Next Article in Journal
A Wearable Navigation Device for Visually Impaired People Based on the Real-Time Semantic Visual SLAM System
Previous Article in Journal
Autonomous Ground Vehicle Lane-Keeping LPV Model-Based Control: Dual-Rate State Estimation and Comparison of Different Real-Time Control Strategies
Previous Article in Special Issue
Graphene Electro-Optical Switch Modulator by Adjusting Propagation Length Based on Hybrid Plasmonic Waveguide in Infrared Band
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

The Highly Uniform Photoresponsivity from Visible to Near IR Light in Sb2Te3 Flakes

1
Department of Physics, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
2
Center of Crystal Research, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
3
Department of Materials and Optoelectronic Science, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
4
Taiwan Consortium of Emergent Crystalline Materials, TCECM, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
5
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, Taipei 10607, Taiwan
*
Author to whom correspondence should be addressed.
Sensors 2021, 21(4), 1535; https://doi.org/10.3390/s21041535
Submission received: 17 December 2020 / Revised: 27 January 2021 / Accepted: 19 February 2021 / Published: 23 February 2021
(This article belongs to the Special Issue The Research and Application of Graphene Phototransducer)

Abstract

:
Broadband photosensors have been widely studied in various kinds of materials. Experimental results have revealed strong wavelength-dependent photoresponses in all previous reports. This limits the potential application of broadband photosensors. Therefore, finding a wavelength-insensitive photosensor is imperative in this application. Photocurrent measurements were performed in Sb2Te3 flakes at various wavelengths ranging from visible to near IR light. The measured photocurrent change was insensitive to wavelengths from 300 to 1000 nm. The observed wavelength response deviation was lower than that in all previous reports. Our results show that the corresponding energies of these photocurrent peaks are consistent with the energy difference of the density of state peaks between conduction and valence bands. This suggests that the observed photocurrent originates from these band structure peak transitions under light illumination. Contrary to the most common explanation that observed broadband photocurrent carrier is mainly from the surface state in low-dimensional materials, our experimental result suggests that bulk state band structure is the main source of the observed photocurrent and dominates the broadband photocurrent.

1. Introduction

The interaction of light with matter has been a widely used technique for light sensors and is widely used in our daily life and scientific research. A system with higher photo-interaction efficiency would lead to higher sensitivity. Therefore, there has been increasing interest in finding a material with higher photoresponsivity. The material band structure is a critical factor to finding the efficiency of the light response and light wavelength. As well as the band structure, the light penetration depth is short, so the photon interaction mainly occurs on the material surface. To enhance the photon interaction, a system with higher surface area is preferred [1,2,3,4,5,6]. In addition to the higher surface ratio, earlier studies have demonstrated that carrier mobility is one of the most critical factors for determining the light responsivity [7,8]. With these considerations in mind, nanostructures and low-dimensional systems with high mobility have been widely investigated [9,10,11,12,13,14,15,16]. Experimental results have revealed that most materials only show high photoresponsivity at particular wavelengths due to their specific band structure, and this limits their potential application. A broadband response along with a high photoresponsivity is important for the potential application across a wide range of wavelengths.
A system with a linear energy-momentum band structure, such as graphene, graphene-based heterostructures, and topological materials, could satisfy these requirements and provide a high photoresponsivity over a broadband wavelength range. Many studies have demonstrated high photoresponsivity over a wide range of wavelengths in these systems. The graphene/transition metal dichalcogenide system shows an extremely fast photoresponse [17]. InGaAs shows a photoresponse for light wavelengths ranging from 400 to 16,000 nm [18]. However, previous reports have revealed that the photoresponses show a strong wavelength-dependent deviation in these systems [10,15,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49]. This inhomogeneous response would limit the potential applications of these materials as broadband photosensors. The search for a system with uniform light responsivity over a wide range of wavelengths is one of the on-going topics in the field of light sensor applications. The light penetration into the material and carriers from the bulk state of the topological material might also contribute to the measured photocurrent. This might be one of the reasons behind the observed strong wavelength dependence in the previous reports [50]. To solve this problem, as well as the surface state band structure, one might take the bulk state band structure into account. We checked the bulk state band structure of BixSb(2−x)TeySe(3−y) topological materials and found that Sb2Te3 might be a good candidate to providing uniformity of the photocurrent over a wide range of wavelengths.
In this work, we study the photoresponse of Sb2Te3 flakes. Our results show that the photocurrent is linear with the light power and the applied electric voltage, and that the photoresponsivity is proportional to the conductance of the flake. According to previous reports, the photoresponsivity is stable for light wavelengths ranging from visible to near-infrared light, and the normalized root means that the square photoresponsivity deviation is about 6%, which is much lower than reported results in low-dimensional systems. The results show that the corresponding energy of the measured photocurrent peaks and dips at different light wavelengths are consistent with the energy difference of the carrier density of state peaks and dips between the conduction and valence band of Sb2Te3 bulk state. This result supports that the observed photoresponsivity is dominated by the carriers of the bulk state and not from the topological surface state. The low-deviated band structure of bulk state leads to the behavior of the observed uniform broadband light responsivity.

