Skip to main content
ACS Omega logoLink to ACS Omega
. 2019 Dec 3;4(25):21509–21515. doi: 10.1021/acsomega.9b03205

Nitrogen-Plasma-Treated Continuous Monolayer MoS2 for Improving Hydrogen Evolution Reaction

Anh Duc Nguyen , Tri Khoa Nguyen †,, Chinh Tam Le , Sungdo Kim , Farman Ullah , Yangjin Lee §, Sol Lee §, Kwanpyo Kim §, Dooyong Lee , Sungkyun Park , Jong-Seong Bae , Joon I Jang #,*, Yong Soo Kim †,*
PMCID: PMC6921679  PMID: 31867547

Abstract

graphic file with name ao9b03205_0003.jpg

Theoretically, the edges of a MoS2 flake and S-vacancy within the lattice have nearly zero Gibbs free energy for hydrogen adsorption, which is essentially correlated to the exchange currents in hydrogen evolution reaction (HER). However, MoS2 possesses insufficient active sites (edges and S-vacancies) in pristine form. Interestingly, active sites can be effectively engineered within the continuous MoS2 sheets by treating it with plasma in a controlled manner. Here, we employed N2 plasma on a large-area continuous-monolayer MoS2 synthesized via metal–organic chemical vapor deposition to acquire maximum active sites that are indeed required for an efficient HER performance. The MoS2 samples with maximum active sites were acquired by optimizing the plasma exposure time. The newly induced edges and S-vacancies were directly verified by high-resolution transmission electron microscopy. The 20 min treated MoS2 sample showed maximum active sites and thereby maximum HER activity, onset overpotential of ∼−210 mV vs reversible hydrogen electrode (RHE), and Tafel slope of ∼89 mV/dec. Clearly, the above results show that this approach can be employed for improving the HER efficiency of large-scale MoS2-based electrocatalysts.

1. Introduction

Hydrogen (H2) gas is an excellent source of clean energy and perceived as a suitable alternative for hydrocarbon-based fuels.14 Hydrogen is abundant in diverse natural compounds such as water, hydrocarbons, and so on. Hydrogen evolution reactions (HERs) from water-splitting reaction that breaks the H–O bond and releases oxygen and hydrogen in gaseous form can be achieved through an electrochemical or a photo-electrochemical process.58 These processes are essentially facilitated by the introduction of a catalytic substance. The best catalytic substances known for efficient HER are the Pt metal and/or Pt-based materials,9 which are not viable economically.2,10 Therefore, finding new cost-effective materials with comparable or even better HER performance is highly desirable.

Two-dimensional (2D) materials, such as MoS2, have been widely investigated for electronics and optoelectronic applications.1113 MoS2 also possesses great potential in terms of HER owing to its structural stability, high electrocatalytic activity, and earth abundance.1416 The electrochemical reaction that produces H2 gas mainly occurs at the outmost layer and/or at the edges of MoS2 where the active sites are located.14,15,1720 The inner layers do not take part in the reaction. Therefore, monolayer MoS2 (1L-MoS2) is more suitable for HER compared to bulk counterpart. Another advantage of 1L-MoS2 is the relatively easy transfer of an electron from the active sites to the electrode due to maximum hopping of electrons in the vertical direction.17,21 This phenomenon essentially reduces the contact resistance required for an efficient reaction. However, crystalline 1L-MoS2 has a low density of active sites in a pristine form that limits the HER performance. Interestingly, active sites in MoS2 can be engineered by various strategies, such as, doping,20,22 forming cracks on the surface,23,24 and creating S-vacancies.25,26 The 1L-MoS2 with a high edge density and defects could be an excellent catalytic material for an efficient HER.

Pertinently, MoS2 grown over an Au foil via conventional chemical vapor deposition (CVD) showed excellent HER performance and Tafel slope of 61 mV/dec attributed to the excellent coupling between the Au substrate and MoS2.27 On the other hand, a relatively cost-effective approach is the enhancement of HER performance by engineering defects in CVD-grown MoS2 on SiO2 substrate using plasma treatment or hydrogen annealing.24 However, 2D MoS2 samples grown by conventional CVD also could not break the limitation in terms of cost-effectiveness, as the sample area is typically limited to ∼1 cm2 or below per synthesis. Such a drawback arises from the utilization of a solid-phase inorganic precursor (mostly MoO3 and sulfur powder). In contrast, the metal–organic CVD (MOCVD) technique uses gas-phase precursors and therefore offers a wafer-scale area of 1L-MoS2 on arbitrary substrates. Moreover, the low reaction rate, low reaction temperature, and precisely controllable flow rate of precursors make MOCVD suitable to yield uniform, conformal, and wafer-scale 2D transition-metal dichalcogenides (TMDs).2833 These properties fulfill the essential requirements for economical and scalable HER electrocatalytic application.

In this article, we synthesized a large-scale 1L-MoS2 via MOCVD and employed N2-plasma treatment to induce more active sites that are required for an efficient HER. The creation of active sites by N2 plasma was confirmed by inspecting the surface morphology and the crystal structure by field emission scanning electron microscopy (FE-SEM) and scanning transmission electron microscopy (STEM). Micro-photoluminescence (μ-PL) and Raman spectroscopies were employed to study the effect of N2 plasma on the optical characteristics of MoS2. Both microscopic and spectroscopic characterization tools clearly showed that increasing the treatment time results in the formation of more active sites (edges and S-vacancies). The optimum treatment time is about 20 min that yielded the best HER activity. We believe that our simple post-treatment technique can be widely employed for improving the HER efficiency of transition-metal dichalcogenide (TMD)-based electrocatalysts in general.

2. Results and Discussion

2.1. Morphological and Structural Analysis

Initially, a large-area crystalline continuous 1L-MoS2 film was grown by MOCVD, as discussed in Section 4.1. The crystalline MoS2 shows a poor HER performance because of the low density of active sites, which are essential for HER activity. Nevertheless, the active sites including edges and S-vacancy sites can be generated by using N2-plasma treatment. We employed N2 plasma on 1L-MoS2 to induce active sites, as schematically depicted in Figure 1b,c. The creation of active sites was comparatively studied by inspecting the surface morphologies of 1L-MoS2 treated with N2 plasma for various times. Evidently, the untreated sample shows a perfectly clean surface with no apparent cracks or deformation (Figure S2a). However, a network of cracks, which work as active sites, begins to appear in the plasma-treated films (Figure S2b,c). The density of cracks was found to be increased with the treatment time. The FE-SEM image of 20 min treated sample clearly shows more cracks compared to 10 min treated sample, as shown in Figure S2b,c, respectively. Further treatment destroys the film (Figure S2d). Although the FE-SEM image of untreated 1L-MoS2 does not show any apparent structural deformation/distortion, some distortions were observed at the nanoscopic level resolution, as shown in the high-angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) image in Figure 1d. These structural distortions essentially happen at the grain boundaries, which arise from the lateral junction of two flakes. Here, it is worth mentioning that the MOCVD technique produces continuous polycrystalline 1L-MoS2 by laterally connecting the neighboring flakes and eventually form a large-area continuous film. The HADDF-STEM images that were taken within the grain exhibited a uniform hexagonal configuration, as shown in Figure S3a,b. Moreover, the as-grown film was a dominantly single layer; however, some negligibly small bilayer regions were also observed (Figure S3c).

Figure 1.

Figure 1

Effect of N2-plasma treatment on the morphology of MOCVD-grown continuous 1L-MoS2. (a) Schematic illustration of the MOCVD system. (b) Schematic illustration of the development of cracks in a pristine continuous 1L-MoS2 when exposed to N2 plasma. (c) A structural model for N2-plasma-treated 1L-MoS2. (d) HAADF-STEM image of pristine 1L-MoS2 across the grain boundary (GB) with the insets corresponding to the diffraction patterns of each grain. HAADF-STEM images of N2-plasma-treated MoS2 observed at (e, f) cracked grain boundary, (g) edge sites, and (h) S-vacancies.

The atomic-scale resolution images of N2-plasma-treated 1L-MoS2 film are depicted in Figures 1e–h and S3d–f, respectively. The high-resolution image taken at the cracked region clearly shows the formation of new edge sites. The edge seems to be terminated at the Mo edge, as depicted in Figure 1g. Another important aspect of plasma treatment is the creation of more S-vacancies in the MoS2 lattice. The HADDF-STEM image taken away from the crack boundary revealed the formation of S-vacancies (Figures 1h and S3f), which were absent in the untreated sample. Although it is well known that S-vacancies ubiquitously exist in CVD-grown MoS2, the density of these intrinsic S-vacancies is not sufficient to ensure a good HER performance. The S-vacancies should be at least up to ∼14% for the optimized Gibbs free energy for hydrogen adsorption (ΔGH* ≈ 0).25 These edges and S-vacancies work as active sites and thereby enhance the HER performance of MoS2.

2.2. Optical Characterization

The effect of N2 plasma on the optical characteristics of 1L-MoS2 was investigated by micro-Raman scattering and PL spectroscopy, as shown in Figure 2a. The wavenumber separation between the two dominant Raman modes, namely, E2g1 and A1g of MoS2, was found to be ∼20.7 cm–1 for the pristine sample, which is the typical Raman signature of a single-layer MoS2. After plasma treatment, the peak positions of both modes did not show a significant shift; however, the peak intensities rapidly drop accompanied by peak broadening, indicating the lattice distortion and defects generation such as cracks and S-vacancies. A similar trend was also observed in the PL spectra. The untreated sample shows a strong PL peak at ∼667 nm corresponding to the radiative recombination of A-excitons across the direct band gap at the K and K′ points in reciprocal space. This again confirms that the as-grown film was essentially a monolayer. The PL intensity also rapidly drops with increasing treatment time, which can be also attributed to enhanced defects sites, created by N2 plasma (Figure 2b).24,34 Moreover, the PL intensity mapping taken over the selected area of 20 μm2 of the MoS2 film before and after 20 min treatment further supports the above results (Figure S4).

Figure 2.

Figure 2

Evolution of Raman modes and PL characteristics of 1L-MoS2 with plasma treatment. (a) Raman spectra shows the evolution of typical Raman modes, namely, E2g1 and A1g of MoS2, and (b) PL spectra as a function of treatment time. The Raman spectra were calibrated by the Si Raman peak. All of the spectroscopic measurements were carried out at the same position.

2.3. Chemical Compositional Analysis

The effect of N2-plasma treatment on elemental bonding states of MoS2 was investigated by X-ray photoelectron spectroscopy (XPS), as shown in Figure 3. A comparative XPS analysis of pristine and 20 min plasma-treated MoS2 sample reveals a slight red shift in S2– 2p1/2 (163.2 eV) and S2– 2p3/2 (161.9 eV) peaks in the treated sample, which is consistent with the previous report,35 as illustrated in Figure 3a. The shift in the peak positions in the treated sample essentially indicates a change in the electronic band structure. Most likely, the Fermi level has been shifted down toward the valence band due to the formation of S-vacancies.36 The S-vacancies might transform the MoS2 from n- to p-type.35,37 Moreover, decrease in the S/Mo ratio from 1.93 of the as-grown MoS2 to 1.71 of the plasma-treated MoS2 further confirmed the generation of S-vacancies by plasma treatment. In addition, the peak located at 397.5 eV (green fitted line in Figure 3b) clearly increased in the plasma-treated sample. This peak can be attributed to the Mo–N bond.22,38,39 However, the quantitative analysis shows that the Mo–N bond position can be easily confusable with the peak position of CNx and CNx–H species, which might be formed during N2-plasma treatment of MoS2 due to the excessive availability of carbon species in ambient environment.35 To exactly confirm that the peak at 397.5 eV indeed corresponds to Mo–N bonding, an annealing process would be needed to completely desorb the carbon species and quickly load to in situ XPS characterization.35 Nevertheless, it was beyond the scope of this work, even though the N-doped MoS2 can be a practical approach to form even more active sites at the interior of the MoS2 lattice. Overall, the XPS results suggest that N2-plasma treatment is a feasible way to cleave the surface of MoS2 and increase the number of active sites in 1L-MoS2 for enhancing HER activity.

Figure 3.

Figure 3

Chemical composition analysis on the effect of plasma treatment. XPS spectra at (a) S 2p and (b) Mo 3p of 1L-MoS2 before and after 20 min treatment.

2.4. HER Performance

Finally, the HER performance of N2-plasma-treated MoS2 samples was evaluated. It is evident that the untreated MoS2 sample exhibited relatively weak HER activities with an onset overpotential at −460 mV vs the reversible hydrogen electrode (RHE), as shown in the iR-corrected linear sweep voltammograms (LSVs) (Figure 4a). Although the ideal MoS2 crystal has very limited active sites available for HER, CVD- or MOCVD-grown MoS2 possesses some S-vacancies.26 These vacancies along with the grain boundaries are indeed responsible for the limited HER activities of untreated MoS2.40 On the other hand, the samples treated with N2 plasma for 10–30 min showed considerable decrease in the onset overpotential (Figure 4a). The smallest value of the onset overpotential of −210 mV was obtained for the 20 min treated sample. The above results can simply be understood in terms of more active sites generated by N2-plasma treatment. As discussed earlier, the plasma treatment induced active sites (edges and S-vacancies) in MoS2 lattice that are beneficial for HER performance. However, the treatment time should be carefully chosen to acquire optimum active sites. In our case, the optimum time was found to be 20 min. Employing plasma for a longer time involves the risk of destroying the film. Nevertheless, the results show that N2 plasma irradiation is an effective way to engineer the active sites in a controlled manner, and thereby improving the HER performance of MoS2. The corresponding iR-corrected Tafel plot was also characterized to manifest the enhancement of the HER performance in plasma-treated MoS2 (Figure 4b). The slope of 89 mV/dec was observed for the optimally treated sample (20 min), which is considerably smaller than 118 mV/dec of the untreated sample. The overpotentials of the samples of overly treated samples (>30 min) were even larger compared to those of the as-grown MoS2 because most of the active area in MoS2 was destroyed. This observation is also consistent with the morphological and optical characterizations. For a comparative analysis, the HER performance of N2-plasma-treated MoS2 grown via MOCVD and recently reported CVD methods is tabulated in Table 1.

Figure 4.

Figure 4

Comparative analysis of HER activities of the pristine and plasma-treated MoS2 samples. (a, b) iR-corrected linear sweep voltammograms and Tafel plots, respectively. (c) Impedance Nyquist plot. (d) CV curves of the 20 min treated sample at various scan rates. (e) Linear fitting of the average capacitive current density vs the scan rate for the as-grown and N2-plasma-treated samples at different treatment times. The slop represents CDL of each sample, which is proportional to the ECSA and the roughness factor. (f) Stable HER performance of N2-plasma-treated samples using the potential vs time plot at −1.5 mA/cm2.

Table 1. Comparison of HER Performances of 2D-MoS2.

catalyst electrode growth method (substrate) post-treatment method onset overpotential (mV) Tafel slope (mV/dec) ref
MoS2 1L glassy carbon CVD (SiO2) O2 plasma –400 162 (24)
MoS2 1L glassy carbon CVD (SiO2) H2 annealing –300 147 (24)
MoS2 1L Au (111) CVD (SiO2) Ar plasma ∼−110 82 (25)
MoS2 1L Au LPCVD (Au)   ∼−100 61 (27)
MoS2 1L flakes high oriented pyrolytic graphite CVD   –286 70 (47)
MoS2, 10 nm thick FTO/glass MOCVD (FTO)   ∼−300 109 (31)
continuous 1L MoS2 graphite foil MOCVD (SiO2) N2 plasma –210 89 this work

Moreover, the charge-transfer resistance (Rct) of the plasma-treated MoS2 was estimated by the electrochemical impedance spectroscopy (EIS) measurement conducted at −0.4 V vs RHE (Figure 4c). The 20 min treated sample clearly shows a smaller semicircle diameter than the rest of the samples, indicating the lowest Rct (18 Ω) by fitting the corresponding EIS spectrum using a Randles circuit model (Figure S5). For the overly treated samples (40 and 50 min), considerably higher Rct values of 202–280 Ω were found, respectively, which is consistent with the linear sweep measurements shown in Figure 4a.

The electrochemically active surface area (ECSA) of catalysts can be evaluated from the value of the double-layer capacitance, CDL, which is known to be proportional to the ECSA and the roughness factor.4144 The cyclic voltammetry (CV) curves were characterized for all of our samples in non-Faradic potential regions at various scan rates from 10 to 70 mV/s. This is shown in Figure 4d for the optimal plasma treatment time of 20 min. The values of CDL were calculated from CV characterizations in Figure 4e, where a 20 min treatment yielded the maximum CDL of 0.74 mF/cm2, which is 4.3 times larger than that of the as-grown sample (0.17 mF/cm2). Our CDL results are also consistent with LSV scan and impedance measurement in Figure 4a,c, respectively, confirming the effect of N2-plasma treatment on the electrochemical performance of MoS2 samples.

Furthermore, the stability of the HER activity was evaluateded using the transient Chrono potentiometric study for 12 h at 1.5 mA/cm2, as shown in Figure 4f. The samples exhibited inappreciable degradation in HER and maintained the overpotential, except those samples that were treated for more than 30 min.

3. Conclusions

In summary, we employed a simple N2-plasma treatment mechanism to generate sufficient active sites in terms of S-vacancies and edges in a large-area 1L-MoS2. This essentially allowed exploiting the large-area 1L-MoS2 film grown via MOCVD for efficient HER activities. The density of active sites could be effectively controlled simply by exposing the sample to N2 plasma for various times. In our case, the sample treated for 20 min showed the best HER performance with the onset overpotential of −210 mV vs RHE and the corresponding Tafel slope of 89 mV/dec. Our proposed method is simple and controllable to maximize the HER efficiency of the wafer-scale working medium, which may also be applied to other members of the TMD family for even further improving the HER efficiency.

4. Experimental Section

4.1. Synthesis of 1L-MoS2

A large-area 1L-MoS2 film was prepared by MOCVD, as schematically illustrated in Figure 1a. In brief, the synthesis was carried out in a sealed quartz tube of 1 inch diameter. A SiO2/Si substrate was placed at the center of the tube. The metal–organic compounds of molybdenum hexacarbonyl (MHC, Mo(CO)6) and diethyl sulfide (DES, (C2H5)2S) were used as gas-phase precursors. These precursors have a high equilibrium vapor pressure at room temperature. Initially, the base pressure (∼1 mTorr) was created in the chamber. The temperatures of MHC and DES holder were kept at 27 and 40 °C, respectively. The temperatures of the precursor’s carrier-line and reaction zone were set to 50 and 500 °C, respectively. 30 sccm of Ar and 3 sccm of H2 was flowed into the reaction chamber. The reaction pressure was set to 60 Torr. The reaction process was started by introducing 1 sccm of DES and MHC. The reaction time was set to 10 h. The detailed schematic description is given in Figure S1a and our previous report.29

4.2. N2-Plasma Treatment

The as-synthesized MoS2 (2 × 3 cm2) was cut into several pieces, and few samples were selected to study the impact of plasma treatment as a function of treatment time (Figure 1b). The electric potential of 20 kV was applied with 100 sccm of N2 flow to generate plasma (Figure S1b). The distance between the sample and the high potential tip was adjusted and the corresponding electrical current was about 0.2 mA. For a comparative study of HER activity, six samples were prepared with different treatment times of 0, 10, 20, 30, 40, and 50 min.

4.3. Characterization

The morphologies of the as-grown and N2-plasma-treated 1L-MoS2 samples were characterized by FE-SEM (JEOL JSM-6500F), high angle annular dark field scanning transmission electron microscopy (HAADF-STEM). STEM samples were prepared without using any polymer backing layer. Quantifoil holey carbon TEM grids were placed onto flakes on SiO2/Si substrate and a drop of isopropyl alcohol was dropped and dried. The SiO2 layer was etched by 30% KOH solution for 4 h and the TEM grids were detached from the substrate. Then, TEM grids were rinsed several times in deionized (DI) water and dried in ambient condition. STEM imaging was performed by a JEOL ARM 200F equipped with image and probe aberration correctors operated at 80 kV. The convergence semiangle was set at ∼25 mrad. HAADF-STEM images were acquired from 80 to 200 mrad range. Optical properties were characterized by micro-Raman and PL spectroscopies using 473 nm excitation source under ambient conditions. The chemical composition analysis was studied by X-ray photoelectron spectroscopy (XPS; Theta Probe AR-XPS System, Thermo Fisher Scientific).

4.4. HER Measurements

The MoS2 samples on SiO2/Si were transferred to a graphite foil that is a working electrode for studying the HER activity (Figure S1c). The films were transferred by the wet transfer method.24,37,45 First, MoS2 (1 × 1 cm2) on SiO2/Si was syringed by 0.3 mL poly methyl methacrylate (PMMA, 5% toluene), spin-coated at 5000 rpm for 30 s, and dried at 90 °C for 30 min. The PMMA/MoS2/SiO2/Si was then dipped in 10 mL of KOH solution (1 M) at 90 °C to etch the SiO2 layer. The PMMA/MoS2 was floated on the solution. After several washings with DI water, the PMMA/MoS2 layer was transferred onto the target electrode (graphite foil). Finally, the PMMA layer was removed by 99.99% warm acetone.

The HER measurements including linear sweep and electrochemical impedance spectroscopy (EIS) were conducted in an electrochemical workstation (IviumStat, Ivium Tech) with three electrodes including a Pt wire as the counter electrode, an Ag/AgCl as the reference electrode, and the MoS2/graphite foil as the working electrode (Figure S1d). For HER stability study, the graphite rod was used as the counter electrode instead of Pt wire to avoid the Pt dissolution and deposition on the working electrode during the HER stability experiment.46 All of the three electrodes were placed in 0.5 M H2SO4 electrolyte to observe the HER activity. The EIS result was fitted using the equivalent circuit evaluator (Ivium Tech) modeled with two constant-phase elements (CPEs).

Acknowledgments

This research was supported by the Basic Science Research Programs (2017R1E1A1A01075350 and 2017R1D1A1B03035539), the Basic Research Lab Program (2014R1A4A1071686), the Priority Research Centers Program (2019R1A6A1A11053838), and the National Research Foundation of Korea (NRF), funded by the Korean government.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.9b03205.

  • Schematics for MOCVD growth, N2-plasma treatment, transfer process, HER characterization; FE-SEM; HAADF-STEM images; PL intensity mapping; equivalent model circuit for fitting EIS results (PDF)

Author Contributions

A.D.N. and T.K.N. contributed equally to this work. The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ao9b03205_si_001.pdf (905.9KB, pdf)

References

  1. Schlapbach L.; Züttel A. Hydrogen-storage Materials for Mobile Applications. Nature 2001, 414, 353–358. 10.1038/35104634. [DOI] [PubMed] [Google Scholar]
  2. Yan Y.; Xia B.; Xu Z.; Wang X. Recent Development of Molybdenum Sulfides as Advanced Electrocatalysts for Hydrogen Evolution Reaction. ACS Catal. 2014, 4, 1693–1705. 10.1021/cs500070x. [DOI] [Google Scholar]
  3. Turner J. A. Sustainable Hydrogen Production. Science 2004, 305, 972–974. 10.1126/science.1103197. [DOI] [PubMed] [Google Scholar]
  4. Cortright R. D.; Davda R. R.; Dumesic J. A. Hydrogen from Catalytic Reforming of Biomass-derived Hydrocarbons in Liquid Wate. Nature 2002, 418, 964–966. 10.1038/nature01009. [DOI] [PubMed] [Google Scholar]
  5. Morales-Guio C. G.; Stern L. A.; Hu X. Nanostructured Hydrotreating Catalysts for Electrochemical Hydrogen Evolution. Chem. Soc. Rev. 2014, 43, 6555–6569. 10.1039/C3CS60468C. [DOI] [PubMed] [Google Scholar]
  6. 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. 10.1021/cs500923c. [DOI] [Google Scholar]
  7. Grätzel M. Photoelectrochemical Cells. Nature 2001, 414, 338–344. 10.1038/35104607. [DOI] [PubMed] [Google Scholar]
  8. Conway B. E.; Tilak B. V. Interfacial Processes Involving Electrocatalytic Evolution and Oxidation of H2, and the Role of Chemisorbed H. Electrochim. Acta 2002, 47, 3571–3594. 10.1016/S0013-4686(02)00329-8. [DOI] [Google Scholar]
  9. Greeley J.; Jaramillo T. F.; Bonde J.; Chorkendorff I. B.; Norskov J. K. Computational High-throughput Screening of Electrocatalytic Materials for Hydrogen Evolution. Nat. Mater. 2006, 5, 909–913. 10.1038/nmat1752. [DOI] [PubMed] [Google Scholar]
  10. Khaselev O.; Turne J. A. A Monolithic Photovoltaic-Photoelectrochemical Device for Hydrogen Production via Water Splitting. Science 1998, 280, 425–427. 10.1126/science.280.5362.425. [DOI] [PubMed] [Google Scholar]
  11. Bhimanapati G. R.; Lin Z.; Meunier V.; Jung Y.; Cha J.; Das S.; Xiao D.; Son Y.; Strano M. S.; Cooper V. R.; Liang L.; Louie S. G.; Ringe E.; Zhou W.; Kim S. S.; Naik R. R.; Sumpter B. G.; Terrones H.; Xia F.; Wang Y.; Zhu J.; Akinwande D.; Alem N.; Schuller J. A.; Schaak R. E.; Terrones M.; Robinson J. A. Recent Advances in Two-Dimensional Materials Beyond Graphene. ACS Nano 2015, 9, 11509–11539. 10.1021/acsnano.5b05556. [DOI] [PubMed] [Google Scholar]
  12. Chhowalla M.; Shin H. S.; Eda G.; Li L. J.; Loh K. P.; Zhang H. The Chemistry of Two-Dimensional Layered Transition Metal Dichalcogenide Nanosheets. Nat. Chem. 2013, 5, 263–275. 10.1038/nchem.1589. [DOI] [PubMed] [Google Scholar]
  13. Wang Q. H.; Kalantar-Zadeh K.; Kis A.; Coleman J. N.; Strano M. S. Electronics and Optoelectronics of Two-dimensional Transition Metal Dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. 10.1038/nnano.2012.193. [DOI] [PubMed] [Google Scholar]
  14. Jaramillo T. F.; Jørgensen K. P.; Bonde J.; Nielsen J. H.; Horch S.; Chorkendorff I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 137, 100–102. 10.1126/science.1141483. [DOI] [PubMed] [Google Scholar]
  15. Hinnemann B.; Moses P. G.; Bonde J.; Jørgensen K. P.; Nielsen J. H.; Horch S.; Chorkendorff I.; Nørskov J. K. Biomimetic Hydrogen Evolution: MoS2 Nanoparticles as Catalyst for Hydrogen Evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. 10.1021/ja0504690. [DOI] [PubMed] [Google Scholar]
  16. Lu Q.; Yu Y.; Ma Q.; Chen B.; Zhang H. 2D Transition-Metal-Dichalcogenide-Nanosheet-Based Composites for Photocatalytic and Electrocatalytic Hydrogen Evolution Reactions. Adv. Mater. 2016, 28, 1917–1933. 10.1002/adma.201503270. [DOI] [PubMed] [Google Scholar]
  17. Yu Y.; Huang S.-Y.; Li Y.; Steinmann S. N.; Yang W.; Cao L. Layer-dependent Electrocatalysis of MoS2 for Hydrogen Evolution. Nano Lett. 2014, 14, 553–558. 10.1021/nl403620g. [DOI] [PubMed] [Google Scholar]
  18. Bonde J.; Moses P. G.; Jaramillo T. F.; Nørskovb J. K.; Chorkendorf I. Hydrogen Evolution on Nano-particulate Transition Metal Sulfides. Faraday Discuss. 2009, 140, 219–231. 10.1039/B803857K. [DOI] [PubMed] [Google Scholar]
  19. Tsai C.; Abild-Pedersen F.; Nørskov J. K. Tuning the MoS2 Edge-site Activity for Hydrogen Evolution via Support Interactions. Nano Lett. 2014, 14, 1381–1387. 10.1021/nl404444k. [DOI] [PubMed] [Google Scholar]
  20. Wang H.; Tsai C.; Kong D.; Chan K.; Abild-Pedersen F.; Nørskov J. K.; Cui Y. Transition-metal Doped Edge Sites in Vertically Aligned MoS2 Catalysts for Enhanced Hydrogen Evolution. Nano Res. 2015, 8, 566–575. 10.1007/s12274-014-0677-7. [DOI] [Google Scholar]
  21. Tang H.; Morrison S. R. Optimization of The Anisotropy of Composite MoS2 Films. Thin Solid Films 1993, 227, 90–94. 10.1016/0040-6090(93)90190-Z. [DOI] [Google Scholar]
  22. Xiao W.; Liu P.; Zhang J.; Song W.; Feng Y. P.; Gao D.; Ding J. Dual-Functional N Dopants in Edges and Basal Plane of MoS2 Nanosheets Toward Efficient and Durable Hydrogen Evolution. Adv. Energy Mater. 2017, 7, 1602086 10.1002/aenm.201602086. [DOI] [Google Scholar]
  23. Xie J.; Zhang H.; Li S.; Wang R.; Sun X.; Zhou M.; Zhou J.; Lou X. W.; Xie Y. Defect-rich MoS2 Ultrathin Nanosheets with Additional Active Edge Sites for Enhanced Electrocatalytic Hydrogen Evolution. Adv. Mater. 2013, 25, 5807–5813. 10.1002/adma.201302685. [DOI] [PubMed] [Google Scholar]
  24. 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. 10.1021/acs.nanolett.5b04331. [DOI] [PubMed] [Google Scholar]
  25. Li H.; Tsai C.; Koh A. L.; Cai L.; Contryman A.; Fragapane A. H.; Zhao J.; Han H. S.; Manoharan H. C.; Abild-Pedersen F.; Nørskov J. K.; Zheng X. Activating and Optimizing MoS2 Basal Planes for Hydrogen Evolution through the Formation of Strained Sulphur Vcancies. Nat. Mater. 2016, 15, 364. 10.1038/nmat4564. [DOI] [PubMed] [Google Scholar]
  26. Voiry D.; Fullon R.; Yang J.; de Carvalho Castro E. S. C.; Kappera R.; Bozkurt I.; Kaplan D.; Lagos M. J.; Batson P. E.; Gupta G.; Mohite A. D.; Dong L.; Er D.; Shenoy V. B.; Asefa T.; Chhowalla M. The Role of Electronic Coupling between Substrate and 2D MoS2 Nanosheets in Electrocatalytic Production of Hydrogen. Nat. Mater. 2016, 15, 1003–1009. 10.1038/nmat4660. [DOI] [PubMed] [Google Scholar]
  27. Shi J.; Ma D.; Han G.-F.; Zhang Y.; Ji Q.; Gao T.; Sun J.; Song X.; Li C.; Zhang Y.; Lang X.-Y.; Zhang Y.; Liu Z. Controllable Growth and Transfer of Monolayer MoS2 on Au Foils and Its Potential Application in Hydrogen Evolution Reaction. ACS Nano 2014, 8, 10196–10204. 10.1021/nn503211t. [DOI] [PubMed] [Google Scholar]
  28. Kang K.; Xie S.; Huang L.; Han Y.; Huang P. Y.; Mak K. F.; Kim C.-J.; Muller D.; Park J. High-mobility Three-atom-thick Semiconducting Films with Wafer-scale Homogeneity. Nature 2015, 520, 656–660. 10.1038/nature14417. [DOI] [PubMed] [Google Scholar]
  29. Nguyen T. K.; Nguyen A. D.; Le C. T.; Ullah F.; Tahir Z.; Koo K. I.; Kim E.; Kim D. W.; Jang J. I.; Kim Y. S. High Photoresponse in Conformally Grown Monolayer MoS2 on a Rugged Substrate. ACS Appl. Mater. Interfaces 2018, 10, 40824–40830. 10.1021/acsami.8b15673. [DOI] [PubMed] [Google Scholar]
  30. Eichfeld S. M.; Hossain L.; Lin Y.-C.; Piasecki A. F.; Kupp B.; Birdwell A. G.; Burke R. A.; Lu N.; Peng X.; Li J.; Azcatl A.; McDonnell S.; Wallace R. M.; Kim M. J.; Mayer T. S.; Redwing J. M.; Robinson J. A. Highly Scalable, Atomically Thin WSe2 Grown via Metal-organic Chemical Vapor Deposition. ACS Nano 2015, 9, 2080–2087. 10.1021/nn5073286. [DOI] [PubMed] [Google Scholar]
  31. Cwik S.; Mitoraj D.; Reyes O. M.; Rogalla D.; Peeters D.; Kim J.; Schütz H. M.; Bock C.; Beranek R.; Devi A. Direct Growth of MoS2 and WS2 Layers by Metal Organic Chemical Vapor Deposition. Adv. Mater. Interfaces 2018, 5, 1800140 10.1002/admi.201800140. [DOI] [Google Scholar]
  32. Kim H.; Ovchinnikov D.; Deiana D.; Unuchek D.; Kis A. Suppressing Nucleation in Metal-Organic Chemical Vapor Deposition of MoS2 Monolayers by Alkali Metal Halides. Nano Lett. 2017, 17, 5056–5063. 10.1021/acs.nanolett.7b02311. [DOI] [PubMed] [Google Scholar]
  33. Kim T.; Mun J.; Park H.; Joung D.; Diware M.; Won C.; Park J.; Jeong S.-H.; Kang S.-W. Wafer-scale Production of Highly Uniform Two-dimensional MoS2 by Metal-Organic Chemical Vapor Deposition. Nanotechnology 2017, 28, 18LT01 10.1088/1361-6528/aa6958. [DOI] [PubMed] [Google Scholar]
  34. Kang N.; Paudel H. P.; Leuenberger M. N.; Tetard L.; Khondaker S. I. Photoluminescence Quenching in Single-Layer MoS2 via Oxygen Plasma Treatment. J. Phys. Chem. C 2014, 118, 21258–21263. 10.1021/jp506964m. [DOI] [Google Scholar]
  35. Azcatl A.; Qin X.; Prakash A.; Zhang C.; Cheng L.; Wang Q.; Lu N.; Kim M. J.; Kim J.; Cho K.; Addou R.; Hinkle C. L.; Appenzeller J.; Wallace R. M. Covalent Nitrogen Doping and Compressive Strain in MoS2 by Remote N2 Plasma Exposure. Nano Lett. 2016, 16, 5437–5443. 10.1021/acs.nanolett.6b01853. [DOI] [PubMed] [Google Scholar]
  36. Ma Q.; Odenthal P. M.; Mann J.; Le D.; Wang C. S.; Zhu Y.; Chen T.; Sun D.; Yamaguchi K.; Tran T.; Wurch M.; McKinley J. L.; Wyrick J.; Magnone K.; Heinz T. F.; Rahman T. S.; Kawakami R.; Bartels L. Controlled Argon Beam-induced Desulfurization of Monolayer Molybdenum Disulfide. J. Phys.: Condens. Matter 2013, 25, 252201 10.1088/0953-8984/25/25/252201. [DOI] [PubMed] [Google Scholar]
  37. McDonnell S.; Addou R.; Buie C.; Wallace R. M.; Hinkle C. L. Defect-Dominated Doping and Contact Resistance in MoS2. ACS Nano 2014, 8, 2880–2888. 10.1021/nn500044q. [DOI] [PubMed] [Google Scholar]
  38. Zhou W.; Hou D.; Sang Y.; Yao S.; Zhou J.; Li G.; Li L.; Liuc H.; Chen S. MoO2 Nanobelts@nitrogen Self-doped MoS2 Nanosheets as Effective Electrocatalysts for Hydrogen Evolution Reaction. J. Mater. Chem. A 2014, 2, 11358. 10.1039/c4ta01898b. [DOI] [Google Scholar]
  39. Li R.; Yang L.; Xiong T.; Wu Y.; Cao L.; Yuan D.; Zhou W. Nitrogen doped MoS2 Nanosheets Synthesized via a Low-temperature Process as Electrocatalysts with Enhanced Activity for Hydrogen Evolution Reaction. J. Power Sources 2017, 356, 133–139. 10.1016/j.jpowsour.2017.04.060. [DOI] [Google Scholar]
  40. Dong S.; Wang Z. Grain Boundaries Trigger Basal Plane Catalytic Activity for the Hydrogen Evolution Reaction in Monolayer MoS2. Electrocatalysis 2018, 9, 744–751. 10.1007/s12678-018-0485-z. [DOI] [Google Scholar]
  41. Boggio R.; Carugati A.; Trasatti S. Electrochemical Surface Properties of Co3O4 Electrodes. J. Appl. Electrochem. 1987, 17, 828–840. 10.1007/BF01007821. [DOI] [Google Scholar]
  42. McCrory C. C.; Jung S.; Ferrer I. M.; Chatman S. M.; Peters J. C.; Jaramillo T. F. Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347–4357. 10.1021/ja510442p. [DOI] [PubMed] [Google Scholar]
  43. Liu G.; Li Z.; Shi J.; Sun K.; Ji Y.; Wang Z.; Qiu Y.; Liu Y.; Wang Z.; Hu P. Black Reduced Porous SnO2 Nanosheets for CO2 Electroreduction with High Formate Selectivity and Low Overpotential. Appl. Catal., B 2020, 260, 118134 10.1016/j.apcatb.2019.118134. [DOI] [Google Scholar]
  44. Zhang Y.; Sun H.; Qiu Y.; Ji X.; Ma T.; Gao F.; Ma Z.; Zhang B.; Hu P. Multiwall Carbon Nanotube Encapsulated Co Grown on Vertically Oriented Graphene Modified Carbon Cloth as Bifunctional Electrocatalysts for Solid-state Zn-air Battery. Carbon 2019, 144, 370–381. 10.1016/j.carbon.2018.12.055. [DOI] [Google Scholar]
  45. van der Zande A. M.; Huang P. Y.; Chenet D. A.; Berkelbach T. C.; You Y.; Lee G.-H.; Heinz T. F.; Reichman D. R.; Muller D. A.; Hone J. C. Grains and Grain Boundaries in Highly Crystalline Monolayer Molybdenum Disulphide. Nat. Mater. 2013, 12, 554–561. 10.1038/nmat3633. [DOI] [PubMed] [Google Scholar]
  46. Liu G.; Qiu Y.; Wang Z.; Zhang J.; Chen X.; Dai M.; Jia D.; Zhou Y.; Li Z.; Hu P. Efficiently Synergistic Hydrogen Evolution Realized by Trace Amount of Pt-Decorated Defect-Rich SnS2 Nanosheets. ACS Appl. Mater. Interfaces 2017, 9, 37750–37759. 10.1021/acsami.7b11413. [DOI] [PubMed] [Google Scholar]
  47. Liu L.; Sun L.; Zheng J.; Xing L.; Gu L.; Wu J.; Wu L.; Ye M.; Liu X.; Wang Q.; Hou S.; Jiang X.; Xie L.; Lu P.; Jiao L. Phase-selective Synthesis of 1T′ MoS2 Monolayers and Heterophase bilayers. Nat. Mater. 2018, 17, 1108–1114. 10.1038/s41563-018-0187-1. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao9b03205_si_001.pdf (905.9KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES