Home Self-powered broadband photodetector based on MoS2/Sb2Te3 heterojunctions: a promising approach for highly sensitive detection
Article Open Access

Self-powered broadband photodetector based on MoS2/Sb2Te3 heterojunctions: a promising approach for highly sensitive detection

  • Hao Wang ORCID logo EMAIL logo , Yaliang Gui ORCID logo , Chaobo Dong , Salem Altaleb , Behrouz Movahhed Nouri , Martin Thomaschewski , Hamed Dalir ORCID logo EMAIL logo and Volker J. Sorger ORCID logo EMAIL logo
Published/Copyright: October 28, 2022
Become an author with De Gruyter Brill

Abstract

Topological insulators have shown great potential for future optoelectronic technology due to their extraordinary optical and electrical properties. Photodetectors, as one of the most widely used optoelectronic devices, are crucial for sensing, imaging, communication, and optical computing systems to convert optical signals to electrical signals. Here we experimentally show a novel combination of topological insulators (TIs) and transition metal chalcogenides (TMDs) based self-powered photodetectors with ultra-low dark current and high sensitivity. The photodetector formed by a MoS2/Sb2Te3 heterogeneous junction exhibits a low dark current of 2.4 pA at zero bias and 1.2 nA at 1V. It shows a high photoresponsivity of >150 mA W−1 at zero bias and rectification of 3 times at an externally applied bias voltage of 1V. The excellent performance of the proposed photodetector with its innovative material combination of TMDs and TIs paves the way for the development of novel high-performance optoelectronic devices. The TIs/TMDs transfer used to form the heterojunction is simple to incorporate into on-chip waveguide systems, enabling future applications on highly integrated photonic circuits.

1 Introduction

Transition metal chalcogenides (TMDs) have been widely explored as active materials for high-performance optoelectronics due to their unique physical properties [13]. Transition metal chalcogenides optoelectronics usually suffers from surface oxidation, high contact resistance, and relatively low mobilities. To alleviate the aforementioned disadvantages of TMD-based photodetectors, recent technologies utilize plasmonic enhancement to improve the light–matter interaction of TMDs, while others use heterostructures to adjust inherent electrical and optical characteristics and graphene contact to aid charge carrier injection and extraction; those all exhibit attractive performance improvements such as higher sensitivity, broader frequency response, improved external quantum efficiencies (EQE), higher detectivity (D*), and a broader spectral photoresponse [48]. However, combining topological insulators (TIs) as a novel quantum materials group with the attractive features of TMDs has remained largely unexplored. Topological insulators themselves exhibit various appealing optoelectronic features [911], which find applications in high-performance photodetection, e.g., due to their unique energy band structure. Furthermore, the topologically protected gapless conductive edge states or surface states in TIs offer extraordinarily high mobility and a broad detection spectrum compared to graphene, which has zero bandgaps [1216]. Topological insulators are also mechanically flexible materials that can be used in wearable technology. Most TMDs are easily oxidized under atmospheric environments, whereas TIs are more stable in normal conditions. As a monolayer, molybdenum disulfide (MoS2) has a direct bandgap of 1.8 eV, whereas, in bulk, it has an indirect bandgap of 1.3 eV [17]. The visible to the infrared absorption spectrum of MoS2 ranges from 350 to 950 nm, which is utilized in various optoelectronic applications such as light harvesting, photovoltaics, and photodetection [1823]. Heterojunctions are widely applied to self-powered photodetectors since the built-in electric field can efficiently separate the electron-hole pairs to enhance the photoresponsivity and speed under zero bias voltage. Different combinations of materials or technics have been demonstrated to improve the performance of the optoelectronic devices [2428]. Recently, a tellurium at selenium roll-to-roll nanotube heterojunction and cadmium sulfide at cadmium selenide core/shell quantum dots showed the enhanced capacity for self-powered broadband photodetection but also significantly improved photocurrent density and stability in different environments [29, 30]. Here, we demonstrate a vertical heterojunction-based photodetector with TIs and TMDs, which provide a promising tool to bring different electrical and optical properties by creating the junction compared to the individual materials. The heterojunction-based photodetector, formed by MoS2 and Sb2Te3 exfoliation and deterministic transfer to a silicon oxide substrate, shows reduced dark current, high responsivity R = I photo/P laser (defined by the ratio between the photocurrent I photo and the incident laser power P laser), and a self-driven detection range from visible to near-infrared wavelengths. The photocarrier separation is accelerated, and photocarrier recombination is suppressed due to the inherent difference in their energy level. Compared to single TMDs, it also exhibits a reduced dark current, which is beneficial for achieving high detection sensitivities at reduced power consumption. The dark current is ∼10 pA, and the light-induced on/off ratio is more than 103 due to the interlayer built-in field and a broadband response from 500 to 900 nm. Besides the demonstrated free-space application, the proposed device has potential in densely integrated applications such as digital to analog converter systems or photonic integrated circuits due to the ease of integration with photonic waveguide technology [3133].

