Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Valley-polarized local excitons in WSe2/WS2 vertical heterostructures

Open Access Open Access

Abstract

Two-dimensional van der Waals heterostructures (vdWHs) are drawing growing interest in the investigation of their valley polarization properties of localized excitons. However, most of the reported vdWHs were made by micro-mechanical peeling, limiting their large-scale production and practical applications. Furthermore, the circular polarization characters of localized excitons in WSe2/WS2 heterostructures remain elusive. Here, a bidirectional-flow physical vapor deposition technique was employed for the synthesis of the WSe2/WS2 type-II vertical heterostructures. The interfaces of such heterojunctions are sharp and clean, making the neutral excitons of the constituent layers quenched, which significantly highlights the luminescence of the local excitons. The circular polarization of localized excitons in this WSe2/WS2 heterostructure was demonstrated by circularly-polarized PL spectroscopy. The degree of the circular polarization of the localized excitons was determined as 7.17% for σ- detection and 4.78% for σ+ detection. Such local excitons play a critical role in a quantum emitter with enhanced spontaneous emission rate that could lead to the evolution of LEDs. Our observations provide valuable information for the exploration of intriguing excitonic physics and the applications of innovative local exciton devices.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Two-dimensional (2D) materials with their unique properties originating from the quantum confinement offer more development choices for modern electronic devices [16]. Up to now, the vdWHs have been evolved into several functional devices such as field-effect transistors [7], high-sensitivity chemical sensors [8], and photodetectors with high performance [911]. Notably, the van der Waals interlayer interaction provides the vdWHs with more novel characteristics than that of the monolayer 2D materials [1216]. One of the new characteristics is the remarkable photoluminescence (PL) quenching phenomenon from neutral excitons occurring in type-II heterojunctions as a result of enhanced PL from the local excitons. So vdWHs provide a unique platform for studying the circular polarization properties of localized excitons [1722]. The present work demonstrates a way of capture and study of the distinct circular polarization properties of the vdWHs [23].

To date, most of the reported vdWHs were made with 2D materials by micro-mechanical peeling, a process limiting their large-scale production and practical applications. Meanwhile, the observation and investigation of circular polarization of local excitons in WSe2/WS2 heterostructures have been challenging. The physical vapor deposition method avoids the above problems because of the closed preparation environment with the protection of inert gas. With the clean interface and high quality, the physical vapor deposition products are often utilized in the excitons research and non-linear optics [2428]. In this article, we demonstrate a bidirectional-flow physical vapor deposition method for growing vertical WSe2/WS2 heterostructures in large quantities. During the preparation process, a reverse carrier flow was used for cooling to reduce undesirable nucleation during the growth of the WSe2/WS2 heterostructures. The sharp and clean interfaces in such heterostructures enable the photoluminescence of neutral excitons in the constituent monolayers to be quenched significantly, creating more local excitons. Additionally, the existence of the circular polarization of local excitons in the WSe2/WS2 heterostructures was successfully proved with the polarization degrees of 7.17% for σ- detection and 4.78% for σ+ detection. In a recent study, a single-photon emission from the localized excitons was observed in monolayer TMDC materials [29]. Local excitons play a critical role in quantum emitters with enhanced spontaneous emission rates [30,31]. The enhancement of the spontaneous radiative emission is crucial for the design of LEDs [32]. The present study of localized excitons in vdWHs may provide insight for the development of new LEDs [33].

The high-quality WSe2/WS2 heterostructures were synthesized by a two-step physical vapor deposition method. In previous experiments, the mechanical transfer method was widely used. While the mechanical transfer method is more primitive, during the transfer process, the nanosheets could be lightly damaged, wrinkled, blistered, and contaminated with organic matter. Therefore, a large number of defects would be introduced. Because of the high density of defect states, multiple defect exciton peaks could be detected within a narrow band range in the same PL spectrum, leading to a rather complex spectrum for the identification and investigation of the local excitons. Also, the density of native defects in the CVD-grown product is much higher than the mechanical exfoliated product, which is not beneficial for the identification of excitons [34]. The physical vapor deposition method, however, has less control over site specific growth as the nucleation spots are rather random. It is also more costly and time consuming compared with the mechanical transfer method.

By contrast, the physical vapor deposition method is efficient in producing high-quality vdWHs with uniformity and few defects. High-quality products also make it possible to study a single defect. Moreover, compared to traditional preparation, the physical vapor deposition method can yield more heterostructures. As shown in Fig. 1(a), the vertical heterostructures were synthesized in a tube furnace (OTF-1200X-II KJ Group) using a physical vapor deposition system at atmospheric pressure. In the traditional CVD growth process, thermal degradation and undesired nucleation are the critical barriers for stable formation of heterostructures [35,36]. A tube furnace with an upgraded pneumatic system could solve these problems [37]. For the first step, the WS2 powder was placed in an aluminum oxide boat, which was propelled into the heating zone of a 1-inch quart tube. Then a piece of polished SiO2 (285nm)/Si was placed (facing up) in another boat at the downstream position as the growth substrate. The center zone of the furnace was heated to 1160 ℃ under atmospheric pressure with a reverse flow of 250 sccm argon gas. When the target temperature was reached, the WS2 vapor source was carried downstream by a forward flow of 100 sccm argon gas for a growth period of 2 minutes. After the WS2 growth, the furnace cooled down naturally with an inverse argon flow of 250 sccm. The second step was the epitaxial growth of WSe2 on top of WS2. The SiO2/Si substrate with the as-grown WS2 nanosheets was placed at the same position in another similar physical vapor deposition system for the WSe2 growth. During the temperature-increasing stage, a reverse flow of 250 sccm cold argon gas cooled the existing WS2 nanosheets to restrain thermal degradation. The center zone of the furnace was heated to 1100 ℃ under atmospheric pressure. After reaching the desired temperature for WSe2, the inverse Ar gas was switched to a 110 sccm forward flow to carry the vaporized WSe2 source to the substrate. Then growth temperature was maintained for 3 minutes for the epitaxial growth of WSe2. Finally, the substrate naturally cooled down to room temperature with the inverse cold argon flow. The growth process of both films was schematically shown in Fig. 1(b), where the blue triangular nanoplate is WS2, and the smaller triangular nanoplate on top of WS2 is WSe2. Due to the different growth temperatures of TMDCs, the order of heterostructure grown by physical vapor deposition method is fixed. The prepared WSe2/WS2 vertical heterostructures are shown in Fig. 1(c). As can be seen, each heterostructure displays two distinct regions of different contrasts. The central triangular region of darker contact is WSe2, sitting on top of a larger triangular region of lighter contrast.

 figure: Fig. 1.

Fig. 1. Growth of the WSe2/WS2 heterostructures. (a) Schematic diagram of the tube furnace used for the growth of the WSe2/WS2 heterostructures in the experiment. The arrows in different colors show two opposite directions of flow, with the green being the forward flow. (b) Schematic diagram of the physical vapor deposition method for the growth of the heterostructures. The top of the diagram shows the deposition process of molecules for the WSe2 layer, while the bottom of the diagram indicates the process of the epitaxial growth of WSe2 on top of the WS2 layer. (c) Optical images of the WSe2/WS2 heterostructures showing the smaller triangular-shaped WSe2 on top of the relatively larger triangular-shaped WS2.

Download Full Size | PDF

Raman and PL studies were conducted at a low-temperature using a WITec Alpha 300R Confocal Raman System to confirm the types of excitons and the layer number of the WSe2 and WS2. A 532 nm laser was applied to irradiate the heterostructure to obtain Raman and PL spectra and mapping images. As shown in Fig. 2(a), the WS2 Raman spectrum matches well with the WS2 spectrum of the WSe2/WS2 heterostructures. There were minor Raman shifts between that of WSe2 and the heterostructure for both the characteristic peaks of out-of-plane A1g and in-plane E2g1 modes as a result of the coupling between WS2 and WSe2. The frequency difference between the A1g and the E2g1 modes of WSe2 is consistent with the splitting of 11-12 cm-1 in the monolayer WSe2. These features in the Raman spectra provide evidence that the heterostructure is composed of monolayer WSe2 and WS2 [38,39]. The optical microscope images (insets of Figs. 2(c)-(e)) of synthesized WSe2/WS2 heterostructures on the SiO2/Si substrate reveal the two concentric triangular regions of WS2 and WSe2 with slightly different optical contrasts. The red regions in the spatially-resolved Raman maps are WS2 obtained with the A1g peak of WS2 at 250 cm-1, while the green regions are WSe2 generated with the E2g peak of WSe2 at 350 cm-1. In the Raman mapping images, the heterostructures show distinct boundary and uniform colors in both the overlapping and non-overlapping regions, demonstrating the high quality of the physical vapor deposition product and the advantages of the bidirectional-flow physical vapor deposition method that enabled the subsequent study on the excitons from these heterostructures.

 figure: Fig. 2.

Fig. 2. Raman mapping and photoluminescence quenching in the WSe2/WS2 heterostructures. (a) Raman spectra of WSe2 and WS2 and their heterostructure. The red line is a spectrum from the WSe2/WS2, and the blue line is the WS2 spectrum obtained from the edge of the heterostructure where there was no WSe2 present. As a comparison, the Raman spectrum (black line) from a monolayer WSe2 sample is also presented. (b) PL spectra of WSe2, WS2, and their heterostructure. (c)-(e) Raman mapping images of the WSe2/WS2 heterostructures. Mapped with their characteristic peaks, the green regions are WSe2, and the red regions are WS2. The insets are optical images of the WSe2/WS2 heterostructures. (f)-(h) PL mapping images of the WSe2/WS2 heterostructures showing photoluminescence quenching resulting from the band-structure of the type-II heterostructure. The insets in (c)-(h) are the optical images of the WSe2/WS2 heterostructures.

Download Full Size | PDF

As previously reported, the type of excitons in the TMDC heterostructures depends on the band alignment of the heterostructures. To probe the alignment of the band structure in the WSe2/WS2 heterostructure in more detail, the heterostructure was investigated by PL spectra (Fig. 2(b)). The PL peak position is to locate at 1.87 eV for WS2, corresponding to the bandgap of 2-3 atomic layers of WS2, while the PL peak is at 1.65 eV for WSe2. Both these PL peaks were observed in the heterostructure, but they were much weaker than that measured in the separate monolayers of WS2 and WSe2. The PL intensities of the heterostructure were quenched by a factor of ∼500-1000. As shown in the PL mapping images of Figs. 2(f)-(h), while the PL intensity was strong and uniform outside the heterostructure region, the PL intensity was too weak to be revealed in the heterostructure region, indicating the type-II band alignment of the WSe2/WS2 heterostructures. For monolayer band structures, carriers are generated under optical excitation and consumed on account of the radiative recombination. Interestingly, in WSe2/WS2 heterostructures, the charge carriers are transferred between the WSe2 and WS2 layers induced by the particular type-II band alignment. The hole and electron carriers are separated into different layers, so the PL intensities of intralayer excitons significantly decrease. This dramatic quenching proves that the heterostructures synthesized by the physical vapor deposition method have a sharp interface that helps the transfer of carriers and highlights the defect excitons of the monolayer materials. As a contrast, the neutral excitons in the WSe2/WS2 heterostructures prepared through chemical vapor deposition and assembling by a PMMA stamping method were not completely quenched because of the contaminated interfaces. There is another possible way to induce the observed quench phenomenon. A direct-to-indirect bandgap transition also could block the recombination of carriers to quench PL of the WSe2 and WS2 nanosheets. However, in the previous report, researchers investigated the origin of PL quench in WSe2/WS2 heterostructures and conluded that such an assumption of bandgap transition is highly unlikely experimentally [27]. The observed local excitons in the current study were due to the significant suppression of the PL intensity of neutral excitons as a result of the band alignment in the WSe2/WS2 heterostructures.

For a deeper understanding of the localized excitons, we further studied the PL properties of the WSe2/WS2 heterostructures under a low-temperature environment (65K-300K). Figure 3 displays the temperature-dependent PL spectra in the different areas of WSe2/WS2 nanoplates. In the WS2 area, there are two emission peaks corresponding to charged exciton peak (X-) and neutral exciton peak (X0) in Fig. 3(a). As the temperature rose, the peak broadened, and the peak position shifted to the low-energy region. The following Varshni equation is commonly used to describe the relationship between the PL peak position and temperature in semiconductors [40]:

$${E_g}(T) = {E_g}(0) - \alpha {T^2}\textrm{/}(T + \beta ), $$

where α and β are fit parameters and ${E_g}(0)$ is the bandgap at absolute zero temperature. The Varshni equation was applied to fit the trend of PL peak positions with the changing of temperature, as shown in Fig. 3(c). For the charged excitons in the WS2 region, the calculated fit parameters are $\alpha \textrm{ = 3}\textrm{.86} \times \textrm{1}{\textrm{0}^{\textrm{ - 4}}}eV/K$, $\beta ={-} 34K$ and ${E_g}(0 )= 2.045\textrm{ }eV$. In Fig. 3(c), the solid red curve shows the fitting result. The calculated parameters are similar to those of a previous study [41]. As the temperature decreased, the peak width of the PL spectrum gradually narrowed, making the identification of the localized excitons easier. Compared with regular excitons, the electrons of the localized excitons were trapped in the bandgap by a potential well (Fig. 3(d)), so the localized excitons (Fig. 3(b)) were induced by defects located below the X- exciton peak of WS2 in the heterostructure region. The interaction between atomic-scale defects and excitons occurred noticeably at atmospheric pressure. Therefore the defect-related peaks can be evolved in the PL spectrum. The defects were induced by the vacancies of sulfur atoms stemming from the oxidation of WS2 molecules after preparation. The current observation and results are consistent with a previous report [42]. Remarkably, the defects of WS2 in such heterojunctions had only one peak, meaning that only a single defect state was introduced, and this is also the advantage of fabricating samples by the physical vapor deposition technique. More crucially, the recognition of local excitons in such heterostructures enabled the study of circular polarization.

 figure: Fig. 3.

Fig. 3. Normalized photoluminescence spectra under variable temperature. (a) The PL spectra of the WS2 region under different temperatures (65K-300K). The dashed line indicates the trends of peak shifting. (b) The PL spectra of the localized excitons in the WSe2/WS2 heterostructures under different temperatures (65K-300K). (c) X- Peak energy as a function of temperature. The solid line indicates a fit using the Varshni equation. (d) Energy level diagram of WS2/WSe2 heterostructure.

Download Full Size | PDF

To better understand the circular polarization of localized excitons, the polarization-resolved photoluminescence spectra of the WSe2/WS2 heterostructures were investigated at T = 65K. In Figs. 4(a)-(d), the circular polarization of localized excitons was observed evidently. For instance, the position of the peak near 1.9 eV revealed that there were localized excitons of WS2 [43]. The degree of the PL polarization can be quantified by the helicity:

$$P = \frac{{{I_{{\sigma ^ - }}} - {I_{{\sigma ^ + }}}}}{{{I_{{\sigma ^ - }}} + {I_{{\sigma ^ + }}}}}$$
where ${I_{{\sigma ^ - }}}$ and ${I_{{\sigma ^ + }}}$ are the peak intensities of the left circularly polarized (σ-) emission light and the right circularly polarized (σ+) emission light.

The four polarization-resolved PL spectra in the heterostructure regions for two samples (sample 1 for Figs. 4(a) and (b); sample 2 for Figs. 4(c) and (d)) reveal that the different degrees of PL valley-polarized emissions at 1.9 eV of the WS2 localized excitons are 7.17%, 4.78%, 6.13%, and -0.39%, respectively. The difference between σ+ and σ- light intensities originated from the valley optical selection rule [4446]. When the sample was excited by σ- light, carriers were generated at − K valley and recombined to emit σ- light, while no carriers were generated at + K valley. However, there was considerable emission of σ+ light at + K valley because of the intervalley scattering (top panel of Fig. 4(e)). On the contrary, only carriers at + K valley were excited when the incident light was σ+ light (bottom panel of Fig. 4(e)).

 figure: Fig. 4.

Fig. 4. Valley polarization of localized excitons. (a) and (c) Polarization-resolved PL spectra for the WSe2/WS2 heterostructures at 65K using σ- excitation. (b) and (d) Polarization-resolved PL spectra for the WSe2/WS2 heterostructures at 65K using σ+ excitation. In (a) and (b), the red line represents the intensity of left-handed light emitted by the heterostructures. The blue line represents the intensity of right-handed light emitted by the heterostructures. The olive-green dotted line indicates the corresponding degree of PL polarization. (e) Schematic diagram of optical selection rules. (f) Schematic diagram of the optical selection rules of the local excitons in WS2. The purple lines indicate defect energy levels.

Download Full Size | PDF

For sample 1, it exhibited stronger emission of σ- light than σ+ light when excited by σ- light (Fig. 4(a)). When excited by σ+ light, the situation was opposite (Fig. 4(b)). As can be seen in Fig. 4(f), one defect energy level exists below the bottom of the conduction band, containing two quasi-degenerated spin-down and spin-up electronic states [47,48]. The transition between the valence band and the defect energy level holds the same optical selection rules as the transition between the valence band and conduction band at the ±K valley [49]. When the incident light was σ- light, carriers generated at -K valley, and then the electrons were trapped at the defect energy level. The electrons localized at the defect level and holes localized in valence band combined radiatively, emitting σ- light. Likewise, the emission of σ+ light was also considerable because of the intervalley scattering(top panel of Fig. 4(f)). The opposite situation can be detected when the sample was excited by σ+ light (bottom panel of Fig. 4(f)). The eventual degrees of polarization were both positive for sample 1. This phenomenon occurred in five of the eight samples. However, there was also another situation, as shown in sample 2 (Figs. 4(c) and 4(d)). When the incident light was σ+ light, the degree of polarization was negative. For sample 2, under the excitation of σ- light, the σ- light emission exhibited more vigorous intensity than that of the σ+ light. Thus, the polarization degree was positive. While the sample was excited by σ+ light, the σ- light emission still exhibited stronger intensity than that of σ+ light and showed a negative polarization degree. The polarization degrees under the excitation of σ- and σ+ lights showed opposite signs. The possible reason might relate to the contributions from multiple excitonic bands and the spin flipping during the charge trapping process [50]. The difference between the two situations originated from a distinct lattice mismatch and the different contributions from excitonic bands. However, the underlying physical principles behind the above phenomenon remain to be further explored.

In summary, we report the synthesis of 2D TMDC heterostructures by a bidirectional-flow physical vapor deposition method. The optical properties of the WSe2/WS2 heterostructures were characterized. The PL quenching phenomenon was distinctly observed and could be attributed to the type-II band alignment. The circular polarization of localized excitons in the heterostructure region was studied under low-temperature conditions, and the trends of PL spectra of the charged excitons in WS2 are in a good agreement with the Varshni’s theory. The novel phenomenon of local excitons needs to be further explored in combination with theoretical calculations. The present work not only helps to understand the behaviors of localized excitons in nano-scale heterostructures but may provide insight towards quantum emitters in TMDC materials.

Funding

National Natural Science Foundation of China (61775241); Central South University (Youth Innovation Team (2019012)); Science, Technology and Innovation Commission of Shenzhen Municipality (JCYJ20180307151237242); Hunan Provincial Science Fund for Distinguished Young Scholars (2020JJ2059); Key Research and Development Project of Hunan Province (2019GK2233); State Key Laboratory of High Performance Complex Manufacturing (ZZYJKT2020-12).

Disclosures

The authors declare that they have no competing interests.

References

1. Y. Ren, Z. Qiao, and Q. Niu, “Topological phases in two-dimensional materials: a review,” Rep. Prog. Phys. 79(6), 066501 (2016). [CrossRef]  

2. Y. Li, Y.-L. Li, B. Sa, and R. Ahuja, “Review of two-dimensional materials for photocatalytic water splitting from a theoretical perspective,” Catal. Sci. Technol. 7(3), 545–559 (2017). [CrossRef]  

3. S. Das, J. A. Robinson, M. Dubey, H. Terrones, and M. Terrones, “Beyond graphene: progress in novel two-dimensional materials and van der Waals solids,” Annu. Rev. Mater. Res. 45(1), 1–27 (2015). [CrossRef]  

4. Y. Liu, C. Zeng, J. Zhong, J. Ding, Z. M. Wang, and Z. Liu, “Spintronics in Two-Dimensional Materials,” Nano-Micro Lett. 12(1), 93 (2020). [CrossRef]  

5. Y. Liu, Y. Gao, S. Zhang, J. He, J. Yu, and Z. Liu, “Valleytronics in transition metal dichalcogenides materials,” Nano Res. 12(11), 2695–2711 (2019). [CrossRef]  

6. J. Yu, X. Kuang, Y. Gao, Y. Wang, K. Chen, Z. Ding, J. Liu, C. Cong, J. He, Z. Liu, and Y. Liu, “Direct Observation of the Linear Dichroism Transition in Two-Dimensional Palladium Diselenide,” Nano Lett. 20(2), 1172–1182 (2020). [CrossRef]  

7. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol. 6(3), 147–150 (2011). [CrossRef]  

8. X. Zhao, T. Huang, S. P. Ping, X. Wu, P. Huang, J. Pan, Y. Wu, and Z. Cheng, “Sensitivity Enhancement in Surface Plasmon Resonance Biochemical Sensor Based on Transition Metal Dichalcogenides/Graphene Heterostructure,” Sensors 18, 2056 (2018). [CrossRef]  

9. W. Yu, S. Li, Y. Zhang, W. Ma, T. Sun, J. Yuan, K. Fu, and Q. Bao, “Near-Infrared Photodetectors Based on MoTe2/Graphene Heterostructure with High Responsivity and Flexibility,” Small 13(24), 1700268 (2017). [CrossRef]  

10. J. Yu, J. Zhong, X. Kuang, C. Zeng, L. Cao, Y. Liu, and Z. Liu, “Dynamic Control of High-Range Photoresponsivity in a Graphene Nanoribbon Photodetector,” Nanoscale Res. Lett. 15(1), 124 (2020). [CrossRef]  

11. J. Zhong, J. Yu, L. Cao, C. Zeng, J. Ding, C. Cong, Z. Liu, and Y. Liu, “High-performance polarization-sensitive photodetector based on a few-layered PdSe2 nanosheet,” Nano Res. 13(6), 1780–1786 (2020). [CrossRef]  

12. A. K. Geim and I. V. Grigorieva, “Van der Waals heterostructures,” Nature 499(7459), 419–425 (2013). [CrossRef]  

13. Y. Liu, Y. Huang, and X. F. Duan, “Van der Waals integration before and beyond two-dimensional materials,” Nature 567(7748), 323–333 (2019). [CrossRef]  

14. P. Liu and B. Xiang, “2D heterostructures based on transition metal dichalcogenides: fabrication, properties and applications,” Sci. Bull. 62(16), 1148–1161 (2017). [CrossRef]  

15. J. Yu, X. Kuang, J. Zhong, L. Cao, C. Zeng, J. Ding, C. Cong, S. Wang, P. Dai, X. Yue, Z. Liu, and Y. Liu, “Observation of double indirect interlayer exciton in WSe2/WS2 heterostructure,” Opt. Express 28(9), 13260–13268 (2020). [CrossRef]  

16. Y. Liu, S. Zhang, J. He, Z. M. Wang, and Z. Liu, “Recent Progress in the Fabrication, Properties, and Devices of Heterostructures Based on 2D Materials,” Nano-Micro Lett. 11(1), 13 (2019). [CrossRef]  

17. F. Ceballos, M. Z. Bellus, H. Y. Chiu, and H. Zhao, “Ultrafast charge separation and indirect exciton formation in a MoS2-MoSe2 van der Waals heterostructure,” ACS Nano 8(12), 12717–12724 (2014). [CrossRef]  

18. J. S. Ross, P. Rivera, J. Schaibley, E. Lee-Wong, H. Yu, T. Taniguchi, K. Watanabe, J. Yan, D. Mandrus, and D. Cobden, “Interlayer exciton optoelectronics in a 2D heterostructure p–n junction,” Nano Lett. 17(2), 638–643 (2017). [CrossRef]  

19. S. Gao, L. Yang, and C. D. Spataru, “Interlayer coupling and gate-tunable excitons in transition metal dichalcogenide heterostructures,” Nano Lett. 17(12), 7809–7813 (2017). [CrossRef]  

20. A. Boulesbaa, K. Wang, M. Mahjouri-Samani, M. Tian, A. A. Puretzky, I. Ivanov, C. M. Rouleau, K. Xiao, B. G. Sumpter, B. G. Sumpter, and D. B. Geohegan, “Ultrafast charge transfer and hybrid exciton formation in 2D/0D heterostructures,” J. Am. Chem. Soc. 138(44), 14713–14719 (2016). [CrossRef]  

21. S. Bettis Homan, V. K. Sangwan, I. Balla, H. Bergeron, E. A. Weiss, and M. C. Hersam, “Ultrafast exciton dissociation and long-lived charge separation in a photovoltaic pentacene–MoS2 van der Waals heterojunction,” Nano Lett. 17(1), 164–169 (2017). [CrossRef]  

22. H. Shi, R. Yan, S. Bertolazzi, J. Brivio, B. Gao, A. Kis, D. Jena, H. G. Xing, and L. Huang, “Exciton dynamics in suspended monolayer and few-layer MoS2 2D crystals,” ACS Nano 7(2), 1072–1080 (2013). [CrossRef]  

23. X. Hong, J. Kim, S.-F. Shi, Y. Zhang, C. Jin, Y. Sun, S. Tongay, J. Wu, Y. Zhang, and F. Wang, “Ultrafast charge transfer in atomically thin MoS 2/WS 2 heterostructures,” Nat. Nanotechnol. 9(9), 682–686 (2014). [CrossRef]  

24. Q. Zeng and Z. Liu, “Novel optoelectronic devices: transition-metal-dichalcogenide-based 2D heterostructures,” Adv. Electron. Mater. 4(2), 1700335 (2018). [CrossRef]  

25. L. Zhang, A. Sharma, Y. Zhu, Y. Zhang, B. Wang, M. Dong, H. T. Nguyen, Z. Wang, B. Wen, and Y. Cao, “Efficient and Layer-Dependent Exciton Pumping across Atomically Thin Organic–Inorganic Type-I Heterostructures,” Adv. Mater. 30(40), 1803986 (2018). [CrossRef]  

26. P. Rivera, J. R. Schaibley, A. M. Jones, J. S. Ross, S. Wu, G. Aivazian, P. Klement, K. Seyler, G. Clark, and N. J. Ghimire, “Observation of long-lived interlayer excitons in monolayer MoSe 2–WSe 2 heterostructures,” Nat. Commun. 6(1), 6242 (2015). [CrossRef]  

27. D. Zhang, Z. Zeng, Q. Tong, Y. Jiang, S. Chen, B. Zheng, J. Qu, F. Li, W. Zheng, and F. Jiang, “Near-Unity Polarization of Valley-Dependent Second-Harmonic Generation in Stacked TMDC Layers and Heterostructures at Room Temperature,” Adv. Mater.1908061 (2020). [CrossRef]  

28. W. Yang, H. Kawai, M. Bosman, B. Tang, J. Chai, W. Le Tay, J. Yang, H. L. Seng, H. Zhu, and H. Gong, “Interlayer interactions in 2D WS 2/MoS 2 heterostructures monolithically grown by in situ physical vapor deposition,” Nanoscale 10(48), 22927–22936 (2018). [CrossRef]  

29. A. Srivastava, M. Sidler, A. V. Allain, D. S. Lembke, A. Kis, and A. Imamoğlu, “Optically active quantum dots in monolayer WSe2,” Nat. Nanotechnol. 10(6), 491–496 (2015). [CrossRef]  

30. V. D. Karanikolas and E. Paspalakis, “Localized exciton modes and high quantum efficiency of a quantum emitter close to a MoS 2 nanodisk,” Phys. Rev. B 96(4), 041404 (2017). [CrossRef]  

31. S. H. Simon, “Spontaneous interlayer exciton coherence in quantum Hall bilayers at v = 1 and v = 2: a tutorial,” Solid State Commun. 134(1-2), 81–88 (2005). [CrossRef]  

32. T. Mueller and E. Malic, “Exciton physics and device application of two-dimensional transition metal dichalcogenide semiconductors,” npj 2D Mater. Appl. 2(1), 29 (2018). [CrossRef]  

33. S. Chichibu, T. Azuhata, T. Sota, and S. Nakamura, “Spontaneous emission of localized excitons in InGaN single and multiquantum well structures,” Appl. Phys. Lett. 69(27), 4188–4190 (1996). [CrossRef]  

34. H. Taghinejad, D. A. Rehn, C. Muccianti, A. A. Eftekhar, M. Tian, T. Fan, X. Zhang, Y. Meng, Y. Chen, and T.-V. Nguyen, “Defect-mediated alloying of monolayer transition-metal dichalcogenides,” ACS Nano 12(12), 12795–12804 (2018). [CrossRef]  

35. Z. Cai, B. Liu, X. Zou, and H.-M. Cheng, “Chemical vapor deposition growth and applications of two-dimensional materials and their heterostructures,” Chem. Rev. 118(13), 6091–6133 (2018). [CrossRef]  

36. Y. Gong, S. Lei, G. Ye, B. Li, Y. He, K. Keyshar, X. Zhang, Q. Wang, J. Lou, and Z. Liu, “Two-step growth of two-dimensional WSe2/MoSe2 heterostructures,” Nano Lett. 15(9), 6135–6141 (2015). [CrossRef]  

37. Z. Zhang, P. Chen, X. Duan, K. Zang, J. Luo, and X. Duan, “Robust epitaxial growth of two-dimensional heterostructures, multiheterostructures, and superlattices,” Sci. Bull. 357(6353), 788–792 (2017). [CrossRef]  

38. A. Hanbicki, M. Currie, G. Kioseoglou, A. Friedman, and B. Jonker, “Measurement of high exciton binding energy in the monolayer transition-metal dichalcogenides WS2 and WSe2,” Solid State Commun. 203, 16–20 (2015). [CrossRef]  

39. W. Shi, M.-L. Lin, Q.-H. Tan, X.-F. Qiao, J. Zhang, and P.-H. Tan, “Raman and photoluminescence spectra of two-dimensional nanocrystallites of monolayer WS2 and WSe2,” 2D Mater. 3(2), 025016 (2016). [CrossRef]  

40. Y. P. Varshni, “Temperature dependence of the energy gap in semiconductors,” Physica 34(1), 149–154 (1967). [CrossRef]  

41. G. Plechinger, P. Nagler, J. Kraus, N. Paradiso, C. Strunk, C. Schüller, and T. Korn, “Identification of excitons, trions and biexcitons in single-layer WS2,” Phys. Status Solidi RRL 9(8), 457–461 (2015). [CrossRef]  

42. A. Hichri, I. B. Amara, S. Ayari, and S. Jaziri, “Exciton, trion and localized exciton in monolayer Tungsten Disulfide,” (2016).

43. Z. He, X. Wang, W. Xu, Y. Zhou, Y. Sheng, Y. Rong, J. M. Smith, and J. H. Warner, “Revealing Defect-State Photoluminescence in Monolayer WS2 by Cryogenic Laser Processing,” ACS Nano 10(6), 5847–5855 (2016). [CrossRef]  

44. T. Cao, J. Feng, J. Shi, Q. Niu, and E. Wang, “MoS_2 as an ideal material for valleytronics: valley-selective circular dichroism and valley Hall effect,” (2011).

45. D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, “Coupled spin and valley physics in monolayers of MoS 2 and other group-VI dichalcogenides,” Phys. Rev. Lett. 108(19), 196802 (2012). [CrossRef]  

46. H. Yu, Y. Wang, Q. Tong, X. Xu, and W. Yao, “Anomalous light cones and valley optical selection rules of interlayer excitons in twisted heterobilayers,” Phys. Rev. Lett. 115(18), 187002 (2015). [CrossRef]  

47. G. Moody, K. Tran, X. Lu, T. Autry, J. M. Fraser, R. P. Mirin, L. Yang, X. Li, and K. L. Silverman, “Microsecond Valley Lifetime of Defect-Bound Excitons in Monolayer WSe2,” Phys. Rev. Lett. 121(5), 057403 (2018). [CrossRef]  

48. L. Linhart, M. Paur, V. Smejkal, J. Burgdörfer, T. Mueller, and F. Libisch, “Localized Intervalley Defect Excitons as Single-Photon Emitters in WSe 2,” Phys. Rev. Lett. 123(14), 146401 (2019). [CrossRef]  

49. S. Refaely-Abramson, D. Y. Qiu, S. G. Louie, and J. B. Neaton, “Defect-Induced Modification of Low-Lying Excitons and Valley Selectivity in Monolayer Transition Metal Dichalcogenides,” Phys. Rev. Lett. 121(16), 167402 (2018). [CrossRef]  

50. W.-T. Hsu, L.-S. Lu, P.-H. Wu, M.-H. Lee, P.-J. Chen, P.-Y. Wu, Y.-C. Chou, H.-T. Jeng, L.-J. Li, M.-W. Chu, and W.-H. Chang, “Negative circular polarization emissions from WSe(2)/MoSe(2) commensurate heterobilayers,” Nat. Commun. 9(1), 1356 (2018). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Growth of the WSe2/WS2 heterostructures. (a) Schematic diagram of the tube furnace used for the growth of the WSe2/WS2 heterostructures in the experiment. The arrows in different colors show two opposite directions of flow, with the green being the forward flow. (b) Schematic diagram of the physical vapor deposition method for the growth of the heterostructures. The top of the diagram shows the deposition process of molecules for the WSe2 layer, while the bottom of the diagram indicates the process of the epitaxial growth of WSe2 on top of the WS2 layer. (c) Optical images of the WSe2/WS2 heterostructures showing the smaller triangular-shaped WSe2 on top of the relatively larger triangular-shaped WS2.
Fig. 2.
Fig. 2. Raman mapping and photoluminescence quenching in the WSe2/WS2 heterostructures. (a) Raman spectra of WSe2 and WS2 and their heterostructure. The red line is a spectrum from the WSe2/WS2, and the blue line is the WS2 spectrum obtained from the edge of the heterostructure where there was no WSe2 present. As a comparison, the Raman spectrum (black line) from a monolayer WSe2 sample is also presented. (b) PL spectra of WSe2, WS2, and their heterostructure. (c)-(e) Raman mapping images of the WSe2/WS2 heterostructures. Mapped with their characteristic peaks, the green regions are WSe2, and the red regions are WS2. The insets are optical images of the WSe2/WS2 heterostructures. (f)-(h) PL mapping images of the WSe2/WS2 heterostructures showing photoluminescence quenching resulting from the band-structure of the type-II heterostructure. The insets in (c)-(h) are the optical images of the WSe2/WS2 heterostructures.
Fig. 3.
Fig. 3. Normalized photoluminescence spectra under variable temperature. (a) The PL spectra of the WS2 region under different temperatures (65K-300K). The dashed line indicates the trends of peak shifting. (b) The PL spectra of the localized excitons in the WSe2/WS2 heterostructures under different temperatures (65K-300K). (c) X- Peak energy as a function of temperature. The solid line indicates a fit using the Varshni equation. (d) Energy level diagram of WS2/WSe2 heterostructure.
Fig. 4.
Fig. 4. Valley polarization of localized excitons. (a) and (c) Polarization-resolved PL spectra for the WSe2/WS2 heterostructures at 65K using σ- excitation. (b) and (d) Polarization-resolved PL spectra for the WSe2/WS2 heterostructures at 65K using σ+ excitation. In (a) and (b), the red line represents the intensity of left-handed light emitted by the heterostructures. The blue line represents the intensity of right-handed light emitted by the heterostructures. The olive-green dotted line indicates the corresponding degree of PL polarization. (e) Schematic diagram of optical selection rules. (f) Schematic diagram of the optical selection rules of the local excitons in WS2. The purple lines indicate defect energy levels.

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

E g ( T ) = E g ( 0 ) α T 2 / ( T + β ) ,
P = I σ I σ + I σ + I σ +
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.