2. Materials and Methods

High purity elements Sb (99.995%) and Te (99.995%) were mixed according to the stoichiometric ratio in a glove box with low oxygen and water to avoid oxidation. The mixed raw materials were put into a vacuum quartz glass tube (pressure ~ 10−3 torr) and sealed inside the quartz glass tube. After that, the materials were melted at 750 °C for 20 h, and then slowly cooled down to 625 °C in 5 h. It took another 60 h to cool the materials down to 565 °C at the cooling rate 1 °C/h. The melting point of Sb2Te3 is 617 °C, so this slow cooling rate ensured that Sb2Te3 could be slowly crystalized. Then, it took 10 h to cool the materials down to room temperature. The crystallized Sb2Te3 was further purified and crystallized by a homemade resistance-heated floating zone furnace (RHFZ). The Sb2Te3 was locally heated up to 617 °C and the Sb2Te3 formed an interface of solid phase and liquid phase. At the same time, we slowly moved the quartz glass tube and changed the position of the solid/liquid interface. During this process, the impurities were moved with the liquid phase and were excluded from the recrystallized Sb2Te3. After this recrystallization process, single crystals of Sb2Te3were naturally cooled down to room temperature. The as-grown crystals were cleaved along the basal plane, producing a silvery shining mirror-like surface, and then prepared for further experiments. The Raman spectrum and electron probe microanalyzer (EPMA) spectrum support that the produced crystal was Sb2Te3 [51]. Details of Raman spectrum and EPMA spectrum are given in Appendix A. Sb2Te3 flakes were obtained by exfoliating bulk crystals using dicing tape and were then dispersed on the insulating SiO2 (300 nm)/n-Si templates with prepatterned Ti/Au circuits. Two platinum (Pt) metal contacts were subsequently deposited on the selected Sb2Te3 flakes using focused-ion beam (FIB) technique. The Chemical Abstracts Service (CAS) numbers of all the chemical and substrates are given in Appendix B. Details of the x-ray diffraction (XRD), Raman, EPMA, SEM, and electrical measurement system are given in Appendix C.

3. Results and Discussion

Figure 1 shows the XRD pattern of the produced Sb2Te3. The Sb2Te3 crystal showed sharp peaks. It revealed the Laue diffraction peaks at the c-axis. The crystal was grown on the [001] direction. The full width at half maximum (FWHM) was between 0.055 and 0.06 deg. The average crystallite size at the c-axis was about 124 nm and the strain was −0.018%. These results indicate that the Sb2Te3 was highly single crystallized. The Sb2Te3 powder was ground from the Sb2Te3 crystal. The XRD pattern of the Sb2Te3 powder was consistent with the Joint Committee on Powder Diffraction Standards (JCPDS) card.
Figure 2 shows the SEM image of a Sb2Te3 flake. It is 5 µm long and ∼5 µm wide. The thickness of a flake is determined by the atomic force microscopy; here, the flake was 260 nm thick (shown in the left-bottom inset of Figure 2).
The left-top inset of Figure 3 shows the linear current–voltage relationship of the flake; the conductivity is about 20.4 S/cm. The right-bottom inset of Figure 3 shows the on–off photocurrent response at different applied electric voltages under 808 nm light wavelength and 50 mW laser power. In this work, we define the photocurrent as the measured current difference between the condition with and without light illumination. The main image from Figure 3 shows that the extracted photocurrent, I P , is linear with the applied electric voltage. These results indicate the ohmic contacts between the Sb2Te3 flake and Pt electrodes, and the metallic behavior of the Sb2Te3 flake.
Figure 4 shows the measured current of our Sb2Te3 flake under light illumination with light power ranging from 1 to 50 mW, which corresponds to the power intensity of 40 to 2000 Wm−2. The photocurrent increases with increasing light power and shows a similar power-dependent photocurrent under three different light wavelengths. The left-top inset of Figure 4 reveals the extracted photocurrent as a function of the light power at three different light wavelengths. It clearly reveals that the photocurrent varies linearly with the light power for three light wavelengths. The higher light intensity means higher photon interaction. The increasing amount of photon interaction will induce more excited carriers, leading to a higher photocurrent. The light intensity is not over the limit of the excited carrier number, and the measured photocurrent will be proportional to the increasing light intensity, resulting in a positive slope for the three different light wavelengths.
To further identify this characteristic in the three different wavelengths, the photocurrent is expressed as the simple power-law relation I P   =   α P β , where the α is a constant for the wavelength of the illuminating light, P is the illuminated light power on the Sb2Te3 flake, and β is a constant related to the transport mechanism and the intrinsic optic characteristics. Table 1 lists α and β for the three different wavelengths. The result reveals that β is 1.03 ± 0.02, which indicates that the behavior of the laser power-dependent photocurrent of the system remains the same at different light wavelengths. The complex carrier scattering process might lead to the non-integer β. Our experiment shows β∼1, which indicates that the carrier transport process is insensitive to the extrinsic light power and interaction between excited carriers. This characteristic is more flexible and stable for a wide range of potential applications. The broadband light response has been observed in many low-dimensional systems [10,15,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49]. These reports show that the observed photocurrent is strongly dependent on light wavelength, which would limit the application potential of the broadband light sensor. Advancing the findings of previous reports, our experimental results reveal that the photocurrent is weakly wavelength dependent [10,15,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49]. The α roughly deviates by 30% at the three different wavelengths, which is much smaller than the reported values. The light power and the measured photocurrent depended upon the effective geometric area of the material. In order to quantitatively determine the performance of the Sb2Te3 flake under illumination, the responsivity, R, is calculated through the following equation:
R   = I P P S ,
where I P , P, and S are the photocurrent, the light intensity, and the effective area, respectively. The obtained R values are 38, 31, and 43 AW−1 for the wavelengths of 405, 532, and 808 nm at a constant applied voltage of 0.1 V. The deviation ratio is about 34%.
To further examine the photocurrent response characteristic at wide wavelength, a photocurrent spectrum measurement was performed. As shown in the top-right inset of Figure 4, the R reveals weak wavelength dependence over a range of wavelengths. The normalized root means square is used to identify the deviation level. The normalized root means square is 0.06, which indicates a 6% deviation for the light wavelength ranging from 300 to 1000 nm. This behavior is more stable than previously reported results in low-dimensional systems. Along with the responsivity, the responsivity stability is a critical factor for the application of broadband photosensors. Table 2 lists the reported broadband light responsivity of various kinds of materials. It shows that most of these materials reveal a large light response deviation on the broadband light wavelength. The photo deviation of the Sb2Te3 flake is a few orders smaller than the reported values of various materials.
The left side of Figure 5 reveals the density of state of the Sb2Te3 bulk state, and it reveals several peaks at valence and conduction bands. The right side of Figure 5 shows the measured photoresponsivity as a function of light wavelength, and it exhibits several photocurrent peaks at different wavelengths. Our analysis shows that the corresponding energies of these measured photoresponsivity peaks and dips at different light wavelengths are consistent with the energy difference of Density of states (DOS) peaks between conduction and valence bands. This suggests that the observed photocurrent originates from these band structure peak transitions under light illumination. Contrary to most previous explanations that the observed broadband photocurrent carrier is mainly from the surface state in topological materials, our experimental result suggests that bulk state carriers are the main source of the observed photocurrent and dominate the broadband photocurrent.
The photocurrent is directly related to the effective carrier mobility, which is determined by both the carriers from the bulk state and surface state. Figure 6 shows the photocurrent responsivity as a function of the conductance of the Sb2Te3 flakes. It shows that the photocurrent responsivity is proportional to the conductance of the Sb2Te3 flakes. This implies that effective conductance is a critical factor to determining the photocurrent responsivity. Detectivity, which is an important figure of merit in evaluating the ability of a photodetector to detect a weak signal, is another important index used to characterize the performance of photodetectors. The specific detectivity ( D * ) is calculated through the following equation:
D * = R S 1 / 2 2 q I d 1 / 2 ,
where R, S, q, and I d are the responsivity, effective area of light illumination, electronic charge, and dark current, respectively. By using the experimental data, the obtained detectivity is about 2   × 10 9 Jones. As shown in the top-right inset of Figure 6, the detectivity is almost insensitive to the flake thickness. The top-left inset of Figure 6 shows that the sensitivity is about 0.02. Our experimental results show that the detectivity and the sensitivity is insensitive to the flake thickness. However, it should be noted that the light penetration length is about 20 nm, which is much shorter than the thickness of our Sb2Te3 flakes. This might be the reason behind the observed thickness-independent detectivity and sensitivity.