2 Results and discussion

To fabricate the device, Sb2Te3 and MoS2 layers were prepared by mechanical exfoliation and transferred to a SiO2/Si substrate by the pick and drop technique. These two layers are stacked by weak van der Waals force (Figure 1a). Three Ti/Au electrodes formed by electron beam evaporation act as the electrical contact pads to inject and extract the photocurrent generated by the photodetector. Two electrodes are placed on the top of the MoS2, and one electrode is placed on the Sb2Te3 (Figure 1b). The compact total junction size is measured to be only 160 μm2. A Schottky barrier is present when the electrode is contacted with the MoS2/Sb2Te3 layers to reach the equilibrium state, and it also generates a tunneling current. The Schottky barrier height is strongly dependent on the metal work function. However, the applied heterostructure minimizes the depletion region due to the atomically thin layer. The photocarriers can be efficiently separated at the heterostructure interface with the built-in electric field in the PN heterojunction or an external electric field from an applied bias, yielding photocurrent across the heterostructure channel. As a result, there is no significant potential barrier in the forward direction, which increases the photodetector’s responsivity (Figure 2a).

Figure 1: 
Sb2Te3/MoS2 pn junction heterostructure photodetector. (a) A schematic representation of the Sb2Te3/MoS2 van der Waals p-n junction photodetector. (b) The optical microscope image of PN junction device.
Figure 1:

Sb2Te3/MoS2 pn junction heterostructure photodetector. (a) A schematic representation of the Sb2Te3/MoS2 van der Waals p-n junction photodetector. (b) The optical microscope image of PN junction device.

Figure 2: 
Material characterization of the Sb2Te3/MoS2 heterojunction. (a) Band structures of the vdW layered MoS2/Sb2Te3 heterojunction (b) SEM image of the fabricated device with point A on Sb2Te3, point B on MoS2 and point C on the MoS2/Sb2Te3 heterojunction. Two electrodes are on top of the MoS2 layer and one on top of the Sb2Te3. (c) As shown in the SEM image, EDS signals were collected at three distinct positions to characterize the materials. (d) Raman spectra were collected from the pure MoS2 and Sb2Te3 and the Sb2Te3/MoS2 PN junction region with a 532 nm laser.
Figure 2:

Material characterization of the Sb2Te3/MoS2 heterojunction. (a) Band structures of the vdW layered MoS2/Sb2Te3 heterojunction (b) SEM image of the fabricated device with point A on Sb2Te3, point B on MoS2 and point C on the MoS2/Sb2Te3 heterojunction. Two electrodes are on top of the MoS2 layer and one on top of the Sb2Te3. (c) As shown in the SEM image, EDS signals were collected at three distinct positions to characterize the materials. (d) Raman spectra were collected from the pure MoS2 and Sb2Te3 and the Sb2Te3/MoS2 PN junction region with a 532 nm laser.

Raman and energy-dispersive X-ray spectroscopy (EDS) measurements are used to characterize the spatial structure and properties of the vertically stacked MoS2/Sb2Te3 system. Raman spectra are collected via 532 nm laser illumination. The Raman signal of MoS2 showed two distinct peaks at 384 cm−1 and 409 cm−1 in the spectrum Figure 2d, which indicated the in-plane Mo-S phonon mode (E1 2g at 384 cm−1) and out-of-plane Mo-S phonon mode (A1g at 409 cm−1) of MoS2, respectively [34, 35]. Raman Spectra evidence the MoS2 multilayer film with a stoke shift of 25 cm−1 [36]. The Raman spectrum of Sb2Te3 consists of different peaks related to Sb2Te3 vibrations: 70 cm−1, 114 cm−1, and 130 cm−1 with two minor peaks centered at 97 cm−1, and 105 cm−1 between, and 146 cm−1, 167 cm−1. The 70 cm−1, 97 cm−1, and 167 cm−1 peaks indicated the A1g and Eg normal modes of the Sb–Te vibrations. The 114 cm−1, 105 cm−1, and 146 cm−1 indicate the Te–Te interactions, which together consist of the whole structure of Sb2Te3 [36]. The thickness of Sb2Te3 is estimated to be above 65 nm [37]. Energy dispersive spectroscopy (EDS) is employed to explore the element information collected from the single layers (the exfoliated MoS2 (point C), and Sb2Te3 (point A) and stacked layers (MoS2/Sb2Te3 [point B]) on the device, respectively (Figure 2b). The corresponding MoS2 EDS spectrum shows the L series peaks of Mo and S Kα peak is very pronounced; therefore, only a widened peak with a tail towards the high energy end can be seen with EDS. The elemental compositions (Sb/Te) of the exfoliated Sb2Te3 layers were also determined by EDS. We observed two weak X-ray emission peaks corresponding to Sb and Te. The spectrum also showed the characteristic peaks of Si, which originates from the used substrate. The absence of any other peaks indicates that the Sb2Te3 layers are formed from Sb and Te, only. The quantitative atomic ratio of Sb and Te is roughly 38.7–61.3%, which is near to the 2:3 stoichiometry ratio in accordance with the EDS investigations of Sb2Te3 [3840].

Prior to the optoelectronic measurements, we characteristic the electrical performance of the MoS2/Sb2Te3 junction by applying a source-drain bias voltage V sd from −1 to 1 V to evaluate the dark current and performance of the van der Waals PN heterojunction. As indicated by the black curve in the I-V measurement (inset of Figure 3a–c), the dark currents are stable and consistent. The dark current is as low as 2.38 pA at zero bias and 1.16 nA at 1 V bias, which benefits from the heterostructure design. To our best knowledge, this is the lowest dark current in a vdW heterojunction photodetector compared to other vdW heterojunction photodetectors [4149]. At a negative bias, the dark current becomes significantly higher than at positive voltages. Furthermore, under on and off bias voltage V sd, the source-drain current, I sd has a rectifying characteristic. This reveals that a vdW PN heterojunction is created at the interface of the stacked MoS2/Sb2Te3, since the rectifying effect is observed under different bias voltages. The diode’s on/off ratio, defined as the ratio between photogenerated current and dark current, is around 5.6 × 103. At room temperature, the ultra-low dark current and high on/off ratio at zero bias enable the development of a highly sensitive self-powered photodetector. The power-dependent IV measurement of the PN heterojunction at the wavelength of 500 nm, 700 nm, and 900 nm has also been performed (inset of Figure 3a–c). At zero bias voltage, the heterojunction separates the photogenerated electro-hole pairs and introduces the photocurrent due to the built-in electric field formed at the MoS2/Sb2Te3 interface. The photocurrent I photo increases with higher optical power illumination because electro-hole pairs generation rates increase. This demonstrates self-powered photodetection at various wavelengths (Figure 3a–c) with the insets in Figure 3a–c showing the measured I photo and I sc as a function of the optical power at various wavelengths. The I photo increases.

Figure 3: 
Electrical characterization of Sb2Te3/MoS2 PN junction heterostructure photodetector. (a), (b), (c) The measured photoresponsivity at different wavelengths, 500 nm, 700 nm, and 900 nm, under −1 to 1 V bias voltage. At zero bias, the photoresponsivity was clearly observed and increased proportionally to the optical power. (d) Normalized responsivity by sweeping the wavelength under 0 bias from 500 to 900 nm with a step size of 10 nm, which indicates a broad wavelength response.
Figure 3:

Electrical characterization of Sb2Te3/MoS2 PN junction heterostructure photodetector. (a), (b), (c) The measured photoresponsivity at different wavelengths, 500 nm, 700 nm, and 900 nm, under −1 to 1 V bias voltage. At zero bias, the photoresponsivity was clearly observed and increased proportionally to the optical power. (d) Normalized responsivity by sweeping the wavelength under 0 bias from 500 to 900 nm with a step size of 10 nm, which indicates a broad wavelength response.

