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

Direct observation of enhanced performance in suspended ReS2 photodetectors

Open Access Open Access

Abstract

Rhenium disulfide (ReS2) has emerged as a promising material for future optoelectric applications due to its extraordinary electrical, mechanical, and optoelectrical properties. However, the ReS2-based photodetectors are severely restricted by their slow response speed (>10 s). Here, we demonstrate a high-performance polarization-sensitive photodetector based on suspended ReS2. Such a transistor shows an n-type behavior with the mobility of about 14.1 cm2V−1s−1, an on/off ratio of 105, and a responsivity of 0.22 A/W. Benefitting from well-developed contact between Au and the ReS2 channel and reduced interface scattering from the Si substrate, the response time of the device can be as short as 83.5 and 325.3µs, respectively, which are three orders of magnitude faster than that reported earlier. Furthermore, the suspended ReS2 photodetector also has the capability to detect polarized light (Imax/Imin ≈ 1.4 at 532 nm) due to the robust in-plane anisotropy of the material. These findings offer an efficient approach for improving the performance of ReS2-based photodetectors.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Two-dimensional (2D) materials have been widely studied because of their unique electrical [13] and optical [48] properties. With high carrier mobility and good mechanical flexibility, graphene has been attracting much attention. However, the practical applications of graphene are limited by its zero bandgap [911]. Recently, transition metal dichalcogenides (TMDs) with the chemical formula of MX2 (M being a transition metal and X being a chalcogen) have emerged as promising materials for electronics and photoelectric applications owing to their sizeable bandgaps and other excellent properties. With a moderate bandgap (while maintaining a direct bandgap from monolayer to the bulk) and intrinsic anisotropy, Rhenium disulfide (ReS2) has received considerable attention among 40 kinds of TMD materials in the field of transistors and photodetectors. For instance, a ReS2 based transistor achieved a high current on/off ratio of ∼104 and an ultra-low electron mobility of 0.44 cm2V−1s−1 using graphene as the source/drain electrodes [12]. Furthermore, photodetectors based on ReS2 [9,13] possess a high responsivity, outstanding external quantum efficiency, and the ability to detect polarized light. However, the slow response time (>100s) of ReS2 originating from its low electron mobility hinders its application. Therefore, to obtain excellent photodetection performance, it is critical to design appropriate ReS2 structures with high electron mobility while maintaining a high current on-off ratio with good polarization sensitivity.

In this article, we report a high-performance phototransistor based on suspended ReS2. The electrical and optoelectronic properties of the suspended ReS2 device were characterized in detail. Specifically, the obtained devices were found to exhibit an excellent current on-off ratio of 105, a high responsivity of 0.22 A/W, and a fast response time of 83.5/325.3µs. Additionally, the devices maintained the capability of ReS2 in detecting polarized light, achieving a linearly dichroic ratio of 1.4. These high performances demonstrate the suspending of ReS2 as a unique and simple way for developing high-performance optoelectronic devices.

Monolayer ReS2 can be mechanically exfoliated through scotch-tape due to its weak van der Waals interlayer interactions [14]. The inset in Fig. 1(a) shows the optical image of monolayer ReS2. The schematic of the complete monolayer ReS2 device is displayed in Fig. 1(b), and the corresponding process of device fabrication is illustrated in Fig. 1(c), including the following steps: (i) The Au film (40 nm) was deposited onto the SiO2/Si substrate by thermal evaporation and standard photolithography, and few-layer ReS2 was mechanically exfoliated from bulk ReS2 onto another SiO2/Si substrate. (ii) With the help of a microscope, the ReS2 nanosheet was transferred onto the two portions of Au film through wet transfer (PPC as dielectrics) [1518]. (iii) The PPC was removed in acetone finally. With such a process, a dual-electrode back-gate field-effect transistor based on suspended ReS2 was completed. It should be emphasized that short channel (<5µm) electrodes were applied to reduce the stress from PPC. In Fig. 1(a), it is clear that there is no obvious damage in the monolayer ReS2 channel. Furthermore, we monitored the few-layer ReS2 quality by acquiring a Raman spectrum continuously throughout the entire fabrication process. Figure 1(d) shows the Raman spectrum of the few-layer ReS2 before (red) and after (blue) the transfer process. Owing to the asymmetric crystal structure of ReS2, more than 12 Raman modes were observed in Fig. 1(d), the same as the reported results on ReS2 [9,14,19,20]. These Raman modes were resulted from the out-of-plane vibrations of Re atoms (Ag-like) and the in-plane vibrations of Re atoms (Eg-like) [16]. PL measurements were also carried out on the few-layer ReS2 before and after the transfer process and were plotted in the inset of Fig. 1(d), respectively. It is noteworthy that the PL spectra of the ReS2 nanosheet reveal a tunable emission peak from 1.49eV to 1.51eV, a feature that might be attributed to the strained ReS2 caused during the device manufacturing process.

 figure: Fig. 1.