4. Conclusions

The photocurrent measurement was performed on Sb2Te3 flakes at various wavelengths ranging from visible to the near IR region. The photocurrent was observed to vary linearly with the applied electric voltage, light power, and conductance of the flakes. Advancing the findings of previous reports that the photoresponse reveals strong wavelength dependence, our measured photocurrent was insensitive to wavelengths from 300 to 1000 nm. The wavelength response deviation was lower than that in all previously reported photoresponse materials and it has the potential to play a role in broadband photosensors. Our analysis showed that the corresponding energies of these photocurrent peaks were consistent with the energy difference of the density of state peaks between conduction and valence bands. This further suggests that the observed photocurrent originated from these band structure peak transitions under light illumination. Contrary to the most common explanation that the observed broadband photocurrent carrier is mainly from the surface state in low-dimensional materials, our experimental result suggests that the bulk state band structure is the main source of the observed photocurrent and dominates the broadband photocurrent. As well as the linear energy-momentum dispersion, one has to take the bulk state band structure into account for the broadband photo-detection material.

Author Contributions

Conceptualization: S.-M.H.; Methodology: J.-L.H.; Validation: S.-M.H. and J.-L.H.; Formal Analysis: J.-L.H.; Investigation: J.-L.H., C.-Y.C., and F.-C.L.; Resources: M.C. and R.-S.C.; Writing—Original Draft Preparation: S.-M.H.; Writing—Review and Editing: M.C., R.-S.C., and J.-L.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Taiwan National Science Council through Grants No. MOST 106-2112-M110-002, MOST 107-2112-M-110-011-MY2, MOST 109-2112-M-110-018, and Center of Crystal Research at National Sun Yat-Sen University. S.M.H. thanks the support of short-term overseas research for scientist and technician from the Taiwan National Science Council, project Grants No. MOST 108-2918-I-110-007.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The details of Raman spectrum and EPMA: The Raman spectrum shows five peaks, E1g: 44.40 cm−1, E2g: 113.96 cm−1, A1(Te): 126.33 cm−1, EH(Te): 141.72 cm−1, and A11g: 165.46 cm−1. That is consistent with Sb2Te3 oscillation frequencies from previous reports [52,53]. EPMA shows that the ratio of Sb to Te at different portions of the single crystal ranges from Sb:Te = 1.95:3 to 2.00:3.

