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

Self-powered, ultraviolet-to-near infrared broadband photodetector based on Ag-doped CsPbI3/PEDOT:PSS heterojunction

Open Access Open Access

Abstract

Metal halide perovskites are studied for photodetection applications because of their outstanding optical and electrical properties. A self-powered ultraviolet-to-near infrared broadband photodetector based on a Ag-doped CsPbI3/PEDOT:PSS heterojunction was investigated. The photodetector using a CsPbI3:Ag/PEDOT:PSS heterostructure with a planar photoconductive structure operated over a broad 355–1560 nm wavelength range in self-powered mode. A terahertz signal was modulated with the CsPbI3:Ag/PEDOT:PSS structure at low optical excitation intensity to investigate its photodetection mechanism. The experimentally designed detector can present images of the letters “C”, “N” and “U” in the visible and near-infrared wavelengths, indicating a potential broadband imaging application.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent years, inorganic perovskites have been developed with a wide variety of morphologies, ranging from zero-dimensional quantum dots to single crystals and three-dimensional textured morphologies [1]. Doping of inorganic perovskites can be performed to give these materials additional functions or improved performance. The different morphologies of these perovskites also have a variety of photoelectric properties [2,3,4]. Metal electrodes can be combined with perovskite nanocrystals to form a basic detector structure [5]. When perovskites are combined with other materials, they provide new ways to manufacture photodetectors based on the use of heterojunctions [6,7], with particular interest being generated by use of the poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) hole transport layer [8,9]. Photodetectors that can operate without an external power source, which are known as self-driving detectors, are also becoming well known [1014]. Typical self-driving detectors use driving mechanisms based on triboelectric, photoelectric [15], and transistor structures [16]. The inorganic hybrid perovskite CsPbX3 (where X = Cl, Br or I) is widely used in these photodetectors. Although CsPbI3 is limited by its narrow band gap of 1.73 eV [17], it still offers several possibilities for development of photodetectors [18], including 0D photodetection development based on devices with quantum dot morphologies [19,20], 1D photodetectors with nanowire morphologies [21,22], and 3D photodetectors based on nanocrystalline films and single crystal structures [23,24].

Most of these photodetectors are usually used for detection in the ultraviolet range, especially for high-energy rays such as X-rays [25]. Additionally, these detectors have more complex structures than conventional detectors [26,27]. Although their parallel structures are more restrictive in terms of carrier mobility than the conventional vertical structure, a hole transport layer composed of PEDOT:PSS can be added to improve the mobility [8,2831]. To enhance the carrier mobility further, dimethyl sulfoxide (DMSO) can be used as a dopant in PEDOT:PSS [32,33,34]. However, reports of self-driven broadband photodetectors with simple structures remain rare in related work.

In this work, a self-powered ultraviolet-to-near infrared broadband photodetector based on a Ag-doped CsPbI3/PEDOT:PSS heterojunction was investigated. The photodetector using a CsPbI3:Ag/PEDOT:PSS heterostructure with a planar photoconductive structure operated over a broad 355–1560 nm wavelength range in self-powered mode. A terahertz signal was modulated with the CsPbI3:Ag/PEDOT:PSS structure at low optical excitation intensity to investigate its photodetection mechanism. [3539] The experimentally designed detector would present images in the visible and near-infrared wavelengths, indicating a potential broadband imaging application.

2. Experimental details

Figure 1(a) shows the CsPbI3:Ag/PEDOT:PSS/quartz structure used for the self-powered ultraviolet-to-near infrared broadband detector. A PEDOT:PSS-doped DMSO solution [34,35] with a 9:1 ratio was spin-coated onto the substrate at 1000 rpm for 9 s, and at 2000rpm for 30 s. Two parallel silver electrodes with a spacing of 1.5 mm were then formed on the PEDOT:PSS layer by vacuum evaporation. The CsPbI3 precursor solution was dispersed into 1 mL of a mixture of DMSO and dimethylformamide (DMF), using 326 mg of cesium iodide (CsI) and 578 mg of lead iodide (PbI2), and the solution was stirred for 30 min at room temperature. Next, silver nanoparticles were dispersed into the CsPbI3 solution at a ratio of 200 mg of silver nanoparticles per 1 ml of solution, and the solution was then stirred for another 10 min at room temperature. The mixed solution was prepared on the PEDOT:PSS:DMSO/quartz surface under spin-coating conditions of 3000 rpm for 15 s and 5000 rpm for 30 s, with an appropriate amount of chlorobenzene being added 5 s into the first 15 s period. The sample was then annealed at 100°C for 15 min. The absorption spectrum of the samples, including those of the Ag-doped/undoped perovskite layers and PEDOT:PSS layer is shown in Fig. 1(b). The transmission spectra of the surfaces and the cross-sections of the CsPbI3:Ag/PEDOT:PSS and CsPbI3/PEDOT:PSS structures were recorded by scanning electron microscopy (SEM), as shown in Fig. 1(c) and (d). The Ag-doped nanoparticles will produce silver iodide in the perovskite, which changes the degree of perovskite crystallinity. Due to the excessive amount of silver nanoparticles, there are still a large number of particles clustered on the surface, which makes the surface morphology rough, which is significantly different from the surface of the undoped sample. The reaction effects of the two samples described above were then measured using an X-ray diffractometer with an accelerating voltage of 10 kV, with results as shown in Fig. 1(e) to (h) [38,40]. The structure was then tested under light irradiation from ultraviolet-to-near infrared. The light and dark currents obtained during the experiment were all controlled via computer control source meter (Keithley 2450) measurements. All measurements were acquired in the air at room temperature.

 figure: Fig. 1.

Fig. 1. (a) Example of the CsPbI3:Ag/PEDOT:PSS structure and experimental setup for the electrical measurements. (b) Absorption spectra of PEDOT:PSS, the CsPbI3:Ag/PEDOT:PSS structure, and the CsPbI3:Ag/PEDOT:PSS structure. SEM images of (c) CsPbI3:Ag/PEDOT:PSS structure and (d) CsPbI3/PEDOT:PSS structure. X-ray diffraction results for (e) CsPbI3:Ag/PEDOT:PSS structure (f) CsPbI3/PEDOT:PSS structure (g) CsPbI3:Ag structure (h) CsPbI3 structure.

Download Full Size | PDF

3. Results and discussion

To investigate the photoelectric properties of the samples, the photo-to-dark current ratio is calculated throughout the testing based on the formulae below to determine the responsivity (R), the external quantum efficiency (EQE), and the detection rate or detectivity (D*), which are all important indicators for photoelectric performance characterization of the photodetector [16,41]:

$$R = \frac{{{I_{Light}} - {I_{Dark}}}}{{{P_{Laser}}\ast S}}$$
$$EQE = \frac{{Rhc}}{{e\lambda }}$$
$${D^\ast } = \frac{R}{{\sqrt {2e{I_{Dark}}} }}$$
where ILight is the photocurrent, IDark is the dark current, PLaser is the optical power density of the incident light, and S is the effective illumination area. In addition, h is Planck's constant, e is the number of charges, c is the speed of light in a vacuum, and λ is the wavelength of the incident light.

The time-resolved photoresponse characteristics of the CsPbI3:Ag/PEDOT:PSS detector to different wavelengths of light are analyzed and studied. Figure 2(a)–(d) show the CsPbI3:Ag/PEDOT:PSS's time-resolved photoresponse curves (I-t) to 450 nm, 780 nm, 1560 nm, and 355 nm irradiation, respectively, under a 1 V bias voltage with 20-s periods. The photocurrents excited by the 450 nm, 780 nm, 1560 nm, and 355 nm light switch rapidly between the high and low levels when the incident light is switched on and off repeatedly. Figure 2(a) and 2(b) show the time-dependent changes in the photo-to-dark current in visible light. As shown in Fig. S1, the I-V characteristic curves of the CsPbI3:Ag/PEDOT:PSS heterojunction under 450 nm light, 780 nm light, 1560 nm light. As shown in Fig. 2(a), under continuous external excitation by light irradiation at 450 nm, the photocurrent change in the sample increases with increasing light intensity. However, because of its large base dark current, the sample’s responsivity is 1.7 mA/W, and its detectivity is 2.54 × 109 Jones. As shown in Fig. 2(b), the photo-to-dark current change in the sample under continuous external excitation by light at 780 nm is much smaller than that at 450 nm. The responsivity and detectivity of the sample in this case are 0.63 mA/W and 2.3 × 108 Jones, respectively. Under the 1560 nm excitation condition in the near infrared, the responsivity and detectivity of the sample are 0.32 mA/W and 1.9 × 108 Jones, respectively. These characteristics are close to those measured under the 780 nm excitation condition. Additionally, the performance under 355 nm pulsed laser excitation is close to that obtained under the 450 nm excitation condition, with a responsivity of 4.92 mA/W and a detectivity of 1.65 × 109 Jones.

 figure: Fig. 2.

