Next Article in Journal
Recent Advances in the Photonic Curing of the Hole Transport Layer, the Electron Transport Layer, and the Perovskite Layers to Improve the Performance of Perovskite Solar Cells
Previous Article in Journal
Tailored Triggering of High-Quality Multi-Dimensional Coupled Topological States in Valley Photonic Crystals
Previous Article in Special Issue
Photoelectric Properties of GaS1−xSex (0 ≤ x ≤ 1) Layered Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Selective Epitaxy of 2D Sb2Te3 onto Monolayer WS2 for Vertical p–n Heterojunction Photodetectors

by
Baojun Pan
1,†,
Zhenjun Dou
2,†,
Mingming Su
1,
Ya Li
1,
Jialing Wu
1,
Wanwan Chang
1,
Peijian Wang
2,
Lijie Zhang
2,
Lei Zhao
3,*,
Mei Zhao
2,* and
Sui-Dong Wang
1
1
Macao Institute of Materials Science and Engineering (MIMSE), MUST-SUDA Joint Research Center for Advanced Functional Materials, Macau University of Science and Technology, Taipa, Macao 999078, China
2
Key Laboratory of Carbon Materials of Zhejiang Province, Institute of New Materials & Industry Technology, College of Chemistry and Materials Engineering, Wenzhou University, Wenzhou 325035, China
3
School of Electronic Engineering, Lanzhou City University, Lanzhou 730070, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(10), 884; https://doi.org/10.3390/nano14100884
Submission received: 10 April 2024 / Revised: 10 May 2024 / Accepted: 13 May 2024 / Published: 19 May 2024

Abstract

:
Two-dimensional transition metal dichalcogenides (2D-TMDs) possess appropriate bandgaps and interact via van der Waals (vdW) forces between layers, effectively overcoming lattice compatibility challenges inherent in traditional heterojunctions. This property facilitates the creation of heterojunctions with customizable bandgap alignments. However, the prevailing method for creating heterojunctions with 2D-TMDs relies on the low-efficiency technique of mechanical exfoliation. Sb2Te3, recognized as a notable p-type semiconductor, emerges as a versatile component for constructing diverse vertical p–n heterostructures with 2D-TMDs. This study presents the successful large-scale deposition of 2D Sb2Te3 onto inert mica substrates, providing valuable insights into the integration of Sb2Te3 with 2D-TMDs to form heterostructures. Building upon this initial advancement, a precise epitaxial growth method for Sb2Te3 on pre-existing WS2 surfaces on SiO2/Si substrates is achieved through a two-step chemical vapor deposition process, resulting in the formation of Sb2Te3/WS2 heterojunctions. Finally, the development of 2D Sb2Te3/WS2 optoelectronic devices is accomplished, showing rapid response times, with a rise/decay time of 305 μs/503 μs, respectively.

1. Introduction

Two-dimensional materials are pivotal in advancing the miniaturization of electronic devices to enhance Moore’s Law, owing to their unique atomic-scale thickness [1,2]. The van der Waals interlayer interaction in these materials overcomes lattice-matching challenges faced in conventional heterojunctions [3], enabling a “Lego-like” stacking approach in layered heterostructures, which serves as an excellent research platform in optics [4], electronics [5], magnetism [6,7], and other fields [8].
Two-dimensional transition metal dichalcogenides (2D-TMDs) are distinguished among the materials investigated due to their layered structure, which includes appropriate bandgaps and chemical stability [9,10]. These characteristics render them highly suitable for optoelectronic device fabrication. In particular, the p–n heterojunction is a crucial structure in photodetectors, as it facilitates efficient separation of photogenerated electrons. Intrinsic two-dimensional transition metal chalcogenides are primarily n-type semiconductors, with p-type semiconductors being less common. For example, monolayer PtSe2 is a p-type semiconductor with an indirect band gap of 1.16 eV [11,12,13]. To achieve high-performance photodetectors, this work introduces Sb2Te3, a p-type semiconductor with a direct band gap of 0.33 eV [14,15], to construct p–n heterojunctions with n-type two-dimensional transition metal chalcogenides. Nevertheless, the current approach to building novel two-dimensional layered material heterostructures primarily relies on the repetitive and inefficient mechanical exfoliation process [16], which is not conducive to mass industrial production.
Considered a method with potential for large-scale, controllable fabrication of 2D-TMD [2], chemical vapor deposition (CVD) is compatible with existing semiconductor processes and can produce high-quality crystals. The synthesis of two-dimensional transition metal dichalcogenide heterostructures via CVD involves both one-step and two-step processes. For instance, Jiao et al. demonstrated a one-step procedure wherein they sulfurized WO3−x/MoO3−x core-shell nanowires, resulting in vertically stacked MoS2/WS2 heterojunctions [17]. In contrast, a two-step CVD method, as employed by Duan et al., involves initial laser etching to create nucleation sites on large-area WSe2 surfaces, followed by epitaxial growth of VSe2 to produce arrayed VSe2/WSe2 vertical heterojunctions [18]. During the synthesis of two-dimensional transition metal chalcogenide heterostructures, the one-step method can easily lead to atomic intermixing within the heterostructures [19]. In contrast, the two-step method requires careful avoidance of surface contamination on the two-dimensional materials produced in the first step to prevent interference with the epitaxial growth of the two-dimensional transition metal chalcogenides in the second step [20,21].
To date, there have been no reports on the preparation of Sb2Te3/WS2 p–n heterojunctions using CVD. Therefore, this paper attempts to fabricate Sb2Te3/WS2 p–n heterojunctions using a two-step chemical vapor deposition method. Initially, to explore the preparation conditions for two-dimensional Sb2Te3, we achieved large-scale fabrication on inert mica substrates as a reference for constructing related Sb2Te3 heterostructures. Due to differences in surface migration barriers, van der Waals layered materials tend to selectively epitaxially grow on surfaces with lower migration barriers [22,23,24]. Therefore, by exploiting the relatively low migration barrier of Sb2Te3 on WS2 surfaces, this work adopts a two-step chemical vapor deposition strategy to selectively epitaxially grow Sb2Te3 on pre-prepared WS2 on SiO2/Si substrates, achieving the assembly of Sb2Te3/WS2 heterojunctions.
Lastly, the two-dimensional Sb2Te3/WS2 heterostructures were utilized in the assembly of photodetectors, achieving in rapid response times of a 302-microsecond rise time and 503-microsecond decay time. This research provides valuable insights into constructing heterojunctions involving Sb2Te3 and other two-dimensional layered materials, contributing to the advancement of high-performance photodetectors. Furthermore, given that two-dimensional materials with atomic-level thickness exhibit excellent mechanical flexibility [25] and high integration, they hold potential applications in industrial flexible robotic sensors.

2. Materials and Methods

2.1. Materials

Tungsten Trioxide Powder (WO3, 99.99%), and Antimony Trioxide Powder (Sb2Te3, 99.9%) were purchased from Alfa Aesar, Waltham, MA, USA. Fluorophlogopite Mica substrates were sourced from Taiyuan Fluorophlogopite Co., Ltd. in Changchun City, China. Chromium Metal (Cr, 99.99%) and Gold Metal (Au, 99.99%) were purchased from Beijing Zhongjin Yanxin Materials Technology Co., Ltd. in Beijing, China. 285 nm of SiO2/Si substrates were obtained from Hefei Kejin Material Co., Ltd. in Hefei, China.

2.2. Synthesis of 2D-Sb2Te3

To synthesize 2D-Sb2Te3, a quartz boat containing 50 mg of Sb2Te3 powder was placed at the center of a single-zone tube furnace (with a quartz tube diameter of 1.0 inch). A mica substrate was placed 5 cm away from the center of the boat. Prior to heating, the system underwent a 30-min purge with Ar gas; once heating commenced, the Ar flow rate was set to 50 standard cubic centimeters per minute (sccm). Subsequently, the temperature was increased at a rate of 25 °C/min up to 650 °C and maintained for 2 min. Ultimately, the quartz tube was swiftly cooled to room temperature, and the sample was extracted.

2.3. Synthesis of Monolayer WS2

To synthesize monolayer WS2, the 280 nm SiO2/Si underwent a sequential treatment involving ultrasonic in acetone, isopropanol, and deionized water, each for 30 min. Subsequent to the ultrasonic treatment, the SiO2/Si was dried using a nitrogen gun, annealed in a muffle furnace at 600 °C for a duration of four hours, allowed to cool naturally to room temperature, and then set aside. Following this, 2.5 mg of WO3 powder was precisely measured and uniformly dispersed on a custom-made graphite trough. A SiO2/Si substrate measuring 1.5 cm × 1.0 cm was delicately positioned 0.5 cm above the WO3 powder in the trough. Additionally, 300 mg of sulfur powder was weighed and placed in a quartz trough fitted with a magnetic pull rod. The graphite trough containing SiO2/Si substrate and the quartz boat containing sulfur powder were positioned at the center and outside the heating region of a one-inch tube furnace. The quartz tube was purged with argon for 30 min to eliminate air and moisture. The argon flow rate was adjusted to 40 sccm, and the temperature was gradually increased at a rate of 25 °C/min up to 900 °C. Upon reaching the specified temperature, the sulfur powder was shifted to an area close to 380 °C, underwent a 5-min reaction, followed by the removal of the sulfur powder, rapid cooling to room temperature, and retrieval of the sample.

2.4. Synthesis of 2D-Sb2Te3/WS2 Heterostructure

For the fabrication of the 2D-Sb2Te3/WS2 heterostructure, a precise amount of 50 mg Sb2Te3 powder was positioned within a quartz boat placed at the central temperature zone of a single-zone tube furnace (with a quartz tube diameter of 1.0 inch). Concurrently, a SiO2/Si substrate, coated with a monolayer of WS2 flakes, was strategically placed 5 cm away from the center of the quartz tube. Before initiating the heating process, the system underwent a thorough 30-min Ar purging, followed by the adjustment of the Ar flow rate to 50 sccm. The temperature was then steadily increased at a rate of 25 °C/min until reaching 650 °C, where it was held for 2 min. Ultimately, the quartz tube was cooled rapidly to ambient temperature, and the sample was extracted.

2.5. Device Fabrication and Testing

For the fabrication of optoelectronic devices, standard e-beam lithography and metal thermal evaporation techniques were used to define electrodes Cr/Au (10/50 nm). Optoelectronic measurements were conducted with the probe station CRX-6.5K from Lake Shore and the semiconductor parameter analyzer Keithley 4200 SCS (Cleveland, OH, USA) at room temperature under a 532 nm laser. Laser power density was collected with a powermeter (Thorlabs GmbH., PM 100D, Dachau, Germany).

2.6. Characterization

The morphology of the samples was examined using optical microscopy (OM, Carl Zeiss Microscopy GmbH, Jena, Germany) and scanning electron microscopy (SEM, Nova Nano SEM 200 FEI, Hillsboro, OR, USA). Their chemical compositions were assessed through Raman and photoluminescence (PL) spectra at room temperature. Raman spectra were obtained using a micro confocal Raman/PL spectrometer (Renishaw in Via, Gloucestershire, UK) with an excitation laser line operating at 532 nm. Flake thickness measurements were conducted using atomic force microscopy (AFM) with a Dimension Icon system from Bruker, San Diego, CA, USA. The crystallinity quality and chemical composition of the Sb2Te3 and Sb2Te3/WS2 flakes were investigated using transmission electron microscopy (TEM, JEOL 2100F, Tokyo, Japan) and X-ray photoelectron spectroscopy (XPS) analysis performed on a PHI 5000 VP III instrument, Woodbury, MN, USA.

2.7. Density Functional Theory (DFT) Calculations

Geometrical optimizations were performed using DFT with Perdew–Burke–Ernzerhof generalized gradient approximation (PBE-GGA) functional employing projector augmented wave (PAW) potentials, implemented in the Vienna Ab-initio Simulation Package (VASP). The band structure of monolayer WS2 and the band structure of 2D Sb2Te3 were calculated separately based on their respective structures. A kinetic cutoff energy of 450 eV was chosen for the planewave basis set. The valence electron configurations for W (5p66s25d4), S (3s23p4), Sb (5s25p3), Te (5s25p4) were employed. The first Brillouin zone was characterized by a Γ-point-centered Monkhorst-Pack k-mesh with a grid configuration of 6 × 6 × 4. The energy convergence criterion was set at 1.0 × 10−4 eV, for both structural optimizations and self-consistent-field (SCF) iteration. The force components convergence criterion was set at −0.02 eV/Å, with the charge density symmetrization employed.

3. Results and Discussion

3.1. Morphological and Structural Characterization of WS2, Sb2Te3, and Sb2Te3/WS2 Heterojunction

This paper describes the fabrication process of Sb2Te3/WS2 heterostructures utilizing a two-step chemical vapor deposition method. Initially, we successfully prepared large-area monolayer WS2 on SiO2/Si substrate. Figure S1a shows an optical image of monolayer WS2 flakes uniformly distributed on SiO2/Si substrate. Additionally, a single WS2 flake was subjected to AFM, revealing a measured thickness of approximately 0.83 nm, consistent with reported monolayer WS2 thicknesses, as shown in Figure S1b [26]. Moreover, Raman spectroscopy was performed on the obtained WS2. Figure S1c presents the Raman characteristic peaks of monolayer WS2, where 349.8 cm−1 represents the in-plane vibrational peak and 419.2 cm−1 represents the out-of-plane vibrational peak of WS2 [27]. Photoluminescence (PL) testing of the WS2 showed a strong 632 nm PL peak, consistent with previous findings of monolayer WS2 [28]. Therefore, the uniformly distributed flakes on the SiO2/Si substrate were confirmed to be monolayer WS2.
Subsequently, to explore the fabrication conditions for two-dimensional Sb2Te3, mica substrates, which lack dangling bonds similar to monolayer WS2 surfaces, were utilized. Figure S2 shows optical images of two-dimensional Sb2Te3 at different positions on a 1.5 cm × 1.5 cm mica substrate. Figure S3a presents an optical image of thin-layer Sb2Te3 crystals, while Figure S3b shows an AFM image of Sb2Te3 with a measured height of 3.8 nm corresponding to four layers of Sb2Te3 [14]. Figure S3c shows the Raman spectrum of Sb2Te3 highlighting the characteristic Raman peaks A11g (69.8 cm−1), E2g (112.1 cm−1), and A21g (164.9 cm−1) [29]. A low-resolution TEM image of two-dimensional Sb2Te3 is depicted in Figure S4a. Figure S4b illustrates a high-resolution image, with a 0.22 nm interplanar spacing, consistent with the reported (100) plane spacing of Sb2Te3 [14]. Figure S4c displays a selected-area electron diffraction pattern of two-dimensional Sb2Te3, consistent with the reported hexagonal pattern. Finally, energy-dispersive spectroscopy testing of the crystals, as shown in Figure S4e,f, revealed the uniform distribution of Sb and Te elements within the flake. Thus, the flakes formed on the mica substrate were identified as large-area, uniformly distributed two-dimensional Sb2Te3.
The preparation process for the Sb2Te3/WS2 vertical heterostructure involved substituting the mica substrate with a single layer of WS2 flakes grown on a SiO2/Si substrate. Figure 1a illustrates the schematic of the synthesis of Sb2Te3/WS2 vertical heterostructures utilizing a two-step chemical vapor deposition (CVD) method. During this process, Sb2Te3 powders were placed at the central end of the quartz reactor, while a SiO2/Si substrate with pre-fabricated WS2 flakes was positioned downstream. The thermal evaporation of Sb2Te3 powders enabled the vertical epitaxial growth of Sb2Te3 over the monolayer WS2 (detailed growth process described in Section 2.4. Synthesis of 2D-Sb2Te3/WS2 Heterostructure). The growth dynamics of Sb2Te3/WS2 vertical heterostructures were meticulously examined. As depicted in Figure 1b, Sb2Te3 selectively grew on the surface of WS2. The pronounced optical contrast vividly illustrates the morphological evolution of Sb2Te3/WS2 grains, progressing from incomplete to complete coverage with increasing growth time from 0 to 90 s. Representative SEM micrographs of pure WS2 samples and Sb2Te3/WS2 heterostructures are shown in Figure 1c and Figure 1d, respectively. The SEM micrograph depicts a typical vertically arranged Sb2Te3/WS2 heterojunction, with Sb2Te3 partially covering the underlying WS2 layer. Evidently, Sb2Te3 flakes tend to deposit on the WS2 surface, with minimal residual sediment observed on the SiO2/Si substrate. AFM was utilized for sample characterization, as illustrated in Figure 1e,f. The thickness measurements of WS2 and Sb2Te3 at 0.857 nm and 1.598 nm, respectively, confirm a monolayer structure for WS2 and a bilayer structure for Sb2Te3, termed as 2QL (with each quintuple layer (QL) in Sb2Te3 arranged in a Te-Sb-Te-Sb-Te sequence) [30,31]. Additionally, the analysis of the heterostructure morphology demonstrates the uniform surface quality of both Sb2Te3 and WS2.
The atomic arrangement of the vertically stacked Sb2Te3/WS2 van der Waals heterojunction was further investigated using TEM coupled with energy-dispersive spectroscopy (EDS). Figure 2a displays a low-magnification TEM image of the as-transferred Sb2Te3/WS2 sample on a copper grid. Figure 2b presents an elemental distribution map of Te, Sb, S, and W originating from the marked sample region in Figure 2a. Notably, W and S elements exhibit uniform distribution throughout the entire area, whereas Sb and Te elements predominantly concentrate within the dark triangular region. Figure S5 presents the EDS spectra obtained from Figure 2a. The high-resolution TEM image in Figure 2c focuses on the region delineated by the red dashed line in Figure 2a. The contrasting regions reveal a lattice spacing of approximately 0.27 nm on the right, corresponding to the (100) plane of hexagonal WS2 [32], and a lattice spacing of 0.21 nm on the left, corresponding to the (100) plane of hexagonal Sb2Te3 [14]. Furthermore, the clarity within the heterojunction of lattice fringe patterns demonstrates its superior crystal quality. The selected area electron diffraction (SAED) pattern featured in Figure 2d provides compelling evidence of the crystallographic structure of the Sb2Te3/WS2 heterojunction. This pattern exhibits two distinct sets of diffraction patterns: one corresponding to the lattice of Sb2Te3, characterized by a spacing of 0.21 nm, and another corresponding to the lattice of WS2, characterized by a spacing of 0.27 nm. These distinct diffraction patterns unequivocally confirm the single-crystal nature of both Sb2Te3 and WS2, thereby underscoring the high-quality crystalline properties exhibited by the heterojunction.
X-ray photoelectron spectroscopy (XPS) analysis (shown in Figure 2e–h) was utilized as a tool for elucidating the chemical composition of the Sb2Te3/WS2 heterostructure. In Figure 2e, distinct peaks observed at binding energies of 33.7 and 35.8 eV correspond to the chemical states of W 4f7/2 and W 4f5/2, respectively. Additionally, peaks observed at binding energies of 163.4 and 164.5 eV are indicative of the chemical states of S 2p3/2 and S 2p1/2 in Figure 2f, respectively [33]. The XPS data for Sb 3d and Te 3d are presented in Figure 2g and Figure 2h, respectively. The fitted curves for Sb 3d at 528.8 and 538.2 eV correspond to Sb 3d3/2 and Sb 3d5/2 [34]. Meanwhile, Te 3d displays a peak at 569.7 eV for Te 3d5/2, and 580.1 eV for Te 3d3/2 [15,35]. Notably, higher binding energy peaks at 573.5 eV and 582.9 eV are attributed to tellurium oxide formation resulting from surface oxidation [36]. The values measured for S 2p, W 4f, Sb 3d, and Te 3d are consistent with reported values for WS2 and Sb2Te3. These analyses suggest that, during the top Sb2Te3 vapor growth process, no additional impurities infiltrate the underlying WS2 flake. Overall, the TEM and XPS investigations affirm the high crystal quality exhibited by the fabricated Sb2Te3/WS2 van der Waals heterostructure.
Raman and photoluminescence (PL) measurements were conducted utilizing a 532 nm laser excitation to comprehensively characterize the Sb2Te3/WS2 vertical heterojunction. Figure 3a exhibits the optical image of the partially covered Sb2Te3/WS2 heterojunction sample. Moving to Figure 3b, the Raman spectrum of the Sb2Te3/WS2 heterostructure reveals distinct Raman peaks corresponding to WS2 (E2g1 at 350 cm−1 and mode A1g at 412 cm−1) [37] and Sb2Te3 (A2u3 at 135 cm−1) [34] within the junction region. This observation serves to validate the vertical configuration of the Sb2Te3/WS2 junction. Further examination through Raman mapping, as shown in Figure 3c,d, provides additional insights. The Raman mapping at 135 cm−1, shows uniform signal intensity distribution at the center of the crystal, significantly stronger than at the periphery, indicating uniform Sb2Te3 in the crystal center. Conversely, the Raman mapping at 350 cm−1 displays a uniform signal intensity distribution at the crystal’s periphery, stronger than at the center, signifying uniform WS2 at the crystal’s outer region. Notably, the stronger signal of WS2 at 350 cm−1, combined with the central Raman spectrum analysis, confirms a vertically stacked Sb2Te3/WS2 structure within the central circle region. Room temperature PL spectra and mapping are presented in Figure 3e,f. The PL mapping depicts strong PL emission at 631 nm from the single WS2 region (blue line in Figure 3e), while the PL of WS2 is significantly quenched in the vertically stacked Sb2Te3/WS2 heterostructure region (red line in Figure 3e). Normalization of the peak intensity of the PL spectra, as illustrated in Figure S6, unveils a redshift in the PL peak of the Sb2Te3/WS2 heterostructure. This redshift is attributed to strong charge transfer between WS2 and Sb2Te3, serving as the primary cause of the significant PL quenching and redshift in the PL peak position [38]. In summary, the successful fabrication of the Sb2Te3/WS2 vertical heterostructure is confirmed through comprehensive Raman and PL characterizations.

3.2. Density Functional Theory Calculations of Band Structures for Monolayer (1L) WS2 and Two-Dimensional (2D) Sb2Te3

Using density functional theory calculations, the band structure of monolayer WS2 and the band structure of 2D Sb2Te3 were individually calculated based on their respective structures, as illustrated in Figure 4a,b. According to existing reports, the conduction band minimum (CBM) and valence band maximum (VBM) of WS2 are, respectively, at −4.39 eV and −6.46 eV [39]; whereas for Sb, the CBM and VBM are located at −4.15 eV and −4.45 eV [40], as demonstrated in Figure 4c. Upon the formation of heterostructures, a type-II band alignment is observed at the junction interface. Upon laser excitation at 532 nm, photoexcited electrons in Sb2Te3 tend to transfer to WS2, while holes in WS2 preferentially migrate to Sb2Te3. This charge separation mechanism hinders recombination within the heterostructure, leading to notable photoluminescence (PL) quenching (Figure 4d) and redshift in the PL peak position (Figure S6), consistent with the observed PL test results [38,39,40,41].

3.3. Optoelectronic Testing of Sb2Te3/WS2

The optoelectronic performance of the Sb2Te3/WS2 nanoflakes was investigated by fabricating Sb2Te3/WS2 based photodetectors on a SiO2/Si substrate, as illustrated in Figure 5a. Employing a Cr/Au electrode, one end was connected to the upper Sb2Te3 layer, while the other was linked to the lower WS2 layer. Figure 5b demonstrates the transfer characteristic curves of the Sb2Te3/WS2 p–n heterojunction. The prevailing n-type transfer curve suggests that electron transport dominates the charge transport within WS2. Figure 5c presents the Ids-Vds curves of the Sb2Te3/WS2 p–n heterojunction under dark conditions and various power levels of 532 nm laser irradiation. With increasing optical power, the photogenerated current increases. Moreover, the power-dependent photoresponse was modeled using the power law ( I p h = α P θ ) to examine the trap states within the Sb2Te3/WS2 nanoflakes [42,43], as depicted in Figure 5d. The fitting coefficient of 0.62 implies that as laser intensity rises, light absorption gradually saturates. To assess the photoresponse speed of the Sb2Te3/WS2 heterojunction photodetector, time-resolved photoresponse measurements were conducted, with results depicted in Figure 5e. The photocurrent demonstrates efficient switching between on and off states by periodic activation and deactivation of the laser, showcasing exceptional stability. Figure 5f shows the rise/decay time of the photocurrent in the photodetector, where the rise/decay time is defined as the time required for the photocurrent to increase from 10% to 90% of the peak value and decrease from 90% to 10% of the peak value, respectively. Further analysis reveals a rise time (τrise) of 305 μs and a decay time (τdecay) of 503 μs for the device, surpassing those of the WSe2/WS2 vertical heterostructure [44]. Given the excellent mechanical flexibility [25] and high integration of two-dimensional layered materials, their potential applications in future industrial flexible robotic sensors warrant additional research into two-dimensional optoelectronic devices.

4. Conclusions

In summary, we achieved a successful synthesis of large-sized two-dimensional Sb2Te3 on mica substrates via chemical vapor deposition. Through subsequent advancements, we attained precise deposition of Sb2Te3 onto the WS2 surface, resulting in the successful fabrication of Sb2Te3/WS2 heterojunctions. Fabricated photodetectors based on Sb2Te3/WS2 showed rapid response times, with a rise time of 305 μs and a decay time of 503 μs. This progress promotes the construction of heterojunctions between two-dimensional Sb2Te3 and various bandgap layered materials. Such progress not only paves the way for the realization of high-performance photodetectors but also propels the application of two-dimensional Sb2Te3-related heterojunctions across diverse fields.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14100884/s1, Figure S1: Optical Microscope (OM), AFM, Raman, PL characterizations of monolayer WS2. Figure S2: large-area, uniformly distributed 2D Sb2Te3 obtained at different spots on the whole 1.5 cm × 1.5 cm mica substrate. Figure S3: AFM and Raman characterizations of 2D Sb2Te3. Figure S4: TEM, SAED, EDS characterizations of 2D Sb2Te3. Figure S5: detailed EDS spectrum of Sb2Te3/WS2 heterostructure. Figure S6: Peak intensity normalization performed on the PL spectra shown in Figure 3e.

Author Contributions

Conceptualization, B.P., Z.D., P.W., L.Z. (Lijie Zhang), L.Z. (Lei Zhao), M.Z. and S.-D.W.; investigation, B.P., Z.D., M.S., Y.L., J.W. and W.C.; methodology, B.P., Z.D., L.Z. (Lei Zhao) and M.Z. validation, B.P., Z.D., M.S., Y.L., J.W. and W.C.; writing—original draft preparation, B.P., Z.D., M.S., Y.L., J.W. and W.C.; writing—review and editing, P.W., L.Z. (Lijie Zhang), L.Z. (Lei Zhao), M.Z. and S.-D.W.; software, P.W.; supervision, P.W., L.Z. (Lijie Zhang), L.Z. (Lei Zhao), M.Z. and S.-D.W.; funding acquisition, P.W., L.Z. (Lijie Zhang), L.Z. (Lei Zhao), M.Z. and S.-D.W.; project administration, L.Z. (Lei Zhao) and M.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology Development Fund of Macao (No. 0016/2022/A), the Science and Technology Development Fund of Macao (No. 0017/2022/AGJ), the National Natural Science Foundation of China (62264007), the National Natural Science Foundation of China (51902061), and the Foundation of Gansu Educational Committee: Innovation Fund Project for University Teachers (2023A-117).

Data Availability Statement

The data presented in this study are available upon request to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shen, P.C.; Su, C.; Lin, Y.; Chou, A.S.; Cheng, C.C.; Park, J.H.; Chiu, M.H.; Lu, A.Y.; Tang, H.L.; Tavakoli, M.M.; et al. Ultralow contact resistance between semimetal and monolayer semiconductors. Nature 2021, 593, 211–217. [Google Scholar] [CrossRef]
  2. Li, T.; Guo, W.; Ma, L.; Li, W.; Yu, Z.; Han, Z.; Gao, S.; Liu, L.; Fan, D.; Wang, Z.; et al. Epitaxial growth of wafer-scale molybdenum disulfide semiconductor single crystals on sapphire. Nat. Nanotechnol. 2021, 16, 1201–1207. [Google Scholar] [CrossRef]
  3. Novoselov, K.S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A.H. 2D materials and van der Waals heterostructures. Science 2016, 353, 9439. [Google Scholar] [CrossRef]
  4. Liang, S.J.; Cheng, B.; Cui, X.; Miao, F. Van der Waals Heterostructures for High-Performance Device Applications: Challenges and Opportunities. Adv. Mater. 2020, 32, 1903800. [Google Scholar] [CrossRef]
  5. Wu, L.; Wang, A.; Shi, J.; Yan, J.; Zhou, Z.; Bian, C.; Ma, J.; Ma, R.; Liu, H.; Chen, J.; et al. Atomically sharp interface enabled ultrahigh-speed non-volatile memory devices. Nat. Nanotechnol. 2021, 16, 882–887. [Google Scholar] [CrossRef]
  6. Cheng, X.; Cheng, Z.; Wang, C.; Li, M.; Gu, P.; Yang, S.; Li, Y.; Watanabe, K.; Taniguchi, T.; Ji, W.; et al. Light helicity detector based on 2D magnetic semiconductor CrI3. Nat. Commun. 2021, 12, 6874. [Google Scholar] [CrossRef]
  7. Jiang, S.; Shan, J.; Mak, K.F. Electric-field switching of two-dimensional van der Waals magnets. Nat. Mater. 2018, 17, 406–410. [Google Scholar] [CrossRef]
  8. Zhang, D.; Liu, Y.; He, M.; Zhang, A.; Chen, S.; Tong, Q.; Huang, L.; Zhou, Z.; Zheng, W.; Chen, M.; et al. Room temperature near unity spin polarization in 2D Van der Waals heterostructures. Nat. Commun. 2020, 11, 4442. [Google Scholar] [CrossRef]
  9. Wang, X.; Song, Z.; Wen, W.; Liu, H.; Wu, J.; Dang, C.; Hossain, M.; Iqbal, M.A.; Xie, L. Potential 2D Materials with Phase Transitions: Structure, Synthesis, and Device Applications. Adv. Mater. 2019, 31, e1804682. [Google Scholar] [CrossRef]
  10. Wang, P.; Selhorst, R.; Emrick, T.; Ramasubramaniam, A.; Barnes, M.D. Bidirectional Electronic Tuning of Single-Layer MoS2 with Conjugated Organochalcogens. J. Phys. Chem. C 2019, 123, 1506–1511. [Google Scholar] [CrossRef]
  11. Wang, G.; Wang, Z.; McEvoy, N.; Fan, P.; Blau, W.J. Layered PtSe2 for Sensing, Photonic, and (Opto-)Electronic Applications. Adv. Mater. 2021, 33, e2004070. [Google Scholar] [CrossRef]
  12. Yang, Y.; Jang, S.K.; Choi, H.; Xu, J.; Lee, S. Homogeneous platinum diselenide metal/semiconductor coplanar structure fabricated by selective thickness control. Nanoscale 2019, 11, 21068–21073. [Google Scholar] [CrossRef]
  13. Li, J.; Kolekar, S.; Ghorbani-Asl, M.; Lehnert, T.; Biskupek, J.; Kaiser, U.; Krasheninnikov, A.V.; Batzill, M. Layer-Dependent Band Gaps of Platinum Dichalcogenides. ACS Nano 2021, 15, 13249–13259. [Google Scholar] [CrossRef]
  14. Liu, H.; Li, D.; Ma, C.; Zhang, X.; Sun, X.; Zhu, C.; Zheng, B.; Zou, Z.; Luo, Z.; Zhu, X.; et al. Van der Waals epitaxial growth of vertically stacked Sb2Te3/MoS2 p–n heterojunctions for high performance optoelectronics. Nano Energy 2019, 59, 66–74. [Google Scholar] [CrossRef]
  15. Li, Y.-L.; Yang, W.-Y.; Peng, Y.-M.; Yao, J.-M.; Zhong, Y.-M.; Zhang, Z.-L.; Wei, M.; Liang, G.-X.; Fan, P.; Zheng, Z.-H. Optimized thermoelectric properties of flexible p-type Sb2Te3 thin film prepared by a facile thermal diffusion method. J. Alloys Compd. 2023, 948, 169730. [Google Scholar] [CrossRef]
  16. Huang, Y.; Pan, Y.H.; Yang, R.; Bao, L.H.; Meng, L.; Luo, H.L.; Cai, Y.Q.; Liu, G.D.; Zhao, W.J.; Zhou, Z.; et al. Universal mechanical exfoliation of large-area 2D crystals. Nat. Commun. 2020, 11, 2453. [Google Scholar] [CrossRef]
  17. Zhang, Q.; Xiao, X.; Zhao, R.; Lv, D.; Xu, G.; Lu, Z.; Sun, L.; Lin, S.; Gao, X.; Zhou, J.; et al. Two-Dimensional Layered Heterostructures Synthesized from Core-Shell Nanowires. Angew. Chem. Int. Ed. 2015, 54, 8957–8960. [Google Scholar] [CrossRef]
  18. Li, J.; Yang, X.; Liu, Y.; Huang, B.; Wu, R.; Zhang, Z.; Zhao, B.; Ma, H.; Dang, W.; Wei, Z.; et al. General synthesis of two-dimensional van der Waals heterostructure arrays. Nature 2020, 579, 368–374. [Google Scholar] [CrossRef]
  19. Gong, Y.; Lei, S.; Ye, G.; Li, B.; He, Y.; Keyshar, K.; Zhang, X.; Wang, Q.; Lou, J.; Liu, Z.; et al. Two-Step Growth of Two-Dimensional WSe2/MoSe2 Heterostructures. Nano Lett. 2015, 15, 6135–6141. [Google Scholar] [CrossRef]
  20. Jiang, X.; Chen, F.; Zhao, S.; Su, W. Recent progress in the CVD growth of 2D vertical heterostructures based on transition-metal dichalcogenides. CrystEngComm 2021, 23, 8239–8254. [Google Scholar] [CrossRef]
  21. Zhang, X.; Huangfu, L.; Gu, Z.; Xiao, S.; Zhou, J.; Nan, H.; Gu, X.; Ostrikov, K.K. Controllable Epitaxial Growth of Large-Area MoS2/WS2 Vertical Heterostructures by Confined-Space Chemical Vapor Deposition. Small 2021, 17, e2007312. [Google Scholar] [CrossRef]
  22. She, Y.; Wu, Z.; You, S.; Du, Q.; Chu, X.; Niu, L.; Ding, C.; Zhang, K.; Zhang, L.; Huang, S. Multiple-Dimensionally Controllable Nucleation Sites of Two-Dimensional WS2/Bi2Se3 Heterojunctions Based on Vapor Growth. ACS Appl. Mater. Interfaces 2021, 13, 15518–15524. [Google Scholar] [CrossRef]
  23. Niu, L.; Li, Y.; Zhao, M.; Liu, Z.; Zhang, M.; Ding, C.; Dou, Z.; She, Y.; Zhang, K.; Luo, Z.; et al. Van der Waals Template-Assisted Low-Temperature Epitaxial Growth of 2D Atomic Crystals. Adv. Funct. Mater. 2022, 32, 2202580. [Google Scholar] [CrossRef]
  24. Zhang, K.; Ding, C.; Pan, B.; Wu, Z.; Marga, A.; Zhang, L.; Zeng, H.; Huang, S. Visualizing Van der Waals Epitaxial Growth of 2D Heterostructures. Adv. Mater. 2021, 33, 2105079. [Google Scholar] [CrossRef]
  25. Zhou, X.; Tian, Z.; Kim, H.J.; Wang, Y.; Xu, B.; Pan, R.; Chang, Y.J.; Di, Z.; Zhou, P.; Mei, Y. Rolling up MoSe2 Nanomembranes as a Sensitive Tubular Photodetector. Small 2019, 15, e1902528. [Google Scholar] [CrossRef]
  26. Zhang, J.; Wang, J.; Chen, P.; Sun, Y.; Wu, S.; Jia, Z.; Lu, X.; Yu, H.; Chen, W.; Zhu, J.; et al. Observation of Strong Interlayer Coupling in MoS2/WS2 Heterostructures. Adv. Mater. 2016, 28, 1950–1956. [Google Scholar] [CrossRef]
  27. Berkdemir, A.; Gutiérrez, H.R.; Botello-Méndez, A.R.; Perea-López, N.; Elías, A.L.; Chia, C.-I.; Wang, B.; Crespi, V.H.; López-Urías, F.; Charlier, J.-C.; et al. Identification of individual and few layers of WS2 using Raman Spectroscopy. Sci. Rep. 2013, 3, 1755. [Google Scholar] [CrossRef]
  28. Gong, Y.; Lin, J.; Wang, X.; Shi, G.; Lei, S.; Lin, Z.; Zou, X.; Ye, G.; Vajtai, R.; Yakobson, B.I.; et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 2014, 13, 1135–1142. [Google Scholar] [CrossRef]
  29. Ma, W.; Wu, T.; Yao, N.; Zhou, W.; Jiang, L.; Qiu, Q.; Li, J.; Huang, Z. Bandgap-independent photoconductive detection in two-dimensional Sb2Te3. Commun. Mater. 2022, 3, 68. [Google Scholar] [CrossRef]
  30. Zheng, Y.; Xia, M.; Cheng, Y.; Rao, F.; Ding, K.; Liu, W.; Jia, Y.; Song, Z.; Feng, S. Direct observation of metastable face-centered cubic Sb2Te3 crystal. Nano Res. 2016, 9, 3453–3462. [Google Scholar] [CrossRef]
  31. Zhang, L.; Song, S.; Xi, W.; Li, L.; Song, Z. Effect of Ti additions on structure and phase stability of Sb2Te3 thin films by experimental and theoretical methods. J. Mater. Sci. Mater. Electron. 2017, 29, 4704–4710. [Google Scholar] [CrossRef]
  32. Lan, C.; Zhou, Z.; Zhou, Z.; Li, C.; Shu, L.; Shen, L.; Li, D.; Dong, R.; Yip, S.; Ho, J.C. Wafer-scale synthesis of monolayer WS2 for highperformance flexible photodetectors by enhanced chemical vapor deposition. Nano Res. 2018, 11, 3371–3384. [Google Scholar] [CrossRef]
  33. Zheng, B.; Zheng, W.; Jiang, Y.; Chen, S.; Li, D.; Ma, C.; Wang, X.; Huang, W.; Zhang, X.; Liu, H.; et al. WO3–WS2 Vertical Bilayer Heterostructures with High Photoluminescence Quantum Yield. J. Am. Chem. Soc. 2019, 141, 11754–11758. [Google Scholar] [CrossRef] [PubMed]
  34. Theekhasuk, N.; Sakdanuphab, R.; Nuthongkum, P.; Pluengphon, P.; Harnwunggmoung, A.; Horprathum, M.; Limsuwan, P.; Sakulkalavek, A.; Sukwisute, P. Improving the thermoelectric properties of thick Sb2Te3 film via Cu doping and annealing deposited by DC magnetron sputtering using a mosaic target. Curr. Appl. Phys. 2021, 31, 7–15. [Google Scholar] [CrossRef]
  35. Zheng, L.; Cheng, X.; Cao, D.; Wang, Q.; Wang, Z.; Xia, C.; Shen, L.; Yu, Y.; Shen, D. Direct growth of Sb2Te3 on graphene by atomic layer deposition. RSC Adv. 2015, 5, 40007–40011. [Google Scholar] [CrossRef]
  36. Sonawane, S.M.; Chaure, S.; Chaure, N.B. Characterization of Sb2Te3 thin films prepared by electrochemical technique. J. Phys. Chem. Solids 2023, 172, 111095. [Google Scholar] [CrossRef]
  37. Wan, Y.; Li, E.; Yu, Z.; Huang, J.-K.; Li, M.-Y.; Chou, A.-S.; Lee, Y.-T.; Lee, C.-J.; Hsu, H.-C.; Zhan, Q.; et al. Low-defect-density WS2 by hydroxide vapor phase deposition. Nat. Commun. 2022, 13, 4149. [Google Scholar] [CrossRef] [PubMed]
  38. Xiao, Y.; Zhang, T.; Zhou, M.; Weng, Z.; Chang, X.; Yang, K.; Liu, J.; Li, J.; Wei, B.; Wang, Z.; et al. Disassembly of 2D Vertical Heterostructures. Adv. Mater. 2019, 31, e1805976. [Google Scholar] [CrossRef]
  39. He, X.; Deng, X.Q.; Sun, L.; Zhang, Z.H.; Fan, Z.Q. Electronic and optical properties and device applications for antimonene/WS2 van der Waals heterostructure. Appl. Surf. Sci. 2022, 578, 151844. [Google Scholar] [CrossRef]
  40. Ko, D.-K.; Kang, Y.; Murray, C.B. Enhanced Thermopower via Carrier Energy Filtering in Solution-Processable Pt–Sb2Te3 Nanocomposites. Nano Lett. 2011, 11, 2841–2844. [Google Scholar] [CrossRef]
  41. Zheng, W.; Zheng, B.; Yan, C.; Liu, Y.; Sun, X.; Qi, Z.; Yang, T.; Jiang, Y.; Huang, W.; Fan, P.; et al. Direct Vapor Growth of 2D Vertical Heterostructures with Tunable Band Alignments and Interfacial Charge Transfer Behaviors. Adv. Sci. 2019, 6, 1802204. [Google Scholar] [CrossRef] [PubMed]
  42. Zhou, X.; Zhou, N.; Li, C.; Song, H.; Zhang, Q.; Hu, X.; Gan, L.; Li, H.; Lü, J.; Luo, J.; et al. Vertical heterostructures based on SnSe2/MoS2 for high performance photodetectors. 2D Mater. 2017, 4, 025048. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Zhao, B.; Shen, D.; Tao, Q.; Li, B.; Wu, R.; Li, B.; Yang, X.; Li, J.; Song, R.; et al. Synthesis of Ultrathin 2D Nonlayered α-MnSe Nanosheets, MnSe/WS2 Heterojunction for High-Performance Photodetectors. Small Struct. 2021, 2, 2100028. [Google Scholar] [CrossRef]
  44. You, W.; Zheng, B.; Xu, Z.; Jiang, Y.; Zhu, C.; Zheng, W.; Yang, X.; Sun, X.; Liang, J.; Yi, X.; et al. Strong interfacial coupling in vertical WSe2/WS2 heterostructure for high performance photodetection. Appl. Phys. Lett. 2022, 120, 181108. [Google Scholar] [CrossRef]
Figure 1. Synthesis of vertically stacked Sb2Te3/WS2 van der Waals heterojunctions: (a) schematic overview of the growth process for the synthesis of Sb2Te3/WS2 heterojunctions using a dual-stage chemical vapor deposition process; (b) schematic diagrams illustrating the growth process of the Sb2Te3/WS2 heterostructures over time, in which the deep purple in the same triangular flake represents the Sb2Te3/WS2 heterostructures, while the light purple represents monolayer WS2 (scale bar: 10 µm); (c) scanning electron microscope (SEM) image exhibiting a typical monolayer WS2 triangular flakes, while (d) exhibits the partially covered vertically stacked Sb2Te3/WS2 heterostructures; (e) Atomic Force Microscope (AFM) images of the partially covered Sb2Te3/WS2 heterojunction; (f) the magnified view of the red square area depicted in (e).
Figure 1. Synthesis of vertically stacked Sb2Te3/WS2 van der Waals heterojunctions: (a) schematic overview of the growth process for the synthesis of Sb2Te3/WS2 heterojunctions using a dual-stage chemical vapor deposition process; (b) schematic diagrams illustrating the growth process of the Sb2Te3/WS2 heterostructures over time, in which the deep purple in the same triangular flake represents the Sb2Te3/WS2 heterostructures, while the light purple represents monolayer WS2 (scale bar: 10 µm); (c) scanning electron microscope (SEM) image exhibiting a typical monolayer WS2 triangular flakes, while (d) exhibits the partially covered vertically stacked Sb2Te3/WS2 heterostructures; (e) Atomic Force Microscope (AFM) images of the partially covered Sb2Te3/WS2 heterojunction; (f) the magnified view of the red square area depicted in (e).
Nanomaterials 14 00884 g001
Figure 2. Atomic configuration of the vertically aligned Sb2Te3/WS2 heterojunction: (a) low-magnification TEM image of a vertically aligned Sb2Te3/WS2 vdW heterojunction, the red dashed rectangular region corresponds to the location in the high-resolution TEM image in (c); (b) a 2D elemental mapping visualizing the distribution of W, S, Sb, and Te within the Sb2Te3/WS2 heterostructure; (c) high-resolution TEM image capturing the interface region of the heterojunction; (d) electron diffraction pattern captured from the aligned region of the Sb2Te3/WS2 vdW heterojunction; (eh) XPS spectra of W 4f, S 2p, Sb 3d, and Te 3d levels in the Sb2Te3/WS2 heterostructure, with black lines denoting measured data and dots representing fitting curves.
Figure 2. Atomic configuration of the vertically aligned Sb2Te3/WS2 heterojunction: (a) low-magnification TEM image of a vertically aligned Sb2Te3/WS2 vdW heterojunction, the red dashed rectangular region corresponds to the location in the high-resolution TEM image in (c); (b) a 2D elemental mapping visualizing the distribution of W, S, Sb, and Te within the Sb2Te3/WS2 heterostructure; (c) high-resolution TEM image capturing the interface region of the heterojunction; (d) electron diffraction pattern captured from the aligned region of the Sb2Te3/WS2 vdW heterojunction; (eh) XPS spectra of W 4f, S 2p, Sb 3d, and Te 3d levels in the Sb2Te3/WS2 heterostructure, with black lines denoting measured data and dots representing fitting curves.
Nanomaterials 14 00884 g002
Figure 3. Utilization of 532 nm laser excitation for Raman and photoluminescence (PL) characterization of the Sb2Te3/WS2 heterojunction: (a) optical photograph of the Sb2Te3/WS2 heterojunction; (b) Raman spectra collected from the red dots (dark covered region at the crystal center) and blue dots (light region at the outer edge of the crystal) in (a); (c,d) frequency-specific Raman mapping of the Sb2Te3/WS2 heterojunction at 135 cm−1 and 350 cm−1, correspondingly; (e) photoluminescence spectra obtained from the isolated WS2 area (blue curve) and the intersected region (red curve); (f) PL cartography of the vertical stacked Sb2Te3/WS2 heterojunction, at 631 nm.
Figure 3. Utilization of 532 nm laser excitation for Raman and photoluminescence (PL) characterization of the Sb2Te3/WS2 heterojunction: (a) optical photograph of the Sb2Te3/WS2 heterojunction; (b) Raman spectra collected from the red dots (dark covered region at the crystal center) and blue dots (light region at the outer edge of the crystal) in (a); (c,d) frequency-specific Raman mapping of the Sb2Te3/WS2 heterojunction at 135 cm−1 and 350 cm−1, correspondingly; (e) photoluminescence spectra obtained from the isolated WS2 area (blue curve) and the intersected region (red curve); (f) PL cartography of the vertical stacked Sb2Te3/WS2 heterojunction, at 631 nm.
Nanomaterials 14 00884 g003
Figure 4. Schematic diagram of theoretical calculations: (a,b) calculated band structures of 1L WS2 and two-dimensional (2D) Sb2Te3; (c) energy band profiles of the 1L WS2 and Sb2Te3 before contract; (d) band alignment illustration displaying the process of charge transfer at the interface of the junction under 532 nm laser exposure.
Figure 4. Schematic diagram of theoretical calculations: (a,b) calculated band structures of 1L WS2 and two-dimensional (2D) Sb2Te3; (c) energy band profiles of the 1L WS2 and Sb2Te3 before contract; (d) band alignment illustration displaying the process of charge transfer at the interface of the junction under 532 nm laser exposure.
Nanomaterials 14 00884 g004
Figure 5. Optoelectronic performances of 2D Sb2Te3/WS2-based photodetectors: (a) schematic representation of the Sb2Te3/WS2-based photodetector—inset: optical image of the Sb2Te3/WS2 device; (b) transfer characteristics of the Sb2Te3/WS2 device in the absence of illumination; (c) Ids-Vds curves of the Sb2Te3/WS2 device with 532 nm laser with power ranging from 0 to 192.18 mW/cm2; (d) relationship between photocurrent (Iph) and illumination intensity levels; (e) response of photocurrent in the p–n junction with light cyclically switched on and off under 532 nm laser illumination; (f) Time-evolved photoresponse of the p–n diode, specifically illustrating the photocurrent rise and decay duration.
Figure 5. Optoelectronic performances of 2D Sb2Te3/WS2-based photodetectors: (a) schematic representation of the Sb2Te3/WS2-based photodetector—inset: optical image of the Sb2Te3/WS2 device; (b) transfer characteristics of the Sb2Te3/WS2 device in the absence of illumination; (c) Ids-Vds curves of the Sb2Te3/WS2 device with 532 nm laser with power ranging from 0 to 192.18 mW/cm2; (d) relationship between photocurrent (Iph) and illumination intensity levels; (e) response of photocurrent in the p–n junction with light cyclically switched on and off under 532 nm laser illumination; (f) Time-evolved photoresponse of the p–n diode, specifically illustrating the photocurrent rise and decay duration.
Nanomaterials 14 00884 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

Pan, B.; Dou, Z.; Su, M.; Li, Y.; Wu, J.; Chang, W.; Wang, P.; Zhang, L.; Zhao, L.; Zhao, M.; et al. Direct Selective Epitaxy of 2D Sb2Te3 onto Monolayer WS2 for Vertical p–n Heterojunction Photodetectors. Nanomaterials 2024, 14, 884. https://doi.org/10.3390/nano14100884

AMA Style

Pan B, Dou Z, Su M, Li Y, Wu J, Chang W, Wang P, Zhang L, Zhao L, Zhao M, et al. Direct Selective Epitaxy of 2D Sb2Te3 onto Monolayer WS2 for Vertical p–n Heterojunction Photodetectors. Nanomaterials. 2024; 14(10):884. https://doi.org/10.3390/nano14100884

Chicago/Turabian Style

Pan, Baojun, Zhenjun Dou, Mingming Su, Ya Li, Jialing Wu, Wanwan Chang, Peijian Wang, Lijie Zhang, Lei Zhao, Mei Zhao, and et al. 2024. "Direct Selective Epitaxy of 2D Sb2Te3 onto Monolayer WS2 for Vertical p–n Heterojunction Photodetectors" Nanomaterials 14, no. 10: 884. https://doi.org/10.3390/nano14100884

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