Appendix B

The CAS numbers of all the chemical and substrates: Antimony (Sb) CAS NO.7440-36-0; Tellurium (Te) CAS NO.13494-80-9; Silicon (Si) CAS NO.7440-21-3; Silicon dioxide (SiO2) CAS NO.7631-86-9; Platinum (Pt) CAS NO.7440-06-4; Gold (Au) CAS NO.7440-57-5; Titanium (Ti) CAS NO.7440-32-6.

Appendix C

X-Ray Diffraction (XRD):
Equipment Model: Siemens D5000 (for powder diffraction measurement); Target: Copper (Cu), Wavelength ( λ K a ) = 0.10540598 nm; Output Power: Voltage = 40 kV, Current = 30 mA; Step Size:   0.05 ° ; Step Time: 1.5 s; Scan Range: 15 ° ~60 ° .
Equipment Model: Bruker D8 (For Crystal diffraction Measurement); Target: Copper (Cu), Wavelength ( λ K a ) = 0.10540598 nm; Output Power: Voltage = 40 kV, Current = 40 mA; Step Size:   0.01 ° ; Step Time: 0.5 s; Scan Range: 15 ° ~60 ° .
Raman:
Equipment Model: Micro-Raman Spectroscopy HR800; Input Laser: He-Cd laser; Wavelength: 633 nm; Scan Range: 25~225 cm−1.
Electron Probe Microanalyzer (EPMA):
Equipment Model: JEOL JXA-8530F Field Emission Electron Probe Microanalyzer; Accelerating Voltage: 10 kV; Magnification: 5000×.
Scanning Electron Microscope (SEM):
Equipment Model: FEI Quanta 3D FEG; Accelerating Voltage: 10 kV; Magnification: 3500×.
Electrical Measurement System:
Equipment Model: Keithley 4200-SCS; Source Type: Voltage; 210; Voltage Range: −0.1~0.1 V and0.005 V Step for Current–Voltage Measurement; Voltage: 0.1 V, 0.2 V, 0.3 V, 0.4 V, 0.5 V and 1 s Step for Current–Time Measurement.

