Next Article in Journal
Green Synthesis, Characterization, Antimicrobial, Anti-Cancer, and Optimization of Colorimetric Sensing of Hydrogen Peroxide of Algae Extract Capped Silver Nanoparticles
Next Article in Special Issue
The Thermoelectric Properties of Monolayer MAs2 (M = Ni, Pd and Pt) from First-Principles Calculations
Previous Article in Journal
Abatement of the Stimulatory Effect of Copper Nanoparticles Supported on Titania on Ovarian Cell Functions by Some Plants and Phytochemicals
Previous Article in Special Issue
Electromagnetic Analysis of Vertical Resistive Memory with a Sub-nm Thick Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Functionalization of Molybdenum Disulfide via Plasma Treatment and 3-Mercaptopropionic Acid for Gas Sensors

1
Department of Control and Instrumentation Engineering, Korea University, 2511 Sejong-ro, Sejong-si 30019, Korea
2
Department of Nanomechatronics Engineering, Pusan University, Busan, 2 Busandaehak-ro 63 beon-gil, Geumjeong-gu, Busan 46241, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(9), 1860; https://doi.org/10.3390/nano10091860
Submission received: 24 August 2020 / Revised: 14 September 2020 / Accepted: 15 September 2020 / Published: 17 September 2020
(This article belongs to the Special Issue 2D Materials for Nanoelectronics)

Abstract

:
Monolayer and multilayer molybdenum disulfide (MoS2) materials are semiconductors with direct/indirect bandgaps of 1.2–1.8 eV and are attractive due to their changes in response to electrical, physicochemical, biological, and mechanical factors. Since the desired electrical properties of MoS2 are known, research on its electrical properties has increased, with focus on the deposition and growth of large-area MoS2 and its functionalization. While research on the large-scale production of MoS2 is actively underway, there is a lack of studies on functionalization approaches, which are essential since functional groups can help to dissolve particles or provide adequate reactivity. Strategies for producing films of functionalized MoS2 are rare, and what methods do exist are either complex or inefficient. This work introduces an efficient way to functionalize MoS2. Functional groups are formed on the surface by exposing MoS2 with surface sulfur vacancies generated by plasma treatment to 3-mercaptopropionic acid. This technique can create 1.8 times as many carboxyl groups on the MoS2 surface compared with previously reported strategies. The MoS2-based gas sensor fabricated using the proposed method shows a 2.6 times higher sensitivity and much lower detection limit than the untreated device.

Graphical Abstract

1. Introduction

As a 2D material, graphene has been demonstrated to exhibit differing electrical and optical properties depending on factors such as electrical and magnetic fields [1], strain [2], stacking geometry [3], and edge chirality [4,5] However, many have realized the applicability limitations of graphene since it does not have a bandgap, which has raised interest in transition-metal dichalcogenide (TMD) materials. TMD materials consist of hexagonal metal atoms (M) sandwiched between two chalcogen atomic layers (X), which form 3D crystals through the lamination of adjacent sheets through van der Waals interactions. TMD materials are potentially useful in electronics owing to their superconductivity [6] and charge density wave effects [7].
Monolayer and multilayer molybdenum disulfide (MoS2), one of the most studied TMDs, are semiconductors with direct/indirect bandgaps of 1.2–1.8 eV [8,9] and are highly attractive due to the considerable environment-dependent changes in their electrical [10,11,12,13], physicochemical [14], biological [15], optical [16], and mechanical properties [17]. Particularly, a field-effect transistor with monolayer MoS2 exhibited highly favorable electrical properties such as a mobility of 200 cm2/(V s), on/off ratio of 108, and sub-threshold swing of 70 mV/dec [11]. Since these electrical properties of MoS2 were revealed, related studies have increased, and research toward extending its applicability has focused on large-area deposition and growth [18,19,20,21] and functionalization [22,23,24,25]. While significant progress has been made toward large-scale MoS2, studies on functionalization techniques have yet to be actively conducted. Functional surface ligands consist of an anchor group that can attach to gas molecules, nanocrystals, biomaterials, etc. Particularly, functional groups can help dissolve the particles and provide adequate reactivity. Therefore, functionalization strategies for these materials are necessary. Detailed strategies must be established since some functionalization techniques for carbon nanotubes are not suitable for TMD materials.
Attempts have been intermittently made to introduce new MoS2 functionalization strategies. Recently, a report described an approach to directly and simply functionalize MoS2 using nitrilotriacetic acid [22]. However, the method requires a transition-metal cation with a high sulfur affinity and octahedral coordination since the direct anchoring of organic ligands to the sulfur surface is impossible. Although their new method was simple, the application of transition-metal cations complicates the process. Chen et al. and Chu et al. introduced methods for functionalizing MoS2 using 4-nitrobenzenediazonium tetrafluoroborate [23] and cysteine [24], respectively, after generating sulfur vacancies, but neither described how to generate the sulfur vacancies or gave detailed experimental methods. The direct synthesis of various organic functional groups on MoS2 thin films has also been reported [25], but the method is less efficient than the proposed method and lacks optimization.

