Next Article in Journal
mTanh: A Low-Cost Inkjet-Printed Vanishing Gradient Tolerant Activation Function
Previous Article in Journal
Full-Bridge DC-DC Converter with Synchronous Rectification Based on GaN Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultra-Thin Al2O3 Grown by PEALD for Low-Power Molybdenum Disulfide Field-Effect Transistors

School of Resources, Environment and Materials, Guangxi University, Nanning 530004, China
*
Author to whom correspondence should be addressed.
J. Low Power Electron. Appl. 2025, 15(2), 26; https://doi.org/10.3390/jlpea15020026
Submission received: 19 March 2025 / Revised: 28 April 2025 / Accepted: 28 April 2025 / Published: 30 April 2025

Abstract

:
The lack of ultra-thin, controllable dielectric layers poses challenges for reducing power consumption in 2D FETs. In this study, plasma-enhanced atomic layer deposition was employed to fabricate a highly reliable, ultra-thin aluminum oxide (Al2O3) dielectric layer with a thickness of 4 nm. The Al2O3 film grown on highly conductive silicon substrates demonstrated a maximum breakdown field of 5.98 MV/cm and a leakage current density as low as 2.48 × 10−7 A/cm2 at 1 MV/cm. MoS2 FETs incorporating this Al2O3 gate dielectric exhibited high-performance n-type characteristics at a low operating voltage of 1 V, achieving a subthreshold swing (SS) of 65 mV/dec, a threshold voltage (Vth) of −0.96 V, a high carrier mobility (μ) of 34.85 cm2·V−1·s−1, and an on/off current ratio exceeding 106. These results highlight the potential of Al2O3 in enabling low-power 2D electronic devices for post-Moore applications.

1. Introduction

The dielectric layer in field-effect transistors (FETs) serves as the fundamental enabler of electrostatic control over channel carriers, making its optimization a cornerstone of semiconductor device physics [1,2,3,4,5]. This critical component operates at the quantum-mechanical frontier, where its thickness approaches atomic dimensions (<1 nm in state-of-the-art nodes) while simultaneously needing to satisfy contradictory requirements: achieving high capacitance density for strong gate control and maintaining robust insulation against leakage currents. Silicon dioxide (SiO2) has dominated microelectronics as the primary gate dielectric for over five decades due to its exceptional insulating properties (bandgap ~ 9 eV), near-perfect Si/SiO2 interface quality with ultra-low defect density (<1010 cm−2 eV−1), and superior process compatibility with thermal oxidation techniques [6,7,8]. However, the relentless scaling of silicon FETs into the sub-nanometer regime has exposed fundamental limitations of SiO2’s intrinsically low dielectric constant (k ≈ 3.9). The central paradox of SiO2 scaling arises from its thickness-dependent quantum tunneling behavior, mathematically described by the Wentzel-Kramers-Brillouin (WKB) approximation [9]:
J e x p ( 4 π t 2 m Φ h )
where J is tunneling current density, t is oxide thickness, m* is effective electron mass, and Φ is barrier height. At thicknesses below 2 nm, direct tunneling currents exceed 102 A/cm2, comparable to semiconductor junction currents, rendering conventional oxides impractical for low-power applications. This creates an untenable trade-off: reducing t to maintain capacitance exponentially increases leakage power dissipation, while thicker layers degrade electrostatic control [10]. For 2 nm SiO2 at 1 V operation, leakage power surpasses 104 W/cm2, inducing catastrophic thermal breakdown in dense integrated circuits (ICs). High-k dielectrics (k > 10) address this through enhanced physical thickness while maintaining equivalent oxide thickness (EOT), effectively suppressing Fowler-Nordheim tunneling by 3–5 orders of magnitude [7]. These limitations manifest in critical reliability issues, including thermal breakdown and yield losses driven by exponential increases in power dissipation at advanced nodes.
This technological impasse has driven industrial adoption of high-k gate dielectrics. Materials with enhanced permittivity enable equivalent capacitance at greater physical thicknesses through the relationship C = kε0A/d, effectively suppressing quantum tunneling leakage while maintaining electrostatic control [1,11,12]. Recent advances in 2D material-based FETs further highlight high-k dielectrics’ critical role in achieving low-voltage operation and enhanced breakdown resilience: Xu et al. [13] demonstrated ultra-thin gadolinium oxychloride (GdOCl) nanosheets synthesized via chloride-assisted chemical vapor deposition (CVD), achieving a high dielectric constant (k = 15.3), exceptional breakdown field strength (>9.9 MV/cm), and minimal gate leakage current (~10−6 A/cm2). Yin et al. [14] developed 2D single-crystalline Gd2O5 through van der Waals epitaxy, exhibiting superior dielectric properties (k = 25.5) with a wide bandgap and ultra-low leakage current (~10−4 A/cm2). Chen et al. [15] reported CVD-grown Bi2SiO5 crystals with k > 30, a large bandgap (~3.8 eV), and van der Waals integration capability. Yi et al. [16] engineered surface-tension-assisted Ga2O3 dielectrics on MoS2 via squeeze printing, achieving k ≈ 30 and a remarkable EOT of ~0.4 nm. Li et al. [17] fabricated 2D perovskite Sr2Nb3O10 dielectrics (k = 24.6) with moderate bandgaps, demonstrating compatibility with diverse 2D channel materials. Despite these material innovations, atomic layer deposition (ALD), grown dielectric films remain the industrial standard due to their unmatched process controllability and scalability [18]. A critical challenge in high-k materials lies in the inherent trade-off between dielectric constant and bandgap (Eg)—higher k values typically correlate with reduced Eg, necessitating meticulous optimization of film quality to suppress leakage currents. Among high-k candidates, Al2O3 emerges as a premier gate dielectric due to its optimal combination of a wide bandgap (~8.7 eV), high k (~9), robust breakdown field (>10 MV/cm), strong interfacial adhesion, and exceptional thermal/chemical stability [19]. The self-limiting growth mechanism of plasma-enhanced ALD (PEALD) enables atomic-level thickness control (<±1% uniformity), producing pinhole-free Al2O3 films with sub-nanometer precision. Bolshakov et al. [20] used an atomic layer deposition process to deposit HfO2 (~10 nm) and Al2O3 (~15 nm) on Si to fabricate MOS capacitors and performed a synthesis gas annealing study to achieve an optimal annealing temperature of 400 °C (1 h) to determine the effect of annealing on interface traps and obtained high-quality dielectric substrates for top-gate MoS2 FETs. Jeong et al. [21] studied the effect of Al2O3 as a passivation layer on the performance of MoS2 TFT, and it is proved that the passivation of Al2O3 film can improve the performance of MoS2. Tanigawa et al. [22] mainly used the THALD method to study the capacitance-voltage (C-V) characteristics of atomic layer deposition Al2O3 film on sulfide MoS2 film and TiN top-gate metal-insulator-semiconductor MOSFET with sulfide MoS2 channel. Hong et al. [23] also tested MoS2 with an Al2O3 passivation layer, and the threshold voltage of the drift of the passivation layer was reduced from 10 V to 7.4 V, which proved the protective effect of the Al2O3 dielectric layer on the MoS2 device.
This work systematically investigates PEALD-optimized Al2O3 dielectric growth using trimethylaluminum [Al(CH3)3] (TMA) and oxygen plasma precursors. Process parameters were varied across growth temperatures: 100 °C, 150 °C, 200 °C, 250 °C, and 300 °C; oxygen plasma power: 20 W, 40 W, 60 W, 80 W, 100 W, and 120 W. The optimized Al2O3 film (200 °C, 80 W) demonstrated exceptional dielectric integrity even at 4 nm thickness, enabling high-performance MoS2 FETs with high breakdown voltage (>5.98 MV/cm), low threshold voltage (Vt = −0.96 V), enhanced carrier mobility (μ > 34.85 cm2·V−1·s−1), near-ideal subthreshold swing (SS = 65 mV/dec), and ultrahigh on/off current ratio (>106). These characteristics satisfy stringent requirements for low-power nanoelectronics, positioning PEALD-Al2O3 as a critical enabler for next-generation 2D material-based devices.

2. Results and Discussion

2.1. Characterization of Al2O3 Thin Films

Atomic layer deposition technology can precisely control film thickness by alternating the flow of precursors in the reaction chamber. To better explore the relationship between cycle number and thickness, Al2O3 films were grown firstly with 25, 35, 40, 60, 80, 100, and 125 cycles under the conditions of a growth temperature of 200 °C and oxygen plasma power of 80 W. Subsequently, the ellipsometer was used to test and fit the thickness of the samples (Figure S1), showing the relationship between cycle number and Al2O3 thickness. It can be observed that as the number of growth cycles increases, the thickness of the Al2O3 film increases linearly, with each cycle growing approximately ~ 0.1 nm of Al2O3, consistent with literature descriptions [24], proving the controllability of the PEALD thickness.
To investigate the suitable growth conditions for Al2O3 films, atomic force microscopy (AFM) characterization was conducted on 12.5 nm Al2O3 films deposited under varying growth parameters, with arithmetic mean roughness (Ra) values extracted through XEI software (Version 5.0.1) analysis of 2 × 2 μm2 scan areas. To identify optimal growth conditions for Al2O3 thin films, AFM characterization was systematically conducted on 12.5 nm thick films deposited under temperature/power conditions with varied parameters: 100 °C/80 W, 200 °C/80 W, 200 °C/20 W, 200 °C/120 W, 300 °C/80 W. As illustrated in Figure 1a, the 200 °C/80 W sample demonstrated superior surface quality, lower than other conditions. This exceptional smoothness correlates with enhanced precursor decomposition efficiency at 200 °C, where thermal energy sufficiently activates TMA ligands without inducing excessive plasma-induced surface damage.
Figure 1b shows the XPS analysis of 12.5 nm thick films grown at 100 °C and 20 W, 200 °C and 80 W, and 300 °C and 120 W, and a 4 nm thick film grown at 200 °C and 80 W. The O1s peak for Al2O3 is at 531.1 eV, and the C-O peak from methyl groups in trimethylaluminum is at 532.4 eV [25]. It can be seen that under the same thickness, the 200 °C and 80 W condition has the fewest oxygen defects, further indicating the density of the Al2O3 film. A thickness-dependent increase in oxygen defect density was observed, with the oxygen vacancy concentration rising by ~14.64% as film thickness decreased from 12.5 nm to 4 nm. This phenomenon is attributed to cycle-dependent precursor reaction kinetics—reduced ALD cycles (from 125 to 40) limit complete ligand exchange between trimethylaluminum methyl groups and oxygen plasma, resulting in residual C-O bonding configurations. The reduction in the number of layers of Al2O3 results in a decrease in the proportion of Al-O bonds in the overall O1s spectra. Nevertheless, even the thinnest 4 nm films exhibited lower defect density compared to 12.5 nm counterparts grown under suboptimal conditions (150 °C/20 W).

2.2. Breakdown Field of Al2O3 Thin Films

To investigate the effect of different growth temperatures and oxygen plasma power on the electrical properties of Al2O3 films. Electrical experiments on Al2O3 films grown under different temperatures and power conditions were conducted. Figure 2a shows the breakdown voltage of 125-cycle Al2O3 films grown with an oxygen plasma power of 80 W at different temperatures (100 °C, 150 °C, 200 °C, 250 °C, and 300 °C). Among these, the breakdown voltage of the Al2O3 film grown at 200 °C is the highest, reaching 6.04 V. Figure 2b shows the breakdown voltage of 125-cycle Al2O3 films grown at 200 °C with different oxygen plasma power levels (20 W, 40 W, 60 W, 80 W, 100 W, and 120 W). The breakdown voltage is highest for the film grown with an oxygen plasma power of 80 W. Figure 2c shows the breakdown voltage of Al2O3 films grown under the optimal conditions of 200 °C and 80 W, with the number of cycles reduced to decrease the film thickness. It can be seen that the breakdown voltage of films with thicknesses of 2.5 nm and 3.5 nm does not exceed 1 V, which does not meet the requirements for gate dielectrics in field-effect transistors. However, when the thickness is 4 nm, the breakdown voltage increases significantly to 2.39 V. Figure 2d–f present statistical box plots of breakdown voltage reproducibility measured across ten discrete locations on PEALD-grown Al2O3 films under varying growth conditions, demonstrating exceptional process stability (details can be found in Figures S2–S4). The tight clustering of breakdown voltages highlights the superior thickness uniformity and defect-free morphology inherent to the PEALD growth mechanism.

2.3. Dielectric Properties of Al2O3 Thin Films

For capacitance-voltage (C-V) measurements, a parallel plate capacitor setup was used, as shown in Figure 3a. A 12.5 nm thick Al2O3 film was grown on heavily doped silicon, and Au electrodes were evaporated. The relative permittivity can be calculated using the formula:
ε r = C × d A × ε 0
where εr is the permittivity of Al2O3, C is the measured capacitance, d is the sample thickness, A is the area of the Au electrode, and ε0 is the permittivity of free space [26,27]. Figure 3a shows the capacitance density of Al2O3 films grown under different conditions over a frequency range of 1 kHz to 1 MHz. It can be observed that the measured capacitance density decreases with increasing voltage and frequency, which is typically observed in incipient ferroelectric ceramics and asymmetric electrode configurations, mainly attributed to the reduction in dipole polarization at high frequencies [28]. Figure 3b shows the capacitance density of Al2O3 films grown under different conditions. It can be seen that for the same thickness of 12.5 nm, the film grown at 200 °C and 80 W has the highest capacitance density, reaching 0.6 μF/cm2, further proving that these are the optimal growth conditions. As the film thickness decreases, the capacitance density of the Al2O3 film increases, reaching 1.73 μF/cm2 at 4 nm.
The relative EOT was evaluated using the formula:
E O T = ε S i O 2 × t ε r
where εSiO2 is the permittivity of SiO2, t is the sample thickness, and εr is the relative permittivity of Al2O3 [10]. Figure 3c shows that as the thickness decreases, both the relative permittivity and the effective oxide thickness of the Al2O3 film decrease. This phenomenon likely arises from dead-layer effects, where interfacial defect layers (0.5–2 nm thick) form at dielectric/electrode interfaces (e.g., metal-oxide junctions) due to lattice mismatch, oxygen vacancy aggregation, and impurity interdiffusion. These structurally compromised regions exhibit constrained dipole polarization capability, rendering them an electrically inactive “dead layer” with diminished response to applied electric fields, consistent with literature reports.
Table 1 summarizes the characteristics of various gate dielectrics obtained through different methodologies, where d denotes the dielectric thickness, Ci represents the areal capacitance in the low-frequency region, and Jleak indicates the leakage current density at 1 MV/cm. For thin-film transistor architectures, an ideal gate dielectric material should exhibit high dielectric breakdown strength and near-zero leakage current density. In this context, Al2O3 films fabricated via plasma-enhanced atomic layer deposition technology demonstrate these desirable characteristics when employed as gate dielectrics in electronic devices, rendering them an excellent candidate. The 4-nm-thick Al2O3 dielectric layer developed in this work exhibits exceptional advantages compared with other reported works: its ultra-thin geometry combined with remarkably high low-frequency areal capacitance. This achievement stems from the exploration of optimal PEALD parameters and interface contact optimization of the Al2O3 dielectric layer, which effectively mitigates suppressed oxygen vacancy aggregation and reduces impurity interdiffusion at the dielectric/electrode interface. These synergistic effects substantially inhibit the formation of interfacial defect layers and significantly alleviate the “dead-layer” effect.