References

  1. Chen, C.; Qiao, H.; Lin, S.; Man Luk, C.; Liu, Y.; Xu, Z.; Song, J.; Xue, Y.; Li, D.; Yuan, J.; et al. Highly responsive MoS2 photodetectors enhanced by graphene quantum dots. Sci. Rep. 2015, 5, 11830. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, R.S.; Wang, W.C.; Chan, C.H.; Lu, M.L.; Chen, Y.F.; Lin, H.C.; Chen, K.H.; Chen, L.C. Photoconduction efficiencies of metal oxide semiconductor nanowires: The material’s inherent properties. Appl. Phys. Lett. 2013, 103, 223107. [Google Scholar] [CrossRef]
  3. Konstantatos, G.; Badioli, M.; Gaudreau, L.; Osmond, J.; Bernechea, M.; de Arquer, F.P.G.; Gatti, F.; Koppens, F.H.L. Hybrid graphene–quantum dot phototransistors with ultrahigh gain. Nat. Nanotechnol. 2012, 7, 363–368. [Google Scholar] [CrossRef]
  4. Roy, K.; Padmanabhan, M.; Goswami, S.; Sai, T.P.; Ramalingam, G.; Raghavan, S.; Ghosh, A. Graphene–MoS2 hybrid structures for multifunctional photoresponsive memory devices. Nat. Nanotechnol. 2013, 8, 826–830. [Google Scholar] [CrossRef] [PubMed]
  5. Soci, C.; Zhang, A.; Xiang, B.; Dayeh, S.A.; Aplin, D.P.R.; Park, J.; Bao, X.Y.; Lo, Y.H.; Wang, D. ZnO Nanowire UV Photodetectors with High Internal Gain. Nano Lett. 2007, 7, 1003–1009. [Google Scholar] [CrossRef]
  6. Zhang, W.; Chuu, C.-P.; Huang, J.-K.; Chen, C.-H.; Tsai, M.-L.; Chang, Y.-H.; Liang, C.-T.; Chen, Y.-Z.; Chueh, Y.-L.; He, J.-H.; et al. Ultrahigh-Gain Photodetectors Based on Atomically Thin Graphene- MoS2 Heterostructures. Sci. Rep. 2014, 4, 3826. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, B.Y.; Liu, T.; Meng, B.; Li, X.; Liang, G.; Hu, X.; Wang, Q.J. Broadband high photoresponse from pure monolayer graphene photodetector. Nat. Commun. 2013, 4, 1811. [Google Scholar] [CrossRef] [Green Version]
  8. Mueller, T.; Xia, F.; Avouris, P. Graphene photodetectors for high-speed optical communications. Nat. Photonics 2010, 4, 297–301. [Google Scholar] [CrossRef] [Green Version]
  9. Hu, P.; Wen, Z.; Wang, L.; Tan, P.; Xiao, K. Synthesis of Few-Layer GaSe Nanosheets for High Performance Photodetectors. ACS Nano 2012, 6, 5988–5994. [Google Scholar] [CrossRef]
  10. Hu, P.; Wang, L.; Yoon, M.; Zhang, J.; Feng, W.; Wang, X.; Wen, Z.; Idrobo, J.C.; Miyamoto, Y.; Geohegan, D.B.; et al. Highly Responsive Ultrathin GaS Nanosheet Photodetectors on Rigid and Flexible Substrates. Nano Lett. 2013, 13, 1649–1654. [Google Scholar] [CrossRef]
  11. Yin, Z.; Li, H.; Li, H.; Jiang, L.; Shi, Y.; Sun, Y.; Lu, G.; Zhang, Q.; Chen, X.; Zhang, H. Single-Layer MoS2 Phototransistors. ACS Nano 2012, 6, 74–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Furchi, M.M.; Polyushkin, D.K.; Pospischil, A.; Mueller, T. Mechanisms of Photoconductivity in Atomically Thin MoS2. Nano Lett. 2014, 14, 6165–6170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Tsai, D.-S.; Liu, K.-K.; Lien, D.-H.; Tsai, M.-L.; Kang, C.-F.; Lin, C.-A.; Li, L.-J.; He, J.-H. Few-Layer MoS2 with High Broadband Photogain and Fast Optical Switching for Use in Harsh Environments. ACS Nano 2013, 7, 3905–3911. [Google Scholar] [CrossRef]
  14. Zhang, W.; Huang, J.-K.; Chen, C.-H.; Chang, Y.-H.; Cheng, Y.-J.; Li, L.-J. High-Gain Phototransistors Based on a CVD MoS2 Monolayer. Adv. Mater. 2013, 25, 3456–3461. [Google Scholar] [CrossRef]
  15. Choi, W.; Cho, M.Y.; Konar, A.; Lee, J.H.; Cha, G.-B.; Hong, S.C.; Kim, S.; Kim, J.; Jena, D.; Joo, J.; et al. High-Detectivity Multilayer MoS2 Phototransistors with Spectral Response from Ultraviolet to Infrared. Adv. Mater. 2012, 24, 5832–5836. [Google Scholar] [CrossRef] [PubMed]
  16. Buscema, M.; Island, J.O.; Groenendijk, D.J.; Blanter, S.I.; Steele, G.A.; van der Zant, H.S.J.; Castellanos-Gomez, A. Photocurrent generation with two-dimensional van der Waals semiconductors. Chem. Soc. Rev. 2015, 44, 3691–3718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Lv, Q.; Lv, R. Two-Dimensional Heterostructures Based on Graphene and Transition Metal Dichalcogenides: Synthesis, Transfer and Applications. Carbon 2019, 145, 240–250. [Google Scholar] [CrossRef]
  18. Passmore, B.S.; Wu, J.; Manasreh, M.O.; Salamo, G.J. Dual broadband photodetector based on interband and intersubband transitions in InAs quantum dots embedded in graded InGaAs quantum wells. Appl. Phys. Lett. 2007, 91, 233508. [Google Scholar] [CrossRef]
  19. Yu, X.-X.; Wang, Y.; Zhang, H.; Zhu, D.; Xiong, Y.; Zhang, W. Fabrication of a Cost Effective and Broadband Self-Powered Photodetector Based on Sb2Te3 and Silicon. Nanotechnology 2019, 30, 345202. [Google Scholar] [CrossRef] [PubMed]
  20. Chander, S.; Purohit, A.; Nehra, A.; Nehra, S.P.; Dhaka, M.S. A Study on Spectral Response and External Quantum Efficiency of Mono-Crystalline Silicon Solar Cell. Int. J. Renew. Energy Res. 2015, 5, 41–44. [Google Scholar]