Fig. 1. (a) Optical microscope image of a suspended ReS2 device. The inset shows a typical optical image of the transistor. (b) Schematic of the complete device. (c) Device-fabrication process. (d) The Raman spectra and PL spectra of ReS2 flakes measured before and after the transfer process. (e) Measured peak intensities of the 210 cm−1 modes of ReS2 after the transfer process as a function of the excitation polarization direction. 0° is defined as the angle at which the polarization direction of the incident light is parallel to the Re-chain.

Download Full Size | PDF

Importantly, for identification of the Re-direction and the convenience of the subsequent measurements [2123], the Mode V (210cm−1) in ReS2 was considered and mentioned in Fig. 1(d). Firstly, Re-chain tends to coincide with one of the cleaved edges of ReS2, and one of them can be defined as 0°. Secondly, the Raman intensity of Mode V reaches its maximum when the laser polarization is parallel to the Re-chain. Therefore, if the maximum occurs at 0°, it means that the Re-chain is parallel to this edge; otherwise, it is the other edge. Here, a half-wave plate was used to manipulate the polarization direction of the incident light. P=0 means that the polarization direction of the incident light is along the Re-chain of ReS2. The corresponding polar plots in Fig. 1(e) show the polarization dependence of the intensity of Mode V measured in the parallel configuration, from which the Re-chain direction was determined.

To investigate the I-V characteristics of suspended ReS2 transistors, the IDS-VDS outputs and IDS-Vg transfer characteristics were determined. Here, two devices were fabricated and named device #1 (suspended ReS2) and #2 (supported ReS2). Evidently, the IDS of device #1 had an increase of one order of magnitude than that of device #2 as shown in Fig. 2(a), attributing to the suspension of ReS2. First, the physical transfer of the ReS2 channel was a gentle and low-energy manufacturing technique that did not require to go through the traditional fabrication processes (such as thermal evaporation and standard photolithography) to prevent defects, residues, and impurities on the 2D ReS2 surface as shown in Fig. 2(b). In other words, the ReS2 channel of device #1 was unbroken due to its gentle preparation method. Second, the ReS2 was not in contact with the substrate, thus leading to the reduction of carrier scattering losses from the SiO2/ReS2. In short, a suspended ReS2 device could eliminate the effects of Fermi-level pinning making the interface to approach an ideal physical model. In addition, the curvature of device #1 was symmetric and linear, indicating the ohmic contact in device #1, while it was Schottky contact in device #2.

 figure: Fig. 2.

Fig. 2. (a) Current-voltage (I-V) characteristics of the device#1 (red) and #2 (blue). (b) Cross-sectional schematics of the transferred ReS2 on top of Au electrode and conventional Au electrodes on top of ReS2. (c) Output characteristics of the device at P = 0 V under different back gate changes from −40 to 40V. (d) Transfer characteristics of the device at P = 0 V under different bias, from 0.1 to 1V.

Download Full Size | PDF

To further understand the contact between the suspended ReS2 and Au, the I-V curves under different Vg values, varied from −40V to 40 V, were plotted in Fig. 2(c). Apparently, the linear curves indicate the well-developed contact between ReS2 and Au in device #1. Figure 2(d) presents the transfer characteristics of the device under different source-drain voltages. The threshold voltage Vth was around −33 V, indicating that the majority carriers were electrons, as observed in most of the TMDs-based transistors [24,25]. At the gate voltages from −40 V to 35 V, the source-drain current was suppressed (∼20 pA) at the source-drain voltage of 0.5 V, and it increased in accordance with the increase in the source-drain voltage from 0.1 to 1 V. Generally, with an increase in the source-drain voltage, the source-drain current would increase, resulting from its proportional relationship with the electric field between two the electrodes. Additionally, under a source-drain bias of 1V, a high on/off ratio of 105 was realized, similar to the reported result [9]. Then the carrier mobility of ReS2 can be extracted from the following formula:

