DOI:
10.1039/D5TC02340H
(Paper)
J. Mater. Chem. C, 2025, Advance Article
Self-powered MoTe2 homojunction photodetector with ultrafast response via h-BN encapsulation and doping regulation
Received
17th June 2025
, Accepted 8th August 2025
First published on 15th August 2025
Abstract
Doping regulation can create spatial variations in carrier type or density by introducing different types or concentrations of impurities, forming stable p–n junction structures. This method maintains lattice continuity and minimizes interfacial defects, offering excellent controllability and process compatibility for large-area fabrication and integration applications. A semi-encapsulated MoTe2 homojunction photodetector was fabricated by partially encapsulating with hexagonal boron nitride (h-BN) followed by high-temperature air annealing. The unencapsulated MoTe2 region undergoes p-type doping from environmental H2O and O2 adsorption, while the encapsulated region remains ambipolar due to environmental isolation. The resultant Fermi level gradient generated an intrinsic electric field at the junction, enhancing photovoltaic performance with improved responsivity (21 mA W−1) and accelerated response speed (54.6/55.2 μs). At a gate voltage (Vgs) of 40 V, the device exhibited a rectification ratio of 2.2 × 103 and a specific detectivity of 1 × 1011 Jones. This work demonstrates an environmentally modulated doping strategy for self-powered photodetectors with ultrafast temporal response and defect-minimized interfaces.
1. Introduction
Two-dimensional (2D) materials are being considered for future photodetector technologies due to their atomic-scale thickness, strong light–matter interactions, and tunable bandgap properties.1–4 van der Waals junctions made from 2D materials can create electric fields that help separate and transfer photogenerated charge carriers efficiently,5–9 leading to improved optoelectronic device performance. These junctions exhibit exceptional performance with ultra-low noise and low power consumption under zero drain bias,10–13 making them ideal for high-performance photodetectors, solar cells, and photovoltaic devices. This advancement in optoelectronic applications is promising for the development of novel technologies.
Lattice mismatch and transfer processes during fabrication can cause interfacial contamination and defects in heterojunctions, leading to degraded device performance.14–18 Homojunction devices offer simpler architectures and better fabrication processes compared to van der Waals heterojunctions. Doping modulation is an effective method for constructing homojunctions, allowing for spatial variation of carrier type or density by introducing impurities with different types or concentrations, forming stable p–n junctions.19–21 This approach preserves lattice continuity, minimizes interfacial defects, and provides excellent controllability and process compatibility for large-area fabrication and integration applications.
The three main doping strategies for 2D materials are surface charge transfer, substitutional doping, and intercalation doping.22 Substitutional doping can cause structural instability and defects,23–25 while intercalation doping may lead to lattice damage and increased complexity.26–29 In contrast, surface charge transfer is a simple and effective method, where charge exchange occurs between adsorbed atoms or molecules and the material surface.30,31 A number of studies have demonstrated the effectiveness of this approach. Nan et al. used defect engineering combined with oxygen chemical adsorption to enhance the photoluminescence of monolayer MoS2, and obtained the conclusion that oxygen-induced p-type doping and reduced nonradiative recombination at defect sites are responsible for the significant PL enhancement.32 Tongay et al. employed thermal annealing and molecular physisorption of O2 and H2O to modulate the photoluminescence of monolayer MoS2, and concluded that electron depletion caused by charge transfer stabilizes neutral excitons and significantly increases PL intensity.33 Similarly, Jürgen Ristein, through a comprehensive analysis of surface transfer doping techniques, explored their impact on the electronic properties of semiconductor materials and concluded that surface transfer doping is an effective strategy for precise control of electrical conductivity.34 Deshun Qu et al. demonstrated that controlled oxygen-pressure thermal annealing (RTA) at 250 °C enables continuous carrier-type modulation in MoTe2 from n-dominant ambipolar to unipolar p-type, where oxygen preferentially passivates tellurium vacancies via Mo–O bond formation, with XPS confirming chemical potential downshift – establishing a defect-engineering paradigm for 2D semiconductor polarity control.35
A partially encapsulated lateral homojunction photodetector was successfully fabricated in this study. The MoTe2 flake was partially encapsulated with h-BN and then annealed at high temperature in air. The exposed region of MoTe2 became p-type doped due to adsorption of H2O and O2 molecules from the air, while the h-BN-encapsulated region maintained its intrinsic ambipolar behavior. The doping-induced Fermi level difference between the two regions created an internal electric field across the junction interface, enhancing the dissociation of photo-induced carriers and resulting in a high responsivity of 21 mA W−1, a specific detectivity of 1 × 1011 Jones, and a fast response time of 54.6 μs for self-powered photodetection.
