Next Article in Journal
Quality Assessment of Reconstructed Cow, Camel and Mare Milk Powders by Near-Infrared Spectroscopy and Chemometrics
Previous Article in Journal
The n-Butanol Extract Obtained from the Inner Bark of Tabebuia rosea (Bertol.) DC, Specioside, and Catalposide Induce Leukemia Cell Apoptosis in the Presence of Apicidin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Performance of MoS2 Field-Effect Transistors Using Self-Assembled Monolayers: A Promising Strategy to Alleviate Dielectric Layer Scattering and Improve Device Performance

1
School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
2
Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China
3
Tianjin Key Laboratory of Imaging and Sensing Microelectronic Technology, School of Microelectronics, Tianjin University, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(17), 3988; https://doi.org/10.3390/molecules29173988
Submission received: 14 July 2024 / Revised: 8 August 2024 / Accepted: 15 August 2024 / Published: 23 August 2024
(This article belongs to the Section Materials Chemistry)

Abstract

:
Field-effect transistors (FETs) based on two-dimensional molybdenum disulfide (2D-MoS2) have great potential in electronic and optoelectronic applications, but the performances of these devices still face challenges such as scattering at the contact interface, which results in reduced mobility. In this work, we fabricated high-performance MoS2-FETs by inserting self-assembling monolayers (SAMs) between MoS2 and a SiO2 dielectric layer. The interface properties of MoS2/SiO2 were studied after the inductions of three different SAM structures including (perfluorophenyl)methyl phosphonic acid (PFPA), (4-aminobutyl) phosphonic acid (ABPA), and octadecylphosphonic acid (ODPA). The SiO2/ABPA/MoS2-FET exhibited significantly improved performances with the highest mobility of 528.7 cm2 V−1 s−1, which is 7.5 times that of SiO2/MoS2-FET, and an on/off ratio of ~106. Additionally, we investigated the effects of SAM molecular dipole vectors on device performances using density functional theory (DFT). Moreover, the first-principle calculations showed that ABPA SAMs reduced the frequencies of acoustic and optical phonons in the SiO2 dielectric layer, thereby suppressing the phonon scattering to the MoS2 channel and further improving the device’s performance. This work provided a strategy for high-performance MoS2-FET fabrication by improving interface properties.

1. Introduction

Two-dimensional (2D) semiconductors, including transition metal dichalcogenides (TMDCs), are highly promising for creating transparent, flexible, and wearable electronics and optoelectronics, owing to their ultra-low dimensions, high carrier mobility, and highly effective light absorption capability [1,2,3]. Among these TMDCs, molybdenum disulfide (MoS2) is one of the most promising semiconductor materials, possessing typical n-type characteristics and high stability [4,5]. In theory, MoS2-based field-effect transistors (FETs) are expected to achieve a significant on/off ratio (>109) and a room-temperature mobility of 410 cm2 V−1 s−1 [6]. Different from the Si semiconductor, the 2D-MoS2 has a smooth surface without any dangling bonds; thus, the channel material MoS2 and the dielectric layer are bonded via van der Waals’ force. However, strong phonon scattering at the semiconductor dielectric layer interface has hindered experimental results from achieving these theoretical expectations [7].
Until now, MoS2 has been combined with high-κ dielectric materials including aluminum oxide (Al2O3) and hafnium oxide (HfO2) for the fabrication of field-effect transistors (FETs) [8]. Unfortunately, the amorphous character of most dielectrics and adverse dielectric/MoS2 interactions pose a challenge for eliminating charge-carrier scattering sites and traps [9,10]. An alternative and attractive approach is to introduce ripples into the lattice structure of the MoS2, where the lattice distortions can reduce electron–phonon scattering in 2D materials and thereby enhance the charge carrier mobility [11]. However, few reports have investigated high-performance MoS2 FETs based on the traditional and cost-effective SiO2 dielectric layer due to its low-κ nature and the associated charge traps [12,13,14]. This limitation has hindered the further development of MoS2-based FETs for next-generation nano-electronic devices.
The self-assembled monolayers (SAMs) of organic molecules have garnered considerable attention in surface and interface engineering owing to their ability to spontaneously form ultrathin molecular films at the substrate interface via chemical or physical processes [15]. By altering the functional groups, SAMs allow the facile tuning of surface energy, dipole moment, and chemical reactivity, making them a versatile tool for tailoring surface properties [16,17,18,19]. Remarkably, SAMs with amine-based functionalities have been shown to promote electron accumulation or enhance the charge transfer between functional SAMs and other materials, thereby enabling the more precise tuning of their characteristics [20,21,22,23,24].
In this study, we presented SiO2/SAMs/MoS2 hybrid-structure FETs and investigated three different types of SAMs to modulate the contact interface property of SiO2/MoS2, including ABPA, PFPA, and ODPA. The performances of three types of SiO2/SAMs/MoS2-FETs were studied and the SiO2/ABPA/MoS2-FET showed the best performances with a high carrier mobility of 528.7 cm2 V−1 s−1 and an on/off ratio of ~106, which was better than that of SiO2/MoS2-FETs without SAMs (mobility of ~70.32 cm2 V−1 s−1, on/off ratio of ~103). After inducing ABPA, the surface potential of SiO2 was calculated via density functional theory (DFT). According to the simulations, the interfacial potential of the SiO2/MoS2 interface was positively enhanced by ABPA and, thus, contributed to more charge accumulation and improved carrier mobility. Additionally, the SAMs could effectively reduce phonon scattering from the SiO2 dielectric layer and SiO2/MoS2 interface, which provided a promising strategy for the fabrication of high-performance MoS2-FETs.

2. Results and Discussion

The MoS2-FET device structure is depicted in Figure 1a, featuring a common bottom-gate configuration. Mechanically exfoliated multilayer MoS2 films served as the semiconductor channel, and the LiF (3 nm)/Au (60 nm) acting as the source/drain contacts were deposited via vacuum evaporation. Recent studies have shown that the contact resistance can be significantly reduced with the incorporation of the LiF layer [25]. The substrate used in this study was heavily p-doped Si (~0.05 Ω·cm), which also served as the common bottom-gate electrode. And, a ~300 nm SiO2 layer was utilized as the dielectric layer.
In the present study, three types of phosphonic acid coupling agents were selected based on their dipole moments and polarity, including (perfluorophenyl) methyl phosphonic acid (PFPA), (4-aminobutyl) phosphonic acid (ABPA), and octadecylphosphonic acid (ODPA). The electrostatic surface potential (ESP) maps of the self-assembled monolayers (SAMs) revealed the variable dipole moments associated with their functional groups. Due to the significant electronegativity of fluorine atoms, PFPA (Figure 1b) can enhance the hole accumulation at the dielectric layer interface, which is beneficial for p-type transport. On the other hand, ABPA containing the NH2 functional group (Figure 1c) and ODPA terminated with CH3 (Figure 1d) exhibit certain electron-donating properties, which can enhance the electron accumulation at the dielectric layer interface to a certain extent and facilitate n-type transport. During the modification processes (Figure S1), PAs reacted with hydroxyl groups and the bridging oxygen on the surface of SiO2 [26]. The chemical states of the carbon atoms in the SAMs molecules were examined using X-ray photoelectron spectroscopy (XPS) analysis, as shown in Figure 2a–c. The binding energy of each sample was calibrated with respect to the peaks at 103.6 eV corresponding to the Si 2p of the pristine SiO2/Si substrate. The C1s peaks of the C-P bonds in PFPA, ABPA, and ODPA were located at 286.1 eV, 286.0 eV, and 286.4 eV, respectively. The slight shifts observed in the C1s peaks may be attributed to the differences in the polarity of the SAM molecules [21]. Moreover, the C1s peak of the C-N group was located at 288 eV for ABPA, and by calculating the fitted areas of the C-N and C-P peaks in Figure 3b, we found that the carbon content ratio of C-P to C-N was approximately 1.5, which is higher than the original ratio of C-P/C-N = 1. This deviation may have been caused by the decomposition of C-N bonds due to irradiation by X-rays during the XPS measurements. In order to gain further insights into the chemical states of the elements within the SAMs, X-ray photoelectron spectroscopy (XPS) analysis was conducted. As illustrated in Figure 2d–f, the survey spectra of the three types of SAMs were characterized by peaks that correspond to their respective elements. Evidently, as depicted in the figure, the long alkyl chain of ODPA leads to a significantly stronger diffraction intensity of its C1s compared to that of ABPA and PFPA. Additionally, the N1s peak was observed at 400.4 eV for ABPA-SiO2, whereas the F1s peak was located at 688.3 eV for PFPA-SiO2. All of the aforementioned findings further validate that the three phosphonic acid molecules have conducted effective self-assembly on the surface of SiO2.
Atomic force microscopy (AFM) measurements were conducted to investigate the morphology of SAM films. The SiO2/Si substrates had high flatness and consistent SAM films after being modified with PFPA (Figure 3a), ABPA (Figure 3b), and ODPA (Figure 3c). According to the analysis of AFM morphologies, the root-mean-square (RMS) roughness of PFPA-SiO2, ABPA-SiO2, and ODPA-SiO2 are 0.43, 0.33, and 0.84 nm, respectively. Figure 3d–f depicted the thicknesses of the self-assembled monolayers (SAMs) of PFPA, ABPA, and ODPA, which measured approximately 0.65 nm, 0.68 nm, and 2.2 nm, respectively. These measurements were in agreement with the height of a single molecule, thus indicating the formation of stable and uniform PA monolayers on the surface of SiO2 through self-assembly [27]. The pink bulges represent multilayer phosphonic acid (PA) molecule aggregations because of the cross-link interactions with other SAM precursors in a liquid environment [28]. Additionally, we conducted water contact angle tests on the three self-assembled monolayers, as shown in Figure S2. The contact angles were 108°(ODPA/SiO2), 103°(PFPA/SiO2), and 69°(ABPA/SiO2), respectively. ODPA/SiO2 and PFPA/SiO2 exhibit excellent hydrophobicity, mainly attributed to the long alkyl chain of ODPA and the F elements in PFPA. Correspondingly, due to the presence of the -NH2 head group in ABPA, it shows a certain degree of hydrophilicity. Overall, all three self-assembled molecules form stable and uniform monolayers on the SiO2 surface.
The performances of MoS2-FET devices modified with three different types of SAMs and a SiO2/MoS2-FET (without SAM) are presented in Figure 4. The drain current (Ids) exhibited a monotonic increase with the gate voltages (Vg) ranging from −40 to +40 V, indicating a typical n-channel transistor behavior. The on/off ratio and subthreshold swing (SS) of the SiO2/MoS2-FET device were extracted to be ∼1 × 103 and ~12 V/dec from the saturation transfer characteristics as shown in the inset of Figure 4a (with a semi-log scale). The PFPA-modified device (Figure 4c) exhibited an improved on/off ratio of 2 × 103 and an SS of ~10 V/dec. Similarly, the ODPA-modified device (Figure 4e) displayed an on/off ratio of ~1.8 × 106 and an SS of approximately 5 V/dec, and the ABPA-modified device showed an on/off ratio of ~1.7 × 106 (Figure 4g), and the lowest SS of ~4 V/dec among these. Furthermore, we investigated the influence of self-assembled molecules on the interface defect density (DIT); the evaluation of DIT can be experimentally carried out or theoretically estimated according to Equation (1) [29]:
S S = k T l n 10 e ( 1 + e 2 C S D I T )
where k is the Boltzmann constant, T is the temperature, e is the elementary charge, CS is the specific capacitance of the gate dielectric, and DIT is the interfacial defect density. Through this relation, these associated defect densities were estimated to be ∼1.43 × 1014 eV−1 cm−2 (SiO2/MoS2-FET), 1.19 × 1014 eV−1 cm−2 (PFPA), 5.91 × 1013 eV−1 cm−2 (ODPA), and 4.71 × 1013 eV−1 cm−2 (ABPA), respectively. These results indicate that all three self-assembled molecules can, to a certain extent, reduce the interface defect density, decrease the capture of carriers, and enable more efficient current switching between the on-state and off-state, thereby achieving a higher on/off current ratio compared to the unmodified device.
The electron field-effect mobility of all four types of MoS2-FETs can be calculated from the transfer characteristics and Equation (2) [11],
μ = 1 C o x L W 1 V d s g m
where Cox is the capacitance (accumulation capacitance) per unit area, L and W are the channel length and width respectively, and gm is the transconductance of the FET. Firstly, the transconductances of the four types of MoS2-FETs were calculated for the SiO2/MoS2-FET (Figure S3a) and those modified with PFPA (Figure S3b), ODPA (Figure S3c), and ABPA (Figure S3d), respectively. As shown in Figure S4, the calculated highest and average mobilities were 70.32 cm2 V−1 s−1 and 54.9 ± 12.3 cm2 V−1 s−1 (SiO2/MoS2-FET), 64.3 cm2 V−1 s−1 and 57.5 ± 6.8 cm2 V−1 s−1 (PFPA), 172.6 cm2 V−1 s−1 and 161.9 ± 9.5 cm2 V−1 s−1 (ODPA), and 528.7 cm2 V−1 s−1 and 511.1 ± 18.2 cm2 V−1 s−1 (ABPA), respectively.
Additionally, the hysteresis properties of these devices were analyzed according to the threshold voltage shift (Figure 5, The red and black arrows in the figure represent the forward and reverse scans respectively), and they were strongly influenced by the dielectric layer and atmospheric conditions [21]. Specifically, the SiO2/MoS2-FET (Figure 5a) and PFPA-(Figure 5b), ODPA-(Figure 5c), and ABPA-(Figure 5d) modified devices exhibited threshold voltage shifts of 26.5 V, 30.9 V, 8 V, and 10 V, respectively. The SiO2/ODPA/MoS2 and SiO2/ABPA/MoS2 FETs had smaller hysteresis compared to the MoS2/SiO2 and SiO2/PFPA/MoS2 FETs. And, the obvious hysteresis in the SAMS-modified FETs might be attributed to the adsorption of H2O or O2 molecules at the channel surface and the dangling bonds with SAMs [30].
The comparisons of the semiconductor performances among the four types of MoS2-FETs are presented in Figure 6. In contrast to SiO2/MoS2-FET, the SS and hysteresis of SiO2/PFPA/MoS2-FET increased, while those of SiO2/ODPA/MoS2-FET and SiO2/ABPA/MoS2-FET decreased significantly (As shown in the blue and pink shaded parts in Figure 6a). Remarkably, the mobility and on/off ratio were enhanced, and the ABPA-treated sample exhibited the highest mobility and on/off ratio among all the samples investigated (As shown in the purple and green shaded parts in Figure 6b), surpassing the mobility values reported in previous studies of MoS2-FETs based solely on SiO2 dielectric layers [31].
Moreover, we analyzed the reasons for the device performance improvements and working mechanisms of SAMs in MoS2 FETs. In general, the enhancement of mobility has been explained by taking the electrostatic potential (VSAMs) generated by SAMs into consideration [32]. The evaluation of the potential can be experimentally carried out or theoretically estimated according to the Helmholtz equation:
V S A M s = N S A M P z ε 0 ε S A M
where NSAMs is the molecular density of the SAMs (the number of molecules per unit area), PZ is the perpendicular component of the dipole moment of the SAMs, ε0 is the vacuum permittivity, and εSAM is the relative permittivity of the SAMs. In the estimation, we assumed that NSAMs was approximately equal to 4.55–5.4 × 1014 cm−2, and εSAM was 2~3 [33,34]. The PZ values of PFPA, ABPA, and ODPA were calculated and are shown in Table 1, according to the density functional theory (DFT) method at the B3LYP/D95 level using the model compounds (Figure 7a).
Evidently, the highest electrostatic potential (VSAMs) was induced by ABPA-SAMs among three molecules, and the corresponding VSAMs formed a space-charge layer at the semiconductor–dielectric interface. As we previously stated, the PFPA-modified substrate had a negative surface potential, and the existing electronic coupling between MoS2 and PFPA-SAMs resulted in a positive space-charge layer in the MoS2 semiconducting layer at the interface (Figure 7b). Electron transport in the MoS2 semiconducting layer could be reduced because the generated positive charge acted as trap sites for electrons injected from the drain electrode. Thus, a larger gate voltage was required to counteract the effects of the generated positive charge by PFPA-SAMs, which made the carrier mobility lower and also increased the hysteresis of the MoS2-FETs. On the contrary, the ABPA-SAMs exhibited a higher positive surface potential compared to the ODPA-SAMs. The magnitude of this difference was quantified as +800 mV for VSAMs = (VABPAVODPA). The electronic coupling between MoS2 and ABPA-SAMs contributed to the creation of a more extensive negative space-charge layer in the MoS2 semiconducting layer at the interface (Figure 7c). This generated negative space-charge layer had a profound impact on carrier migration by enhancing it effectively. To promote the accumulation of charges in the dielectric layer during n-channel operation, it was crucial to create negative charges within the MoS2 semiconducting layer. This was achieved through the generation of a space-charge layer, which was brought about by the action of SAMs. Our analysis suggested that the development of this space-charge layer was a necessary condition for realizing the enhanced performance of MoS2 FETs.
To validate the effect of self-assembled molecules on the phonon scattering of the dielectric layers, we carried out density functional theory (DFT) calculations of SiO2 and ABPA-modified SiO2, and the computational methods of the phonon density of states (DOS) are shown in the Supplementary Materials (SII). The SiO2-terminated surface along the (101) direction was considered as a model atomic structure of unmodified SiO2 (Figure 8a), where an ABPA molecule was bound to the SiO2 (101) surface, which was modeled as modified SiO2 (Figure 8b). As shown in Figure 8c, phonons were renormalized due to the interaction between SiO2 and ABPA after binding. An obvious decrease in the frequency of acoustic phonons was observed and a renormalization of optical phonons was as well. This renormalization of optical phonons across a wider range of frequencies weakens their intensity due to the breaking of degeneracy induced by the formation of a covalent bond with ABPA-SiO2. The estimated effective density of scattering provided insight into the impact of renormalized phonons on mobility [35]. These renormalized phonons dramatically decrease the effective density of scattering in SiO2, thus resulting in strongly suppressed electron–phonon scattering and a high-mobility MoS2 FET based on the dielectric layer of SiO2.

3. Conclusions

To summarize, we presented a straightforward approach to modulate the insulating layer of SiO2 in MoS2 FETs using self-assembled monolayers (SAMs), resulting in an impressive transistor performance. Our results demonstrated that the dipole moments of SAMs gave rise to varying values and orientations of space-charge layers. Notably, the PFPA-modified substrate generated a positive space-charge layer, leading to larger hysteresis and lower mobility in MoS2 FETs compared to ODPA- and ABPA-modified substrates. Conversely, the ABPA-modified device exhibited high mobility up to 528.7 cm2 V−1 s−1, surpassing the mathematically predicted limit for the mobility of n-type MoS2 FETs. This exceptional mobility was primarily attributed to the formation of a larger and more negative space-charge layer, which promoted the accumulation of charges in the dielectric layer, as well as a reduction in phonon scattering of the insulating layer. These findings underline the potential of our approach for high-performance electronic devices, which is compatible with existing 2D semiconductor methods and offers a straightforward method for the manufacturing of 2D semiconductor-based devices.

4. Experimental Section

Device Fabrication: FETs in the bottom-gate configuration were fabricated using contact photolithography on the top of p-doped Si substrate with 300 nm of thermally grown SiO2 (Cox = 11.7 nF/cm2). The SiO2/Si substrates modified with three types of phosphonic acid molecules (the modification methods are shown in detail in the Supplementary Materials (SI)). MoS2 layers were then fabricated on the Si/SiO2/SAMs layer via the mechanical exfoliation method. The source and drain regions were patterned on the MoS2 layers via the electron-beam lithography (EBL, Raith e-Line Plus) system. The channel dimensions were 10 mm in width and 10 mm in length. Finally, a 3 nm LiF and 60 nm Au metal layer were deposited via e-beam evaporation, followed by a lift-off process to form the source/drain electrodes.
Instruments and Measurements: The topography of SAM films was characterized by AFM (Bruker Dimension Icon, Billerica, MA, USA). The quality of the SiO2/SAMs was characterized by XPS (Thermo Scientific ESCALAB 250Xi, Waltham, MA, USA). The electrical analyses of the MoS2 were conducted using a semiconductor parameter analyzer (Keithley 4200 SCS, Cleveland, OH, USA) under ambient conditions.
Materials: (Perfluorophenyl)methyl phosphonic acid (PFPA), (4-aminobutyl) phosphonic acid (ABPA), and octadecylphosphonic acid (ODPA) were purchased from Sigma-Aldrich, Shanghai, China.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29173988/s1: The details of the modification process of SiO2 using SAMs. The details of computational methods in ABPA-modified devices. Figure S1. Schematic diagram of the reaction mechanism of phosphonic acids with the hydroxyl groups and bridging oxygen on the SiO2 surface. Figure S2. Transconductance of SAM-modified FETs. Refs. [36,37,38,39,40,41,42] are cited in supplementary file.

Author Contributions

Conceptualization, L.C.; methodology, L.C. and J.W.; formal analysis, L.C.; resources, J.W.; data curation, L.C. and J.W.; writing—original draft preparation, L.C. and J.W.; writing—review and editing, S.W.; supervision, G.Q.; Funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 52173180), the Haihe Laboratory of Sustainable Chemical Transformations (ZYTS202103), and the Key Research and Development Project of Tianjin (19ZXNCGX00020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jones, A.M.; Yu, H.; Ghimire, N.J.; Wu, S.; Aivazian, G.; Ross, J.S.; Zhao, B.; Yan, J.; Man-drus, D.G.; Xiao, D.; et al. Optical generation of excitonic valley coherence in monolayer WSe2. Nat. Nanotechnol. 2013, 8, 634–638. [Google Scholar] [CrossRef]
  2. Pak, S.; Lee, J.; Lee, Y.; Jang, A.; Ahn, S.; Ma, K.Y.; Cho, Y.; Hong, J.; Lee, S.; Jeong, H.Y.; et al. Strain-mediated interlayer coupling effects on the excitonic behaviors in an epitaxially grown MoS2/WS2 van der waals heterobilayer. Nano Lett. 2017, 17, 5634–5640. [Google Scholar] [CrossRef] [PubMed]
  3. Chhowalla, M.; Jena, D.; Zhang, H. Two-dimensional semiconductors for transistors. Nat. Rev. Mater. 2016, 1, 16052. [Google Scholar] [CrossRef]
  4. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Integrated circuits and logic operations based on single-layer MoS2. ACS Nano 2011, 5, 9934–9938. [Google Scholar] [CrossRef] [PubMed]
  5. Das, S.; Appenzeller, J. Where does the current flow in two-dimensional layered systems? Nano Lett. 2013, 13, 3396–3402. [Google Scholar] [CrossRef] [PubMed]
  6. Kaasbjerg, K.; Thygesen, K.S.; Jacobsen, K.W. Phonon-limited mobility in n-type single-layer MoS2 from first principles. Phys. Rev. B 2012, 85, 115317. [Google Scholar] [CrossRef]
  7. Ma, N.; Jena, D. Charge scattering and mobility in atomically thin semiconductors. Phys. Rev. X 2014, 4, 011043. [Google Scholar] [CrossRef]
  8. Illarionov, Y.Y.; Knobloch, T.; Jech, M.; Lanza, M.; Akinwande, D.; Vexler, M.I.; Mueller, T.; Lemme, M.C.; Fiori, G.; Schwierz, F.; et al. Insulators for 2D nanoelectronics: The gap to bridge. Nat. Commun. 2020, 11, 3385. [Google Scholar] [CrossRef]
  9. Kim, H.G.; Lee, H. Atomic layer deposition on 2D materials. Chem. Mater. 2017, 29, 3809–3826. [Google Scholar] [CrossRef]
  10. Li, W.; Zhou, J.; Cai, S.; Yu, Z.; Zhang, J.; Fang, N.; Li, T.; Wu, Y.; Chen, T.; Xie, X.; et al. Uniform and ultrathin high-kappa gate dielectrics for two-dimensional electronic devices. Nat. Electron. 2019, 2, 563–571. [Google Scholar] [CrossRef]
  11. Ng, H.K.; Xiang, D.; Suwardi, A.; Hu, G.; Yang, K.; Zhao, Y.; Liu, T.; Cao, Z.; Liu, H.; Li, S.; et al. Improving carrier mobility in two-dimensional semiconductors with rippled materials. Nat. Electron. 2022, 5, 489–496. [Google Scholar] [CrossRef]
  12. Liu, Y.; Guo, J.; Zhu, E.; Liao, L.; Lee, S.; Ding, M.; Shakir, I.; Gambin, V.; Huang, Y.; Duan, X. Approaching the Schottky–Mott limit in van der Waals metal–semiconductor junctions. Nature 2018, 557, 696–700. [Google Scholar] [CrossRef] [PubMed]
  13. Baugher, B.W.H.; Churchill, H.O.H.; Yang, Y.; Jarillo-Herrero, P. Intrinsic electronic transport properties of high-quality monolayer and bilayer MoS2. Nano Lett. 2013, 13, 4212–4216. [Google Scholar] [CrossRef] [PubMed]
  14. Yu, Z.; Pan, Y.; Shen, Y.; Wang, Z.; Ong, Z.; Xu, T.; Xin, R.; Pan, L.; Wang, B.; Sun, L.; et al. Towards intrinsic charge transport in monolayer molybdenum disulfide by defect and interface engineering. Nat. Commun. 2014, 5, 5290. [Google Scholar] [CrossRef]
  15. Paniagua, S.A.; Giordano, A.J.; Smith, O.L.; Barlow, S.; Li, H.; Armstrong, N.R.; Pemberton, J.E.; Bredas, J.; Ginger, D.; Marder, S.R. Phosphonic acids for interfacial engineering of transparent conductive oxides. Chem. Rev. 2016, 116, 7117–7158. [Google Scholar] [CrossRef]
  16. Timpel, M.; Li, H.; Nardi, M.V.; Wegner, B.; Frisch, J.; Hotchkiss, P.J.; Marder, S.R.; Barlow, S.; Bredas, J.; Koch, N. Electrode work function engineering with phosphonic acid monolayers and molecular acceptors: Charge redistribution mechanisms. Adv. Funct. Mater. 2018, 28, 1704438. [Google Scholar] [CrossRef]
  17. Benneckendorf, F.S.; Hillebrandt, S.; Ullrich, F.; Rohnacher, V.; Hietzschold, S.; Jaensch, D.; Freudenberg, J.; Beck, S.; Mankel, E.; Jaegermann, W.; et al. Structure-property relationship of phenylene-based self-assembled monolayers for record low work function of indium tin oxide. J. Phys. Chem. Lett. 2018, 9, 3731–3737. [Google Scholar] [CrossRef]
  18. Wang, B.; Di Carlo, G.; Turrisi, R.; Zeng, L.; Stallings, K.; Huang, W.; Bedzyk, M.J.; Beverina, L.; Marks, T.J.; Facchetti, A. The dipole moment inversion effects in self-assembled nanodielectrics for organic transistors. Chem. Mater. 2017, 29, 9974–9980. [Google Scholar] [CrossRef]
  19. Hotchkiss, P.J.; Jones, S.C.; Paniagua, S.A.; Sharma, A.; Kippelen, B.; Armstrong, N.R.; Marder, S.R. The modification of indium tin oxide with phosphonic acids: Mechanism of Binding, Tuning of Surface Properties, and Potential for Use in Organic Electronic Applications. Acc. Chem. Res. 2012, 45, 337–346. [Google Scholar] [CrossRef] [PubMed]
  20. Kuo, L.; Sangwan, V.K.; Rangnekar, S.V.; Chu, T.C.; Lam, D.; Zhu, Z.; Richter, L.J.; Li, R.; Szydłowska, B.M.; Downing, J.R.; et al. All-printed ultrahigh-responsivity MoS2 nanosheet photodetectors enabled by megasonic exfoliation. Adv. Mater. 2022, 34, 2203772. [Google Scholar] [CrossRef]
  21. Yokota, K.; Takai, K.; Enoki, T. Carrier control of graphene driven by the proximity effect of functionalized self-assembled monolayers. Nano Lett. 2011, 11, 3669–3675. [Google Scholar] [CrossRef]
  22. 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]
  23. Nasr, J.R.; Simonson, N.; Oberoi, A.; Horn, M.W.; Robinson, J.A.; Das, S. Low-power and ultra-thin MoS2 photodetectors on glass. ACS Nano 2020, 14, 15440–15449. [Google Scholar] [CrossRef]
  24. Lu, D.; Chen, Y.; Kong, L.; Luo, C.; Lu, Z.; Tao, Q.; Song, W.; Ma, L.; Li, Z.; Li, W.; et al. Strain-plasmonic coupled broadband photodetector based on monolayer MoS2. Small 2022, 18, 2107104. [Google Scholar] [CrossRef]
  25. Cho, H.; Kang, D.; Lee, Y.; Bae, H.; Hong, S.; Cho, Y.; Kim, K.; Yi, Y.; Park, J.H.; Im, S. Dramatic reduction of contact resistance via ultrathin LiF in two-Dimensional MoS2 field effect transistors. Nano Lett. 2021, 21, 3503–3510. [Google Scholar] [CrossRef]
  26. Bardecker, J.A.; Ma, H.; Kim, T.; Huang, F.; Liu, M.S.; Cheng, Y.; Ting, G.; Jen, A.K.Y. Self-assembled electroactive phosphonic acids on ITO: Maximizing hole-injection in polymer light-emitting diodes. Adv. Funct. Mater. 2008, 18, 3964–3971. [Google Scholar] [CrossRef]
  27. Cheng, H.; Huai, J.; Cao, L.; Li, Z. Novel self-assembled phosphonic acids monolayers applied in N-channel perylene diimide (PDI) organic field effect transistors. Appl. Surf. Sci. 2016, 378, 545–551. [Google Scholar] [CrossRef]
  28. Rozlosnik, N.; Gerstenberg, M.C.; Larsen, N.B. Effect of solvents and concentration on the formation of a self-assembled monolayer of octadecylsiloxane on silicon (001). Langmuir 2003, 19, 1182–1188. [Google Scholar] [CrossRef]
  29. Choi, Y.; Kim, H.; Yang, J.; Shin, S.W. Proton-Conductor-Gated MoS2 Transistors with Room Temperature Electron Mobility of >100 cm2 V−1 s−1. Chem. Mater. 2018, 30, 4527–4535. [Google Scholar] [CrossRef]
  30. Liu, N.; Baek, J.; Kim, S.M.; Hong, S.; Hong, Y.K.; Kim, Y.S.; Kim, H.; Kim, S.; Park, J. Improving the stability of high-performance multilayer MoS2 field-effect transistors. ACS Appl. Mater. Interfaces 2017, 9, 42943–42950. [Google Scholar] [CrossRef]
  31. Guo, Y.; Wei, X.; Shu, J.; Liu, B.; Yin, J.; Guan, C.; Han, Y.; Gao, S.; Chen, Q. Charge trapping at the MoS2-SiO2 interface and its effects on the characteristics of MoS2 metal-oxide-semiconductor field effect transistors. Appl. Phys. Lett. 2015, 106, 103109. [Google Scholar] [CrossRef]
  32. Cui, X.; Lee, G.; Kim, Y.D.; Arefe, G.; Huang, P.Y.; Lee, C.; Chenet, D.A.; Zhang, X.; Wang, L.; Ye, F.; et al. Multi-terminal transport measurements of MoS2 using a van der Waals heterostructure device platform. Nat. Nanotechnol. 2015, 10, 534–540. [Google Scholar] [CrossRef]
  33. Aghamohammadi, M.; Rödel, R.; Zschieschang, U.; Ocal, C.; Boschker, H.; Weitz, R.T.; Barrena, E.; Klauk, H. Threshold-voltage shifts in organic transistors due to self-assembled monolayers at the dielectric: Evidence for electronic coupling and dipolar effects. ACS Appl. Mater. Interfaces 2015, 7, 22775–22785. [Google Scholar] [CrossRef]
  34. Chung, Y.; Verploegen, E.; Vailionis, A.; Sun, Y.; Nishi, Y.; Murmann, B.; Bao, Z. Controlling electric dipoles in nanodielectrics and its applications for enabling air-stable n-channel organic transistors. Nano Lett. 2011, 11, 1161–1165. [Google Scholar] [CrossRef]
  35. Cheng, L.; Zhang, C.; Liu, Y. Why Two-dimensional semiconductors generally have low electron mobility. Phys. Rev. Lett. 2020, 125, 177701. [Google Scholar] [CrossRef]
  36. Kresse, G.; Furthmuller, J. Efficient iIterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  37. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  38. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  39. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  40. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H.J. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. Chem. Phys. 2010, 132, 154104. [Google Scholar]
  41. Henkelman, G.; Uberuaga, B.P.; Jonsson, H.J. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. Chem. Phys. 2000, 113, 9901. [Google Scholar] [CrossRef]
  42. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
Figure 1. The structure of SiO2/SAM/MoS2-FET. (a) Schematic illustration of MoS2-FET. The chemical structures and electrostatic surface potential (ESP) maps of three types of SAMs: (b) PFPA, (c) ABPA, and (d) ODPA.
Figure 1. The structure of SiO2/SAM/MoS2-FET. (a) Schematic illustration of MoS2-FET. The chemical structures and electrostatic surface potential (ESP) maps of three types of SAMs: (b) PFPA, (c) ABPA, and (d) ODPA.
Molecules 29 03988 g001
Figure 2. X-ray photoelectron spectroscopy (XPS) analysis of C 1s spectra in (a) PFPA/SiO2, (b) ABPA/SiO2, and (c) ODPA/SiO2, respectively. XPS survey spectra of (d) PFPA/SiO2, (e) ABPA/SiO2, and (f) ODPA/SiO2, respectively.
Figure 2. X-ray photoelectron spectroscopy (XPS) analysis of C 1s spectra in (a) PFPA/SiO2, (b) ABPA/SiO2, and (c) ODPA/SiO2, respectively. XPS survey spectra of (d) PFPA/SiO2, (e) ABPA/SiO2, and (f) ODPA/SiO2, respectively.
Molecules 29 03988 g002
Figure 3. Surface morphologies of (a) SiO2/PFPA, (b) SiO2/ABPA, and (c) SiO2/ODPA. Section analyses of SAM: (d) PFPA, (e) ABPA, and (f) ODPA.
Figure 3. Surface morphologies of (a) SiO2/PFPA, (b) SiO2/ABPA, and (c) SiO2/ODPA. Section analyses of SAM: (d) PFPA, (e) ABPA, and (f) ODPA.
Molecules 29 03988 g003
Figure 4. The transfer and output properties of the MoS2-FETs. (a,b) SiO2/MoS2-FET (without SAMs) (c,d) SiO2/PFPA/MoS2-FET, (e,f) SiO2/ODPA/MoS2-FET, and (g,h) SiO2/ABPA/MoS2-FET.
Figure 4. The transfer and output properties of the MoS2-FETs. (a,b) SiO2/MoS2-FET (without SAMs) (c,d) SiO2/PFPA/MoS2-FET, (e,f) SiO2/ODPA/MoS2-FET, and (g,h) SiO2/ABPA/MoS2-FET.
Molecules 29 03988 g004
Figure 5. The hysteresis properties of the MoS2-FETs: (a) SiO2/MoS2-FET, (b) PFPA, (c) ODPA, and (d) ABPA.
Figure 5. The hysteresis properties of the MoS2-FETs: (a) SiO2/MoS2-FET, (b) PFPA, (c) ODPA, and (d) ABPA.
Molecules 29 03988 g005
Figure 6. The performance comparisons of the SiO2/SAMs/MoS2-FETs; (a) the subthreshold swing and hysteresis properties. (b) The mobility and on/off ratio of all devices.
Figure 6. The performance comparisons of the SiO2/SAMs/MoS2-FETs; (a) the subthreshold swing and hysteresis properties. (b) The mobility and on/off ratio of all devices.
Molecules 29 03988 g006
Figure 7. Calculation results: (a) Calculation of dipole moments of the SAMs molecules. Schematic of SAM-induced charged surface and formation of a space-charge layer at the interface for (b) PFPA and (c) ABPA.
Figure 7. Calculation results: (a) Calculation of dipole moments of the SAMs molecules. Schematic of SAM-induced charged surface and formation of a space-charge layer at the interface for (b) PFPA and (c) ABPA.
Molecules 29 03988 g007
Figure 8. DFT-simulated relaxed structures of (a) unmodified SiO2 and (b) ABPA-modified SiO2. (c) Calculated phonon DOS in unmodified SiO2 and ABPA-modified SiO2, with frequencies of acoustic (optical) phonons.
Figure 8. DFT-simulated relaxed structures of (a) unmodified SiO2 and (b) ABPA-modified SiO2. (c) Calculated phonon DOS in unmodified SiO2 and ABPA-modified SiO2, with frequencies of acoustic (optical) phonons.
Molecules 29 03988 g008
Table 1. DFT-calculated molecular dipole (Pz) and calculated (VSAMs) surface potentials of PFPA-, ABPA-, and ODPA-SAMs.
Table 1. DFT-calculated molecular dipole (Pz) and calculated (VSAMs) surface potentials of PFPA-, ABPA-, and ODPA-SAMs.
SAMsP (debye)PZ (debye)VSAMs (mV)
PFPA−0.980.72−496 to −589
ABPA2.152.04+1405 to +1668
ODPA1.441.06+730 to+868
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cao, L.; Wei, J.; Li, X.; Wang, S.; Qin, G. Enhancing the Performance of MoS2 Field-Effect Transistors Using Self-Assembled Monolayers: A Promising Strategy to Alleviate Dielectric Layer Scattering and Improve Device Performance. Molecules 2024, 29, 3988. https://doi.org/10.3390/molecules29173988

AMA Style

Cao L, Wei J, Li X, Wang S, Qin G. Enhancing the Performance of MoS2 Field-Effect Transistors Using Self-Assembled Monolayers: A Promising Strategy to Alleviate Dielectric Layer Scattering and Improve Device Performance. Molecules. 2024; 29(17):3988. https://doi.org/10.3390/molecules29173988

Chicago/Turabian Style

Cao, Li, Junqing Wei, Xianggao Li, Shirong Wang, and Guoxuan Qin. 2024. "Enhancing the Performance of MoS2 Field-Effect Transistors Using Self-Assembled Monolayers: A Promising Strategy to Alleviate Dielectric Layer Scattering and Improve Device Performance" Molecules 29, no. 17: 3988. https://doi.org/10.3390/molecules29173988

Article Metrics

Back to TopTop