$$\mu = \frac{L}{{WC{V_{DS}}}}\frac{{d{I_{DS}}}}{{d{V_G}}}\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; $$
where L, W, $d{I_{DS}}/d{V_G}\; $, and C are the channel length, channel width, the maximum of the slope, and the capacitance between the channel and the back gate per unit area (C=εoεr/d; εo is the vacuum permittivity; εr is the relative permittivity, and d is the thickness of dielectric), respectively. It is noteworthy that both SiO2 and the air underneath the suspended 2D ReS2 function as the dielectric, which can be regarded as connected in series. Therefore, C is the sum of the ${C_{Si{O_2}}}$ and ${C_{air}}$:
$$C = {C_{Si{O_2}}} \bullet {C_{air}}/({C_{Si{O_2}}} + {C_{air}})$$
The mobility of the suspended ReS2 FET was calculated to be 14.1 cm2V−1s−1, a distinct improvement compared to most of the published reports [12,26,27].

To investigate the photoresponse of device#1, the measuremnts were acquired in the dark and under laser illumination with a diameter of 0.5mm, separately. As shown in Fig. 3(a), the device displayed linear I−V characteristics under the bias voltages from −1 to 1V. The IDS increased linearly with the source-drain bias, which can be attributed to the increase in carrier drift velocity and reduction of the carrier transit time [24], leading to a more significant channel current. Furthermore, as shown in the inset of Fig. 3(a), the photocurrent Iph (Iph = Ilight − Idark) also displayed a linear relationship with the bias voltage varying from 0 to 1 V, confirming that a significant photocurrent can be obtained through a stronger bias voltage. In Fig. 3(b), under illumination with a light density of 10.68 mW/cm2, the VCNP (voltage of charge-neutral point) shifted to a higher negative voltage (≈−39 V) compared with that in dark (≈−34 V), suggesting the existence of trap states within ReS2. These trap states can trap photo-generated electrons and generate a gate electric field to adjust the channel conductance.

 figure: Fig. 3.

Fig. 3. (a) Output characteristics of the device at Vg = 0V under different illumination powers from 0 to 15.96 mW/cm2. (b) Transfer characteristics of the device at Vg = 0 V under different illumination powers from 0 to 10.68 mW/cm2. (c) Photocurrent response of the ReS2 photodetector under various light powers, fixing VDS at 1 V and Vg at 0 V. (d) The photocurrent response of device#1 and #2 revealing the rising time and falling time. The photocurrent response of device #2 was scaled for comparison. (e) Rise and fall times of the high-resolution photocurrent response of the suspended device. (f) Photoswitching before and after 500 cycles demonstrating stability of the suspended ReS2.

Download Full Size | PDF

The transient photoresponse was obtained by measuring the change of ID during on and off switching of illumination, realized through an optical chopper. As displayed in Fig. 3(c), when the frequency and source-drain voltage were set as 2Hz and 1V, respectively, the photoresponse curves of suspended ReS2 under illumination with different laser intensities, demonstrated excellent repeatability and periodicity of photoswitching. It is clear that the photocurrent increased gradually with the increasing of light intensity under a fixed VDS. For instance, the Ilight of the device under a light density of 20.2 mW/cm2 was approximately three times larger than that under 6.23 mW/cm2. Here, the rising (falling) time is defined as the time interval required for the response to rise (decay) from 10% (90%) to 90% (10%). As the light was turned on (off), ID increased (dropped) rapidly to the maximum (minimum) with a rising (falling) time of only ≈ 15 ms (11 ms). An oscilloscope was introduced into the detection system to obtain the fastest response speed of the suspended ReS2 device. The rising and falling times were measured to be 83.5 and 325.3µs, respectively (Fig. 3(e)), which are over three orders of magnitude compared to that of the supported device#2 (4 and 13s) and those obtained in previous work. Furthermore, the curve in Fig. 3(f) can keep nearly unchanged after 500 cycles of illumination in ambient conditions indicating the superior stability of the suspended ReS2. An essential performance metric of photodetectors is the responsivity, $R = \frac{{{I_{ph}}}}{{P\frac{{{S_s}}}{{{S_L}}}}}$, where Iph, P, SL, and SS represent photocurrent, the power density of incident light, the area of the laser beam, and the area of ReS2 channel, respectively. As shown in Fig. 3(d), the responsivity for device #1 was extracted to be 0.22 A/W [21,28,29], which is superior to device#2 (0.05 A/W) and previously reported photodetectors based on other 2D materials. It is worth mentioning that optical absorption was anticipated to play a certain role for the enhanced photo-responsivity. From the Fig. 4(a), we can see that an excitonic peak located at ≈826nm is clearly identified in the spectrum, which is consistent with the PL peak at ≈1.501 eV (1.51 eV in Fig. 1(d)). The difference in absorption spectrum between the suspended ReS2 and the supported ReS2, originating from the reflection of photons (the inset in Fig. 4(a)), indicates the significance of optical absorption.