  21. Lei, S.; Ge, L.; Najmaei, S.; George, A.; Kappera, R.; Lou, J.; Chhowalla, M.; Yamaguchi, H.; Gupta, G.; Vajtai, R.; et al. Evolution of the Electronic Band Structure and Efficient Photo-Detection in Atomic Layers of InSe. ACS Nano 2014, 8, 1263–1272. [Google Scholar] [CrossRef]
  22. Yang, Z.; Jie, W.; Mak, C.-H.; Lin, S.; Lin, H.; Yang, X.; Yan, F.; Lau, S.P.; Hao, J. Wafer-Scale Synthesis of High-Quality Semiconducting Two-Dimensional Layered InSe with Broadband Photoresponse. ACS Nano 2017, 11, 4225–4236. [Google Scholar] [CrossRef]
  23. Feng, W.; Gao, F.; Hu, Y.; Dai, M.; Li, H.; Wang, L.; Hu, P. High-performance and flexible photodetectors based on chemical vapor deposition grown two-dimensional In2Se3 nanosheets. Nanotechnology 2018, 29, 445205. [Google Scholar] [CrossRef] [PubMed]
  24. Mech, R.K.; Solanke, S.V.; Mohta, N.; Rangarajan, M.; Nath, D.N. In2Se3 Visible/Near-IR Photodetector With Observation of Band-Edge in Spectral Response. IEEE Photonics Technol. Lett. 2019, 31, 905–908. [Google Scholar] [CrossRef]
  25. Long, M.; Gao, A.; Wang, P.; Xia, H.; Ott, C.; Pan, C.; Fu, Y.; Liu, E.; Chen, X.; Lu, W.; et al. Room temperature high-detectivity mid-infrared photodetectors based on black arsenic phosphorus. Sci. Adv. 2017, 3, e1700589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wu, J.; Koon, G.K.W.; Xiang, D.; Han, C.; Toh, C.T.; Kulkarni, E.S.; Verzhbitskiy, I.; Carvalho, A.; Rodin, A.S.; Koenig, S.P.; et al. Colossal Ultraviolet Photoresponsivity of Few-Layer Black Phosphorus. ACS Nano 2015, 9, 8070–8077. [Google Scholar] [CrossRef]
  27. Buscema, M.; Groenendijk, D.J.; Blanter, S.I.; Steele, G.A.; van der Zant, H.S.J.; Castellanos-Gomez, A. Fast and Broadband Photoresponse of Few-Layer Black Phosphorus Field-Effect Transistors. Nano Lett. 2014, 14, 3347–3352. [Google Scholar] [CrossRef] [Green Version]
  28. Patel, A.B.; Chauhan, P.; Machhi, H.K.; Narayan, S.; Sumesh, C.K.; Patel, K.D.; Soni, S.S.; Jha, P.K.; Solanki, G.K.; Pathak, V.M. Transferrable thin film of ultrasonically exfoliated MoSe2 nanocrystals for efficient visible-light photodetector. Phys. E Low-dimensional Syst. 2020, 119, 114019. [Google Scholar] [CrossRef]
  29. Zhang, W.; Chiu, M.-H.; Chen, C.-H.; Chen, W.; Li, L.-J.; Wee, A.T.S. Role of Metal Contacts in High-Performance Phototransistors Based on WSe2 Monolayers. ACS Nano 2014, 8, 8653–8661. [Google Scholar] [CrossRef] [Green Version]
  30. Pataniya, P.; Zankat, C.K.; Tannarana, M.; Sumesh, C.K.; Narayan, S.; Solanki, G.K.; Patel, K.D.; Pathak, V.M.; Jha, P.K. Paper-Based Flexible Photodetector Functionalized by WSe2 Nanodots. ACS Appl. Nano Mater. 2019, 2, 2758–2766. [Google Scholar] [CrossRef]
  31. Perea-López, N.; Elías, A.L.; Berkdemir, A.; Castro-Beltran, A.; Gutiérrez, H.R.; Feng, S.; Lv, R.; Hayashi, T.; López-Urías, F.; Ghosh, S.; et al. Photosensor Device Based on Few-Layered WS2 Films. Adv. Funct. Mater. 2013, 23, 5511–5517. [Google Scholar] [CrossRef]
  32. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef] [PubMed]
  33. Yu, X.; Yu, P.; Wu, D.; Singh, B.; Zeng, Q.; Lin, H.; Zhou, W.; Lin, J.; Suenaga, K.; Liu, Z.; et al. Atomically thin noble metal dichalcogenide: A broadband mid-infrared semiconductor. Nat. Commun. 2018, 9, 1545. [Google Scholar] [CrossRef]
  34. Tao, Y.; Wu, X.; Wang, W.; Wang, J. Flexible photodetector from ultraviolet to near infrared based on a SnS2 nanosheet microsphere film. J. Mater. Chem. C 2015, 3, 1347–1353. [Google Scholar] [CrossRef]
  35. Zhou, X.; Gan, L.; Tian, W.; Zhang, Q.; Jin, S.; Li, H.; Bando, Y.; Golberg, D.; Zhai, T. Ultrathin SnSe2 Flakes Grown by Chemical Vapor Deposition for High-Performance Photodetectors. Adv. Mater. 2015, 27, 8035–8041. [Google Scholar] [CrossRef] [PubMed]
  36. Ogawa, S.; Shimatani, M.; Fukushima, S.; Okuda, S.; Kanai, Y.; Ono, T.; Matsumoto, K. Broadband photoresponse of graphene photodetector from visible to long-wavelength infrared wavelengths. Opt. Eng. 2019, 58, 057106. [Google Scholar] [CrossRef]
  37. Unsuree, N.; Selvi, H.; Crabb, M.G.; Alanis, J.A.; Parkinson, P.; Echtermeyer, T.J. Visible and infrared photocurrent enhancement in a graphene-silicon Schottky photodetector through surface-states and electric field engineering. 2D Mater. 2019, 6, 041004. [Google Scholar] [CrossRef] [Green Version]
  38. Sorifi, S.; Moun, M.; Kaushik, S.; Singh, R. High-Temperature Performance of a GaSe Nanosheet-Based Broadband Photodetector. ACS Appl. Electron. Mater. 2020, 2, 670–676. [Google Scholar] [CrossRef]
  39. Cao, Y.; Cai, K.; Hu, P.; Zhao, L.; Yan, T.; Luo, W.; Zhang, X.; Wu, X.; Wang, K.; Zheng, H. Strong enhancement of photoresponsivity with shrinking the electrodes spacing in few layer GaSe photodetectors. Sci. Rep. 2015, 5, 8130. [Google Scholar] [CrossRef] [Green Version]
  40. Liang, F.; Wang, C.; Luo, C.; Xia, Y.; Wang, Y.; Xu, M.; Wang, H.; Wang, T.; Zhu, Y.; Wu, P.; et al. Ferromagnetic CoSe broadband photodetector at room temperature. Nanotechnology 2020, 31, 374002. [Google Scholar] [CrossRef] [PubMed]
  41. Lee, J.W.; Kim, D.Y.; Baek, S.; Yu, H.; So, F. Inorganic UV–Visible–SWIR Broadband Photodetector Based on Monodisperse PbS Nanocrystals. Small 2016, 12, 1328–1333. [Google Scholar] [CrossRef]
  42. Chen, S.; Shehzad, K.; Qiao, X.; Luo, X.; Liu, X.; Zhang, Y.; Zhang, X.; Xu, Y.; Fan, X. A high performance broadband photodetector based on (SnxSb1−x)Se3 nanorods with enhanced electrical conductivity. J. Mater. Chem. C 2018, 6, 11078–11085. [Google Scholar] [CrossRef]
  43. Hasan, M.R.; Arinze, E.S.; Singh, A.K.; Oleshko, V.P.; Guo, S.; Rani, A.; Cheng, Y.; Kalish, I.; Zaghloul, M.E.; Rao, M.V.; et al. An Antimony Selenide Molecular Ink for Flexible Broadband Photodetectors. Adv. Electron. Mater. 2016, 2, 1600182. [Google Scholar] [CrossRef] [Green Version]
  44. Wang, Z.; Safdar, M.; Jiang, C.; He, J. High-Performance UV–Visible–NIR Broad Spectral Photodetectors Based on One-Dimensional In2Te3 Nanostructures. Nano Lett. 2012, 12, 4715–4721. [Google Scholar] [CrossRef]
  45. Yan, S.; Zhou, L.; Shi, Y.; Wang, B.; Wang, J.; Xu, X. Solvothermal Synthesis of Indium Telluride Nanowires and Its Photoelectrical Property. J. Nanosci. Nanotechnol. 2015, 15. [Google Scholar] [CrossRef] [PubMed]
  46. Jiang, T.; Zang, Y.; Sun, H.; Zheng, X.; Liu, Y.; Gong, Y.; Fang, L.; Cheng, X.; He, K. Broadband High-Responsivity Photodetectors Based on Large-Scale Topological Crystalline Insulator SnTe Ultrathin Film Grown by Molecular Beam Epitaxy. Adv. Opt. Mater. 2017, 5, 1600727. [Google Scholar] [CrossRef]
  47. Yang, J.; Yu, W.; Pan, Z.; Yu, Q.; Yin, Q.; Guo, L.; Zhao, Y.; Sun, T.; Bao, Q.; Zhang, K. Ultra-Broadband Flexible Photodetector Based on Topological Crystalline Insulator SnTe with High Responsivity. Small 2018, 14, 1802598. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Q.; Li, C.-Z.; Ge, S.; Li, J.-G.; Lu, W.; Lai, J.; Liu, X.; Ma, J.; Yu, D.-P.; Liao, Z.-M.; et al. Ultrafast Broadband Photodetectors Based on Three-Dimensional Dirac Semimetal Cd3As2. Nano Lett. 2017, 17, 834–841. [Google Scholar] [CrossRef] [Green Version]
  49. Niu, Y.; Wang, B.; Chen, J.; Wu, D. Ultra-broadband and highly responsive photodetectors based on a novel EuBiTe3 flake material at room temperature. J. Mater. Chem. C 2018, 6, 713–716. [Google Scholar] [CrossRef]
  50. Zhang, H.; Zhang, X.; Liu, C.; Lee, S.-T.; Jie, J. High-Responsivity, High-Detectivity, Ultrafast Topological Insulator Bi2Se3/Silicon Heterostructure Broadband Photodetectors. ACS Nano 2016, 10, 5113–5122. [Google Scholar] [CrossRef]
  51. Jian, H.-L. Investigation of Crystal Properties of Topological Insulator Bi(x)Sb(2−X)Te3 by Melting Growth Method. Master’s Thesis, National Sun Yat-Sen University, Kaohsiung, Taiwan, 4 September 2018. [Google Scholar]
  52. Shahil, K.M.F.; Hossain, M.Z.; Goyal, V.; Balandin, A.A. Micro-Raman Spectroscopy of Mechanically Exfoliated Few-Quintuple Layers of Bi2Te3, Bi2Se3, and Sb2Te3 Materials. J. Appl. Phys. 2012, 111, 054305. [Google Scholar] [CrossRef] [Green Version]
  53. Zybala, R.; Mars, K.; Mikuła, A.; Boguslawski, J.; Sobon, G.; Sotor, J.; Schmidt, M.; Kaszyca, K.; Chmielewski, M.; Ciupinski, L.; et al. Synthesis and Characterization of Antimony Telluride for Thermoelectric and Optoelectronic Applications. Arch. Metall. Mater. 2017, 62. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The X-ray diffraction (XRD) pattern of the produced Sb2Te3. The sharp peaks indicate that the Sb2Te3 is highly single crystalline. The Sb2Te3 powder was ground from the Sb2Te3 crystal. The XRD pattern of the Sb2Te3 powder is consistent with the JCPDS card.
Figure 1. The X-ray diffraction (XRD) pattern of the produced Sb2Te3. The sharp peaks indicate that the Sb2Te3 is highly single crystalline. The Sb2Te3 powder was ground from the Sb2Te3 crystal. The XRD pattern of the Sb2Te3 powder is consistent with the JCPDS card.
Sensors 21 01535 g001
Figure 2. The SEM picture of the Sb2Te3 flake for the photocurrent experiment. Two Pt contacts were deposited on the Sb2Te3 flake. The left-bottom inset shows an Atomic Force Microscope (AFM) picture of the Sb2Te3 flake and the AFM thickness profile of the flake. The thickness is about 260 nm.
Figure 2. The SEM picture of the Sb2Te3 flake for the photocurrent experiment. Two Pt contacts were deposited on the Sb2Te3 flake. The left-bottom inset shows an Atomic Force Microscope (AFM) picture of the Sb2Te3 flake and the AFM thickness profile of the flake. The thickness is about 260 nm.
Sensors 21 01535 g002
