Skip to content
BY 4.0 license Open Access Published by De Gruyter January 21, 2022

Plasmon-enhanced photoluminescence from MoS2 monolayer with topological insulator nanoparticle

  • Dikun Li ORCID logo , Hua Lu ORCID logo EMAIL logo , Yangwu Li , Shouhao Shi , Zengji Yue and Jianlin Zhao
From the journal Nanophotonics

Abstract

Topological insulators (TI), as a kind of fantastic nanomaterial with excellent electrical and optical properties, have attracted particular attention due to the promising applications in optoelectronic devices. Herein, we experimentally demonstrated the interaction between light and molybdenum disulfide (MoS2) monolayer with an antimony telluride (Sb2Te3) TI nanoparticle. It was found that photoluminescence (PL) emission and Raman scattering signal can be boosted by 5 and 8 folds in MoS2 monolayer integrated with the TI nanoparticle, respectively. The measured and simulated dark-field scattering spectra illustrated that the enhancement of light–matter interaction could be derived from the generation of localized surface plasmons on the TI nanoparticle with distinctly boosted electric field. We also found that there exists a redshift of 5 nm for the enhanced PL peak, which could be attributed to the formation of trions in MoS2 induced by plasmon doping. This work would provide a new pathway for the applications of TI nanoparticles in the optoelectronics, especially light–matter interaction enhancement.

1 Introduction

Two-dimensional (2D) layered semiconductor nanomaterials with excellent optical, electrical, mechanical, and thermal properties have attracted broad attentions in electronics, optoelectronics, and photonics [1], [2], [3]. As a kind of typical 2D nanomaterials, transition-metal dichalcogenides (TMDCs) have found considerable applications in optoelectronic functionalities and devices, for instance photodetection [45], modulation [6, 7], frequency conversion [8], and lasering [9]. The energy band gaps of TMDCs are abundant and of dependence on atomic layer numbers [10]. For example, molybdenum disulfide (MoS2) bulk has an indirect bandgap of about 1.29 eV, but MoS2 monolayer presents a direct bandgap of about 1.9 eV [11]. The special bandgap characteristics are beneficial for the realization of the field effect transistors with an ultra-high on/off ratio [12]. Recently, MoS2 monolayer with attractive properties of light absorption, chemical stability, photoemission, and excitonic binding injects new vitality into solar cells [13], transistors [14], photocatalysis [15], and light emission [16, 17], and so on. However, the challenge facing us is that the atomic-layer structure of MoS2 with weak light–matter interaction prohibits the development of high-performance optoelectronic functionalities, for example photoluminescence (PL) emission [18]. Improving light interaction with MoS2 is crucial for highly efficient PL emission.

Different from traditional insulators/semiconductors and metals, topological insulators (TIs) are electronic materials with a bulk bandgap insulator and topologically protected conducting (metal-like) states on their edge or surface [19]. Due to the spin–orbit coupling effect in the bulk, the Dirac cores without the bandgap are protected by time-reversal symmetry. These unique properties contribute to the observation of exotic physical phenomena, such as quantum spin Hall effect [20], carrier backscattering avoidance [21], and Majorana fermions [22]. Recently, the topological surface states have been found in three-dimensional (3D) Sb2Te3, Bi2Te3, and Bi2Se3 materials [23]. These 3D materials exhibit unique electronic and optical features containing broad operating wavelength range, external tunability, and compatibility with optical elements [24], and play an important role in the realization of photonic Weyl points [25] and nanometric hologram [26]. Especially, the 3D TIs with conducting surface states enable the excitation of surface plasmons with lower loss and higher figure of merit in the ultraviolet and partly visible ranges compared with traditional metals [27, 28]. TIs with the generation of surface plasmons offer a new platform for the enhancement of light–matter interaction [29, 30].

Herein, we experimentally demonstrated the enhancement of interaction between light and MoS2 monolayer assisted by the Sb2Te3 topological insulator nanoparticle. The results show that the PL emission and Raman scattering can be enhanced by 5 and 8 folds in MoS2 monolayer with the TI nanoparticle, respectively. The dark field scattering spectrum denotes that the enhancement of light–MoS2 interaction results in the generation of localized surface plasmons with a strong electric field on the TI nanoparticle. The experimental measurement agrees well with the numerical simulations. A redshift is observed for the enhanced PL peak, which can be derived from the increased proportion of trions in MoS2 induced by plasmon doping. These results will pave a new avenue for exploring the enhancement of interaction between light and atomic-layer nanomaterials, especially for light emission.

