Next Article in Journal
Evaluation on Improvement Effect of Different Anti-Stripping Agents on Pavement Performance of Granite–Asphalt Mixture
Previous Article in Journal
Effect of Internal Donors on Raman and IR Spectroscopic Fingerprints of MgCl2/TiCl4 Nanoclusters Determined by Machine Learning and DFT
Previous Article in Special Issue
Monolayer Twisted Graphene-Based Schottky Transistor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Gate Work Function and Electrode Gap on Wide Band-Gap Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors

1
Semiconductor Devices and Circuits Laboratory, Advanced Institute of Convergence Technology, Seoul National University, Suwon 16229, Korea
2
Research Center for Materials, Components and Equipment, Advanced Institute of Convergence Technology, Seoul National University, Suwon 16229, Korea
*
Author to whom correspondence should be addressed.
Materials 2022, 15(3), 913; https://doi.org/10.3390/ma15030913
Submission received: 27 December 2021 / Revised: 20 January 2022 / Accepted: 22 January 2022 / Published: 25 January 2022

Abstract

:
We present technology computer aided design (TCAD) results for wide band-gap Sn-doped α-Ga2O3 metal–semiconductor field-effect transistors (MESFETs). In particular, the effect of gate work function and electrode gap length on the electrical characteristics is demonstrated for a thorough understanding of the behavior of such devices. The gate work function significantly affects the reverse bias drain current under the gate-current dominant regime, whereas a gate-source/drain gap larger than 0.1 µm has a negligible effect on the drain current.

1. Introduction

During the past decade, ultrawide bandgap Ga2O3 semiconductors with a bandgap of 4.5~5.3 eV (depending on the crystal structure) have been investigated as an alternative to SiC and GaN (3.3 and 3.4 eV, respectively) for high-power electronic device applications [1,2]. Among five different phases of Ga2O3 (α, β, γ, ε, and δ) [3], the orthorhombic β phase is the most stable thermodynamically, while the rhombohedral corundum α phase is semi-stable [4]. On the other hand, the band gap of α-Ga2O3 is 5.3 eV [5,6], which is wider compared to β-Ga2O3 4.9 eV [1,7,8], promising a higher breakdown field.
Conventionally, β-Ga2O3 has been grown via molecular beam epitaxy [1,2] on a β-Ga2O3 substrate grown from the melt [9]. However, it is difficult to produce a β-Ga2O3 wafer with a diameter large enough for practical application due to easy formation of cleavages such that the wafer size is limited to four inches [9]. Recently, mist chemical vapor deposition (Mist-CVD) has been introduced as a non-vacuum solution-process heteroepitaxy for α-Ga2O3 on mass-produced sapphire (Al2O3) wafers up to six inches, with a similar a crystal structure to α-Ga2O3 [10,11,12,13,14]. Being able to lift α-Ga2O3 off of the sapphire substrate and bond it to other substrates with high thermal conductivity (such as SiC, AlN, diamond, etc.) provides an additional advantage in high power switching and RF applications over β-Ga2O3 with a low thermal conductivity [9]. Despite these promising results on the epitaxial growth of α-Ga2O3 on sapphire, there are few demonstrations of electronic devices based on α-Ga2O3 [15,16,17].
A high-quality Silver oxide AgOx Schottky contact was incorporated into Sn-doped α-Ga2O3 metal-semiconductor field-effect transistors (MESFET) [16] in order to achieve high rectifying Schottky contact at the gate–semiconductor interface. The use of the non-metallic gate electrode for the α-Ga2O3 MESFET enables the formation of the gate electrode and the metallic source/drain contact at the same plane (i.e., a coplanar configuration). In 2019, an in-depth experimental study on the oxidized metal Schottky contacts (of which the work function ranges from 4.70 to 5.80 eV) including AgOx, on β-Ga2O3 was reported [18]. However, a similar work on the oxidized metal Schottky contacts on α-Ga2O3 has not been reported yet. Therefore, the design strategy for optimal operation is lacking.
In this study, we begin by conducting a study on the effect of the gate work function on the electrical characteristics of wide band-gap Sn-doped α-Ga2O3 MESFET for a broad range of the work function, from 4.40 to 5.80 eV. The optimal gate work function found thusly will be applied while varying the source/drain-gate gap length between 0.1 to 2.0 µm. Electrical characteristics issued from these parameters will then be discussed, deepening our knowledge of the optimal configuration of such a device.

2. Materials and Methods

2.1. Metal–Semiconductor Field-Effect Transistor (MESFET)

A metal–semiconductor field-effect transistor (MESFET) consists of a substrate, a semiconductor layer, the gate electrode (G), and the source (S) and drain (D) electrodes (Figure 1). For a coplanar structure, the channel length L is defined as the distance between the S and D electrode; hence, L = LG + 2 × Lgap, where LG is the gate length and Lgap is the gap between the S/D electrode and the G electrode. The channel width is denoted by W. The thickness of the semiconductor layer is denoted by ds.
The energy structure of an n-type MESFET is determined by the conduction and the valence band edge level, EC and EV; the total density of states for the conduction and valence band of the semiconductor, NC and NV; the donor level, ED; the total density of states for donor ND of the n-type dopant; and the work function of S/D and G, WS/D and WG, respectively. The dielectric constant εs, electron and hole mobility, respectively μe and μh, and electron effective mass, me, describe the electrical properties of the semiconductor.
A complete list of parameters used for the simulation is provided in Table 1. The values correspond to the Sn-doped α-Ga2O3 MESFET with Ti as the S/D electrodes and AgOx as the gate electrode. α-Ga2O3 is amenable to n-type doping by Sn as well [10,19,20], which enhances the free electron concentration and hence the mobility, and facilitates charge carrier injection at the source/drain. The values for ED and ND were taken from [16]. The typical value and the range of WG were determined considering the reported values in [18,21]. The value for WS/D was taken from [22]. Note that edge dislocation could present in the α-Ga2O3 epitaxy layer, around 107 (epitaxial lateral overgrowth) ~ 1010 cm−2 (Mist-CVD), depending on the deposition methods [9], which lowers electron mobility, i.e., 1.3 cm2V−1s−1 for high edge dislocation density [5,15,23] and 24 cm2V−1s−1 for low edge dislocation density [24] compared to the theoretical value 300 cm2V−1s−1. In this study, the effect of dislocation is considered by carrier mobility.
Although there is a lack of study on the effect of defects at the interface between α-Ga2O3 and metal/oxidized metal, for oxide semiconductors the most likely defects are oxygen vacancies, VO, formed by chemical reactions during metal deposition [25,26]. The number of VO is smaller at the semiconductor–oxidized metal interface compared to the semiconductor–metal interface because of the oxygen-rich deposition conditions for oxidized metal layer [18], which is likely to prevent Fermi level pinning by VO. Therefore, in this study, the effect of Fermi level pinning is not considered at the semiconductor–gate interface.

2.2. Numerical Simulation

The numerical simulation of MESFET resolves the coupled drift–diffusion current equation and the Poisson’s equation to obtain the current–voltage characteristics and the current density, charge carrier, and potential distribution. We adopted TCAD software Atlas from Silvaco, Santa Clara, CA, USA. [27]. It is an advantage of numerical simulation that the work function can be varied without altering other physical parameters, which is difficult to achieve experimentally. We considered the Schottky barrier lowering and tunneling models computed by Wentzel-Kramers-Brillouin approximation [28] at both the source/drain-semiconductor and the gate–semiconductor junction. This allows description of the charge carrier injection at the Schottky junction with a large injection barrier.
In order to investigate the effect of the gate work function, WG, we varied the latter from 4.4 to 5.8 eV by 0.2 eV while fixing the source/drain-gate gap to 1.0 µm and the gate length to 8.0 µm. Then, in order to investigate the effect of the source/drain-gate gap, Lgap, Lgap was varied as 0.1, 0.2, 0.5, 1.0, and 2.0 µm. Concomitantly, the gate length changed accordingly, as 9.8, 9.6, 9.0, 8.0, and 6.0 µm, as the channel length L was fixed to 10 µm. The gate work function WG for the second simulation set was fixed at 5.4 eV.