Proportionally to the illuminated optical power, which gives evidence of the photovoltaic effect in the vdW PN heterojunction. The responsivity of the Sb2Te3/MoS2 PN junction-based heterostructure photodetector is measured from −1 V to +1 V at the wavelength of 500 nm, 700 nm, and 900 nm using the free-space optical setup. With a positive bias voltage, an external electric field is formed at the junction interface, increasing the separation efficiency of the photogenerated carriers and, therefore, the responsivity. At +1 V bias, the responsivity was amplified 1.5 to 3 times depending on the wavelength and optical power. The responsivity grows with the applied voltage and gradually saturates depending on the incident optical power. Under zero bias, the significant broadband photocurrent of the PN junction photodetector was observed within a wavelength range from 500 to 900 nm (Figure 3d). The highest responsivity under zero bias was 170 mA W−1 at 66 nm.

3 Conclusions

In conclusion, we proposed and experimentally demonstrated a self-powered, extremely sensitive photodetector based on a novel combination of a topological insulator and TMDs van der Waals PN heterojunction. Due to the current rectifying property of the PN heterojunction, the constructed device exhibits extremely low dark current and high responsivities. Under zero bias, the dark current of the device is 2.4 pA, and the responsivity of 170 mA W−1. By increasing the bias voltage to 1 V, the responsivity was magnified thrice due to the external field amplification. This outstanding performance of the proposed photodetector, with its novel material combination, paves the way for further research and applications based on other TMDs and Tis in optoelectronic applications. The PN heterojunction transfer can be easily integrated into waveguide systems, allowing for future applications in PICs.

4 Methods

4.1 Device fabrication

The Sb2Te3/MoS2 PN junction heterostructure is formed using 2D flakes exfoliated from the bulk crystals and transferred by the pick and drop transfer system on Si/SiO2 substrate. The electrical contact pads were formed using e-beam lithography (Raith Pioneer EBL), and electron beam evaporation was employed to deposit the Ti/Au (5 nm/50 nm) electrodes on the produced heterostructures. The lift-off was performed by acetone, then rinsing in isopropyl alcohol and nitrogen drying in RT (room temperature).

4.2 Device experimentation

The tunable (NKT SUPERCONTINUUM Compact) source and the source meter (Keithly 2600B) were used for electrical response measurements of Sb2Te3/MoS2 PN junction heterostructure devices. The laser beam was focused on the devices by an objective lens. The Raman was performed using a 532 nm laser source at room temperature.


Corresponding authors: Hao Wang, Department of Electrical & Computer Engineering, The George Washington University, 800 22nd Street NW 5000 Science & Engineering Hall, Washington, DC 20052, USA; and Optelligence LLC, 10703 Marlboro Pike, Upper Marlboro, MD 20772, USA, E-mail: ; Hamed Dalir, Optelligence LLC, 10703 Marlboro Pike, Upper Marlboro, MD 20772, USA, E-mail: ; and Volker J. Sorger, Department of Electrical & Computer Engineering, The George Washington University, 800 22nd Street NW 5000 Science & Engineering Hall, Washington, DC 20052, USA, E-mail:

Funding source: AFOSR

Award Identifier / Grant number: FA9550-21-1-0425

Acknowledgments

This work was performed in part at the George Washington University Nanofabrication and Imaging Center (GWNIC). We thank Professor S. Solares and his student Mr. M. McCraw from the George Washington University for sharing their resources for AFM measurements.

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This study was funded by AFOSR (FA9550-21-1-0425).

