Next Article in Journal
Individual Effects of Alkali Element and Wire Structure on Metal Transfer Process in Argon Metal-Cored Arc Welding
Next Article in Special Issue
Ultrasensitive Detection of Malachite Green Isothiocyanate Using Nanoporous Gold as SERS Substrate
Previous Article in Journal
Influence of Microstructure and Mechanical Properties of Dissimilar Rotary Friction Welded Inconel to Stainless Steel Joints
Previous Article in Special Issue
Electrochemical SEIRAS Analysis of Imidazole-Ring-Functionalized Self-Assembled Monolayers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of GaN in the Heterostructure WS2/GaN for SERS Applications

1
Department of Electronic Engineering, National Changhua University of Education, No. 2, Shi-Da Road, Changhua 50074, Taiwan
2
International College of Semiconductor Technology, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
*
Author to whom correspondence should be addressed.
Materials 2023, 16(8), 3054; https://doi.org/10.3390/ma16083054
Submission received: 10 March 2023 / Revised: 7 April 2023 / Accepted: 11 April 2023 / Published: 12 April 2023

Abstract

:
In the application of WS2 as a surface–enhanced Raman scattering (SERS) substrate, enhancing the charge transfer (CT) opportunity between WS2 and analyte is an important issue for SERS efficiency. In this study, we deposited few-layer WS2 (2–3 layers) on GaN and sapphire substrates with different bandgap characteristics to form heterojunctions using a chemical vapor deposition. Compared with sapphire, we found that using GaN as a substrate for WS2 can effectively enhance the SERS signal, with an enhancement factor of 6.45 × 104 and a limit of detection of 5 × 10−6 M for probe molecule Rhodamine 6G according to SERS measurement. Analysis of Raman, Raman mapping, atomic force microscopy, and SERS mechanism revealed that The SERS efficiency increased despite the lower quality of the WS2 films on GaN compared to those on sapphire, as a result of the increased number of transition pathways present in the interface between WS2 and GaN. These carrier transition pathways could increase the opportunity for CT, thus enhancing the SERS signal. The WS2/GaN heterostructure proposed in this study can serve as a reference for enhancing SERS efficiency.

1. Introduction

Surface-enhanced Raman scattering (SERS) technique is a powerful tool for detecting and characterizing molecules at very low concentrations [1,2]. However, the SERS signal can still be weak for some analytes and in some experimental conditions. Improving the SERS signal would enable more sensitive detection and better characterization of molecules, which can have important applications in fields such as biomedicine, environmental monitoring, and materials science [3,4,5,6]. Additionally, a better understanding of the mechanism behind SERS enhancement is important for further optimizing the technique and developing new applications. The SERS effect is generally accepted to occur through the following two mechanisms: the electromagnetic mechanism (EM) and the chemical mechanism (CM). The use of noble metals in SERS methods is commonly adopted to achieve a significant enhancement in signal due to high electromagnetic fields around metallic structures [7,8] in the case of the EM effect. The metallic structure can provide a boost of several orders of magnitude to SERS signals. The CM effect, on the other hand, refers to the chemical interaction between the analyte molecule and the substrate [9,10]. Generally, the SERS signal of the CM effect is much lower than that of the EM effect. In recent years, several studies have demonstrated that 2D materials such as graphene, boron nitride, and molybdenum disulfide (MoS2) can be utilized as SERS substrates for biomedical detections [11,12].
Among 2D materials used for SERS measurements, tungsten disulfide (WS2) is a two–dimensional layered material that has recently attracted great interest in electronic and optoelectronic applications due to its high electron mobility, direct band gap, and strong absorption properties in the visible and near–infrared regions [13,14,15,16]. Recent research has also pointed out that the high sensitivity, good chemical stability, and excellent biocompatibility of WS2 make it highly promising for SERS applications in detecting probe molecules [17,18,19]. The mechanism behind the use of WS2 as a SERS substrate to enhance the SERS signal intensity of the analyte is attributed to a CM, where the exciton generated under light illumination undergoes a CT process between the energy levels of the analyte and WS2 at their interface, causing a significant enhancement of Raman signals achieved through charge transitions [20].
Compared to the noble metals such as Au and Ag, which are commonly used for enhancing the SERS signal through the electromagnetic mechanism, the enhancement effect of the SERS signal using CM of WS2 is weaker and limited [21]. Therefore, it is necessary to enhance the signal intensity by using nanostructures. However, the process of producing nanostructured WS2 is complex, and the size is difficult to control [22]. Therefore, we propose using a simple semiconductor heterostructure of WS2/GaN to effectively enhance the SERS signal. In this study, we utilized a metal-organic chemical vapor deposition (MOCVD) system to grow GaN on sapphire, followed by the deposition of a few-layer WS2 film on both GaN and sapphire substrates using a furnace with chemical vapor deposition (CVD) progress. After that, we compared the SERS performance of the two substrates using rhodamine (R6G) as the analyte. Our analysis, which included Raman spectroscopy, Raman mapping, atomic force microscopy (AFM), scanning electron microscopy (SEM), and X-ray photoelectron spectroscopy (XPS), showed that the quality of the WS2 films grown on GaN was inferior to those grown on sapphire. Despite this, we observed that using GaN as a substrate for WS2 can significantly enhance the SERS signal, with an enhancement factor of 6.45 × 104 and a limit of detection of 5 × 10−6 M for the probe molecule R6G, as confirmed by SERS measurements. Through analysis of the CT mechanism, it can be inferred that an appropriate semiconductor heterostructure such as WS2/GaN can effectively provide more paths for carrier transition, making the effect of CT more intense and further enhancing the SERS signal intensity. A series of results could be expected to contribute to the development of using semiconductor heterostructures as SERS substrates.

2. Materials and Methods

2.1. Fabrication of WS2/GaN Heterostructure

A close–coupled showerhead MOCVD system was utilized to grow regular thin films of c–plane GaN on c–plane sapphire. A GaN nucleation layer with a thickness of around 30 nm was deposited by MOCVD at a growth temperature of 530 °C and a pressure of 600 mbar on c–plane sapphire substrates. The initial GaN layer was deposited at a growth temperature of 1030 °C and a pressure of 300 mbar. The temperature and pressure for the main GaN layer were set at 1050 °C and 150 mbar, respectively, with a V/III ratio of approximately 1200. The thickness of the GaN was approximately 2 μm. Afterward, the WS2 film was uniformly grown on both GaN and sapphire substrates positioned at the center of the furnace through the CVD process. The Ar/H2 carrier gas ratio was about 3:1 (150 sccm and 50 sccm). During growing few-layer WS2, the growth gas used was a mixture of H2S (10% to Ar) and WF6, with a flow ratio of H2S/WF6 equal to 200:1.

2.2. Preparation of R6G Molecules

A solution of R6G (ACROS Organics, 99%, Geel, Belgium) in water with concentrations ranging from 10−2 to 10−6 M was prepared and dip–coated onto the WS2/GaN, WS2/sapphire, and sapphire substrates. The 10−2 M R6G solution was also placed onto a flat sapphire as a reference substrate, thereby allowing measurement of the EF. The specimens were subjected to ambient conditions and dried on a hot plate at 70 °C for 5 min.

2.3. Characterizations

A confocal Raman microscopy system (Tokyo Instruments, Nanofinder 30, Tokyo, Japan) was used to perform Raman spectra and SERS analyses. The excitation was carried out by a He–Ne laser, and the laser power used was 0.1 mW for as–grown samples and 0.3 mW for the reference substrate. The laser spot size was adjusted to 3 μm using a microscope objective with a magnification of 100× and a numerical aperture of 0.9. The acquisition time for laser measurements was configured to 20 s, while the spectrometer employed a grating with a density of 300 lines per millimeter. The resolution of Raman spectrometer was about 0.5 cm−1. The spectrometer’s charge–coupled device was cooled to around −50 °C, utilizing a thermoelectric cooling chip to minimize measurement noise. Prior to measurement, the spectra were calibrated using the peak (520 cm−1) of the Si on the bulk Si substrate. The surface morphology of the samples was examined using an atomic force microscope (Veeco DI–3100, Plainview, NY, USA) through an AFM scan. SEM (FEI Helios 1200+, Hillsboro, OR, USA) revealed the morphologies and nanostructures of the as-grown samples. The chemical configurations of the samples were determined using an XPS instrument (ULVAC-PHI Phi V5000, Chigasaki, Japan) equipped with an Al Kα X-ray source and applied to the samples for analysis.

3. Results and Discussion

WS2 is a two-dimensional material in which the layers interact with each other through van der Waals forces [23]. Even though there is a bonding force between WS2 and the underlying GaN atoms, the resulting stress has a negligible effect on the arrangement of the lattice above WS2. Thus, we were unable to use X-ray diffraction to analyze the aforementioned information and instead were only able to further analyze the thickness of WS2 on GaN and sapphire using Raman spectroscopy. The molecule vibration modes of the WS2 grown on GaN and sapphire substrates under identical CVD conditions were examined through Raman spectroscopy analysis. The Raman spectra depicted in Figure 1a reveal the typical E12g and A1g vibration mode peaks of WS2 for both WS2/GaN and WS2/sapphire samples [24], indicating successful deposition of the WS2 layer on the respective substrates. A widely recognized phenomenon in the field is that the difference in wavenumber between the two vibration modes E12g and A1g of WS2 exhibits a decreasing trend with the reduction in WS2 thickness from bulk to few-layered structures [25,26]. Based on the observed difference in wavenumber of approximately 63 cm−1 for the WS2/GaN sample and 62 cm−1 for the WS2/sapphire sample, it can be inferred that the WS2 thin film formed on both substrates has a thickness ranging from 2 to 3 layers. On the other hand, it has been reported by McCreary et al. that the quality of WS2 thin film can be determined by observing the full–width at half–maximum (FWHM) of the Raman A1g peak [27]. Therefore, the A1g peaks of the two samples mentioned above were fitted with Gaussian curves to obtain accurate FWHM values. As shown in Figure 1b, the FWHM of the WS2 A1g peak grown on sapphire was 4.07 cm−1, while that on GaN was approximately 4.85 cm−1. This result suggests that the WS2 grown on sapphire has better thin film quality compared to that grown on GaN. The reason for this will be discussed later.
In order to further compare the uniformity of WS2 grown on GaN and sapphire substrates in terms of Raman results, we further utilized Raman mapping to obtain the spatial distribution of the two peaks E12g and A1g, as well as the wavenumber difference (Δω) between the two peaks. As shown in Figure 2, when WS2 was grown on sapphire, the uniformity of both E12g and A1g peaks was significantly higher than that of GaN. In addition, the results of Δω also showed that the thickness uniformity was better when grown on sapphire. Overall, the thickness range of WS2 grown on GaN was larger, which also verified the results of Figure 1a. Furthermore, our AFM measurement results, shown in Figure 3a,b, indicate that the root mean square (RMS) values of surface roughness for WS2/GaN and WS2/sapphire are 0.39 and 0.11, respectively. This suggests that the surface of WS2/GaN is rougher than that of WS2/sapphire, consistent with the aforementioned Raman and Raman mapping results. The two AFM images in Figure 3c,d provide further evidence that the surface of GaN is not as flat as that of sapphire, as the former has undergone MOCVD epitaxial growth while the latter has been polished. The RMS roughness of GaN is approximately 0.09, while that of sapphire is 0.06. We also conducted additional SEM observations, as shown in Figure S2a,b in the Supplementary Materials. The WS2 grown on GaN appeared significantly rougher than the WS2 grown on sapphire. It can be observed that the SEM results are consistent with the AFM results. This explains why the subsequent growth of WS2 on GaN, in terms of surface uniformity and thickness uniformity, is inferior to that on sapphire. Another possible reason is that during the furnace CVD growth process, due to its higher surface activity compared to sapphire, GaN is more prone to adsorb some of the native contaminants in the furnace, such as O, C, and S atoms [28], which can affect the interface smoothness of the subsequently grown WS2 film. Therefore, it can be concluded that the WS2 grown on GaN exhibits inferior lattice quality and surface roughness compared to that grown on sapphire. Regarding this aspect, we provide XPS data and have a detailed discussion below.
To further understand the reason for the poor surface quality of WS2 grown on GaN, we performed XPS experiments to analyze the content of S, C, and O atoms. The results of XPS are shown below in Figure 4a–d. Based on the signals of (a) S2p, (b) S2s, and (c) C1s, it can be concluded that when WS2 was grown on GaN, both the amount of S and C atoms were higher than when it was grown on sapphire. This result is consistent with our AFM results mentioned above. Compared to sapphire, GaN is more prone to absorb S, C, and other atoms on its surface during the growth of WS2, which leads to poorer surface quality. On the other hand, as shown in Figure 4d, there are relatively more O atoms in the WS2 on the sapphire structure. It is well known that sapphire has a higher affinity for oxygen compared to GaN [29]. This is because sapphire has a more polar surface with a higher surface energy, making it easier for oxygen atoms to be absorbed or captured by the surface. In contrast, GaN has a nonpolar surface with lower surface energy, which makes it less likely for oxygen atoms to be captured or absorbed. Another possible reason for the higher concentration of O atoms could be the presence of O atoms in the sapphire substrate itself.
We conducted SERS measurements on these two types of SERS substrates using R6G of 10−4 M as the analyte, as shown in Figure 5a. The results indicate that under this range of wavenumbers, the main R6G SERS signals measured on WS2/GaN are significantly stronger than those on WS2/sapphire, with the peak at 611 cm−1 being particularly prominent [30]. Although WS2/sapphire has better film quality, WS2/GaN shows a more prominent performance as a SERS substrate. Moreover, we used this WS2/GaN heterostructure substrate to measure different concentrations of R6G solution, as shown in Figure 5b. It can be observed that a weak 611 cm−1 peak signal is still present at 5 × 10−6 M, but no R6G signal can be detected at 10−6 M, indicating that the LOD of WS2/GaN, in our case, can reach 5 × 10−6 M. Generally, the EF is an important metric for evaluating the performance of SERS because it quantifies the level of signal enhancement that occurs when analyte molecules are adsorbed onto either a semiconductor or plasmonic substrate. EF is calculated by comparing the Raman signal intensity of a given molecule on a plasmonic substrate with that of the same molecule in its bulk form. A high EF indicates that the plasmonic substrate provides a strong enhancement effect on the Raman signal of the analyte, leading to a more sensitive detection method. Therefore, the EF value is commonly used as a figure of merit to compare the performance of different SERS substrates and to optimize the SERS experimental conditions. According to the calculation formula of EF shown in the following equation [31]:
EF = I SERS I REF × N R E F N S E R S ,
where I is the Raman spectral intensity and N is the number of molecules, respectively; SERS and REF denote the values obtained from WS2/GaN and the reference sample of the sapphire substrate, respectively. As the laser spot size and exposure time were constant for all measurements, the ratio of NREF/NSERS was equivalent to the R6G concentration ratio for both samples. According to the above equation, it can be determined that the EF for the peak 611 cm−1 of R6G measured using the WS2/GaN SERS substrate can reach 6.45 × 104. Therefore, in terms of overall SERS performance, WS2/GaN performs better than WS2/sapphire. In the following discussion, we will explore the advantages of using GaN in the WS2 heterostructure for SERS measurement based on our experimental results. We also conducted measurements on the relationship between the 611 cm−1 peak intensity and different R6G concentrations (ranging from 10−4 M to 5 × 10−6 M) for both samples. The results are presented in Figure S3 in the Supplementary Materials. It is apparent from the figure that the peak intensity WS2/GaN was consistently higher than that of WS2/sapphire in this R6G concentration range, indicating that our measurements were highly reliable.
In general, the enhancement mechanism of semiconductors as SERS substrates mainly comes from the CT between the photoexcited carriers and the analyte in the band gap of the semiconductor [32]. Lee et al. reported that WS2 demonstrates better SERS performance as the number of atomic layers decreases [33]. From a band perspective, our Raman results indicate that WS2 has a thickness of only about 2–3 atomic layers and a direct bandgap of about 1.6 eV [34]. The laser energy used in our experiments is approximately 1.96 eV, as shown in Figure 6, which can provide carrier transition pathways (2), (3), (5), and (6). The corresponding energy levels are the lowest unoccupied molecular orbital (LUMO) and highest occupied molecular orbital (HOMO) of R6G, as well as the conduction band edge energy level (Ec) and valence band edge energy level (Ev) of WS2. The energy differences between these pathways are all smaller than the energy of the laser, allowing WS2 to effectively enhance the SERS signal intensity of R6G. On the other hand, when GaN is introduced as a substrate for WS2 to form the WS2/GaN heterostructure, the Raman results in Figure S1 show that the laser intensity can penetrate through the thin WS2 film and measure the oscillation frequency signals E2 and A1 of GaN [35]. This demonstrates that the laser can also excite carriers in the interface between WS2 and GaN with energy differences smaller than the laser energy. Compared to sapphire, GaN has a bandgap of approximately 3.4 eV, which at the interface with 1.6 eV of WS2, can provide additional effective carrier transition pathways (7)–(9). This allows the excited electrons to have the opportunity to migrate to the interface between R6G and WS2, increasing the probability of CT and enhancing the SERS signal intensity. For the purpose of comparison, we have listed the various pathways and their corresponding energy levels in Table 1, along with the possibility of carrier transition.
In addition to the transition pathways between the WS2/GaN heterostructure mentioned above that contribute to the strong SERS signal, other possibilities still need to be considered. Our AFM results show that the surface of WS2 grown on GaN is rougher than that grown on sapphire. Therefore, the higher roughness of WS2 on GaN also provides a larger contact area with the probe molecule R6G, which could enhance the SERS signal. Furthermore, in the above discussions, we mentioned that the native impurities O, C, and S in the furnace tube can be easily adsorbed onto the GaN surface, forming additional deep levels in the GaN bandgap. Therefore, these impurities may provide additional transition pathways for both WS2 and GaN, thereby increasing the SERS intensity. However, our current experimental results do not confirm the contribution of these reasons to the enhancement of the SERS signal. Therefore, in the future, we will design further experiments to analyze these possible factors.

4. Conclusions

In conclusion, we propose the use of GaN as a substrate for few-layer WS2 to form WS2/GaN heterostructures as SERS substrates and compare their thin film quality and SERS performance with WS2/sapphire. Through a series of experiments, including Raman, Raman mapping, AFM, and SERS mechanism investigation, our results showed that the quality of few-layer WS2 film grown on GaN was inferior to that on sapphire. However, due to the high surface roughness associated with high surface contact and the additional carrier transition paths in the WS2/GaN interface, WS2/GaN heterostructures could effectively enhance SERS signals in detecting the R6G target molecule compared to WS2/sapphire, with an EF value of 6.45 × 104 and a LOD of 5 × 10−6 M. Therefore, our study suggests that designing appropriate semiconductor heterostructure SERS substrates for different target molecules can effectively increase the chance of CT and enhance SERS signals.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16083054/s1, Figure S1: Additional Raman results of WS2/GaN heterostructure. Figure S2: SEM images of different samples: (a) WS2/GaN, and (b) WS2/sapphire. Figure S3. A relationship between the Raman peak intensities at 611 cm−1 and the R6G concentrations (from 5 × 10−6 to 10−4 M).

Author Contributions

Conceptualization and writing—original draft preparation: T.-S.K.; methodology: E.-T.L., C.-A.D. and T.-S.K.; software: E.-T.L., C.-A.D. and T.-S.K.; validation: T.-S.K., E.-T.L. and C.-A.D.; formal analysis: T.-S.K., E.-T.L. and C.-A.D.; investigation: T.-S.K., Y.-T.H., E.-T.L. and C.-A.D.; resources: T.-S.K. and Y.-T.H.; data curation: T.-S.K. and Y.-T.H.; writing—review and editing: T.-S.K. and Y.-T.H.; visualization: T.-S.K.; supervision: T.-S.K.; project administration: T.-S.K.; funding acquisition: T.-S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Science and Technology Council of Taiwan (grant number 1111-2221-E-018-016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Acknowledgments

The authors thank AIXTRON (Aachen, Germany) for help with the deposition of the GaN thin films, Taiwan Semiconductor Research of Institute (Hsinchu, Taiwan) for the help with the measurement of XPS, Roots Technology Co., Ltd. (Hsinchu, Taiwan) for help with the deposition of the few-layer WS2 and J. Shieh (National United University), as well as J.-Y. Chen (National Chung Hsing University) for assistance with Raman instrumentation and for the useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Santhoshkumar, S.; Murugan, E. Rationally designed SERS AgNPs/GO/g-CN nanohybrids to detect methylene blue and Hg2+ ions in aqueous solution. Appl. Surf. Sci. 2021, 533, 149544. [Google Scholar] [CrossRef]
  2. Tzeng, Y.; Lin, B.-Y. Silver SERS Adenine Sensors with a Very Low Detection Limit. Biosensors 2020, 10, 53. [Google Scholar] [CrossRef]
  3. Lai, H.; Li, G.; Xu, F.; Zhang, Z. Metal–organic frameworks: Opportunities and challenges for surface-enhanced Raman scattering–A review. J. Mater. Chem. C 2020, 8, 2952–2963. [Google Scholar] [CrossRef]
  4. Guo, J.; Zeng, F.; Guo, J.; Ma, X. Preparation and application of microfluidic SERS substrate: Challenges and future perspectives. J. Mater. Sci. Technol. 2020, 15, 96–103. [Google Scholar] [CrossRef]
  5. Liu, Y.; Ma, H.; Han, X.X.; Zhao, B. Metal–semiconductor heterostructures for surface-enhanced Raman scattering: Synergistic contribution of plasmons and charge transfer. Mater. Horiz. 2021, 8, 370–382. [Google Scholar] [CrossRef]
  6. Alhmoud, H.; Brodoceanu, D.; Elnathan, R.; Kraus, T.; Voelcker, N.H. A MACEing silicon: Towards single-step etching of defined porous nanostructures for biomedicine. Prog. Mater. Sci. 2021, 116, 100636. [Google Scholar] [CrossRef]
  7. Freeman, R.G.; Grabar, K.C.; Allison, K.J.; Bright, R.M.; Davis, J.A.; Guthrie, A.P.; Hommer, M.B.; Jackson, M.A.; Smith, P.C.; Walter, D.G.; et al. Self-Assembled Metal Colloid Monolayers: An Approach to SERS Substrates. Science 1995, 267, 1629–1632. [Google Scholar] [CrossRef]
  8. Singh, J.P.; Chu, H.Y.; Abell, J.; Tripp, R.A.; Zhao, Y. Flexible and mechanical strain resistant large area SERS active substrates. Nanoscale 2012, 4, 3410–3414. [Google Scholar] [CrossRef]
  9. Xia, L.; Chen, M.; Zhao, X.; Zhang, Z.; Xia, J.; Xu, H.; Sun, M. Visualized method of chemical enhancement mechanism on SERS and TERS. J. Raman Spectrosc. 2014, 45, 533–540. [Google Scholar] [CrossRef]
  10. Chen, R.; Jensen, L. Interpreting the chemical mechanism in SERS using a Raman bond model. J. Chem. Phys. 2020, 152, 024126. [Google Scholar] [CrossRef]
  11. Qiu, H.; Li, Z.; Gao, S.; Chen, P.; Zhang, C.; Jiang, S.; Li, H. Large-area MoS2 thin layers directly synthesized on Pyramid-Si substrate for surface-enhanced Raman scattering. RSC Adv. 2015, 5, 83899–83905. [Google Scholar] [CrossRef]
  12. Dai, Z.; Xiao, X.; Wu, W.; Zhang, Y.; Liao, L.; Guo, S.; Jiang, C. Plasmon-driven reaction controlled by the number of graphene layers and localized surface plasmon distribution during optical excitation. Light Sci. Appl. 2015, 4, e342–e348. [Google Scholar] [CrossRef]
  13. Ernandes, C.; Khalil, L.; Almabrouk, H.; Pierucci, D.; Zheng, B.; Avila, J.; Dudin, P.; Chaste, J.; Oehler, F.; Pala, M.; et al. Indirect to direct band gap crossover in two–dimensional WS2(1−x)Se2x alloys. NPJ 2D Mater. Appl. 2021, 5, 7. [Google Scholar] [CrossRef]
  14. Wang, Y.; Tutuc, E.; Sohier, T.; Watanabe, K.; Taniguchi, T.; Verstraete, M.J. Electron mobility in monolayer WS2 encapsulated in hexagonal boron–nitride. Appl. Phys. Lett. 2021, 118, 102105. [Google Scholar] [CrossRef]
  15. Zhang, D.; Xiong, Y.; Chai, J.; Liu, T.; Ba, X.; Cheng, J.; Ullah, S.; Zheng, G.; Yan, M.; Cao, M. Synergetic dielectric loss and magnetic loss towards superior microwave absorption through hybridization of few–layer WS2 nanosheets with NiO nanoparticles. Sci. Bull. 2020, 65, 138–146. [Google Scholar] [CrossRef]
  16. Zhang, X.; Dong, Q.; Li, Z.; Jing, X.; Liu, R.; Liu, B.; Zhao, T.; Lin, T.; Li, Q.; Liu, B. Significant pressure–induced enhancement of photoelectric properties of WS2 in the near–infrared region. Mater. Res. Lett. 2022, 10, 547–555. [Google Scholar] [CrossRef]
  17. Ghopry, S.A.; Alamri, M.A.; Goul, R.; Sakidja, R.; Wu, J.Z. Extraordinary Sensitivity of Surface–Enhanced Raman Spectroscopy of Molecules on MoS2 (WS2) Nanodomes/Graphene van der Waals Heterostructure Substrates. Adv. Optical Mater. 2019, 7, 1801249. [Google Scholar] [CrossRef]
  18. Ghopry, S.A.; Alamri, M.; Goul, R.; Cook, B.; Sadeghi, S.M.; Gutha, R.R.; Sakidja, R.; Wu, J.Z. Au Nanoparticle/WS2 Nanodome/Graphene van der Waals Heterostructure Substrates for Surface–Enhanced Raman Spectroscopy. ACS Appl. Nano Mater. 2020, 3, 2354–2363. [Google Scholar] [CrossRef]
  19. Song, Y.; Huang, H.C.; Lu, W.; Li, N.; Su, J.; Cheng, S.B.; Lai, Y.; Chen, J.; Zhan, J. Ag@WS2 quantum dots for Surface Enhanced Raman Spectroscopy: Enhanced charge transfer induced highly sensitive detection of thiram from honey and beverages. Food Chem. 2022, 344, 128570. [Google Scholar] [CrossRef]
  20. Fu, S.; Fossé, I.d.; Jia, X.; Xu, J.; Yu, X.; Zhang, H.; Zheng, W.; Krasel, S.; Chen, Z.; Wang, Z.M.; et al. Long–lived charge separation following pump–wavelength–dependent ultrafast charge transfer in graphene/WS2 heterostructures. Sci. Adv. 2021, 7, eabd9061. [Google Scholar] [CrossRef]
  21. Shin, Y.; Kim, J.; Jang, Y.; Ko, E.; Lee, N.S.; Yoon, S.; Kim, M.H. Vertically–Oriented WS2 Nanosheets with a Few Layers and Its Raman Enhancements. Nanomaterials 2020, 10, 1847. [Google Scholar] [CrossRef]
  22. Berkdemir, A.; Gutiérrez, H.R.; Botello–Méndez, A.R.; Perea-López, N.; Elías, L.A.; 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]
  23. Zheng, D.; Dong, X.; Lu, J.; Niu, Y.; Wang, H. Ultrafast and hypersensitized detection based on van der Waals connection in two-dimensional WS2/Si structure. Appl. Surf. Sci. 2022, 574, 151662. [Google Scholar] [CrossRef]
  24. Sinha, S.; Arora, S.K.; Sathe, V. Temperature Dependent Phononic Response of Liquid Phase Exfoliated Few Layered WS2 Nanosheets. AIP Conf. Proc. 2019, 2115, 030197. [Google Scholar]
  25. Zhang, Y.; Zhang, Y.; Ji, Q.; Ju, J.; Yuan, H.; Shi, J.; Gao, T.; Ma, D.; Liu, M.; Chen, Y.; et al. Controlled Growth of High–Quality Monolayer WS2 Layers on Sapphire and Imaging Its Grain Boundary. ACS Nano 2013, 7, 8963–8971. [Google Scholar] [CrossRef]
  26. Qiao, S.; Yang, H.; Bai, Z.; Peng, G.; Zhang, X. Thickness–dependent structural and electrical properties of WS2 nanosheets obtained via the ALD–grown WO3 sulfurization technique as a channel material for field–effect transistors. ACS Omega 2021, 6, 34429–34437. [Google Scholar]
  27. McCreary, A.; Berkdemir, A.; Wang, J.; Nguyen, M.A.; Elías, A.L.; Perea-López, N.; Fujisawa, K.; Kabius, B.; Carozo, V.; Cullen, D.A.; et al. Distinct photoluminescence and Raman spectroscopy signatures for identifying highly crystalline WS2 monolayers produced by different growth methods. J. Mater. Res. 2016, 31, 931–944. [Google Scholar] [CrossRef]
  28. Henning, A.; Bartl, J.D.; Zeidler, A.; Qian, S.; Bienek, O.; Jiang, C.-M.; Paulus, C.; Rieger, B.; Stutzmann, M.; Sharp, I.D. Aluminum Oxide at the Monolayer Limit via Oxidant–Free Plasma–Assisted Atomic Layer Deposition on GaN. Adv. Funct. Mater. 2021, 31, 2101441. [Google Scholar] [CrossRef]
  29. Cui, Y.; Niu, X.; Zhou, J.; Wang, Z.; Wang, R.; Zhang, J. Effect of chloride ions on the chemical mechanical planarization efficiency of sapphire substrate. ECS J. Solid State Sci. Technol. 2019, 8, 488–495. [Google Scholar] [CrossRef]
  30. He, X.N.; Gao, Y.; Mahjouri-Samani, M.; Black, P.N.; Allen, J.; Mitchell, M.; Xiong, W.; Zhou, Y.S.; Jiang, L.; Lu, Y.F. Surface–enhanced Raman spectroscopy using gold–coated horizontally aligned carbon nanotubes. Nanotechnology 2012, 23, 205702. [Google Scholar] [CrossRef]
  31. Chen, M.; Solarska, R.; Li, M. Additional Important Considerations in Surface–Enhanced Raman Scattering Enhancement Factor Measurements. J. Phys. Chem. C 2023, 127, 2728–2734. [Google Scholar] [CrossRef]
  32. Lombardi, J.R.; Birke, R.L. Theory of Surface–Enhanced Raman Scattering in Semiconductors. J. Phys. Chem. C 2014, 118, 11120–11130. [Google Scholar] [CrossRef]
  33. Lee, Y.; Kim, H.; Lee, J.; Yu, S.H.; Hwang, E.; Lee, C.; Ahn, J.-H.; Cho, J.H. Enhanced Raman Scattering of Rhodamine 6G Films on Two–Dimensional Transition Metal Dichalcogenides Correlated to Photoinduced Charge Transfer. Chem. Mater. 2016, 28, 180–187. [Google Scholar] [CrossRef]
  34. Gutiérrez, H.R.; Perea-López, N.; Elías, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; Lopez-Urías, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary Room–Temperature Photoluminescence in Triangular WS2 Monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef]
  35. Bagnall, K.R.; Wang, E.N. Contributed Review: Experimental characterization of inverse piezoelectric strain in GaN HEMTs via micro–Raman spectroscopy. Rev. Sci. Instrum. 2016, 87, 061501. [Google Scholar] [CrossRef]
Figure 1. (a) Raman spectra of few-layer WS2 grown on GaN and sapphire. (b) Enlarged range of the A1g peak in Figure 1a and corresponding Gaussian fitting result.
Figure 1. (a) Raman spectra of few-layer WS2 grown on GaN and sapphire. (b) Enlarged range of the A1g peak in Figure 1a and corresponding Gaussian fitting result.
Materials 16 03054 g001
Figure 2. Raman mapping results of E12g, A1g, and Δω for WS2/sapphire and WS2/GaN.
Figure 2. Raman mapping results of E12g, A1g, and Δω for WS2/sapphire and WS2/GaN.
Materials 16 03054 g002
Figure 3. AFM images of different samples: (a) WS2/GaN, (b) WS2/sapphire, (c) GaN, and (d) sapphire.
Figure 3. AFM images of different samples: (a) WS2/GaN, (b) WS2/sapphire, (c) GaN, and (d) sapphire.
Materials 16 03054 g003
Figure 4. (a) S2p, (b) S2s, (c) C1s and (d) O1s core level XPS spectra of both WS2/GaN and WS2/sapphire.
Figure 4. (a) S2p, (b) S2s, (c) C1s and (d) O1s core level XPS spectra of both WS2/GaN and WS2/sapphire.
Materials 16 03054 g004
Figure 5. (a) Comparison of SERS results for measuring 10−4 M R6G using WS2/GaN and WS2/sapphire. (b) SERS results for different concentrations of R6G using WS2/GaN.
Figure 5. (a) Comparison of SERS results for measuring 10−4 M R6G using WS2/GaN and WS2/sapphire. (b) SERS results for different concentrations of R6G using WS2/GaN.
Materials 16 03054 g005
Figure 6. Schematic diagram of CT pathways between R6G molecules, few-layer WS2 and GaN.
Figure 6. Schematic diagram of CT pathways between R6G molecules, few-layer WS2 and GaN.
Materials 16 03054 g006
Table 1. Corresponding energy differences and probabilities of carrier transition pathways indicated in Figure 6.
Table 1. Corresponding energy differences and probabilities of carrier transition pathways indicated in Figure 6.
Pathways(1)(2)(3)(4)(5)(6)(7)(8)(9)(10)(11)
ΔE (eV)2.31.50.82.20.11.41.70.31.73.13.4
PossibilityXOOXOOOOOXX
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

Ko, T.-S.; Lin, E.-T.; Ho, Y.-T.; Deng, C.-A. The Role of GaN in the Heterostructure WS2/GaN for SERS Applications. Materials 2023, 16, 3054. https://doi.org/10.3390/ma16083054

AMA Style

Ko T-S, Lin E-T, Ho Y-T, Deng C-A. The Role of GaN in the Heterostructure WS2/GaN for SERS Applications. Materials. 2023; 16(8):3054. https://doi.org/10.3390/ma16083054

Chicago/Turabian Style

Ko, Tsung-Shine, En-Ting Lin, Yen-Teng Ho, and Chen-An Deng. 2023. "The Role of GaN in the Heterostructure WS2/GaN for SERS Applications" Materials 16, no. 8: 3054. https://doi.org/10.3390/ma16083054

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