 figure: Fig. 4.

Fig. 4. (a) Absorption spectrum of suspended ReS2 and supported ReS2. Inset: the reflection process of photons. (b) Photocurrent mapping of the suspended ReS2 device under 0° (top) and 90° (bottom) polarization, showing prominent linear dichroic photodetection. Schematic band diagrams of the Au-ReS2-Au structure in a thermal equilibrium state (c) and under a source−drain bias (d).

Download Full Size | PDF

The fast response time and superior stability under the illumination of suspended ReS2 offered us an opportunity to study polarization-sensitive photocurrent mapping of suspended ReS2. Here, the polarized direction of the incident light at 532 nm was adjusted by a half-wave plate. We performed scanning photocurrent measurements, achieved by scanning the surface of the photodetector with a focused 300nm light spot on the suspended device at VDS = 1V under 0° (top) polarization and 90° (bottom) polarization. It was clear intuitively from Fig. 4(b) that there was a remarkable difference in the current intensity between 0° polarization and 90° polarization, originating from the more significant absorption of ReS2 channel under the 0° polarization [21,2830]. The anisotropic photocurrent ratio, defined as Imax/Imin, was calculated to be 1.4, which is comparable with that of other photodetectors based on traditional 2D materials [3135]. The results indicate that suspended ReS2 still has enormous potential in the polarization-sensitive photodetector. Note that the presence of the strongest photoresponse area was in the boundary of ReS2 and Au. The observed phenomenon of suspended ReS2 photodetector can be explained by the energy band diagrams in Fig. 4(b) and 4(c). When the first Schottky junction was exposed to the light spot, electron-hole pairs generated in the ReS2 and photo-generated carriers were separated, resulting from the built-in field between ReS2­ and Au, however, the holes were trapped in ReS2. Under a source-drain voltage, the holes continued to accumulate, and the electrons could recirculate along the ReS2 channel, lowering the Schottky barrier. Therefore, a strong photoresponsivity was observed. When the light spot was shed on the other Schottky junction, the reverse built-in electric field caused the electrons and holes to drift in opposite directions as compared to the first junction. Under the same bias, the separation of the electron-hole pairs was suppressed, thus causing a weak photoresponsivity.

In summary, we successfully fabricated high-performance photodetector based on suspended ReS2, with a high current on/off ratio (105) and a high responsivity (0.22 A/W) at room temperature. By suspending the ReS2 on the Au electrodes, the response times (83.5/325.3µs) of our device are three orders of magnitude faster than that of ReS2 reported earlier due to the reduced interface scattering and ohmic contact. Furthermore, the photodetector based on suspended ReS2 exhibited a strong polarization-dependent photoresponse with a linearly dichroic ratio of 1.4, which is comparable with those of other reported 2D materials. Our results provide the means for increasing the response speed of photodetectors by using suspended 2D materials as channels that offer insights into their potential applications in optical and optoelectronic devices.

Funding

National Natural Science Foundation of China (61775241); Youth Innovation Promotion Association of the Chinese Academy of Sciences (2019012); Hunan Provincial Science Fund for Distinguished Young Scholars (2020JJ2059); Key Project of Research and Development Plan of Hunan Province (2019GK2233); Science and Technology Innovation Basic Research Project of Shenzhen (JCYJ20180307151237242); Hunan Province Graduate Research and Innovation Project (CX20190177); Fundamental Research Funds for Central Universities of the Central South University (1053320182462); Innovative and entrepreneurial projects created by teachers and students of Central South University (2018gczd031); Project of State Key Laboratory of High Performance Complex Manufacturing, Central South University (ZZYJKT2020-12).

Disclosures

The authors declare that they have no competing interests.

References

1. L. Fang, H. T. Chen, X. M. Yuan, H. Huang, G. Chen, L. Li, J. N. Ding, J. He, and S. H. Tao, “Quick Optical Identification of the Defect Formation in Monolayer WSe2 for Growth Optimization,” Nanoscale Res. Lett. 14(1), 274 (2019). [CrossRef]  

2. L. Fang, X. M. Yuan, K. W. Liu, L. Li, P. Zhou, W. Ma, H. Huang, J. He, and S. H. Tao, “Direct bilayer growth: a new growth principle for a novel WSe2 homo-junction and bilayer WSe2 growth,” Nanoscale 12(6), 3715–3722 (2020). [CrossRef]  

3. Y. Liu, L. Cao, J. Zhong, J. Yu, J. He, and Z. Liu, “Synthesis of bismuth selenide nanoplates by solvothermal methods and its stacking optical properties,” J. Appl. Phys. 125(3), 035302 (2019). [CrossRef]  

4. L. K. Cao, J. H. Zhong, J. Yu, C. Zeng, J. N. Ding, C. X. Cong, X. F. Yue, Z. W. Liu, and Y. P. Liu, “Valley-polarized local excitons in WSe2/WS2 vertical heterostructures,” Opt. Express 28(15), 22135–22143 (2020). [CrossRef]  

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

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

7. Y. Liu, K. Tom, X. Zhang, S. Lou, Y. Liu, and J. Yao, “Alloying effect on bright–dark exciton states in ternary monolayer MoxW1–xSe2,” New J. Phys. 19(7), 073018 (2017). [CrossRef]  

8. Y. Liu, Q. Xia, J. He, and Z. Liu, “Direct Observation of High Photoresponsivity in Pure Graphene Photodetectors,” Nanoscale Res. Lett. 12(1), 93 (2017). [CrossRef]  

9. X. C. Yu, Z. G. Dong, Y. P. Liu, T. Liu, J. Tao, Y. Q. Zeng, J. K. W. Yang, and Q. J. Wang, “A high performance, visible to mid-infrared photodetector based on graphene nanoribbons passivated with HfO2,” Nanoscale 8(1), 327–332 (2016). [CrossRef]  

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

11. 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]  

12. A. Dathbun, Y. Kim, S. Kim, Y. Yoo, M. S. Kang, C. Lee, and J. H. Cho, “Large-Area CVD-Grown Sub-2 V ReS2 Transistors and Logic Gates,” Nano Lett. 17(5), 2999–3005 (2017). [CrossRef]  

13. E. Zhang, Y. B. Jin, X. Yuan, W. Y. Wang, C. Zhang, L. Tang, S. S. Liu, P. Zhou, W. D. Hu, and F. X. Xiu, “ReS2-Based Field-Effect Transistors and Photodetectors,” Adv. Funct. Mater. 25(26), 4076–4082 (2015). [CrossRef]  

14. S. Tongay, H. Sahin, C. Ko, A. Luce, W. Fan, K. Liu, J. Zhou, Y.-S. Huang, C.-H. Ho, J. Yan, D. F. Ogletree, S. Aloni, J. Ji, S. Li, J. Li, F. M. Peeters, and J. Wu, “Monolayer behaviour in bulk ReS2 due to electronic and vibrational decoupling,” Nat. Commun. 5(1), 3252 (2014). [CrossRef]  

15. R. Frisenda, E. Navarro-Moratalla, P. Gant, D. P. De Lara, P. Jarillo-Herrero, R. V. Gorbachev, and A. Castellanos-Gomez, “Recent progress in the assembly of nanodevices and van der Waals heterostructures by deterministic placement of 2D materials,” Chem. Soc. Rev. 47(1), 53–68 (2018). [CrossRef]  

16. Y. Feng, W. Zhou, Y. Wang, J. Zhou, E. Liu, Y. Fu, Z. Ni, X. Wu, H. Yuan, F. Miao, B. Wang, X. Wan, and D. Xing, “Raman vibrational spectra of bulk to monolayer ReS2 with lower symmetry,” Phys. Rev. B 92(5), 054110 (2015). [CrossRef]  

17. 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]  

18. P. J. Zomer, M. H. D. Guimarães, J. C. Brant, N. Tombros, and B. J. van Wees, “Fast pick up technique for high quality heterostructures of bilayer graphene and hexagonal boron nitride,” Appl. Phys. Lett. 105(1), 013101 (2014). [CrossRef]  

