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

High-responsivity, self-driven visible-near infrared Sb2Se3 nanorod array photodetector

Open Access Open Access

Abstract

Anisotropic antimony selenide (Sb2Se3) semiconductor has received considerable attention due to its unique one-dimensional crystal structure and corresponding superior and anisotropic optical and electronic properties. It is a promising material for a wide range of applications related to electronics and optoelectronics. Herein, we demonstrate a high-performance and self-powered Sb2Se3 nanorod array-based core/shell heterojunction detector fabricated on glass substate. The detector shows a wide spectral photoresponse range from visible to near-infrared (405-980 nm). The detector yields a detectivity of as high as 2.06×1012 Jones in the visible light (638 nm) and that of 1.82×1012 Jones (830 nm) at zero bias. Due to the strong built-in filed and excellent carrier transport, the detector exhibits ultrafast response speed at both rise (30 μs) and decay (68 μs) processes. Further analysis demonstrates that the noise is mainly generated from the 1/f noise in the low frequency range, while it is affected by the shot noise and generation-recombination noise in high frequency.

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

1. Introduction

Antimony selenide (Sb2Se3) has received considerable attention as potential light-absorbing material in the optoelectronic devices due to its superior optoelectronic properties such as proper optical bandgap(1.1-1.3eV), long carrier lifetime (∼ 70 ns) [1], large absorption coefficient (lager than 105 cm-1 at visible light) [28], and decent carrier mobility in proper direction (about 10 cm2V-1s-1) [4,911], as well as its environmental-friendly [12], low-cost and abundant constituents [13]. Moreover, Sb2Se3 is a simple binary compound with stable orthorhombic phase, which allows avoiding the complexities associated with phase-control during the fabrication of Sb2Se3-based semiconductor material and related devices.

The most attractive feature of Sb2Se3 is its one-dimensional crystal structure and the corresponding strong anisotropy in optical, electrical and defect properties. The crystalline Sb2Se3 consists of accumulated (Sb4Se6)n nanoribbons stacked through weak van der Waals along the [100] and [010] directions, whereas strong covalent bonds connecting antimony and selenium atoms in the [001] direction [2,14]. This results in strongly anisotropic carrier mobility depending on the transport direction. Moreover, theoretical calculations reveals that the surfaces parallel to the (001) direction, such as (110) and (120) direction, have lowest formation energies and are terminated by dangling-free surfaces, which results in benign grain boundary for the [hk1]-oriented Sb2Se3 grains [13,14]. It has been proved to effectively improve the carrier transport and suppress recombination for the [hk1] orientation preference Sb2Se3 devices [15,16,17].

Low dimensional Sb2Se3 nanowires and nano flakes have been synthesized and demonstrated impressive broadband photodetection performance [18,19,20]. For instance, the responsivity of Sb2Se3 flake-based photodetector reached 4.2 A/W with ms-scaled response times [21]. For Sb2Se3 thin film, its orientation could be controlled by the substrate surface states, fabrication techniques, etc. For example, the [hk0]-oriented Sb2Se3 films were deposited on mica substrate or quartz, where the chains were grown parallel to the substrate and allowed carriers transport within the covalently bonded ribbon. It is easier than that in other directions due to avoiding photogenerated carriers hopping against interchain barrier. The photodetectors based on Sb2Se3 films also show high responsivity (∼100 mA W-1), detectivity (∼1010 Jones) and tens of ms of response time [22]. However, these photoconductive type Sb2Se3 photodetectors don’t work without the external bias voltage. Sb2Se3 heterojunction photodetectors, especially devices based on the Sb2Se3 layer grown perpendicular to the substrate, could be driven by the build-in field and take advantage of the decent carrier transport properties along the Sb2Se3 chains. Such characteristics are beneficial to the separation and transport of photogenerated carriers for high performance photodetectors. Jiang et al. fabricated the Sb2Se3/ZnO heterojunction photodetector, which exhibited an optimal responsivity of 0.76 A W-1, a detectivity of 2.45×1011 Jones, and rise/decay time of 39/118 μs [23]. Chen et al. also reported Sb2Se3/CdS planar heterojunction detector exhibiting high performance with 0.42 A W-1 photoresponsivity, and 95 dB linear dynamic range [24].

In this work, we demonstrate strongly-anisotropic Sb2Se3 nanorod array (NRA) based photodetector in the configuration of Glass/Mo/MoSe2/Sb2Se3 NRAs/CdS/Window layers, where the Sb2Se3 NRA is the light-absorbing layer. The device exhibits excellent detection performance in the visible to near-infrared region (405-980 nm). A built-in field in the junction lead to ultrafast and sensitive photodetection without any external bias voltage. A detectivity (D*) of as high as 2.06×1012 Jones at 638 nm (or 1.82 ×1012 Jones at 830 nm) and a high responsivity of 46.44 A/W at 638 nm (or 41 A/W at 830 nm) under the bias of 0 V is achieved, which is the higher than most Sb2Se3-based detectors in previous reference. The noise for this device is also discussed in the last section.

2. Experimental section

The Sb2Se3 NRA photodetector were fabricated in a configuration of Glass/Mo/MoSe2/Sb2Se3 NRAs/CdS/Window layers. At first, Mo layer was deposited on the glass substrate by magnetron sputtering. Then, the as-deposited Mo-coated glass was transferred into a selenium-containing chamber for the formation of ∼10 nm thick MoSe2. The Sb2Se3 NRAs layer were deposited by a homemade close spaced sublimation (CSS) system with a growth process of our previous recipe [17]. An approximately 80-nm-thick CdS buffer layer was deposited by chemical bath deposition to form Sb2Se3/CdS heterojunction. Window layers of i-ZnO and ZnO:Al were deposited by RF magnetron sputtering to a thickness of 70 and 300 nm, respectively. Finally, gold electrodes were thermally evaporated using the resistance evaporation system.

The surface and cross-section morphologies were characterized by field-emission SEM (FEI Nova NANOSEM 450 field-emission microscope). The optical reflectance spectra were recorded using a Perkin-Elmer Lambda 950 spectrophotometer. The photoelectrial performance of Sb2Se3 NRA photodetector were measured by a self-built test platform including a Keithley 2400 source meter, optical chopper (C-995 optical chopper), semiconductor laser with different wavelengths (405, 532, 638, 780, 830 and 980 nm), and the results were recorded by an oscilloscope (Agilent Technologies infinivision DSO-X4022A).

3. Results and discussion

In this work, Sb2Se3 NRAs based photodiode detector was fabricated in a substrate configuration of Glass/Mo/MoSe2/Sb2Se3 absorber/CdS/TCO/Au. The Sb2Se3 NRAs have a thickness of 1.0∼1.2 μm, and the diameter of single Sb2Se3 nanorod is about 150-200 nm. Figure 1(a) shows the completed Sb2Se3 NRAs based photovoltaic detector, and Fig. 1(b) illustrates the corresponding device structure. For the completed detector, the original nanorod array patterns basically remain even after the deposition of following buffer layer and ZnO window layers. The Sb2Se3 NRAs detector appears black to naked eye, suggesting a very low surface reflectance in a widely spectral region. As shown in Fig. 1(c), the surface reflectivity is lower than 5% in a wide range. It hints that the Sb2Se3 NRA device has excellent light-trapping effect, which allows more photons to be absorbed and convert into electrical signal. Figure 1(d) displays the XRD patterns of the Sb2Se3 NRAs. The (211) and (221) diffraction peaks were observed stronger than others in the XRD pattern, hinting that the sample have a [hk1] orientation preference. Moreover, the XRD pattern also exhibits strong (101) and (002) peaks at 23.64 and 45.57 degree, respectively. It suggests that the growth of Sb2Se3 NRAs is along the c-axis ([001]) direction [9,25]. The texture coefficient (TC) of the diffraction peaks for our Sb2Se3 NRAs is calculated on the following equation [14]

$$\textrm{T}{\textrm{C}_{\textrm{hkl}}} = {\; }\frac{{{\textrm{I}_{({\textrm{hkl}} )}}}}{{{\textrm{I}_{0({\textrm{hkl}} )}}}}/( {\frac{1}{\textrm{N}}\mathop \sum \limits_{\textrm{i} = 1}^\textrm{N} \frac{{{\textrm{I}_{({{\textrm{h}_\textrm{i}}{\textrm{k}_\textrm{i}}{\textrm{I}_\textrm{i}}} )}}}}{{{\textrm{I}_{0({{\textrm{h}_\textrm{i}}{\textrm{k}_\textrm{i}}{\textrm{I}_\textrm{i}}} )}}}}} )$$
where I(hkl) is the diffraction peak intensity of (hkl) plane in the measured XRD pattern, while I0(hkl) is the diffraction peak intensity of (hkl) plane in the standard XRD pattern (JCPDS 15-0861). The texture coefficients (TCs) were also calculated and displayed in Fig. 1(e). As shown in the supplemental document (Table S1 and Figure S1), the TC value of (101), (002) is about 2.0, higher than that of the Sb2Se3 thin films. Besides, the peaks of (120), (240), (340) peaks could be detected, suggesting that the (Sb4Se6)n ribbons in the thin films were grown titled to the substrate [15,25,26]. Moreover, the Sb2Se3 also shows anisotropy in carrier transport. It is difficult for the carriers’ transport in [010] or [100] direction, because they have to hop between the ribbons held together by van der Waals forces. Whereas, the carrier transport along the covalent bond direction ([001]) is easier due to the overlapping electron clouds.

 figure: Fig. 1.

Fig. 1. (a) Corss-sectional SEM image of the fabricated Sb2Se3 nanorod array (NRA)-based photodetector. (b) The corresponding schematic diagram of the Sb2Se3 NRA photodetector. (c) Reflectance spectra of the Sb2Se3 NRA photodetector. (d) The X-ray diffraction (XRD) pattern of Sb2Se3 NRAs. (e) The calculated texture coefficients (TCs) for the Sb2Se3 NRAs.

Download Full Size | PDF

Figures 2(a) and 2(b) depicts the current-voltage (I-V) characteristics of the Sb2Se3 NRA photodiode detector in the dark and under light illumination (10 mW cm-2), respectively. The dark current at zero point is only 8.05 nA, whereas the dark current at reverse bias direction (-0.5 V) and at forward bias direction (+0.5 V) are 3.0 μA and 0.26 mA, respectively. The device demonstrates a rectification ratio of 87 at ${\mp} $0.5V under dark condition. This means the low leakage current and high diode quality in the Sb2Se3 NRA photodiode detector. Under the weak white light illumination, the photocurrent nearly keeps almost a constant value when the reverse bias decrease from -0.6 V to 0.3 V, suggesting the photo-generated carriers, driven by the built-in electrical field, could be collected without any bias. It demonstrates that the Sb2Se3 NRA device could be acted as a self-powered photodetector. As shown in Figure S2, the heterojunction displays a built-in field of 428 mV.

 figure: Fig. 2.

Fig. 2. (a, b) I-V characteristics of the photodetectors measured in the dark (a) and under white light illumination (b) (10 mW cm-2), respectively. (c) Schematic illustration of the energy-band diagram of the Sb2Se3 NRA photodetector. (d) Photoresponse of the photodetector irradiated under different wavelength from 405 to 980 nm. The laser was turn on/off regularly to measure the time-dependent response of the photodetectors. No external bias voltage was applied.

Download Full Size | PDF

To understand the transport of photogenerated carriers, Fig. 2(c) shows the energy band diagram of the Sb2Se3 NRA detector. The n-type CdS layer form core/shell heterojunction with p-type Sb2Se3 NRA absorber, allowing the efficient separation of photogenerated electron-hole pairs at the junction interface. The holes and electrons drift and are collected by the Mo back contact and front Au contact through ZnO window layers. Figure 2(d) depicts the photoresponse of the Sb2Se3 NRA photodetector from visible (405 nm) light to near-infrared (980) light with the intensity of 636 mW/cm2 without any voltage bias, where the continuous laser was switched on and off by a chopper. The device presents photoresponse in a wide spectral range from visible (405 nm) to near-infrared (980 nm) region, due to the narrow bandgap (1.24 eV) of the Sb2Se3 absorber. Moreover, the photoresponse shows a wavelength dependent behavior, where the red (638nm, 780nm) light could excite higher current under the same light intensity.

In addition, the photodetector also exhibits highly power intensity-dependent photoresponse for the incident laser, as shown in Fig. 3(a), where the photocurrent increases almost linearly with the light intensity ranging from 12 to 2547 mW/cm2 under 638 nm light. The continuous temporal response cycles of the detector were further measured under 638 nm light illumination with a light intensity of 636 mW/cm2 at zero bias, by repeatedly switching the incident laser on/off at a frequency of 5 kHz, as shown in Fig. 3(b). One can see that the current on/off switching behavior retains reversible conversion between high conduction and low conduction status with the laser on and off, showing highly cycling stability.

 figure: Fig. 3.

Fig. 3. (a) Photoresponse of the Sb2Se3 NRA photodetector under 638 nm laser irradiation with power from 12 to 2547 mW/cm2. (b) time-dependent photoresponse of the photodetector under 638 nm laser with the intensity of 636 mW/cm2 at 0 V bias. (c) and (d) the detailed rise and decay time curve of the detector.

Download Full Size | PDF

To quantitatively evaluate the response rate of the detector, the detailed rise and decay time curves are measured and displayed in Fig. 3(c) and 3(d), which can be defined as the time required to reach 90% and drop to 10% of the maximum photocurrent, respectively. The rise and decay times were determined to be 30 μs and 68 μs, respectively, at a light frequency of 5 KHz. This indicates that the ultrafast charge carrier generation and transport. It is noteworthy that this response rate is much quickly than the most reported on Sb2Se3-based detectors, including both photoconductor [22,2737] or photodiode detectors [2423]. The faster response speed of Sb2Se3 NRA photodetector is attributed to the unique device configuration: (i) The [001]-oriented mono-crystalline Sb2Se3 nanorod ensures the fast transport of photogenerated carrier along the c-axis direction due to its strongly anisotropic crystal structure and corresponding excellent electrical properties [17,13]. (ii) Strong built-in electric field of heterojunction between Sb2Se3 NR absorber and CdS buffer.

The power-dependent photodetector performance was further characterized by calculating the photoresponsivity ($\textrm{R}$) and detectivity (${\textrm{D}^{\ast }}$), respectively. The responsivity $\textrm{R}$ indicates how efficient the detector responds to the optical signal. It can be evaluated by the Eq. (2)

$$\textrm{R} = \frac{{{\textrm{I}_{\textrm{ph}}}}}{{\textrm{PA}}}$$
Where ${\textrm{I}_{\textrm{ph}}} = {\textrm{I}_{\textrm{light}}} - {\textrm{I}_{\textrm{dark}}}$, ${\textrm{I}_{\textrm{light}}}$ is the photocurrent, ${\textrm{I}_{\textrm{dark}}}$ is the dark current, A is the active area of the photodetector (A=0.00785 cm2), and P is the given light intensity (mW/cm2) of the incident light, respectively. The detectivity (${\textrm{D}^{\ast }}$), indicating the ability of a detector to detect weak optical signals, was calculated by the following Eq. (3)
$${\textrm{D}^{\ast }} = \frac{{{\textrm{A}^{1/2}}\textrm{R}}}{{{{({2\textrm{q}{\textrm{I}_\textrm{d}}} )}^{1/2}}}}$$
Figure 4(a), b plot the $\textrm{R}$ and ${\textrm{D}^{\ast }}{\; }$ at different light wavelengths as a function of light intensity. Clearly, both R and ${\textrm{D}^{\ast }}{\; }$ decreases with the light intensity for all the light wavelengths according to the varying tendencies in Fig. 4(a), and 4b. Figure 4(c) plots the wavelength-dependent $\textrm{R}$ and ${\textrm{D}^{\ast }}$ of the device. As shown, the maximum value of R in the visible region and NIR region was estimated to be 46.44 A/W (638 nm) and 41 A/W (830 nm), respectively, at a light intensity of 12 mW/cm2. Moreover, the maximum ${\textrm{D}^{\ast }}$ was about 2.06×1012 Jones (638 nm) and 1.82×1012 Jones (830 nm), respectively. This suggests that the Sb2Se3 NRA photodetector is extremely sensitive to small optical input signals. The performance of our Sb2Se3 NRA photodetector is compared with other reported Sb2Se3-based detectors in Table 1. The photoresponse speed of Sb2Se3 NRA photodetector is much faster than single Sb2Se3 nanwire and Sb2Se3 nanorods [20,28,34]. It could be ascribed to the fabrication technique, the difference in carrier transport distance, the built-in field, and so on. We also compared the detection performance between the Sb2Se3 thin film-based and Sb2Se3 NRA-based photodetectors both in the visible and NIR region. As shown in Figure S3, the responsibility of the detector based on Sb2Se3 NARs at 638 nm and 830 nm are 23.85 and 33.26 A/W, respectively. The detectivity at 638 and 830 nm are 1.06×1012 and 1.48×1012 Jones, respectively. As shown in Figure S4, the detector based on Sb2Se3 thin films exhibits slower rise and decay speed than the device based on Sb2Se3 NRAs under the same test conductions. The high detection performance for the Sb2Se3 NRA based photodetector could be ascribed two points. At first, the anisotropic crystal structure of Sb2Se3 NRAs and the corresponding benign grain boundary for each Sb2Se3 nanorod. The carrier transport along the covalent bond direction ([001]) is easier due to the overlapping electron clouds. Whereas, it is difficult for the carriers’ transport in [010] or [100] direction, because it is required to hop between the ribbons held together by van der Waals forces. Secondly, the grain boundaries of the vertical Sb2Se3 rod are terminated with the intrinsically benign (hk0) surfaces. This suppresses the carrier recombination at the interface of the CdS/Sb2Se3 core/shell junction, resulting in a high detectivity and great response times.

 figure: Fig. 4.

Fig. 4. (a) Responsivity of the device as a function of light intensity irradiated under different wavelength from 405 to 980 nm. (b) Detectivity of the device as a function of light intensity irradiated under different wavelength from 405 to 980 nm. (c) Wavelength-dependent responsivity and detectivity of the device. (d) Measured dark current noise at various frequencies of photodetectors with 0 V bias. The measured instrument noise, calculated 1/f noise, shot noise, and generation-recombination noise.

Download Full Size | PDF

Tables Icon

Table 1. Comparison of optoelectronic performance of Sb2Se3-based photodetectors.

Finally, we evaluate the noise spectrum for the Sb2Se3 NRA photodetector. The noise is associated with the leakage (shunt) current, temperature, and other elements of static and dynamic disorder induced noises, which usually cannot be derived simply from the dark current of the detector [24,42,43]. As shown in Fig. 4(d), the measured total noise power density is frequency-dependent and distributed in the range of 10−12∼10−13 $\textrm{A H}{\textrm{z}^{ - 1/2}}$ at 0 V bias. The measured noise was mainly generated from the $1/\textrm{f}$ noise in the low frequency range (4-100 Hz), which could be ascribed to the substantial carrier scattering at the grain boundary in the Sb2Se3 absorber [24]. In the frequency range of 100-5000 Hz, the total noise is also affected the shot noise (∼3.5×10−14 $\textrm{A H}{\textrm{z}^{ - 1/2}}$) and generation-recombination (G-R) noise (2-7×10−14 $\textrm{A H}{\textrm{z}^{ - 1/2}}$).

4. Conclusion

In conclusion, we have demonstrated a Sb2Se3 NRA-based photodetector in the configuration of Glass/Mo/MoSe2/Sb2Se3 NRA/CdS/TCO. The Sb2Se3 NRA detector exhibits a high responsivity of ∼45 A/W under the bias of 0 V, and the detectivity is above 1012 Jones in the visible and near-infrared wavelength range. Moreover, the rise time and fall time were determined to be 30 μs and 68 μs at 638 nm, which is faster than most of the Sb2Se3-based photodetector.

Funding

Natural Science Foundation for Distinguished Young Scholars of Hebei Province of China (F2019201289); National Natural Science Foundation of China (61804040); Education Commission of Hebei Province (ZD2019037).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. K. Wang, C. Chen, H. Y. Liao, S. Y. Wang, J. Tang, M. C. Beard, and Y. Yang, “Both free and trapped carriers contribute to photocurrent of Sb2Se3 solar cells,” J. Phys. Chem. Lett. 10(17), 4881–4887 (2019). [CrossRef]  

2. Y. Zhou, M. Y. Leng, Z. Xia, J. Zhong, H. B. Song, X. S. Liu, B. Yang, J. P. Zhang, J. Chen, K. H. Zhou, J. B. Han, Y. B. Cheng, and J. Tang, “Solution-processed antimony selenide heterojunction solar cells,” Adv. Energy Mater. 4(8), 1301846 (2014). [CrossRef]  

3. G. X. Liang, Z. H. Zheng, P. Fan, J. T. Luo, J. G. Hu, X. H. Zhang, H. L. Ma, B. Fan, Z. K. Luo, and D. P. Zhang, “Thermally induced structural evolution and performance of Sb2Se3 films and nanorods prepared by an easy sputtering method,” Sol. Energy Mater. Sol. Cells 174, 263–270 (2018). [CrossRef]  

4. C. C. Yuan, L. J. Zhang, W. F. Liu, and C. F. Zhu, “Rapid thermal process to fabricate Sb2Se3 thin film for solar cell application,” Sol. Energy 137, 256–260 (2016). [CrossRef]  

5. M. Birkett, W. M. Linhart, J. Stoner, L. J. Phillips, K. Durose, J. Alaria, J. D. Major, R. Kudrawiec, and T. D. Veal, “Band gap temperature-dependence of close-space sublimation grown Sb2Se3 by photo-reflectance,” APL Mater. 6(8), 084901 (2018). [CrossRef]  

6. X. M. Wang, R. F. Tang, C. Y. Wu, C. F. Zhu, and T. Chen, “Development of antimony sulfide-selenide Sb-2(S, Se)(3)-based solar cells,” J. Energy Chem. 27(3), 713–721 (2018). [CrossRef]  

7. K. Shen, C. Z. Ou, T. L. Huang, H. B. Zhu, J. J. Li, Z. Q. Li, and Y. Mai, “Mechanisms and modification of nonlinear shunt leakage in Sb2Se3 thin film solar cells,” Sol. Energy Mater. Sol. Cells 186, 58–65 (2018). [CrossRef]  

8. C. Chen, W. Q. Li, Y. Zhou, C. Chen, M. Luo, X. S. Liu, K. Zeng, B. Yang, C. W. Zhang, J. B. Han, and J. Tang, “Optical properties of amorphous and polycrystalline Sb2Se3 thin films prepared by thermal evaporation,” Appl. Phys. Lett. 107(4), 043905 (2015). [CrossRef]  

9. H. W. Lei, J. J. Chen, Z. J. Tan, and G. J. Fang, “Review of recent progress in antimony chalcogenide-based solar cells: Materials and Devices,” Sol. RRL 3(6), 1900026 (2019). [CrossRef]  

10. A. Shongalova, M. R. Correia, J. P. Teixeira, J. P. Leitao, J. C. Gonzalez, S. Ranjbar, S. Garud, B. Vermang, J. M. V. Cunha, P. M. P. Salome, and P. A. Fernandes, “Growth of Sb2Se3 thin films by selenization of RF sputtered binary precursors,” Sol. Energy Mater. Sol. Cells 187, 219–226 (2018). [CrossRef]  

11. S. L. Benjamin, C. H. de Groot, A. L. Hector, R. Huang, E. Koukharenko, W. Levason, and G. Reid, “Chemical vapour deposition of antimony chalcogenides with positional and orientational control: precursor design and substrate selectivity,” J. Mater. Chem. C 3(2), 423–430 (2015). [CrossRef]  

12. L. Wang, D. B. Li, K. H. Li, C. Chen, H. X. Deng, L. Gao, Y. Zhao, F. Jiang, L. Y. Li, F. Huang, Y. S. He, H. S. Song, G. D. Niu, and J. Tang, “Stable 6%-efficient Sb2Se3 solar cells with a ZnO buffer layer,” Nat. Energy 2(4), 17046 (2017).

13. C. Chen, D. C. Bobela, Y. Yang, S. C. Lu, K. Zeng, C. Ge, B. Yang, L. Gao, Y. Zhao, M. C. Beard, and J. Tang, “Characterization of basic physical properties of Sb2Se3 and its relevance for photovoltaics,” Front. Optoelectron. 10(1), 18–30 (2017). [CrossRef]  

14. Y. Zhou, L. Wang, S. Y. Chen, S. K. Qin, X. S. Liu, J. Chen, D. J. Xue, M. Luo, Y. Z. Cao, Y. B. Cheng, E. H. Sargent, and J. Tang, “Thin-film Sb2Se3 photovoltaics with oriented one-dimensional ribbons and benign grain boundaries,” Nat. Photonics 9(6), 409–415 (2015). [CrossRef]  

15. K. Li, C. Chen, S. Lu, C. Wang, S. Wang, Y. Lu, and J. Tang, “Orientation Engineering in low-dimensional crystal-structural materials via seed screening,” Adv. Mater. 31(44), 1903914 (2019). [CrossRef]  

16. H. Deng, Y. Zeng, M. Ishaq, S. Yuan, H. Zhang, X. Yang, M. Hou, U. Farooq, J. Huang, K. Sun, R. Webster, H. Wu, Z. Chen, F. Yi, H. Song, X. Hao, and J. Tang, “Quasiepitaxy strategy for efficient full-inorganic Sb2S3 solar cells,” Adv. Funct. Mater. 29(31), 1901720 (2019). [CrossRef]  

17. Z. Li, X. Liang, G. Li, H. Liu, H. Zhang, J. Guo, J. Chen, K. Shen, X. San, W. Yu, R. E. I. Schropp, and Y. Mai, “9.2%-efficient core-shell structured antimony selenide nanorod array solar cells,” Nat. Commun. 10(1), 125 (2019). [CrossRef]  

18. M. R. Hasan, E. S. Arinze, A. K. Singh, V. P. Oleshko, S. Guo, A. Rani, Y. Cheng, I. Kalish, M. E. Zaghloul, and M. V. Rao, “An antimony selenide molecular ink for flexible broadband photodetectors,” Adv. Electron. Mater. 2(9), 1600182 (2016). [CrossRef]  

19. Y. Liang, Y. Wang, J. Wang, S. Wu, D. Jiang, and J. Lian, “High-performance flexible photodetectors based on single-crystalline Sb2Se3 nanowires,” RSC Adv. 6(14), 11501–11506 (2016). [CrossRef]  

20. Z. Ma, S. Chai, Q. Feng, L. Li, X. Li, L. Huang, D. Liu, J. Sun, R. Jiang, T. Zhai, and H. Xu, “Chemical vapor deposition growth of high crystallinity Sb2Se3 nanowire with strong anisotropy for near-infrared photodetectors,” Small 15(9), 1805307 (2019). [CrossRef]  

21. M. Zhao, J. Su, Y. Zhao, P. Luo, F. Wang, W. Han, Y. Li, X. Zu, L. Qiao, and T. Zhai, “Sodium-mediated epitaxial growth of 2D ultrathin Sb2Se3 flakes for broadband photodetection,” Adv. Funct. Mater. 30(13), 1909849 (2020). [CrossRef]  

22. X. X. Wen, Z. H. Lu, L. Valdman, G. C. Wang, M. Washington, and T. M. Lu, “High-crystallinity epitaxial Sb2Se3 thin films on mica for flexible near-infrared photodetectors,” ACS Appl. Mater. Interfaces 12(31), 35222–35231 (2020). [CrossRef]  

23. J. Jiang, Y. Guo, X. Weng, F. Long, Y. Xin, Y. Lu, Z. Ye, S. Ruan, and Y.-J. Zeng, “A tailorable polarity-flipping response in self-powered, flexible Sb2Se3/ZnO bilayer photodetectors,” J. Mater. Chem. C 9(14), 4978–4988 (2021). [CrossRef]  

24. C. Chen, K. Li, F. Li, B. Wu, P. Jiang, H. Wu, S. Lu, G. Tu, Z. Liu, and J. Tang, “One-dimensional Sb2Se3 enabling a highly flexible photodiode for light-source-free heart rate detection,” ACS Photonics 7(2), 352–360 (2020). [CrossRef]  

25. S. N. Park, S. Y. Kim, S. J. Lee, S. J. Sung, K. J. Yang, J. K. Kang, and D. H. Kim, “Controlled synthesis of (hk1) preferentially oriented Sb2Se3 rod arrays by co-evaporation for photovoltaic applications,” J. Mater. Chem. A 7(45), 25900–25907 (2019). [CrossRef]  

26. P. Fan, G.-J. Chen, S. Chen, Z.-H. Zheng, M. Azam, N. Ahmad, Z.-H. Su, G.-X. Liang, X.-H. Zhang, and Z.-G. Chen, “Quasi-vertically oriented Sb2Se3 thin-film solar cells with open-circuit voltage exceeding 500 mV prepared via close-space sublimation and selenization,” ACS Appl. Mater. Interfaces 13(39), 46671–46680 (2021). [CrossRef]  

27. W. Wu, Y. Li, L. Liang, Q. Hao, J. Zhang, H. Liu, and C. Liu, “Enhanced broadband responsivity of Ni-Doped Sb2Se3 nanorod photodetector,” J. Phy. Chem. C 123(23), 14781–14789 (2019). [CrossRef]  

28. G. H. Chen, J. Zhou, J. Zuo, and Q. Yang, “Organometallically anisotropic growth of ultralong Sb2Se3 nanowires with highly enhanced photothermal response,” ACS Appl. Mater. Interfaces 8(4), 2819–2825 (2016). [CrossRef]  

29. T. Y. Zhai, M. F. Ye, L. Li, X. S. Fang, M. Y. Liao, Y. F. Li, Y. Koide, Y. Bando, and D. Golberg, “Single-crystalline Sb2Se3 nanowires for high-performance field emitters and photodetectors,” Adv. Mater. 22(40), 4530–4533 (2010). [CrossRef]  

30. Y. Xin, J. C. Jiang, Y. F. Lu, H. W. Liang, Y. J. Zeng, and Z. Z. Ye, “Self-powered broad spectral photodetector with ultrahigh responsivity and fast response based on Sb2Se3/VO2 heterojunction,” Adv. Mater. Interfaces 8(10), 2100058 (2021). [CrossRef]  

31. G. H. Chen, W. L. Wang, C. D. Wang, T. Ding, and Q. Yang, “Controlled synthesis of ultrathin Sb2Se3 nanowires and application for flexible photodetectors,” Adv. Sci. 2(10), 1500109 (2015). [CrossRef]  

32. Y. Q. Liu, M. Zhang, F. X. Wang, and G. B. Pan, “Facile microwave-assisted synthesis of uniform Sb2Se3 nanowires for high performance photodetectors,” J. Mater. Chem. C 2(2), 240–244 (2014). [CrossRef]  

33. S. Chen, X. S. Qiao, F. X. Wang, Q. Luo, X. H. Zhang, X. Wan, Y. Xu, and X. P. Fan, “Facile synthesis of hybrid nanorods with the Sb2Se3/AgSbSe2 heterojunction structure for high performance photodetectors,” Nanoscale 8(4), 2277–2283 (2016). [CrossRef]  

34. W. Wu, M. J. Wang, L. M. Liang, Y. F. Yan, X. R. Zheng, Q. Y. Hao, H. Liu, J. Zhang, and C. C. Liu, “Sb(2)Se(3)nanorods with length dependent photoresponses,” Journal of Alloys and Compounds 726, 254–260 (2017). [CrossRef]  

35. M. J. Wang, W. Wu, X. R. Zheng, X. Hou, C. C. Liu, Q. Y. Hao, and H. Liu, “Preparation of uniform Sb2Se3 nanorods by hot-injection polyol process and their application in photodetector,” Mater. Lett. 193, 191–194 (2017). [CrossRef]  

36. B. W. Zhao, K. K. Li, Y. X. Ren, T. J. Dai, and X. Z. Liu, “A fluorine based plasma surface modification process opens up a new avenue to improve the film conductivity and optoelectronic properties of Sb2Se3 broadband photodetector,” Sol. Energy 195, 454–460 (2020). [CrossRef]  

37. D. Choi, Y. Jang, J. Lee, G. H. Jeong, D. Whang, S. W. Hwang, K. S. Cho, and S. W. Kim, “Diameter-controlled and surface-modified Sb2Se3 nanowires and their photodetector performance,” Sci Rep 4, 7 (2014).

38. G. Li, Z. Q. Li, J. W. Chen, X. Chen, S. Qiao, S. F. Wang, Y. Xu, and Y. H. Mai, “Self-powered, high-speed Sb2Se3/Si heterojunction photodetector with close spaced sublimation processed Sb2Se3 layer,” Journal of Alloys and Compounds 737, 67–73 (2018). [CrossRef]  

39. J. L. Wang, S. P. Li, T. T. Wang, F. Guan, L. J. Zhao, L. H. Li, J. H. Zhang, and G. J. Qiao, “Solution-processed Sb2Se3 on TiO2 thin films toward oxidation-and moisture-resistant, self-powered photodetectors,” ACS Appl. Mater. Interfaces 12(34), 38341–38349 (2020). [CrossRef]  

40. L. Zhiqiang, Z. Hongbing, G. Yuting, N. Xiaona, C. Xu, Z. Chong, Z. Wen, L. Xiaoyang, Z. Dong, C. Jingwei, and M. Yaohua, “Efficiency enhancement of Sb2Se3 thin-film solar cells by the co-evaporation of Se and Sb2Se3,” Appl. Phy. Express 9(5), 052302 (2016). [CrossRef]  

41. H. B. Song, T. Y. Li, J. Zhang, Y. Zhou, J. J. Luo, C. Chen, B. Yang, C. Ge, Y. Q. Wu, and J. Tang, “Highly anisotropic Sb2Se3 nanosheets: gentle exfoliation from the bulk precursors possessing 1D crystal structure,” Adv. Mater. 29(29), 1700441 (2017). [CrossRef]  

42. Y. Fang, A. Armin, P. Meredith, and J. Huang, “Accurate characterization of next-generation thin-film photodetectors,” Nat. Photonics 13(1), 1–4 (2019). [CrossRef]  

43. Y. Liu, H. Ye, Y. Zhang, K. Zhao, Z. Yang, Y. Yuan, H. Wu, G. Zhao, Z. Yang, J. Tang, Z. Xu, and S. Liu, “Surface-tension-controlled crystallization for high-quality 2D perovskite single crystals for ultrahigh photodetection,” Matter 1(2), 465–480 (2019). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Document

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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) Corss-sectional SEM image of the fabricated Sb2Se3 nanorod array (NRA)-based photodetector. (b) The corresponding schematic diagram of the Sb2Se3 NRA photodetector. (c) Reflectance spectra of the Sb2Se3 NRA photodetector. (d) The X-ray diffraction (XRD) pattern of Sb2Se3 NRAs. (e) The calculated texture coefficients (TCs) for the Sb2Se3 NRAs.
Fig. 2.
Fig. 2. (a, b) I-V characteristics of the photodetectors measured in the dark (a) and under white light illumination (b) (10 mW cm-2), respectively. (c) Schematic illustration of the energy-band diagram of the Sb2Se3 NRA photodetector. (d) Photoresponse of the photodetector irradiated under different wavelength from 405 to 980 nm. The laser was turn on/off regularly to measure the time-dependent response of the photodetectors. No external bias voltage was applied.
Fig. 3.
Fig. 3. (a) Photoresponse of the Sb2Se3 NRA photodetector under 638 nm laser irradiation with power from 12 to 2547 mW/cm2. (b) time-dependent photoresponse of the photodetector under 638 nm laser with the intensity of 636 mW/cm2 at 0 V bias. (c) and (d) the detailed rise and decay time curve of the detector.
Fig. 4.
Fig. 4. (a) Responsivity of the device as a function of light intensity irradiated under different wavelength from 405 to 980 nm. (b) Detectivity of the device as a function of light intensity irradiated under different wavelength from 405 to 980 nm. (c) Wavelength-dependent responsivity and detectivity of the device. (d) Measured dark current noise at various frequencies of photodetectors with 0 V bias. The measured instrument noise, calculated 1/f noise, shot noise, and generation-recombination noise.

Tables (1)

Tables Icon

Table 1. Comparison of optoelectronic performance of Sb2Se3-based photodetectors.

Equations (3)

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

T C hkl = I ( hkl ) I 0 ( hkl ) / ( 1 N i = 1 N I ( h i k i I i ) I 0 ( h i k i I i ) )
R = I ph PA
D = A 1 / 2 R ( 2 q I d ) 1 / 2
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.