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

References

[1] Y. Xue, Y Zhang, Y Liu, et al.., “Scalable production of a Few-Layer MoS2/WS2 vertical heterojunction array and its application for photodetectors,” ACS Nano, vol. 10, pp. 573–580, 2016. https://doi.org/10.1021/acsnano.5b05596.Search in Google Scholar PubMed

[2] Y. Huang, F. Zhuge, J. Hou, et al.., “Van der Waals coupled organic molecules with monolayer MoS2 for fast response photodetectors with gate-tunable responsivity,” ACS Nano, vol. 12, pp. 4062–4073, 2018. https://doi.org/10.1021/acsnano.8b02380.Search in Google Scholar PubMed

[3] Y. Q. Bie, G. Grosso, M. Heuck, et al.., “A MoTe2-based light-emitting diode and photodetector for silicon photonic integrated circuits,” Nat. Nanotechnol., vol. 12, pp. 1124–1129, 2017. https://doi.org/10.1038/nnano.2017.209.Search in Google Scholar PubMed

[4] R. Cheng, D. Li, H. Zhou, et al.., “Electroluminescence and photocurrent generation from atomically sharp WSe2/MoS2 heterojunction p-n diodes,” Nano Lett., vol. 14, pp. 5590–5597, 2014. https://doi.org/10.1021/nl502075n.Search in Google Scholar PubMed PubMed Central

[5] O. Lopez-Sanchez, E. Alarcon Llad, V. Koman, A. Fontcuberta, I. Morral, A. Radenovic, and A. Kis, “Light generation and harvesting in a van der waals heterostructure,” ACS Nano, vol. 8, pp. 3042–3048, 2014. https://doi.org/10.1021/nn500480u.Search in Google Scholar PubMed PubMed Central

[6] W. Zhang, M. H. Chiu, C. H. Chen, W. Chen, L. J. Li, and A. T. S. Wee, “Role of metal contacts in high-performance phototransistors based on WSe2 monolayers,” ACS Nano, vol. 8, pp. 8653–8661, 2014. https://doi.org/10.1021/nn503521c.Search in Google Scholar PubMed

[7] H. Huang, J. Wang, W. Hu, et al.., “Highly sensitive visible to infrared MoTe2 photodetectors enhanced by the photogating effect,” Nanotechnology, vol. 27, pp. 1–7, 2016. https://doi.org/10.1088/0957-4484/27/44/445201.Search in Google Scholar PubMed

[8] J. Ning, J. C. Martinez, J. Momand, et al.., “Differences in Sb2Te3 growth by pulsed laser and sputter deposition,” Acta Mater., vol. 200, pp. 811–820, 2020. https://doi.org/10.1016/j.actamat.2020.09.035.Search in Google Scholar

[9] R. A. Ganeev, V. S. Popov, A. I. Zvyagin, et al.., “Exfoliated Bi2Te3 nanoparticle suspensions and films: morphological and nonlinear optical characterization,” Nanophotonics, vol. 10, pp. 3857–3870, 2021. https://doi.org/10.1515/nanoph-2021-0335.Search in Google Scholar

[10] Q. Yin, G. Si, J. Li, et al.., “Self-powered topological insulator Bi2Te3/Ge heterojunction photodetector driven by long-lived excitons transfer,” Nanotechnology, vol. 33, p. 255502, 2022. https://doi.org/10.1088/1361-6528/ac5df7.Search in Google Scholar PubMed

[11] S. Ruan, X. Lin, H. Chen, et al.., “Terahertz probe of nonequilibrium carrier dynamics and ultrafast photocurrents in the topological insulator Sb2Te3,” Appl. Phys. Lett., vol. 118, p. 011102, 2021. https://doi.org/10.1063/5.0027930.Search in Google Scholar

[12] 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

[13] Y.L. Chen, J. G. Analytis, J.-H. Chu, et al.., “Experimental Realization of a Three-Dimensional Topological Insulator, Bi2Te3,” Science, vol. 325, no. 5937, pp. 178–181, 2009. https://doi.org/10.1126/science.1173034.Search in Google Scholar PubMed

[14] L. Fu, C. L. Kane, and E. J. Mele, “Topological insulators in three dimensions,” Phys. Rev. Lett., vol. 98, pp. 1–4, 2007. https://doi.org/10.1103/physrevlett.98.106803.Search in Google Scholar PubMed

[15] S. Z. Butler, S. M. Hollen, L. Cao, et al.., “Progress, challenges, and opportunities in two-dimensional materials beyond graphene,” ACS Nano, vol. 7, pp. 2898–2926, 2013. https://doi.org/10.1021/nn400280c.Search in Google Scholar PubMed

[16] H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, and S.-C. Zhang, “Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface,” Nat. Phys., vol. 5, pp. 438–442, 2009.10.1038/nphys1270Search in Google Scholar

[17] I. A. Rahman and A. Purqon, “First principles study of molybdenum disulfide electronic structure,” J. Phys. Conf. Ser., vol. 877, p. 012026, 2017. https://doi.org/10.1088/1742-6596/877/1/012026.Search in Google Scholar

[18] W. Zhang, J. K. Huang, C. H. Chen, Y. H. Chang, Y. J. Cheng, and L. J. Li, “High-gain phototransistors based on a CVD MoS2 monolayer,” Adv. Mater., vol. 25, pp. 3456–3461, 2013. https://doi.org/10.1002/adma.201301244.Search in Google Scholar PubMed

[19] K. Roy, M. Padmanabhan, S. Goswami, et al.., “Graphene-MoS2 hybrid structures for multifunctional photoresponsive memory devices,” Nat. Nanotechnol., vol. 8, pp. 826–830, 2013. https://doi.org/10.1038/nnano.2013.206.Search in Google Scholar PubMed

[20] 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

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

[22] M. Bernardi, M. Palummo, and J. C. Grossman, “Extraordinary sunlight absorption and one nanometer thick photovoltaics using two-dimensional monolayer materials,” Nano Lett., vol. 13, pp. 3664–3670, 2013. https://doi.org/10.1021/nl401544y.Search in Google Scholar PubMed

[23] J. Lou, X. Xu, and P. D. Ye, “Black phosphorus monolayer MoS2 diode,” ACS Nano, vol. 8, pp. 8292–8299, 2014. https://doi.org/10.1021/nn5027388.Search in Google Scholar PubMed

[24] W. Huang, M. Wang, L. Hu, C. Wang, Z. Xie, and H. Zhang, “Recent advances in semiconducting monoelemental Selenium nanostructures for device applications,” Adv. Funct. Mater., vol. 30, pp. 1–25, 2020. https://doi.org/10.1002/adfm.202003301.Search in Google Scholar

[25] W. Huang, J. Zhu, M. Wang, et al.., “Emerging mono-elemental bismuth nanostructures: controlled synthesis and their versatile applications,” Adv. Funct. Mater., vol. 31, pp. 1–34, 2021. https://doi.org/10.1002/adfm.202007584.Search in Google Scholar

[26] W. Huang, L. Hu, Y. Tang, Z. Xie, and H. Zhang, “Recent advances in functional 2D MXene-based nanostructures for next-generation devices,” Adv. Funct. Mater., vol. 30, pp. 1–32, 2020. https://doi.org/10.1002/adfm.202005223.Search in Google Scholar

[27] S. Li, Y. Chen, H. Liu, et al.., “Graphdiyne materials as nanotransducer for in vivo photoacoustic imaging and photothermal therapy of tumor,” Chem. Mater., vol. 29, pp. 6087–6094, 2017. https://doi.org/10.1021/acs.chemmater.7b01965.Search in Google Scholar

[28] M. Wang, J. Zhu, Y. Zi, et al.., “Functional two-dimensional black phosphorus nanostructures towards next-generation devices,” J. Mater. Chem. A, vol. 9, pp. 12433–12473, 2021. https://doi.org/10.1039/d1ta02027g.Search in Google Scholar

[29] W. Huang, Y. Zhang, Q. You, et al.., “Enhanced photodetection properties of tellurium@selenium roll-to-roll nanotube heterojunctions,” Small, vol. 15, pp. 1–10, 2019. https://doi.org/10.1002/smll.201900902.Search in Google Scholar PubMed

[30] Y. Zi, J. Zhu, M. Wang, et al.., “CdS@CdSe core/shell quantum dots for highly improved self-powered photodetection performance,” Inorg. Chem., vol. 60, pp. 18608–18613, 2021. https://doi.org/10.1021/acs.inorgchem.1c03023.Search in Google Scholar PubMed

[31] M. H. Tahersima, Z. Ma, Y. Gui, et al.., “Coupling-enhanced dual ITO layer electro-absorption modulator in silicon photonics,” Nanophotonics, vol. 8, pp. 1559–1566, 2019. https://doi.org/10.1515/nanoph-2019-0153.Search in Google Scholar