Figure 3. The left-top inset shows the linear current–voltage relationship, which indicates the ohmic contact between the Sb2Te3 flake and Pt electrodes. The right-bottom inset shows the measured photocurrent at different applied voltages. The main figure shows the extracted photocurrent as a function of applied voltage. The photocurrent varies linearly with the applied voltage.
Figure 3. The left-top inset shows the linear current–voltage relationship, which indicates the ohmic contact between the Sb2Te3 flake and Pt electrodes. The right-bottom inset shows the measured photocurrent at different applied voltages. The main figure shows the extracted photocurrent as a function of applied voltage. The photocurrent varies linearly with the applied voltage.
Sensors 21 01535 g003
Figure 4. The photocurrent as a function of the light power at three wavelengths. The left-top inset shows that the photocurrent varies linearly with the light power at three wavelengths. The right-top inset shows that the responsivity is insensitive to wavelengths from 300 to 1000 nm.
Figure 4. The photocurrent as a function of the light power at three wavelengths. The left-top inset shows that the photocurrent varies linearly with the light power at three wavelengths. The right-top inset shows that the responsivity is insensitive to wavelengths from 300 to 1000 nm.
Sensors 21 01535 g004
Figure 5. The left side shows the band structure of the Sb2Te3 bulk state. The right side shows the photocurrent as a function of light wavelength. The corresponding energies of these photocurrent peaks at different light wavelengths are consistent with the energy difference of band structure peaks between conduction and valence bands.
Figure 5. The left side shows the band structure of the Sb2Te3 bulk state. The right side shows the photocurrent as a function of light wavelength. The corresponding energies of these photocurrent peaks at different light wavelengths are consistent with the energy difference of band structure peaks between conduction and valence bands.
Sensors 21 01535 g005
Figure 6. The responsivity is proportional to the conductance of the flake. The top-left inset shows that the responsivity is insensitive to the flake thickness. The top-right inset shows that the detectivity is insensitive to the flake thickness.
Figure 6. The responsivity is proportional to the conductance of the flake. The top-left inset shows that the responsivity is insensitive to the flake thickness. The top-right inset shows that the detectivity is insensitive to the flake thickness.
Sensors 21 01535 g006
Table 1. List of the fitting parameters of the laser power-dependent photocurrent for three light wavelengths.
Table 1. List of the fitting parameters of the laser power-dependent photocurrent for three light wavelengths.
Laser Wavelength405 (nm)532 (nm)808 (nm)
α   A / W 383143
β 1.030.991.05
Table 2. List of the wavelength range and the deviation ratio of photocurrent for reported broadband light materials.
Table 2. List of the wavelength range and the deviation ratio of photocurrent for reported broadband light materials.
MaterialWavelength (nm)Deviation Ratio (%)Reference
Sb2Te3300∼100034This work
Sb2Te3/Si/Sb2Te3365∼94098.17[19]
Si350∼1050109.41[20]
InSe400∼800180[21]
InSe370∼980266[22]
In2Se3250∼700252[23]
In2Se3400∼900104[24]
AsP2500∼8000121[25]
BP300∼6802449[26]
BP405∼940156[27]
MoSe2400∼60087[28]
WSe2500∼800217[29]
WSe2390∼88070[30]
WS2450∼650235[31]
MoS2400∼690542[32]
MoS2455∼850178[15]
PtSe22500∼10,000373[33]
SnS2300∼700130[34]
SnSe2500∼700286[35]
Graphene600∼96007334[36]
Graphene1100∼16501443[37]
GaSe220∼650165[38]
GaSe400∼800368[39]
GaS245∼610278[10]
CoSe450∼4100143[40]
CoSe450∼950132[40]
PbS400∼2000186[41]
InGaAs400∼16,000397[18]
InGaAs400∼1000220[18]
Sb2Se3300∼1100267[42]
Sb2Se3300∼1100226[43]
In2Te3350∼1090129[44]
In2Te3350∼1000403[45]
SnTe405∼2003120[46]
SnTe405∼80891[46]
SnTe254∼650399[47]
Cd3As2532∼10,600926[48]
Cd3As2532∼940140[48]
EuBiTe3370∼1550112[49]
EuBiTe3370∼1064110[49]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, S.-M.; Hung, J.-L.; Chou, M.; Chen, C.-Y.; Liu, F.-C.; Chen, R.-S. The Highly Uniform Photoresponsivity from Visible to Near IR Light in Sb2Te3 Flakes. Sensors 2021, 21, 1535. https://doi.org/10.3390/s21041535

AMA Style

Huang S-M, Hung J-L, Chou M, Chen C-Y, Liu F-C, Chen R-S. The Highly Uniform Photoresponsivity from Visible to Near IR Light in Sb2Te3 Flakes. Sensors. 2021; 21(4):1535. https://doi.org/10.3390/s21041535

Chicago/Turabian Style

Huang, Shiu-Ming, Jai-Lung Hung, Mitch Chou, Chi-Yang Chen, Fang-Chen Liu, and Ruei-San Chen. 2021. "The Highly Uniform Photoresponsivity from Visible to Near IR Light in Sb2Te3 Flakes" Sensors 21, no. 4: 1535. https://doi.org/10.3390/s21041535

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop