Next Article in Journal
ZnO Nanoparticles with Controllable Ce Content for Efficient Photocatalytic Degradation of MB Synthesized by the Polyol Method
Next Article in Special Issue
Emerging Energy Harvesting Technology for Electro/Photo-Catalytic Water Splitting Application
Previous Article in Journal
Editorial: Special Issue on “Flavin Monooxygenases”
Previous Article in Special Issue
Reducing the Photodegradation of Perovskite Quantum Dots to Enhance Photocatalysis in CO2 Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Hydrogen Evolution Reaction in Surface Functionalized MoS2 Monolayers

1
Department of Physics, Sungkyunkwan University (SKKU), Suwon 16419, Korea
2
School of Materials Science and Engineering, Kookmin University, Seoul 02707, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2021, 11(1), 70; https://doi.org/10.3390/catal11010070
Submission received: 28 December 2020 / Revised: 4 January 2021 / Accepted: 4 January 2021 / Published: 6 January 2021

Abstract

:
Monolayered, semiconducting MoS2 and their transition metal dichalcogenides (TMDCs) families are promising and low-cost materials for hydrogen generation through electrolytes (HER, hydrogen evolution reaction) due to their high activities and electrochemical stability during the reaction. However, there is still a lack of understanding in identifying the underlying mechanism responsible for improving the electrocatalytic properties of theses monolayers. In this work, we investigated the significance of controlling carrier densities in a MoS2 monolayer and in turn the corresponding electrocatalytic behaviors in relation to the energy band structure of MoS2. Surface functionalization was employed to achieve p-doping and n-doping in the MoS2 monolayer that led to MoS2 electrochemical devices with different catalytic performances. Specifically, the electron-rich MoS2 surface showed lower overpotential and Tafel slope compared to the MoS2 with surface functional groups that contributed to p-doping. We attributed such enhancement to the increase in the carrier density and the corresponding Fermi level that accelerated HER and charge transfer kinetics. These findings are of high importance in designing electrocatalysts based on two-dimensional TMDCs.

Graphical Abstract

1. Introduction

Hydrogen (H2) is emerging to be an essential energy solution for a next-generation and sustainable energy system that can replace fossil fuels due to its clean, renewable, and affordable characteristics [1,2]. Nevertheless, the majority of hydrogen fuels is produced from fossil fuels by reforming procedures. In order to avoid fossil fuels in hydrogen production, a number of hydrogen production methods are in development, especially in eco-friendly ways. One environment-friendly method to produce H2 is by electrochemically evolving H2 by water splitting using sunlight or electricity generated by wind turbines or solar cells [3]. Splitting water molecules into hydrogen occurs by electrochemical hydrogen evolution reaction (HER, 2H+ + 2e → H2) on the surface of the catalyst’s electrodes, and the efficiency of hydrogen production can be increased by decreasing the overpotential needed to drive the hydrogen evolution reaction (HER) [4]. Current hydrogen production from water splitting generally relies on noble metal catalysts. In particular, platinum (Pt)-based catalysts are known to display an excellent catalytic performance with its balanced hydrogen adsorption/desorption energy, small overpotential, and long cycling stability in terms of favorable HER behavior. However, due to high cost and limited supply of platinum, extensive efforts have been paid to seek high efficiency and cost-effective non-noble metal alternatives such as transition metal sulfides and oxides, nickel alloys, and so forth [5].
Molybdenum disulfide (MoS2), which is mostly studied among a hexagonally packed layered structure of transition metal dichalcogenides (TMDCs), has been considered to be a promising candidate for catalyzing electrochemical hydrogen production from water due to its relatively low cost, its element abundance in earth, high catalytic activity, and good electrochemical stability [6,7,8]. Up to now, numerous studies have focused on improving the catalytic performance of MoS2 by increasing the concentration of exposed sulfur edges (catalytic active sties) and constructing the hierarchical morphology and structure of MoS2 [1,4,9,10,11,12]. In particular, the MoS2 monolayer shows a dramatic increase in catalytic performance compared to its bulk counterpart, which can be attributed to the large density of unsaturated molybdenum in basal planes and edge sites as well as lower internal resistance than its bulk counterpart, that subsequently lead to efficient charge transfer kinetics [13,14].
Due to its semiconducting nature, the HER behaviors of the MoS2 monolayer can be significantly affected by the change in energy level of MoS2, which can then lead to different charge transport kinetics. Simultaneously, controlling the energy level of the MoS2 monolayer can result in different HER behavior by modifying the charge exchange process between the electrocatalyst and protons as well as the charge transfer dynamics between the electrode and the electrocatalyst [5,14]. Consequently, the changes in the electronic band structure of the MoS2 monolayer can strongly affect the overall electrocatalytic performance of MoS2 catalysts. However, so far, relatively little attention has been paid to methods that modify the electrocatalytic performance of the MoS2 monolayer and to the role of the modulation of energy level of MoS2 in HER.
In this work, we investigated the influence of the surface-chemistry on the modulation of the charge carrier density of the MoS2 monolayer associated with Fermi level position and the corresponding electrocatalytic behaviors. To demonstrate a clear relationship between the surface chemistry, fermi-level position, and corresponding HER performance, a self-assembled monolayer (SAM) was employed to effectively donate and withdraw electrons from the MoS2 monolayer. We found that the surface modification to the MoS2 monolayer effectively modulated charge carrier density and the resulting Fermi-level position, which was confirmed through Raman and photoluminescence (PL) measurement as well as electrical measurement by fabricating back-gated field-effect transistors (FETs). The catalytic performance of the surface modified MoS2 monolayer was measured by contact with the Ti/Au electrode with on-chip device configuration and exposing the MoS2 surface only to the electrolyte. We demonstrated efficient catalytic activities of the MoS2 monolayer when the Fermi-level of MoS2 shifted upward, which was observed through improved onset potential, decreased overpotential, and decreased Tafel slope. We attributed such enhancement to the higher density of free carriers that translated into higher conductivity and better charge transfer kinetics. These findings present an important pathway toward designing catalysts based on TMDC monolayers.

2. Results and Discussion

Monolayer MoS2 was directly grown on a 300 nm SiO2/Si substrate using our thermal chemical vapor deposition (CVD) method [15,16], as shown in Figure S1a, and the resulting optical image of MoS2 is shown in Figure S1b. In order to find whether the synthesized MoS2 was monolayered, Raman and PL analysis was conducted as shown in Figure S2. Figure S2a shows the two characteristic Raman peaks measured around 383 cm−1 and 402 cm−1, corresponding to in-plane (E12g) and out-of-plane (A1g) vibrational modes, respectively. The difference between the two modes was found to be around 19 cm−1, which confirmed that the synthesized MoS2 was single-layered [17]. We also measured PL as shown in Figure S2b, and the distinct strong PL emission was measured, which corresponded to the direct bandgap nature of a single-layer MoS2. In order to investigate the influence of the charge carrier density of the MoS2 monolayer on the electrocatalytic behaviors, the well-known SAM surface functionalization technique was employed in this study. Octadecyltrichlosilane (ODTS, electron withdrawing) and (3-Aminopropyl)triethoxysilane (APTES, electron donating) organic molecules were used as they can effectively tune the carrier concentrations, electrical, and optical properties of TMDC monolayers and other nanomaterials [18,19,20]. The surface functionalization of the as-grown MoS2 monolayers with ODTS and APTES was performed in a hexane solution containing the functionalized organic molecules in an argon-filled glove box as shown in Figure 1a,b. Note that SAM molecules donate and withdraw electrons through built-in molecular dipoles [19]. The charge transfer from and/or to organic molecules then effectively modulates the charge carrier densities of the MoS2 monolayer, which then results in the rise and/or lower Fermi energy level as shown in Figure 1c. Compared to pristine MoS2, the Fermi-level of APTES-functionalized MoS2 was shifted upward, while it shifted downward for ODTS-functionalized MoS2, and the changes in the carrier density and Fermi level are supported in the next paragraphs. It should be highlighted that the SAM technique is effective and a nondestructive method of changing carrier concentrations of the MoS2 monolayers.
In order to confirm and analyze the SAM functionalized-induced change of the carrier density using either ODTS or APTES, we first measured the change in the Raman and PL spectrum as they were strongly perturbed by the change in the carrier density of the MoS2 monolayers [18]. Raman spectra of the pristine MoS2 (pristine-MoS2) and functionalized MoS2 (APTES-MoS2, and ODTS-MoS2) were measured, which is shown in Figure 2a. We clearly observed that the A1g Raman mode of the functionalized MoS2 monolayers was dependent on the functionalized molecules, while the E12g mode was rather unaffected by the SAM doping. As shown in Figure 2b, the A1g vibrational mode of pristine was 400.2 cm−1, which then increased to 401 cm−1 for ODTS-MoS2 and decreased to 399.3 cm−1 for APTES-MoS2. Moreover, it was found that the A1g mode stiffened for ODTS-MoS2 and softened for APTES-MoS2, evidenced by the change in its linewidth. The doping-dependence of A1g vibrational mode of MoS2 can be understood as an electron–phonon coupling, in which electron doping leads to strong electron–phonon coupling, which then softens the A1g vibrational mode, and the trend shown in the results were in good agreement with previous studies [21].
The functionalized MoS2 monolayer was further characterized using PL spectroscopy. PL spectra of pristine-MoS2, ODTS-MoS2, and APTES-MoS2 were obtained (Figure 2c). It was clearly noticeable that the PL spectrum of ODTS-MoS2 was blue-shifted (6 nm), and the PL intensity was increased, while for the APTES-MoS2, there was a decrease in the PL intensity and the red-shift (7 nm) of the peak wavelength. Furthermore, the full-width at half-maximum (FWHM) of the PL peak of the ODTS-MoS2 layer was decreased, while it was largely increased for APTES-MoS2. These trends of changes in the PL of the MoS2 layer upon surface functionalization can be attributed to the change of carrier concentrations: more electrons in the MoS2 monolayer decrease the radiative recombination rates of exciton and makes recombination of trions dominant, and these results are consistent with previous reports that measured the carrier density-dependent PL of MoS2 [22].
In order to characterize the SAM functionalization-induced change in the charge carrier density and the corresponding Fermi level, electrical properties of MoS2 FETs were measured for the pristine and modified MoS2 FETs. The devices employed for this study was fabricated using photolithography, followed by the deposition of Ti/Au (5/45 nm) electrode layers as shown in Figure 3a. Figure 3b shows the representative drain-source current, Ids, versus gate voltage, Vg, on a logarithmic scale (transfer curve) at a drain-source voltage of Vds = 0.1 V. We could easily notice that the APTES-MoS2 device showed noticeably high on-current and conductance compared to the pristine-MoS2 device, while low on-current and conductance was measured for the ODTS-MoS2 device. These trends of conductance with respect to the SAM functionalization were in good agreement with Figure S3, which shows the representative drain-source current versus drain-source voltage at varying gate voltages (output curve). In addition, it was found that the threshold voltages were negatively shifted for APTES-MoS2 devices, while threshold voltages shifted toward more positive voltages for ODTS-MoS2. For statistical analysis of the devices, the resulting statistical chart of 10 control devices are shown in Figure 3. The results presented in the transfer and output curves demonstrate that SAM functionalization is effective in donating electrons (APTES) and withdrawing electrons (ODTS) from the MoS2 monolayer and consequently changes the Fermi energy level of the MoS2 monolayer. To further analyze and compare the change in the free carrier density, we calculated and compared the amount of change in carrier density when MoS2 devices were functionalized with SAM using the electrical results shown in Figure 3b and the parallel-plate capacitor model, N d o p i n g = C | V t h | e , where C is the capacitance of the HfO2 gate oxide layer (309.9 nF/cm2), V t h is the change in the threshold voltage of the functionalized devices with respect to the pristine device, and e is the elementary charge. The N d o p i n g was calculated to be 3.874   ×   10 12 and 2.905   ×   10 12 cm−2 for the APTES-MoS2 device and ODTS-MoS2 device, respectively, and was consistent with previously reported results [18,23,24,25]. Therefore, it is evident from the shift of threshold voltages and change in the carrier density that the SAM functionalization-induced doping of MoS2 was successfully employed, which consequently affected the Fermi energy level of theMoS2 monolayers.
Having confirmed the changes in the carrier density and Fermi level upon surface functionalization, we now focus on elucidating the role of the carrier density in the HER catalytic behaviors of the MoS2 monolayer. As shown in Figure 4a, the HER catalytic behaviors of the pristine and surface functionalized MoS2 monolayers were measured by aa three electrodes system, which is the setup typically employed to measure HER catalysis [26]. A probe station was used to electrically connect the working electrode (WE), counter electrode (CE), and reference electrode (RE). Ti/Au electrodes deposited on a chip was used as the contact and current collector. Droplet of electrolyte solution (0.5M H2SO4) was applied on the surface of MoS2 catalysts. In order to sorely investigate the catalytic activity of the monolayered MoS2, we used photoresist to open the MoS2 active sites only while the Ti/Au electrodes were covered by the photoresist, making the electrodes inert to the acid electrolyte as shown in Figure 4b. Note that this experimental design and employing an on-chip device in the electrochemical measurement made it possible to characterize 2D materials by accurately measuring the area of the active sites involved in the HER reaction [12]. Furthermore, electrochemically inert SiO2 was used as the substrate to exclude any HER contribution from the substrate and to sorely focus on the catalytic performance of the active MoS2.
Figure 4c shows the linear sweep curves of the pristine MoS2 monolayer and surface functionalized MoS2 monolayers at a scan rate of 10 mV/s. From the polarization curve, the distinctively different HER activities were noticed for the pristine and surface functionalized MoS2 monolayers. Compared to the catalytic behavior of the pristine monolayer MoS2, the catalytic performance was improved for APTES-MoS2, while it deteriorated for ODTS-MoS2, and each device performance was characterized for multiple cycles for the stability of functionalized molecules (Figure S4). The overpotential of the MoS2 monolayers, which is measured at the catalytic current density of 10 mA/cm2, were calculated to be 637 mV for ODTS-MoS2, 445 mV for pristine-MoS2, and 382 mV APTES-MoS2. Likewise, as shown in Figure 4d, the Tafel slope was calculated to be 162 mV/decade for ODTS-MoS2, 133 mV/decade for pristine-MoS2, and 110 mV/decade for APTES-MoS2. The increase in the carrier density and the rise of Fermi level close to the conduction band of the MoS2 monolayer makes them favorable for H2 production. Controlling the energy level of MoS2 can activate or deactivate the electrochemical HER kinetics by the surface functionalization of the MoS2 monolayer [5,14]. Furthermore, as evidenced in the output curve of Figure S3, the rise in Fermi level decreased the potential barrier between the electrode and the MoS2 catalyst, which facilitated the charge injection from the electrode to the MoS2 and resulted in enhanced catalytic activities of the MoS2 monolayer [14,27]. Thus, the increase in the carrier density and the corresponding Fermi level accelerate the HER kinetics and must be considered when designing the catalysts based on 2D semiconductors.

3. Materials and Methods

3.1. Synthesis of Monolayer MoS2

The MoS2 monolayer was directly grown on a SiO2 (300 nm)/Si substrate using our thermal CVD process. To summarize, inside a 2-inch tube furnace, a MoO3 precursor (0.05 mg, >99% Sigma Aldrich, Seoul, South Korea) and sulfur powders (200 mg, >99.98% Sigma Aldrich, Seoul, South Korea) on a alumina boat, which was placed on the upstream and downstream, respectively, were used to synthesize the MoS2 monolayers. The growth substrate (SiO2/Si) was placed faced down on the boat containing the MoO3 precursors. The CVD furnace was heated at a rate of 18.75 °C/min up to 750 °C in an argon-filled environment, and the temperature of the furnace was maintained for 10 min for the synthesis and cooled down to room temperature.

3.2. Formation of Octadecyltrichlorosilane (ODTS) and (3-Aminopropyl)-Triethoxysilane (APTES) on a MoS2 Monolayer

Octadecyltrichlorosilane (ODTS, Sigma Aldrich Seoul, South Korea, >90%) and (3-Aminopropyl)-triethoxysilane (APTES, Sigma Aldrich, Seoul, South Korea, 99 %) were used to surface functionalize the as-grown MoS2 in an argon-filled glove box by dipping the as-grown substrates into a 10 mL of hexane solution that contained 50 μL of ODTS or APTES. The substrates were soaked for 1 h and rinsed with hexane, followed by drying with nitrogen. The substrates were then put on a hot plate and baked at 120 °C inside a glove box.

3.3. Characterization of the Doped MoS2

The PL and Raman measurements were performed using a 532 nm laser on a Witec Confocal Raman Spectroscopy. Laser power of 20 μW and 1 mW were used for PL and Raman analysis, respectively. A ×50 objective lens, which corresponded to the spot size of around 1 μm2, was used.

3.4. Device Fabrication and Measurement

The polysterene (PS MW ~192,000, Sigma Aldrich, Seoul, South Korea) coated MoS2 monolayers were detached from the growth substrate using DI water. The detached MoS2/PS film was then transferred onto a 20 nm HfO2/Si substrate that was used as a global back gate. The electrode pads in the MoS2 FETs were patterned using a photolithography process and AZ 5214E MicroChem photoresist, followed by the deposition of titanium (5 nm) and gold (40 nm) layers using a thermal evaporator. After a lift-off process, another photolithography process was performed in order to open the active catalytic area in MoS2. A semiconductor characterization system (4200-SCS, Keithley, Solon, OH, USA) was used to measure the electrical properties of MoS2, and the catalytic activities of the MoS2 monolayers were measured using µAUTOLABIII. In our device measurement experimental setup, monolayer MoS2 acts as the working electrode and the Ti/Au (5 nm/45 nm) electrode works as a current collector. Ag/AgCl and Pt wire with the diameter of 0.5 mm were used as the reference and counter electrode, respectively, and the distance between the electrodes was set to be larger than 5 mm. The active area was calculated based on the area of the triangular MoS2 exposed through photolithography.

4. Conclusions

In summary, we investigated the role of charge carrier density associated with the Fermi energy level on the electrocatalytic behaviors of the MoS2 monolayer. The carrier density of MoS2 was modulated to be n-doped (APTES-MoS2) and p-doped (ODTS-MoS2) through the SAM doping technique, and the electrocatalytic performance was measured by fabricating an on-chip device that could accurately measure the catalytic activity of the MoS2 monolayer. The catalytic activity of the electron rich APTES-MoS2 was found to be enhanced, while it was deteriorated for ODTS-MoS2, suggesting that the increase in charge carrier density associated with the upshift of Fermi-level position helps the charge transfer kinetics that lead to decreased overpotential and Tafel slope. We believe that the results presented in this work are important in designing HER catalysts based on two-dimensional TMDCs.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/11/1/70/s1, Figure S1: (a) Illustration of the CVD synthesis process for a MoS2 monolayer. (b) An optical image of the as-grown MoS2 monolayers. Scale bar: 20 μm, Figure S2: (a) Raman and (b) PL spectrum of MoS2 monolayer, Figure S3: Output curves of (a) APTES-MoS2, (b) pristine-MoS2, (c) ODTS-MoS2, Figure S4: Polarization curves of (a) ODTS-MoS2 and (b) APTES-MoS2 plotted before and after 10 cycles while the electrolyte was replaced every cycle.

Author Contributions

S.P. and J.L. contributed equally to this work. S.P. and J.L. performed experiments and wrote the paper. J.H. and S.C. conceived and deigned the experiments. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation (NRF) of Korea (2019R1A2C1005930, 2020R1F1A1068979).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects Engineered Monolayer MoS2 for Improved Hydrogen Evolution Reaction. Nano Lett. 2016, 16, 1097–1103. [Google Scholar] [CrossRef]
  2. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.W.; Asefa, T.; Shenoy, V.B.; Eda, G.; Chhowalla, M. Conducting MoS2 Nanosheets as Catalysts for Hydrogen Evolution Reaction. Nano Lett. 2013, 13, 6222–6227. [Google Scholar] [CrossRef]
  3. Tsai, C.; Abild-Pedersen, F.; Norskov, J.K. Tuning the MoS2 Edge-Site Activity for Hydrogen Evolution via Support Interactions. Nano Lett. 2014, 14, 1381–1387. [Google Scholar] [CrossRef] [PubMed]
  4. Kibsgaard, J.; Jaramillo, T.F.; Besenbacher, F. Building an appropriate active-site motif into a hydrogen-evolution catalyst with thiomolybdate [Mo3S13](2-) clusters. Nat. Chem. 2014, 6, 248–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Shi, Y.; Zhou, Y.; Yang, D.R.; Xu, W.X.; Wang, C.; Wang, F.B.; Xu, J.J.; Xia, X.H.; Chen, H.Y. Energy Level Engineering of MoS2 by Transition-Metal Doping for Accelerating Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 15479–15485. [Google Scholar] [CrossRef] [PubMed]
  6. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jorgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Norskov, J.K. Biornimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef]
  7. Merki, D.; Hu, X.L. Recent developments of molybdenum and tungsten sulfides as hydrogen evolution catalysts. Energy Environ. Sci. 2011, 4, 3878–3888. [Google Scholar] [CrossRef] [Green Version]
  8. Benck, J.D.; Hellstern, T.R.; Kibsgaard, J.; Chakthranont, P.; Jaramillo, T.F. Catalyzing the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide Nanomaterials. ACS Catal. 2014, 4, 3957–3971. [Google Scholar] [CrossRef]
  9. Li, G.Q.; Zhang, D.; Qiao, Q.; Yu, Y.F.; Peterson, D.; Zafar, A.; Kumar, R.; Curtarolo, S.; Hunte, F.; Shannon, S.; et al. All The Catalytic Active Sites of MoS2 for Hydrogen Evolution. J. Am. Chem. Soc. 2016, 138, 16632–16638. [Google Scholar] [CrossRef]
  10. Xie, J.F.; Zhang, H.; Li, S.; Wang, R.X.; Sun, X.; Zhou, M.; Zhou, J.F.; Lou, X.W.; Xie, Y. Defect-Rich MoS2 Ultrathin Nanosheets with Additional Active Edge Sites for Enhanced Electrocatalytic Hydrogen Evolution. Adv. Mater. 2013, 25, 5807. [Google Scholar] [CrossRef]
  11. Li, H.; Tsai, C.; Koh, A.; Cai, L.; Contryman, A.W.; Fragapane, A.H.; Zhao, J.; Han, H.; Manoharan, H.C.; Abild-Pedersen, F.; et al. Activating and optimizing MoS2 basal planes for hydrogen evolution through the formation of strained sulphur vacancies. Nat. Mater. 2015, 15, 48–53. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, J.; Wu, J.J.; Guo, H.; Chen, W.B.; Yuan, J.T.; Martinez, U.; Gupta, G.; Mohite, A.; Ajayan, P.M.; Lou, J. Unveiling Active Sites for the Hydrogen Evolution Reaction on Monolayer MoS2. Adv. Mater. 2017, 29, 1701955. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, Y.F.; Huang, S.Y.; Li, Y.P.; Steinmann, S.N.; Yang, W.T.; Cao, L.Y. Layer-Dependent Electrocatalysis of MoS2 for Hydrogen Evolution. Nano Lett. 2014, 14, 553–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Voiry, D.; Fullon, R.; Yang, J.E.; Silva, C.D.C.E.; Kappera, R.; Bozkurt, I.; Kaplan, D.; Lagos, M.J.; Batson, P.E.; Gupta, G.; et al. The role of electronic coupling between substrate and 2D MoS2 nanosheets in electrocatalytic production of hydrogen. Nat. Mater. 2016, 15, 1003–1009. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, J.; Pak, S.; Giraud, P.; Lee, Y.W.; Cho, Y.; Hong, J.; Jang, A.R.; Chung, H.S.; Hong, W.K.; Jeong, H.; et al. Thermodynamically Stable Synthesis of Large-Scale and Highly Crystalline Transition Metal Dichalcogenide Monolayers and their Unipolar n–n Heterojunction Devices. Adv. Mater. 2017, 29, 1702206. [Google Scholar] [CrossRef]
  16. Sahoo, P.K.; Memaran, S.; Xin, Y.; Balicas, L.; Gutierrez, H.R. One-pot growth of two-dimensional lateral heterostructures via sequential edge-epitaxy. Nature 2018, 553, 63–67. [Google Scholar] [CrossRef] [Green Version]
  17. Jung, S.W.; Pak, S.; Lee, S.; Reimers, S.; Mukherjee, S.; Dudin, P.; Kim, T.K.; Cattelan, M.; Fox, N.; Dhesi, S.S.; et al. Spectral functions of CVD grown MoS2 monolayers after chemical transfer onto Au surface. Appl. Surf. Sci. 2020, 532, 147390. [Google Scholar] [CrossRef]
  18. Pak, S.; Jang, A.R.; Lee, J.; Hong, J.; Giraud, P.; Lee, S.; Cho, Y.; An, G.H.; Lee, Y.W.; Shin, H.S.; et al. Surface functionalization-induced photoresponse characteristics of monolayer MoS2 for fast flexible photodetectors. Nanoscale 2019, 11, 4726–4734. [Google Scholar] [CrossRef]
  19. Najmaei, S.; Zou, X.; Er, D.; Li, J.; Jin, Z.; Gao, W.; Zhang, Q.; Park, S.; Ge, L.; Lei, S.; et al. Tailoring the Physical Properties of Molybdenum Disulfide Monolayers by Control of Interfacial Chemistry. Nano Lett. 2014, 14, 1354–1361. [Google Scholar] [CrossRef]
  20. Kobayashi, S.; Nishikawa, T.; Takenobu, T.; Mori, S.; Shimoda, T.; Mitani, T.; Shimotani, H.; Yoshimoto, N.; Ogawa, S.; Iwasa, Y. Control of carrier density by self-assembled monolayers in organic field-effect transistors. Nat. Mater. 2004, 3, 317–322. [Google Scholar] [CrossRef]
  21. Chakraborty, B.; Bera, A.; Muthu, D.V.S.; Bhowmick, S.; Waghmare, U.V.; Sood, A.K. Symmetry-dependent phonon renormalization in monolayer MoS2 transistor. Phys. Rev. B 2012, 85, 161403. [Google Scholar] [CrossRef] [Green Version]
  22. Mak, K.; He, K.; Lee, C.; Lee, G.; Hone, J.; Heinz, T.F.; Shan, J. Tightly bound trions in monolayer MoS2. Nat. Mater. 2012, 12, 207–211. [Google Scholar] [CrossRef] [PubMed]
  23. Li, Y.; Xu, C.-Y.Y.; Hu, P.; Zhen, L. Carrier control of MoS2 nanoflakes by functional self-assembled monolayers. ACS Nano 2013, 7, 7795–7804. [Google Scholar] [CrossRef] [PubMed]
  24. Kang, D.H.; Kim, M.S.; Shim, J.; Jeon, J.; Park, H.Y.; Jung, W.S.; Yu, H.Y.; Pang, C.H.; Lee, S.; Park, J.H. High-Performance Transition Metal Dichalcogenide Photodetectors Enhanced by Self-Assembled Monolayer Doping. Adv. Funct. Mater. 2015, 25, 4219–4227. [Google Scholar] [CrossRef]
  25. Pak, S.; Cho, Y.; Hong, J.; Lee, J.; Lee, S.; Hou, B.; An, G.H.; Lee, Y.W.; Jang, J.E.; Im, H.; et al. Consecutive Junction-Induced Efficient Charge Separation Mechanisms for High-Performance MoS2/Quantum Dot Phototransistors. ACS Appl. Mater. Interfaces 2018, 10, 38264–38271. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, J.; Yan, M.; Zhao, K.; Liao, X.; Wang, P.; Pan, X.; Yang, W.; Mai, L. Field Effect Enhanced Hydrogen Evolution Reaction of MoS2 Nanosheets. Adv. Mater. 2017, 29, 1604464. [Google Scholar] [CrossRef]
  27. Pak, S.; Lee, J.; Jang, A.-R.; Kim, S.; Park, K.-H.; Sohn, J.I.; Cha, S. Strain Engineering of Contact Energy Barriers and Photoresponse Behaviors in Monolayer MoS2 Flexible Devices. Adv. Funct. Mater. 2020, 22, 2002023. [Google Scholar] [CrossRef]
Figure 1. (a) A schematic illustration of the surface functionalization of either ODTS (Octadecyltrichlosilane) or APTES ((3-Aminopropyl)triethoxysilane) in a hexane solution. (b) ODTS and APTES molecules that withdraw and donate electrons from the MoS2 monolayer. (c) Energy band diagram of MoS2 monolayer depicting the change in the Fermi-level upon surface functionalization with either ODTS or APTES.
Figure 1. (a) A schematic illustration of the surface functionalization of either ODTS (Octadecyltrichlosilane) or APTES ((3-Aminopropyl)triethoxysilane) in a hexane solution. (b) ODTS and APTES molecules that withdraw and donate electrons from the MoS2 monolayer. (c) Energy band diagram of MoS2 monolayer depicting the change in the Fermi-level upon surface functionalization with either ODTS or APTES.
Catalysts 11 00070 g001
Figure 2. Raman and photoluminesence characterization of MoS2 monolayer functionalized with ODTS and APTES. (a) Raman and (c) PL spectra taken from pristine and surface functionalized MoS2. (b) Raman peak (E12g mode and A1g mode) shift and (d) PL peak shift (primary axis) and full width at half maximum (secondary axis) with respect to the functionalized molecules.
Figure 2. Raman and photoluminesence characterization of MoS2 monolayer functionalized with ODTS and APTES. (a) Raman and (c) PL spectra taken from pristine and surface functionalized MoS2. (b) Raman peak (E12g mode and A1g mode) shift and (d) PL peak shift (primary axis) and full width at half maximum (secondary axis) with respect to the functionalized molecules.
Catalysts 11 00070 g002
Figure 3. Effect of surface functionalization on the carrier density of the MoS2 monolayer. (a) An optical image of the field effect transistors based on the MoS2 monolayer fabricated using photolithography. (b) Transfer curve at 0.1 V drain-source voltage acquired from the devices. (c) A change in threshold voltages for 10 control devices before and after surface functionalization. Note that threshold voltages shifted toward more negative voltages for the APTES-MoS2 devices and shifted toward more positive voltages for ODTS-MoS2 devices.
Figure 3. Effect of surface functionalization on the carrier density of the MoS2 monolayer. (a) An optical image of the field effect transistors based on the MoS2 monolayer fabricated using photolithography. (b) Transfer curve at 0.1 V drain-source voltage acquired from the devices. (c) A change in threshold voltages for 10 control devices before and after surface functionalization. Note that threshold voltages shifted toward more negative voltages for the APTES-MoS2 devices and shifted toward more positive voltages for ODTS-MoS2 devices.
Catalysts 11 00070 g003
Figure 4. Electrochemical setup for measuring the HER activity of the MoS2 monolayers. (a) Photograph of the electrochemical device. (b) An optical microscope image of the electrochemical device. The Ti/Au electrode was covered by photoresists to make it inert to catalytic activities. (c) Polarization curves and (d) Tafel plots measured for pristine and surface-functionalized MoS2 monolayers.
Figure 4. Electrochemical setup for measuring the HER activity of the MoS2 monolayers. (a) Photograph of the electrochemical device. (b) An optical microscope image of the electrochemical device. The Ti/Au electrode was covered by photoresists to make it inert to catalytic activities. (c) Polarization curves and (d) Tafel plots measured for pristine and surface-functionalized MoS2 monolayers.
Catalysts 11 00070 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pak, S.; Lim, J.; Hong, J.; Cha, S. Enhanced Hydrogen Evolution Reaction in Surface Functionalized MoS2 Monolayers. Catalysts 2021, 11, 70. https://doi.org/10.3390/catal11010070

AMA Style

Pak S, Lim J, Hong J, Cha S. Enhanced Hydrogen Evolution Reaction in Surface Functionalized MoS2 Monolayers. Catalysts. 2021; 11(1):70. https://doi.org/10.3390/catal11010070

Chicago/Turabian Style

Pak, Sangyeon, Jungmoon Lim, John Hong, and SeungNam Cha. 2021. "Enhanced Hydrogen Evolution Reaction in Surface Functionalized MoS2 Monolayers" Catalysts 11, no. 1: 70. https://doi.org/10.3390/catal11010070

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