Next Article in Journal
Effect of Local Topography on Cell Division of Staphylococcus spp.
Previous Article in Journal
Morphological and Electrochemical Properties of ZnMn2O4 Nanopowders and Their Aggregated Microspheres Prepared by Simple Spray Drying Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Growth of PbS/PbI2 Heterojunction and Its Photoelectric Properties

1
School of Weapons Science and Technology, Xi’an Technological University, Xi’an 710032, China
2
School of Optoelectronic Engineering, Xi’an Technological University, Xi’an 710032, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(4), 681; https://doi.org/10.3390/nano12040681
Submission received: 14 January 2022 / Revised: 4 February 2022 / Accepted: 15 February 2022 / Published: 18 February 2022

Abstract

:
In this paper, PbI2 thin films with a uniform surface morphology and compact structure were prepared by adjusting the spin coating process parameters. On such a basis, the PbS/PbI2 heterojunction was fabricated on the PbI2 surface by the method of in situ chemical replacement growth. The results show that the PbS/PbI2 heterojunction grown by this method has a clear interface and is closely combined. The introduction of a PbS layer enables its spectral response range to cover the visible and near-infrared regions. Compared with the PbI2 thin film device, its responsivity is increased by three orders of magnitude, its response time reduced by 42%, and its recovery time decreased by nearly 1/2 under 450 nm illumination. In the case that there is no response for the PbI2 thin film device under 980 nm illumination, the specific detectivity of the PbS/PbI2 heterojunction device still amounts to 1.8 × 108 Jones. This indicates that the in situ chemical replacement is a technique that can construct a high-quality heterojunction in a simple process. PbS/PbI2 heterojunction fabricated by this method has a visible–near-infrared light detection response range, which provides a new idea for creating visible–near-infrared common-path detection systems.

1. Introduction

As an important part of photoelectric information collection, photoelectric detectors are widely used in vehicle driving, motion sensing, security alarms, infrared induction imaging and other aspects [1,2,3,4,5,6,7]. Photoelectric detection systems working within different wavelength ranges have their own advantages in terms of their capability and application. However, with the continuous development of science and technology as well as the emergence of various unforeseen circumstances and interference technologies, users’ requirements for photodetectors are growing. Therefore, photodetectors that only respond in a single-wave band are far from meeting their actual needs. Efficiently obtaining accurate multi-band information in complex environments is extremely important. Thus, the photoelectric detection system with dual-band or multi-band fusion were invented [8,9,10,11,12]. However, the previously developed dual-band or multi-band detection systems are usually composed of two or more independent detection systems. Such systems are costly in terms of production and large in volume, which runs counter to the idea of high integration, miniaturization and lightweight required for detection systems. Therefore, collecting dual-band or multi-band information with a single detector to achieve broad-spectrum common-path detection is the key to solving the above problems [13,14,15].
In order to expand the response range of photoelectric devices and improve their responsivity, a good few scientists have proposed many schemes of referential meaning. For example, the heavily doped silicon photoelectric device fabricated by Aurore J. Said et al. achieved an external quantum efficiency of 3000% at 12 V bias voltage, with a detection range from the visible region to the near-infrared region [16]. Although such a method of doping traditional silicon substrate and preparing a microstructure can widen the response range of the device, the preparation method is complex and will bring about a substantial number of defects in the silicon material, thus amplifying the dark current of the photoelectric device and decreasing its responsivity and sensitivity. Therefore, there exist some limits for the photoelectric device fabricated with a single material. Wang et al. doped PTB7-Th in the P3HT:PC71BM system and successfully obtained a highly sensitive polymer photoelectric device with a spectral response range from the ultraviolet region to the near-infrared region [17]. Photoelectric devices fabricated with organic semiconductors have an ultra-wide detection spectral range, but the instability of organic materials will make their service life much shorter than those fabricated with inorganic materials. Zhuo et al. prepared GaN thin films on sapphire substrates in the method of MOCVD, and stacked GaN with β-In2Se3 obtained by mechanical exfoliation to fabricate the photodetector with a heterojunction structure [18]. It achieved responses in the UV and near-infrared regions with a responsivity of 1.6 A/W under 365 nm and that of 0.03 A/W under 850 nm as well. The process of fabricating heterojunction is complex. The size of the device is limited by that of the tube furnace chamber, and the cleanliness of the stack surface is highly required in the physical stacking process. Mukherjee et al. realized the in situ growth of high-quality 0D/2D mixed dimensional heterojunction structure in MoS2 by the low-temperature hydrothermal method [19]. Therefore, the device prepared with the heterojunction had responses at 400~1600 nm with a specific detectivity of 1012 Jones. The process of fabricating heterojunction by this method is simpler than that by the chemical vapor deposition (CVD) method, and the as-prepared suspension can be fabricated with large devices by spin coating, spraying and other methods. However, two-dimensional materials in the suspension are small in size, stacking may arise under certain circumstances, and the agglomeration of quantum dots may occur to some extent. In addition, the preparation of the carrier transport layer is required as an auxiliary for its preparation with large devices.
In this paper, a simpler method for fabricating visible–near-infrared photodetectors was proposed. The heterojunction was constructed through the in situ growth of the PbS layer on the surface of the PbI2 thin film. Therein, the two materials were responsible for the absorption of and response in different wavebands, respectively. The spectral response range of the photodetectors was expanded. The compact PbI2 thin film also had excellent carrier mobility, which can act as a carrier migration layer. This study provides an idea and method with great application potential for fabricating array photodetectors with all-day and wide-spectrum response.