2. Results and discussion
2.1. Structural configuration of the MoTe2 homojunction
A MoTe2 homojunction photodetector was constructed as shown in Fig. 1a. Contacts were placed on both the exposed and protected parts of the MoTe2 layer, with the underlying SiO2/Si serving as the back gate. The fabrication process is illustrated in Fig. S1. A h-BN flake was dry-transferred onto a clean SiO2/Si substrate, followed by the dry transfer of a MoTe2 flake onto the h-BN, thus partially encapsulating the MoTe2 from the bottom. After spin-coating, photolithography was performed using a UV maskless lithography machine (TuoTuo Technology (Suzhou) Co., Ltd), followed by metal deposition and lift-off. A 50 nm-thick Au layer was deposited on both ends of the MoTe2 flake as metal electrodes. A top h-BN flake was dry-transferred onto the MoTe2 to complete the top encapsulation. The device was annealed in air at 250 °C for 15 minutes, causing the exposed MoTe2 region to adsorb H2O and O2 molecules, resulting in p-type doping of that region. Fig. S2 demonstrates that MoTe2 is most sensitive to the annealing temperature of 250 °C, and gradually transitions to p-type conductivity with increasing annealing time. Fig. 1b shows an optical microscopy image of the homojunction channel, with the blue area representing the bottom h-BN, the white region corresponding to MoTe2, and the yellow area representing the top h-BN. Fig. 1c presents AFM image of the device, providing its height profile and contour. The measured thicknesses of the bottom h-BN, MoTe2, and top h-BN are 17 nm, 40 nm, and 60 nm, respectively. The AFM topography clearly visualizes the interface region, highlighting the high quality of the semi-encapsulated MoTe2 homojunction. The Raman spectrum is a powerful tool for evaluating the lattice integrity and defect density of two-dimensional materials. In Fig. 1d, the three main vibrational modes (A1g at 174.63 cm−1, E2g1 at 237.87 cm−1, and B2g1 at 291.97 cm−1) from both the encapsulated and exposed regions show highly consistent peak positions, full widths at half maximum (FWHM), and relative intensities, which also match well with the characteristic features reported in ref. 36. If lattice damage had occurred during the annealing or surface adsorption doping processes, it would typically lead to peak shifts, broadening, or the emergence of defect-related peaks. However, no such changes were observed in our measurements. Therefore, the two comparable Raman spectra not only demonstrate the high crystalline quality of the MoTe2 film but also further confirm that the surface adsorption doping did not cause any structural damage to the crystal. The XPS characterization of MoTe2 before and after annealing is shown in Fig. S3. The Mo 3d5/2 peak is located at approximately 228.17 eV, the Te 3d5/2 peak at around 572.23 eV, and the O 1s peak remains stable near 532.9 eV. These values closely match those reported for pure 2H-phase MoTe2 in the literature. After annealing, the Mo 3d5/2 peak shifts toward higher binding energy accompanied by a broadening of the FWHM, indicating partial oxidation of Mo. Meanwhile, the Te 3d5/2 peak shifts from 572.2 eV to 572.85 eV with a similar increase in FWHM, suggesting partial oxidation of Te to TeO2.
 |
| Fig. 1 (a) Schematic diagram of MoTe2 homojunction photodetector. (b) Optical image. (c) Atomic force microscope (AFM) image. (d) Raman spectra of encapsulated MoTe2 region and exposed MoTe2 region (after annealing). | |
2.2. Electrical performance of the device
The output characteristics of the encapsulated and exposed regions of MoTe2 after annealing are shown in Fig. 2a. Encapsulated MoTe2 retains its original characteristics, forming ohmic contact with the gold electrodes. The current in the exposed MoTe2 region increases by more than 10 times after annealing. Fig. 2b shows the corresponding transfer characteristics. Encapsulated MoTe2 maintains its intrinsic bipolar behavior, while the exposed MoTe2 exhibits degenerate p-type doping with significantly increased hole concentration. By placing the electrodes on the encapsulated and exposed regions of MoTe2 respectively, a p–n homojunction was successfully fabricated. Fig. S4 presents the transfer and output characteristics, as well as the weak photoresponse of the device before annealing, confirming that no junction was formed prior to annealing and thus no self-powered photodetection capability was observed. Output characteristics under different gate voltages were measured to evaluate the performance of this homojunction, as shown in Fig. 2c. The homojunction exhibits pronounced rectifying behavior under positive gate bias, with a rectification ratio as high as 2.2 × 103 at Vgs = 40 V. This rectification behavior confirms the formation of a built-in electric field, critical for achieving excellent self-powered optoelectronic performance. Fig. S5 re-evaluates the photoresponse characteristics of the device after three months of storage, revealing no significant degradation in photocurrent under the same illumination intensity. Moreover, over 100 cycles of photoresponse testing were conducted, further demonstrating the excellent stability of the device.
 |
| Fig. 2 Electrical performance of the device. (a) Output characteristic curves of the encapsulated and exposed regions of MoTe2 after annealing. (b) Transfer characteristic curves of the encapsulated and exposed regions of MoTe2 after annealing. (c) Output characteristic curves of the device under different gate voltages. | |
2.3. Optoelectronic performance of the device
Fig. 3a shows the output characteristics of the homojunction device under 637 nm laser illumination at different power levels. The corresponding relationship between photocurrent and applied bias voltage is extracted and presented in Fig. 3b. The curves exhibit a clear shift from the origin, with a high photocurrent of up to 20 μA even at zero bias. In which, the photocurrent Iph is calculated using the following equation:here, Ilight represents the current measured under illumination, while Idark denotes the dark current. Fig. 3c displays the open-circuit voltage and short-circuit current under 637 nm laser illumination with varying power levels. Both the open-circuit voltage and short-circuit current increase with increasing optical power, further demonstrating the excellent self-powered optoelectronic performance of the device. The optical response curves of different power under 637 nm laser irradiation are shown in Fig. 3d, where the photocurrent increases rapidly with increasing light intensity. Moreover, photoresponsivity (R) serves as a crucial and essential parameter for assessing the photoresponse performance of photodetectors. R is calculated using the following equation: |
 | (2) |
 |
| Fig. 3 Self-powered optoelectronic performance of the device. (a) Photocurrent output characteristics under 637 nm laser illumination at different power levels. (b) Photocurrent as a function of bias voltage. (c) Open-circuit voltage and short-circuit current under 637 nm laser illumination. (d) Optical response curves of different power under 637 nm laser irradiation. (e) Photocurrent and photoresponsivity under 637 nm laser illumination. (f) The specific detectivity under 637 nm laser illumination. (g) Optical response curves under laser illumination at different wavelengths. (h) Fitted α values under laser illumination at different wavelengths. (i) The maximum photoresponsivity and the specific detectivity under different wavelengths of lasers irradiation. | |
Fig. 3e further presents the self-powered photocurrent and responsivity under 637 nm laser illumination (Vgs = 0 V, Vds = 0 V) at different optical power levels. The responsivity initially increases with rising optical power and then saturates, reaching a stable value of 16 mA W−1 at a power of 1.13 μW. Fig. 3f shows the variation of specific detectivity with optical power under illumination of different wavelengths (Vgs = 0 V, Vds = 0 V). The specific detectivity initially increases with increasing optical power and then decreases, reaching a maximum value of 1 × 1011 Jones at an optical power of 1.13 μW. Specific detectivity (D*) is a key parameter used to assess a device's sensitivity to weak optical signals. Considering that the dominant source of background noise originates from the dark current, D* can be calculated using the following expression:
|
 | (3) |
In which, e is the charge of an electron, A is the area of the laser spot, and R is the photoresponsivity.37 Fig. 3g presents the photoresponse curves under 447 nm, 637 nm, and 940 nm lasers illumination, all with an optical power of 5 μW. The photocurrents under 447 nm and 637 nm illumination are approximately 80 times higher than that under 940 nm, indicating that the device has a strong detection capability for visible light. In addition, the correlation between Iph and Pin is described by the power-law function:
In an ideal photodiode, the photocurrent is linearly proportional to the incident power.38 The fitted α values under different wavelength illuminations are shown in Fig. 3h. Under 447 nm and 637 nm illumination, α is slightly less than 1, while under 940 nm illumination, α reaches 1.2. For visible light, the device exhibits nearly linear photoresponse, indicating minimal carrier recombination and efficient charge separation at the interface, which can be caused by the inherent electric field across the interface. Under 940 nm infrared light illumination, the observed superlinear response (α = 1.2) may result from mechanisms such as trap filling and thermal excitation. Fig. 3i shows the maximum self-powered responsivity and specific detectivity under different wavelength illuminations (Vgs = 0 V, Vds = 0 V). The device achieves a maximum self-powered responsivity of 21 mA W−1 under 447 nm illumination and a maximum detectivity of 1 × 1011 Jones under 637 nm illumination. The device demonstrates high responsivity and specific detectivity in the visible range, while both metrics drop significantly under near-infrared illumination. This is because visible photons possess enough energy to drive interband transitions across the bandgap of MoTe2, whereas 940 nm photons (∼1.32 eV) have energies close to or even lower than the indirect bandgap of multilayer MoTe2, resulting in lower excitation efficiency. The light output curves at different wavelengths, along with the corresponding photocurrent and responsivity, are provided in Fig. S6.
The response time of a photodetector is a key parameter that determines its bandwidth, temporal resolution, and high-speed performance. As shown in Fig. 4a and b, when Vds = 0 V, the rise time is 54.6 μs and the decay time is 55.2 μs. Fig. 4c displays the device's response under different switching frequencies, further demonstrating its capability to detect ultrafast optical signals and maintain stable operation under high-frequency conditions. Fig. 4d and e show the photocurrent response of the homojunction device under 532 nm laser illumination at modulation frequencies of 2 kHz and 9 kHz, respectively. The device exhibits fast and highly repeatable switching behavior, indicating excellent stability and reliability when processing optical signals at varying frequencies. The cutoff frequency refers to the frequency at which the output signal amplitude drops to 3 dB below its initial value as the pulse laser modulation frequency increases. As shown in Fig. 4f, the photocurrent response as a function of modulation frequency reveals a cutoff frequency of 10 kHz for the device. Table 1 compares the performance of this work with that of state-of-the-art photodetectors, showing that our device demonstrates overall superior performance.
 |
| Fig. 4 Self-powered response time of the device. (a) Rise time. (b) Decay time. (c) Rise and decay times under different modulation frequencies. Fast and repeatable photocurrent switching under 532 nm laser illumination at modulation frequencies of (d) 2 kHz and (e) 9 kHz, respectively. (f) Normalized photocurrents as a function of frequencies. | |
Table 1 Performance comparison of self-powered photodetectors based on MoTe2 and other photodetectors
Materials/device structure |
Wavelength spectrum range [nm] |
Maximum R [A W−1] |
Maximum D* [Jones] |
Rise/fall [ms] |
Ref. |
MoTe2 homojunction |
447–940 |
0.021 |
1.0 × 1011 |
0.054/0.055 |
This work |
MoTe2 homojunction |
— |
0.00063 |
1.4 × 109 |
15/17 |
39 |
GeSe/SnSe |
— |
0.12 |
5.4 × 1011 |
23/61 |
40 |
Graphene/InSe/MoS2 vertical heterostructure |
— |
0.11 |
1.0 × 1010 |
<1 |
41 |
Bi2O2Se |
420–940 |
0.033 |
9.9 × 1010 |
40/50 |
42 |
PdSe2/NbSe2 |
405–980 |
0.027 |
9.6 × 107 |
0.0016/0.0019 |
43 |
BP/MoS2 |
450–700 |
0.000034 |
— |
1000/10 0000 |
44 |
Se-P/P3HT |
300–900 |
0.034 |
6.5 × 1011 |
16/27 |
45 |
q-2D/3D CsPbBr3 |
— |
0.16 |
1.81 × 107 |
0.18/0.12 |
46 |
As illustrated in Fig. 5(a), the optical microscope image features red coordinate axes aligned with the reference system used in the photocurrent mapping analysis. Fig. 5(b) presents the spatially resolved photocurrent mapping image acquired under a 1 V bias voltage. The photocurrent signal exhibits pronounced localization along the periphery of the top h-BN encapsulation layer, as evidenced by the intense red-orange contrast in this region. This spatial distribution suggests the existence of a robust built-in electric field at the top h-BN/MoTe2 interface, which facilitates efficient separation of photogenerated electron–hole pairs. In stark contrast, the bottom h-BN-encapsulated boundary displays negligible photocurrent response, appearing as a uniform blue-green background in the false-color map. This asymmetry strongly corroborates that adsorbed H2O and O2 molecules on the exposed MoTe2 surface induce Fermi level pinning, resulting in p-type doping. Consequently, a vertical p–n junction forms between the environmentally doped exposed region and the protected h-BN-encapsulated area. The absence of measurable photocurrent at the bottom interface further excludes potential artifacts from substrate-induced carrier scattering or interfacial strain effects, confirming the intrinsic origin of the observed photovoltaic behavior.
 |
| Fig. 5 (a) Optical microscope image. (b) Photocurrent mapping image. | |
2.4. Formation mechanism of the MoTe2 homojunction
Fig. 6a schematically illustrates the pre-annealing configuration of a partially h-BN encapsulated MoTe2 device, with its equilibrium energy band diagram shown in Fig. 6c. In this pristine state, the exposed MoTe2 surface remains free of molecular adsorbates, preserving the material's intrinsic ambipolar transport properties and resulting in flat energy bands (ΔΦ = 0) across the MoTe2/h-BN heterostructure. Following atmospheric annealing (Fig. 6b), H2O and O2 molecules chemisorb onto the exposed MoTe2 surface, acting as electron acceptors via the redox reaction MoTe2 + O2/H2O → MoTe2+ + e−, which lowers the Fermi level, inducing degenerate p-type doping in the exposed region. In contrast, the h-BN-encapsulated area retains intrinsic ambipolarity with EF near mid-gap. This spatial carrier polarity modulation creates a type-II band alignment at the lateral heterojunction (Fig. 6d), where a Fermi level discontinuity drives band bending, generating a built-in electric field. Under illumination, this field enables efficient photogenerated carrier separation, reduces Schottky barrier asymmetry, and enhances minority carrier extraction.