[32] Y. Gui, B. M. Nouri, M. Miscuglio, R. Amin, and H. Wang, “Monolithic ITO Mach – Zehnder interferometer modulator enabling 3500 times higher packing density,” Nanophotonics, vol. 11, no. 17, pp. 4001–4009, 2022. https://doi.org/10.1515/nanoph-2021-0796.Search in Google Scholar

[33] R. Amin, J. K. George, H. Wang, et al.., “An ITO-graphene heterojunction integrated absorption modulator on Si-photonics for neuromorphic nonlinear activation,” APL Photonics, vol. 6, no. 12, p. 120801, 2021. https://doi.org/10.1063/5.0062830.Search in Google Scholar

[34] H. Li, J. Wu, Z. Yin, and H. Zhang, “Preparation and applications of mechanically exfoliated single-layer and multilayer MoS2 and WSe2 nanosheets,” Acc. Chem. Res., vol. 47, pp. 1067–1075, 2014. https://doi.org/10.1021/ar4002312.Search in Google Scholar PubMed

[35] H. Li, Q. Zhang, C. C. R. Yap, et al.., “From bulk to monolayer MoS2: evolution of Raman scattering,” Adv. Funct. Mater., vol. 22, pp. 1385–1390, 2012. https://doi.org/10.1002/adfm.201102111.Search in Google Scholar

[36] M. Thripuranthaka, R. V. Kashid, C. Sekhar Rout, and D. J. Late, “Temperature dependent Raman spectroscopy of chemically derived few layer MoS2 and WS2 nanosheets,” Appl. Phys. Lett., vol. 104, no. 8, p. 081911, 2014. https://doi.org/10.1063/1.4866782.Search in Google Scholar

[37] G. Hao, X. Qi, Y. Fan, et al.., “Spiral growth of topological insulator Sb2Te3 nanoplates,” Appl. Phys. Lett., vol. 102, pp. 1–5, 2013. https://doi.org/10.1063/1.4773587.Search in Google Scholar

[38] G. Zhang, H. Qin, J. Teng, et al.., “Quintuple-layer epitaxy of thin films of topological insulator Bi2Se3,” Appl. Phys. Lett., vol. 95, pp. 95–98, 2009. https://doi.org/10.1063/1.3200237.Search in Google Scholar

[39] K.M.F. Shahil, M. Z. Hossain, V. Goyal, and A. A. Balandin, “Micro-Raman spectroscopy of mechanically exfoliated few-quintuple layers of Bi2Te3,” J. Appl. Phys., vol. 111, no. 5, p. 054305, 2012.10.1063/1.3690913Search in Google Scholar

[40] G. C. Sosso, S. Caravati, and M. Bernasconi, “Vibrational properties of crystalline Sb2Te3 from first principles,” J. Phys. Condens. Matter, vol. 21, no. 9, p. 095410, 2009. https://doi.org/10.1088/0953-8984/21/9/095410.Search in Google Scholar PubMed

[41] J. D. Yao, Z. Q. Zheng, J. M. Shao, and G. W. Yang, “Stable, highly-responsive and broadband photodetection based on large-area multilayered WS2 films grown by pulsed-laser deposition,” Nanoscale, vol. 7, pp. 14974–14981, 2015. https://doi.org/10.1039/c5nr03361f.Search in Google Scholar PubMed

[42] A. Varghese, D. Saha, K. Thakar, et al.., “Near-direct bandgap WSe2/ReS2 type-II pn heterojunction for enhanced ultrafast photodetection and high-performance photovoltaics,” Nano Lett., vol. 20, pp. 1707–1717, 2020. https://doi.org/10.1021/acs.nanolett.9b04879.Search in Google Scholar PubMed

[43] Z. Q. Fan, X. W. Jiang, J. Chen, and J. W. Luo, “Improving performances of in-plane transition-metal dichalcogenide Schottky barrier field-effect transistors,” ACS Appl. Mater. Interfaces, vol. 10, pp. 19271–19277, 2018. https://doi.org/10.1021/acsami.8b04860.Search in Google Scholar PubMed

[44] J. W. Chen, S. T. Lo, S. C. Ho, et al.., “A gate-free monolayer WSe2 pn diode,” Nat. Commun., vol. 9, no. 1, p. 3143, 2018. https://doi.org/10.1038/s41467-018-05326-x.Search in Google Scholar PubMed PubMed Central

[45] J. Xiao, Y. Zhang, H. Chen, N. Xu, and S. Deng, “Enhanced performance of a monolayer MoS2/WSe2 heterojunction as a photoelectrochemical cathode,” Nano-Micro Lett., vol. 10, no. 4, p. 60, 2018. https://doi.org/10.1007/s40820-018-0212-6.Search in Google Scholar PubMed PubMed Central

[46] L. Wang, J. Jie, Z. Shao, et al.., “MoS2/Si heterojunction with vertically standing layered structure for ultrafast, high-detectivity, self-driven visible-near infrared photodetectors,” Adv. Funct. Mater., vol. 25, pp. 2910–2919, 2015. https://doi.org/10.1002/adfm.201500216.Search in Google Scholar

[47] R. Tian, X. Gan, C. Li, et al.., “Chip-integrated van der Waals PN heterojunction photodetector with low dark current and high responsivity,” Light Sci. Appl., vol. 11, no. 1, p. 101, 2022. https://doi.org/10.1038/s41377-022-00784-x.Search in Google Scholar PubMed PubMed Central

[48] M. Dai, H. Chen, F. Wang, et al.., “Ultrafast and sensitive self-powered photodetector featuring self-limited depletion region and fully depleted channel with van der Waals contacts,” ACS Nano, vol. 14, pp. 9098–9106, 2020. https://doi.org/10.1021/acsnano.0c04329.Search in Google Scholar PubMed

[49] N. Huo, J. Kang, Z. Wei, S. S. Li, J. Li, and S. H. Wei, “Novel and enhanced optoelectronic performances of multilayer MoS2 – WS2 heterostructure transistors,” Adv. Funct. Mater., vol. 24, pp. 7025–7031, 2014. https://doi.org/10.1002/adfm.201401504.Search in Google Scholar

Received: 2022-08-10
Accepted: 2022-10-16
Published Online: 2022-10-28

© 2022 the author(s), published by De Gruyter, Berlin/Boston

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

Articles in the same Issue

  1. Frontmatter
  2. Editorial
  3. Novel two-dimensional materials based bio-nanophotonics
  4. Reviews
  5. Carbon nanomaterials for phototherapy
  6. Two-dimensional transition metal carbides and nitrides (MXenes) based biosensing and molecular imaging
  7. Two-dimensional photonic MXene nanomedicine
  8. Two-dimensional nanostructures based ‘-onics’ and ‘-omics’ in personalized medicine
  9. Emerging nanophotonic biosensor technologies for virus detection
  10. Research Articles
  11. Three birds with one stone: oxygen self-supply engineering palladium nanocluster/titanium carbide hybrid for single-NIR laser-triggered synergistic photodynamic-photothermal therapy
  12. Oxygen-supplied electrotherapy for enhanced photodynamic synergistic therapy overcomes hypoxia tumor microenvironment
  13. NIR-II light-activated two-photon squaric acid dye with Type I photodynamics for antitumor therapy
  14. Highly active selenium nanotherapeutics combined with metformin to achieve synergistic sensitizing effect on NK cells for osteosarcoma therapy
  15. Self-powered broadband photodetector based on MoS2/Sb2Te3 heterojunctions: a promising approach for highly sensitive detection
  16. Fluorescein-derived carbon dots with chitin-targeting for ultrafast and superstable fluorescent imaging of fungi
  17. A pH-sensitive liposomal co-delivery of fingolimod and ammonia borane for treatment of intracerebral hemorrhage
  18. Exploring magneto-optic properties of colloidal two-dimensional copper-doped CdSe nanoplatelets
  19. Nanoparticle-on-mirror pairs: building blocks for remote spectroscopies
  20. Mesoporous biophotonic carbon spheres with tunable curvature for intelligent drug delivery
  21. Manganese-functionalized MXene theranostic nanoplatform for MRI-guided synergetic photothermal/chemodynamic therapy of cancer
  22. Copper decorated Ti3C2 nanosystem with NIR-II-induced GSH-depletion and reactive oxygen species generation for efficient nanodynamic therapy
Downloaded on 25.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2022-0413/html
Scroll to top button