3. Results and Discussion

3.1. Effect of Gate Work Function Variation on Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors

3.1.1. Current–Voltage (I-V) Characteristics

Figure 2 shows the simulated current–voltage (I-V) characteristics of MESFETs with various gate work functions. The gate current IG and drain current ID are plotted as a function of the gate-source voltage VGS. When VGS is larger than the on voltage, Von, and smaller than a certain voltage (~7 V, which is similar to the drain-source, VDS), Von < VGSVDS and ID dominates over IG. When VGSVDS, IG dominates over ID because the drain-gate diode is now forward biased. When VGS is smaller than the on voltage Von, IG is a dominant factor.
The on-off ratio, defined as the ratio of ID at VGS = 7 V to that at VGS = −7 V, increases as the gate Fermi level is lowered. When WG ≤ 5.0 eV, the drain current ID under the IG-dominant regime becomes comparable and even larger than that under the normal-operation regime. Therefore, the device could not be used as a switching element. When WG > 5.0 eV, the on-off ratio is around 101 ~ 107, showing good rectification behavior. In summary, the degree of electron injection into the gate electrode on the drain side, as will be shown in the following sections, determines the level of off-current, and hence the on-off ratio of the transistor.
In addition, Von should be as close as possible to 0 V to guarantee functional transistor behavior. Thus, a gate work function of WG = 5.4 eV is the optimal condition. This condition was used to analyze the effect of the source/drain-gate gap Lgap.

3.1.2. Current Density Distribution and Vector

Figure 3a,b provides direct evidence that the current flows into the gate electrode under the IG-dominant regime and the off regime. In particular, the current density is high at the edge of gate on the drain side (black boxes). On the other hand, the current flows out from the gate electrode (Figure 3c). The current coming from the drain joins that coming from the gate, and flows into the source, which establishes the current path of the device under the normal-operation regime.

3.1.3. Carrier Concentration and Potential Distribution

The semiconductor under the gate electrode is fully or partially depleted, whereas the semiconductor under the source/drain-gate gap is accumulated and the charge carrier concentration is high (n ~ 1017 cm−3 ) (Figure 4a–c). In addition, the potential difference is −7 V between G and S and −14 V between G and D (under the IG-dominant regime, Figure 4d), and 0 V between G and S and −7 V between G and D (under the off regime, Figure 4e). Thus, the current flows into the gate under the IG-dominant regime and the off regime.

3.2. Effect of Source/Drain-Gate Gap Variation on Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors

3.2.1. Current–Voltage (I-V) Characteristics

In general, the current–voltage characteristics for all cases of Lgap between 0.1 to 2.0 µm (shown in Figure 5) feature the typical I-V characteristics of MESFET, with a greater ID compared to IG when VGS is higher than Von and a greater IG compared to ID when VGS is lower than Von.
In detail, ID decreases as Lgap increases in the IG-dominant regime, whereas ID increases as Lgap increases in the normal-operation regime. However, IG decreases as Lgap increases in both the IG-dominant regime and the normal-operation regime. It is noticeable that the I-V characteristics for Lgap = 0.1 µm are significantly different, with a longer Lgap = 0.2, 0.5, 1.0, and 2.0 µm. Such differences are explained in the following sections by considering the current path of the device with the current density, charge concentration and potential distribution. It can be inferred that under the normal operation regime electron transport under the gate–source gap does not deteriorate the current unless the carrier concentration under the gap is maintained at a high enough level.

3.2.2. Current Density Distribution and Vector

Figure 6 shows the current density distribution and its vector in the IG-dominant regime, off regime, and normal-operation regime for Lgap = 0.1 µm (Figure 6a–c) and Lgap = 2.0 µm (Figure 6d–f). As discussed in Section 3.2.1, the current flows into the gate electrode from the drain electrode under the IG-dominant regime and the off regime. For both values of Lgap, the current density is high at the edge of the gate on the drain side (highlighted by the black rectangle) under the IG-dominant regime and the off regime. In the normal-operation regime, the drain current joins the gate current to flow into the source.
The differences in current–voltage characteristics between Lgap = 0.1 µm and the other cases can be elucidated by the current path. In the IG-dominant regime, the current flows from the source and drain electrodes toward the gate electrodes. Therefore, a smaller gap length between the gate and source/drain electrodes increases both ID and IG by providing a shorter resistive path to the gate electrode. In the case of the off regime, the current flows from the drain electrode to the source electrode while being leaked in the gate channel area. A smaller Lgap decreases both ID and IG due to a longer gate current path between the drain and source electrodes. Noticeably, a greater difference in the drain current ID compared to IG undermines leakage of the drain current while crossing the gate channel area. Lastly, in the normal-operation regime the current flows from the drain and gate electrodes toward the source electrode. Therefore, there is no crowding of current at the frontier of the gate electrode and drain–gate electrode gap. Due to this phenomenon, the current density ID remains almost constant when Lgap > 0.1 µm. Meanwhile, the gate current ID shows a drastic difference in cases where Lgap is 0.1 µm. This could originate from the fact that the current from both the gate and the drain accumulates itself at the edge of the source electrode. A more plausible explanation can be made by referring to the charge carrier concentration and potential distribution, as detailed in the following section.

3.2.3. Carrier Concentration and Potential Distribution

The simulation results of the carrier concentration distribution n(x, y), shown in Figure 7, reveal that the semiconductor under the gate electrode is either fully depleted in the IG-dominant regime or partially depleted in the off regime and normal-operation regime for both cases of Lgap. On the other hand, a high carrier concentration up to ~1017 cm−3 is observed beneath the electrodes gap, where the effect of the gate field is out of reach. This phenomenon is more pronounced in the case of a larger Lgap. For Lgap = 0.1 µm when the gap becomes comparable to a few Debye length, the effect of the gate field is present in the gap, as reported in [29]. In this case, the carrier concentration in the gap becomes approximately 106 cm−3 lower than 1017 cm−3 by several orders of magnitude.
Figure 8 shows the potential distribution in the device for an electrode gap of Lgap = 0.1 and 2.0 µm. Similar to the observation in Section 3.1.3, the greater potential difference between gate and drain under the IG-dominant regime (Figure 8a,d) and off regime (Figure 8b,e) justifies the high current density concentration at the edge of the gate electrode from the drain electrode. In the normal-operation regime (Figure 8c,f), a greater potential difference is found at the edge of the source electrode from the gate electrode. Thereby, the high current density flows in this area.

4. Conclusions

In this study, we have described the effects of the gate work function and electrode gap on the electrical characteristics of Sn-doped α-Ga2O3 MESFETs using TCAD software. The gate work function significantly changes the current level of the IG-dominant regime, hence the rectification ratio. The existence and the mechanism of the gate current under the IG-dominant regime were illustrated by simulated current density distribution and vector as well as by charge carrier and potential distribution, allowing for determination of a theoretical optimal gate work function value of a coplanar MESFET. As for the electrode gap, the simulation results of the current vector enabled us to understand the current path in Sn-doped α-Ga2O3 MESFETs. It is imperative to respect a certain amount of gap distance between electrodes of at least than 0.1 µm to prevent the effect of the gate field in the gap region. Considering that most research efforts have been focused on the deposition and characterization of an Sn-doped α-Ga2O3 heteroepitaxial layer, this study on device simulation will help to translate such knowledge concerning α-Ga2O3 heteroepitaxy into device design, fabrication and optimization for further improvement of device performance.

Author Contributions

Conceptualization, S.J.; methodology, S.J.; formal analysis, H.-S.R.; data curation, H.-S.R.; writing—original draft preparation, H.-S.R.; writing—review and editing, S.J., S.H.K.; supervision, S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Materials, Components and Equipment Research Program, funded by the Gyeonggi Province and by Basic Science Research Program through the National Research Foundation of the Ministry of Science and ICT of Korea under grant NRF-2021R1F1A1064384.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Higashiwaki, M.; Sasaki, K.; Kuramata, A.; Masui, T.; Yamakoshi, S. Gallium oxide (Ga2O3) metal-semiconductor field-effect transistors on single-crystal β-Ga2O3 (010) substrates. Appl. Phys. Lett. 2012, 100, 013504. [Google Scholar] [CrossRef]
  2. Higashiwaki, M.; Sasaki, K.; Kamimura, T.; Hoi Wong, M.; Krishnamurthy, D.; Kuramata, A.; Masui, T.; Yamakoshi, S. Depletion-mode Ga2O3 metal-oxide-semiconductor field-effect transistors on β-Ga2O3 (010) substrates and temperature dependence of their device characteristics. Appl. Phys. Lett. 2013, 103, 123511. [Google Scholar] [CrossRef] [Green Version]
  3. Pearton, S.J.; Yang, J.; Cary, P.H.; Ren, F.; Kim, J.; Tadjer, M.J.; Mastro, M.A. A review of Ga2O3 materials, processing, and devices. Appl. Phys. Rev. 2018, 5, 011301. [Google Scholar] [CrossRef] [Green Version]
  4. Fujita, S.; Oda, M.; Kaneko, K.; Hitora, T. Evolution of corundum-structured III-oxide semiconductors: Growth, properties, and devices. Jpn. J. Appl. Phys. 2016, 55, 1202A3. [Google Scholar] [CrossRef]
  5. Shinohara, D.; Fujita, S. Heteroepitaxy of corundum-structured α-Ga2O3 thin films on α-Al2O3 substrates by ultrasonic mist chemical vapor deposition. Jpn. J. Appl. Phys. 2008, 47, 7311–7313. [Google Scholar] [CrossRef]
  6. Mondal, A.K.; Mohamed, M.A.; Ping, L.K.; Taib, M.F.M.; Samat, M.H.; Haniff, M.A.S.M.; Bahru, R. First-principles studies for electronic structure and optical properties of p-type calcium doped α-Ga2O3. Materials 2021, 14, 604. [Google Scholar] [CrossRef]
  7. Kaneko, K.; Nomura, T.; Kakeya, I.; Fujita, S. Fabrication of highly crystalline corundum-structured α-(Ga 1-xFex)2O3alloy thin films on sapphire substrates. Appl. Phys. Express 2009, 2, 075501. [Google Scholar] [CrossRef]
  8. Usseinov, A.; Koishybayeva, Z.; Platonenko, A.; Pankratov, V.; Suchikova, Y.; Akilbekov, A.; Zdorovets, M.; Purans, J.; Popov, A.I. Vacancy Defects in Ga2O3: First-Principles Calculations of Electronic Structure. Materials 2021, 14, 7384. [Google Scholar] [CrossRef]
  9. Ahmadi, E.; Oshima, Y. Materials issues and devices of α- and β-Ga2O3. J. Appl. Phys. 2019, 126, 160901. [Google Scholar] [CrossRef] [Green Version]
  10. Kawaharamura, T.; Dang, G.T.; Furuta, M. Successful growth of conductive highly crystalline Sn-doped α-Ga2O3 thin films by fine-channel mist chemical vapor deposition. Jpn. J. Appl. Phys. 2012, 51, 040207. [Google Scholar] [CrossRef]
  11. Xu, Y.; Zhang, C.; Cheng, Y.; Li, Z.; Cheng, Y.; Feng, Q.; Chen, D.; Zhang, J.; Hao, Y. Influence of carrier gases on the quality of epitaxial corundum-structured α-Ga2O3 films grown by mist chemical vapor deposition method. Materials 2019, 12, 3670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Cheng, Y.; Xu, Y.; Li, Z.; Zhang, J.; Chen, D.; Feng, Q.; Xu, S.; Zhou, H.; Zhang, J.; Hao, Y.; et al. Heteroepitaxial growth of α-Ga2O3 thin films on a-, c- and r-plane sapphire substrates by low-cost mist-CVD method. J. Alloys Compd. 2020, 831, 154776. [Google Scholar] [CrossRef]
  13. Ha, M.-T.; Kim, K.-H.; Shin, Y.-J.; Jeong, S.-M.; Bae, S.-Y. Leidenfrost Motion of Water Microdroplets on Surface Substrate: Epitaxy of Gallium Oxide via Mist Chemical Vapor Deposition. Adv. Mater. Interfaces 2021, 8, 2001895. [Google Scholar] [CrossRef]
  14. Park, S.Y.; Ha, M.T.; Kim, K.H.; Van Lich, L.; Shin, Y.J.; Jeong, S.M.; Kwon, S.H.; Bae, S.Y. Enhanced thickness uniformity of large-scale α-Ga2O3 epilayers grown by vertical hot-wall mist chemical vapor deposition. Ceram. Int. 2021, 48, 5075–5082. [Google Scholar] [CrossRef]
  15. Oda, M.; Tokuda, R.; Kambara, H.; Tanikawa, T.; Sasaki, T.; Hitora, T. Schottky barrier diodes of corundum-structured gallium oxide showing on-resistance of 0.1mωcm2 grown by MIST EPITAXY®. Appl. Phys. Express 2016, 9, 021101. [Google Scholar] [CrossRef]
  16. Dang, G.T.; Kawaharamura, T.; Furuta, M.; Allen, M.W. Mist-CVD Grown Sn-Doped α-Ga2O3 MESFETs. IEEE Trans. Electron Devices 2015, 62, 3640–3644. [Google Scholar] [CrossRef]
  17. Hu, Z.; Nomoto, K.; Li, W.; Jinno, R.; Nakamura, T.; Jena, D.; Xing, H. 1.6 kV Vertical Ga2O3 FinFETs with Source-Connected Field Plates and Normally-off Operation. Proc. Int. Symp. Power Semicond. Devices ICs 2019, 2019, 483–486. [Google Scholar] [CrossRef]
  18. Hou, C.; Gazoni, R.M.; Reeves, R.J.; Allen, M.W. Oxidized Metal Schottky Contacts on (010) β-Ga 2O3. IEEE Electron Device Lett. 2019, 40, 337–340. [Google Scholar] [CrossRef]
  19. Nishinaka, H.; Yoshimoto, M. Mist Chemical Vapor Deposition of Single-Phase Metastable Rhombohedral Indium Tin Oxide Epitaxial Thin Films with High Electrical Conductivity and Transparency on Various α-Al2O3 Substrates. Cryst. Growth Des. 2018, 18, 4022–4028. [Google Scholar] [CrossRef]
  20. Akaiwa, K.; Fujita, S. Electrical conductive corundum-structured α-Ga2O3 Thin films on sapphire with tin-doping grown by spray-assisted mist chemical vapor deposition. Jpn. J. Appl. Phys. 2012, 51, 070203. [Google Scholar] [CrossRef]
  21. Santana, L.A.; Reséndiz, L.M.; Díaz, A.I.; Hernandez-Cuevas, F.J.; Aleman, M.; Hernandez-Como, N. Schottky barrier diodes fabricated with metal oxides AgOx/IGZO. Microelectron. Eng. 2020, 220, 111182. [Google Scholar] [CrossRef]
  22. Hölzl, J.; Schulte, F.K. Work function of metals. In Springer Tracts in Modern Physics; Springer: Berlin/Heidelberg, Germany, 1979; pp. 1–150. [Google Scholar]
  23. Kaneko, S.K.; Kawanowa, H.; Ito, H.; Fujita, S. Evaluation of misfit relaxation in α-Ga2O3 epitaxial growth on α-Al2O3. Jpn. J. Appl. Phys. 2012, 51, 020201. [Google Scholar] [CrossRef]
  24. Jinno, R.; Uchida, T.; Kaneko, K.; Fujita, S. Reduction in edge dislocation density in corundum-structured α-Ga2O3 layers on sapphire substrates with quasi-graded α-(Al,Ga)2O3 buffer layers. Appl. Phys. Express 2016, 9, 071101. [Google Scholar] [CrossRef]
  25. Robertson, J.; Sharia, O.; Demkov, A.A. Fermi level pinning by defects in HfO2-metal gate stacks. Appl. Phys. Lett. 2007, 91, 132912. [Google Scholar] [CrossRef]
  26. Allen, M.W.; Durbin, S.M. Influence of oxygen vacancies on Schottky contacts to ZnO. Appl. Phys. Lett. 2008, 92, 122110. [Google Scholar] [CrossRef]
  27. Atlas Device Simulator 2016. Available online: http://www.silvaco.com/ (accessed on 1 November 2021).
  28. Atlas User’s Manual, Silvaco, Inc.: Santa Clara, CA, USA, 2014.
  29. Sandborn, P.A.; East, J.R.; Haddad, G.I. Diffusion effects in short-channel gaas mesfets. Solid-State Electron. 1989, 32, 191–198. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) A schematic diagram for the device structure of an Sn-doped α-Ga2O3 MESFET; (b) the energy diagram of the corresponding device.
Figure 1. (a) A schematic diagram for the device structure of an Sn-doped α-Ga2O3 MESFET; (b) the energy diagram of the corresponding device.
Materials 15 00913 g001
Figure 2. Simulated gate current, IG, and drain current, ID for various gate work functions WG. WG was varied from 4.4 to 5.8 eV by 0.2 eV. An IG-dominant region is observed for the gate voltages smaller than the on voltage VGS < Von.
Figure 2. Simulated gate current, IG, and drain current, ID for various gate work functions WG. WG was varied from 4.4 to 5.8 eV by 0.2 eV. An IG-dominant region is observed for the gate voltages smaller than the on voltage VGS < Von.
Materials 15 00913 g002
Figure 3. Simulated total current density Jtot(x, y) for (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). The arrows represent the simulated total current density vector. A magnified view of the semiconductor region near the gate electrode (18 µm ≤ x ≤ 20 µm for (a,b) and 9 µm ≤ x ≤ 11 µm for (c)) is shown for (ac). The work function of the gate is WG = 5.4 eV. The red boxes indicate the region where the total current density is high.
Figure 3. Simulated total current density Jtot(x, y) for (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). The arrows represent the simulated total current density vector. A magnified view of the semiconductor region near the gate electrode (18 µm ≤ x ≤ 20 µm for (a,b) and 9 µm ≤ x ≤ 11 µm for (c)) is shown for (ac). The work function of the gate is WG = 5.4 eV. The red boxes indicate the region where the total current density is high.
Materials 15 00913 g003
Figure 4. (ac) Simulated carrier concentration distribution for n(x, y) (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated potential distribution V(x, y) for (d) VGS = −7 V and VDS = 7 V, (e) VGS = 0 V and VDS = 7 V, (f) VGS = 7 V and VDS = 7 V. The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (18 µm ≤ x ≤ 20 µm for (a,b,d,e) and 9 µm ≤ x ≤ 11 µm for (c,f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Figure 4. (ac) Simulated carrier concentration distribution for n(x, y) (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated potential distribution V(x, y) for (d) VGS = −7 V and VDS = 7 V, (e) VGS = 0 V and VDS = 7 V, (f) VGS = 7 V and VDS = 7 V. The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (18 µm ≤ x ≤ 20 µm for (a,b,d,e) and 9 µm ≤ x ≤ 11 µm for (c,f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Materials 15 00913 g004
Figure 5. Simulated gate current IG and drain current ID for various source/drain-gate gaps, Lgap. Lgap was divided into five separate cases of 0.1, 0.2, 0.5, 1.0 and 2.0 µm.
Figure 5. Simulated gate current IG and drain current ID for various source/drain-gate gaps, Lgap. Lgap was divided into five separate cases of 0.1, 0.2, 0.5, 1.0 and 2.0 µm.
Materials 15 00913 g005
Figure 6. (ac) Simulated total current density Jtot(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated total current density Jtot(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The arrows represent the simulated total current density vector. A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV. The black boxes indicate the region where the total current density is high.
Figure 6. (ac) Simulated total current density Jtot(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated total current density Jtot(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The arrows represent the simulated total current density vector. A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV. The black boxes indicate the region where the total current density is high.
Materials 15 00913 g006
Figure 7. (ac) Simulated carrier concentration distribution n(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated carrier concentration distribution n(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Figure 7. (ac) Simulated carrier concentration distribution n(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated carrier concentration distribution n(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Materials 15 00913 g007
Figure 8. (ac) Simulated potential distribution V(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated potential distribution V(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Figure 8. (ac) Simulated potential distribution V(x, y) for Lgap = 0.1 µm (a) VGS = −7 V and VDS = 7 V (IG-dominant regime), (b) VGS = 0 V and VDS = 7 V (off regime), (c) VGS = 7 V and VDS = 7 V (normal-operation regime). (df) Simulated potential distribution V(x, y) for Lgap = 2.0 µm (d) VGS = −7 V and VDS = 7 V (IG-dominant regime), (e) VGS = 0 V and VDS = 7 V (off regime), (f) VGS = 7 V and VDS = 7 V (normal-operation regime). The entire semiconductor layer is shown for all panels (af). A magnified view of the semiconductor region near the gate electrode (19.8 µm ≤ x ≤ 20.1 µm for (a,b), 9.9 µm ≤ x ≤ 10.2 µm for (c), 17 µm ≤ x ≤ 21 µm for (d,e), 9 µm ≤ x ≤ 13 µm for (f)) is shown for all panels. The work function of the gate is WG = 5.4 eV.
Materials 15 00913 g008
Table 1. Parameters used for TCAD Simulation.
Table 1. Parameters used for TCAD Simulation.
NameSymbolValueUnit
Channel lengthL10µm
Gate lengthLG6.0~9.8µm
Source/drain-gate gapLgap0.1~2.0µm
Source and drain lengthLS/D10µm
Channel widthW262µm
Semiconductor thicknessds210nm
Conduction band edge levelEC3.00eV
Valence band edge levelEV7.30eV
Total density of states for conduction bandNC4.97 × 1018cm−3
Total density of states for valence bandNV4.97 × 1018cm−3
Total density of states for donorND3.00 × 1017cm−3
Donor levelED1.10eV
Source/drain work functionWS/D4.33eV
Gate work functionWG4.40~5.80eV
Semiconductor relative dielectric constantεs10-
Electron mobilityμe1.3cm2V−1s−1
Hole mobilityμh1.3cm2V−1s−1
Electron effective massme0.34-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ro, H.-S.; Kang, S.H.; Jung, S. The Effect of Gate Work Function and Electrode Gap on Wide Band-Gap Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors. Materials 2022, 15, 913. https://doi.org/10.3390/ma15030913

AMA Style

Ro H-S, Kang SH, Jung S. The Effect of Gate Work Function and Electrode Gap on Wide Band-Gap Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors. Materials. 2022; 15(3):913. https://doi.org/10.3390/ma15030913

Chicago/Turabian Style

Ro, Han-Sol, Sung Ho Kang, and Sungyeop Jung. 2022. "The Effect of Gate Work Function and Electrode Gap on Wide Band-Gap Sn-Doped α-Ga2O3 Metal–Semiconductor Field-Effect Transistors" Materials 15, no. 3: 913. https://doi.org/10.3390/ma15030913

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