2 Materials and methods

The high-quality Sb2Te3 bulk material was grown from high purity Sb and Te powders with an atomic ratio of 2:3 by the melting-slow cooling method. The small Sb2Te3 bulk is placed on transparent tape and peeled repeatedly using the tapes until it is changed into particles. After this, we stick the tape onto the polydimethylsiloxane (PDMS) film, and peel the tape slowly enough to make the smaller particles stay on the PDMS. Then, the Sb2Te3 nanoparticle with suitable size is found and transferred to the MoS2 layer by using optical microscope and micromanipulation system. The MoS2 layer was grown on the silicon wafer in a quartz tube by using chemical vapor deposition (CVD) method. During the transfer process, the temperature of the transfer platform can be kept at about 60 °C. Figure 1(a) shows the 3D diagram of MoS2 layer with a TI nanoparticle. To characterize the crystalline and stoichiometry of TI material, we measured the X-ray diffraction (XRD, Shimadzu) pattern of Sb2Te3 bulk with a scanning angle range from 10° to 85° and a degree step of 0.02° by using Cu targets at 3 kW power, as can be seen in Figure 1(b). It is found that the XRD pattern is aligned along the (001) plane, which denotes that the Sb2Te3 material is single-crystalline. Figure 1(c) depicts the X-ray spectrum of the grown Sb2Te3 measured by energy dispersive X-ray spectrometer (EDS) in scanning electron microscope (SEM, FEI Verios G4). We can see that the material contains Sb and Te elements with a molar ratio of about 2:3. The inset of Figure 1(c) shows the SEM image of grown Sb2Te3. Moreover, we measured the Raman-shift spectrum of Sb2Te3 employing confocal Raman microscope (WITec alpha300R) with a 532 nm wavelength laser, as shown in Figure 1(d). The Raman peaks at 67, 89, 117, 137, and 162 cm−1 are consistent with those of Sb2Te3 single crystal [31]. The profiles of MoS2 and Sb2Te3 nanoparticle can be measured by using atomic force microscope (AFM, Bruker Dimension Icon) and SEM. The Raman shift spectra, dark-field scattering spectra, and PL emission of MoS2 layer with the TI nanoparticle were measured with confocal Raman microscope. In addition, transmission electron microscope (TEM, FEI Talos F200X) was used to further characterize the morphology and structure of Sb2Te3 material. TEM image in high-angle annular dark field (HAADF) mode and SAED pattern of Sb2Te3 are shown in Figure 1(e) and (f), respectively. These results verify the good single-crystal characteristic of Sb2Te3. The finite-difference time-domain (FDTD) method is used to numerically simulate the optical spectra and field distribution of the Sb2Te3 TI nanoparticle [32].

Figure 1: 
Structure and characterization of Sb2Te3 material.
(a) 3D diagram of the MoS2 monolayer with a Sb2Te3 TI nanoparticle. (b) X-ray diffraction (XRD) patterns of Sb2Te3 crystal (aligned along the (001) plane) with a scanning angle range from 10° to 85° and a degree step of 0.02°. (c) X-ray spectrum of Sb2Te3 measured employing energy dispersive X-ray spectrometer (EDS) in scanning electron microscope (SEM). The inset shows the SEM top view of grown Sb2Te3 for EDS. (d) Raman scattering spectrum of Sb2Te3 excited with a 532 nm wavelength laser. (e) Transmission electron microscope (TEM) image in high-angle annular dark field (HAADF) mode of Sb2Te3. The scale bar is 1 μm. (f) Selected area electron diffraction (SAED) pattern of Sb2Te3. The scale bar is 10 nm−1.
Figure 1:

Structure and characterization of Sb2Te3 material.

(a) 3D diagram of the MoS2 monolayer with a Sb2Te3 TI nanoparticle. (b) X-ray diffraction (XRD) patterns of Sb2Te3 crystal (aligned along the (001) plane) with a scanning angle range from 10° to 85° and a degree step of 0.02°. (c) X-ray spectrum of Sb2Te3 measured employing energy dispersive X-ray spectrometer (EDS) in scanning electron microscope (SEM). The inset shows the SEM top view of grown Sb2Te3 for EDS. (d) Raman scattering spectrum of Sb2Te3 excited with a 532 nm wavelength laser. (e) Transmission electron microscope (TEM) image in high-angle annular dark field (HAADF) mode of Sb2Te3. The scale bar is 1 μm. (f) Selected area electron diffraction (SAED) pattern of Sb2Te3. The scale bar is 10 nm−1.

