Next Article in Journal
A Study on the Color Prediction of Ancient Chinese Architecture Paintings Based on a Digital Color Camera and the Color Design System
Previous Article in Journal
Plant Management and Soil Improvement in Specialty Crop Production
Previous Article in Special Issue
Highly Stable CsPbI3 Perovskite Quantum Dots Enabled by Single SiO2 Coating toward Down-Conversion Light-Emitting Diodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mixed-Dimensional Heterostructure Photodetector Based on Bi2O2Se Nanosheets and PbS Quantum Dots

School of Physics and Electronic Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2024, 14(13), 5914; https://doi.org/10.3390/app14135914 (registering DOI)
Submission received: 4 June 2024 / Revised: 29 June 2024 / Accepted: 4 July 2024 / Published: 6 July 2024
(This article belongs to the Special Issue Challenges and Future Trends of Low-Dimensional Materials)

Abstract

:
Due to their exceptional electronic and optical properties, two-dimensional materials have emerged as one of the most promising candidates for future optoelectronic detection. However, optoelectronic detectors based on two-dimensional transition metal materials still face challenges due to factors such as limited absorption coefficients and carrier recombination. In this study, we combine two-dimensional Bi2O2Se with PbS quantum dots to prepare a hybrid heterojunction, effectively broadening the detection range and significantly enhancing the photoresponse rate. The hybrid photodetector exhibited a remarkable photoresponsivity of 14.89 A/W at 450 nm and demonstrated broadband detection capabilities from visible (405 nm) to near-infrared (1350 nm) light illumination. Moreover, the hybrid device showed reduced photocurrent response and recovery times, highlighting its improved performance over bare Bi2O2Se photodetectors. This work underscores the potential of hybrid heterojunctions for enhancing optoelectronic detection capabilities, paving the way for advanced applications in various fields.

1. Introduction

Due to its crucial applications in fields like video imaging, biomedical imaging, network security, optoelectronic communication, and motion detection, there is a pressing need for optoelectronic devices with high response speed, broad-wavelength detection, and versatile integration [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. Conventional detectors are unable to meet the escalating demands in various domains, including high-frequency communication, information security, and cutting-edge biomedical imaging. Two-dimensional (2D) materials, as layered low-dimensional substances, enable efficient optoelectronic detection through interlayer van der Waals interactions. Moreover, traditional optoelectronic detectors typically require low-temperature operation and complex fabrication processes, but using 2D materials simplifies this process. Two-dimensional semiconductor materials have unique advantages in the field of optoelectronic detection: (I) The preparation methods of two-dimensional semiconductor materials are relatively simple and can be prepared by methods such as chemical vapor deposition (CVD) [15] or mechanical exfoliation [2,16]. (II) The stacking between layers of two-dimensional layered semiconductor materials is not limited by lattice matching [17]. (III) The bandgap of most two-dimensional layered semiconductor materials can be adjusted by controlling the number of material layers [18]. Different types of two-dimensional semimetal materials have their unique advantages in the field of optoelectronic detection. Graphene, due to its extremely narrow bandgap, exhibits a wide optical response range, but this also results in a significant dark current [19,20,21]. Transition metal dichalcogenides (TMDs) come in various types, and their bandgaps can be adjusted based on the number of layers, yet most materials primarily respond to light in the ultraviolet and visible regions [22,23,24,25,26]. Black phosphorus (BP) demonstrates excellent optoelectronic properties, but it is prone to volatilization in air [27]. Hexagonal boron nitride (hBN) can be used for ultraviolet photodetection due to its wide bandgap and is often employed as a dielectric material. However, the preparation process for hBN is quite intricate [28]. To enhance the optoelectronic performance of two-dimensional transistors, researchers have explored methods such as etching and doping to modify the structure of two-dimensional materials. Among these, two-dimensional transition metal compounds, particularly Bi2O2Se, stand out due to their exceptional optoelectronic properties [29].
Bi2O2Se possesses excellent electron mobility and environmental stability. Nevertheless, Bi2O2Se photodetectors face challenges like a limited response range and higher dark current [30]. Constructing heterostructures, such as 2D/2D and 0D/2D junctions, offers a promising approach to enhance Bi2O2Se photodetectors, improving their on/off ratio and reducing dark current [31,32,33,34,35,36,37,38,39,40]. Heterojunctions between Bi2O2Se and other 2D or zero-dimensional (0D) materials hold potential to extend the detection range and boost photoresponsivity [30]. Due to the bandgap limitation, Bi2O2Se detection performance in the infrared spectrum exhibits a noticeable decline. Quantum dots (QDs), on the other hand, can achieve effective response in the infrared spectrum by controlling magnetization. There are many types of 0D narrow-bandgap colloidal quantum dots, whose bandgap size is related to the size of the QDs, and they possess a high extinction coefficient in the near-infrared band but relatively low electron mobility. Heterojunctions formed by QDs and 2D materials can simultaneously leverage the advantages of both materials: the high-mobility 2D materials serve as carrier transport channels, while the QDs provide the device with broadband light absorption capabilities. Additionally, the band bending formed between QDs and 2D materials due to the heterojunction structure can facilitate the efficient separation of photogenerated charges, reducing recombination probability and thus enhancing the photodetection performance. Therefore, a combination of the two can effectively broaden the detection range and improve the responsivity of photodetection.
In this work, we present the construction of a type-II Bi2O2Se/PbS QDs hybrid heterojunction by using chemical vapor deposition and spin-coating. Comparative analysis with the 2D Bi2O2Se photodetector reveals a substantial enhancement in photoresponsivity for the hybrid heterojunction, achieving a remarkable 14.89 A/W at 450 nm. Furthermore, the Bi2O2Se-PbS QD-based photodetector demonstrates broadband photoresponsivity spanning from visible (405 nm) to near-infrared (1350 nm) light illumination. Notably, compared to the bare Bi2O2Se photodetector, the hybrid photodetector exhibits significantly reduced photocurrent response and recovery times.

2. Materials and Methods

2.1. Synthesis of Bi2O2Se Nanosheets

Few-layer Bi2O2Se was prepared via the chemical vapor deposition (CVD) method. The furnace tube for dual-temperature zone CVD has a diameter of 10 cm and a length of 25 cm. Bi2O3 powder and Bi2Se3 powder were placed at the hot center of the first temperature zone and 5 cm upstream from it, respectively, with a molar mass ratio of 2:1 between Bi2O3 and Bi2Se3. Mica substrates were placed at the hot center of the second temperature zone. First, we flushed the quartz tube with Ar gas three times to expel the air from the tube. The temperature of the first heating zone was set to increase from room temperature to 580 °C over 35 min and then maintained at 580 °C for 30 min. The temperature of the second heating zone was set to increase from room temperature to 680–730 °C over 35 min and then maintained at 680–730 °C for 30 min. The entire CVD process for preparing the two-dimensional Bi2O2Se film was conducted under Ar conditions, with an argon flow rate of 120 sccm. The pressure inside the quartz tube was maintained at 250 torr by controlling the vacuum pump.

2.2. Characterization

To evaluate the prepared Bi2O2Se nanosheets, PbS QDs, and hybrid PbS/Bi2O2Se heterostructure, atomic force microscopy (AFM) (Innova, Bruker, Billerica, MA, USA), laser micro-Raman spectroscopy (DXR, Thermo Fisher, Waltham, MA, USA), and multi-functional X-ray photoelectron spectroscopy (XPS) (ESCALAB QXi, Thermo) were employed. To obtain the energy band alignment information between PbS QDs and Bi2O2Se, ultraviolet photoelectron spectroscopy (UPS) (ESCALAB 250XI, Thermo) measurements were conducted. To investigate the photodetection performance of Bi2O2Se photodetectors and hybrid photodetectors, we used a photodetection testing platform to test the devices. The photodetection testing platform includes a xenon lamp light source (GLORIA X500A, Zolix, Beijing, China), a probe station (DCH-E2, Metatest Corporation, Nanjing, China), an optical system (Omni-λ-i, Zolix), and a source meter (MODELS 2614B, Keythley, Cleveland, OH, USA).

2.3. Preparation of the Hybrid Photodetector

The Bi2O2Se and hybrid Bi2O2Se/PbS photodetectors were fabricated using electron beam evaporation, with Au (50 nm) as the contact electrodes. The electrode preparation requires the design of a mask for the sample material. The distance between electrodes is approximately 30 μm. Prior to electrode fabrication, alignment of the electrode pattern on the mask with the Bi2O2Se sample material is achieved using an optical microscope, followed by securing the mask to the substrate using high-temperature resistant tape. Finally, the assembly is placed into the electron beam evaporation equipment for depositing a 50 nm thick layer of gold.
To sensitize the device with PbS, the bare Bi2O2Se device was spin-coated with a layer of PbS QDs. A 25 μL solution of octane PbS–OA was spin-coated onto the entire substrate at 2500 rpm for 30 s. This process was repeated twice, and the sample was then annealed at 100 °C for 15 min. All electrical and optical measurements were performed using a probe station with a source meter. The devices were illuminated with a xenon lamp light source, and power calibration was conducted using a standard silicon detector.