19. D. A. Chenet, O. B. Aslan, P. Y. Huang, C. Fan, A. M. van der Zande, T. F. Heinz, and J. C. Hone, “In-Plane Anisotropy in Mono- and Few-Layer ReS2 Probed by Raman Spectroscopy and Scanning Transmission Electron Microscopy,” Nano Lett. 15(9), 5667–5672 (2015). [CrossRef]  

20. E. Liu, Y. Fu, Y. Wang, Y. Feng, H. Liu, X. Wan, W. Zhou, B. Wang, L. Shao, C.-H. Ho, Y.-S. Huang, Z. Cao, L. Wang, A. Li, J. Zeng, F. Song, X. Wang, Y. Shi, H. Yuan, H. Y. Hwang, Y. Cui, F. Miao, and D. Xing, “Integrated digital inverters based on two-dimensional anisotropic ReS2 field-effect transistors,” Nat. Commun. 6(1), 6991 (2015). [CrossRef]  

21. Y. H. Chang, W. Zhang, Y. Zhu, Y. Han, J. Pu, J. K. Chang, W. T. Hsu, J. K. Huang, C. L. Hsu, M. H. Chiu, T. Takenobu, H. Li, C. I. Wu, W. H. Chang, A. T. S. Wee, and L. J. Li, “Monolayer MoSe2 Grown by Chemical Vapor Deposition for Fast Photodetection,” ACS Nano 8(8), 8582–8590 (2014). [CrossRef]  

22. Y. Choi, K. Kim, S. Y. Lim, J. Kim, J. M. Park, J. H. Kim, Z. Lee, and H. Cheong, “Complete determination of the crystallographic orientation of ReX2 (X = S, Se) by polarized Raman spectroscopy,” Nanoscale Horiz. 5(2), 308–315 (2020). [CrossRef]  

23. M. Hafeez, L. Gan, A. S. Bhatti, and T. Y. Zhai, “Rhenium dichalcogenides (ReX2, X = S or Se): an emerging class of TMDs family,” Mater. Chem. Front. 1(10), 1917–1932 (2017). [CrossRef]  

24. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive photodetectors based on monolayer MoS2,” Nat. Nanotechnol. 8(7), 497–501 (2013). [CrossRef]  

25. J. Xia, X. Huang, L. Z. Liu, M. Wang, L. Wang, B. Huang, D. D. Zhu, J. J. Li, C. Z. Gu, and X. M. Meng, “CVD synthesis of large-area, highly crystalline MoSe2 atomic layers on diverse substrates and application to photodetectors,” Nanoscale 6(15), 8949–8955 (2014). [CrossRef]  

26. X. B. Li, F. F. Cui, Q. L. Feng, G. Wang, X. S. Xu, J. X. Wu, N. N. Mao, X. Liang, Z. Y. Zhang, J. Zhang, and H. Xu, “Controlled growth of large-area anisotropic ReS2 atomic layer and its photodetector application,” Nanoscale 8(45), 18956–18962 (2016). [CrossRef]  

27. W. G. Liao, W. Wei, Y. Tong, W. K. Chim, and C. X. Zhu, “Low-Frequency Noise in Layered ReS2 Field Effect Transistors on HfO2 and Its Application for pH Sensing,” ACS Appl. Mater. Interfaces 10(8), 7248–7255 (2018). [CrossRef]  

28. W. Choi, M. Y. Cho, A. Konar, J. H. Lee, G. B. Cha, S. C. Hong, S. Kim, J. Kim, D. Jena, J. Joo, and S. Kim, “High-Detectivity Multilayer MoS2 Phototransistors with Spectral Response from Ultraviolet to Infrared,” Adv. Mater. 24(43), 5832–5836 (2012). [CrossRef]  

29. C. Y. Lan, C. Li, Y. Yin, and Y. Liu, “Large-area synthesis of monolayer WS2 and its ambient-sensitive photo-detecting performance (vol 7, pg 5974, 2015),” Nanoscale 7(42), 17980 (2015). [CrossRef]  

30. A. Arora, J. Noky, M. Druppel, B. Jariwala, T. Deilmann, R. Schneider, R. Schmidt, O. Del Pozo-Zamudio, T. Stiehm, A. Bhattacharya, P. Kruger, S. M. de Vasconcellos, M. Rohlfing, and R. Bratschitsch, “Highly Anisotropic in-Plane Excitons in Atomically Thin and Bulklike 1 T ‘-ReSe2,” Nano Lett. 17(5), 3202–3207 (2017). [CrossRef]  

31. X. T. Wang, Y. Y. Li, L. Huang, X. W. Jiang, L. Jiang, H. L. Dong, Z. M. Wei, J. B. Li, and W. P. Hu, “Short-Wave Near-Infrared Linear Dichroism of Two-Dimensional Germanium Selenide,” J. Am. Chem. Soc. 139(42), 14976–14982 (2017). [CrossRef]  

32. E. Z. Zhang, P. Wang, Z. Li, H. F. Wang, C. Y. Song, C. Huang, Z. G. Chen, L. Yang, K. T. Zhang, S. H. Lu, W. Y. Wang, S. S. Liu, H. H. Fang, X. H. Zhou, H. G. Yang, J. Zou, X. G. Wan, P. Zhou, W. D. Hu, and F. X. Xiu, “Tunable Ambipolar Polarization-Sensitive Photodetectors Based on High-Anisotropy ReSe2 Nanosheets,” ACS Nano 10(8), 8067–8077 (2016). [CrossRef]  

33. Z. Tian, C. L. Guo, M. X. Zhao, R. R. Li, and J. M. Xue, “Two-Dimensional SnS: A Phosphorene Analogue with Strong In-Plane Electronic Anisotropy,” ACS Nano 11(2), 2219–2226 (2017). [CrossRef]  

34. Z. D. Zhang, J. H. Yang, K. Zhang, S. Chen, F. H. Mei, and G. Z. Shen, “Anisotropic photoresponse of layered 2D SnS-based near infrared photodetectors,” J. Mater. Chem. C 5(43), 11288–11293 (2017). [CrossRef]  

35. H.-X. Zhong, S. Gao, J.-J. Shi, and L. Yang, “Quasiparticle band gaps, excitonic effects, and anisotropic optical properties of the monolayer distorted 1 T diamond-chain structures ReS2 and ReSe2,” Phys. Rev. B: Condens. Matter Mater. Phys. 92(11), 115438 (2015). [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. (a) Optical microscope image of a suspended ReS2 device. The inset shows a typical optical image of the transistor. (b) Schematic of the complete device. (c) Device-fabrication process. (d) The Raman spectra and PL spectra of ReS2 flakes measured before and after the transfer process. (e) Measured peak intensities of the 210 cm−1 modes of ReS2 after the transfer process as a function of the excitation polarization direction. 0° is defined as the angle at which the polarization direction of the incident light is parallel to the Re-chain.
Fig. 2.
Fig. 2. (a) Current-voltage (I-V) characteristics of the device#1 (red) and #2 (blue). (b) Cross-sectional schematics of the transferred ReS2 on top of Au electrode and conventional Au electrodes on top of ReS2. (c) Output characteristics of the device at P = 0 V under different back gate changes from −40 to 40V. (d) Transfer characteristics of the device at P = 0 V under different bias, from 0.1 to 1V.
Fig. 3.
Fig. 3. (a) Output characteristics of the device at Vg = 0V under different illumination powers from 0 to 15.96 mW/cm2. (b) Transfer characteristics of the device at Vg = 0 V under different illumination powers from 0 to 10.68 mW/cm2. (c) Photocurrent response of the ReS2 photodetector under various light powers, fixing VDS at 1 V and Vg at 0 V. (d) The photocurrent response of device#1 and #2 revealing the rising time and falling time. The photocurrent response of device #2 was scaled for comparison. (e) Rise and fall times of the high-resolution photocurrent response of the suspended device. (f) Photoswitching before and after 500 cycles demonstrating stability of the suspended ReS2.
Fig. 4.
Fig. 4. (a) Absorption spectrum of suspended ReS2 and supported ReS2. Inset: the reflection process of photons. (b) Photocurrent mapping of the suspended ReS2 device under 0° (top) and 90° (bottom) polarization, showing prominent linear dichroic photodetection. Schematic band diagrams of the Au-ReS2-Au structure in a thermal equilibrium state (c) and under a source−drain bias (d).

Equations (2)

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

μ = L W C V D S d I D S d V G
C = C S i O 2 C a i r / ( C S i O 2 + C a i r )
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.