,49,50] are cited in the Supplementary Materials.

Author Contributions

W.A. and M.H. prepared the samples. W.A. and Y.X. performed the spectral experiments. W.A. and Y.L. analyzed and processed experiment data. W.A and Z.L. designed the experiment and wrote the manuscript. Y.L. and Z.L. offered the funding. W.A. and Y.L. contributed equally as the first authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (No. 62275076, 92163135) and the Open Project Program of Wuhan National Laboratory for Optoelectronics (No. 2022WNLOKF003).

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors would like to thank Sajid Ur Rehman for his efforts regarding the DFT calculations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yin, X.; Tang, C.S.; Zheng, Y.; Gao, J.; Wu, J.; Zhang, H.; Chhowalla, M.; Chen, W.; Wee, A.T. 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]
  2. Zhang, Y.; Chang, T.-R.; Zhou, B.; Cui, Y.-T.; Yan, H.; Liu, Z.; Schmitt, F.; Lee, J.; Moore, R.; Chen, Y. Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2. Nat. Nanotechnol. 2014, 9, 111–115. [Google Scholar]
  3. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  4. Sitt, A.; Hadar, I.; Banin, U. Band-gap engineering, optoelectronic properties and applications of colloidal heterostructured semiconductor nanorods. Nano Today 2013, 8, 494–513. [Google Scholar] [CrossRef]
  5. Hadi, M.A.; Islam, M.N.; Podder, J. Indirect to direct band gap transition through order to disorder transformation of Cs2AgBiBr6 via creating antisite defects for optoelectronic and photovoltaic applications. RSC Adv. 2022, 12, 15461–15469. [Google Scholar] [CrossRef]
  6. Chaves, A.; Azadani, J.G.; Alsalman, H.; da Costa, D.R.; Frisenda, R.; Chaves, A.J.; Song, S.H.; Kim, Y.D.; He, D.; Zhou, J.; et al. Bandgap engineering of two-dimensional semiconductor materials. NPJ 2D Mater. Appl. 2020, 4, 29. [Google Scholar] [CrossRef]
  7. Singh, E.; Kim, K.S.; Yeom, G.Y.; Nalwa, H.S. Atomically thin-layered molybdenum disulfide (MoS2) for bulk-heterojunction solar cells. ACS Appl. Mater. Interfaces 2017, 9, 3223–3245. [Google Scholar] [CrossRef] [PubMed]
  8. Tartakovskii, A. Excitons in 2D heterostructures. Nat. Rev. Phys. 2020, 2, 8–9. [Google Scholar] [CrossRef]
  9. Jadczak, J.; Kutrowska-Girzycka, J.; Schindler, J.J.; Debus, J.; Watanabe, K.; Taniguchi, T.; Ho, C.-H.; Bryja, L. Investigations of electron-electron and interlayer electron-phonon coupling in van der Waals hBN/WSe2/hBN heterostructures by photoluminescence excitation experiments. Materials 2021, 14, 399. [Google Scholar] [PubMed]
  10. Qiao, H.; Liu, H.; Huang, Z.; Hu, R.; Ma, Q.; Zhong, J.; Qi, X. Tunable electronic and optical properties of 2D monoelemental materials beyond graphene for promising applications. Energy Environ. Mater. 2021, 4, 522–543. [Google Scholar]
  11. Kim, H.; Ahn, G.H.; Cho, J.; Amani, M.; Mastandrea, J.P.; Groschner, C.K.; Lien, D.-H.; Zhao, Y.; Ager III, J.W.; Scott, M.C. Synthetic WSe2 monolayers with high photoluminescence quantum yield. Sci. Adv. 2019, 5, 4728. [Google Scholar] [CrossRef] [PubMed]
  12. Wen, W.; Wu, L.; Yu, T. Excitonic lasers in atomically thin 2D semiconductors. ACS Mater. Lett. 2020, 2, 1328–1342. [Google Scholar] [CrossRef]
  13. Ahmad, W.; Gong, Y.; Abbas, G.; Khan, K.; Khan, M.; Ali, G.; Shuja, A.; Tareen, A.K.; Khan, Q.; Li, D. Evolution of low-dimensional material-based field-effect transistors. Nanoscale 2021, 13, 5162–5186. [Google Scholar] [PubMed]
  14. Ahsan, M.A.; He, T.; Noveron, J.C.; Reuter, K.; Puente-Santiago, A.R.; Luque, R. Low-dimensional heterostructures for advanced electrocatalysis: An experimental and computational perspective. Chem. Soc. Rev. 2022, 51, 812–828. [Google Scholar] [CrossRef] [PubMed]
  15. Kanade, C.K.; Seok, H.; Kanade, V.K.; Aydin, K.; Kim, H.-U.; Mitta, S.B.; Yoo, W.J.; Kim, T. Low-temperature and large-scale production of a transition metal sulfide vertical heterostructure and its application for photodetectors. ACS Appl. Mater. Interfaces 2021, 13, 8710–8717. [Google Scholar] [PubMed]
  16. Do, T.-N.; Idrees, M.; Amin, B.; Hieu, N.N.; Phuc, H.V.; Hoa, L.T.; Nguyen, C.V. First principles study of structural, optoelectronic and photocatalytic properties of SnS, SnSe monolayers and their van der Waals heterostructure. Chem. Phys. 2020, 539, 110939. [Google Scholar] [CrossRef]
  17. Kim, S.E.; Mujid, F.; Rai, A.; Eriksson, F.; Suh, J.; Poddar, P.; Ray, A.; Park, C.; Fransson, E.; Zhong, Y. Extremely anisotropic van der Waals thermal conductors. Nature 2021, 597, 660–665. [Google Scholar] [CrossRef]
  18. Li, F.; Feng, Y.; Li, Z.; Ma, C.; Qu, J.; Wu, X.; Li, D.; Zhang, X.; Yang, T.; He, Y.; et al. Rational kinetics control toward universal growth of 2D vertically stacked heterostructures. Adv. Mater. 2019, 31, 1901351. [Google Scholar]
  19. Halim, N.D.; Zaini, M.S.; Talib, Z.A.; Liew, J.Y.C.; Kamarudin, M.A. Study of the electron-phonon coupling in PbS/MnTe quantum dots based on temperature-dependent photoluminescence. Micromachines 2022, 13, 443. [Google Scholar]
  20. Vaquero, D.; Salvador-Sánchez, J.; Clericò, V.; Diez, E.; Quereda, J. The low-temperature photocurrent spectrum of monolayer MoSe2: Excitonic features and gate voltage dependence. Nanomaterials 2022, 12, 322. [Google Scholar] [CrossRef]
  21. Yamada, Y.; Yoshimura, T.; Ashida, A.; Fujimura, N.; Kiriya, D. Strong photoluminescence enhancement from bilayer molybdenum disulfide via the combination of UV irradiation and superacid molecular treatment. Appl. Sci. 2021, 11, 3530. [Google Scholar] [CrossRef]
  22. Zhao, K.; He, D.; Fu, S.; Bai, Z.; Miao, Q.; Huang, M.; Wang, Y.; Zhang, X. Interfacial coupling and modulation of van der Waals heterostructures for nanodevices. Nanomaterials 2022, 12, 3418. [Google Scholar] [CrossRef] [PubMed]
  23. Wu, H.; Lin, M.-L.; Leng, Y.-C.; Chen, X.; Zhou, Y.; Zhang, J.; Tan, P.-H. Probing the interfacial coupling in ternary van der Waals heterostructures. NPJ 2D Mater. Appl. 2022, 6, 87. [Google Scholar]
  24. Kumar, A.; Kumar, V.; Romeo, A.; Wiemer, C.; Mariotto, G. Raman spectroscopy and in situ XRD probing of the thermal decomposition of Sb2Se3 thin films. J. Phys. Chem. C 2021, 125, 19858–19865. [Google Scholar] [CrossRef]
  25. Ono, R.; Imai, S.; Kusama, Y.; Hamada, T.; Hamada, M.; Muneta, I.; Kakushima, K.; Tsutsui, K.; Kano, E.; Ikarashi, N. Elucidation of PVD MoS2 film formation process and its structure focusing on sub-monolayer region. Jpn. J. Appl. Phys. 2022, 61, 1023. [Google Scholar] [CrossRef]
  26. Arafat, A.; Islam, M.S.; Ferdous, N.; Islam, A.J.; Sarkar, M.M.H.; Stampfl, C.; Park, J. Atomistic reaction mechanism of CVD grown MoS2 through MoO3 and H2S precursors. Sci. Rep. 2022, 12, 16085. [Google Scholar]
  27. Teli, A.; Beknalkar, S.; Mane, S.; Bhat, T.; Kamble, B.; Patil, S.; Sadale, S.; Shin, J. Electrodeposited crumpled MoS2 nanoflakes for asymmetric supercapacitor. Ceram. Int. 2022, 48, 29002–29010. [Google Scholar] [CrossRef]
  28. Zheng, X.; Zhu, Y.; Sun, Y.; Jiao, Q. Hydrothermal synthesis of MoS2 with different morphology and its performance in thermal battery. J. Power Sources 2018, 395, 318–327. [Google Scholar]
  29. Huang, M.; Ali, W.; Yang, L.; Huang, J.; Yao, C.; Xie, Y.; Sun, R.; Zhu, C.; Tan, Y.; Liu, X. Multifunctional optoelectronic synapses based on arrayed MoS2 monolayers emulating human association memory. Adv. Sci. 2023, 10, 2300120. [Google Scholar] [CrossRef]
  30. Jia, H.; Yang, R.; Nguyen, A.E.; Alvillar, S.N.; Empante, T.; Bartels, L.; Feng, P.X.-L. Large-scale arrays of single-and few-layer MoS2 nanomechanical resonators. Nanoscale 2016, 8, 10677–10685. [Google Scholar] [CrossRef]
  31. Li, H.; Zhang, Q.; Yap, C.C.R.; Tay, B.K.; Edwin, T.H.T.; Olivier, A.; Baillargeat, D. From bulk to monolayer MoS2: Evolution of Raman scattering. Adv. Funct. Mater. 2012, 22, 1385–1390. [Google Scholar] [CrossRef]
  32. Zeng, H.; Liu, G.-B.; Dai, J.; Yan, Y.; Zhu, B.; He, R.; Xie, L.; Xu, S.; Chen, X.; Yao, W. Optical signature of symmetry variations and spin-valley coupling in atomically thin tungsten dichalcogenides. Sci. Rep. 2013, 3, 1608. [Google Scholar] [CrossRef]
  33. Ramasubramaniam, A. Large excitonic effects in monolayers of molybdenum and tungsten dichalcogenides. Phys. Rev. B 2012, 86, 115409. [Google Scholar] [CrossRef]
  34. Jelken, J.; Lambin, C.D.; Avilés, M.A.O.; Lagugné-Labarthet, F. Real-time observation of photo-oxidation of single MoS2 flakes using stochastic optical reconstruction microscopy. J. Phys. Chem. C 2023, 127, 14270–14282. [Google Scholar]
  35. Di Bartolomeo, A.; Kumar, A.; Durante, O.; Sessa, A.; Faella, E.; Viscardi, L.; Intonti, K.; Giubileo, F.; Martucciello, N.; Romano, P. Temperature-dependent photoconductivity in two-dimensional MoS2 transistors. Mater. Today Nano 2023, 100382. [Google Scholar] [CrossRef]
  36. Wu, Z.Y.; Zhuang, J.-H.; Lin, Y.-T.; Chou, Y.-H.; Wu, P.C.; Wu, C.-L.; Chen, P.; Hsu, H.-C. One-and two-photon excited photoluminescence and suppression of thermal quenching of CsSnBr3 microsquare and micropyramid. ACS Nano 2021, 15, 19613–19620. [Google Scholar] [CrossRef]
  37. Chen, M.; Zhou, B.; Wang, F.; Xu, L.; Jiang, K.; Shang, L.; Hu, Z.; Chu, J. Interlayer coupling and the phase transition mechanism of stacked MoS2/TaS2 heterostructures discovered using temperature dependent Raman and photoluminescence spectroscopy. RSC Adv. 2018, 8, 21968–21974. [Google Scholar] [CrossRef] [PubMed]
  38. Varshni, Y.P. Temperature dependence of the energy gap in semiconductors. Physica 1967, 34, 149–154. [Google Scholar] [CrossRef]
  39. O’donnell, K.P.; Chen, X. Temperature dependence of semiconductor band gaps. Appl. Phys. Lett. 1991, 58, 2924–2926. [Google Scholar] [CrossRef]
  40. Hu, Z.; Bao, Y.; Li, Z.; Gong, Y.; Feng, R.; Xiao, Y.; Wu, X.; Zhang, Z.; Zhu, X.; Ajayan, P.M. Temperature dependent Raman and photoluminescence of vertical WS2/MoS2 monolayer heterostructures. Sci. Bull. 2017, 62, 16–21. [Google Scholar] [CrossRef] [PubMed]
  41. Liu, B.; Fathi, M.; Chen, L.; Abbas, A.; Ma, Y.; Zhou, C. Chemical vapor deposition growth of monolayer WSe2 with tunable device characteristics and growth mechanism study. ACS Nano 2015, 9, 6119–6127. [Google Scholar] [PubMed]
  42. Yao, Z.; Liu, J.; Xu, K.; Chow, E.K.; Zhu, W. Material synthesis and device aspects of monolayer tungsten diselenide. Sci. Rep. 2018, 8, 5221. [Google Scholar] [CrossRef] [PubMed]
  43. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar]
  44. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef] [PubMed]
  45. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  46. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. Chem. Phys. 2010, 15, 132. [Google Scholar]
  47. Li, Y.; Li, Y.-L.; Araujo, C.M.; Luo, W.; Ahuja, R. Single-layer MoS2 as an efficient photocatalyst. Catal. Sci. Technol. 2013, 3, 2214–2220. [Google Scholar]
  48. Zhang, L.; Huang, L.; Yin, T.; Yang, Y. Strain-induced tunable band offsets in blue phosphorus and WSe2 van der Waals heterostructure. Crystals 2021, 11, 470. [Google Scholar]
  49. Su, X.; Ju, W.; Zhang, R.; Guo, C.; Zheng, J.; Yong, Y.; Li, X. Bandgap engineering of MoS2/MX2 (MX2 = WS2, MoSe2 and WSe2) heterobilayers subjected to biaxial strain and normal compressive strain. RSC Adv. 2016, 6, 18319–18325. [Google Scholar]
  50. 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]
Figure 1. Structure and morphology of MoS2/WSe2 heterostructure. (a) Schematic of MoS2 and WSe2 monolayers. (b) SEM and (c) optical images of the vertically stacked heterostructure with scale bars of 10 and 9 µm, respectively. (d) Height profiles of MoS2 and WSe2 monolayers confirmed by AFM.
Figure 1. Structure and morphology of MoS2/WSe2 heterostructure. (a) Schematic of MoS2 and WSe2 monolayers. (b) SEM and (c) optical images of the vertically stacked heterostructure with scale bars of 10 and 9 µm, respectively. (d) Height profiles of MoS2 and WSe2 monolayers confirmed by AFM.
Nanomaterials 13 02349 g001
Figure 2. Raman and PL spectra of MoS2/WSe2 heterostructures. (a) Raman shifts of MoS2 and WSe2 monolayers and their heterostructures. (b) PL spectra of MoS2 and WSe2 monolayers with the main peaks at 655 nm and 758 nm, respectively. Mapping images of PL intensity of heterostructure samples at a wavelength of (c) 758 nm and (d) 655 nm.
Figure 2. Raman and PL spectra of MoS2/WSe2 heterostructures. (a) Raman shifts of MoS2 and WSe2 monolayers and their heterostructures. (b) PL spectra of MoS2 and WSe2 monolayers with the main peaks at 655 nm and 758 nm, respectively. Mapping images of PL intensity of heterostructure samples at a wavelength of (c) 758 nm and (d) 655 nm.
Nanomaterials 13 02349 g002
Figure 3. Temperature-dependent Raman spectra of MoS2/WSe2 heterostructures. (a,b) The variation of Raman modes of MoS2 and WSe2 at varying temperatures of 79–473 K. The vertically black and orange dashed lines are drawn to highlight specific Raman modes and indicate temperature increase, respectively. (c) Temperature-induced Raman shifts of three characteristic peaks showing linear fitting curves. (d) The full width at half maximum of Raman peaks at different temperatures.
Figure 3. Temperature-dependent Raman spectra of MoS2/WSe2 heterostructures. (a,b) The variation of Raman modes of MoS2 and WSe2 at varying temperatures of 79–473 K. The vertically black and orange dashed lines are drawn to highlight specific Raman modes and indicate temperature increase, respectively. (c) Temperature-induced Raman shifts of three characteristic peaks showing linear fitting curves. (d) The full width at half maximum of Raman peaks at different temperatures.
Nanomaterials 13 02349 g003
Figure 4. Temperature-dependent photoluminescence. (a,b) PL spectra of MoS2 and WSe2 at various temperatures. The peak position shifts red, and the intensity decreases as temperature increases form 79 to 473 K. (c,d) Integrated PL intensities of corresponding regions of MoS2 and WSe2, respectively. Solid lines depict experimental data, while the scattered points show the corresponding Arrhenius fitting. (e,f) Temperature-dependent shifting of PL peaks detected for MoS2 and WSe2 regions in the heterostructures. Red lines and dotted blue lines represent fitting curves for Equations (2) and (3).
Figure 4. Temperature-dependent photoluminescence. (a,b) PL spectra of MoS2 and WSe2 at various temperatures. The peak position shifts red, and the intensity decreases as temperature increases form 79 to 473 K. (c,d) Integrated PL intensities of corresponding regions of MoS2 and WSe2, respectively. Solid lines depict experimental data, while the scattered points show the corresponding Arrhenius fitting. (e,f) Temperature-dependent shifting of PL peaks detected for MoS2 and WSe2 regions in the heterostructures. Red lines and dotted blue lines represent fitting curves for Equations (2) and (3).
Nanomaterials 13 02349 g004
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

Ali, W.; Liu, Y.; Huang, M.; Xie, Y.; Li, Z. Temperature-Dependent Phonon Scattering and Photoluminescence in Vertical MoS2/WSe2 Heterostructures. Nanomaterials 2023, 13, 2349. https://doi.org/10.3390/nano13162349

AMA Style

Ali W, Liu Y, Huang M, Xie Y, Li Z. Temperature-Dependent Phonon Scattering and Photoluminescence in Vertical MoS2/WSe2 Heterostructures. Nanomaterials. 2023; 13(16):2349. https://doi.org/10.3390/nano13162349

Chicago/Turabian Style

Ali, Wajid, Ye Liu, Ming Huang, Yunfei Xie, and Ziwei Li. 2023. "Temperature-Dependent Phonon Scattering and Photoluminescence in Vertical MoS2/WSe2 Heterostructures" Nanomaterials 13, no. 16: 2349. https://doi.org/10.3390/nano13162349

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