3. Results and Discussion

Bi2O2Se forms a tetragonal structure at room temperature. With a spatial group symmetry of I4/mmm (a = b = 3.88 Å, c = 12.16 Å), eight Bi protones are located at the top point of the cube, as illustrated in Figure 1a. The [Bi2O2]n2n+ cation layer and [Se]n2n− anionic layer are alternately stacked along the C-axis by weak electrostatic interaction, with a layer thickness of 0.608 nm [29,39]. Unlike the common van der Waals layered materials, Bi2O2Se is ion-layered, but it still exhibits the characteristics of typical 2D materials. The experiment utilized chemical vapor deposition (CVD) to synthesize two-dimensional Bi2O2Se, with Bi2O3 and Bi2Se3 powders as precursors. Since Bi2O2Se is the only stable phase on the eutectic line of Bi2Se3 and Bi2O3, the reaction will not result in the formation of other impurities or the occurrence of side reactions. Bi2O3 powder is adsorbed and diffused onto the substrate to react with Bi2Se3 to form Bi2O2Se; this mechanism can provide an improved local environment for high-quality growth. Typically, CVD preparation of 2D materials uses silicon dioxide as the substrate. The interlayer bonding force in Bi2O2Se is electrostatic. Using silicon dioxide as the substrate would promote the vertical growth of Bi2O2Se, preventing the formation of few-layer nanosheets. The choice of freshly cleaved fluorophlogopite mica [KMg3(AlSi3O10)F2] as the substrate for preparing 2D Bi2O2Se is due to its flat surface, surface inertness, and high thermal stability. Additionally, there is a strong electrostatic interaction between the K+ layers in fluorophlogopite mica and the Se2− layers in Bi2O2Se, which promotes the lateral growth of Bi2O2Se [40,41]. A diagram of the CVD system processing the sample here with a double-temperature zone is shown in Figure 1b. Bi2Se3 powder and Bi2O3 powder are used as source materials and placed in a hot center for evaporation; the distance between the two quartz boats is 5 cm. Freshly cleaved fluorophlogopite mica [KMg3(AlSi3O10)F2] is placed in the hot center of the second warm zone, 10–13 cm away from the center of the first warm zone. As illustrated in Figure S1, when the substrate temperature is 680 °C, the precursor absorption rate on the substrate is relatively fast, allowing for the formation of larger Bi2O2Se nanosheets [29]. As the temperature increases, the precursor absorption rate decreases, resulting in smaller Bi2O2Se nanosheets. Further increases in temperature can lead to nanosheet decomposition and poorer crystallinity. In the experiment, electrodes were prepared using electron beam evaporation. Given that the mask electrode spacing is 30 μm, we chose to control the substrate temperature between 680 and 690 °C to obtain large-area Bi2O2Se nanosheets.
The two-dimensional Bi2O2Se nanosheets selected for heterojunction fabrication are shown in Figure 1c. The morphology of individual Bi2O2Se nanosheets is characterized by regular square shapes, with a side length of approximately 35 μm. The thickness of the nanosheets, as measured by atomic force microscopy (AFM), is approximately 6 nm, as shown in Figure 1d. The Raman spectra of two-dimensional Bi2O2Se is shown in Figure 1e. Due to the A1g vibration modes of bi, the spectra show a single characteristic peak with a center of ≈159 cm−1, indicating good uniformity of the 2D Bi2O2Se [35,36,37]. To further confirm that the synthesized material is Bi2O2Se, it was also characterized by energy-dispersive X-ray photoelectron spectroscopy (XPS). As described in Figure 1f–h, the peaks of Bi 4f7/2 and Bi 4f5/2, which are located at 158.8 eV and 164.2 eV, respectively. By performing Lorentzian fitting on the broad peak near 54 eV using XPS software (Avantage V5.9931), it is possible to resolve the peaks at 53.2 eV and 54.2 eV, which correspond to the Se 3d5/2 and 3d3/2 peaks, respectively. The O 1s spectrum contains two peaks that are located at 532.1 eV and 529.8 eV; these two peaks correspond to adsorbed oxygen and lattice oxygen [37]. All bonding energies are consistent with the elements in Bi2O2Se.
Subsequently, a Au electrode with a thickness of 50 nm was fabricated via electron beam evaporation deposition for establishing an electrical connection with Bi2O2Se (Figure S3a). As shown in Figure 2a, the photocurrent of the Bi2O2Se photodetector increases with the enhancement of light intensity at 550 nm. Due to the bandgap of about 0.8 eV in few-layer Bi2O2Se, the device’s dark current reaches 10−6 A. As shown in Figure 2b, Bi2O2Se exhibits excellent photoresponse curves in the 405–1000 nm range, while almost no photocurrent is detected beyond 1000 nm. Figure 2c shows the responsivity and detectivity of the Bi2O2Se photodetector from 405 nm to 1000 nm. The experiment found that the maximum responsivity and detectivity of the few-layer Bi2O2Se photodetector are both at 600 nm, with a responsivity reaching 0.604 A/W and a detectivity reaching 7.61 × 109 Jones. Thus, the detection range of the few-layer Bi2O2Se photodetector prepared in the experiment is 405–1000 nm, mainly demonstrating its excellent detection capability in the visible light region. In the near-infrared region, however, the photodetection capability of Bi2O2Se is very weak, mainly due to the intrinsic indirect bandgap of Bi2O2Se [42,43], which limits its application in the field of photodetection. As shown in Figure 2d, the τrisedecay values of the Bi2O2Se photodetector are 0.63/1.27 s (Vds = 1 V).
However, by taking advantage of the high mobility of Bi2O2Se, an efficient broadband photodetector can be constructed by integrating it with QDs with high extinction coefficients [30]. As shown in Figure 3a, we further prepared photosensitive PbS QDs with a characteristic diameter of approximately 4.13 nm on the pre-fabricated Bi2O2Se transistor. These QDs possess a narrow bandgap of 1.1 electron volts, enabling spectral detection up to 1300 nm. In Figure 3b, the absorption of the quantum dots exhibits a first exciton absorption peak at approximately 1100 nm. The results are consistent with the predicted bandgap. The experiment involved spin-coating PbS QDs onto Bi2O2Se nanosheets, as illustrated in Figure 3c. Subsequently, the sample was annealed at 90 °C under Ar gas conditions for 15 min to induce solvent removal and atomic restructuring of the quantum dot surfaces, releasing surface partial stress defects, thereby enhancing the photodetection performance of the quantum dots. After two repeated spin-coating cycles, the optimal thickness of the PbS quantum dot layer was achieved, ensuring a dense distribution of the PbS quantum dots on the detector.