2. Materials and Methods

Tri-layer MoS2 films grown through CVD were obtained from 6carbon (Shenzhen, China). Gold paste (C2041206P2) as a gas sensor pad was purchased from Gwent (Pontypool, UK). 3-Mercaptopropionic acid (MPA) was purchased from Sigma-Aldrich (St. Louis, MO, USA), and a 40 mM MPA solution was prepared using a mixture of 99% ethanol and deionized (DI) water.
The fabricated MoS2-based gas sensor was characterized by scanning electron microscopy (SEM) (Hitachi, Tokyo, Japan), and the change in resistance in response to the constituent gas concentration was measured using an LCR meter (Hioki, Nahano, Japan). The chemical composition for each experimental step was investigated using X-ray photoelectron spectroscopy (XPS) (Thermo, MA, U.S.A). The sulfur vacancies and MoS2 thickness after plasma treatment were measured by Raman spectroscopy (LabRAM, Horiba KOREA, Bucheon, Korea) and atomic force microscopy (AFM) (Shimadzu, Kyoto, Japan), respectively.
The MoS2 film grown by CVD on a 3 mm × 3 mm SiO2/Si substrate was plasma-treated at 10 W for 2 s in an Ar+ atmosphere to create sulfur vacancies. The plasma-treated MoS2 film was then exposed to the MPA solution overnight to form coordinate bonds between the HS groups in MPA and sulfur vacancies. To fabricate the gas sensor pad, the gold paste was screen-printed onto both ends of the functionalized MoS2 film. Figure 1 shows a schematic diagram and SEM image of the functionalized MoS2-based gas sensor.

3. Results and Discussion

The presence of functional groups on a sensing material is a crucial factor in determining its performance since they can increase the sensitivity. Figure 2 shows the functionalization process of the MoS2 films. The first step involves the generation of sulfur vacancies by plasma treatment. Since MoS2 is a 2D material, it is highly susceptible to physical damage, and thus a weak and chemically inactive plasma treatment should be performed. Since the minimum power of our home-built plasma equipment is 10 W, the plasma power was fixed at 10 W, and Ar ions were used as plasmons for their chemical inactivity. As shown in Figure 3a, AFM analysis was conducted to examine the physical changes in the MoS2 film over the treatment period. The thickness of a three-layer MoS2 film is theoretically 2.3 nm [8,9,26], which was the same as the thickness measured by AFM analysis. As the treatment time increased, the MoS2 film was physically etched at a rate of 0.6 Å/s. The etch rate of the Ar plasma was notably slower than those reported for previous studies using oxygen plasma [11], indicating that Ar plasma is more suitable for the generation of sulfur vacancies in MoS2 films.
Figure 3b shows the Raman spectra of the MoS2 films that were plasma treated for different durations. Numerous theoretical studies have indicated that the frequency difference between the E12g and the A1g Raman modes is dependent on the thickness of the MoS2 film [18]. The layers of MoS2 are held together via van der Waals forces. As the number of layers decreases, less incident energy is lost through the vertical vibrations of the molecules, resulting in a decrease in the frequency of the A1g peak [18,19,20,21]. Conversely, the E12g peak is affected more by the interatomic coulomb potential, which decreases with fewer layers due to the reduction in interactions between atoms. As a result, the E12g frequency increases. This increase in E12g frequency corresponding to horizontal vibrations is less sensitive to the number of layers than the decrease in A1g frequency corresponding to vertical vibrations. The A1g peak occurring above the bi-layer rather than a single layer is less sensitive to the MoS2 thickness. However, the E12g peak is sensitive to the MoS2 thickness. The reduced sensitivity to thickness of the E12g mode has been explained by the less-efficient inter-layer coupling of the in-plane phonons [27]. As the MoS2 becomes the generation of sulfur vacancies was confirmed to be 2 s, since the physical thickness of sulfur is 0.176 nm thinner, the E12g frequency increases and the A1g frequency decreases, thereby reducing the frequency difference. This frequency difference for the MoS2 film treated for 8 s was reduced from 23.21 for the untreated film to 21.89 cm−1. A previous paper reported this frequency difference for a bilayer MoS2 film as 21.61 cm−1 [28]. Therefore, the Raman analysis indicates that the treatment time required to etch one layer was 8 s in the proposed method. Thus, to form sulfur vacancies in a 3-layer MoS2 film without removing a layer, a treatment time of less than 8 s is required. As demonstrated by the AFM analysis (Figure 3a), the treatment time for the generation of sulfur vacancies was confirmed to be 2 s, since the physical thickness of sulfur is 0.176 nm.
The XPS S 2p and Mo 3d deconvolutions further supported the generation of sulfur vacancies by the plasma treatment (Figure 4). The Mo 3 d 5 2 and S 2 p 3 2 peaks of the MoS2 film appeared at 229.3 and 162.2 eV, respectively, which have previously indicated an MoS2 film with three layers [25]. The S/Mo ratio can be calculated as the area ratio of the S 2p orbital representing the number of S atoms and the Mo 3d orbital representing the number of Mo atoms. Using the integral, the area of each orbital according to the XPS result was calculated, and the S/Mo ratio was calculated to compare before and after plasma treatment. The untreated MoS2 film showed the ideal S/Mo ratio of 1.91, while that of the plasma-treated MoS2 film was 1.51, demonstrating the formation of sulfur vacancies [29].
For the functionalization of the MoS2 surface, the MoS2 film with sulfur vacancies was exposed to an MPA solution. MoS2 has been used as a hydrodesulphurization catalyst because surface sulfur vacancies tend to form covalent bonds with sulfur-containing groups. As a result, molecules with thiol functional groups chemically bond to the sulfur vacancies formed on the MoS2 surface [30,31]. The chemical components of the MoS2 films with sulfur vacancies treated in solutions of MPA at different concentrations were measured by XPS to determine the optimized MPA concentration (Figure 5a). The measured absorbance spectra presented four peaks; the first at 285 eV was assigned to sp2-hybridized carbon atoms, and the other three at 286.06, 288.51, and 288.97 eV were assigned to oxygen-containing hydroxyl (C−O), carbonyl (C=O), and carboxyl groups (−COO), respectively. The presence of these oxygen-containing group features implies that the functional groups were stably bonded to the MoS2 film [32]. From the XPS results for different MPA concentrations, the atomic ratio corresponding to carboxyl groups increased for concentrations up to 55 mM and then decreased, potentially because high concentrations of MPA may form inter-MPA disulfide bridges (Figure 5b).
Figure 5 also compares the effectiveness of the proposed and previous method for the formation of functional groups. The exfoliation process by n-butyllithium deforms the crystal structure of MoS2 and introduces defects mainly at the edges of the exfoliated 2D material [25]. Conversely, the defects introduced by the plasma treatment were evenly generated on the MoS2 surface, and thus more defects were created compared with the exfoliation process. As previously mentioned, since thiol functional groups covalently bond to the defect sites, the number of defect sites is a crucial factor in the functionalization of the nanosheets. As illustrated by the C 1s deconvolution in Figure 5, approximately 1.8 times more carboxyl groups were produced using the proposed strategy compared with the amount generated by the previously published methods.
As shown in Figure 6, the analytical performance characteristics of the functionalized MoS2 films were confirmed using an NH3 gas sensor. The electrical properties of MoS2 are highly sensitive to charge transfer between adsorbed molecules and the 2D films. According to theoretical studies, NH3 interacts weakly with perfect films and induces little charge transfer [33,34]. From the theoretical studies discussed above, it can be expected that the functionalized MoS2-based gas sensor has a higher sensitivity than the pristine MoS2-based sensor. Therefore, the adsorption of NH3 onto defect-free films is difficult, whereas MoS2 films with functional groups can easily adsorb the gas. Additionally, the adsorption barrier of functionalized MoS2 films is further lowered by pre-dissociated oxygen atoms, which results in an increase in the charge transfer rate [35]. Both the untreated and functionalized devices exhibited an increase in resistance upon exposure to NH3 gas [36]. Figure 6c presents the calibrated response curves for different gas concentrations of the untreated and functionalized MoS2-based sensors. The device with untreated MoS2 had a sensitivity of 0.0075 at a gas concentration range of 0–22 ppm. On the other hand, the functionalized MoS2-based sensor showed a sensitivity of 0.02 for the same range, which is 2.6 times higher. This enhanced sensitivity is attributable to a reduced electron-withdrawing ability from the formation of hydrogen bonds between the functional groups and polar NH3 molecules [37]. Moreover, the device with the functionalized MoS2 film showed a detection limit two times lower than that of the untreated device.

4. Conclusions

In this study, a technique for the stable functionalization of MoS2 was investigated. Although it is difficult to produce functional groups on 2D materials, it is possible to stably form functional groups on MoS2 films through treatment with weak plasma and successive exposure to MPA. Sulfur vacancies were formed on the MoS2 film surface by a minimal plasma treatment (10 W and 2 s), and then carboxyl groups were formed through coordinate bonds at the sulfur vacancies using an MPA solution. The stable functionalization of MoS2 using this strategy was confirmed by the performance of an NH3 sensor. The functionalized MoS2-based device at low gas concentrations exhibited a 2.6 times greater sensitivity than the device with the untreated MoS2 film since the adsorption of NH3 onto perfect films is not only difficult but also causes weak charge transfer. Furthermore, the MoS2 films with functional groups provided lower detection limits since they allow for easier gas adsorption. The proposed strategy is suitable for the fabrication of sensors based on other 2D materials in addition to MoS2-based devices.

Author Contributions

Conceptualization, J.H.K., K.-B.P., S.C.R. and N.K.M.; Data Curation, D.K.K., W.S.S. and J.-H.H.; Formal Analysis, D.K.K., W.S.S., J.-H.H. and J.H.K.; Writing—Original Draft Preparation, D.K.K., W.S.S. and J.-H.H.; Writing—Review and Editing, J.H.K., S.C.R., K.-B.P. and N.K.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2015R1A6A3A04060326).

Acknowledgments

W.S.S. and D.K.K. contributed equally to this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.B.; Tang, T.T.; Girit, C.; Hao, Z.; Martin, M.C.; Zettl, A.; Crommie, M.F.; Shen, Y.R.; Wang, F. Direct observation of a widely tunable bandgap in bilayer graphene. Nature 2009, 459, 820–823. [Google Scholar] [CrossRef] [PubMed]
  2. Ni, Z.H.; Yu, T.; Luo, Z.Q.; Wang, Y.Y.; Liu, L.; Wong, C.P.; Miao, J.M.; Huang, W.; Shen, Z.X. Probing charged impurities in suspended graphene using raman spectroscopy. ACS Nano 2009, 3, 569–574. [Google Scholar] [CrossRef] [PubMed]
  3. Yan, W.; He, W.Y.; Chu, Z.D.; Liu, M.X.; Meng, L.; Dou, R.F.; Zhang, Y.F.; Liu, Z.F.; Nie, J.C.; He, L. Strain and curvature induced evolution of electronic band structures in twisted graphene bilayer. Nat. Commun. 2013, 4, 1–7. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, Y.N.; Zhan, D.; Liu, L.; Suo, H.; Ni, Z.H.; Thuong, T.N.; Zhao, C.; Shen, Z.X. Thermal dynamics of graphene edges investigated by polarized raman spectroscopy. ACS Nano 2011, 5, 147–152. [Google Scholar] [CrossRef] [Green Version]
  5. Lloyd, D.; Liu, X.H.; Boddeti, N.; Cantley, L.; Long, R.; Dunn, M.L.; Bunch, J.S. Adhesion, Stiffness, and Instability in Atomically Thin MoS2 Bubbles. Nano Lett. 2017, 17, 5329–5334. [Google Scholar] [CrossRef] [Green Version]
  6. Gamble, F.R.; Silbernagel, B.G. Anisotropy of the proton spin–lattice relaxation time in the superconducting intercalation complex TaS2(NH3): Structural and bonding implications. J. Chem. Phys. 1975, 63, 2544. [Google Scholar] [CrossRef]
  7. Clerc, F.; Battaglia, C.; Cercellier, H.; Monney, C.; Berger, H.; Despont, L.; Garnier, M.G.; Aebi, P. Fermi surface of layered compounds and bulk charge density wave systems. J. Phys.-Condens. Mat. 2007, 19, 355002. [Google Scholar] [CrossRef]
  8. Splendiani, A.; Sun, L.; Zhang, Y.B.; Li, T.S.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  9. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [Green Version]
  10. Lembke, D.; Bertolazzi, S.; Kis, A. Single-Layer MoS2 electronics. Acc. Chem. Res. 2015, 48, 100–110. [Google Scholar] [CrossRef]
  11. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  12. Raybaud, P.; Hafner, J.; Kresse, G.; Kasztelan, S.; Toulhoat, H. Structure, energetics, and electronic properties of the surface of a promoted MoS2 catalyst: An ab initio local density functional study. J. Catal. 2000, 190, 128–143. [Google Scholar] [CrossRef]
  13. Pierucci, D.; Henck, H.; Ben Aziza, Z.; Naylor, C.H.; Balan, A.; Rault, J.E.; Silly, M.G.; Dappe, Y.J.; Bertran, F.; Le Fevre, P.; et al. Tunable doping in hydrogenated single layered molybdenum disulfide. ACS Nano 2017, 11, 1755–1761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Lukowski, M.A.; Daniel, A.S.; Meng, F.; Forticaux, A.; Li, L.S.; Jin, S. Enhanced hydrogen evolution catalysis from chemically exfoliated metallic MoS2 nanosheets. J. Am. Chem. Soc. 2013, 135, 10274–10277. [Google Scholar] [CrossRef] [PubMed]
  15. Yin, W.Y.; Yu, J.; Lv, F.T.; Yan, L.; Zheng, L.R.; Gu, Z.J.; Zhao, Y.L. Functionalized Nano-MoS2 with peroxidase catalytic and near-infrared photothermal activities for safe and synergetic wound antibacterial applications. ACS Nano 2016, 10, 11000–11011. [Google Scholar] [CrossRef] [PubMed]
  16. Tedeschi, D.; Blundo, E.; Felici, M.; Pettinari, G.; Liu, B.Q.; Yildrim, T.; Petroni, E.; Zhang, C.; Zhu, Y.; Sennato, S.; et al. Controlled micro/nanodome formation in proton-irradiated bulk transition-metal dichalcogenides. Adv. Mater 2019, 31, 1903795. [Google Scholar] [CrossRef]
  17. Li, H.; Tsai, C.; Koh, A.L.; Cai, L.L.; Contryman, A.W.; Fragapane, A.H.; Zhao, J.H.; Han, H.S.; 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. 2016, 15, 48–53. [Google Scholar] [CrossRef]
  18. 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. [Google Scholar] [CrossRef]
  19. Choudhary, N.; Park, J.; Hwang, J.Y.; Choi, W. Growth of large-scale and thickness-modulated MoS2 nanosheets. ACS Appl. Mater. Int. 2014, 6, 21215–21222. [Google Scholar] [CrossRef] [PubMed]
  20. Hussain, S.; Shehzad, M.A.; Vikraman, D.; Khan, M.F.; Singh, J.; Choi, D.C.; Seo, Y.; Eom, J.; Lee, W.G.; Jung, J. Synthesis and characterization of large-area and continuous MoS2 atomic layers by RF magnetron sputtering. Nanoscale 2016, 8, 4340–4347. [Google Scholar] [CrossRef]
  21. Kang, K.; Xie, S.E.; Huang, L.J.; Han, Y.M.; 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. [Google Scholar] [CrossRef]
  22. Tahir, M.N.; Zink, N.; Eberhardt, M.; Therese, H.A.; Kolb, U.; Theato, P.; Tremel, W. Overcoming the insolubility of molybdenum disulfide nanoparticles through a high degree of sidewall functionalization using polymeric chelating ligands. Angew. Chem. Int. Ed. 2006, 45, 4809–4815. [Google Scholar] [CrossRef]
  23. Chu, X.M.S.; Yousaf, A.; Li, D.O.; Tang, A.L.A.; Debnath, A.; Ma, D.; Green, A.A.; Santos, E.J.G.; Wang, Q.H. Direct covalent chemical functionalization of unmodified two-dimensional molybdenum disulfide. Chem. Mater. 2018, 30, 2112–2128. [Google Scholar] [CrossRef]
  24. Chen, X.; Berner, N.C.; Backes, C.; Duesberg, G.S.; McDonald, A.R. Functionalization of two-dimensional MoS2: On the reaction between MoS2 and organic thiols. Angew. Chem. Int. Ed. 2016, 55, 5803–5808. [Google Scholar] [CrossRef] [PubMed]
  25. Zhou, L.; He, B.Z.; Yang, Y.; He, Y.G. Facile approach to surface functionalized MoS2 nanosheets. Rsc Adv. 2014, 4, 32570–32578. [Google Scholar] [CrossRef]
  26. Nan, H.Y.; Wang, Z.L.; Wang, W.H.; Liang, Z.; Lu, Y.; Chen, Q.; He, D.W.; Tan, P.H.; Miao, F.; Wang, X.R.; et al. Strong Photoluminescence Enhancement of MoS2 through Defect Engineering and Oxygen Bonding. Acs Nano 2014, 8, 5738–5745. [Google Scholar] [CrossRef] [Green Version]
  27. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous lattice vibrations of single- and few-layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [Green Version]
  28. Makarova, M.; Okawa, Y.; Aono, M. Selective adsorption of thiol molecules at sulfur vacancies on MoS2(0001), followed by vacancy repair via S-C dissociation. J. Phys. Chem. C 2012, 116, 22411–22416. [Google Scholar] [CrossRef]
  29. Li, X.; Zhu, H.W. Two-dimensional MoS2: Properties, preparation, and applications. J. Mater. 2015, 1, 33–44. [Google Scholar] [CrossRef] [Green Version]
  30. Xu, J.; Chen, L.; Dai, Y.W.; Cao, Q.; Sun, Q.Q.; Ding, S.J.; Zhu, H.; Zhang, D.W. A two-dimensional semiconductor transistor with boosted gate control and sensing ability. Sci. Adv. 2017, 3, e1602246. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, Y.; Song, J.G.; Park, Y.J.; Ryu, G.H.; Lee, S.J.; Kim, J.S.; Jeon, P.J.; Lee, C.W.; Woo, W.J.; Choi, T.; et al. Self-Limiting layer synthesis of transition metal dichalcogenides. Sci. Rep. UK 2016, 6, 18754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Brown, N.M.D.; Cui, N.Y.; McKinley, A. An XPS study of the surface modification of natural MoS2 following treatment in an RF-oxygen plasma. Appl. Surf. Sci. 1998, 134, 11–21. [Google Scholar] [CrossRef]
  33. Zhao, J.J.; Buldum, A.; Han, J.; Lu, J.P. Gas molecule adsorption in carbon nanotubes and nanotube bundles. Nanotechnology 2002, 13, 195–200. [Google Scholar] [CrossRef]
  34. Andzelm, J.; Govind, N.; Maiti, A. Nanotube-based gas sensors—Role of structural defects. Chem. Phys. Lett. 2006, 421, 58–62. [Google Scholar] [CrossRef] [Green Version]
  35. Ham, S.W.; Hong, H.P.; Kim, J.H.; Min, S.J.; Min, N.K. Effect of oxygen plasma treatment on carbon nanotube-based sensors. J. Nanosci. Nanotechnol. 2014, 14, 8476–8481. [Google Scholar] [CrossRef] [PubMed]
  36. Dolui, K.; Rungger, I.; Sanvito, S. Origin of the n-type and p-type conductivity of MoS2 monolayers on a SiO2 substrate. Phys. Rev. B 2013, 87, 165402. [Google Scholar] [CrossRef] [Green Version]
  37. Watts, P.C.P.; Mureau, N.; Tang, Z.N.; Miyajima, Y.; Carey, J.D.; Silva, S.R.P. The importance of oxygen-containing defects on carbon nanotubes for the detection of polar and non-polar vapours through hydrogen bond formation. Nanotechnology 2007, 18, 175701. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic representation and (b) SEM image of the gas sensor prepared with an MoS2 film functionalized by plasma treatment for 2 s at 10 W and exposure to an 3-Mercaptopropionic acid (MPA) solution overnight.
Figure 1. (a) Schematic representation and (b) SEM image of the gas sensor prepared with an MoS2 film functionalized by plasma treatment for 2 s at 10 W and exposure to an 3-Mercaptopropionic acid (MPA) solution overnight.
Nanomaterials 10 01860 g001
Figure 2. Schematic representation of the efficient and stable functionalization of an MoS2 film.
Figure 2. Schematic representation of the efficient and stable functionalization of an MoS2 film.
Nanomaterials 10 01860 g002
Figure 3. (a) Thickness and (b) Raman spectra of MoS2 films that were plasma treated for different durations. The frequency difference between Raman shifts indicates that the etching of the MoS2 film depended on the treatment time. In addition, the AFM results indicate that the sulfur vacancies were generated in 2 s.
Figure 3. (a) Thickness and (b) Raman spectra of MoS2 films that were plasma treated for different durations. The frequency difference between Raman shifts indicates that the etching of the MoS2 film depended on the treatment time. In addition, the AFM results indicate that the sulfur vacancies were generated in 2 s.
Nanomaterials 10 01860 g003
Figure 4. Chemical compositions of MoS2 films (a) before and (b) after plasma treatment. After the plasma treatment, the S/Mo ratio decreased from 1.91 to 1.51, indicating the creation of sulfur vacancies.
Figure 4. Chemical compositions of MoS2 films (a) before and (b) after plasma treatment. After the plasma treatment, the S/Mo ratio decreased from 1.91 to 1.51, indicating the creation of sulfur vacancies.
Nanomaterials 10 01860 g004
Figure 5. (a) C 1s deconvolution curves and (b) atomic ratios of films treated with different MPA concentrations. The MPA concentration was optimized at 55 mM. The MoS2 film treated at the optimized concentration had 1.8 times more carboxyl groups than the film prepared by the previously reported strategy.
Figure 5. (a) C 1s deconvolution curves and (b) atomic ratios of films treated with different MPA concentrations. The MPA concentration was optimized at 55 mM. The MoS2 film treated at the optimized concentration had 1.8 times more carboxyl groups than the film prepared by the previously reported strategy.
Nanomaterials 10 01860 g005
Figure 6. Responses of (a) untreated and (b) functionalized MoS2-based sensors for different NH3 concentrations at room temperature. (c) Variation in R/R0 of the MoS2 films exposed to different NH3 concentrations, where R0 is the resistance of the device in the outgassed state and R is the steady-state resistance when exposed to the NH3 gas.
Figure 6. Responses of (a) untreated and (b) functionalized MoS2-based sensors for different NH3 concentrations at room temperature. (c) Variation in R/R0 of the MoS2 films exposed to different NH3 concentrations, where R0 is the resistance of the device in the outgassed state and R is the steady-state resistance when exposed to the NH3 gas.
Nanomaterials 10 01860 g006

Share and Cite

MDPI and ACS Style

Seo, W.S.; Kim, D.K.; Han, J.-H.; Park, K.-B.; Ryu, S.C.; Min, N.K.; Kim, J.H. Functionalization of Molybdenum Disulfide via Plasma Treatment and 3-Mercaptopropionic Acid for Gas Sensors. Nanomaterials 2020, 10, 1860. https://doi.org/10.3390/nano10091860

AMA Style

Seo WS, Kim DK, Han J-H, Park K-B, Ryu SC, Min NK, Kim JH. Functionalization of Molybdenum Disulfide via Plasma Treatment and 3-Mercaptopropionic Acid for Gas Sensors. Nanomaterials. 2020; 10(9):1860. https://doi.org/10.3390/nano10091860

Chicago/Turabian Style

Seo, Won Seok, Dae Ki Kim, Ji-Hoon Han, Kang-Bak Park, Su Chak Ryu, Nam Ki Min, and Joon Hyub Kim. 2020. "Functionalization of Molybdenum Disulfide via Plasma Treatment and 3-Mercaptopropionic Acid for Gas Sensors" Nanomaterials 10, no. 9: 1860. https://doi.org/10.3390/nano10091860

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