 |
| Fig. 6 Fabrication process of the homojunction structure regulated by H2O and O2 molecule adsorption on MoTe2 surface and its band evolution. (a) Schematic illustration of the initial partially encapsulated MoTe2. (b) Schematic of the device after annealing in air, leading to adsorption of H2O and O2 molecule. (c) Energy band diagram of the initial partially encapsulated MoTe2 (Vds = 0 V, Vgs = 0 V). (d) Energy band diagram of the device after annealing (Vds = 0 V, Vgs = 0 V). | |
3. Conclusion
We used a dry transfer technique to partially encapsulate MoTe2. After annealing in air, H2O and O2 molecules were adsorbed onto the unencapsulated region, causing it to exhibit p-type behavior. The encapsulated part of MoTe2 retained its intrinsic ambipolar characteristics, creating a p–n homojunction. This allowed for the separation of photogenerated carriers and resulted in a device with high responsivity, detectivity, and ultrafast response speed under zero bias. This approach offers a promising method for constructing high-performance 2D homojunction devices through interfacial molecular modulation, with potential applications in flexible optoelectronics, image sensors, imaging systems, and environmental sensing platforms.
4. Experimental section
MoTe2 and h-BN were both obtained via mechanical exfoliation. MoTe2 was then transferred onto h-BN using a dry transfer technique. After spin coating, photolithography, metal deposition, and lift-off processes, 50 nm of Au was deposited on both ends of the MoTe2 as metal electrodes. A maskless UV lithography system (produced by Tuotuo Technology Co., Ltd, Suzhou) was used for the photolithography process. The thickness of the samples was measured using a scanning probe microscope equipped with an AFM module. Raman spectra were acquired using a micro-Raman spectrometer with a 532 nm laser. The photoelectric performance was measured using a Keithley 2634B source meter.
Author contributions
Kangwei Shao: investigation, data curation, writing original draft. Haiyan Nan: investigation, methodology, and formal analysis. Renxian Qi: validation, formal analysis. Chenxin Jiang and Haomin Wang: investigation (characterization), data curation, writing – review & editing. Zhengjin Weng and Jialing Jian: supervision, writing – review & editing. Shaoqing Xiao and Xiaofeng Gu: funding acquisition, supervision.
Conflicts of interest
The authors declare no competing financial interest.
Data availability
The data supporting this article have been included as part of the SI. Please refer to the supporting information for details on device fabrication, the effect of annealing time on electrical properties, XPS analysis of MoTe2 before and after annealing, electrical characteristics and self-powered photoresponse of the unannealed device, stability tests, and photoresponse under lasers of different wavelengths. See DOI: https://doi.org/10.1039/d5tc02340h
Acknowledgements
This work is supported by the National Science Foundation under Grants 62074070, 52203356, 62364003 and 62104084, the Natural Science Foundation of Jiangsu Province, China under Grants BK20221534 and BK20221065, the Natural Science Foundation of Jiangxi Province, China under Grants 20224BAB202035, Open Research Fund of State Key Laboratory of Materials for Integrated Circuits (No. SKLJC-K2024-X03).
References
- A. Ahmed, M. Z. Iqbal, A. Dahshan, S. Aftab, H. H. Hegazy and E. Yousef, Recent advances in 2D transition metal dichalcogenide-based photodetectors: a review, Nanoscale, 2024, 16, 2097–2120 RSC
. - Y. Luo, J. X. Zhao, A. Fieramosca, Q. B. Guo, H. F. Kang, X. Z. Liu, T. C. H. Liew, D. Sanvitto, Z. Y. An, S. Ghosh, Z. Y. Wang, H. X. Xu and Q. H. Xiong, Strong light-matter coupling in van der Waals materials, Light: Sci. Appl., 2024, 13, 203 CrossRef CAS PubMed
. - Z. Wang, K. V. Sreekanth, M. Zhao, J. P. Nong, Y. C. Liu and J. H. Teng, Two-dimensional materials for tunable and nonlinear metaoptics, Adv. Photonics, 2024, 6, 034001 CAS
. - S. Q. Yan, X. L. Zhu, J. J. Dong, Y. H. Ding and S. S. Xiao, 2D materials integrated with metallic nanostructures: fundamentals and optoelectronic applications, Nanophotonics, 2020, 9, 1877–1900 CrossRef CAS
. - S. B. Homan, V. K. Sangwan, I. Balla, H. Bergeron, E. A. Weiss and M. C. Hersam, Ultrafast exciton dissociation and long-lived charge separation in a photovoltaic pentacene-MoS2 van der Waals heterojunction, Nano Lett., 2017, 17, 164–169 CrossRef PubMed
. - C. H. Jin, E. Y. Ma, O. Karni, E. C. Regan, F. Wang and T. F. Heinz, Ultrafast dynamics in van der Waals heterostructures, Nat. Nanotechnol., 2018, 13, 994–1003 CrossRef CAS PubMed
. - S. N. Li, J. J. Zhou, J. X. Xiong, S. X. Yang, J. L. Zhang, W. J. Fan and J. B. Li, Two dimensional CuInP2S6/h-BN/MoTe2 van der Waals heterostructure phototransistors with double gate control, J. Mater. Chem. C, 2025, 13, 2378–2387 RSC
. - Y. Z. Li, J. Y. Pan, C. X. Yan, J. X. Li, W. Xin, Y. T. Zhang, W. Z. Liu, X. F. Liu, H. Y. Xu and Y. C. Liu, Efficient carrier multiplication in self-powered near-ultraviolet γ-InSe/graphene heterostructure photodetector with external quantum efficiency exceeding 161%, Nano Lett., 2024, 24, 7252–7260 CrossRef CAS PubMed
. - W. Zheng, Y. S. Xu, L. L. Zheng, C. Yang, N. Pinna, X. H. Liu and J. Zhang, MoS2 van der Waals p–n junctions enabling highly selective room-temperature NO2 sensor, Adv. Funct. Mater., 2020, 30, 2000435 CrossRef CAS
. - M. J. Dai, H. Y. Chen, F. K. Wang, M. S. Long, H. M. Shang, Y. X. Hu, W. Li, C. Y. Ge, J. Zhang, T. Y. Zhai, Y. Q. Fu and P. A. Hu, Ultrafast and sensitive self-powered photodetector featuring self-limited depletion region and fully depleted channel with van der Waals contacts, ACS Nano, 2020, 14, 9098–9106 CrossRef CAS PubMed
. - E. Reato, P. Palacios, B. Uzlu, M. Saeed, A. Grundmann, Z. Y. Wang, D. S. Schneider, Z. X. Wang, M. Heuken, H. Kalisch, A. Vescan, A. Radenovic, A. Kis, D. Neumaier, R. Negra and M. C. Lemme, Zero-bias power-detector circuits based on MoS2 field-effect transistors on wafer-scale flexible substrates, Adv. Mater., 2022, 34, 2108469 CrossRef CAS PubMed
. - Z. W. Wang, Z. Z. Yin, Z. Y. Yang, F. K. Shan, J. Huang and D. D. Hao, Research progress of self-powered photodetectors based on halide perovskites, Chem. Eng. J., 2024, 501, 157512 CrossRef CAS
. - W. Zheng, X. H. Liu, J. Y. Xie, G. C. Lu and J. Zhang, Emerging van der Waals junctions based on TMDs materials for advanced gas sensors, Coord. Chem. Rev., 2021, 447, 214151 CrossRef CAS
. - X. C. Liu, D. S. Qu, H. M. Li, I. Moon, F. Ahmed, C. Kim, M. Lee, Y. Choi, J. H. Cho, J. C. Hone and W. J. Yoo, Modulation of quantum tunneling via a vertical two-dimensional black phosphorus and molybdenum disulfide p-n junction, ACS Nano, 2017, 11, 9143–9150 CrossRef CAS PubMed
. - A. A. Murthy, T. K. Stanev, J. D. Cain, S. Q. Hao, T. LaMountain, S. Kim, N. Speiser, K. Watanabe, T. Taniguchi, C. Wolverton, N. P. Stern and V. P. Dravid, Intrinsic transport in 2D heterostructures mediated through h-BN tunneling contacts, Nano Lett., 2018, 18, 2990–2998 CrossRef CAS PubMed
. - D. N. Shanks, F. Mahdikhanysarvejahany, C. Muccianti, A. Alfrey, M. R. Koehler, D. G. Mandrus, T. Taniguchi, K. Watanabe, H. Y. Yu, B. J. LeRoy and J. R. Schaibley, Nanoscale trapping of interlayer excitons in a 2D semiconductor heterostructure, Nano Lett., 2021, 21, 5641–5647 CrossRef CAS PubMed
. - C. D. Zhang, M. Y. Li, J. Tersoff, Y. M. Han, Y. S. Su, L. J. Li, D. A. Muller and C. K. Shih, Strain distributions and their influence on electronic structures of WSe2-MoS2 laterally strained heterojunctions, Nat. Nanotechnol., 2018, 13, 152–158 CrossRef CAS PubMed
. - K. K. Paul, J. H. Kim and Y. H. Lee, Hot carrier photovoltaics in van der Waals heterostructures, Nat. Rev. Phys., 2021, 3, 178–192 CrossRef CAS
. - R. Frisenda, A. J. Molina-Mendoza, T. Mueller, A. Castellanos-Gomez and H. S. J. van der Zant, Atomically thin p-n junctions based on two-dimensional materials, Chem. Soc. Rev., 2018, 47, 3339–3358 RSC
. - Y. F. Kang, Y. F. Pei, D. He, H. Xu, M. J. Ma, J. L. Yan, C. Z. Jiang, W. Q. Li and X. H. Xiao, Spatially selective p-type doping for constructing lateral WS2 p-n homojunction via low-energy nitrogen ion implantation, Light: Sci. Appl., 2024, 13, 127 CrossRef CAS PubMed
. - X. M. Zheng, Y. H. Wei, J. X. Liu, S. T. Wang, J. Shi, H. Yang, G. Peng, C. Y. Deng, W. Luo, Y. Zhao, Y. Z. Li, K. L. Sun, W. Wan, H. P. Xie, Y. L. Gao, X. A. Zhang and H. Huang, A homogeneous p-n junction diode by selective doping of few layer MoSe2 using ultraviolet ozone for high-performance photovoltaic devices, Nanoscale, 2019, 11, 13469–13476 RSC
. - P. Luo, F. W. Zhuge, Q. F. Zhang, Y. Q. Chen, L. Lv, Y. Huang, H. Q. Li and T. Y. Zhai, Doping engineering and functionalization of two-dimensional metal chalcogenides, Nanoscale Horiz., 2019, 4, 26–51 RSC
. - J. Jiang, T. Xu, J. P. Lu, L. T. Sun and Z. H. Ni, Defect engineering in 2D materials: precise manipulation an improved functionalities, Research, 2019, 2019, 4641739 Search PubMed
. - R. Kumar, A. K. Shringi, H. J. Wood, I. M. Asuo, S. Oturak, D. E. Sanchez, T. S. K. Sharma, R. Chaurasiya, A. Mishra, W. M. Choi, N. Y. Doumon, I. Dabo, M. Terrones and F. Yan, Substitutional doping of 2D transition metal dichalcogenides for device applications: current status, challenges and prospects, Mater. Sci. Eng. R, 2025, 163, 100946 CrossRef
. - A. W. Robertson, Y.
C. Lin, S. S. Wang, H. Sawada, C. S. Allen, Q. Chen, S. Lee, G. D. Lee, J. Lee, S. Han, E. Yoon, A. I. Kirkland, H. Kim, K. Suenaga and J. H. Warner, Atomic structure and spectroscopy of single metal (Cr, V) substitutional dopants in monolayer MoS2, ACS Nano, 2016, 10, 10227–10236 CrossRef CAS PubMed
. - M. Rajapakse, B. Karki, U. O. Abu, S. Pishgar, M. R. K. Musa, S. M. S. Riyadh, M. Yu, G. Sumanasekera and J. B. Jasinski, Intercalation as a versatile tool for fabrication, property tuning, and phase transitions in 2D materials, npj 2D Mater. Appl., 2021, 5, 30 CrossRef CAS
. - S. Q. Shao, G. R. M. Tainton, W. J. Kuang, N. Clark, R. Gorbachev, A. Eggeman, I. V. Grigorieva, D. J. Kelly and S. J. Haigh, Nanoscale disorder and deintercalation evolution in K-doped MoS2 analysed via in situ TEM, Adv. Funct. Mater., 2023, 33, 224390 Search PubMed
. - R. J. Yang, L. Mei, Z. Y. Lin, Y. Y. Fan, J. Lim, J. H. Guo, Y. J. Liu, H. S. Shin, D. Voiry, Q. Y. Lu, J. Li and Z. Y. Zeng, Intercalation in 2D materials and in situ studies, Nat. Rev. Chem., 2024, 8, 410–432 CrossRef CAS PubMed
. - J. Yu, J. W. Fu, H. C. Ruan, H. Wang, Y. M. Yu, J. P. Wang, Y. H. He, J. S. Wu, F. W. Zhuge, Y. Ma and T. Y. Zhai, Tailoring lithium intercalation pathway in 2D van der Waals heterostructure for high-speed edge-contacted floating-gate transistor and artificial synapses, InfoMat, 2024, 6, e12599 CrossRef CAS
. - R. H. Sun, S. Y. Sun, X. Liang, H. Y. Gong, X. S. Zhang, Y. Li, M. Gao, D. W. Li and G. C. Xu, Surface charge transfer doping of MoS2 monolayer by molecules with aggregation-induced emission effect, Nanomaterials, 2022, 12, 164 Search PubMed
. - S. S. Withanage, B. Chamlagain, A. C. Johnston and S. I. Khondaker, Charge transfer doping of 2D PdSe2 thin film and Its application in fabrication of heterostructures, Adv. Electron. Mater., 2021, 7, 2001057 CrossRef CAS
. - H. Y. Nan, Z. L. Wang, W. H. Wang, Z. Liang, Y. Lu, Q. Chen, D. W. He, P. H. Tan, F. Miao, X. R. Wang, J. L. Wang and Z. H. Ni, Strong photoluminescence enhancement of MoS2 through defect engineering and oxygen bonding, ACS Nano, 2014, 8, 5738–5745 CrossRef CAS PubMed
. - S. Tongay, J. Zhou, C. Ataca, J. Liu, J. S. Kang, T. S. Matthews, L. You, J. B. Li, J. C. Grossman and J. Q. Wu, Broad-range modulation of light emission in two-dimensional semiconductors by molecular physisorption gating, Nano Lett., 2013, 13, 2831–2836 CrossRef CAS PubMed
. - J. Ristein, Surface transfer doping of semiconductors, Science, 2006, 313, 1057–1058 CrossRef CAS PubMed
. - D. S. Qu, X. C. Liu, M. Huang, C. Lee, F. Ahmed, H. Kim, R. S. Ruoff, J. Hone and W. J. Yoo, Carrier-type modulation and mobility improvement of thin MoTe2, Adv. Mater., 2017, 29, 1606433 CrossRef PubMed
. - Q. Wang, Y. W. Song, Y. Q. Ran, Y. P. Li, Y. Pan and Y. Ye, Coplanar MoS2-MoTe2 heterojunction with the same crystal orientation, Small, 2024, 19, 2308635 CrossRef PubMed
. - C. Xie, C. Mak, X.
M. Tao and F. Yan, Photodetectors based on two-dimensional layered materials beyond graphene, Adv. Funct. Mater., 2017, 27, 1603886 CrossRef
. - Y. W. Zhang, K. K. Ma, C. Zhao, W. Hong, C. J. Nie, Z. J. Qiu and S. Wang, An ultrafast WSe2 photodiode based on a lateral p-i-n homojunction, ACS Nano, 2021, 15, 4405–4415 CrossRef CAS PubMed
. - Y. X. Hu, J. Wang, M. Tamtaji, Y. Feng, T. W. Tang, M. Amjadian, T. Kang, M. Y. Xu, X. Y. Shi, D. X. Zhao, Y. L. Mi, Z. T. Luo and L. An, Adv. Mater., 2025, 37, 2404013 CrossRef CAS PubMed
. - Y. L. Mao, T. Z. Deng, Y. X. Li and F. He, Appl. Phys. Lett., 2024, 124, 181106 CrossRef CAS
. - Z. S. Chen, Z. L. Zhang, J. Biscaras and A. Shukla, J. Mater. Chem. C, 2018, 6, 12407–12412 RSC
. - P. K. Roy, K. Mosina, S. Hengtakaeh, K. J. Sarkar, V. Mazánek, J. Luxa and Z. Sofer, ACS Appl. Nano Mater., 2024, 7, 24377–24387 CrossRef CAS PubMed
. - C. Su, M. Y. Li, H. Yan, Y. Zhang, H. Li, W. H. Fan, W. J. Bai, X. J. Liu, Q. G. Wang and S. G. Yin, ACS Appl. Mater. Interfaces, 2025, 17, 5213–5222 CrossRef PubMed
. - A. Mazzotti, O. Durante, S. De Stefano, L. Viscardi, A. Pelella, O. Kharsah, L. Daniel, S. Sleziona, M. Schleberger and A. Di Bartolomeo, Adv. Opt. Mater., 2025, 13, 2500811 CrossRef CAS
. - X. W. Yu, Y. X. Huang, P. F. Li, S. L. Feng, X. Wan, Y. F. Jiang and P. P. Yu, Nanomaterials, 2024, 14, 1923 CrossRef CAS PubMed
. - C. Y. Sung, W. C. Chen, C. L. Liu, B. H. Lin, Y. C. Lin and W. C. Chen, Adv. Opt. Mater., 2024, 12, 2303241 CrossRef CAS
.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.