Fig. 2. CsPbI3:Ag/ PEDOT:PSS sample I-t characteristics under 1 V bias for excitation at (a) 450 nm, (b) 780 nm, (c) 1560 nm, and (d) 355 nm.

Download Full Size | PDF

Figure 3(a)–(d) show the CsPbI3:Ag/PEDOT:PSS's time-resolved photoresponse curves (I-t) to 450 nm, 780 nm, 1560 nm, and 355 nm irradiation, respectively, under a 0 V bias voltage with 20-s periods. The results in Fig. 3 show that the CsPbI3:Ag/PEDOT:PSS structure can be used to construct a broadband self-driven detector for operation over the UV to near infrared range. As shown in Fig. 4, we calculated the responsivity (R), the external quantum efficiency (EQE) and, as the most important indicator, the detectivity (D*) of the CsPbI3:Ag/PEDOT:PSS/quartz structure under the 0 V bias for the different wavelength bands of light. As shown in Fig. 4(a)–4(c), the responsivity of the sample is 1.21 × 10−3 mA/W under continuous external excitation by light at 450 nm, and the detection rate is 4.9 × 106 Jones. Under continuous light excitation at 780 nm, the responsivity of the sample is 1.14 × 10−3 mA/W, and the detectivity is 3.37 × 106 Jones, as shown in Fig. 4(d)–4(f). As shown in Fig. 4(g)–4(h), although the energy of the near-infrared light at 1560 nm is lower, the responsivity of the structure at that wavelength is 1.22 × 10−3 mA/W and the detectivity is 1.49 × 107 Jones. Because the energy of the ultraviolet light is stronger, the responsivity of the sample under 355 nm pulsed light reaches 2.62 × 10−3 mA/W, which is numerically higher than the corresponding figures for visible light and near infrared light. Similarly, the detectivity at this wavelength is 5.17 × 107 Jones, which is the highest among the band we investigated.

 figure: Fig. 3.

Fig. 3. CsPbI3:Ag/ PEDOT:PSS sample I-t characteristics under 0 V bias for excitation at (a) 450 nm, (b) 780 nm, (c) 1560 nm, and (d) 355 nm.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. CsPbI3:Ag/ PEDOT:PSS sample under 0 V bias. (a) 450 nm: R, (b) 450 nm: EQE, (c) 450 nm: D*, (d) 780 nm: R, (e) 780 nm: EQE;, (f) 780 nm: D*, (g) 1560 nm: R, (h) 1560 nm: EQE, (i) 1560 nm: D*, (j) 355 nm: R, (k) 355 nm: EQE. (l) 355 nm: D*.

Download Full Size | PDF

Based on the results presented above, a self-driven broadband photodetector with a photoconductive structure was constructed and measured. For the mechanism of this device, the heterojunction is formed between the PEDOT:PSS layer (with 9:1 DMSO doping) and the silver-doped CsPbI3 layer, with a channel for carrier transport being formed in heterojunction, was assumed. To confirm this assumption, we performed comparison measurement under the 0 V bias, with results as shown in Fig. 5. Figure 5(a) shows that the photocurrent of the undoped sample still varies with the light intensity under 450 nm laser irradiation. The responsivity and detectivity of this sample are 9.12 × 10−4 mA/W and 2.43 × 107 Jones, respectively. As shown in Fig. 5(b), the responsivity of the undoped sample at 780 nm is 6.21 × 10−4 mA/W and its detectivity is 1.44 × 107 Jones. In the visible light band, the undoped sample and the silver-doped sample show similar results. However, under the 1560 nm laser irradiation, as shown in Fig. 5(c), the sample can only respond under the external excitation light irradiation at a power density of close to 1.6 W/cm2. In addition, the photocurrent of the sample changed insignificantly with changes in the light intensity in this case. Moreover, the undoped sample could not detect the 355 nm pulsed light. These results clearly show that doping with the silver nanoparticles allows the sample to detect light over a broadband range. Furthermore, we retained the silver-doped perovskite structure and removed the PEDOT:PSS layer for the next measurement. After the removal of the carrier transport layer, the sample with this structure can only respond to continuous light at 450 nm with little changes in photocurrent, as illustrated in Fig. 5(d). The responsivity and detectivity in this case are 5.41 × 10−4 mA/W and 2.14 × 107 Jones, respectively. The results changed again when both the silver-doped perovskite and the PEDOT:PSS layer were removed. Under excitation from 450 nm, 780 nm, and 1560 nm continuous light and 355 nm light, no signals were detected. Obviously, the CsPbI3:Ag/PEDOT:PSS structure forms a heterojunction, and only this structure can complete the broadband detection over the range from ultraviolet to near infrared. Thus, a self-driven broadband detector with a photoconductive structure was successfully constructed.

 figure: Fig. 5.

Fig. 5. CsPbI3/ PEDOT:PSS sample I-t characteristics under 0 V bias for excitation at (a) 450 nm, (b) 780 nm, and (c) 1560 nm. (d) CsPbI3:Ag sample I-t characteristics under 0 V bias for excitation at 450 nm.

Download Full Size | PDF

To verify the response mechanism of the CsPbI3:Ag/PEDOT:PSS detector to visible light and near-infrared light, the carrier response of the perovskite heterojunction was investigated at different wavelengths via terahertz (THz) time-domain spectroscopy, as shown in Fig. 6(a). The terahertz transmission time-domain and frequency-domain graphs that were obtained from the testing are used to characterize the terahertz transmission. When the CsPbI3:Ag/PEDOT:PSS is excited by 450 nm light, the terahertz transmission signal drops significantly, as shown in both Fig. 6(b) and Fig. 6(c), which means that large numbers of carriers are generated by the excitation, thereby reducing the terahertz transmission by 24%. When the CsPbI3:Ag/PEDOT:PSS is excited by 780 nm light, the terahertz transmission signal drops, as shown in both Fig. 6(d) and Fig. 6(e), which means that less numbers of carriers are generated by the excitation, thereby reducing the terahertz transmission by 4%. In CsPbI3:Ag/PEDOT:PSS structure, after doping with silver nanoparticles, the CsPbI3 structure and the positive silver ions form a P-type doping profile, which then forms a P-N junction [38]. Excessive silver nanoparticles will produce surface plasmon resonance effect after the surface is irradiated by laser, which increases the generation of carriers [42,43]. Doping of silver nanoparticles results in AgI, which reduced the Fermi level of perovskite, resulting in an energy level match between perovskite and PEDOT:PSS [42,44]. The results show that the THz signal peak decreases and conductivity of the sample increases as the excited light intensity increases, which caused the photocurrent increased.

 figure: Fig. 6.

Fig. 6. (a) Schematic of the terahertz time-domain spectroscopic (THz-TDS) system and the sample in CsPbI3:Ag/PEDOT:PSS. (b) THz transmission amplitude and (c) transmission power through CsPbI3:Ag/PEDOT:PSS sample with the 450 nm incident light. (d) THz transmission amplitude and (e) transmission power through CsPbI3:Ag/PEDOT:PSS sample with the 780 nm incident light.

Download Full Size | PDF

Because of the broadband detection capability of the hybrid CsPbI3:Ag/PEDOT:PSS structure, it can be developed further to provide an imaging capability that ranges from visible to the near infrared. As shown in Fig. 7(a), we fabricated 4 × 4 arrayed parallel single-slit devices separated on the 2 cm×2 cm×1 mm perovskite heterojunction structure. In this way, the 16 independent detection units can form 16 pixel points. We can control the different units by light excitation to present the pattern required. English letters including “C”, “N”, and “U” were imaged as shown in Fig. 7(b) and (c). The colored areas show the pixels that are responding to the excitation light, with a function similar to a “1”. Additionally, the white parts represent the pixels that are not under light excitation. We used the 450 nm laser (as shown in Fig. 7(b)) and the 1560 nm laser (as shown in Fig. 7(c)) to illustrate the imaging effects of the device in the visible and near infrared bands. The results show that this structure can be extended for use as an imaging device.

 figure: Fig. 7.

Fig. 7. (a) Schematic imaging device. Imaging results for “C”, “N”, and “U” under (b) 450 nm and (c) 1560 nm laser irradiation.

Download Full Size | PDF

4. Conclusions

In summary, a self-powered ultraviolet-to-near infrared broadband photodetector based on a Ag-doped CsPbI3/PEDOT:PSS heterojunction was investigated. The photodetector using a CsPbI3:Ag/PEDOT:PSS heterostructure with a planar photoconductive structure operated over a broad 355–1560 nm wavelength range in self-powered mode. A terahertz signal was modulated with the CsPbI3:Ag/PEDOT:PSS structure at low optical excitation intensity to investigate its photodetection mechanism. The experimentally designed detector can present images of the letters “C”, “N” and “U” in the visible and near-infrared wavelengths, indicating a potential broadband imaging application.

Funding

Project of High-level Teachers in Beijing Municipal Universities in the Period of 13th Five-year Plan; Beijing Municipal Natural Science Foundation (4202013); National Natural Science Foundation of China (61505125, 62175168).

Acknowledgments

This research was supported by National Natural Science Foundation of China (61505125, 62175168), Nature Science Foundation of Beijing Municipality (4202013) and High-level Teachers in Beijing Municipal Universities in the Period of 13th Five–year Plan.

Disclosures

The authors declare no conflict of interest.

Data availability

All data included in this study are available upon request by contact with the corresponding author.

Supplemental document

See Supplement 1 for supporting content.

References

1. C. Zhang, D. B. Kuang, and W. Q. Wu, “A Review of Diverse Halide Perovskite Morphologies for Efficient Optoelectronic Applications,” Small Methods 4(2), 1900662 (2020). [CrossRef]  

2. E. A. Gaulding, J. Hao, H. S. Kang, E. M. Miller, S. N. Habisreutinger, Q. Zhao, A. Hazarika, C. P. Sercel, and M. J. Luther, “Conductivity tuning via doping with electron donating and withdrawing molecules in perovskite CsPbI3 nanocrystal films,” Adv. Mater. 31(27), 1902250 (2019). [CrossRef]  

3. H. Wang and D. H. Kim, “Perovskite-based photodetectors: materials and devices,” Chem. Soc. Rev. 46(17), 5204–5236 (2017). [CrossRef]  

4. L. Wu, K. Chen, W. Huang, Z. Lin, J. Zhao, X. Jiang, Y. Ge, F. Zhang, Q. Xiao, Z. Guo, Y. Xiang, J. Li, Q. Bao, and H. Zhang, “Perovskite CsPbX3: a promising nonlinear optical material and its applications for ambient all-optical switching with enhanced stability,” Adv. Opt. Mater. 6(19), 1800400 (2018). [CrossRef]  

5. Z. Li, F. Zhou, H. Yao, Z. Ci, Z. Yang, and Z. Jin, “Halide perovskites for high-performance X-ray detector,” Mater. Today 48, 155–175 (2021). [CrossRef]  

6. U. Bansode, A. Rahman, and S. Ogale, “Low-temperature processing of optimally polymer-wrapped α-CsPbI3 for self-powered flexible photo-detector application,” J. Mater. Chem. C 7(23), 6986–6996 (2019). [CrossRef]  

7. H. J. Na, N. K. Cho, J. Park, S. E. Lee, E. G. Lee, C. Im, and Y. S. Kim, “A visible light detector based on a heterojunction phototransistor with a highly stable inorganic CsPbIxBr3−x perovskite and In-Ga-Zn-O semiconductor double-layer,” J. Mater. Chem. C 7(45), 14223–14231 (2019). [CrossRef]  

8. S. Mu, Q. Ye, X. Zhang, S. Huang, and J. You, “Polymer hole-transport material improving thermal stability of inorganic perovskite solar cells,” Front. Optoelectron 13(3), 265–271 (2020). [CrossRef]  

9. J. Huang, K. X. Wang, J. J. Chang, Y. Y. Jiang, Q. S. Xiao, and Y. Li, “Improving the efficiency and stability of inverted perovskite solar cells with dopamine-copolymerized PEDOT:PSS as a hole extraction layer,” J. Mater. Chem. A 5(26), 13817–13822 (2017). [CrossRef]  

10. L. Li, F. Zhang, S. Ye, X. Peng, Z. Sun, J. Lian, L. Liu, J. Qu, and J. Song, “Self-powered photodetectors based on CsxDMA1-xPbI3 perovskite films with high detectivity and stability,” Nano Energy 71, 104611 (2020). [CrossRef]  

11. C. Paduani and A. M. Rappe, “Tuning the gap of lead-based halide perovskites by introducing superalkali species at the cationic sites of ABX3-type structure,” Phys. Chem. Chem. Phys. 19(31), 20619–20626 (2017). [CrossRef]  

12. Y. Li, Y. Zhang, Z. Chen, Q. Li, T. Li, M. Li, H. Zhao, Q. Sheng, W. Shi, and J. Yao, “Self-powered, flexible, and ultrabroadband ultraviolet-terahertz photodetector based on a laser-reduced graphene oxide/CsPbBr3 composite,” Photonics Res. 8(8), 1301–1308 (2020). [CrossRef]  

13. R. Saraf and V. Maheshwari, “Self-powered photodetector based on electric-field-induced effects in MAPbI3 perovskite with improved stabilit,” ACS Appl. Mater. Interfaces 10(25), 21066–21072 (2018). [CrossRef]  

14. W. Tian, H. Zhou, and L. Li, “Hybrid organic–inorganic perovskite photodetectors,” Small 13(41), 1702107 (2017). [CrossRef]  

15. L. Su, Z. X. Zhao, H. Y. Li, J. Yuan, Z. L. Wang, G. Z. Cao, and G. Zhu, “High-performance organolead halide perovskite-based self-powered triboelectric photodetector,” ACS Nano 9(11), 11310–11316 (2015). [CrossRef]  

16. T. Pang, R. Jia, Y. Wang, K. Sun, Z. Hu, Y. Zhu, S. Luan, and Y. Zhang, “Self-powered behavior based on the light-induced self-poling effect in perovskite-based transport layer-free photodetectors,” J. Mater. Chem. C 7(3), 609–616 (2019). [CrossRef]  

17. W. Ahmad, J. Khan, G. Niu, and J. Tang, “Inorganic CsPbI3 perovskite-based solar cells:A choice for a tandem device,” Sol. RRL 1(7), 1700048 (2017). [CrossRef]  

18. J. Deng, J. Li, Z. Yang, and M. Wang, “All-inorganic lead halide perovskites: a promising choice for photovoltaics and detectors,” J. Mater. Chem. C 7(40), 12415–12440 (2019). [CrossRef]  

19. Q. Xu, C. Li, J. Nie, Y. Guo, X. Wang, B. Zhang, and X. Ouyang, “Highly Sensitive and Stable X-ray Detector Based on a 0D Structural Cs4PbI6 Single Crystal,” J. Phys. Chem. Lett. 12(1), 287–293 (2021). [CrossRef]  

20. J. Lu, X. Sheng, G. Tong, Z. Yu, X. Sun, L. Yu, X. Xu, J. Wang, J. Xu, Y. Shi, and K. Chen, “Ultrafast Solar-Blind Ultraviolet Detection by Inorganic Perovskite CsPbX3 Quantum Dots Radial Junction Architecture,” Adv. Mater. 29(23), 1700400 (2017). [CrossRef]  

21. X. Wu, J. Sun, H. Shao, Y. Zhai, L. Li, W. Chen, J. Zhu, B. Dong, L. Xu, D. Zhou, W. Xu, H. Song, and X. Bai, “Self-powered UV photodetectors based on CsPbCl3 nanowires enabled by the synergistic effect of acetate and lanthanide ion passivation,” Chem. Eng. J. 426, 131310 (2021). [CrossRef]  

22. B. B. Zhang, X. Liu, B. Xiao, A. B. Hafsia, K. Gao, Y. Xu, J. Zhou, and Y. Chen, “High-performance X-ray detection based on one-dimensional inorganic halide perovskite CsPbI3,” J. Phys. Chem. Lett. 11(2), 432–437 (2020). [CrossRef]  

23. B. Xin, N. Alaal, S. Mitra, A. Subahi, Y. Pak, D. Almalawi, N. Alwadai, S. Lopatin, and I. S. Roqan, “Identifying Carrier Behavior in Ultrathin Indirect-Bandgap CsPbX3 Nanocrystal Films for Use in UV/Visible-Blind High-Energy Detectors,” Small 16(43), 2004513 (2020). [CrossRef]  

24. F. Maddalena, A. Xie, X. Y. Chin, R. Begum, M. E. Witkowski, M. Makowski, B. Mahler, W. Drozdowski, S. V. Springham, R. S. Rawat, N. Mathews, C. Dujardin, M. D. Birowosuto, C. Dang, Deterministic Light Yield, and Fast Scintillation, “and Microcolumn Structures in Lead Halide Perovskite Nanocrystals,” J. Phys. Chem. C 125(25), 14082–14088 (2021). [CrossRef]  

25. C. Zou, Y. Xi, C. Y. Huang, E. G. Keeler, T. Feng, S. Zhu, L. D. Pozzo, and L. Y. Lin, “A highly sensitive UV–vis–NIR all-inorganic perovskite quantum dot phototransistor based on a layered heterojunction,” Adv. Opt. Mater. 6(14), 1800324 (2018). [CrossRef]  

26. F. Mei, D. Sun, S. Mei, J. Feng, Y. Zhou, J. Xu, and X. Xiao, “Recent progress in perovskite-based photodetectors: the design of materials and structures,” Adv. Phys.: X 4(1), 1592709 (2019). [CrossRef]  

27. G. E. Eperon, G. M. Paternò, R. J . Sutton, A. Zampetti, A. A. Haghighirad, F. Cacialli, and H. J. Snaith, “Inorganic caesium lead iodide perovskite solar cells,” J. Mater. Chem. A 3(39), 19688–19695 (2015). [CrossRef]  

28. Q. Niu, W. Huang, J. Tong, H. Lv, Y. Deng, Y. Ma, Z. Zhao, R. Xia, W. Zeng, Y. Min, and W. Huang, “Understanding the mechanism of PEDOT:PSS modification via solvent on the morphology of perovskite films for efficient solar cells,” Synth. Met. 243, 17–24 (2018). [CrossRef]  

29. J. M. Yun, J. S. Yeo, J. Kim, H. G. Jeong, D. Y. Kim, Y. J. Noh, S. S. Kim, B. C. Ku, and S. I. Na, “Solutio-processable reduced graphene oxide as a novel alternative to PEDOT:PSS hole transport layers for highly efficient and stable polymer solar cells,” Adv. Mater. 23(42), 4923–4928 (2011). [CrossRef]  

30. A. A. Syed, C. Y. Poon, H. W. Li, and F. Zhu, “A sodium citrate-modified-PEDOT:PSS hole transporting layer for performance enhancement in inverted planar perovskite solar cells,” J. Phys. Chem. C 7(18), 5260–5266 (2019). [CrossRef]  

31. M. Vosgueritchian, D. J. Lipomi, and Z. Bao, “Highly conductive and transparent PEDOT:PSS films with a fluorosurfactant for stretchable and flexible transparent electrodes,” Adv. Funct. Mater. 22(2), 421–428 (2012). [CrossRef]  

32. G. Wang, S. I. Na, T. W. Kim, Y. Kim, S. Park, and T. Lee, “Effect of PEDOT:PSS–molecule interface on the charge transport characteristics of the large-area molecular electronic junctions,” Org. Electron. 13(5), 771–777 (2012). [CrossRef]  

33. H. Zhou, M. H. Chua, Q. Zhu, and J. Xu, “High-performance PEDOT:PSS-based thermoelectric composites,” Composites Commun. 27, 100877 (2021). [CrossRef]  

34. S. Suchol, C. Esther, M. R. Sandro, D. P. Adam, V. Z. Aliaksandr, F. O. C. Timothy, and J. L. Darren, “Effect of Broken Conjugation on the Stretchability of Semiconducting Polymers,” Adv. Funct. Mater. 25(3), 427–436 (2015). [CrossRef]  

35. A. Nahata, J. T. Yardley, and T. F. Heinz, “Free-space electro-optic detection of continuous-wave terahertz radiation,” Appl. Phys. Lett. 75(17), 2524–2526 (1999). [CrossRef]  

36. D. Y. Choi, H. W. Kang, H. J. Sung, and S. S. Kim, “Annealing-free, flexible silver nanowire-polymer composite electrodes via a continuous two-step spray-coating method,” Nanoscale 5(3), 977–983 (2013). [CrossRef]  

37. M. Tonouchi, “Cutting-edge terahertz technology,” Nat. Photonics 1(2), 97–105 (2007). [CrossRef]  

38. B. Liu, J. Liu, L. Lv, Y. Hou, J. Shen, and B. Zhang, “Stable Terahertz In Situ Photo-Writable Electrically Erasable Memory with a CsPbI3:Ag/SnO2/PEDOT:PSS Hybrid Structure,” ACS Appl. Electron. Mater. 3(2), 1006–1014 (2021). [CrossRef]  

39. X. Wang, J. Weng, M. Xu, X. Li, T. Yang, Y. Hou, J. Shen, and B. Zhang, “Morphology engineering of a hybrid perovskite for active terahertz memory modulation,” Opt. Express 30(2), 2626–2635 (2022). [CrossRef]  

40. S. V. Barma, S. R. Rondiya, B. R. Bade, M. P. Nasane, S. B. Jathar, K. B. Kore, P. U. Tandale, D. S. Nilegave, S. R. Jadkar, and A. M. Funde, “Phase stability investigation of CsPbI3 perovskite for solar cell application,” AIP Conf. Proc. 2335, 040001 (2021). [CrossRef]  

41. T. Yang, F. Li, and R. Zheng, “Recent progress on cesium lead halide perovskites for photodetection applications,” ACS Appl. Electron. Mater. 1(8), 1348–1366 (2019). [CrossRef]  

42. L. Xiong, B. Liu, D. Liu, L. Lv, Y. Hou, J. Shen, and B. Zhang, “An in situ rewritable electrically-erasable photo-memory device for terahertz waves,” Nanoscale 12(5), 3343–3350 (2020). [CrossRef]  

43. C. An, J. Wang, J. Liu, S. Wang, and Q. H. Zhang, “Plasmonic enhancement of photocatalysis over Ag incorporated AgI hollow nanostructures,” RSC Adv. 4(5), 2409–2413 (2014). [CrossRef]  

44. S. Shahbazi, C. M. Tsai, S. Narra, C. Y. Wang, H. S. Shiu, S. Afshar, N. Taghavinia, and E. W. G. Diau, “Ag doping of organometal lead halide perovskites: morphology modification and p-type character,” J. Phys. Chem. C 121(7), 3673–3679 (2017). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Figure S1. I-V characteristic curves of the CsPbI3:Ag/PEDOT:PSS heterojunction at (a) 450 nm, (b) 780 nm and (c) 1560 nm.

Data availability

All data included in this study are available upon request by contact with the corresponding author.

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 (7)

Fig. 1.
Fig. 1. (a) Example of the CsPbI3:Ag/PEDOT:PSS structure and experimental setup for the electrical measurements. (b) Absorption spectra of PEDOT:PSS, the CsPbI3:Ag/PEDOT:PSS structure, and the CsPbI3:Ag/PEDOT:PSS structure. SEM images of (c) CsPbI3:Ag/PEDOT:PSS structure and (d) CsPbI3/PEDOT:PSS structure. X-ray diffraction results for (e) CsPbI3:Ag/PEDOT:PSS structure (f) CsPbI3/PEDOT:PSS structure (g) CsPbI3:Ag structure (h) CsPbI3 structure.
Fig. 2.
Fig. 2. CsPbI3:Ag/ PEDOT:PSS sample I-t characteristics under 1 V bias for excitation at (a) 450 nm, (b) 780 nm, (c) 1560 nm, and (d) 355 nm.
Fig. 3.
Fig. 3. CsPbI3:Ag/ PEDOT:PSS sample I-t characteristics under 0 V bias for excitation at (a) 450 nm, (b) 780 nm, (c) 1560 nm, and (d) 355 nm.
Fig. 4.
Fig. 4. CsPbI3:Ag/ PEDOT:PSS sample under 0 V bias. (a) 450 nm: R, (b) 450 nm: EQE, (c) 450 nm: D*, (d) 780 nm: R, (e) 780 nm: EQE;, (f) 780 nm: D*, (g) 1560 nm: R, (h) 1560 nm: EQE, (i) 1560 nm: D*, (j) 355 nm: R, (k) 355 nm: EQE. (l) 355 nm: D*.
Fig. 5.
Fig. 5. CsPbI3/ PEDOT:PSS sample I-t characteristics under 0 V bias for excitation at (a) 450 nm, (b) 780 nm, and (c) 1560 nm. (d) CsPbI3:Ag sample I-t characteristics under 0 V bias for excitation at 450 nm.
Fig. 6.
Fig. 6. (a) Schematic of the terahertz time-domain spectroscopic (THz-TDS) system and the sample in CsPbI3:Ag/PEDOT:PSS. (b) THz transmission amplitude and (c) transmission power through CsPbI3:Ag/PEDOT:PSS sample with the 450 nm incident light. (d) THz transmission amplitude and (e) transmission power through CsPbI3:Ag/PEDOT:PSS sample with the 780 nm incident light.
Fig. 7.
Fig. 7. (a) Schematic imaging device. Imaging results for “C”, “N”, and “U” under (b) 450 nm and (c) 1560 nm laser irradiation.

Equations (3)

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

R = I L i g h t I D a r k P L a s e r S
E Q E = R h c e λ
D = R 2 e I D a r k
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.