Next, we will investigate the performance of the PbS QDs/Bi2O2Se photodetector in detecting signals in the visible light and infrared spectrum. The schematic representation of this device is shown in Figure 4a. As shown in Figure 4b, introducing PbS QDs into Bi2O2Se was observed to increase the channel current by approximately 10 times under dark conditions, a phenomenon attributed to the electron transfer from PbS QDs to Bi2O2Se. Under excitation at 550 nm, the photocurrent (3648 nA) in the hybrid photodetector was nearly 13 times that of bare Bi2O2Se (280 nA). This significant enhancement in photocurrent is attributed to the charge transfer between PbS and Bi2O2Se, along with a higher electron mobility in Bi2O2Se compared to PbS QDs. This enhancement is due to the increased number of carriers within Bi2O2Se, leading to a reduction in the transit time within the transport channel and thereby enhancing the photodetection performance of the hybrid photodetector [44]. Figure 4c illustrates the variation of photocurrent at 550 nm under different light intensities.
Responsivity and detectivity are important metrics for assessing the performance of photodetectors, and they can be obtained based on the following equations [45,46,47]:
R = I p h P A
D * = A R 2 e I d
where P is the light power intensity, R is the responsivity, e is the electron charge, Id is dark current, and A is the effective device area. As the light intensity increases, the photocurrent of the hybrid photodetector also increases continuously, and the hybrid heterojunction photodetector also exhibits a fast response speed. However, we observed a gradual decrease in the responsivity and detectivity of the hybrid photodetector with increasing light intensity (Figure 4d). This phenomenon can be attributed to two reasons: (I) When the light intensity reaches a certain level, the responsivity and detectivity of the photodetector itself have already reached their maximum values, and further increases in light intensity do not correspondingly enhance the responsivity and detectivity. (II) At high light intensities, photogenerated carriers in the photodetector undergo recombination, where carriers excited by light recombine, leading to a decrease in the photodetection performance of the photodetector.
As the key figures of merit of photodetectors, the response speed (τrise: photocurrent increases from 10 to 90% of its final value)/decay time (τdecay: photocurrent falls from 90 to 10% of its beginning value) are also characterized. As displayed in Figure 2d, τrisedecay values of 0.105/0.651 s (Vds = 1 V). Compared to the 0.63/1.27 s of bare Bi2O2Se (Figure 2d), the response speed and decay time of the hybrid heterojunction have been significantly improved. Figure 4f shows the photoresponse spectra of the compared devices. Compared to bare Bi2O2Se, the hybrid photodetector exhibited a 20-fold improvement in responsiveness below 1000 nm due to the detection range of PbS quantum dots covering 405–1350 nm (Figure S4). At λ > 1000 nm, the responsiveness of bare Bi2O2Se is nearly undetectable, whereas the PbS/Bi2O2Se hybrid photodetector exhibits significant photoresponse.
The charge transfers between Bi2O2Se and PbS QDs is determined by the band alignment at the interface. To obtain the band information of the prepared PbS QDs and Bi2O2Se, ultraviolet photoelectron spectroscopy (UPS) measurements were conducted. In Figure 4g, valence band spectra and the second cutoff energy of Bi2O2Se and PbS are presented. The work functions (W) of PbS and Bi2O2Se can be calculated from the second cutoff energy (Ecut) using W= hν − Ecut, where hν = 21.22 eV represents the photon energy. The work functions of PbS and Bi2O2Se were found to be 4.2 eV and 4.35 eV, respectively. Additionally, the valence band positions of PbS and Bi2O2Se were determined to be 0.49 eV and 0.51 eV below the Fermi level (EF) of each material based on the valence band spectra, thus establishing the valence band positions of PbS quantum dots and Bi2O2Se. Therefore, based on the bandgaps of PbS (1.1 eV) and Bi2O2Se (0.8 eV), along with the estimated work functions and valence band positions, the energy band diagram shown in Figure 4h was constructed. PbS and Bi2O2Se form a type-II hybrid heterojunction, where electrons are transferred from PbS to Bi2O2Se due to the lower work function of PbS, forming an internal electric field that enhances the photodetection performance of Bi2O2Se.
Figure 5a illustrates the charge flow diagram of the PbS/Bi2O2Se hybrid photodetector under a 1 V bias. When the device is illuminated, both PbS quantum dots and Bi2O2Se generate photogenerated carriers. Due to the type-II heterojunction formed between PbS quantum dots and Bi2O2Se, photogenerated electrons in the PbS quantum dots transfer to the Bi2O2Se. Bi2O2Se acts as the current transport layer throughout the device. The transfer of electrons from the PbS quantum dots enhances the electron mobility in Bi2O2Se, shortening the carrier transport time (τtransit), thereby increasing the photoconductive gain of the entire device [44]. Consequently, we can observe that the responsivity and specific detectivity of the PbS/Bi2O2Se hybrid photodetector have improved by an order of magnitude compared to the standalone Bi2O2Se photodetector.
The PbS QDs and two-dimensional Bi2O2Se form a type-II heterojunction, resulting in the creation of an internal electric field within the device. This field can promote the separation of carriers, improving the device’s photoresponse speed. However, experimental results reveal that the response time of the hybrid heterojunction device is slower compared to devices constructed with Bi2O2Se and other two-dimensional materials in type-II heterojunctions. This slower response time is attributed to Förster resonance in the 0D/2D heterojunction. Förster resonance energy transfer (FRET) is a highly distance-dependent non-radiative energy transfer mechanism [48,49]; the energy transfer diagram is shown in Figure 5b. In the PbS QDs/Bi2O2Se hybrid photodetector, when PbS quantum dots, acting as energy donors, are photoexcited to an excited state, they can transfer their energy non-radiatively to Bi2O2Se through dipole–dipole interactions. This causes electrons in the valence band of Bi2O2Se to transition to the conduction band. In the PbS QDs/Bi2O2Se hybrid photodetector, the primary factor enhancing photoresponse time is the capture of photogenerated carriers by defect states in Bi2O2Se rather than the separation of photogenerated carriers at the interface. Therefore, while the photoresponse speed of the hybrid heterojunction photodetector is improved compared to a standalone Bi2O2Se photodetector, it does not match the improvement observed in other type-II heterojunctions.

4. Conclusions

In summary, we successfully synthesized 2D layered Bi2O2Se samples using low-pressure chemical vapor deposition and demonstrated a broadband and high-performance infrared photodetector based on the hybrid structure of high-mobility 2D Bi2O2Se and PbS QDs. The fast photoresponse (rise/fall time) of the hybrid photodetector reached 0.15/0.651 s (compared to 0.63/1.27 s for bare Bi2O2Se), attributed to the rapid photogenerated carrier separation at the formed type-II interface. Compared to few-layer Bi2O2Se photodetectors, the hybrid photodetector expanded its photoresponse range to near 1350 nm. The responsivity (R) of the hybrid photodetector within 1000 nm was nearly 20 times higher than that of few-layer Bi2O2Se (reaching 14.89 A/W at 450 nm). The construction of the PbS quantum dots and Bi2O2Se heterojunction further validates the feasibility of the theoretical approach of combining photosensitive materials with channel materials. By utilizing the high light response of the photosensitive material, incident photons are efficiently converted into photogenerated electrons, and the high-speed charge transfer of the channel material allows photogenerated carriers to cycle multiple times within their lifetime, achieving high gain. This device structure holds broad prospects for future applications in the field of photodetection, offering wide spectral response range, high responsivity, miniaturization, and portability. To further enhance the optoelectronic performance of the hybrid heterojunction, future research should focus on the intrinsic properties of 2D materials and the surface defects of QDs. Improving the preparation and surface treatment processes of QDs and 2D materials will broaden the spectral response range and enhance the device’s performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app14135914/s1, Figure S1. OM images of Bi2O2Se growth at different substrate temperatures. Figure S2. I–V curves of the bare Bi2O2Se under different wavelengths. Figure S3. Bare Bi2O2Se device and its optoelectronic detection performance. Figure S4. Optoelectronic performance of the bare PbS device.

Author Contributions

Conceptualization, B.Z.; material preparation, B.Z.; data curation, B.Z. and W.L.; writing—original draft preparation, B.Z.; writing—review and editing, Z.W., Y.X. and Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 12204198).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article and Supplementary Materials, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tyagi, D.; Wang, H.; Huang, W.; Hu, L.; Tang, Y.; Guo, Z.; Ouyang, Z.; Zhang, H. Recent advances in two-dimensional-material-based sensing technology toward health and environmental monitoring applications. Nanoscale 2020, 12, 3535–3559. [Google Scholar] [CrossRef] [PubMed]
  2. Dong, T.; Simões, J.; Yang, Z. Flexible photodetector based on 2D materials: Processing, architectures, and applications. Adv. Mater. Interfaces 2020, 7, 1901657. [Google Scholar] [CrossRef]
  3. Konstantatos, G. Current status and technological prospect of photodetectors based on two-dimensional materials. Nat. Commun. 2018, 9, 5266. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, M.; Ye, P.; Wang, M.; Wang, L.; Wu, C.; Xu, J.; Chen, Y. 2D/2D Bi-MOF-derived BiOCl/MoS2 nanosheets S-scheme heterojunction for effective photocatalytic degradation. J. Environ. Chem. Eng. 2022, 10, 108436. [Google Scholar] [CrossRef]
  5. An, C.; Nie, F.; Zhang, R.; Ma, X.; Wu, D.; Sun, Y.; Hu, X.; Sun, D.; Pan, L.; Liu, J. Two-dimensional material-enhanced flexible and self-healable photodetector for large-area Photodetection. Adv. Funct. 2021, 31, 2100136. [Google Scholar] [CrossRef]
  6. Cheng, Z.; Zhao, T.; Zeng, H. 2D material-based photodetectors for infrared imaging. Small Sci. 2022, 2, 2100051. [Google Scholar] [CrossRef]
  7. Lim, Y.R.; Han, J.K.; Kim, S.K.; Lee, Y.B.; Yoon, Y.; Kim, S.J.; Min, B.K.; Kim, Y.; Jeon, C.; Won, S.; et al. Roll-to-Roll Production of Layer-Controlled Molybdenum Disulfide: A Platform for 2D Semiconductor-Based Industrial Applications. Adv. Mater. 2018, 30, 1705270. [Google Scholar] [CrossRef] [PubMed]
  8. An, J.; Wang, B.; Shu, C.; Wu, W.; Sun, B.; Zhang, Z.; Li, D.; Li, S. Research development of 2D materials based photodetectors towards mid-infrared regime. Nano Sel. 2021, 2, 527–540. [Google Scholar] [CrossRef]
  9. Wang, B.; Zhong, S.P.; Bin Zhang, Z.; Zheng, Z.Q.; Zhang, Y.P.; Zhang, H. Broadband photodetectors based on 2D group IVA metal chalcogenides semiconductors. Appl. Mater. Today 2019, 15, 115–138. [Google Scholar] [CrossRef]
  10. Rehman, A.; Park, S.-J. State of the art two-dimensional materials-based photodetectors: Prospects, challenges and future outlook. J. Ind. Eng. Chem. 2020, 89, 28–46. [Google Scholar] [CrossRef]
  11. Xu, H.; Cao, C.; Shui, X.; Gu, J.; Sun, Y.; Ding, L.; Lin, Y.; Shi, W.; Wei, B. Discrimination and control of the exciton-recombination region of thermal-stressed blue organic light-emitting diodes. Phys. Chem. Chem. Phys. 2023, 25, 2742–2746. [Google Scholar] [CrossRef]
  12. Tang, Z.; Lü, Z.; Zheng, Y.; Wang, J. Management of exciton recombination zone and energy loss for 4CzTPN-based organic light-emitting diodes via engineering hosts. Phys. B Condens. Matter 2022, 644, 414206. [Google Scholar] [CrossRef]
  13. Tasaki, S.; Nishimura, K.; Toyoshima, H.; Masuda, T.; Nakamura, M.; Nakano, Y.; Itoi, H.; Kambe, E.; Kawamura, Y.; Kuma, H. Realization of ultra-high-efficient fluorescent blue OLED. J. Soc. Inf. Disp. 2022, 30, 441–451. [Google Scholar] [CrossRef]
  14. Rozenman, G.G.; Peisakhov, A.; Zadok, N. Dispersion of organic exciton polaritons—A novel undergraduate experiment. Eur. J. Phys. 2022, 43, 035301. [Google Scholar] [CrossRef]
  15. Shen, P.-C.; Lin, Y.; Wang, H.; Park, J.-H.; Leong, W.S.; Lu, A.-Y.; Palacios, T.; Kong, J. CVD technology for 2-D materials. IEEE Trans. Electron Devices 2018, 65, 4040–4052. [Google Scholar] [CrossRef]
  16. Yi, M.; Shen, Z. A review on mechanical exfoliation for the scalable production of graphene. J. Mater. Chem. A 2015, 3, 11700–11715. [Google Scholar] [CrossRef]
  17. Liu, Y.; Huang, Y.; Duan, X. Van der Waals integration before and beyond two-dimensional materials. Nature 2019, 567, 323–333. [Google Scholar] [CrossRef]
  18. Chaves, A.; Azadani, J.G.; Alsalman, H.; Da Costa, D.R.; Frisenda, R.; Chaves, A.J.; Song, S.H.; Kim, Y.D.; He, D.; Zhou, J.; et al. Bandgap engineering of two-dimensional semiconductor materials. NPJ 2D Mater. Appl. 2020, 4, 29. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Liu, T.; Meng, B.; Li, X.; Liang, G.; Hu, X.; Wang, Q.J. Broadband high photoresponse from pure monolayer graphene photodetector. Nat. Commun. 2013, 4, 1811. [Google Scholar] [CrossRef]
  20. Liu, C.-H.; Chang, Y.-C.; Norris, T.B.; Zhong, Z. Graphene photodetectors with ultra-broadband and high responsivity at room temperature. Nat. Nanotechnol. 2014, 9, 273–278. [Google Scholar] [CrossRef]
  21. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed]
  22. Taffelli, A.; Dirè, S.; Quaranta, A.; Pancheri, L. MoS2 based photodetectors: A review. Sensors 2021, 21, 2758. [Google Scholar] [CrossRef] [PubMed]
  23. Ashtar, M.; Marwat, M.A.; Yang, Y.; Cao, D. Two-dimensional ZnO/ZrXY (X, Y = Br, Cl and F) van der Waals heterostructures as promising photocatalysts for high efficiency water splitting. Int. J. Hydrogen Energy 2023, 48, 32797–32805. [Google Scholar] [CrossRef]
  24. Idrees, M.; Amin, B.; Chen, Y.; Yan, X. Computation insights of MoS2-CrXY (X ≠ YS, Se, Te) van der waals heterostructure for Spintronic and photocatalytic water Splitting applications. Int. J. Hydrogen Energy 2024, 51, 1217–1228. [Google Scholar] [CrossRef]
  25. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef] [PubMed]
  26. Selamneni, V.; Sahatiya, P. Mixed dimensional transition metal dichalcogenides (TMDs) vdW heterostructure based photodetectors: A review. Microelectron. Eng. 2023, 269, 111926. [Google Scholar] [CrossRef]
  27. Lin, S.; Li, Y.; Qian, J.; Lau, S.P. Emerging opportunities for black phosphorus in energy applications. Mater. Today Energy 2019, 12, 1–25. [Google Scholar] [CrossRef]
  28. Basu, N.; Dutta, A.; Singh, R.; Bayazeed, M.; Parmar, A.S.; Som, T.; Lahiri, J. Substrate roughness and crystal orientation-controlled growth of ultra-thin BN films deposited on Cu foils. Appl. Phys. A 2022, 128, 392. [Google Scholar] [CrossRef]
  29. Wu, J.; Yuan, H.; Meng, M.; Chen, C.; Sun, Y.; Chen, Z.; Dang, W.; Tan, C.; Liu, Y.; Yin, J.; et al. High electron mobility and quantum oscillations in non-encapsulated ultrathin semiconducting Bi2O2Se. Nat. Nanotechnol. 2017, 12, 530–534. [Google Scholar] [CrossRef]
  30. Luo, P.; Zhuge, F.; Wang, F.; Lian, L.; Liu, K.; Zhang, J.; Zhai, T. PbSe quantum dots sensitized high-mobility Bi2O2Se nanosheets for high-performance and broadband photodetection beyond 2 μm. ACS Nano 2019, 13, 9028–9037. [Google Scholar] [CrossRef]
  31. Yang, T.; Li, X.; Wang, L.; Liu, Y.; Chen, K.; Yang, X.; Liao, L.; Dong, L.; Shan, C.X. Broadband photodetection of 2D Bi2O2Se–MoSe2 heterostructure. J. Mater. Sci. 2019, 54, 14742–14751. [Google Scholar] [CrossRef]
  32. Sun, L.; Xu, Y.; Yin, T.; Wan, R.; Ma, Y.; Su, J.; Zhang, Z.; Liu, N.; Li, L.; Zhai, T.; et al. Van der Waals heterostructure of Bi2O2Se/MoTe2 for high-performance multifunctional devices. Nano Energy 2024, 119, 109047. [Google Scholar] [CrossRef]
  33. Yu, M.; Fang, C.; Han, J.; Liu, W.; Gao, S.; Huang, K. Construction of Bi2O2Se/Bi2Se3 van der Waals heterostructures for self-powered and broadband photodetectors. ACS Appl. Mater. 2022, 14, 13507–13515. [Google Scholar] [CrossRef]
  34. Tao, L.; Li, S.; Yao, B.; Xia, M.; Gao, W.; Yang, Y.; Wang, X.; Huo, N. Raman anisotropy and polarization-sensitive photodetection in 2D Bi2O2Se–WSe2 heterostructure. ACS Omega 2021, 6, 34763–34770. [Google Scholar] [CrossRef]
  35. Sagar, R.U.R.; Khan, U.; Galluzzi, M.; Aslam, S.; Nairan, A.; Anwar, T.; Ahmad, W.; Zhang, M.; Liang, T. Transfer-free growth of Bi2O2Se on silicon dioxide via chemical vapor deposition. ACS Appl. Electron. 2020, 2, 2123–2131. [Google Scholar] [CrossRef]
  36. Chen, Y.; Ma, W.; Tan, C.; Luo, M.; Zhou, W.; Yao, N.; Wang, H.; Zhang, L.; Xu, T.; Tong, T.; et al. Broadband Bi2O2Se photodetectors from infrared to terahertz. Adv. Funct. 2021, 31, 2009554. [Google Scholar] [CrossRef]
  37. Khan, U.; Luo, Y.; Tang, L.; Teng, C.; Liu, J.; Liu, B.; Cheng, H.M. Controlled vapor–solid deposition of millimeter-size single crystal 2D Bi2O2Se for high-performance phototransistors. Adv. Funct. 2019, 29, 1807979. [Google Scholar] [CrossRef]
  38. Tong, T.; Chen, Y.; Qin, S.; Li, W.; Zhang, J.; Zhu, C.; Zhang, C.; Yuan, X.; Chen, X.; Nie, Z.; et al. Sensitive and ultrabroadband phototransistor based on two-dimensional Bi2O2Se nanosheets. Adv. Funct. 2019, 29, 1905806. [Google Scholar] [CrossRef]
  39. Wu, J.; Tan, C.; Tan, Z.; Liu, Y.; Yin, J.; Dang, W.; Wang, M.; Peng, H. Controlled synthesis of high-mobility atomically thin bismuth oxyselenide crystals. Nano Lett. 2017, 17, 3021–3026. [Google Scholar] [CrossRef]
  40. Fu, Q.; Zhu, C.; Zhao, X.; Wang, X.; Chaturvedi, A.; Zhu, C.; Wang, X.; Zeng, Q.; Zhou, J.; Liu, F.; et al. Ultrasensitive 2D Bi2O2Se phototransistors on silicon substrates. Adv. Mater. 2019, 31, 1804945. [Google Scholar] [CrossRef]
  41. Li, J.; Wang, Z.; Wen, Y.; Chu, J.; Yin, L.; Cheng, R.; Lei, L.; He, P.; Jiang, C.; Feng, L.; et al. High-performance near-infrared photodetector based on ultrathin Bi2O2Se nanosheets. Adv. Funct. 2018, 28, 1706437. [Google Scholar] [CrossRef]
  42. Wu, M.; Zeng, X.C. Bismuth oxychalcogenides: A new class of ferroelectric/ferroelastic materials with ultra high mobility. Nano Lett. 2017, 17, 6309–6314. [Google Scholar] [CrossRef]
  43. Hen, C.; Wang, M.; Wu, J.; Fu, H.; Yang, H.; Tian, Z.; Tu, T.; Peng, H.; Sun, Y.; Xu, X.; et al. Electronic structures and unusually robust bandgap in an ultrahigh-mobility layered oxide semiconductor, Bi2O2Se. Sci. Adv. 2018, 4, eaat8355. [Google Scholar]
  44. Buscema, M.; Island, J.O.; Groenendijk, D.J.; Blanter, S.I.; Steele, G.A.; van der Zant, H.S.J.; Castellanos-Gomez, A. Photocurrent generation with two-dimensional van der Waals semiconductors. Chem. Soc. Rev. 2015, 44, 3691–3718. [Google Scholar] [CrossRef]
  45. Zhang, Z.; Yang, J.; Zhang, K.; Chen, S.; Mei, F.; Shen, G. Anisotropic photoresponse of layered 2D SnS-based near infrared photodetectors. J. Mater. Chem. C 2017, 5, 11288–11293. [Google Scholar] [CrossRef]
  46. Correa-Baena, J.-P.; Luo, Y.; Brenner, T.M.; Snaider, J.; Sun, S.; Li, X.; Jensen, M.A.; Hartono, N.T.P.; Nienhaus, L.; Wieghold, S.; et al. Homogenized halides and alkali cation segregation in alloyed organic-inorganic perovskites. Science 2019, 363, 627–631. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, D.; Wang, Y.; Zeng, L.; Jia, C.; Wu, E.; Xu, T.; Shi, Z.; Tian, Y.; Li, X.; Tsang, Y.H. Design of 2D layered PtSe2 heterojunction for the high-performance, room-temperature, broadband, infrared photodetector. ACS Photonics 2018, 5, 3820–3827. [Google Scholar] [CrossRef]
  48. Li, M.; Chen, J.; Routh, P.K.; Zahl, P.; Nam, C.; Cotlet, M. Distinct optoelectronic signatures for charge transfer and energy transfer in quantum dot–MoS2 hybrid photodetectors revealed by photocurrent imaging microscopy. Adv. Funct. Mater. 2018, 28, 1707558. [Google Scholar] [CrossRef]
  49. Gough, J.J.; McEvoy, N.; O’Brien, M.; Bell, A.P.; McCloskey, D.; Boland, J.B.; Coleman, J.N.; Duesberg, G.S.; Bradley, A.L. Dependence of Photocurrent Enhancements in Quantum Dot (QD)-Sensitized MoS2 Devices on MoS2 Film Properties. Adv. Funct. Mater. 2018, 28, 1706149. [Google Scholar] [CrossRef]
Figure 1. Synthesis and characterization of 2D Bi2O2Se nanosheets. (a) Schematic diagram of the structure of Bi2O2Se. (b) Schematic illustration of the CVD growth of Bi2O2Se. (c) Optical microscopy image of the Bi2O2Se sample. (d) AFM image. (e) Raman spectrum of the Bi2O2Se sample. (fh) XPS spectra of the few-layer Bi2O2Se on mica.
Figure 1. Synthesis and characterization of 2D Bi2O2Se nanosheets. (a) Schematic diagram of the structure of Bi2O2Se. (b) Schematic illustration of the CVD growth of Bi2O2Se. (c) Optical microscopy image of the Bi2O2Se sample. (d) AFM image. (e) Raman spectrum of the Bi2O2Se sample. (fh) XPS spectra of the few-layer Bi2O2Se on mica.
Applsci 14 05914 g001
Figure 2. Optoelectronic detection performance of the bare Bi2O2Se device. (a) I–V curves of the bare Bi2O2Se under different light intensities. (b) A typical device response at 550–1000 nm. (c) Responsivity and detectivity of the bare Bi2O2Se detector at 405–1000 nm. (d) The dynamic response speed of the device under 550 nm at P = 2.33 mW/cm2.
Figure 2. Optoelectronic detection performance of the bare Bi2O2Se device. (a) I–V curves of the bare Bi2O2Se under different light intensities. (b) A typical device response at 550–1000 nm. (c) Responsivity and detectivity of the bare Bi2O2Se detector at 405–1000 nm. (d) The dynamic response speed of the device under 550 nm at P = 2.33 mW/cm2.
Applsci 14 05914 g002
Figure 3. Preparation and characterization of PbS QDs/Bi2O2Se. (a) TEM images of PbS QDs. (b) The absorption spectrum of the PbS quantum dot solution used shows the first excitation peak at around 1100 nm. (c) Schematic diagram of the preparation process for the PbS QDs/Bi2O2Se hybrid heterojunction.
Figure 3. Preparation and characterization of PbS QDs/Bi2O2Se. (a) TEM images of PbS QDs. (b) The absorption spectrum of the PbS quantum dot solution used shows the first excitation peak at around 1100 nm. (c) Schematic diagram of the preparation process for the PbS QDs/Bi2O2Se hybrid heterojunction.
Applsci 14 05914 g003
Figure 4. Photodetection performance and band alignment of the PbS/Bi2O2Se hybrid photodetector. (a) The tetragonal crystal structure of Bi2O2Se. (b) I–V curves of bare Bi2O2Se and PbS/Bi2O2Se hybrid photodetectors under dark conditions and illumination at a wavelength of 550 nm (2.33 mW/cm2). (c) A typical device response at 550 nm under varied light intensities from 2.33 mW/cm2 to 0.2 mW/cm2. (d) Responsivity and detectivity of the Bi2O2Se/PbS hybrid photodetector under illumination by a 550 nm laser at various power levels with a bias of 1 V. (e) The dynamic response speed of the device under 550 nm at P = 2.33 mW/cm2. (f) Response spectra of Bi2O2Se and Bi2O2Se/PbS hybrid photodetectors. (g) UPS test graphs of bare PbS and Bi2O2Se. (h) Drawing on the estimated valence band positions and work function values from UPS, a band diagram is depicted illustrating the carrier transfer situation between PbS and Bi2O2Se upon contact.
Figure 4. Photodetection performance and band alignment of the PbS/Bi2O2Se hybrid photodetector. (a) The tetragonal crystal structure of Bi2O2Se. (b) I–V curves of bare Bi2O2Se and PbS/Bi2O2Se hybrid photodetectors under dark conditions and illumination at a wavelength of 550 nm (2.33 mW/cm2). (c) A typical device response at 550 nm under varied light intensities from 2.33 mW/cm2 to 0.2 mW/cm2. (d) Responsivity and detectivity of the Bi2O2Se/PbS hybrid photodetector under illumination by a 550 nm laser at various power levels with a bias of 1 V. (e) The dynamic response speed of the device under 550 nm at P = 2.33 mW/cm2. (f) Response spectra of Bi2O2Se and Bi2O2Se/PbS hybrid photodetectors. (g) UPS test graphs of bare PbS and Bi2O2Se. (h) Drawing on the estimated valence band positions and work function values from UPS, a band diagram is depicted illustrating the carrier transfer situation between PbS and Bi2O2Se upon contact.
Applsci 14 05914 g004
Figure 5. Schematic of the mechanism for the PbS QDs/Bi2O2Se hybrid photodetector. (a) Photogenerated carrier flow diagram of the hybrid photodetector at Vds = 1 V. (b) Diagram of non-radiative energy transfer in PbS QDs and defect capture of carriers in Bi2O2Se.
Figure 5. Schematic of the mechanism for the PbS QDs/Bi2O2Se hybrid photodetector. (a) Photogenerated carrier flow diagram of the hybrid photodetector at Vds = 1 V. (b) Diagram of non-radiative energy transfer in PbS QDs and defect capture of carriers in Bi2O2Se.
Applsci 14 05914 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, B.; Liu, W.; Wang, Z.; Xie, Y.; Chen, Y. Mixed-Dimensional Heterostructure Photodetector Based on Bi2O2Se Nanosheets and PbS Quantum Dots. Appl. Sci. 2024, 14, 5914. https://doi.org/10.3390/app14135914

AMA Style

Zhang B, Liu W, Wang Z, Xie Y, Chen Y. Mixed-Dimensional Heterostructure Photodetector Based on Bi2O2Se Nanosheets and PbS Quantum Dots. Applied Sciences. 2024; 14(13):5914. https://doi.org/10.3390/app14135914

Chicago/Turabian Style

Zhang, Bin, Weijing Liu, Zhongxuan Wang, Yuee Xie, and Yuanping Chen. 2024. "Mixed-Dimensional Heterostructure Photodetector Based on Bi2O2Se Nanosheets and PbS Quantum Dots" Applied Sciences 14, no. 13: 5914. https://doi.org/10.3390/app14135914

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop