Path-dependent hydrogen evolution reaction via selective etching of bilayer MoS2 catalysts

Min Jung a, Jungmoon Lim a, Junsung Byeon a, Taehun Kim a, Younghoon Lim a, Hongju Park a, Jaesik Eom a, Seungsub Lee a, Sangyeon Pak *b and SeungNam Cha *a
aDepartment of Physics, Sungkyunkwan University (SKKU), Suwon, Gyeonggi-do 16419, Republic of Korea. E-mail: chasn@skku.edu
bDepartment of Electronic and Electrical Engineering, Hongik University, Seoul 04066, Republic of Korea. E-mail: spak@hongik.ac.kr

Received 8th May 2024 , Accepted 13th August 2024

First published on 22nd August 2024


Abstract

Molybdenum disulfide (MoS2) is being considered as a promising candidate for replacing noble metal catalysts in the hydrogen evolution reaction (HER). However, low active site density, semiconducting properties, and electrochemically inert basal planes limit the catalytic performance of MoS2 catalysts. Therefore, various strategies have been developed for activating basal planes in MoS2 catalysts. Here, we propose an efficient laser technology for activating basal planes and creating controlled artificial patterns. The precisely controlled laser etching technique allows for etching only the top layer of the bilayer MoS2 catalyst, and using this technique, the appropriately etched area increases the active sites without any degradation of catalytic performance, enabling efficient carrier injection. The catalytic behavior of line patterned MoS2 exhibited line density dependent catalytic performance. Compared with pristine MoS2 catalysts, the etched catalysts with 5 lines in a specific area (L5-MoS2) showed a decreased overpotential and Tafel slope up to 65.1% and 73.7%, respectively. Moreover, in the case of L5-MoS2 devices, the vertical contact between the etched line and electrode exhibited improved catalytic performance with a low Tafel slope and charge transfer resistance, which is mainly attributed to the facilitated charge transfer induced rapid charge exchange process at the electrolyte/catalyst interface. This effective strategy proposes a controllable defect engineering technique for improving the catalytic performance in 2D TMDC catalysts.


1 Introduction

Recently, two-dimensional transition metal dichalcogenides (2D TMDCs) have drawn tremendous attention due to their catalytically active properties in the hydrogen evolution reaction (HER).1–6 Molybdenum disulfide (MoS2), among these 2D TMDCs, is known as a promising substitute for Pt due to its reasonable cost, elemental abundance in nature, and thermodynamically neutral Gibbs free energy (ΔGH*) at the active Mo sites. However, unlike Pt, MoS2 has a disadvantage of low catalytically active site density, which is restricted to the unsaturated Mo sites (edge and S-vacancy sites) within the MoS2 crystals.7–9 Additionally, the semiconducting nature of MoS2 crystals showed poor electrical conductivity and a large band gap, which gave rise to sluggish kinetics of the HER. Therefore, various strategies have been employed to increase the active sites of MoS2 through defect engineering, doping, fabricating heterostructures, etc.10–13

Defect engineering has emerged as a promising technology for activating basal planes in MoS2 catalysts. Among them, thermal etching can effectively generate atomic vacancies through a chemical vapor deposition (CVD) method. However, in the case of thermal etching, it is hard to specifically control the density of defects, which leads to crystal deformation. The chemical etching method offers simple and efficient strategies for creating atomic defect sites in MoS2 catalysts through a chemical reaction. However, excessive chemical treatment leads to unwanted crystal structure degradation and chemical doping, which modifies the intrinsic material properties. Therefore, simple and controllable defect engineering strategies are needed for improving the catalytic activity of the basal planes of MoS2 catalysts.

In this study, we propose controlled laser etching techniques to precisely control the density of active sites on the basal plane of bilayer MoS2, which can significantly enhance the hydrogen evolution reaction efficiency of MoS2. Through carefully controlled laser intensity, we can selectively etch only the top layer of bilayer MoS2 catalysts, thereby enabling efficient charge injection from the maintained bottom layer to active sites. We confirmed that the electrochemical hydrogen evolution performance improved as the line density increased in the same area. Laser etching increases the number of exposed Mo-edge sites in the MoS2 crystal, which possesses ideal proton adsorption energy. Thus, the catalytic activity was improved with a lowered Tafel slope and overpotential in the etched MoS2 catalysts. In addition, the differences in catalytic performance were further confirmed by modifying the current path of the line pattern. The perpendicularly contacted bilayer MoS2 catalysts exhibited optimized catalytic performance with a low overpotential (329 mV at 100 mA cm−2) and a low Tafel slope of (72 mV dec−1), which is attributed to bottom-layer-induced facilitated charge transport in the catalytically active top layer in bilayer MoS2 catalysts. This effective strategy provides a controllable defect engineering technology for improving the catalytic performance in semiconducting 2D TMDC catalysts.

2 Results and discussion

Fig. 1a illustrates a schematic of the laser etching process of bilayer MoS2 catalysts. In previous studies, laser etching in 2D materials was attributed to the oxidation mediated thermal sublimation of crystals. Therefore, when the bilayer MoS2 was irradiated with a 532 nm laser with moderate intensity (∼12 mW μm−2) under ambient conditions, we were able to selectively etch only the top layer in the bilayer MoS2 catalysts.14,15 As a result, a line-pattern was formed in the top layer of the bilayer MoS2 through line scanning of the laser, and this line-pattern can expose the active Mo-sites at the top layer of the bilayer MoS2, as shown in Fig. 1b.16 Therefore, as shown in Fig. 1b, by adjusting the line density, the active site density at the basal plane can be modulated. In addition, the remaining bottom layer contributes to transporting the charge carriers to active sites, preventing catalytic performance degradation due to retarded electron transport. Increasing the supply of electrons also improves the overall hydrogen generation performance of MoS2.17,18 Through this controllable laser etching technique, catalytically active Mo-edge sites were created on the top layer of bilayer MoS2, while leaving the bottom layer intact and unaltered.
image file: d4ta03197k-f1.tif
Fig. 1 (a) Schematic illustration of the bilayer MoS2 etching technique with a 532 nm laser. (b) A schematic diagram of the process of generating hydrogen from the catalytically active etched area in the MoS2 bilayer.

To investigate the formation of the line-pattern in the bilayer MoS2 crystal, we utilized the Raman, photoluminescence (PL) and atomic force microscopy (AFM) analyses. At first, before laser etching, we confirmed the characteristics of chemical vapor deposition (CVD) grown bilayer MoS2 catalysts by using various spectroscopic analyses, as shown in Fig. S1. Line-patterned bilayer MoS2 was transferred on the target substrate, and the optical image of the sample is shown in Fig. 2a. The height profile between the top layer and etched area in the AFM mapping image (Fig. 2b) exhibited a thickness of 0.8 nm, which corresponds to the thickness of the MoS2 monolayer.19–23 Therefore, it was confirmed that only top layer MoS2 was etched through the precisely controlled laser etching technique.


image file: d4ta03197k-f2.tif
Fig. 2 (a) Optical microscopy image of etched bilayer MoS2 and (b) atomic force microscopy (AFM) mapping image. Inset exhibits the corresponding height profile. (c) Raman and (d) photoluminescence mapping image of etched bilayer MoS2. Layer-dependent (e) Raman and (f) PL spectra from a top layer, etched site, and the bottom layer of etched bilayer MoS2. The scale bar for all figures is 10 μm.

In this etching process, the laser induced deformation of the crystal structure was caused by thermal degradation.14,24,25 Therefore, under these harsh conditions, MoS2 may experience lattice deformation, additional chemical contamination, and excessive layer etching. To investigate the chemical state changes and residual strain effects in the bilayer MoS2 crystal during laser etching, we utilized Raman and PL spectroscopic analyses. Increasing the number of layers in MoS2 resulted in notable spectral shifts for both the E12g and A1g Raman modes (Fig. 2e). In Fig. 2e, there was no discernible difference observed in the A1g and E12g peaks between the etched site and the bottom layer, which implies that the only top layer was etched through a precisely controlled laser etching technique. However, in contrast to the peaks seen in the etched site and the bottom layer, each individual peak in the top layer showed an increased separation and intensity. This phenomenon can be attributed to several factors. The A1g peak's shift provided a clear indication of the enhanced stiffness in van der Waals interlayer interactions, attributed to the strengthening of restoring forces with increasing layer thickness.26,27 Meanwhile, the E12g peak's shift underscored the softening of the wavelength due to intensified Coulomb interaction effects. Additionally, as shown in Fig. 2c, an observable increase in Raman spectra intensity was correlated with the increasing number of layers, a phenomenon attributed to the optical interference that occurs as the laser penetrates the MoS2 structure.26,28,29 In addition, there was no difference in the full width at half maximum (FWHM) of the two A1g peaks, and there was no satellite peak, confirming that there was no chemical contamination or lattice strain in the etched MoS2. In the case of the PL mapping image, the dark image indicates the bilayer MoS2 crystal, which was attributed to indirect band gap nature induced non-radiative recombination in the bilayer MoS2 crystal (Fig. 2f).26,30 As shown in Fig. 2d and f, the etched area also showed similar intensity to the bottom layer MoS2 crystal, which indicates that the laser etching technique was successfully conducted for peeling only the top layer of the MoS2 crystal.31 Through the above processes, it was confirmed that selective etching was successfully performed without chemical damage and strain effect on the etched MoS2.19,32–38

To investigate the line density dependent catalytic behavior in line-patterned bilayer MoS2 catalysts, we examined the hydrogen evolution catalytic performance with a three-electrode system.7 In Fig. S2a, the electrochemical measurement setup was composed of a graphite rod (counter electrode), Ag/AgCl electrode (reference electrode), and line-patterned MoS2 bilayer (working electrode). To compare the catalytic activity of line-patterned bilayer MoS2 catalysts, we open specific areas (59.1 μm2) of the MoS2 catalysts by using photolithography, as shown in Fig. S2b. In the electrochemical measurements, as shown in Fig. 3a, the line density dependent catalytic activity was clearly observed in the linear sweep voltammetry (LSV) curve. In the case of pristine MoS2 (L0), the overpotential at 100 mA cm−2 was measured at 942 mV, which corresponds to the poor catalytic activity of the basal plane of MoS2 catalysts. However, after the line pattern was created, the overpotential at 100 mA cm−2 gradually decreased to 703 mV, 504 mV and 329 mV for 2-line (L2), 3-line (L3) and 5-line (L5) patterned bilayer MoS2 catalysts, respectively. The Tafel slope also gradually decreased with increasing line density in the Tafel plot (Fig. 3b), achieving its lowest value (72 mV dec−1) with L5-MoS2 catalysts. These results indicate that the catalytic activity in bilayer MoS2 catalysts was gradually improved depending on the number of lines at the basal plane, created through laser etching.


image file: d4ta03197k-f3.tif
Fig. 3 Electrochemical hydrogen evolution performance of line-patterned bilayer MoS2 catalysts. (a) Polarization curve, (b) Tafel plot and (c) Nyquist plot when the number of etched lines is 1 to 5. (d) Etched line dependent overpotential and Tafel slope.

The line density dependent charge exchange process at the catalyst/electrolyte interface was further investigated through electrochemical impedance spectroscopy (EIS). The Nyquist plot in Fig. 3c shows a remarkable change of charge transfer resistance (Rct) between pristine (L0) and line patterned (L2, L3, L5) bilayer MoS2 catalysts. The pristine MoS2 catalysts measured high Rct at 46.7 Ω cm2, which indicates poor electrochemical activity in the basal plane of MoS2 catalysts. In contrast, the Rct gradually decreased when the line density increased. Compared with pristine MoS2 (L0-MoS2), L5-MoS2 showed the lowest Rct with 4.84 Ω cm2. These results indicate that the laser etching-induced activated MoS2 basal plane effectively exchanges the charge carriers with the ions in the electrolyte, resulting in reduced charge transfer resistance.11,39–41

In respect of etched MoS2 catalysts, excessive etching leads to degradation of the catalytic performance and electrical conductivity, despite increased catalytically active site density.40,42,43 Therefore, effective charge carrier injection into catalytically active sites is crucial for activating etched sites. Thus, we utilized bilayer MoS2 catalysts for investigating the current path dependent catalytic behavior in line-patterned bilayer MoS2 catalysts. As shown in Fig. 4a and b, we fabricated two types of devices with different current paths: (1) parallel-contact: the etched lines are parallel to the electrode. Therefore, the charge carriers were injected from the bottom layer to the line-patterned top layer of the MoS2 catalysts and (2) perpendicular-contact: the etched lines are perpendicular to the electrode (Fig. S3). Therefore, the charge carriers were injected directly into both the bottom and top layers of MoS2 catalysts. In electrochemical measurements, 5-line-patterned MoS2 (5L-MoS2) catalysts exhibited contact direction dependent catalytic behavior. The perpendicularly contacted MoS2 catalysts exhibited better catalytic performance compared to parallel contacted MoS2 catalysts. In Fig. 4c, the overpotential was measured at 136 mV and 326 mV at 10 mA cm−2 for perpendicularly and parallel contacted MoS2 catalysts, respectively. Furthermore, the Tafel slope also exhibited consistent behavior in both catalysts (Fig. 4d). The perpendicularly contacted MoS2 catalysts exhibited a decreased Tafel slope with 72 mV dec−1 compared to parallel contacted MoS2 catalysts (134 mV dec−1). EIS measurements also showed lowered Rct in perpendicularly contacted MoS2 catalysts (4.84 Ω cm2). In parallel contacted catalysts, the charge carriers are only injected through bottom layer MoS2. In this case, the charge carriers transport through interlayer gaps between bottom layer and top layer MoS2 catalysts. Therefore, the interlayer gap induced high potential barrier leads to hopping transport from the bottom layer to the etched MoS2 site, which corresponds to high resistance and sluggish charge transport.44 The sluggish supply of charge carriers limits the activation of catalytically active etched sites in bilayer MoS2 catalysts, which was related to the high overpotential and large charge transfer resistance in electrochemical measurements. Moreover, the etched line, which was far from the electrode, was hard to supply sufficient charge carriers from the electrode.


image file: d4ta03197k-f4.tif
Fig. 4 The schematic illustration of contact direction dependent device architecture. (a) Parallel contacted and (b) perpendicularly contacted MoS2 electrochemical cells. (c) Polarization curve, (d) Tafel plot and (e) Nyquist plot dependent when the etching direction and electrode direction are vertical and horizontal.

In contrast, in the case of perpendicularly contacted catalysts, the charge carriers transport from the electrode to both layers (bottom layer and etched layer) in bilayer MoS2 catalysts. Therefore, the supply of sufficient charge carriers in perpendicularly contacted bilayer MoS2 accelerates the electrochemical reaction at the catalytically active etched sites in bilayer MoS2 catalysts. The facilitated charge transport induced rapid electrochemical reaction at the catalyst/electrolyte interface gave rise to low charge transfer resistance, and this catalytic behavior was attributed to low in-plane resistance in the MoS2 crystal. Moreover, the increased active edge sites in the basal plane leads to a facilitated hydrogen adsorption/desorption process, which corresponds to a lowered Tafel slope. As a result, we confirmed that the optimization of catalytic activity at the etched active sites was realized through efficient charge transport through the line pattern in the top layer of bilayer MoS2 catalysts.7

3 Conclusions

In this research, we verified that the catalytically inert MoS2 basal plane can be activated through a simple laser etching technique under ambient conditions. Successful etching of only the top layer of the bilayer MoS2 crystal was realized through moderate layer intensity, which was verified by using AFM, Raman, and PL analyses. As a result, it was confirmed that a clean etching process without any change in the chemical state was successfully implemented. Exposing the catalytically active Mo sites at the basal plane through a laser etching technique showed an improvement in catalytic performance as the line density increased. Moreover, the optimized catalytic performance was realized by a vertically contacted electrode, which effectively injected charge carriers through the catalytically active line-pattern in bilayer MoS2 catalysts. In summary, our research proposed an efficient method to improve the catalytic behavior of 2D TMDC materials, while analyzing the effect of charge injection characteristics at the active sites on catalytic performance.

4 Experimental section

4.1 MoS2 synthesis

To precisely control the growth layers of MoS2 crystals, we employed a solution-processed synthesis method through a chemical vapor deposition (CVD) process. Molybdenum oxide (MoO3) powder was dissolved in ammonium hydroxide (NH4OH) solution. We utilized a two-zone furnace CVD system. Sulfur powder (Sigma-Aldrich) was placed upstream and heated to 200 °C. The MoO3 solution, along with a Si/SiO2 substrate, was placed downstream. The growth process was carried out at 800 °C under a continuous flow of Ar + H2 gas at 65 sccm for 20 minutes.20,22,34

4.2 MoS2 characterization and line-patterning method

The atomic force microscopy (AFM) images were obtained using a commercial AFM setup (XE7, Park Systems). The Raman and PL measurements were carried out using a 532 nm laser on a Witec Confocal Raman spectrometer. Optical spectroscopic measurements were carried out using a 532 nm wavelength laser at an intensity of 2 mW with an irradiation time of 60 s. Laser etching was conducted with a laser intensity of 14 mW and an irradiation time of 0.1 s μm−1 at room temperature and air ambient.

4.3 Electrochemical measurements

The fabricated microcell was measured in a three-electrode system. Photolithography was performed so that only the etched part of MoS2 reached the electrolyte. Thereafter, 0.5 M H2SO4 solution was dropped on the sample to perform measurements. The polarization curve was measured at a scan rate of 5 mV s−1 when it was 0 to −1 V compared to the reference electrode.
ERHE = E(Ag/AgCl) + 218 mV
With the above equation, the reversible hydrogen electrode (ERHE) was corrected in consideration of the potential of 0.5 M H2SO4, which is a reference electrode and an acid electrolyte.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

This work was supported by the National Research Foundation (NRF) of Korea (RS-2023-00220471 and RS-2024-00355424).

References

  1. L. X. Lin, P. Sherrell, Y. Q. Liu, W. Lei, S. W. Zhang, H. J. Zhang, G. G. Wallace and J. Chen, Adv. Energy Mater., 2020, 10, 1903870 CrossRef CAS.
  2. Q. Fu, J. C. Han, X. J. Wang, P. Xu, T. Yao, J. Zhong, W. W. Zhong, S. W. Liu, T. L. Gao, Z. H. Zhang, L. L. Xu and B. Song, Adv. Mater., 2021, 33, 1907818 CrossRef CAS.
  3. Q. P. Lu, Y. F. Yu, Q. L. Ma, B. Chen and H. Zhang, Adv. Mater., 2016, 28, 1917–1933 CrossRef CAS PubMed.
  4. B. Mohanty, P. Bhanja and B. K. Jena, Mater. Today Energy, 2022, 23, 100902 CrossRef CAS.
  5. P. Bhanja, B. Mohanty, S. Chongdar, A. Bhaumik, B. K. Jena and S. Basu, ACS Appl. Energy Mater., 2021, 4, 12827–12835 CrossRef CAS.
  6. R. I. Mohanty, L. Pradhan, S. Chongdar, S. Basu, P. Bhanja and B. K. Jena, Catal. Today, 2023, 424, 113789 CrossRef CAS.
  7. D. Voiry, R. Fullon, J. E. Yang, C. D. C. E. Silva, R. Kappera, I. Bozkurt, D. Kaplan, M. J. Lagos, P. E. Batson, G. Gupta, A. D. Mohite, L. Dong, D. Q. Er, V. B. Shenoy, T. Asefa and M. Chhowalla, Nat. Mater., 2016, 15, 1003–1009 CrossRef CAS PubMed.
  8. M. R. Gao, M. K. Y. Chan and Y. G. Sun, Nat. Commun., 2015, 6, 7493 CrossRef PubMed.
  9. C. Tsai, H. Li, S. Park, J. Park, H. S. Han, J. K. Norskov, X. L. Zheng and F. Abild-Pedersen, Nat. Commun., 2017, 8, 1–8 CrossRef PubMed.
  10. J. Lim, T. Kim, J. Byeon, K. H. Park, J. Hong, S. Pak and S. Cha, J. Mater. Chem. A, 2022, 10, 23274–23281 RSC.
  11. J. Xu, G. L. Shao, X. Tang, F. Lv, H. Y. Xiang, C. F. Jing, S. Liu, S. Dai, Y. G. Li, J. Luo and Z. Zhou, Nat. Commun., 2022, 13, 2193 CrossRef CAS PubMed.
  12. Z. L. Zheng, L. Yu, M. Gao, X. Y. Chen, W. Zhou, C. Ma, L. H. Wu, J. F. Zhu, X. Y. Meng, J. T. Hu, Y. C. Tu, S. S. Wu, J. Mao, Z. Q. Tian and D. H. Deng, Nat. Commun., 2020, 11, 3315 CrossRef CAS PubMed.
  13. Y. Z. Zhang, J. Du, Z. Q. Wang, M. Luo, Y. Tian, T. Fujita, Q. K. Xue and M. W. Chen, ACS Appl. Energy Mater., 2018, 1, 2183–2191 CrossRef CAS.
  14. Y. Rho, J. Y. Pei, L. T. Wang, Z. L. Su, M. Eliceiri and C. P. Grigoropoulos, ACS Appl. Mater. Interfaces, 2019, 11, 39385–39393 CrossRef CAS PubMed.
  15. L. L. Hu, X. Y. Shan, Y. L. Wu, J. M. Zhao and X. H. Lu, Sci. Rep., 2017, 7, 15538 CrossRef PubMed.
  16. A. Castellanos-Gomez, M. Barkelid, A. M. Goossens, V. E. Calado, H. S. J. van der Zant and G. A. Steele, Nano Lett., 2012, 12, 3187–3192 CrossRef CAS PubMed.
  17. D. J. Li, U. N. Maiti, J. Lim, D. S. Choi, W. J. Lee, Y. Oh, G. Y. Lee and S. O. Kim, Nano Lett., 2014, 14, 1228–1233 CrossRef CAS.
  18. H. R. Cai, L. F. Xiong, B. Wang, D. L. Zhu, H. J. Hao, X. J. Yu, C. Li and S. C. Yang, Chem. Eng. J., 2022, 430, 132824 CrossRef CAS.
  19. S. Pak, A. R. Jang, J. Lee, J. Hong, P. Giraud, S. Lee, Y. Cho, G. H. An, Y. W. Lee, H. S. Shin, S. M. Morris, S. Cha, J. I. Sohn and J. M. Kim, Nanoscale, 2019, 11, 4726–4734 RSC.
  20. J. Lee, S. Pak, P. Giraud, Y. W. Lee, Y. Cho, J. Hong, A. R. Jang, H. S. Chung, W. K. Hong, H. Y. Jeong, H. S. Shin, L. G. Occhipinti, S. M. Morris, S. Cha, J. I. Sohn and J. M. Kim, Adv. Mater., 2017, 29, 1702206 CrossRef PubMed.
  21. S. Pak, Y. Cho, J. Hong, J. Lee, S. Lee, B. Hou, G. H. An, Y. W. Lee, J. E. Jang, H. Im, S. M. Morris, J. I. Sohn, S. Cha and J. M. Kim, ACS Appl. Mater. Interfaces, 2018, 10, 38264–38271 CrossRef CAS.
  22. J. Lee, S. Pak, Y. W. Lee, Y. Park, A. R. Jang, J. Hong, Y. Cho, B. Hou, S. Lee, H. Y. Jeong, H. S. Shin, S. M. Morris, S. Cha, J. I. Sohn and J. M. Kim, ACS Nano, 2019, 13, 13047–13055 CrossRef CAS.
  23. B. Kim, S. Pak, H. W. Choi, Y. J. Choi, A. R. Jang, J. Lee, Y. T. Chun, S. Cha and J. I. Sohn, 2D Mater., 2019, 6, 025017 CrossRef CAS.
  24. E. Y. Gu, Q. Y. Wang, Y. W. Zhang, C. X. Cong, L. G. Hu, P. F. Tian, R. Liu, S. L. Zhang and Z. J. Qiu, AIP Adv., 2017, 7, 125329 CrossRef.
  25. R. Rani, D. Dimple, N. Jena, A. Kundu, A. De Sarkar and K. S. Hazra, Appl. Phys. Lett., 2017, 110, 083101 CrossRef.
  26. S. M. Shinde, K. P. Dhakal, X. Chen, W. S. Yun, J. Lee, H. Kim and J. H. Ahn, NPG Asia Mater., 2018, 10, e468 CrossRef CAS.
  27. B. Chakraborty, H. S. S. R. Matte, A. K. Sood and C. N. R. Rao, J. Raman Spectrosc., 2013, 44, 92–96 CrossRef CAS.
  28. H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier and D. Baillargeat, Adv. Funct. Mater., 2012, 22, 1385–1390 CrossRef CAS.
  29. C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone and S. Ryu, ACS Nano, 2010, 4, 2695–2700 CrossRef CAS PubMed.
  30. A. Splendiani, L. Sun, Y. B. Zhang, T. S. Li, J. Kim, C. Y. Chim, G. Galli and F. Wang, Nano Lett., 2010, 10, 1271–1275 CrossRef CAS PubMed.
  31. S. Kang, Y. S. Kim, J. H. Jeong, J. Kwon, J. H. Kim, Y. Jung, J. C. Kim, B. Kim, S. Hyun, P. Y. Huang, J. C. Hone, H. Y. Jeong, J. W. Park, C. H. Lee and G. H. Lee, ACS Appl. Mater. Interfaces, 2021, 13, 1245–1252 CrossRef CAS PubMed.
  32. S. Pak, S. Kim, J. Lim, T. Kim, K. H. Park and S. Cha, J. Phys. Chem. C, 2023, 127, 4689–4695 CrossRef CAS.
  33. S. Pak, J. Lee, Y. W. Lee, A. R. Jang, S. Ahn, K. Y. Ma, Y. Cho, J. Hong, S. Lee, H. Y. Jeong, H. Im, H. S. Shin, S. M. Morris, S. Cha, J. I. Sohn and J. M. Kim, Nano Lett., 2017, 17, 5634–5640 CrossRef CAS.
  34. S. W. Jung, S. Pak, S. Lee, S. Reimers, S. Mukherjee, P. Dudin, T. K. Kim, M. Cattelan, N. Fox, S. S. Dhesi, C. Cacho and S. Cha, Appl. Surf. Sci., 2020, 532, 147390 CrossRef CAS.
  35. S. Pak, J. Lee, A. R. Jang, S. Kim, K. H. Park, J. I. Sohn and S. Cha, Adv. Funct. Mater., 2020, 30, 2002023 CrossRef CAS.
  36. S. Pak, J. Lim, J. Hong and S. Cha, Catalysts, 2021, 11, 70 CrossRef CAS.
  37. S. Pak, S. Jang, T. Kim, J. Lim, J. S. Hwang, Y. Cho, H. Chang, A. R. Jang, K. H. Park, J. Hong and S. Cha, Adv. Mater., 2021, 33, 2102091 CrossRef CAS PubMed.
  38. S. Pak, J. Son, T. Kim, J. Lim, J. Hong, Y. Lim, C. J. Heo, K. B. Park, Y. W. Jin, K. H. Park, Y. Cho and S. Cha, Nanotechnology, 2023, 34, 015702 CrossRef CAS PubMed.
  39. A. Bala, A. Sen, Y. H. Kim, Y. M. Kim, S. Gandla, H. Park and S. Kim, J. Phys. Chem. C, 2022, 126, 9696–9703 CrossRef CAS.
  40. X. Wang, Y. W. Zhang, H. N. Si, Q. H. Zhang, J. Wu, L. Gao, X. F. Wei, Y. Sun, Q. L. Liao, Z. Zhang, K. Ammarah, L. Gu, Z. Kang and Y. Zhang, J. Am. Chem. Soc., 2020, 142, 4298–4308 CrossRef CAS PubMed.
  41. Z. G. Wang, Q. Li, H. X. Xu, C. Dahl-Petersen, Q. Yang, D. J. Cheng, D. P. Cao, F. Besenbacher, J. V. Lauritsen, S. Helveg and M. D. Dong, Nano Energy, 2018, 49, 634–643 CrossRef CAS.
  42. G. L. Ye, Y. J. Gong, J. H. Lin, B. Li, Y. M. He, S. T. Pantelides, W. Zhou, R. Vajtai and P. M. Ajayan, Nano Lett., 2016, 16, 1097–1103 CrossRef CAS PubMed.
  43. J. F. Xie, J. J. Zhang, S. Li, F. Grote, X. D. Zhang, H. Zhang, R. X. Wang, Y. Lei, B. C. Pan and Y. Xie, J. Am. Chem. Soc., 2013, 135, 17881–17888 CrossRef CAS.
  44. Y. F. Yu, S. Y. Huang, Y. P. Li, S. N. Steinmann, W. T. Yang and L. Y. Cao, Nano Lett., 2014, 14, 553–558 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available: Characterization section, AFM mapping image of bilayer MoS2, Raman and PL spectra mapping of CVD grown MoS2 bilayers, image of the setup for electrochemical analysis, optical image of the etched MoS2 catalyst. See DOI: https://doi.org/10.1039/d4ta03197k
These authors contributed equally.

This journal is © The Royal Society of Chemistry 2024