2. Materials and Methods

Figure 1 shows the specific preparation process of the PbI2 thin film and PbI2/PbS heterojunction. The PbI2 solution was prepared at room temperature with N, N-dimethylformamide (DMF) (Alfa Aesar, Shanghai, China) and dimethyl sulfoxide (DMSO) (Alfa Aesar, Shanghai, China) mixed solvent. The prepared solution was magnetically stirred for 12 h at 70 °C and 100 r/min, and was aged for 48 h. Then, the prepared solution was spin-coated on the surface of the cleaned sodium calcium glass at a rotational speed of 6000 r/min for 30 s, and annealed at 80 °C for 20 min to obtain the PbI2 thin film. Finally, the PbI2 thin film was put into 1 mg/mL of Na2S (Alfa Aesar, Shanghai, China) ethanol solution and taken out, and the PbI2/PbS heterojunction was obtained after the surface of the sample was cleaned with anhydrous ethanol (Alfa Aesar, Shanghai, China) and dried. The purities of PbI2, Na2S, DMF, and DMSO are 99, 98, 99.9, and 99%, respectively.
The phase characteristics of samples was tested through X-ray diffraction (XRD) analysis (D8 ADVANCE, Bruker AXS, Karlsruhe, German). The morphological property was observed with a field-emission scanning electron microscopy (GeminiSEM 500, Zeiss, Aalen, Germany). The laser Raman spectroscope (LabRAM HR Evolution, HORIBA, Fukuoka, Japan) was used for Raman testing and analysis with a 532 nm laser for excitation. The absorption spectra were collected by using the UV-NIR spectrometer (JASCO V-570 UV/vis/NIR, JASCO, Tokyo, Japan). Spectral response characteristics of samples were measured with optical spectrum analyzer (DSR-F4-XIAN, Zolix, Beijing, China). The I–V curves and time-resolved photoresponse of the samples were obtained with a self-organized detection system which includes two digital source tables (2280 and 2450 Keithley, Cleveland, OH, USA), and two monochromatic light emitting diodes (450 and 980 nm, Shanghai, China). The optical power meter (843-R, Newport, CA, USA) was used for light intensity calibration.

3. Results and Discussion

In this study, PbI2 thin film was used as precursor, and a layer of PbS thin film was formed in situ on the surface of the PbI2 thin film through the replacement reaction between PbI2 and Na2S, thus forming the PbS/PbI2 heterojunction. As shown in Figure 1, before and after the replacement reaction, the surface color of the sample changed from light yellow to dark brown due to the difference in absorption spectra between PbS and PbI2.
In order to obtain a PbS/PbI2 heterojunction with excellent performance, the quality of the PbI2 thin film as a precursor is essential. In this paper, the effects of PbI2 concentration and the ratio of dimethyl sulfoxide (DMSO) to dimethyl formamide (DMF) on the quality of PbI2 thin film in its preparation were systematically studied. As shown in Figure S1a, a transparent, clear and precipitate-free PbI2 solution can only be obtained when the ratio of DMF to DMSO is 9:1 after aging for two days. In the solvent with such a ratio, PbI2 concentration cannot exceed 1.75 M, otherwise precipitation will be produced, as shown in Figure S1b. In accordance with Figures S2 and S3, as the concentration of PbI2 solution increases from 1 M to 1.75 M, PbI2 thin film becomes more compact and smoother, and its thickness rises from approximately 70 to approximately 300 nm. The visible–near-infrared spectrophotometer test shows that the PbI2 thin film prepared with 1.75 M of PbI2 solution has the best visible light absorption, as shown in Figure S4a.
In an effort to investigate the effects of the in situ growing time of PbS on the photon absorbing capacity of the PbS/PbI2 heterojunction in the near-infrared band, the growing time was set to 5, 10, 15, 20, 25, and 30 s, respectively. As shown in Figure S4b, the PbS/PbI2 heterojunction has the best absorption in the near-infrared band when the growing time was 20 s. Furthermore, if the growing time exceeds 30 s, some part of the PbI2 thin film substrates will be dissolved and exfoliated by anhydrous ethanol, as shown in Figure S5.
In view of the above results, the process parameters of preparing PbI2 thin film and PbS/PbI2 heterojunction were determined in the paper as follows: the PbI2 precursor thin film was prepared with the solvent of which the ratio of DMF to DMSO was 9:1 and PbI2 solution with a concentration of 1.75 M; PbI2 precursor thin film was soaked in Na2S ethanol solution for 20 s. The SEM images of PbI2 thin film and PbS/PbI2 heterojunction prepared by the above process parameters is shown in Figure 2, they both have smooth and compact surfaces, as well as a clear heterojunction interface with the thicknesses of the PbS layer and PbI2 layer being approximately 25 and 230 nm, respectively.
To verify whether the samples were the target products, they were characterized with XRD and Raman spectra. Figure 3a shows the XRD curves of the prepared PbI2 thin film and PbS/PbI2 heterojunction. Compared with the powder diffraction standard cards of PbI2 and PbS, it was found that there exists a one-to-one correspondence between the XRD diffraction peak of the PbI2 thin film and the standard cards of PbI2. A very weak diffraction peak appears at 30° in the XRD spectra of the PbS/PbI2 heterojunction. In comparison with the standard card of PbS, this diffraction peak should correspond to the (200) crystalline of PbS. Since the diffraction peak of PbS is very weak, it is impossible to completely determine whether the PbS layer has been deposited on the surface of the sample. Therefore, the samples are also characterized with Raman spectra. Figure 3b shows the Raman spectra curve of the as-prepared PbS/PbI2 heterojunction, and the inset indicates the comparison of the Raman spectra curves between PbI2 thin film and PbS/PbI2 heterojunction within the range of 50~300 cm−1. According to the curves, the Raman peaks of PbI2 are 70.57, 93.97, 109.32, 165.69, and 214.26 cm−1, respectively [20,21,22]. Through comparison, it was found that the Raman peaks of PbS/PbI2 heterojunction also appear at 70, 93, and 109 cm−1, and the vibration peak at 1377.55 cm−1 is a unique Raman peak of PbS [23,24]. Based on the characterization and analysis of XRD and Raman, it can be concluded that the PbS layer is indeed generated on the surface of PbI2, hence forming PbS/PbI2 heterojunction.
Figure 4a shows the visible–near-infrared absorption spectra curves of the PbI2 thin film and PbS/PbI2 heterojunction. In accordance with the figure, the optical absorption of the PbS/PbI2 heterojunction in the visible and near-infrared regions was significantly improved compared with that of PbI2 thin film. The insets in Figure 4a are the band gap width of the PbI2 thin film and PbS/PbI2 heterojunction fitted by Tauc plot method [25,26,27]. PbI2 and PbS/PbI2 are the direct band gaps with widths of 2.41 and 0.7 eV, respectively. The measured band gap of the PbI2 thin film is close to the theoretical value (2.32 eV). However, the band gap of PbS/PbI2 is mainly determined by the upper PbS, so the fitting results deviate from the theoretical band gap of PbS (0.37 eV). There are two possible reasons: firstly, PbS quantum dots which are formed through the in situ replacement reaction have a quantum confinement effect, leading to an increase in the band gap; secondly, because the PbS layer is very thin, the measured absorption spectra curve is the result of the combined action of PbS and PbI2, resulting in a larger fitting band gap. Figure 4b is the normalized photoresponse curve of PbS/PbI2 heterojunction at 30 V external bias. It can be obviously observed that PbS/PbI2 heterojunction has a response in the visible and near-infrared region, and the responsivity within the range of 400~515 nm is obviously higher than that in the range after 515 nm. The main reason for that is PbS and PbI2 can jointly generate the photocurrent within the range of 400~515 nm, while only PbS creates the photocurrent after 515 nm (2.41 eV, the energy gap of PbI2, corresponds to the absorption cutoff wavelength of 515 nm).
In order to further study the characteristics of the visible–near–infrared photoresponse of the PbI2 thin film and PbS/PbI2 heterojunction devices, an interdigital electrode was etched with a channel width of 0.20 mm on FTO glass, and the PbI2 thin film and PbS/PbI2 heterojunction were prepared on it, as shown in Figure 5a. Figure 5b shows the I–V curves of the PbS/PbI2 heterojunction device under 450 and 980 nm illumination. It can be obviously observed that the PbS/PbI2 heterojunction device has obvious characteristics of photosensitive resistance, and the photocurrent under 450 nm is clearly greater than that under 980 nm. The reason is that the photocurrent under 450 nm is jointly provided by PbI2 and PbS, while the photocurrent under 980 nm is only provided by PbS, which is consistent with the results shown in Figure 4b. Moreover, the photocurrent of the PbS/PbI2 heterojunction device is amplified with the increase in external bias voltage, which is a typical characteristic of photosensitive devices. However, when the voltage exceeds 30 V, the device will break down. Therefore, the external bias voltage applied in this study was determined as 30 V.
Under the bias voltage of 30 V, the PbI2 thin film and PbS/PbI2 heterojunction devices were periodically irradiated by monochromatic light at 450 and 980 nm wavelengths, respectively. The obtained dynamic time-resolved optical response curves are shown in Figure 6a,b. By comparison, it can be found that both devices have responses under 450 nm illumination. It is worth noting that the response current of PbS/PbI2 heterojunction is nearly three orders of magnitude higher than that of the PbI2 thin film. However, the PbI2 thin film device fails to respond under 980 nm illumination, while the PbS/PbI2 heterojunction device still has a photocurrent of 10 μA. This shows that the fabricated PbS/PbI2 heterojunction device couples the photoelectric properties of the two materials, and retains the photoresponse capabilities of PbI2 and PbS in the visible and near-infrared bands, respectively. As shown in Figure 6c, the response and recovery times of the PbI2 thin film device are 0.69 s and 0.80 s, respectively, under 450 nm illumination. Under the same conditions, those of the PbS/PbI2 heterojunction device are 0.40 s and 0.42 s, respectively. The response time is reduced by 42%, and the recovery time nearly by 1/2. The reason for that is because the heterojunction can better separate photogenerated carriers and accelerate the hole and electron transport and collection. The response and recovery times of the PbS/PbI2 heterojunction device are 0.25 and 0.45 s, respectively, under 980 nm illumination, as shown in Figure 6d. Compared with that under 450 nm, the response time of the device is shorter, which indicates that the PbI2 layer has excellent carrier transport properties.
The responsivities of the above two devices were calculated by using the formula:
R = Ilight/Pin
(Ilight refers to the photogenerated current, and Pin the incident light power). Then, the specific detectivity of the device was calculated through formula:
D * = R S 2 q I d a r k
where R refers to the responsivity, S the effective light area, q the charge constant, and Idark the dark current [28]. Figure 7 shows the responsivity and specific detectivity of the PbI2 thin film and the PbS/PbI2 heterojunction devices under incident light at different wavelengths and power densities. According to Figure 7a,b, the responsivities of the two devices increase with the decrease in light power density under 450 nm illumination, indicating that both devices have excellent characteristics of low-illumination response. The main reason is because at a low light power density, the defects of the material will capture some charges to produce a local electric field, leading to the photogating effect. This makes carriers operate in the channel and complicates the chances of recombining them, so that the carriers’ lifetime is prolonged, resulting in high gain under low light. The combination rate of the carriers increases under high power illumination, and the defects in the material are filled by carriers excited by strong light resulting in a decrease in responsivity [29,30,31]. Particularly, when the light power density is as low as 0.00667 mw/cm2, the responsivity and specific detectivity of the PbI2 thin film device are 2400 μA/W and 1.8 × 1010 Jones, respectively, while those of the PbS/PbI2 heterojunction device are 740 mA/W and 1.4 × 1011 Jones, respectively. Compared with the PbI2 thin film device, the responsivity of the PbS/PbI2 heterojunction device improves by more than 300 times and the specific detectivity by over 8 times.
Figure 7c,d show the responsivity and specific detectivity of the PbS/PbI2 heterojunction device at different light power densities under 980 nm illumination. Values of its responsivity and specific detectivity rise first and then fall as the light power density grows. When the light power density is 43.92 mw/cm2, the responsivity and specific detectivity are 1.36 mA/W and 1.8 × 108 Jones, respectively. In summary, the PbS/PbI2 heterojunction device has low illumination response characteristics in the visible region. Moreover, its photoresponse range extends to the near-infrared region. This device has a full-day-wide spectral detection capability.

4. Conclusions

In this study, the PbS/PbI2 heterojunction structure was successfully prepared by the method of growing PbS on a PbI2 thin film by in situ chemical substitution reaction. Compared with the PbI2 thin film, the spectral response range of the photoelectric device with the PbS/PbI2 heterojunction structure expanded from the previous visible light band to the visible–near-infrared band; the responsivity of the device in the visible light band rose from 2.4 × 103 μA/W to 7.4 × 102 mA/W, with an increase of two orders of magnitudes; the specific detectivity improved from 1.8 × 1010 Jones to 1.4 × 1011 Jones, an increase of one order of magnitude; and the response/recovery time decreased by 45% from 1.49 to 0.82 s. Moreover, the response/recovery time, responsivity, and specific detectivity of the PbS/PbI2 heterojunction device in the near-infrared band also reached 0.7 s, 1.09 mA/W, and 1.8 × 108 Jones, respectively. The PbS/PbI2 heterojunction device has noticeably low light response characteristics, especially in the visible region. These strengths preview the great potential of PbS/PbI2 heterojunction devices in the development of visible–near-infrared all-day spectral photoelectric devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12040681/s1, Figure S1: (a) Solubility of 1.25 M PbI2 in mixed solvents with different ratios of DMF to DMSO after ageing for 48 h; (b) Solubility of PbI2 with various concentrations when the ratio of DMF to DMSO is 9:1; Figure S2: SEM images of PbI2 thin film prepared with PbI2 solutions with various concentrations (a) 1M; (b) 1.25 M; (c) 1.50 M; and (d) 1.75 M; Figure S3: Relationship between the concentration of PbI2 solution and the thickness of PbI2 thin film; Figure S4: (a) Normalized absorbance of PbI2 thin film prepared with PbI2 solutions with various concentrations; (b) Absorbance of PbI2/PbS heterojunction at different growing times; Figure S5: SEM image of PbI2/PbS heterojunction after growing for 30 s.

Author Contributions

Conceptualization, S.Y.; data curation, J.Z.; investigation, Y.K. and H.L.; project administration, J.H.; validation, J.Z.; writing—original draft, S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key Research and Development Program of Shaanxi Province (2019ZDLGY16-01).

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.

References

  1. Barve, A.V.; Lee, S.J.; Noh, S.K.; Krishna, K. Review of current progress in quantum dot infrared photodetectors. Laser Photonics Rev. 2010, 4, 738–750. [Google Scholar] [CrossRef]
  2. Hu, W.D.; Li, Q.; Chen, X.S.; Lu, W. Recent progress on advanced infrared photodetectors. Acta Phys. Sin. 2019, 68, 120701. [Google Scholar] [CrossRef]
  3. Li, Y.; Shi, Z.; Lei, L.; Zhang, F.; Li, X.J. Highly Stable Perovskite Photodetector Based on Vapor-Processed Micrometer-Scale CsPbBr3 Microplatelets. Chem. Mater. 2018, 30, 6744–6755. [Google Scholar] [CrossRef]
  4. Yusoff, A.; Kim, H.P.; Li, X.; Kim, J.; Jang, J.; Nazeeruddin, M.K. Field-Effect Transistors: Ambipolar Triple Cation Perovskite Field Effect Transistors and Inverters. Adv. Mater. 2017, 29, 1602940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. He, J.; Qiao, K.; Gao, L.; Song, H.; Hu, L.; Jiang, S.; Zhong, J.; Tang, J. Synergetic Effect of Silver Nanocrystals Applied in PbS Colloidal Quantum Dots for High-Performance Infrared Photodetectors. ACS Photonics 2014, 1, 936–943. [Google Scholar] [CrossRef]
  6. Yu, H.; Liu, S.; Baek, S.; Kim, D.Y.; Dong, C.; So, F. Solution-processed copper oxide interlayers for broadband PbS quantum-dot photodiodes. J. Mater. Chem. C 2016, 4, 11205–11211. [Google Scholar] [CrossRef]
  7. Gong, M.G.; Liu, Q.F.; Goul, R.; Ewing, D.; Casper, M.; Stramel, A.; Elliot, A.; Wu, J.Z. Printable Nanocomposite FeS2-PbS Nanocrystals/Graphene Heterojunction Photodetectors for Broadband Photodetection. ACS Appl. Mater. Interfaces 2017, 9, 27801–27808. [Google Scholar] [CrossRef]
  8. Jin, Z.; Hu, C.; Lan, Y.; Cao, Y.L.; Deng, H.; Yang, X.K.; Wang, J.; Song, H.S. Epitaxial Growth of PbS Nanocrystals from PbI2 Nanosheet Templates and Its Application in Fast Near infrared Photodetectors. Adv. Opt. Mater. 2020, 8, 2001319. [Google Scholar] [CrossRef]
  9. Schneider, D.S.; Bablich, A.; Lemme, M.C. Flexible hybrid graphene/a-Si:H multispectral photodetectors. Nanoscale 2017, 9, 287–290. [Google Scholar] [CrossRef] [Green Version]
  10. Benyahia, K.; Djeffal, F.; Ferhati, H.; Bendjerad, A.; Saidi, A. Self-powered photodetector with improved and broadband multispectral photoresponsivity based on ZnO-ZnS composite. J. Alloy. Compd. 2020, 859, 158242. [Google Scholar] [CrossRef]
  11. Sarkar, S.S.; Bera, S.; Hassan, S.; Sapra, R.K.; Ray, S.K. MoSe2-Cu2-xS/GaAs Heterostructure-Based Self-Biased Two Color- Band Photodetectors with High Detectivity. J. Phys. Chem. C 2021, 125, 10768–10776. [Google Scholar] [CrossRef]
  12. Song, J.Z.; Cui, Q.Z.; Li, J.H.; Xu, J.Y.; Zeng, H.B. Ultralarge All-Inorganic Perovskite Bulk Single Crystal for High-Performance Visible–Infrared Dual-Modal Photodetectors. Adv. Opt. Mater. 2017, 5, 1700157. [Google Scholar] [CrossRef]
  13. Fan, X.; Huang, Y.; Ren, X.; Duan, X.; Qi, W.; Cai, S.; Xia, Z. Hybrid integrated photodetector with flat-top steep-edge spectral response. Appl. Opt. 2012, 51, 5767. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, J.F.; Zens, T.; Hu, J.J.; Becla, P.; Kimerling, L.C.; Agarwal, A.M. Monolithically integrated, resonant-cavity-enhanced dual-band mid-infrared photodetector on silicon. Appl. Phys. Lett. 2012, 100, 211106. [Google Scholar] [CrossRef]
  15. Liu, X.Q.; Jiang, L.; Zou, X.M.; Xiao, X.H.; Guo, S.S.; Jiang, C.Z.; Liu, X.; Fan, Z.Y.; Hu, W.D.; Chen, X.S.; et al. Scalable integration of indium zinc oxide/photosensitive-nanowire composite thin-film transistors for transparent multicolor photodetectors array. Adv. Mater. 2014, 26, 2919–2924. [Google Scholar] [CrossRef]
  16. Said, A.J.; Recht, D.; Sullivan, J.T.; Warrender, J.M.; Buonassisi, T.; Persans, P.D.; Aziz, M.J. Extended infrared photoresponse and gain in chalcogen-supersaturated silicon photodiodes. Appl. Phys. Lett. 2011, 99, 073503. [Google Scholar] [CrossRef]
  17. Li, H.G.; Wu, G.; Shi, M.M.; Yang, L.G.; Chen, H.Z.; Wang, M. ZnO/poly(9,9-dihexylfluorene) based inorganic/organic hybrid ultraviolet photodetector. Appl. Phys. Lett. 2008, 93, 383. [Google Scholar] [CrossRef]
  18. Mukherjee, S.; Jana, S.; Sinha, T.K.; Das, S.; Ray, S.K. Infrared tunable, two colour-band photodetectors on flexible platforms using 0D/2D PbS–MoS2 hybrids. Nanoscale Adv. 2019, 1, 3279–3287. [Google Scholar] [CrossRef] [Green Version]
  19. Solanke, S.V.; Rathkanthiwar, S.; Kalra, A.; Mech, R.K.; Rangarajan, M.; Raghavan, S.; Nath, D.N. Integration of multi-layered materials with wide bandgap semiconductors for multi-spectral photodetectors: Case for MoS2/GaN and β-In2Se3/GaN. Semicond Sci. Technol. 2019, 34, 075020. [Google Scholar] [CrossRef]
  20. Baltog, I.; Baibarac, M.; Lefrant, S. Quantum well effect in bulk PbI2 crystals revealed by the anisotropy of photoluminescence and Raman spectra. J. Phys. Condens. Matter 2009, 21, 025507. [Google Scholar] [CrossRef]
  21. Sears, W.M.; Klein, M.L.; Morrison, J.A. Polytypism and the vibrational properties of PbI2. Phys. Rev. B 1979, 19, 2305–2313. [Google Scholar] [CrossRef] [Green Version]
  22. Frisenda, R.; Island, J.O.; Lado, J.L.; Giovanelli, E.; Gant, P.; Nagler, P.; Bange, S.; Lupton, J.M.; Schüller, C.; MolinaMendoza, A. Characterization of highly crystalline lead iodide nanosheets prepared by room-temperature solution processing. Nanotechnology 2017, 28, 455703. [Google Scholar] [CrossRef]
  23. Krauss, T.D.; Wise, F.W.; Tanner, D.B. Observation of coupled vibrational modes of a semiconductor nanocrystal. Phys. Rev. Lett. 1996, 76, 1376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Nanda, K.K.; Sarangi, S.N.; Sahu, S.N.; Deb, S.K.; Behera, S.N. Raman spectroscopy of PbS nanocrystalline semiconductors. Phys. B Condens. Matter 1999, 262, 31–39. [Google Scholar] [CrossRef]
  25. Tauc, J. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Status Solidi 1966, 3, 37–46. [Google Scholar]
  26. Mohd, S.; Mohd, T.K.; Ganesh, V.; Yahia, I.S.; Bakhtiar, U.H.; Abdullah, A.; Shankaragouda, P.P.; Maidur, S.R.; Alfaify, S. Influence of Dy doping on key linear, nonlinear and optical limiting characteristics of SnO2 films for optoelectronic and laser applications. Opt. Laser Technol. 2018, 108, 609–618. [Google Scholar]
  27. Chen, R.T.; Robinson, D. Electro-optic and all-optical phase modulator on an indium tin oxide single-mode waveguide. Appl. Phys. Lett. 1992, 60, 1541–1543. [Google Scholar] [CrossRef] [Green Version]
  28. Gong, X.; Tong, M.; Xia, Y.; Cai, W.; Moon, J.S.; Cao, Y.; Yu, G.; Shieh, C.L.; Nilsson, B.; Heeger, A.J. High-Detectivity Polymer Photodetectors with Spectral Response from 300 nm to 1450 nm. Science 2009, 325, 1665–1667. [Google Scholar] [CrossRef]
  29. Mcdonald, S.A.; Konstantatos, G.; Zhang, S.; Cyr, P.W.; Klem, E.J.D.; Levin, L.; Sargent, E.H. Solution-processed PbS quantum dot infrared photodetectors and photovoltaics. Nat. Mater. 2005, 4, 138–142. [Google Scholar] [CrossRef]
  30. Greenham, N.C.; Peng, X.; Alivisatos, A.P. Charge separation and transport in conjugated-polymer/semiconductor-nanocrystal composites studied by photoluminescence quenching and photoconductivity. Phys. Rev. B 1996, 54, 17628. [Google Scholar] [CrossRef]
  31. 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]
Figure 1. Schematic of preparing the PbS/PbI2 heterojunction.
Figure 1. Schematic of preparing the PbS/PbI2 heterojunction.
Nanomaterials 12 00681 g001
Figure 2. SEM patterns of the PbI2 thin film and PbS/PbI2 heterojunction: (a) top view of the PbI2 thin film; (b) top view of the PbS/PbI2 heterojunction; and (c) cross-section of the PbS/PbI2 heterojunction.
Figure 2. SEM patterns of the PbI2 thin film and PbS/PbI2 heterojunction: (a) top view of the PbI2 thin film; (b) top view of the PbS/PbI2 heterojunction; and (c) cross-section of the PbS/PbI2 heterojunction.
Nanomaterials 12 00681 g002aNanomaterials 12 00681 g002b
Figure 3. XRD and Raman curves of PbI2 thin film and PbS/PbI2 heterojunction: (a) XRD curves; (b) Raman curves and the inset is the comparison of Raman spectra between PbI2 thin film and PbS/PbI2 heterojunction within the range of 50–300 cm−1.
Figure 3. XRD and Raman curves of PbI2 thin film and PbS/PbI2 heterojunction: (a) XRD curves; (b) Raman curves and the inset is the comparison of Raman spectra between PbI2 thin film and PbS/PbI2 heterojunction within the range of 50–300 cm−1.
Nanomaterials 12 00681 g003
Figure 4. (a) Vis–NIR absorption spectra of the PbI2 thin film and PbS/PbI2 heterojunction, and the insets are the band gap fitting curves of the PbI2 thin film and PbS/PbI2 heterojunction; and (b) normalized responsivity of the PbS/PbI2 heterojunction.
Figure 4. (a) Vis–NIR absorption spectra of the PbI2 thin film and PbS/PbI2 heterojunction, and the insets are the band gap fitting curves of the PbI2 thin film and PbS/PbI2 heterojunction; and (b) normalized responsivity of the PbS/PbI2 heterojunction.
Nanomaterials 12 00681 g004
Figure 5. (a) Schematic of the PbS/PbI2 heterojunction device structure; and (b) I−V curves of the PbS/PbI2 heterojunction device under 450 and 980 nm incident light.
Figure 5. (a) Schematic of the PbS/PbI2 heterojunction device structure; and (b) I−V curves of the PbS/PbI2 heterojunction device under 450 and 980 nm incident light.
Nanomaterials 12 00681 g005
Figure 6. Dynamic time-resolved photoresponse and response/recovery plots of the PbI2 thin film device and the PbS/PbI2 heterojunction device under periodical illumination by monochromatic light source at different wavelengths and with 30 V bias voltage: (a) dynamic time-resolved photoresponse under 450 nm illumination; (b) dynamic time-resolved photoresponse under 980 nm illumination; (c) response/recovery plots under 450 nm illumination; and (d) response/recovery plot under 980 nm illumination.
Figure 6. Dynamic time-resolved photoresponse and response/recovery plots of the PbI2 thin film device and the PbS/PbI2 heterojunction device under periodical illumination by monochromatic light source at different wavelengths and with 30 V bias voltage: (a) dynamic time-resolved photoresponse under 450 nm illumination; (b) dynamic time-resolved photoresponse under 980 nm illumination; (c) response/recovery plots under 450 nm illumination; and (d) response/recovery plot under 980 nm illumination.
Nanomaterials 12 00681 g006
Figure 7. Responsivity and specific detectivity of the PbI2 thin film device and the PbS/PbI2 heterojunction device under illumination at different wavelengths and the power density of incident light at 30 V bias voltage: (a) relationship between responsivity and power density under 450 nm illumination; (b) relationship between specific detectivity and power density under 450 nm illumination; (c) relationship between responsivity and power density under 980 nm illumination; and (d) relationship between specific detectivity and power density under 980 nm illumination.
Figure 7. Responsivity and specific detectivity of the PbI2 thin film device and the PbS/PbI2 heterojunction device under illumination at different wavelengths and the power density of incident light at 30 V bias voltage: (a) relationship between responsivity and power density under 450 nm illumination; (b) relationship between specific detectivity and power density under 450 nm illumination; (c) relationship between responsivity and power density under 980 nm illumination; and (d) relationship between specific detectivity and power density under 980 nm illumination.
Nanomaterials 12 00681 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, S.; Han, J.; Zhang, J.; Kong, Y.; Liu, H. In Situ Growth of PbS/PbI2 Heterojunction and Its Photoelectric Properties. Nanomaterials 2022, 12, 681. https://doi.org/10.3390/nano12040681

AMA Style

Yang S, Han J, Zhang J, Kong Y, Liu H. In Situ Growth of PbS/PbI2 Heterojunction and Its Photoelectric Properties. Nanomaterials. 2022; 12(4):681. https://doi.org/10.3390/nano12040681

Chicago/Turabian Style

Yang, Shangxun, Jun Han, Jin Zhang, Yingxiu Kong, and Huan Liu. 2022. "In Situ Growth of PbS/PbI2 Heterojunction and Its Photoelectric Properties" Nanomaterials 12, no. 4: 681. https://doi.org/10.3390/nano12040681

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