3 Results and discussion

Figure 2(a) shows the SEM image of the CVD-grown MoS2 layer with a Sb2Te3 TI nanoparticle. The inset of Figure 2(a) depicts the optical microscopy image of the MoS2 layer with a Sb2Te3 TI nanoparticle. We measured the profile of the MoS2 layer with the TI nanoparticle, as shown in Figure 2(b). The AFM-measured profiles in Figure 2(c) show that the Sb2Te3 nanoparticle is approximately a rectangular shape with the length of a = 700 nm, width of b = 165 nm, and height of h = 30 nm. To clarify the MoS2 layer number, we measured the Raman shift spectrum of MoS2 layer with a 532 nm laser irradiation. It is shown that there exist obvious Raman peaks at about 385.3 and 404.6 cm−1, which correspond to the vibration mode E12g in the excitation plane and the vibration mode A1g outside the plane of MoS2 monolayer, respectively [33]. The frequency difference between the two peaks is 19.3 cm−1, which is similar to the reported result of monolayer MoS2 [34]. Subsequently, we demonstrated the PL emission and Raman scattering of MoS2 monolayer with the TI nanoparticle for exploring the light interaction with MoS2 by using confocal Raman microscope with the objective lens of 100× and 532 nm excitation laser. As depicted in Figure 3(a), there exist two PL emission peaks at 620 and 665 nm wavelengths for MoS2 on the substrate, which corresponds to the B and A direct excitonic transitions of MoS2 monolayer at the K point of Brillouin zone [35]. The excitonic behaviors of PL emission in MoS2 may be changed under the condition of low temperature [36]. The inset of Figure 3(a) shows the PL intensity image obtained by integrating PL spectral intensities around the 665 nm wavelength at the mapping area (3.4 μm × 3.4 μm) covering the TI nanoparticle. We can see an obvious maximum PL intensity at the position of TI nanoparticle, which is improved by 5 folds compared to the MoS2 monolayer without the TI nanoparticle. The value is remarkable compared with the PL enhancement of CdSe/ZnS QDs on the Bi2Te3 nanoplate [30]. It is worth noting that the area of PL enhancement is larger than that of TI nanoparticle, which results in the limited laser beam with the diameter of ∼500 nm. When the size of TI nanoparticle is larger than the diameter of laser beam, the PL emission can also be enhanced around the TI particle due to the plasmonic generation at the side of TI particle. Figure 3(b) shows the Raman spectra of MoS2 monolayer with and without the TI nanoparticle. The Raman intensity is strongest at the center of TI nanoparticle, and decreases at the positions away from the TI center. The Raman scattering signal can be considerably enhanced by 8 folds for the MoS2 monolayer with the TI nanoparticle. It is also found in Figure 3(a) that the PL peak presents a redshift of 5 nm for the A exciton emission. To elucidate the reason, we fit the PL emission spectra by means of the multi-Lorentzian function, where a kind of quasiparticle state (i.e., A trion) can be considered [37]. Figure 3(c) shows the fitting results of normalized PL spectra of A exciton, A trion, and B exciton for the MoS2 monolayer without the TI nanoparticle. We can see that the A excitons dominate in the PL emission for MoS2 monolayer on the silicon substrate. There exists the lower proportion for A trions in MoS2 on the substrate at 675 nm wavelength. However, the proportion of A trions dramatically increases for MoS2 monolayer with the TI nanoparticle, as shown in Figure 3(d). This behavior could be derived from the electron doping of MoS2 from the TI nanoparticle with the radiation of the light field [37]. The increased A trions at the longer wavelength result in the redshift of PL emission peak. When the Sb2Te3 is covered on MoS2, the binding energies of excitons will decrease with the increase of relative dielectric constant on MoS2 [38]. This results in the blueshift of A exciton and A trion of MoS2 PL emission, as shown in Figure 3(c) and (d).

Figure 2: 
Characterization of MoS2 layer with the TI nanoparticle.
(a) SEM top view of MoS2 layer with a Sb2Te3 TI nanoparticle and the corresponding optical microscopy image. (b) AFM-measured MoS2 layer with the Sb2Te3 TI nanoparticle. In (a) and (b), the scale bars are 1 μm. (c) Height profiles of Sb2Te3 nanoparticle along two white dotted lines in (b). (d) Raman shift spectrum of MoS2 layer excited with a 532 nm wavelength laser.
Figure 2:

Characterization of MoS2 layer with the TI nanoparticle.

(a) SEM top view of MoS2 layer with a Sb2Te3 TI nanoparticle and the corresponding optical microscopy image. (b) AFM-measured MoS2 layer with the Sb2Te3 TI nanoparticle. In (a) and (b), the scale bars are 1 μm. (c) Height profiles of Sb2Te3 nanoparticle along two white dotted lines in (b). (d) Raman shift spectrum of MoS2 layer excited with a 532 nm wavelength laser.

Figure 3: 
Enhanced PL emission and Raman shift spectra of MoS2 layer.
(a) PL emission spectra of MoS2 monolayer with and without the Sb2Te3 TI nanoparticle. The inset shows the PL intensity mapping image around the Sb2Te3 TI nanoparticle. (b) Raman shift spectra of MoS2 monolayer with and without the Sb2Te3 TI nanoparticle. (c) Normalized PL spectra of A exciton, A− trion, and B exciton in MoS2 monolayer on the substrate. (d) Normalized PL spectra of A exciton, A− trion, and B exciton in MoS2 monolayer with the Sb2Te3 TI nanoparticle. The PL spectra in (c) and (d) are obtained by fitting the experimental data using the multi-Lorentzian function.
Figure 3:

Enhanced PL emission and Raman shift spectra of MoS2 layer.

(a) PL emission spectra of MoS2 monolayer with and without the Sb2Te3 TI nanoparticle. The inset shows the PL intensity mapping image around the Sb2Te3 TI nanoparticle. (b) Raman shift spectra of MoS2 monolayer with and without the Sb2Te3 TI nanoparticle. (c) Normalized PL spectra of A exciton, A trion, and B exciton in MoS2 monolayer on the substrate. (d) Normalized PL spectra of A exciton, A trion, and B exciton in MoS2 monolayer with the Sb2Te3 TI nanoparticle. The PL spectra in (c) and (d) are obtained by fitting the experimental data using the multi-Lorentzian function.

To clarify the mechanism of the above results, we measured the dark-field scattering spectrum of the Sb2Te3 TI nanoparticle on MoS2 monolayer using confocal Raman microscope, as shown in Figure 4(a). It is found that there exists a scattering spectral peak at the wavelength of 550 nm with a peak width of about 120 nm. The inset of Figure 4(a) depicts the dark-field image of the MoS2 monolayer with the Sb2Te3 TI nanoparticle. The scattering spectrum is achieved through dividing the scattering intensity from the sample by the incident light spectrum (measured by mirror reflector). To verify it, we numerically simulate the scattering spectrum of the Sb2Te3 TI nanoparticle using the FDTD method [32]. In FDTD simulations, we use an effective modeling method of dark-field scattering for the nanoparticles proposed by Jiang [39]. Two confocal Gaussian beams with a phase of π are used to construct an annular light source (dark-field light source), which is impinging on the Sb2Te3 TI nanoparticle. A power monitor with a finite area is set above the source. The perfectly matched layer absorbing boundary condition is set for the six planes in the computational area [39]. The relative permittivities of Sb2Te3 TI surface and bulk states are simultaneously considered in the simulations [40]. It is worth noting that both the surface and bulk states for Sb2Te3 present negative relative permittivities in the visible range of interest, which can contribute to the generation of surface plasmons. The thickness of TI surface layer is set as 2.6 nm [41]. The relative permittivity of silicon is set as the experiment data in Ref. [41]. The simulated scattering spectrum can be calculated by dividing the power passing through the monitor from the Sb2Te3 nanoparticle by that from the mirror reflector. As shown in Figure 4(b), the calculated scattering spectrum is in excellent agreement with the experimental result. The inset of Figure 4(b) depicts the simulated electric field distribution of |E| at 532 nm wavelength, which reveals the excitation of localized surface plasmons with a distinct field enhancement on the TI nanoparticle. The localized surface plasmons of TI nanoparticle can benefit for improving light–MoS2 interaction [30]. Therefore, it is understandable that the distinct enhancement of PL emission and Raman scattering can be observed in MoS2 monolayer with the TI nanoparticle [42]. The localized surface plasmons with the free-electron oscillation contribute to the doping of MoS2 monolayer with the TI nanoparticle, resulting in the increased A trions and PL peak intensity. Thus, the charge transfer will exist between the MoS2 monolayer with Sb2Te3 TI nanoparticle. The strain of MoS2 on the nanocone substrate will give rise to the redshift of Raman and PL emission peaks [43]. In our structure, the MoS2 grown on a flat substrate is covered on TI nanoparticle; the redshift of the PL peak induced by the strain can be neglected.

Figure 4: 
Dark-field scattering spectra of TI nanoparticle.
(a) Experimentally measured dark-field scattering spectrum of the Sb2Te3 TI nanoparticle on the MoS2 monolayer. The inset shows the dark-field image of MoS2 monolayer with a Sb2Te3 TI nanoparticle. (b) Numerically simulated scattering spectrum of the Sb2Te3 TI nanoparticle. The inset shows the distribution of electric field |E| for the Sb2Te3 nanoparticle at the 532 nm wavelength. The scale bar is 100 nm. The experiment and numerical results are obtained by the confocal Raman microscope and FDTD simulation, respectively.
Figure 4:

Dark-field scattering spectra of TI nanoparticle.

(a) Experimentally measured dark-field scattering spectrum of the Sb2Te3 TI nanoparticle on the MoS2 monolayer. The inset shows the dark-field image of MoS2 monolayer with a Sb2Te3 TI nanoparticle. (b) Numerically simulated scattering spectrum of the Sb2Te3 TI nanoparticle. The inset shows the distribution of electric field |E| for the Sb2Te3 nanoparticle at the 532 nm wavelength. The scale bar is 100 nm. The experiment and numerical results are obtained by the confocal Raman microscope and FDTD simulation, respectively.

4 Conclusions

In this paper, we have demonstrated the interaction between light and MoS2 monolayer integrated with a Sb2Te3 TI nanoparticle. Through the fixed-point transfer method, the TI nanoparticle is transferred to the CVD-grown MoS2 monolayer. The experimental results show that the PL emission and Raman scattering intensity with a 532 nm excitation laser can be improved by 5 and 8 folds for MoS2 monolayer with the TI nanoparticle. The dark-field scattering spectrum illustrates that the TI nanoparticle can support the generation of localized surface plasmons with a distinct electric field enhancement. The experimental measurement is consistent with the FDTD numerical simulations. The improvement of PL emission and Raman scattering of MoS2 monolayer can be attributed to the formation of localized surface plasmons on the TI nanoparticle. The localized surface plasmons also contribute to the 5 nm redshift of PL emission peak with the increase of A trions proportion in MoS2 monolayer by plasmon doping. The crystal quality of MoS2 can be improved using the modified CVD growth method [44], [45], [46]. The improvement of MoS2 quality would enhance Raman and PL emission in the systems. This work provides a new pathway for the applications of TI materials in light–matter interaction and high-performance optoelectronic functionalities based on atomic-layer materials.


Corresponding author: Hua Lu, MOE Key Laboratory of Material Physics and Chemistry under Extraordinary Conditions, Key Laboratory of Light-Field Manipulation and Information Acquisition, Ministry of Industry and Information Technology, and Shaanxi Key Laboratory of Optical Information Technology, School of Physical Science and Technology, Northwestern Polytechnical University, Xi’an 710129, China, E-mail:

Funding source: National Natural Science Foundation of China http://dx.doi.org/10.13039/501100001809

Award Identifier / Grant number: 11974283

Award Identifier / Grant number: 61705186

Award Identifier / Grant number: 11774290

Award Identifier / Grant number: 11634010

Funding source: “Double first-class” construction fund project

Award Identifier / Grant number: 0651021GH0201113

Funding source: Natural Science Basic Research Plan in Shaanxi Province of China

Award Identifier / Grant number: 2020JM-130

Funding source: National Key R&D Program of China

Award Identifier / Grant number: 2017YFA0303800

Acknowledgment

The authors thank the Analytical & Testing Center of Northwestern Polytechnical University for the as well as AFM, SEM, and XRD measurements. Thanks for the supply of MoS2 from Dr. Q. Feng at Northwestern Polytechnical University.

  1. Author contribution: D. L. carried out the measurement of sample and simulations as well as drafted the manuscript. H. L. conceived the idea, characterized the TI material, analyzed the results, and improved the manuscript. Y. L. and S. S. took part in the transfer of material. Z. Y. fabricated the TI material. J. Z. discussed the results and promoted the manuscript presentation. All authors substantially contributed to the manuscript.

  2. Research funding: The authors would like to thank the funding support from the National Key R&D Program of China (2017YFA0303800), National Natural Science Foundation of China (11974283, 61705186, 11774290 and 11634010), Natural Science Basic Research Plan in Shaanxi Province of China (2020JM-130), and “Double first-class” construction fund project (0651021GH0201113).

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] P. Hu, Z. Wen, L. Wang, P. Tan, and K. Xiao, “Synthesis of few-layer GaSe nanosheets for high performance photodetectors,” ACS Nano, vol. 6, pp. 5988–5994, 2012. https://doi.org/10.1021/nn300889c.Search in Google Scholar PubMed

[2] H. Guo, C. Lan, Z. Zhou, P. Sun, D. Wei, and C. Li, “Transparent, flexible, and stretchable WS2 based humidity sensors for electronic skin,” Nanoscale, vol. 9, pp. 6246–6253, 2017. https://doi.org/10.1039/c7nr01016h.Search in Google Scholar PubMed

[3] C. Lan, Z. Zhou, Z. Zhou, et al.., “Wafer-scale synthesis of monolayer WS2 for high-performance flexible photodetectors by enhanced chemical vapor deposition,” Nano Res., vol. 11, pp. 3371–3384, 2018. https://doi.org/10.1007/s12274-017-1941-4.Search in Google Scholar

[4] S. Li, X. Chen, F. Liu, et al.., “Enhanced performance of a CVD MoS2 photodetector by chemical in situ n-type doping,” ACS Appl. Mater. Interfaces, vol. 11, pp. 11636–11644, 2019. https://doi.org/10.1021/acsami.9b00856.Search in Google Scholar PubMed

[5] P. Xiao, J. Mao, K. Ding, et al.., “Solution-processed 3D RGO-MoS2/pyramid Si heterojunction for ultrahigh detectivity and ultra-broadband photodetection,” Adv. Mater., vol. 30, p. 1801729, 2018. https://doi.org/10.1002/adma.201801729.Search in Google Scholar PubMed

[6] S. Yu, X. Wu, Y. Wang, X. Guo, and L. Tong, “2D materials for optical modulation: challenges and opportunities,” Adv. Mater., vol. 29, p. 1606128, 2017. https://doi.org/10.1002/adma.201606128.Search in Google Scholar PubMed

[7] Z. Sun, A. Martinez, and F. Wang, “Optical modulators with 2D layered materials,” Nat. Photonics, vol. 10, pp. 227–238, 2016. https://doi.org/10.1038/nphoton.2016.15.Search in Google Scholar

[8] J. Shi, P. Yu, F. Liu, et al.., “3R MoS2 with broken inversion symmetry: a promising ultrathin nonlinear optical device,” Adv. Mater., vol. 29, p. 1701486, 2017. https://doi.org/10.1002/adma.201701486.Search in Google Scholar PubMed

[9] Y. Ye, Z. J. Wong, X. Lu, et al.., “Monolayer excitonic laser,” Nat. Photonics, vol. 9, pp. 733–737, 2015. https://doi.org/10.1038/nphoton.2015.197.Search in Google Scholar

[10] J. Gusakova, X. Wang, L. L. Shiau, et al.., “Electronic Properties of Bulk and Monolayer TMDs: theoretical Study Within DFT Framework (GVJ-2e Method),” Phys. Status Solidi A, vol. 214, p. 1700218, 2017. https://doi.org/10.1002/pssa.201700218.Search in Google Scholar

[11] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS2: a new direct-gap semiconductor,” Phys. Rev. Lett., vol. 105, p. 136805, 2010. https://doi.org/10.1103/physrevlett.105.136805.Search in Google Scholar PubMed

[12] O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive photodetectors based on monolayer MoS2,” Nat. Nanotechnol., vol. 8, pp. 497–501, 2013. https://doi.org/10.1038/nnano.2013.100.Search in Google Scholar PubMed

[13] M.-L. Tsai, S.-H. Su, J.-K. Chang, et al.., “Monolayer MoS2 heterojunction solar cells,” ACS Nano, vol. 8, pp. 8317–8322, 2014. https://doi.org/10.1021/nn502776h.Search in Google Scholar PubMed

[14] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol., vol. 6, pp. 147–150, 2011. https://doi.org/10.1038/nnano.2010.279.Search in Google Scholar PubMed

[15] K. Chang, X. Hai, H. Pang, et al.., “Targeted synthesis of 2H-and 1T-phase MoS2 monolayers for catalytic hydrogen evolution,” Adv. Mater., vol. 28, pp. 10033–10041, 2016. https://doi.org/10.1002/adma.201603765.Search in Google Scholar PubMed

[16] S. Butun, S. Tongay, and K. Aydin, “Enhanced light emission from large-area monolayer MoS2 using plasmonic nanodisc arrays,” Nano Lett., vol. 15, pp. 2700–2704, 2015. https://doi.org/10.1021/acs.nanolett.5b00407.Search in Google Scholar PubMed

[17] J. S. Ross, P. Klement, A. M. Jones, et al.., “Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions,” Nat. Nanotechnol., vol. 9, pp. 268–272, 2014. https://doi.org/10.1038/nnano.2014.26.Search in Google Scholar PubMed

[18] S. Mouri, Y. Miyauchi, and K. Matsuda, “Tunable photoluminescence of monolayer MoS2 via chemical doping,” Nano Lett., vol. 13, pp. 5944–5948, 2013. https://doi.org/10.1021/nl403036h.Search in Google Scholar PubMed

[19] J. E. Moore, “The birth of topological insulators,” Nature, vol. 464, pp. 194–198, 2010. https://doi.org/10.1038/nature08916.Search in Google Scholar PubMed

[20] M. Koenig, S. Wiedmann, C. Bruene, et al.., “Quantum spin hall insulator state in HgTe quantum wells,” Science, vol. 318, pp. 766–770, 2007.10.1126/science.1148047Search in Google Scholar PubMed

[21] M. Z. Hasan and C. L. Kane, “Colloquium: topological insulators,” Rev. Mod. Phys., vol. 82, pp. 3045–3067, 2010. https://doi.org/10.1103/revmodphys.82.3045.Search in Google Scholar

[22] L. Fu and C. L. Kane, “Superconducting proximity effect and Majorana fermions at the surface of a topological insulator,” Phys. Rev. Lett., vol. 100, p. 096407, 2008. https://doi.org/10.1103/PhysRevLett.100.096407.Search in Google Scholar PubMed

[23] Y. Xia, D. Qian, D. Hsieh, et al.., “Observation of a large-gap topological-insulator class with a single Dirac cone on the surface,” Nat. Phys., vol. 5, pp. 398–402, 2009. https://doi.org/10.1038/nphys1274.Search in Google Scholar

[24] Y. Wu, C. Li, X. Hu, Y. Ao, Y. Zhao, and Q. Gong, “Applications of topological photonics in integrated photonic devices,” Adv. Opt. Mater., vol. 5, p. 1700357, 2017. https://doi.org/10.1002/adom.201700357.Search in Google Scholar

[25] E. Goi, Z. Yue, B. P. Cumming, and M. Gu, “Observation of type I photonic Weyl points in optical frequencies,” Laser Photon. Rev., vol. 12, p. 1700271, 2018. https://doi.org/10.1002/lpor.201700271.Search in Google Scholar

[26] Z. Yue, G. Xue, J. Liu, Y. Wang, and M. Gu, “Nanometric holograms based on a topological insulator material,” Nat. Commun., vol. 8, p. 15354, 2017. https://doi.org/10.1038/ncomms15354.Search in Google Scholar PubMed PubMed Central

[27] Z. Yue, H. Ren, S. Wei, J. Lin, and M. Gu, “Angular-momentum nanometrology in an ultrathin plasmonic topological insulator film,” Nat. Commun., vol. 9, p. 4413, 2018. https://doi.org/10.1038/s41467-018-06952-1.Search in Google Scholar PubMed PubMed Central

[28] H. Lu, S. Dai, Z. Yue, et al.., “Sb2Te3 topological insulator: surface plasmon resonance and application in refractive index monitoring,” Nanoscale, vol. 11, pp. 4759–4766, 2019. https://doi.org/10.1039/c8nr09227c.Search in Google Scholar PubMed

[29] Z. Yue, B. Cai, L. Wang, X. Wang, and M. Gu, “Intrinsically core-shell plasmonic dielectric nanostructures with ultrahigh refractive index,” Sci. Adv., vol. 2, 2016, Art no. e1501536. https://doi.org/10.1126/sciadv.1501536.Search in Google Scholar PubMed PubMed Central

[30] M. Zhao, J. Zhang, N. Gao, et al.., “Actively tunable visible surface plasmons in Bi2Te3 and their energy-harvesting applications,” Adv. Mater., vol. 28, pp. 3138–3144, 2016. https://doi.org/10.1002/adma.201506367.Search in Google Scholar PubMed

[31] R. Zybala, K. Mars, A. Mikula, et al.., “Synthesis and characterization of antimony telluride for thermoelectric and optoelectronic applications,” Arch. Metall. Mater., vol. 62, pp. 1067–1070, 2017.10.1515/amm-2017-0155Search in Google Scholar

[32] A. Taflove and S. C. Hagness, Computational Electrodynamics: The Finite-Difference Time-Domain Method, 2nd ed., Boston, Artech House, 2000.Search in Google Scholar

[33] Y. Liu, H. Nan, X. Wu, et al.., “Layer-by-Layer thinning of MoS2 by plasma,” ACS Nano, vol. 7, pp. 4202–4209, 2013. https://doi.org/10.1021/nn400644t.Search in Google Scholar PubMed

[34] D. Pierucci, H. Henck, C. H. Naylor, et al.., “Large area molybdenum disulphide-epitaxial graphene vertical Van der Waals heterostructures,” Sci. Rep., vol. 6, p. 26656, 2016. https://doi.org/10.1038/srep26656.Search in Google Scholar PubMed PubMed Central

[35] A. Splendiani, L. Sun, Y. Zhang, et al.., “Emerging photoluminescence in monolayer MoS2,” Nano Lett., vol. 10, pp. 1271–1275, 2010. https://doi.org/10.1021/nl903868w.Search in Google Scholar PubMed

[36] Y. Wang, L. Deng, Q. Wei, et al.., “Spin-valley locking effect in defect states of monolayer MoS2,” Nano Lett., vol. 20, pp. 2129–2136, 2020. https://doi.org/10.1021/acs.nanolett.0c00138.Search in Google Scholar PubMed

[37] K. F. Mak, K. He, C. Lee, et al.., “Tightly bound trions in monolayer MoS2,” Nat. Mater., vol. 12, pp. 207–211, 2013. https://doi.org/10.1038/nmat3505.Search in Google Scholar PubMed

[38] Y. Lin, X. Ling, L. Yu, et al.., “Dielectric screening of excitons and trions in single-layer MoS2,” Nano Lett., vol. 14, pp. 5569–5576, 2014. https://doi.org/10.1021/nl501988y.Search in Google Scholar PubMed

[39] L. Jiang, T. Yin, Z. Dong, et al.., “Accurate modeling of dark-field scattering spectra of plasmonic nano structures,” ACS Nano, vol. 9, pp. 10039–10046, 2015. https://doi.org/10.1021/acsnano.5b03622.Search in Google Scholar PubMed

[40] H. Lu, Z. Yue, Y. Li, et al.., “Magnetic plasmon resonances in nanostructured topological insulators for strongly enhanced light-MoS2 interactions,” Light Sci. Appl., vol. 9, p. 191, 2020. https://doi.org/10.1038/s41377-020-00429-x.Search in Google Scholar PubMed PubMed Central

[41] E. D. Palik, Handbook of Optical Constants of Solids, New York, Academic Press, 1991.Search in Google Scholar

[42] Y. Yu, T. H. Xiao, Y. Z. Wu, et al.., “Roadmap for single-molecule surface-enhanced Raman spectroscopy,” Adv. Photonics, vol. 2, p. 014002, 2020. https://doi.org/10.1117/1.ap.2.1.014002.Search in Google Scholar

[43] H. Li, A. Contryman, X. Qian, et al.., “Optoelectronic crystal of artificial atoms in strain-textured molybdenum disulphide,” Nat. Commun., vol. 6, p. 7381, 2015. https://doi.org/10.1038/ncomms8381.Search in Google Scholar PubMed PubMed Central

[44] J. Chen, W. Tang, B. Tian, et al.., “Chemical vapor deposition of high-quality large-sized MoS2 crystals on silicon dioxide substrates,” Adv. Sci., vol. 3, p. 1500033, 2016. https://doi.org/10.1002/advs.201600033.Search in Google Scholar PubMed PubMed Central

[45] W. Wang, H. Shu, J. Wang, Y. Cheng, P. Liang, and X. Chen, “Defect passivation and photoluminescence enhancement of monolayer MoS2 crystals through sodium halide-assisted chemical vapor deposition growth,” ACS Appl. Mater. Interfaces, vol. 12, pp. 9563–9571, 2020. https://doi.org/10.1021/acsami.9b19224.Search in Google Scholar PubMed

[46] Q. Wang, N. Li, J. Tang, et al.., “Wafer-scale highly oriented monolayer MoS2 with large domain sizes,” Nano Lett., vol. 20, pp. 7193–7199, 2020. https://doi.org/10.1021/acs.nanolett.0c02531.Search in Google Scholar PubMed

Received: 2021-11-05
Revised: 2021-12-18
Accepted: 2021-12-30
Published Online: 2022-01-21

© 2022 Dikun Li et al., published by De Gruyter, Berlin/Boston

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

Downloaded on 8.12.2023 from https://www.degruyter.com/document/doi/10.1515/nanoph-2021-0685/html
Scroll to top button