2.4. Characterization of Top-Gated MoS2 FETs with Al2O3 Thin Films

To further investigate the application of ultra-thin Al2O3 films as dielectric layers in MOSFETs, MoS2 field-effect transistors were fabricated with a 4 nm thick Al2O3 film as the dielectric layer. Figure 4a shows the fabrication process of the MOSFET, where a top-contact process was used. Au bottom electrodes were evaporated, MoS2 was transferred, and a 4 nm thick Al2O3 dielectric layer was grown at 200 °C and 80 W. Subsequently, top Au electrodes were transferred onto the dielectric layer. Figure 4b shows the thickness of the MoS2 material used. Figure 4c,d show the scanning electron microscope (SEM) and energy dispersive spectrometer (EDS) images of the transistor, respectively. The morphology and elemental composition of the device can be observed, with the bottom and top electrodes being Au, the Al2O3 dielectric layer distributed uniformly, and the outline of the MoS2 layer clearly visible.

2.5. Electrical Performance of MoS2 FETs

To test the transistor’s performance, electrical measurements were conducted on the MoS2 field-effect transistor. The transfer and output characteristics are shown in Figure 5a and Figure 5b, respectively. From Figure 5a, it can be seen that the MoS2 transistor exhibits typical n-type behavior, with a threshold voltage estimated to be −0.96 V. The device achieves an on/off ratio greater than 106, a high carrier mobility of 34.85 cm2·V−1·s−1, and a low subthreshold swing of 65 mV/dec within the measurement range of −1 to 1 V. Figure 5b shows the output characteristics, which exhibit linear behavior at low voltages, indicating good Ohmic contact between the electrodes and the MoS2 and current saturation at high voltages. Owing to the low carrier concentration of MoS2, there remains significant potential for enhancement in the metal-semiconductor contacts within the conventional top-contact architecture. The integration approach of bottom-contact electrodes allows the top gate to exert precise control over the semiconductor material that interfaces with the bottom electrodes. This capability serves as an effective means to optimize and improve the metal-semiconductor contacts, thereby addressing the existing limitations in the traditional architecture [34]. At the same time, the coincident area between the top gate electrode and the source leakage electrode is much smaller than the coincident area of the capacitor, and we measured that the coincident area between the electrodes in the MOSFET is less than 50 μm2 (Figure 4c), and there is no breakdown of the 10 V voltage in the MOSFET leakage test (Figure S5).

3. Materials and Methods

The fabrication of MoS2 MOSFETs commenced with bottom electrode patterning via proximity photolithography, followed by gold metallization through thermal evaporation to define source/drain contacts with a 5.5 μm channel width. Few-layer MoS2 flakes were then mechanically exfoliated using PDMS stamps and precisely aligned onto the prefabricated Au electrodes. The Al2O3 films were deposited using alternating exposures of [Al(CH3)3] and oxygen plasma, with a growth rate of 0.14 nm per cycle. Each deposition cycle comprised the following steps: TMA exposure (0.06 s), Ar purging (12 s), oxygen plasma exposure (12 s), and final Ar purging (12 s), totaling 40 s per cycle. The process utilized Ar as the carrier gas under a chamber pressure of 0.01 Torr. Finally, top-gate Au electrodes were transfer-printed onto the Al2O3 surface using PDMS-mediated van der Waals assembly, completing the MOSFET architecture.

4. Conclusions

In summary, high-quality Al2O3 films were grown successfully using PEALD by controlling the growth temperature and oxygen plasma power. By growing high-quality Al2O3 films on silicon substrates, wide bandgap, high permittivity, low leakage current, and high breakdown strength were achieved, meeting the requirements for advanced low-power devices. Based on this, high-performance MoS2 field-effect transistors were fabricated with low working voltage, extremely low subthreshold swing, and high on/off ratio, demonstrating their potential in low-power electronic applications.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/jlpea15020026/s1. Figure S1: (a–d) Ellipsometric Fitting Analysis of Al2O3 Thin Films. (e) PEALD growth cycles vs. thickness correlation. Figure S2: A repeatability experiment on the breakdown voltage of 12.5 nm thick Al2O3 thin films under different temperatures at 80 W oxygen plasma. (a) 100 °C; (b) 150 °C; (c) 200 °C; (d) 250 °C; (e) 300 °C. Figure S3: A repeatability experiment on the breakdown voltage of 12.5 nm thick Al2O3 thin films under different oxygen plasma power levels at 200 °C. (a) 20 W; (b) 40 W; (c) 60 W (d) 100 W; (e) 120 W. Figure S4: A repeatability experiment on the breakdown voltage of Al2O3 thin films of different thicknesses under the conditions of 200 °C and 80 W oxygen plasma power. (a) 4 nm; (b) 6 nm; (c) 8 nm (d) 10 nm; (e) 12.5 nm. Figure S5: (a) Current density of MoS2 FETs growing at a 4 nm Al2O3 top gate dielectric layer. (b) Stress-time transfer curve of a MoS2 FETs with a bias voltage of 1 V.

Author Contributions

Conceptualization, resources, and methodology, H.H.; preparation, S.S.; electrical test, D.M.; characterization, B.Y. and G.L.; analysis, N.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guangxi Science and Technology Major Project, grant number AA23073018.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Acknowledgments

Acknowledges support from the Analysis and Testing Center of Guangxi University for the data testing support provided.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
FETsField-effect transistors
EOTEquivalent oxide thickness
CVDChemical vapor deposition
AFMAtomic force microscopy
TMATrimethylaluminum [Al(CH3)3]
SEMScanning electron microscope
EDSEnergy dispersive spectrometer
SSSubthreshold swing
VTThreshold voltage
μMobility
RaArithmetic mean roughness
ICsIntegrated circuits
PDMSPolydimethylsiloxane

References

  1. Das, S.; Sebastian, A.; Pop, E.; McClellan, C.J.; Franklin, A.D.; Grasser, T.; Knobloch, T.; Illarionov, Y.; Penumatcha, A.V.; Appenzeller, J.; et al. Transistors based on two-dimensional materials for future integrated circuits. Nat. Electron. 2021, 4, 786–799. [Google Scholar] [CrossRef]
  2. Wang, S.; Liu, X.; Xu, M.; Liu, L.; Yang, D.; Zhou, P. Two-dimensional devices and integration towards the silicon lines. Nat. Mater. 2022, 21, 1225–1239. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, S.; Liu, K.; Xu, Y.; Liu, L.; Li, H.; Zhai, T. Gate Dielectrics Integration for 2D Electronics: Challenges, Advances, and Outlook. Adv. Mater. 2023, 35, e2207901. [Google Scholar] [CrossRef]
  4. Lin, Y.C.; Lin, C.M.; Chen, H.Y.; Vaziri, S.; Bao, X.; Woon, W.Y.; Wang, H.; Liao, S.S. Dielectric Material Technologies for 2-D Semiconductor Transistor Scaling. IEEE Trans. Electron Devices 2023, 70, 1454–1473. [Google Scholar] [CrossRef]
  5. 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]
  6. Kilby, J.S. Invention of the integrated circuit. IEEE Trans. Electron Devices 1976, 23, 648–654. [Google Scholar] [CrossRef]
  7. Kingon, A.I.; Maria, J.-P.; Streiffer, S.K. Alternative dielectrics to silicon dioxide for memory and logic devices. Nature 2000, 406, 1032–1038. [Google Scholar] [CrossRef]
  8. Muller, D.A.; Sorsch, T.; Moccio, S.; Baumann, F.H.; Evans-Lutterodt, K.; Timp, G. The electronic structure at the atomic scale of ultrathin gate oxides. Nature 1999, 399, 758–761. [Google Scholar] [CrossRef]
  9. Luisier, M.; Klimeck, G. Simulation of nanowire tunneling transistors: From the Wentzel–Kramers–Brillouin approximation to full-band phonon-assisted tunneling. J. Appl. Phys. 2010, 107, 084507. [Google Scholar] [CrossRef]
  10. Wong, H.; Iwai, H. On the scaling of subnanometer EOT gate dielectrics for ultimate nano CMOS technology. Microelectron. Eng. 2015, 138, 57–76. [Google Scholar] [CrossRef]
  11. Wang, B.; Huang, W.; Chi, L.; Al-Hashimi, M.; Marks, T.J.; Facchetti, A. High-k Gate Dielectrics for Emerging Flexible and Stretchable Electronics. Chem. Rev. 2018, 118, 5690–5754. [Google Scholar] [CrossRef] [PubMed]
  12. 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-κ gate dielectrics for two-dimensional electronic devices. Nat. Electron. 2019, 2, 563–571. [Google Scholar] [CrossRef]
  13. Xu, W.; Jiang, J.; Chen, Y.; Tang, N.; Jiang, C.; Yang, S. Single-crystalline High-κ GdOCl dielectric for two-dimensional field-effect transistors. Nat. Commun. 2024, 15, 9469. [Google Scholar] [CrossRef]
  14. Yin, L.; Cheng, R.; Wan, X.; Ding, J.; Jia, J.; Wen, Y.; Liu, X.; Guo, Y.; He, J. High-κ monocrystalline dielectrics for low-power two-dimensional electronics. Nat. Mater. 2024, 24, 197–204. [Google Scholar] [CrossRef]
  15. Chen, J.; Liu, Z.; Dong, X.; Gao, Z.; Lin, Y.; He, Y.; Duan, Y.; Cheng, T.; Zhou, Z.; Fu, H.; et al. Vertically grown ultrathin Bi2SiO5 as high-κ single-crystalline gate dielectric. Nat. Commun. 2023, 14, 4406. [Google Scholar] [CrossRef]
  16. Yi, K.; Qin, W.; Huang, Y.; Wu, Y.; Feng, S.; Fang, Q.; Cao, X.; Deng, Y.; Zhu, C.; Zou, X.; et al. Integration of high-κ native oxides of gallium for two-dimensional transistors. Nat. Electron. 2024, 7, 1126–1136. [Google Scholar] [CrossRef]
  17. Li, S.; Liu, X.; Yang, H.; Zhu, H.; Fang, X. Two-dimensional perovskite oxide as a photoactive high-κ gate dielectric. Nat. Electron. 2024, 7, 216–224. [Google Scholar] [CrossRef]
  18. Zhang, X.-H.; Domercq, B.; Wang, X.; Yoo, S.; Kondo, T.; Wang, Z.L.; Kippelen, B. High-performance pentacene field-effect transistors using Al2O3 gate dielectrics prepared by atomic layer deposition (ALD). Org. Electron. 2007, 8, 718–726. [Google Scholar] [CrossRef]
  19. Ma, P.; Du, L.; Wang, Y.; Jiang, R.; Xin, Q.; Li, Y.; Song, A. Low voltage operation of IGZO thin film transistors enabled by ultrathin Al2O3 gate dielectric. Appl. Phys. Lett. 2018, 112, 023501. [Google Scholar] [CrossRef]
  20. Bolshakov, P.; Zhao, P.; Azcatl, A.; Hurley, P.K.; Wallace, R.M.; Young, C.D. Electrical characterization of top-gated molybdenum disulfide field-effect-transistors with high-k dielectrics. Microelectron. Eng. 2017, 178, 190–193. [Google Scholar] [CrossRef]
  21. Jeong, S.H.; Liu, N.; Park, H.; Hong, Y.K.; Kim, S. Temperature-Dependent Electrical Properties of Al2O3-Passivated Multilayer MoS2 Thin-Film Transistors. Appl. Sci. 2018, 8, 424. [Google Scholar] [CrossRef]
  22. Tanigawa, H.; Matsuura, K.; Muneta, I.; Hoshii, T.; Kakushima, K.; Tsutsui, K.; Wakabayashi, H. Enhancement-mode accumulation capacitance–voltage characteristics in TiN/ALD-Al2O3/sputtered-MoS2 top-gated stacks. Jpn. J. Appl. Phys. 2020, 59, SMMC01. [Google Scholar] [CrossRef]
  23. Hong, W.; Sik, O.D.; Choi, S.-Y. Passivation layer effect on the positive bias temperature instability of molybdenum disulfide thin film transistors. J. Inf. Disp. 2021, 22, 13–19. [Google Scholar] [CrossRef]
  24. Chang, Y.-H.; Yu, M.-J.; Lin, R.-P.; Hsu, C.-P.; Hou, T.-H. Abnormal positive bias stress instability of In–Ga–Zn–O thin-film transistors with low-temperature Al2O3 gate dielectric. Appl. Phys. Lett. 2016, 108, 033502. [Google Scholar] [CrossRef]
  25. Iatsunskyi, I.; Kempiński, M.; Jancelewicz, M.; Załęski, K.; Jurga, S.; Smyntyna, V. Structural and XPS characterization of ALD Al2O3 coated porous silicon. Vacuum 2015, 113, 52–58. [Google Scholar] [CrossRef]
  26. Movva, H.C.P.; Rai, A.; Kang, S.; Kim, K.; Fallahazad, B.; Taniguchi, T.; Watanabe, K.; Tutuc, E.; Banerjee, S.K. High-Mobility Holes in Dual-Gated WSe2 Field-Effect Transistors. ACS Nano 2015, 9, 10402–10410. [Google Scholar] [CrossRef]
  27. Wang, J.I.J.; Yang, Y.; Chen, Y.-A.; Watanabe, K.; Taniguchi, T.; Churchill, H.O.H.; Jarillo-Herrero, P. Electronic Transport of Encapsulated Graphene and WSe2 Devices Fabricated by Pick-up of Prepatterned hBN. Nano Lett. 2015, 15, 1898–1903. [Google Scholar] [CrossRef]
  28. Koval, V.; Viola, G.; Zhang, M.; Faberova, M.; Bures, R.; Yan, H. Dielectric relaxation and conductivity phenomena in ferroelectric ceramics at high temperatures. J. Eur. Ceram. Soc. 2024, 44, 2886–2902. [Google Scholar] [CrossRef]
  29. Groner, M.D.; Fabreguette, F.H.; Elam, J.W.; George, S.M. Low-Temperature Al2O3 Atomic Layer Deposition. Chem. Mater. 2004, 16, 639–645. [Google Scholar] [CrossRef]
  30. Chen, R.; Zhou, W.; Zhang, M.; Kwok, H.S. High performance self-aligned top-gate ZnO thin film transistors using sputtered Al2O3 gate dielectric. Thin Solid Films 2012, 520, 6681–6683. [Google Scholar] [CrossRef]
  31. Liu, A.; Liu, G.-X.; Shan, F.-K.; Zhu, H.-H.; Shin, B.C.; Lee, W.J.; Cho, C.R. High-Performance InTiZnO Thin-Film Transistors Deposited by Magnetron Sputtering. Chin. Phys. Lett. 2013, 30, 127301. [Google Scholar] [CrossRef]
  32. Kim, H.W.; Oh, C.; Jang, H.; Kim, M.Y.; Kim, B.S. Influence of oxygen-related defects on In-Ga-Sn-O semiconductor due to plasma-enhanced atomic layer deposition of Al2O3 for low-temperature thin-film transistor in terms of electrical properties. J. Alloys Compd. 2022, 918, 165649. [Google Scholar] [CrossRef]
  33. Li, P.; Yang, J.; Ding, X.; Li, X.; Zhang, J. High-Performance of InGaZnO TFTs With an Ultrathin 5-nm Al2O3 Gate Dielectric Enabled by a Novel Atomic Layer Deposition Method. IEEE J. Electron Devices Soc. 2024, 12, 121–126. [Google Scholar] [CrossRef]
  34. Han, L.; Huang, Y.; Tang, W.; Chen, S.; Zhao, J.; Guo, X. Reducing contact resistance in bottom contact organic field effect transistors for integrated electronics. J. Phys. D Appl. Phys. 2019, 53, 014002. [Google Scholar] [CrossRef]
Figure 1. (a) AFM surface topography of Al2O3 films under varied deposition parameters (temperature/power), with insets specifying growth conditions and arithmetic mean roughness (Ra). (b) XPS O1s spectra of Al2O3 films under distinct growth conditions.
Figure 1. (a) AFM surface topography of Al2O3 films under varied deposition parameters (temperature/power), with insets specifying growth conditions and arithmetic mean roughness (Ra). (b) XPS O1s spectra of Al2O3 films under distinct growth conditions.
Jlpea 15 00026 g001
Figure 2. (a) Breakdown field characteristics of 12.5 nm Al2O3 films grown at varying temperatures with fixed 80 W oxygen plasma power. (b) Breakdown field dependence on oxygen plasma power (20–120 W range) for 12.5 nm Al2O3 films deposited at 200 °C. (c) Thickness-dependent breakdown field evolution (4–12.5 nm) for films grown under optimized 200 °C/80 W conditions. (df) Statistical reproducibility of breakdown fields measured at ten discrete locations across Al2O3 films under each condition, presented as box-whisker plots with median values.
Figure 2. (a) Breakdown field characteristics of 12.5 nm Al2O3 films grown at varying temperatures with fixed 80 W oxygen plasma power. (b) Breakdown field dependence on oxygen plasma power (20–120 W range) for 12.5 nm Al2O3 films deposited at 200 °C. (c) Thickness-dependent breakdown field evolution (4–12.5 nm) for films grown under optimized 200 °C/80 W conditions. (df) Statistical reproducibility of breakdown fields measured at ten discrete locations across Al2O3 films under each condition, presented as box-whisker plots with median values.
Jlpea 15 00026 g002
Figure 3. (a) Dielectric constant and EOT of Al2O3 films with varying thicknesses grown under optimized conditions (200 °C, 80 W oxygen plasma). (b) Frequency-dependent capacitance density of Al2O3 capacitors under varied growth parameters, with inset illustrating the parallel-plate capacitor configuration. (c) Voltage-dependent capacitance density of Al2O3 films, lower panel: 12.5 nm films across different conditions; upper panel: thickness-scaled films deposited under optimized protocols.
Figure 3. (a) Dielectric constant and EOT of Al2O3 films with varying thicknesses grown under optimized conditions (200 °C, 80 W oxygen plasma). (b) Frequency-dependent capacitance density of Al2O3 capacitors under varied growth parameters, with inset illustrating the parallel-plate capacitor configuration. (c) Voltage-dependent capacitance density of Al2O3 films, lower panel: 12.5 nm films across different conditions; upper panel: thickness-scaled films deposited under optimized protocols.
Jlpea 15 00026 g003
Figure 4. (a) Fabrication process flow of back-gate top-contact MoS2. (b) AFM topography of the MoS2 channel material, revealing a flake thickness of ~8 nm. (c) SEM image of the FETs structure: red dashed lines outline the MoS2 channel (thickness: 8 nm), yellow dashed lines demarcate the top-gate Au electrodes, with a channel geometry of L/W = 5.5/6 μm. (d) EDS elemental mapping confirming spatial distribution of Mo, S, Au, and Al (from Al2O3 dielectric).
Figure 4. (a) Fabrication process flow of back-gate top-contact MoS2. (b) AFM topography of the MoS2 channel material, revealing a flake thickness of ~8 nm. (c) SEM image of the FETs structure: red dashed lines outline the MoS2 channel (thickness: 8 nm), yellow dashed lines demarcate the top-gate Au electrodes, with a channel geometry of L/W = 5.5/6 μm. (d) EDS elemental mapping confirming spatial distribution of Mo, S, Au, and Al (from Al2O3 dielectric).
Jlpea 15 00026 g004
Figure 5. (a) Transfer characteristics (ID-VG curve) of the device, yielding a subthreshold swing (SS) of 65 mV dec−1, a threshold voltage (VT) of −0.96 V, an Ion/Ioff of 106, and a μ of 34.85 cm2·V−1·s−1. Among them, the blue dotted line represents the steepness of the subthreshold region. (b) Output characteristics (ID-VD curve) of the same device, demonstrating gate-modulated current saturation behavior.
Figure 5. (a) Transfer characteristics (ID-VG curve) of the device, yielding a subthreshold swing (SS) of 65 mV dec−1, a threshold voltage (VT) of −0.96 V, an Ion/Ioff of 106, and a μ of 34.85 cm2·V−1·s−1. Among them, the blue dotted line represents the steepness of the subthreshold region. (b) Output characteristics (ID-VD curve) of the same device, demonstrating gate-modulated current saturation behavior.
Jlpea 15 00026 g005
Table 1. Summary of dielectric properties for Al2O3 dielectric.
Table 1. Summary of dielectric properties for Al2O3 dielectric.
MaterialMethodD (nm)Temp. (°C)Ci (nF/cm)Jleak (A/cm2)Ref.
Al2O3ALD433320110−7[29]
Al2O3PVD541RT117-[30]
Al2O3PLD80RT59.810−8[31]
Al2O3PEALD1401505710−9[32]
Al2O3ALD15400420-[20]
Al2O3ALD525066010−8[33]
Al2O3PEALD4200173010−7This work
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

Sun, S.; Ma, D.; Ye, B.; Liu, G.; Luo, N.; Huang, H. Ultra-Thin Al2O3 Grown by PEALD for Low-Power Molybdenum Disulfide Field-Effect Transistors. J. Low Power Electron. Appl. 2025, 15, 26. https://doi.org/10.3390/jlpea15020026

AMA Style

Sun S, Ma D, Ye B, Liu G, Luo N, Huang H. Ultra-Thin Al2O3 Grown by PEALD for Low-Power Molybdenum Disulfide Field-Effect Transistors. Journal of Low Power Electronics and Applications. 2025; 15(2):26. https://doi.org/10.3390/jlpea15020026

Chicago/Turabian Style

Sun, Shiwei, Dinghao Ma, Boxi Ye, Guanshun Liu, Nanting Luo, and Hao Huang. 2025. "Ultra-Thin Al2O3 Grown by PEALD for Low-Power Molybdenum Disulfide Field-Effect Transistors" Journal of Low Power Electronics and Applications 15, no. 2: 26. https://doi.org/10.3390/jlpea15020026

APA Style

Sun, S., Ma, D., Ye, B., Liu, G., Luo, N., & Huang, H. (2025). Ultra-Thin Al2O3 Grown by PEALD for Low-Power Molybdenum Disulfide Field-Effect Transistors. Journal of Low Power Electronics and Applications, 15(2), 26. https://doi.org/10.3390/jlpea15020026

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