Next Article in Journal
Evaluation of Feasibility on Dental Zirconia—Accelerated Aging Test by Chemical Immersion Method
Next Article in Special Issue
Crystal Structure, Luminescence and Electrical Conductivity of Pure and Mg2+-Doped β-Ga2O3-In2O3 Solid Solutions Synthesized in Oxygen or Argon Atmospheres
Previous Article in Journal
Single-Ion Counting with an Ultra-Thin-Membrane Silicon Carbide Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in Gallium Oxide Field-Effect Transistors for High-Power and RF Applications

1
Department of Electrical Engineering, George Mason University, Fairfax, VA 22030, USA
2
Nanoscale Device and Characterization Division, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA
3
Quantum Science & Engineering Center, George Mason University, Fairfax, VA 22030, USA
*
Author to whom correspondence should be addressed.
Materials 2023, 16(24), 7693; https://doi.org/10.3390/ma16247693
Submission received: 30 October 2023 / Revised: 21 November 2023 / Accepted: 23 November 2023 / Published: 18 December 2023
(This article belongs to the Special Issue Ultra-Wide Bandgap Semiconductor Materials and Devices)

Abstract

:
Power electronics are becoming increasingly more important, as electrical energy constitutes 40% of the total primary energy usage in the USA and is expected to grow rapidly with the emergence of electric vehicles, renewable energy generation, and energy storage. New materials that are better suited for high-power applications are needed as the Si material limit is reached. Beta-phase gallium oxide (β-Ga2O3) is a promising ultra-wide-bandgap (UWBG) semiconductor for high-power and RF electronics due to its bandgap of 4.9 eV, large theoretical breakdown electric field of 8 MV cm−1, and Baliga figure of merit of 3300, 3–10 times larger than that of SiC and GaN. Moreover, β-Ga2O3 is the only WBG material that can be grown from melt, making large, high-quality, dopable substrates at low costs feasible. Significant efforts in the high-quality epitaxial growth of β-Ga2O3 and β-(AlxGa1−x)2O3 heterostructures has led to high-performance devices for high-power and RF applications. In this report, we provide a comprehensive summary of the progress in β-Ga2O3 field-effect transistors (FETs) including a variety of transistor designs, channel materials, ohmic contact formations and improvements, gate dielectrics, and fabrication processes. Additionally, novel structures proposed through simulations and not yet realized in β-Ga2O3 are presented. Main issues such as defect characterization methods and relevant material preparation, thermal studies and management, and the lack of p-type doping with investigated alternatives are also discussed. Finally, major strategies and outlooks for commercial use will be outlined.

1. Introduction

The power semiconductor market observed a 30% growth in 2022, and continual growth is expected as more electrical energy passes through power electronics, approximately 30% in 2019 and up to 80% in the next decade [1,2]. High-power semiconductor applications are classified into high-power (low-frequency) or high-frequency, RF. As silicon power devices reach their limit at breakdown voltages up to 6.5 kV and have a high temperature capability up to 200 °C [3], wide-bandgap (WBG) materials offer improved efficiency, large power ratings, high switching speeds, and RF performance. While SiC and GaN have been the dominant WBG semiconductors with commercially available devices, ultra-wide-bandgap (UWBG) beta-phase gallium oxide (β-Ga2O3) is emerging as a material for next-generation high-power and RF electronics.
With its bandgap of 4.7–4.9 eV, large theoretical breakdown field of 8 MV cm−1, and high electron saturation velocity of 2 × 107 cm s−1, β-Ga2O3 has a Baliga figure of merit (BFOM), indicating DC conduction losses, as well as a Johnson figure of merit (JFOM) for RF performance, higher than those of GaN and SiC [4,5,6,7]. Additionally, the ability to grow bulk substrates from the melt gives β-Ga2O3 a significant cost advantage over SiC and GaN [8]. However, difficult challenges face β-Ga2O3 in the form of a lack of shallow p-type dopants and low thermal conductivity, which is especially difficult for high-power applications where heat dissipation is essential. Heterostructures with p-type oxides have been fabricated with high performance; however, most of the research on β-Ga2O3 has focused on unipolar devices.
Great strides have been made in both high-power and RF β-Ga2O3 field-effect transistors (FETs) with continually improving material quality and fabrication processes. High-power lateral FETs have reported breakdown voltages up to 10 kV and BFOMs near 1 GW cm−2, while vertical devices have yet to achieve similar performance. The factors limiting vertical FETs are largely the lack of p-type dopants, which minimizes current-blocking capabilities, gate dielectric quality, stability, and robustness [9]. Many of the FET structural and material improvements, discussed in Section 3, that have been tested on lateral FETs can similarly be applied to vertical devices. Most β-Ga2O3 FETs are depletion-mode (D-mode), or normally on, due to the unipolar nature of β-Ga2O3 devices. The off-state leakage in D-mode FETs is more prominent than in enhancement-mode (E-mode), or normally off, FETs; however, their fabrication is more difficult and often requires band bending a heterointerface to deplete the existing channel. RF FETs are predominantly lateral devices with thin channel layers and highly scaled gate lengths for strong gate control and reduced parasitics. Techniques such as delta-doping and modulation doping are used to form a two-dimensional electron gas (2DEG) with a high carrier concentration and mobility. Maximum oscillating frequencies near 50 GHz with high breakdown electric fields have been reported, showing potential for high-power RF β-Ga2O3 FETs in the future. With the tremendous progress that has already been accomplished in the field, β-Ga2O3 presents itself as a strong candidate for high-power and high-frequency applications, but not without its challenges to overcome.
Previous review articles on β-Ga2O3 FETs have reported the chronological development in device design and performance [10], or focused specifically on RF FETs [7], E-mode FETs [11], or vertical GaN and β-Ga2O3 FETs [9]. Other review papers have covered FETs designed for both high-power and RF applications [12,13]. This review is formatted to aid current and future β-Ga2O3 high-power and RF FET researchers by separately discussing the different steps of FET fabrication, ranging from structures, materials, ohmic contacts, gate dielectrics, and material preparation. An overview of various material and FET defect characterization techniques is presented as well.
This article presents a comprehensive overview of the progress in β-Ga2O3 FETs, current challenges, and potential strategies to overcome them. Section 2 discusses the crystal structure and material properties, including FOM comparisons, bulk and epitaxial growth, and the doping of β-Ga2O3. Section 3 reviews many of the recent transistor designs for both high-power and RF applications. Section 3.1 focuses on structures that have been implemented, as well as proposed structures through technology computer-aided design (TCAD). Section 3.2 summarizes FETs with different channel and substrate materials such as semi-insulating homoepitaxial or heterostructure layers in the channel, and high thermally conductive substrates. Section 3.3 reviews metals and processes for high-quality ohmic contact formation and Section 3.4 gives an overview of different gate dielectrics used in β-Ga2O3 FETs. Section 4 discusses the importance of defect engineering, various characterization methods, and material preparation to improve interface quality. Section 5 gives a high-level overview of the current challenges and steps needed to add β-Ga2O3 devices to the market. Section 6 briefly summarizes the most promising applications and trends of β-Ga2O3 FETs. Section 7 then summarizes the advancements in β-Ga2O3 and provides an outlook of the future of β-Ga2O3.

2. Crystal Growth and Material Properties of β-Ga2O3

2.1. Different Phases

In 1952, Roy et al. discovered five polymorphs of Ga2O3 (α, β, γ, δ, and ε) using a gallia gel–water system, and determined that the β phase is the stable form [14]. Using first-principles calculations, Yoshioka et al. found that the theoretical formation energies of the different phases are in the order of β < ε < α < δ < γ, confirming that β-Ga2O3 is stable, while the other polymorphs show metastable behavior and transform into β-Ga2O3 at high temperatures [15]. In 2013, Playford et al. discovered another metastable phase (κ) via the thermal decomposition of Ga5O7(OH) above 500 °C [16]. The phase transitions compiled from Roy et al. and Playford et al. are shown in ref. [17].
The crystal structure of β-Ga2O3 is monoclinic and belongs to the C2/m space group with lattice constants of a = 12.2 Å, b = 3.0 Å, c = 5.8 Å, α = 90°, β = 104°, and γ = 90° (Figure 1a). The unique structure has two Ga sites, one with a tetrahedral geometry and one with an octahedral geometry, as well as three O sites, leading to high anisotropy in many of its material properties [18,19,20,21].

2.2. Material Properties

The band structure of β-Ga2O3, calculated using first-principles density functional theory (DFT) (Figure 1b), shows an indirect gap of 4.84 eV and a direct gap of 4.88 eV; however, β-Ga2O3 is largely considered a direct-gap semiconductor because of the close proximity of the gaps. The conduction band dispersion estimates an electron effective mass of ≈0.28 me, where me is the rest electron mass. The valence band, however, exhibits almost no dispersion and, therefore, exhibits a very large hole effective mass due to the localized self-trapping of holes [19,23].
Experimentally observed bandgaps range between 4.7 and 4.9 eV [19,24], projecting a critical breakdown electric field, Ebr, of 6–8 MV cm−1. Various figures of merit (FOM), discussed below, have been developed to compare semiconductors for high-power applications. The Baliga FOM (BFOM) is an estimate of DC conduction losses in a material and is defined as both ε·µ·Ebr3, where ε is the material dielectric constant and µ is the carrier mobility, and as Vbr2 Ron,sp−1 for devices, where Vbr is the breakdown voltage and Ron,sp is the specific on-resistance. The theoretical BFOM of β-Ga2O3 is approximately 28 GW cm−2 and ≈3214 times larger than that of Si. Other power device metrics include the Johnson FOM (JFOM), which represents the power–frequency capability; the Baliga high-frequency FOM (BHFFOM), which is a measure of switching losses; the Keyes FOM for thermal capability for power density and speed; and the Huang chip area manufacturing FOM (HCAFOM) as an indicator of chip area requirements. The material properties and FOMs of β-Ga2O3 compared to other materials are summarized in Table 1 [6,12,25].
It is important to note the low and anisotropic thermal conductivity in β-Ga2O3 of 27.0 W m−1 K−1 in the [010] direction and 10.9 W m−1 K−1 in [100] [26]. The difference in thermal conductivity of [010] β-Ga2O3 and [100] β-Ga2O3 might not seem large compared to those of other (ultra)-wide-bandgap ((U)WBG) materials; however, simulations have shown that the max temperature rise in devices has a decreasing rate dependence on thermal conductivity, and that ≈105 °C and ≈61 °C max temperature rises are simulated for [100] and [010] β-Ga2O3, respectively. On the other hand, the simulated max temperature rises for SiC and diamond are ≈34 °C and ≈30 °C, respectively [27].
At low doping densities below 1018–1019, electron interactions with polar longitudinal optical (LO) phonons is identified as the dominant scattering mechanism, limiting the theoretical bulk mobility to ≤250 cm2 V−1 s−1, while at higher doping concentrations, impurity scattering is dominant [28,29,30]. Even though β-Ga2O3 has a low mobility, β-Ga2O3 maintains higher FOMs than GaN and SiC because of the square or cubic dependence on breakdown voltage and only a linear dependence on mobility.
Table 1. Material properties and FOMs, relative to Si, of β-Ga2O3 compared to other semiconductors [6,12,31].
Table 1. Material properties and FOMs, relative to Si, of β-Ga2O3 compared to other semiconductors [6,12,31].
Material PropertiesSiGaAs4H-SiCGaNβ-Ga2O3Diamond
Bandgap, Eg (eV)1.11.43.33.44.95.5
Dielectric Constant, ε11.812.99.79105.5
Breakdown field, Ebr (MV cm−1)0.30.42.53.3810
Electron mobility, µ (cm2 V−1 s−1)1480840010001250200–2502000
Saturation velocity, vsat (107 cm s−1)11.222.51.8–21
Thermal conductivity, λ (W m−1 K−1)1505527021010.9–271000
BFOM = ε r μ E b r 3 114.7317846321424,660
JFOM = E b r 2 v s 2 / 4 π 2 11.8278108928441100
BHFFOM = μ E b r 2 110.146.3100.8142.21501
Keyes   FOM = λ c · v s / ( 4 π ε ) 1 / 2 10.33.61.80.241.5
HCAFOM = ε μ 0.5 E b r 2 154885279619

2.3. Crystal Growth

One of the greatest advantages of β-Ga2O3 is the potential for ultra-low-cost, large-size (diameter 100–150 mm), high-quality substrates made possible via melt growth. β-Ga2O3 is the only WBG semiconductor that can be grown from the melt, and therefore, the cost of a β-Ga2O3 wafer is expected to be approximately 80% cheaper than that of SiC [8]. The different bulk crystal growth techniques are edge-defined film-fed growth (EFG) [32,33], Czochralski (CZ) [34], vertical Bridgman (VB) [35,36], floating zone (FZ) [37,38], and Verneuil [39,40]. From all the methods, EFG has so far grown larger-diameter substrates with high quality, low defect densities, and a relatively wide doping range [22,41].

2.4. Epitaxial Growth

The main methods of β-Ga2O3 thin film growth that have been developed include molecular beam epitaxy (MBE), plasma-assisted MBE (PAMBE), metal–organic chemical vapor deposition (MOCVD), halide vapor-phase epitaxy (HVPE), and low-pressure chemical vapor deposition (LPCVD). MBE has the advantage of growing very high-quality thin films with less impurities and precise control over the growth rate and doping (1016–1020 cm−3). It suffers, however, from low growth rates of 0.05–0.18 µm h−1 that make it impractical for thick epitaxial layers used in vertical devices, but ideal for lateral thin-channel devices. PAMBE uses an activated oxygen source to help the growth of β-Ga2O3 thin films, and has been shown to reduce background (unintentional) impurity concentrations [42,43,44]. MOCVD, also called metal–organic vapor-phase epitaxy (MOVPE), also produces high-purity thin films with controlled doping (1017–8 × 1019 cm−3) at higher growth rates of 0.8 µm h−1 and less costly than MBE, making MOCVD conducive for large-scale production. HVPE has a minimum doping concentration at the order of 1015 cm−3 and a significantly higher growth rate, with the maximum reported rate of 250 µm h−1. It is therefore used in thick epi layer growth for vertical devices [45]. The tradeoff for the higher growth rates in HVPE is lower-quality thin films with rougher surfaces and more defects. LPCVD is a scalable and lower-cost method that produces high-quality thin films with growth rates ranging from 0.5 to 10 µm h−1, controlled doping in the range of 1017–1019 cm−3, and a heterostructure capability [46,47]. LPCVD is the least used of the three growth techniques, but can provide a path for scalable, production-level β-Ga2O3 wafers. Additionally, MBE, MOCVD, and LPCVD can grow heterostructures, unlike HVPE. More detail on these growth methods can be found in refs. [17,48,49].

2.5. Doping Strategies

DFT calculations have been used to find the energy levels of various impurities, oxygen vacancies (VO), and gallium vacancies (VGa) in the β-Ga2O3 bandgap. Oxygen vacancies act as deep donors more than 1 eV below the conduction band (EC), and gallium vacancies act as deep acceptors more than 1 eV above the valence band (EV) [50,51]. These vacancies do not contribute to conduction, but only act as doping compensation. Shallow donors found via DFT include SiGa(I) (Si impurity in GaI site), GeGa(I), SnGa(II), ClO(I), and FO(I), with energy levels very near EC [52]; however, the majority of experimentally used donors are Si, Sn, and Ge [53,54]. Acceptor impurities such as N, Sr, Zn, Cd, Ca, Be, Mg, and Fe all have levels more than 1.3 eV above EV, indicating that p-type doping is not possible, and is a major challenge in the development of β-Ga2O3 devices [50,55]. Deep acceptors are used to form highly resistive semi-insulating layers.
The donor levels of Si and Ge in MBE, LPCVD, CZ, and EFG samples, calculated using temperature-dependent Hall and conductivity measurements, ranged from 15 to 31 meV below EC, indicating shallow donors, while the Mg and Fe levels were located at EC—0.86 eV and EC—1.1 eV, respectively [53]. Mobility dependence on carrier concentration is expected to flatten at 250 cm2 V−1 s−1 as the carrier concentration approaches 1015 cm−3, and drops significantly as it increases above 1017 cm−3 (Figure 2) [31].
While p-type behavior is not available when using conventional methods, some groups have observed hole conduction when compensating donors are reduced [56]. The p-type conductivity of thin-film β-Ga2O3 by reducing the mean free path of carriers using amphoteric Zn doping was able to achieve ultra-high breakdown fields of 13.2 MV cm−1, beyond that of the theoretical β-Ga2O3 [57]. Another technique to modulate between high n-type and p-type conductivity was developed via controlled H incorporation, where p-type conductivity with an acceptor state 42 meV above EV was observed after direct H diffusion, and n-type conductivity with a donor state 20 meV below EC was observed after filling up oxygen vacancies by annealing in O2 [58].

3. β-Ga2O3 FET Designs

The following section reviews many of the current FET designs, including their structure, channel materials, substrate materials, ohmic contact formation, and gate dielectrics. Their process steps, use cases, advantages, and disadvantages are discussed. The tables below compare the many different device designs for D-mode high-power (Table 2), E-mode high-power (Table 3), and D-/E-mode RF applications (Table 4). Table 4 also includes the RF performance of mature GaN HEMTs and emerging hydrogen-terminated diamond HEMTs to illustrate the differences in performance of other material systems to β-Ga2O3.

3.1. β-Ga2O3 FET Structures

3.1.1. MESFETs and Delta Doping

The metal–semiconductor field-effect transistor (MESFET) in Figure 3a, fabricated by Higashiwaki et al., was the first demonstrated single-crystal β-Ga2O3 transistor [101]. Many of the following MESFETs, reported by the Rajan group, incorporated delta doping [60,88,102,103,104,105,106]. The delta-doping technique was first developed in 2017 by Krishnamoorthy et al. [102] in attempts to improve Si doping during PAMBE epi layer growth. The Si source oxidized quickly, reducing the Si doping level in the β-Ga2O3, creating doping spikes. Pulsing the Si shutter for 1 s over a period of 1 min removed the oxide and created uniform, highly doped regions with UID spacers (Figure 3b), leading to the delta-doping method for β-Ga2O3 devices. This resulted in a 2D electron gas (2DEG), high-electron-mobility transistor (HEMT) behavior with an increased carrier sheet concentration and mobility, as well as reduced contact and sheet resistances. These improvements and lower gate capacitance of MESFETs compared to MOSFETs lend delta-doped MESFETs more for RF applications. Regrown ohmic contacts, discussed in the Section “Regrown Layers”, are required for delta-doped FETs to reach the 2DEG because it is surrounded by UID β-Ga2O3. The Rajan group fabricated delta-doped MESFETs using regrown contacts, gate-connected field plates (GFP), and highly scaled T-gate structures with gate lengths (LG) down to 120 nm to improve their low- and high-frequency performance, with a BFOM of 118 MW cm−2 [60], a mobility of 95 cm2 V−1 s−1 [104], and a current gain cutoff frequency (fT) of 27 GHz (Figure 3c) [88]. The GFP and T-gate structures are also discussed in Section 3.1.5 and Section 3.1.7, respectively.
Bhattacharyya et al. reported high-performing, non-delta-doped lateral MESFETs using a combination of regrown ohmic contacts for a low contact resistivity of 8.3 × 10−7 Ω cm2 [107], GFP for a Vbr up to 4.4 kV [108,109], and a fin-shaped channel design surrounded by variable-temperature MOCVD-grown layers, achieving a mobility of 184 cm2 V−1 s−1, negligible hysteresis, and a BFOM of 0.95 GW cm−2 [63]. Passivating layers such as Al2O3 and SiNx can also be used to improve both low-frequency BFOMs and high-frequency Huang’s Material Figure of Merit (HMFOM) [63,66]. The current highest-reported FET breakdown voltage of 10 kV was achieved using a MESFET design with a T-gate structure, source-connected field plates (SFPs), SiNx passivation, oxygen annealing (OA), Si ion implantation, UID buffer layers surrounding the channel region, and B implantation for device isolation [59]. These device improvements will be discussed in more detail in later sections.

3.1.2. Self-Aligned Gate (SAG) FETs

The self-aligned gate (SAG) FET design is a well-known process developed to reduce series resistance and aggressively scale devices by reducing the source–gate spacing (LSG), essentially eliminating the source–gate access region. The earliest β-Ga2O3 SAG FETs, developed by AFRL, were designed by first depositing the Al2O3 gate dielectric using plasma-assisted atomic layer deposition (PA-ALD), which acts as the ion implantation cap. A refractory metal gate of W or W/Cr, able to withstand the high ion activation temperature, is patterned to protect the gate and drift regions. The source–gate and drain–gate access regions are then very highly doped via Si ion implantation and activated at 900 °C using rapid thermal annealing (RTA) for 2 min in N2 ambient conditions [89,110]. The gate metal is then etched via reactive ion etching (RIE) from the drift region and the ohmic contacts are formed (Figure 4a). A low contact resistance (RC) of 1.5 Ω mm, a sheet charge density (ns) of 4.96 × 1012 cm−2, and a Hall mobility of 48.4 cm2 V−1 s−1 were measured in these devices [110]. RF load–pull continuous wave (CW) power measurements were reported for these early SAG FETs, with a high output power (Pout), transducer gain (GT), and power-added efficiency (PAE) of 715 mW mm−1, 13 dB, and 23.4%, respectively, at 1 GHz (Figure 4a) [89,111].
A recent PAMBE-grown delta-doped SAG FET structure incorporated delta doping, in situ Ga etching for gate recess, and in situ Al2O3 gate dielectric growth, achieving sub-100 nm source–gate and gate–drain access regions [61]. A 30 nm Mg-doped layer was initially grown to compensate the Si impurities at the substrate/epi interface, followed by a 500 nm UID buffer layer. Then, two delta-doping layers, 5 nm apart, another 40 nm UID layer, and a 45 nm n++ cap layer were grown. The n++ cap layers were used as an alternative to ion implantation, enabling SAG. The SAG fabrication process (Figure 4b) began with ohmic contact fabrication and the plasma-enhanced chemical vapor deposition (PECVD) of SiNx with patterning to expose the gate region. The sample was placed into an MBE system, where in situ Ga etching of the n++ cap layer was performed at a substrate temperature of 550 °C and Ga flux of 1.5 × 10−7 Torr. Ga droplets were removed at a temperature of 600 °C, followed by 10 nm in situ Al2O3 deposition at a temperature of 400 °C. Conformal ex situ ALD was used to uniformly deposit 60 nm of Al2O3 in the gate and sidewall regions. Anisotropic RIE and isotropic BOE wet-etching of the Al2O3 reduced the gate and sidewall dielectric thickness to 20 nm and 50 nm, respectively. This FET outperformed the initial SAG FETs, with a source–gate access resistance of 1.3 Ω mm, an ns of 2.8 × 1013 cm−2, and a mobility of 65 cm2 V−1 s−1, leading to record peak DC and pulsed drain currents of 560 mA mm−1 and 895 mA mm−1, respectively, for a lateral FET on a native β-Ga2O3 substrate. The FET exhibited a high gate leakage and low current on/off ratio due to a low-quality gate dielectric or remaining Ga droplets at the interface. A drain current droop was observed in the DC measurements, indicating excessive self-heating, as shown by the red dashed line in Figure 4b. The use of SAG FETs has not yet been translated to vertical devices but is expected to improve performance for both low- and high-frequency operation.

3.1.3. Trench/Recessed-Gate FETs

Another FET design, first implemented by AFRL in 2017 [90], is the trench or recessed gate to scale devices down to sub-µm gate lengths for improved RF performance. The FET in ref. [90] was fabricated on an n+ 180 nm channel layer with an n++ 25 nm ohmic cap layer grown via MOVPE. Post-ohmic contact formation, the n++ cap layer was etched and 200 nm of SiO2 was deposited via PECVD as the passivation and field-plate dielectric. A 0.7 µm gate region was patterned on the SiO2 and etched via RIE nearly halfway into the epi layer, followed by ALD-Al2O3 deposition as the gate dielectric, Ni/Au gate stack evaporation, and the evaporation of interconnects (Figure 5a). The measured cutoff frequency (fT) and maximum oscillating frequency (fMAX) were 3.3 GHz and 12.9 GHz [90].
E-mode FETs are useful in reducing off-state power loss; however, they are difficult to fabricate in β-Ga2O3 due to the lack of p-type doping, large hole effective mass, and hole self-trapping. The recessed-gate approach is among the few early methods used to achieve E-mode operation by etching into the channel region, such that the remaining channel is fully depleted due to band bending at the oxide/epi and epi/substrate interfaces [112,113]. Chabak et al. [80] studied the band bending of a 200 nm Si-doped 5.5 × 1017 cm−3 epi due to 5.5 × 1012 cm−2 surface states at the SiO2/β-Ga2O3 interface, noting an approximately 100 nm depletion, and 34 nm of depletion due to the Fe-doped substrate. The E-mode FET with a threshold voltage (Vth) of +2 V was realized due to a gate recess of 140 nm (Figure 5b). E-mode recessed-gate FETs with an epi thickness of 200 nm, etch depth of 180 nm, and an LG of 2 µm were also reported to have high switching characteristics, with turn-on/off delay times of 4.0 ns/11.8 ns and rise/fall times of 24.6 ns/82.2 ns (Figure 5c). The longer fall time is attributed to low electron mobility and slow discharging from interface states. While switching losses are reduced with higher switching speeds, a high on-resistance (Ron), identified based on the VDS of ≈5 V in the top plot of Figure 5c, results in high on-state power losses that might be more limiting than switching losses [114]. A high Ron and increased power losses are observed for many trench FETs; however, further LG scaling by incorporating SAG can reduce the channel resistance contribution.
Figure 5. (a) First recessed-gate FET reported with sub-µm LG. Reprinted with permission from [90]. (b) First E-mode recessed-gate FET and respective transfer and output curves. © (2018) IEEE. Reprinted with permission from [80]. (c) Switching characteristics of a recessed-gate lateral FET where turn-on delay time, td(on), is defined as the time between 0.1Vgs and 0.1Ids. Likewise, td(off) is the time between 0.9Vgs and 0.9Ids. Similarly, the rise time, tr, is the time between 0.1Ids and 0.9Ids, and the fall time, tf, is the time between 0.9Ids and 0.1Ids. © (2019) IEEE. Reprinted with permission from [114].
Figure 5. (a) First recessed-gate FET reported with sub-µm LG. Reprinted with permission from [90]. (b) First E-mode recessed-gate FET and respective transfer and output curves. © (2018) IEEE. Reprinted with permission from [80]. (c) Switching characteristics of a recessed-gate lateral FET where turn-on delay time, td(on), is defined as the time between 0.1Vgs and 0.1Ids. Likewise, td(off) is the time between 0.9Vgs and 0.9Ids. Similarly, the rise time, tr, is the time between 0.1Ids and 0.9Ids, and the fall time, tf, is the time between 0.9Ids and 0.1Ids. © (2019) IEEE. Reprinted with permission from [114].
Materials 16 07693 g005
Various TCAD studies have been reported on the effects of doping and recess depth in E-mode recessed-gate FETs. The band diagram variation, electron concentration, and Vth at different channel thicknesses (epi thickness–recess depth) is shown in Figure 6a. At zero gate and drain bias, the electron concentration drops off very quickly as the channel thickness drops below 80 nm, down to ≈108 cm−3 at 50 nm, because of the oxide/epi and substrate/epi depletion. A threshold voltage near 0 V is observed for a 75 nm channel thickness going up to +4 V for 50 nm [115]. Decreasing the doping concentration in the channel layer is observed to both decrease the peak drain current and increase the Vth, such that a doping of 1 × 1016 cm−3 causes E-mode behavior (Figure 6b). A larger recess depth (smaller channel thickness) increased the Vth from ≈−50 V to near 0 V from the I-V transfer curves but decreased the drain current in the I-V output curves (Figure 6b) [116]. At high VGS and VDS, however, the drain current is nearly equal, indicating that the recess depth has little actual effect on the peak drain current. A slightly different trench FET design using body and epitaxial drift layers with different dopings, as well as recess through the entire drift layer, is shown in Figure 6c [117]. From I-V transfer curves with a drift layer doping of 3 × 1017 cm−3 and varying body doping from 1 × 1013 to 1 × 1017 cm−3, E-mode operation was only realizable for a body doping of 1 × 1015 cm−3 or less, showing a larger current and more negative Vth for higher doping concentrations. A 2D view of the electron concentration at a VGS of 0 V, NBody of 1 × 1015 cm−3, and NDrift of 3 × 1017 cm−3 (Figure 6c) shows normally off conditions due to the full depletion of the body layer from band bending at the oxide/body interface.

3.1.4. FinFETs

The first FinFET structures were lateral devices designed by Chabak et al. in 2016 using inductively coupled plasma (ICP) over-etching into the substrate to create thin 300 nm triangular-shaped fins as the channel (Figure 7a) [113]. E-mode operation was enabled by channel depletion due to the gate, with the I-V transfer curve shown in Figure 7a. Substrate conduction, shown by the red curve in Figure 7a, was observed and attributed to uncompensated carriers at the substrate surface. Hu et al. fabricated various vertical single-fin E-mode FETs [118,119,120] with current densities reaching 1 kA cm−2, a Vbr of 1.6 kV, and a low subthreshold slope (SS) of 80 mV dec−1, giving an interface trap state density (Dit) of >6 × 1011 cm−2 eV−1. Interface traps were observed to reduce the field-effect mobility and current density by depleting the channel, as well as limiting breakdown by exacerbating drain-induced barrier lowering (DIBL) [118]. Single- and multi-fin E-mode FETs were later fabricated by Li et al. with single/multi-fin current densities of 2 kA cm−2/230 A cm−2, a Ron,sp of 35.2 mΩ cm2/25.2 mΩ cm2, and BFOMs of 172 MW cm−2/280 MW cm−2 for a fin width (Wfin) of 0.15 µm. Another advantage of multi-fin FETs is that, unlike single-fin FETs, current spreading does not drastically change the active area, making the BFOM and Ron,sp less ambiguous. The fabrication was performed on a 10 µm HVPE epi layer with doping of 2 × 1015 cm−3 grown on a conductive substrate. First, the epi layer was Si-ion implanted and activated at 1000 °C for the source ohmic contact, followed by e-beam lithography and dry etching to form sub-µm fin channels. A Ti/Au stack was deposited on the backside as the drain contact and a 35 nm ALD-Al2O3 was used as the gate dielectric. A sputtered Cr gate metal and 120 nm ALD-Al2O3 spacer were patterned with an SAG process. Finally, a Ti/Al/Pt stack was sputtered, forming the source and source-connected field plate. The devices were measured before and after post-deposition annealing (PDA) at 350 °C in N2, resulting in significant improvements (Figure 7b) [81]. It has been shown that Vth strongly decreases with increasing Wfin and ND (Figure 7c), giving a small window for normally off devices. The previously mentioned FinFETs were fabricated with sub-0.5 µm Wfin and ND below 1 × 1016 cm−3, enabling a positive Vth. Because Si doping below 3 × 1015 is difficult in epi growth, a resistive layer via nitrogen doping of 1 × 1016 cm−3 during HVPE growth was shown to significantly minimize the Vth dependence on Wfin and achieve normally off operation for Wfin up to 2 µm (Figure 7c) [121].
Other vertical FinFETs have been reported on (100) oriented substrates to potentially reduce intrinsic growth defects, even though the majority are fabricated on (001) substrates [122]. Lateral tri-gate FinFETs have also reported high RF performance [91], high BFOMs of 0.95 GW cm−2, and mobilities of 184 cm2 V−1 s−1 using high/low temperature MOCVD growths [63], as mentioned in Section 3.1.1.
A highly selective wet-etching technique, metal-assisted chemical etching (MacEtch), is an attractive, damage-free alternative to dry etching that is typically used in FinFET fabrication [123]. More details on the MacEtch techniques and chemical reactions can be found in Ref. [124]. Lateral FinFETs fabricated via MacEtch have recently been reported (Figure 8a), with an aspect ratio of 4.2:1, an Ron,sp of 6.5 mΩ cm2, and a BFOM of 21 MW cm−2 [125]. The lowest SS, Vth, and hysteresis of 87.2 mV dec−1, −6.9 V, and 24 mV, respectively, were measured on FinFETs with a 90° orientation from the [102] direction (Figure 8a). A previous study showed that fins perpendicular to [102] had the most vertical sidewalls and lowest Dit of 2.73 × 1011 cm−2 eV−1 [126]. DC I-V measurements of these FinFETs at high temperatures up to 298 °C (Figure 8b) reported increasing off-state currents and a lower on/off ratio, attributed to thermionic emission from source to drain, a decrease in Vth by ≈20 V due to trapping/de-trapping at the gate metal/oxide and oxide/semiconductor interfaces, and increasing hysteresis up to 4.29 V and SS up to 1.35 V dec−1, indicating thermal degradation of the interface or dielectric [127].

3.1.5. Gate-Connected Field Plates

It is widely known that field plates can improve device breakdown by reducing the peak electric field near the contact edges. The gate-connected field plate (GFP) extends into the gate–drain access region, where most of the voltage drop occurs, “spreading” the electric field. The first β-Ga2O3 GFP FET was reported by Wong et al., using SiO2 as the FP dielectric (Figure 9a). TCAD simulations of the peak electric field for various field-plate heights, hFP, and field-plate drain lengths, LFP,D, at the drain edge of the gate (top), depicted by the symbol x in the diagram, and drain edge of FP (bottom), depicted by the symbol * in the diagram, are shown in Figure 9a. Increasing the LFP,D is seen to quickly reduce the electric field at the gate edge, while having little effect on the field at the FP edge. However, as the hFP increases, the field at the gate edge rises while the field at the FP edge falls, indicating an ideal window for hFP [128].
Since then, other SiO2 FPs, SiO2 composite FPs with polymer passivation, and SiNx FPs passivated with SiO2 have been reported [64,92,108,109,129,130,131] with some of the highest breakdown voltages and BFOMs of 8.56 kV and 355 MW cm−2, respectively. SiNx is better suited both to spread electric fields due to its higher dielectric constant and in mitigating virtual gate effects originally discovered in AlGaN/GaN HEMTs [132], but is also mentioned as a possible mechanism of current dispersion [133] and series resistance increases [134] in β-Ga2O3 FETs.
Figure 9. (a) GFP FET cross-section with the symbols x and * indicating peak electric fields in the channel. Plots of simulated breakdown electric field dependence on LFP,D and hFP are shown at the location of symbol x (top plot) and * (bottom plot). © (2016) IEEE. Reprinted with permission from [128]. (b) FET with composite PECVD-SiO2/ALD-SiO2 GFP and SU8 passivation used to increase Vbr. © (2020) IEEE. Reprinted with permission from [131]. (c) GFP FET similar to that in (b) but with SU8 as part of the FP and vacuum annealing, increasing Vbr and reducing Ron. © (2022) IEEE. Reprinted with permission from [64].
Figure 9. (a) GFP FET cross-section with the symbols x and * indicating peak electric fields in the channel. Plots of simulated breakdown electric field dependence on LFP,D and hFP are shown at the location of symbol x (top plot) and * (bottom plot). © (2016) IEEE. Reprinted with permission from [128]. (b) FET with composite PECVD-SiO2/ALD-SiO2 GFP and SU8 passivation used to increase Vbr. © (2020) IEEE. Reprinted with permission from [131]. (c) GFP FET similar to that in (b) but with SU8 as part of the FP and vacuum annealing, increasing Vbr and reducing Ron. © (2022) IEEE. Reprinted with permission from [64].
Materials 16 07693 g009
Zeng et al. in the Singisetti group used a composite FP composed of a thick 350 nm PECVD-SiO2 below a denser, high-quality 50 nm ALD-SiO2 layer to improve breakdown [129,130]. Sharma et al., also in the Singisetti group, then improved on the GFP design by adding a polymer, SU8, passivation layer on the composite FP and S/D regions, reaching some of the highest reported breakdown voltages of 8.03 kV (Figure 9b) and 8.56 kV (Figure 9c) [64,131]. The high Ron led to low BFOMs, but vacuum annealing before FP deposition resulted in a ×10 reduction in Ron, with little change in Vbr (Figure 9c) [64].

3.1.6. Source-Connected Field Plates

Source-connected field plates (SFPs) are another viable FP strategy where the source metal extends past the gate, and can be considered a better field-spreading method across the gate region and at the drain side of the gate [135,136]. One of the first SFP FETs measured a BFOM of 50.4 MW cm−2 in 2019 (Figure 10a) [137]. Simulated electric field profiles in Figure 10a show field spreading and a reduced peak electric field using an SFP. A T-gate structure can be used in conjunction with SFPs for further field management, achieving higher BFOMs of 277 MW cm−2 (Figure 10b [65]) and a record Vbr of 10 kV (Figure 10c [59]).

3.1.7. T-Gates

T-gates, as mentioned previously, are unique in that they not only improve breakdown as a field-plate structure [59,62], but also improve RF results in thin-channel FETs by decreasing the LG while maintaining a large cross-section. This reduces the gate access resistance and decreases the electron transport time, but does not degrade the noise figure [138]. Various T-gate RF FET structures are shown in Figure 11a–e that incorporate recessed gates with SiO2 FP dielectrics (Figure 11a [139]), air FP dielectrics with implanted channels (Figure 11b,c [94,140]), MESFET with Al2O3 passivation (Figure 11d [66]), and SiO2 gate dielectrics with SiNx passivation (Figure 11e [93]). The FET in Figure 11e has the highest-reported fmax to date of 48 GHz and a high breakdown field of 5.4 MV cm−1. A T-gate MESFET with an fT of 27 GHz was also discussed and shown in Figure 3b [88]. RF FETs with a T-gate structure must use highly scaled LG, typically in the range of 100–300 nm, for peak RF performance.

3.1.8. Semiconductor-on-Insulator (SOI)

Another important property of β-Ga2O3 is the anisotropic cleavage planes, making the (100) plane easy to exfoliate as a nanomembrane, similar to graphene. This makes the fabrication of β-Ga2O3 devices on different substrates, or heterojunctions with unconventional materials such as transition metal dichalcogenides (TMDs), much simpler. The first SOI β-Ga2O3 FET was reported in 2014, with an exfoliated β-Ga2O3 layer placed on a p+ Si wafer with 285 nm of thermally grown SiO2 acting as the gate oxide [141]. The SOI FET was then fabricated via back-gate metal and top-source/drain ohmic contact deposition. The corresponding I-V curves proved that channels could be created using the mechanical exfoliation of β-Ga2O3. Other p+ back-gate SOI FETs have been fabricated and studied [68,82,142,143,144,145,146,147,148,149,150,151]. One advantage of SOI FETs compared to non-SOI FETs is that more devices can be fabricated from β-Ga2O3 wafers and, therefore, studies on transport, irradiation, thermal effects, etc., can be more cheaply and readily performed. Different methods of tuning the Vth have been implemented, such as by varying the β-Ga2O3 channel layer thickness [82,145], fluorine plasma [147], adding a p-type material such as p-SnO on the channel of back-gate FETs [85], and using a fixed back-gate bias (VBG) on top-gate FETs. A top-gate FET with VBG shows variation in transconductance (gm) and Vth with VBG, obtaining a Vth of 0 V for VBG ≈ 6 V (Figure 12a) [152]. Other studies about defect impacts on current dispersion [148], proton irradiation [144], scattering mechanisms [153], and device improvement using thermal management via different thermally conductive substrates have been reported using SOI FETs [71,73,154,155,156,157,158,159]. These will be discussed more in the Sections “AlN/GO”, “SiC/GO”, and “Diamond/GO”. High-performing SOI FETs have also been realized, with the highest reported mobility of 191 cm2 V−1 s−1 (Figure 12b [85]), high current densities reaching 1.5 A mm−1 (Figure 12c [68]), an SS of 61 mV dec−1 very near the thermionic limit using a TMD-TaS2/β-Ga2O3 heterojunction (Figure 12d [160]), a Vbr up to 800 V [67], and BFOMs of 100 MW cm−2. The high BFOM of 100 MW cm−2 is achieved by a β-Ga2O3-on-SiC FET using ion cutting, a novel heterogenous wafer-scale integration technique for β-Ga2O3 on SiC [156,159].
Other SOI FETs have integrated β-Ga2O3 nanomembranes with various p-type 2D materials such as WSe2 [161,162], MoTe2 [162], and black phosphorus (BP) [163] to act as p-type gates, and large work function materials such as NbS2 and TaS2 to improve SS (61 mV dec−1) and off-state behavior [160]. Double-gate FETs using top-gate dielectrics such as HfO2 [152], h-BN [164], and bottom-gate dielectrics as the SiO2 on p-Si wafers have been utilized for improved gate control and Vth tuning. Monolithically integrated top and bottom graphene gates with both E-mode and D-mode FETs on the same layer are one of the first mentions of a β-Ga2O3 logic circuit with both E-mode and D-mode FETs [165].

3.1.9. Other Novel Structures

Most of the designs discussed in the previous sections were first developed in the early stages of β-Ga2O3 devices and iteratively improved upon. With the help of TCAD simulation, novel structures with high potential can be initially proposed. One of the recently proposed structures (2022–2023) includes a lateral field-plated MOSFET with a self-aligned trench vertical gate (Figure 13a [166]). The channel from source to drain starts with a UID β-Ga2O3 buffer layer with ion implantation in the source region. The UID buffer layer separates the ion-implanted source to the n+ horizontal channel, reaching the drain. The trench portion of the gate falls below the n+ channel into the UID buffer layer. The dominant channel becomes the vertical UID portion, which is highly controlled and not restricted by high-resolution photolithography. This structure has been proposed for AlGaN/GaN HEMTs and shown to improve drain current and transconductance [167]. The similar β-Ga2O3 device incorporates a GFP with a SiO2 FP dielectric to improve Vbr and current uniformity in the channel.
Another proposed device is a lateral MOSFET with an SFP, where the FP dielectric is air/SiNx and the FP makes contact with the SiNx in the gate–drain drift region (Figure 13b [168]). The electric field plot (Figure 13b) mentions device 1 (the proposed device), device 2 incorporating a GFP, and device 3 with no FP. The proposed device exhibits a higher BFOM of ≈2.2 GW cm−2 compared to 1.6 GW cm−2 for device 2 and 106 MW cm−2 for device 3. The capacitances, Cgd and Cgs, of the air gap device were slightly higher than the non-FP device, resulting in a slightly lower BHFOM, but an overall much larger JFOM of 7.8 THz V.
Gate-all-around (GAA) FETs are another newer FET used to scale below 10 nm in Si, but has not yet been realized in β-Ga2O3. A proposed β-(AlGa)2O3/Ga2O3 GAA FP HEMT is simulated with a high Pout of ≈22 kW and an fT of 2.4 GHz, showing potential for future GAA β-Ga2O3 FETs (Figure 13c [169]).
The lack of p-doping and very low hole mobility in β-Ga2O3 has limited most devices to unipolar operation. However, recently, there has been a TCAD investigation of potential β-Ga2O3 heterojunction bipolar transistors (HBTs) using p-type oxides. An npn with p-CuO2 showed HBT behavior and current gain (Figure 13d), but both the current gain and the breakdown electric field are strongly limited by interface traps and the low bandgap of CuO2 (2.1 eV) [170]. They mention that other p-oxides like NiO could apply, and that an (AlxGa1−x)2O3 layer could be used as the emitter to reduce the electron barrier into the base. For future designs, they propose specifications for minority carrier transport, emitter-base CBO, and a threshold for interface trap state density.

3.2. Channel and Substrate Materials

This section will mainly discuss β-Ga2O3 FETs designed by using different materials and processes as opposed to, in the previous section, FET structures and patterning.

3.2.1. Current Aperture Vertical Transistors and U-Shaped Trench MOSFETs

Vertical FETs are more suited for high-power applications than lateral FETs due to their higher power densities and smaller size, because breakdown voltage scales with drift layer thickness, as opposed to lateral devices where it scales with LGD, sacrifice the chip area. Current aperture vertical transistors (CAVETs), motivated by their counterparts in Si [171], SiC [172], and GaN [173], use current-blocking layers (CBLs) to reduce off-state drain currents and improve on/off ratios. The CBLs can either surround the source from the drift layer for E-mode purposes (Figure 14a [83]), or leave an opening/aperture for carriers to access the channel. For the later type, E-/D-modes are dependent on the doping of the channel [69,83,174,175,176]. CAVETs with only the channel doping, nch, varying from 5 × 1017 to 1.5 × 1018 cm−3 show E-mode behavior for a doping of 5 × 1017 cm−3 and D-mode behavior for the other dopings (Figure 14b [174]). Varying the aperture length, Lap, creates diode-like behavior with increasing on-voltage, Von, as the Lap decreases, possibly due to a fixed sheet charge of 1011–1012 cm−2 originating from defects diffused from the CBL (Figure 14c [176]). Semi-insulating CBLs using Mg2+ ion implantation were first used, but resulted in high leakage currents due to large Mg diffusion during the activation anneal [175]. The nitrogen thermal diffusivity in β-Ga2O3 is much lower compared to Mg [177], and therefore, CBL layers formed via N2+ ion implantation have exhibited less leakage [69,174,176,178].
The U-shaped trench-gate MOSFETs (UMOSFETs) are known to increase the packing density and reduce input capacitance when compared to planar vertical FETs. They have only recently been studied in the β-Ga2O3 material system using CBLs (Figure 14d), with both oxygen annealing, discussed in the following section, and N ion implantation [84,179]. Oxygen annealing formed an insulating CBL but increased the contact and sheet resistance in the n+ layers used for source ohmic contacts, while N implantation did not have these setbacks and showed higher current densities. One major drawback of these devices so far is the low on/off ratio of ≈7 × 104.

3.2.2. Oxygen Annealing

Oxygen annealing (OA), mentioned in the previous sections, increases the resistivity of n-type β-Ga2O3. The intended mechanism is reducing oxygen vacancies, which act as deep donors, thereby increasing acceptor compensation [180]. However, there is still some uncertainty in the true mechanism due to the multiple Ga and O sites in the unit cell, and complex substitutions that can occur at high temperatures [181,182]. OA was first observed to reduce the effect of a secondary conducting channel due to the UID layer, improving the pinch-off, output power density, and high-speed performance [95]. While there are few reports of OA FETs, there is potential for the future incorporation of OA (Figure 10c) [59,84,183].

3.2.3. Heterostructures

(AlxGa1−x)2O3/β-Ga2O3 Modulation-Doped FETs

The (AlxGa1−x)2O3/β-Ga2O3 (AlGO/GO) heterostructure has been studied intensively because of the observed carrier confinement near the AlGO/GO interface due to the conduction band offset (CBO) of ≈0.6 eV [184,185]. The first modulation-doped FETs (MODFETs) used a Ge-doped AlGO layer between two UID-AlGO layers to generate a 2DEG below UID-β-Ga2O3 [186]. However, the majority of later MODFETs incorporated delta doping in the AlGO layer which, along with the AlGO/GO CBO, generated a 2DEG inside the UID-β-Ga2O3 near the AlGO/GO interface. This avoids reductions in mobility due to the dopants of the 2DEG (Figure 15a) [187,188]. Like delta-doped FETs, regrown ohmic contacts are typically used to reach the 2DEG.
Various enhancements—such as double-heterostructure MODFETs with a UID β-Ga2O3 quantum well (Figure 15b [189]) and reduced spacer lengths, defined as the distance between the β-Ga2O3 and delta doping, down to 1 nm [190,191]—are used to increase carrier concentration. The 2DEG charge density typically is in the range of 1 × 1012–5 × 1012 cm−2, while the highest of 1.1 × 1013 cm−2 was measured in a double heterostructure, 1 nm spacer, high-k gate dielectric MODFET [79]. Field plates [70] and high-k gate dielectrics have shown to improve breakdown up to 5.5 MV cm−1 [79].
A modification of the MODFET, named the heterostructure FET (HFET), uses a heavily doped AlGO spacer as opposed to a delta-doping layer with a UID spacer (Figure 15c). E-beam lithography scaled LSG down to 55 nm, reducing the parasitic resistance, reporting near-record fT and fmax values of 30 GHz and 37 GHz, respectively, and minimal RF degradation at temperatures up to 250 °C [96,97].

AlN/GO

AlN is potentially a better alternative to AlGO/GO MODFETs with its higher CBO to β-Ga2O3 of ≈1.7–1.86 eV and polarization-induced charge, leading to larger 2DEG concentrations of 3 × 1013–5 × 1013 cm−2 [192,193,194]. Currently, no AlN/β-Ga2O3 HEMTs have been fabricated and reported; however, multiple TCAD simulations show promise, with much higher frequency operations up to an fT of 166 GHz and fmax of 142 GHz [195,196,197].
The thermal heat sink advantages of AlN, with a thermal conductivity of ≈320 W m−1 K−1 [25], to β-Ga2O3 has also been studied in SOI MOSFETs on an AlN/Si substrate for heat dissipation, showing little-to-no current dispersion between DC and pulsed I-V (Figure 16a [71]).

SiC/GO

SiC, with its high thermal conductivity of 370 W m−1 K−1, is mainly used for heat dissipation when forming a heterostructure with β-Ga2O3 [25]. TCAD simulations of p-SiC replacing the semi-insulating substrate in β-Ga2O3 FETs reduced peak temperatures by 100 °C. High breakdown voltages and on-currents were maintained by increasing the SiC thickness and the β-Ga2O3 doping to avoid premature breakdown in SiC [198,199]. Both the growth [200] and process/integration methods [156,201] have been developed for β-Ga2O3/SiC heterojunctions. Fabricated transistors have used the SiC layer only as a thermal heat sink (Figure 16b) [72,159], while others have used p-SiC to also behave as a back gate [158]. Figure 16b shows a composite SiC/β-Ga2O3 MOSFET formed via fusion bonding [201] with a large reduction in the temperature rise with a SiC substrate compared to a β-Ga2O3 substrate [72].

Diamond/GO

Diamond has one of the highest, if not the highest, thermal conductivity of any semiconductor at approximately 2290 W m−1 K−1. For this reason, diamond is commonly used in high-power applications for heat dissipation, such as cooling mechanisms with micro channels demonstrated in GaN [5]. TCAD simulations have confirmed the advantages of using a nanocrystalline diamond (NCD) substrate compared to β-Ga2O3 or SiC substrates [202]. In β-Ga2O3, thermal studies of both exfoliated nanomembranes on diamond substrates and the ALD growth of polycrystalline β-Ga2O3 on diamond substrates have been realized, with the nanomembrane reporting better thermal boundary conductance (TBC) due to the low thermal conductivity of polycrystalline β-Ga2O3 [157,203]. SOI MOSFETs on diamond exhibited higher drain currents of 980 mA mm−1 and a 60% reduction in temperature rise compared to similar devices on a sapphire substrate (Figure 16c) [73,154]. Additionally, a study on various device thermal-cooling methods concluded that the best performing solution in terms of the lowest temperature rise and thermal resistance consisted of cooling devices from the top contact nearest to the junction, named junction-side cooling, through flip-chip hetero-integration via thermal bumps to a diamond carrier, and NCD passivation with a thermal conductivity of 400 W m−1 K−1, labeled as FC3 in Figure 16d [204].
Figure 16. Thermal studies using AlN, SiC, and diamond. (a) SOI FET on an AlN/Si substrate showing effective heat dissipation when comparing DC and pulsed I-V. Reprinted from [71]. (b) SiC/GO composite wafer MOSFET with significant reduced temperatures at high power densities. Reprinted with permission from [72]. Copyright 2023 American Chemical Society. (c) MOSFET with diamond substrate and I-V curves compared to FETs, with other substrates showing significant current dispersion caused by self-heating. Reproduced from [73]. CC BY 4.0. (d) Simulation and comparison of device-level cooling methods. © (2019) IEEE. Reprinted with permission from [204].
Figure 16. Thermal studies using AlN, SiC, and diamond. (a) SOI FET on an AlN/Si substrate showing effective heat dissipation when comparing DC and pulsed I-V. Reprinted from [71]. (b) SiC/GO composite wafer MOSFET with significant reduced temperatures at high power densities. Reprinted with permission from [72]. Copyright 2023 American Chemical Society. (c) MOSFET with diamond substrate and I-V curves compared to FETs, with other substrates showing significant current dispersion caused by self-heating. Reproduced from [73]. CC BY 4.0. (d) Simulation and comparison of device-level cooling methods. © (2019) IEEE. Reprinted with permission from [204].
Materials 16 07693 g016

3.3. Source and Drain Ohmic Contacts

This section covers the design of ohmic contacts, including metals, processes, and techniques, to decrease contact resistance and improve ohmic behavior.

3.3.1. Metals and Processes

By far, the most common metal stacks for ohmic contact formation are Ti/Au or Ti/Al/Ni/Au due to the low metal work function of Ti (≈4.3 eV) [101,205]. In the earliest reports of β-Ga2O3 devices, BCl3 RIE was performed following metal evaporation and lift-off [206], but more recent reports use RTA at 400–500 °C for 1 min in N2. Comprehensive studies on the Ti/Au interfacial reactions found that the interdiffusion of Ti and Au, as well as a thin Ti-TiOx interlayer partially lattice-matched to β-Ga2O3, is responsible for ohmic contact formation (Figure 17a) [207,208]. Additionally, Si ion implantation and RIE improved the thermal stability and lowered resistivity in Ti/Au ohmic contacts [209]. Ti/Au ohmic contacts were observed to perform better in the (001) and (−201) orientations compared to (010), potentially due to more dangling bonds and higher surface energy in the (001) and (−201) directions [210].
Yao et al. investigated the capability of forming ohmic contacts with various metals such as Ti, In, Ag, Sn, W, Mo, Sc, Zn, and Zr, with and without an Au capping layer, concluding that Ti/Au with RTA at 400 °C for 1 min in Ar resulted in ohmic behavior with the lowest contact resistance [211]. Temperatures at 500+ °C degraded the Ti/Au contact and increased the resistivity. In, with its work function of 4.1 eV, also showed ohmic behavior after annealing at 600 °C for 1 min in Ar, but is not practical due to its low melting point. All other metals exhibited pseudo-ohmic or non-ohmic behavior, concluding that Ti/Au is the ideal metal stack of the nine used.
Other metals that have been found to form ohmic contacts on β-Ga2O3 are Mg/Au, with a work function of 3.8 eV [212]. The electrode resistance was found to increase with increasing annealing temperature from 300 to 500 °C due to Mg oxidation. Annealing at 300 °C and 500 °C found that the current density was not consistent 37 days later, while that at 400 °C was the same, indicating long-term stability. In order to withstand high temperatures, a refractory metal alloy TiW ohmic contact is feasible to highly doped (1 × 1019 cm−3) β-Ga2O3 [213].
Figure 17. (a) Formation of a defective β-Ga2O3 and TiOx layer enabling ohmic contacts. Reproduced from [207]; licensed under a Creative Commons Attribution (CC BY) license. (b) Effective doping vs. annealing temperatures of Si implantation and I-V curves for various Si concentrations after annealing at 950 °C. Reproduced from [214]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (c) Regrown ohmic contacts reducing contact resistance as shown through TLM measurements of the regrown layer (TLM-A) and to the channel layer (TLM-B). Reproduced from [107]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved.
Figure 17. (a) Formation of a defective β-Ga2O3 and TiOx layer enabling ohmic contacts. Reproduced from [207]; licensed under a Creative Commons Attribution (CC BY) license. (b) Effective doping vs. annealing temperatures of Si implantation and I-V curves for various Si concentrations after annealing at 950 °C. Reproduced from [214]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (c) Regrown ohmic contacts reducing contact resistance as shown through TLM measurements of the regrown layer (TLM-A) and to the channel layer (TLM-B). Reproduced from [107]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved.
Materials 16 07693 g017

3.3.2. Improvements

Decreasing the ohmic contact resistance improves both low- and high-frequency performance and is thus a critical component of FET design.

Ion Implantation

Si ion implantation was the first technique, starting in 2013 [214,215], to reduce contact resistance, and has been used in many future FET designs to form n+ layers in the source/drain regions before metal evaporation. In β-Ga2O3, ion implantation is achieved by first depositing a thick cap layer and patterning to expose the implant regions. Secondly, the implant depth and dose must be determined iteratively, followed by ion activation. For Si implantation, ions are typically activated at 900–950 °C for 30 min in N2 [66,216]. Figure 17b shows the annealing temperature effects on Si ion implant concentrations ranging from 1019 to 1020 cm−3, with an ideal temperature window in the range of 900–1000 °C. The I-V curves show the least resistance at a Si concentration of 5 × 1019 cm−3 for annealing at 950 °C [214]. Some use various implant steps for either more conductive regions or for semi-insulating dopants, like N or Mg, such as in CBL formation or some E-mode FETs [112,128,174,175]. Ge ion implantation for n+ regions has recently been investigated with pulsed RTA from 900 to 1200 °C for dopant activation. While one of the lowest specific contact resistivity (ρC) values of 4.8 × 10−7 Ω cm2 was reported at 1100 °C, a strong redistribution of the ions was observed along with an increased surface roughness [217].

Regrown Layers

Regrown layers are an alternative to ion implantation that avoid damage caused by the high-energy ions and high annealing temperatures. The fabrication process for regrown layers is outlined as follows: a sacrificial layer (typically SiO2) is patterned, followed by etching through the SiO2 and a portion of the epitaxial layer in the exposed source/drain regions. The sample is then placed in a growth chamber, and an n+ layer for ohmic contact evaporation is grown, followed by wet-etching the sacrificial layer to lift-off the regrown layers outside the source/drain regions [104,107,108,109,187,189,218]. Besides avoiding the damage caused by ion implantation, regrown layers are commonly used to contact a 2DEG in a delta-doped FET or MODFET. Initial Si delta doping before regrowth is used to neutralize any F ions, incorporated during dry etching, that could deplete the channel, confirmed to have low contact resistance using the transfer length method (TLM) (Figure 17c [107]). Recently, ultra-low ρC values for both β-Ga2O3 and β-(AlxGa1−x)2O3 of 1.62 × 10−7 Ω cm2 and 5.86 × 10−6 Ω cm2, respectively, have been reported, using dopings up to 3.2 × 1020 cm−3 [219].

Intermediate Layers

In WBG semiconductors, adding intermediate layers with lower bandgaps and/or higher doping concentrations can reduce the barrier for carrier transport to and from the contact. In β-Ga2O3, the most common intermediate layers are indium–tin–oxide (ITO) and aluminum–zinc–oxide (AZO). Both ITO and AZO, deposited via sputtering, have been used to form ohmic contacts. The ohmic contacts were formed on a 2 × 1017–3 × 1017 cm−3 doped β-Ga2O3 epi with varying annealing temperatures of 900–1150 °C and 500–600 °C for ITO [220,221], and 400–600 °C for AZO [222].

Diffusion Doping (Spin-on-Glass)

Diffusion doping, or spin-on-glass (SOG) doping, is one of the least common ways to both dope β-Ga2O3 and improve ohmic contact resistance. It provides a lower-cost and simpler fabrication process than the typical ion implantation or regrowth methods, as well as more predictable diffusion during activation and peak doping at the surface, ideal for ohmic contacts. The process begins with Sn-doped SOG spin-coated on a β-Ga2O3 epi layer and RTA at 1200 °C for 3–5 min in N2 to activate the dopants, followed by the removal of the SOG layer via dipping in buffered HF (BFH) for 10 min. A low ρC of 2.1 × 10−5 Ω cm2 was reported, and lateral MOSFETs reported improved peak current density and transconductance as well as a high thermal stability up to 200 °C using SOG [223,224].

3.4. Gate Dielectrics

3.4.1. Materials and Processes

The choice of gate dielectric material is vital to high-performing β-Ga2O3 FETs. The main material used is (PA)-ALD Al2O3 because of its large bandgap of 6.4–6.9 eV, with both electron- and hole-blocking capabilities, and similar composition to β-Ga2O3 [225]. The first MOSFETs, as well as many of the MOSFETs discussed previously, use Al2O3 as their gate dielectric [205,215]. While the main deposition method is (PA)-ALD, some initial reports of using in situ MOCVD-grown Al2O3 immediately after β-Ga2O3 epi growth have garnered attention, measuring lower interface defect densities and higher-quality Al2O3 to improve breakdown characteristics [226,227].
The second most common gate dielectric is SiO2, with the advantage of a higher bandgap around 9 eV but a lower dielectric constant, which becomes important in distributing electric field profiles, discussed later. The deposition of SiO2 can be performed either via PECVD or ALD [129,130].
Low defect densities have been reported for Al2O3 using solvent, O2 plasma, piranha, and BHF surface cleaning before ALD deposition and in situ forming-gas post-deposition annealing (PDA) at 250 °C [228], as well as SiO2 with a solvent only [229] or both solvent and piranha cleaning [230].

3.4.2. p-Gates

The more recently investigated gate dielectrics are p-type materials. The main materials used are p-NiO, p-GaN, and p-SnO. P-gated FETs, referred to as heterojunction FETs (HJ-FETs), are unique in that they provide vertical channel depletion for pinch-off while also increasing Vbr due to the pn junction between the gate and channel. This allows for thicker and higher doped channel regions, leading to higher currents and lower Ron, without a reduction in Vbr.

p-NiO

P-NiO has gained significant attention as a candidate for pn heterojunctions, with a wide bandgap of 3.7–4.0 eV and controllable p-doping in the range of 1016–1019 cm−3. Additionally, the highest-recorded BFOM of 13.2 GW cm−2 in β-Ga2O3 was recently reported on p-NiO/n-β-Ga2O3 heterojunction diodes [231]. P-NiO is typically sputtered at room temperature with a hole concentration modulated using the Ar/O2 ratio [75,232]. The theoretical CBO and VBO of p-NiO to β-Ga2O3 is expected to be 2.2 eV and 3.3 eV, respectively, while experimentally observed values vary greatly due to the polycrystalline p-NiO from sputtering [74,233]. P-NiO gated FETs have reported BFOMs of 0.39 GW cm−2 [74], and with incorporated recessed-gate p-NiO bi-layers, a T-gate structure with a SiO2 FP, and piranha treatment, a negligible hysteresis of 4 mV, an SS of 66 mV dec−1, and a BFOM of 0.74 GW cm−2 were realized (Figure 18a [75]). The recessed gate is also useful in creating E-mode FETs [234]. One difficulty, however, is the increased gate leakage which occurs when the pn junction becomes forward-biased. A proposed solution, studied initially through TCAD simulations and verified using fabricated devices, is the addition of a dielectric layer between the p-NiO and gate metal to suppress gate leakage when the pn junction is forward-biased [76,235]. A 20 nm ALD-SiO2 interlayer reduced the gate leakage by six orders of magnitude, maintained an on/off ratio of 106, and improved the gate swing from 3 V to 13 V compared to the non-interlayer FET (Figure 18b).
As an aside, FETs using reduced surface electric field (RESURF) and superjunction (SJ) techniques incorporating p-NiO have shown to increase breakdown, although more work is required to improve their BFOMs [236,237].

p-GaN

P-GaN is another material that can be used as a p-gate to β-Ga2O3; however, it has only been studied in TCAD simulations. Similar to p-NiO, it is primarily to improve Ron without sacrificing Vbr, as well as to enable normally off operation by depleting the channel underneath. Increasing the doping and/or thickness of the p-GaN layer increases the Vth, as more charge is available to deplete the channel [194,238]. Increased p-GaN doping is shown to have little effect on gm, while an increase in p-GaN thickness reduces gm due to reduced gate control. In a vertical FinFET, the addition of p-GaN as the gate metal also improves Vbr when compared to Schottky gate metals such as Ni/Au because of the higher work function of GaN, leading to a vertical electron barrier of ≈5 eV for p-GaN gates compared to ≈2.5 eV for a Ni/Au gate metal (Figure 18c). Additionally, the Vbr of the p-GaN-gate FinFET is more resistant to Wfin increases compared to the Ni/Au-gate FinFET [239].
Figure 18. (a) P-NiO-gated HJ-FET with BFOM of 0.74 GW cm−2, ultra-low SS, and negligible hysteresis due to piranha treatment. Reproduced from [75], with the permission of AIP Publishing. (b) Addition of SiO2 between gate metal and p-NiO increases pn turn-on and enables larger gate swing. © (2023) IEEE. Reprinted with permission from [76]. (c) EC electron barrier of FinFET with conventional (Con., blue) Ni/Au gate contact compared with proposed (pro., red) device using p-GaN as the gate metal, simulated through TCAD. Reproduced from [239]. © IOP Publishing. Reproduced with permission. All rights reserved. (d) HJ-FET with p-SnO-gate dielectric grown via PAMBE. Reproduced from [77]. CC BY 4.0.
Figure 18. (a) P-NiO-gated HJ-FET with BFOM of 0.74 GW cm−2, ultra-low SS, and negligible hysteresis due to piranha treatment. Reproduced from [75], with the permission of AIP Publishing. (b) Addition of SiO2 between gate metal and p-NiO increases pn turn-on and enables larger gate swing. © (2023) IEEE. Reprinted with permission from [76]. (c) EC electron barrier of FinFET with conventional (Con., blue) Ni/Au gate contact compared with proposed (pro., red) device using p-GaN as the gate metal, simulated through TCAD. Reproduced from [239]. © IOP Publishing. Reproduced with permission. All rights reserved. (d) HJ-FET with p-SnO-gate dielectric grown via PAMBE. Reproduced from [77]. CC BY 4.0.
Materials 16 07693 g018

p-SnO

MBE-grown p-SnO is reported to have a low bandgap of 0.7 eV and a type-I band alignment to β-Ga2O3, with CBO and VBO values of 0.49 eV and 3.70 eV, respectively. However, HJ-FETs have reported breakdown electric fields of ≈2 MV cm−1, possibly due to the high-quality MBE-grown SnO (Figure 18d) [77,240]. Another report of p-SnO was in an SOI FET using sputtered p-SnO as a back depletion layer, but not as a gate. This shifted the Vth to +40 V, and resulted in record high mobilities of 191 cm2 V−1 s−1 (Figure 12b [85]).

3.4.3. High-k Gate Dielectrics

High-k dielectrics are a heavily studied field in β-Ga2O3 devices, particularly for their incredible field-spreading capability [241]. HfO2, BaTiO3 (BTO), and SrTiO3 (STO) are the most used high-k dielectrics, where BTO and STO are considered extreme-k dielectrics because their dielectric constants can be on the order of 300. Early reports of high-k FETs used ALD-HfO2, but both FETs suffered from high interface trap densities [86,242]. A later report of SOI FETs with ALD-HfO2 gate dielectrics reported near-ideal behavior, with negligible hysteresis, an SS of 64 mV dec−1, and an on/off ratio of 108. This result was attributed to the high-temperature HfO2 deposition of 350 °C, forming a high-quality polycrystalline layer (Figure 12a [152]).
Xia et al. discuss the ability of extreme-k dielectrics in reducing the peak electric field and premature breakdown due tunneling by increasing the barrier length in heterojunction diodes [243]. Kalarickal et al. developed an electrostatic model of extreme-k dielectrics in FETs using the polarization of the charge to create highly uniform electric field profiles in the channel and enable both a high sheet charge density and an average breakdown field (Figure 19a). The modeled FET was then fabricated to show the improvement in the average breakdown field up to 4 MV cm−1 and a high sheet charge density of 1.6 × 1013 cm−2 using extreme-k dielectric BTO with a dielectric constant of 235 [78]. Kalarickal et al. improved on their FET design by first adding a 12.5 nm low-k dielectric of ALD-Al2O3 on the epitaxial layer to reduce interface traps and protect the surface during the sputtering of extreme-k BTO, resulting in an average breakdown field of 5.5 MV cm−1 and a BFOM of 408 MW cm−2 (Figure 19b [79]).

3.4.4. Multi-Stack Gate Dielectrics

As previously mentioned, multi-stack gates can provide the advantages of both materials, such as in ref. [79], with a low-k Al2O3/extreme-k BTO multi-stack gate dielectric. Comparisons between the Al2O3/HfO2 and HfO2/Al2O3 gate stacks indicates reduced gate leakage in the HfO2/Al2O3/β-Ga2O3 because of the better carrier-blocking capability of the Al2O3. Both stacks show increased dielectric constants and similar Dit [244]. Another study comparing polycrystalline (p-)/amorphous (a-) HfO2 with p-HfO2/a-Al2O3 gate stacks showed that the p-HfO2/a-HfO2 stack had a lower Dit, as evidenced by a distorted energy band in p-HfO2/a-Al2O3, a larger effective barrier height of 1.62 eV, and a hard breakdown field of 9.1 MV cm−1 compared to 4.9 MV cm−1 of the p-HfO2/a-Al2O3 stack (Figure 19c). The HfO2 bilayer stack also outperforms the single-layer p-HfO2 in breakdown because of better leakage suppression [245]. Similarly, the higher bandgap of SiO2 reduced gate leakage by 800× and increased the breakdown electric field by 1.7× when added as an interlayer between Al2O3/β-Ga2O3 [230]. Multi-gate stacks have also been used for ferroelectric charge storage via Al2O3/Hf0.5Zr0.5O2 (Al2O3/HZO) and Al2O3/HfO2/Al2O3/HZO, where the HZO polarization traps charge at the Al2O3/HZO interface for the first stack and in the HfO2 in the second stack (Figure 19d) [87,246].
Figure 19. (a) The charge profile after applying a negative gate bias on a 2DEG with an extreme-k BTO gate dielectric. © (2021) IEEE. Reprinted with permission from [78]. (b) Double-heterostructure AlGO/GO MODFET with low-k Al2O3/extreme-k BTO gate dielectric reaching average breakdown electric fields of 5.5 MV cm−1. © (2021) IEEE. Reprinted with permission from [79]. (c) Multi-stack gate dielectrics of p-HfO2/a-HfO2 with superior breakdown compared to the p-HfO2/a-Al2O3 stack. © (2021) IEEE. Reprinted with permission from [245]. (d) Ferroelectric charge storage due to multi-stack gates and polarization trapping by Hf0.5Zr0.5O2. Reproduced from [87], with the permission of AIP Publishing.
Figure 19. (a) The charge profile after applying a negative gate bias on a 2DEG with an extreme-k BTO gate dielectric. © (2021) IEEE. Reprinted with permission from [78]. (b) Double-heterostructure AlGO/GO MODFET with low-k Al2O3/extreme-k BTO gate dielectric reaching average breakdown electric fields of 5.5 MV cm−1. © (2021) IEEE. Reprinted with permission from [79]. (c) Multi-stack gate dielectrics of p-HfO2/a-HfO2 with superior breakdown compared to the p-HfO2/a-Al2O3 stack. © (2021) IEEE. Reprinted with permission from [245]. (d) Ferroelectric charge storage due to multi-stack gates and polarization trapping by Hf0.5Zr0.5O2. Reproduced from [87], with the permission of AIP Publishing.
Materials 16 07693 g019

4. Defect Engineering

4.1. Defects

Defects can severely degrade device performance and reliability and are thus an important area of study for any semiconductor device. WBG and UWBG semiconductors such as β-Ga2O3 require somewhat different characterization methods to study deep traps, since room-temperature and high-temperature measurements still only observe a small portion of the bandgap, and therefore photon excitation is typically used to obtain trap information throughout the bandgap. The following discusses many of the various trap characterization methods for β-Ga2O3, as well as the material preparations used to reduce traps in β-Ga2O3.

4.1.1. Characterization

Deep-level transient spectroscopy (DLTS) and deep-level optical spectroscopy (DLOS) are powerful techniques to determine the energies and concentrations of deep-level traps. DLTS and DLOS are based on the principle that the capture and emission of carriers from traps in the space charge region (SCR) varies the measured capacitance. Therefore, capacitance transients are typically used to determine trap energy levels and their capture and emission rates at a specific energy level, determined by the temperature in DLTS and photon energy in DLOS. DLOS is required for (U)WBG semiconductors because most DLTS systems are limited to ≈1 eV below EC or above EV, which is insufficient to fully characterize (U)WBG materials. The traps found in β-Ga2O3 via DLTS and DLOS are shown in Figure 20a. More details on the DLTS/DLOS principle and deep-level defects in β-Ga2O3 are reported in ref. [247].
While DLTS/DLOS mainly characterize bulk traps, photo-assisted C-V (PCV) can be used for extracting deep-level interface and dielectric bulk traps at β-Ga2O3/dielectric interfaces. Two main PCV methods have been reported, where the first uses above-bandgap light and compares ΔV the dark and UV C-V curves vs. surface potential to calculate Dt, which is the sum of the interface trap state density (Dit) and the dielectric bulk trap density (nbul). An average Dit is found from the y-intercept of Dt vs. tox (Figure 20b). Note that the dark CV curve is measured from accumulation to depletion after being held in accumulation for 10 min for all interface traps to fill. In depletion, the device is exposed to UV light to excite electrons from all interface traps, and held in depletion for 10 min in the dark after UV exposure so that the generated holes move to the Al2O3/β-Ga2O3 interface [248]. The second method uses at least two sub-bandgap light sources to empty interface traps at two energies below EC. The resulting flatband voltage shift (ΔVfb) determines the Dit. The second method has the advantage that no e–h pairs are generated, which can be a source of error, and a more precise Dit can be found as opposed to an average value [249].
Figure 20. (a) Bulk trap levels found via DLTS and DLOS. Reproduced from [247]. © IOP Publishing. Reproduced with permission. All rights reserved. (b) PCV method using above-gap light for Dt and extrapolation to obtain an average Dit. Reproduced from [248], with the permission of AIP Publishing. (c) Photo I-V to obtain donor and acceptor interface trap state densities. © (2018) IEEE. Reprinted with permission from [250]. (d) Stress I-V and trapped charge from stress C-V. Reproduced from [251], with the permission of AIP Publishing. (e) Pulsed I-V showing current dispersion (left) without passivation and significant improvement after SiNx passivation (right). (f) Fitted I-V transient to variable range-hopping mechanism for time constant and temperature-dependent measurements for activation energy. © (2021) IEEE. Reprinted with permission from [133].
Figure 20. (a) Bulk trap levels found via DLTS and DLOS. Reproduced from [247]. © IOP Publishing. Reproduced with permission. All rights reserved. (b) PCV method using above-gap light for Dt and extrapolation to obtain an average Dit. Reproduced from [248], with the permission of AIP Publishing. (c) Photo I-V to obtain donor and acceptor interface trap state densities. © (2018) IEEE. Reprinted with permission from [250]. (d) Stress I-V and trapped charge from stress C-V. Reproduced from [251], with the permission of AIP Publishing. (e) Pulsed I-V showing current dispersion (left) without passivation and significant improvement after SiNx passivation (right). (f) Fitted I-V transient to variable range-hopping mechanism for time constant and temperature-dependent measurements for activation energy. © (2021) IEEE. Reprinted with permission from [133].
Materials 16 07693 g020
A photo I-V characterization method was developed to determine donor and acceptor interface trap state densities by comparing the subthreshold drain current in dark and UV conditions, attributing the range Von < VGS < Vfb due to donor traps and Vfb < VGS < Vth due to acceptor traps [250].
Stress measurements are another trapping-characterization method, where stress I-V determines the Vth instability and stress C-V quantifies the trapped charge. A study of stress measurements on a β-Ga2O3 MOSFET observed a logarithmic dependence of Vth on stress time, as well as full recovery after 365 nm of UV illumination. The stress C-V at various temperatures also indicated that the trapping followed an inhibition model, where a trapped electron inhibits neighboring charge trapping due to coulombic repulsion [251]. Monitoring Vth shifts and Ron under positive bias stress (PBS) and negative bias stress (NBS) can help to identify traps contributing to the FET instability and degradation. This has been studied in recessed-gate [252], p-NiO-gate [253], and β-Ga2O3/SiC FETs [254]. PBS-induced instability is primarily caused by border traps in the gate oxide, while NBS-induced instability is attributed to both interface states and border traps [252]. In p-NiO gated FETs, a high VGS or long stress time permanently negatively shifted the Vth by ionizing interface dipoles, which neutralized ionized charges in the depletion region [253]. The PBS of β-Ga2O3/SiC FETs fabricated via H+ ion implantation, as described in [156], positively shifted the Vth for short stress times due to electron trapping by the interface and border traps. For longer stress times, however, negative Vth shifts and an increase in carrier concentration were observed and attributed to the generation of shallow donors by H+ interstitials or the H passivation of deep-acceptor gallium vacancies [254].
Pulsed I-V, studied in nearly all material systems including SiC and GaN, is useful in that it can isolate the effects of buffer traps (in the case of drain pulsing) and surface/interface traps (in the case of gate pulsing) [255]. In some β-Ga2O3 studies, the drain lag had no DC-RF dispersion, while the gate lag exhibited significant current dispersion, indicating that surface traps near the β-Ga2O3/Al2O3 interface under the gate and in the drain–gate access region heavily impacted the RF performance, while buffer traps had minimal effects [96,133]. SiNx passivation is seen to improve current dispersion in gate-lag measurements. Additionally, the fitting of the drain current transient provides information on the trap time response and trapping/de-trapping mechanism (Shockley–Read–Hall, variable range hopping, etc.), and temperature-dependent pulsed measurements can give trap activation energies.

4.1.2. Material Preparation

It is widely known in β-Ga2O3 that there is Si contamination at the substrate/epi interface (Figure 21a), which can create a parasitic secondary channel, adding parasitic resistance and capacitance [216]. Additionally, the band bending at the substrate/epi interface depletes the channel [109], and semi-insulating impurities can diffuse into the channel region. Therefore, the growth of a buffer layer is useful in mitigating both the parasitic channels from Si contamination and channel depletion from the substrate. In delta-doped MESFETs, Fe-doped semi-insulating substrates reduced both the charge density and mobility with decreasing buffer thickness as well as increased RF dispersion (Figure 21b [103]). A secondary-ion mass spectroscopy (SIMS) depth profile showed that Fe impurities diffused nearly 200 nm into the epi layer, making the buffer thickness a critically important feature in lateral β-Ga2O3 FETs. To mitigate the effects of the second parasitic channel, a Mg delta-doping layer was grown at 420 °C using a 10 s open/30 s close/10 s open pulsing shutter scheme near the substrate/epi interface to compensate any Si impurity concentrations and reduce device leakage. An improved leakage current by six orders of magnitude and negligible hysteresis in transfer I-V curves was reported (Figure 21c [106]).
In another study, two buffer trap levels at EC—0.7 eV and EC—0.8 eV, associated with Fe-doped substrates, were observed using isothermal constant drain current DLTS (CID-DLTS) (Figure 21d). The former was observed in MESFETs with buffer layers of both 100 nm and 600 nm, while the latter was not seen in the 600 nm buffer layer MESFET, leading to the conclusion that the trap at EC—0.8 eV is correlated with Fe diffusion into the buffer layer, while the trap at EC—0.7 eV is consistent with a point defect source observed in β-Ga2O3. The RF dispersion was not significantly reduced and an increased Vth variation was observed in the MESFET with the 600 nm buffer, indicating that the EC—0.7 eV trap was dominant in Ron degradation and Vth instability [105].
Techniques to reduce defect densities and their effects on device performance include pre-dielectric deposition cleans, post-deposition (PDA) and post-metallization annealing (PMA), in situ dielectric growth, and MacEtch device fabrication. Piranha treatment has shown to reduce the surface roughness and interface trapped charge, Qit, from 1.4 × 1012 cm−2 to 3.2 × 1011 cm−2. PDA at 500 °C in O2 or N2/O2 after piranha treatment showed a highly improved interface quality and lowered the average Dit to 2.3 × 1011 cm−2 eV−1 (O2 PDA) [256]. A study of PDA and PMA temperatures on interface quality observed that PMA at 300 °C in N2 after low-temperature PDA from 300 °C to 600 °C shifted the Vfb to near-ideal values and fixed the charge to the order of 1 × 1011 cm−2; however, PMA had little effect when the PDA temperature was increased from 700 to 900 °C due to Ga and Al interdiffusion. PMA significantly reduced the shallow Dit states for all PDA temperatures (lowest for PDA at 300 °C), but had no effect on the deep Dit. The deep Dit states decreased only with increasing PDA from 300 to 900 °C, in contrast to the increasing density of shallow Dit states with increasing PDA temperature (Figure 22a [249]). Islam et al. recently reported on a solvent (S), O2 plasma, and piranha (P) surface-cleaning method, followed by BHF (B) surface etch, PE-ALD Al2O3, and in situ forming-gas PDA (FG-PDA) at 250 °C to achieve a ΔVfb of 300 mV and 80 mV for the first and second C-V cycle, respectively, as well as stable accumulation at positive VG, while other comparative samples showed large first- and/or second-cycle hysteresis and no accumulation (Figure 22b [228]). Another study on surface damage removal after ICP used post-tetramethyl ammonium hydroxide (TMAH) and self-reaction etching (SRE) using Ga flux in an MBE chamber at 900 °C, after an observed reduced and negative growth rate with increased Ga flux, reducing Dit to 7.3 × 1011 cm−2 eV−1 [257]. The SRE method removed surface damage and improved C-V characteristics (Figure 22c). SRE can be useful when performing in situ gate dielectric growth, as initially reported by another group, achieving high breakdown fields of 5.8 MV cm−1 and an average Dit of 6.4 × 1011 cm−2 eV−1 [226].
MacEtch is not quite a material preparation method, but an alternative to FinFET fabrication that avoids dry-etch-induced damage, with hysteresis of only 9.7 mV and SS of 87.2 mV dec−1 [125].
A negligible hysteresis of 4 mV, µs switching, and a near-record low SS of 66 mV dec−1 was recently achieved in recessed p-NiOx-gated FETs using only piranha surface treatment [75]. This provides a potential path for maximizing β-Ga2O3 FET performance while maintaining ultra-low interface defect densities.

5. Current Challenges and Major Strategies

5.1. Lack of p-Type Doping

The unavailability of shallow acceptors in β-Ga2O3 prevents the fabrication of homoepitaxy pn diodes, guard rings, and superjunction devices. Heterojunction devices with p-type materials such p-NiO, p-GaN, p-SnO, and p-CuO2 have been investigated with some of the highest reported BFOMs, as discussed in the previous sections. Additionally, the interface properties in p-NiO-gate FETs reveal no hysteresis, low SS, and large mobility, indicating a high-quality interface and minimal trap scattering. While p-NiO is currently the most promising of the p-type materials for heterojunctions, its deposition via sputtering and polycrystallinity can result in nonuniformity and a low yield, which needs further investigation.
Some groups have reported p-doping using amphoteric Zn and H diffusion (see Section 2.5); however, such devices have yet to be reported. The feasibility of these techniques in high-performing diodes and FETs need to be verified for these techniques to be readily adopted in the high-power and RF markets.

5.2. Low Thermal Conductivity

The low thermal conductivity is a major issue in β-Ga2O3 devices, especially for high-power applications where self-heating is unavoidable. Various thermal studies using high-thermally conductive substrates have been reported and are discussed in the GO/AlN, GO/SiC, and GO/Diamond sections, and illustrated in Figure 16. There seem to be promising solutions using the hetero-integration of thermally conductive substrates via SiC ion-cutting technology or via flip-chip to a diamond carrier, thermal bumps, and NCD passivation.

5.3. Monolithic and Heterogeneous Integration

So far, most β-Ga2O3 devices have been stand-alone; however, it is imperative to incorporate them within circuitry to fully realize the potential of β-Ga2O3. Monolithic integration refers to circuitry designed on the same sample, which has so far been demonstrated with an inverter with SOI D-/E-mode graphene-gated FETs. The amplifier capability of RF FETs, obtained by determining their gain, output power, and power-added-efficiency using CW power measurements, highlights the utility of these FETs in integrated circuits (Table 4). Heterogeneous integration has mainly been realized with SOI FETs on high-thermally conductive substrates.
While the mechanical exfoliation of β-Ga2O3 nanomembranes is used in most devices, it is best-suited for proof of concept, and not for large-scale production. For β-Ga2O3, two methods of wafer-to-wafer bonding have been demonstrated, including ion cutting using H+ implantation, tested on SiC and Si substrates, [156] and low-temperature fusion bonding on SiC using a SiNx interlayer [201]. Another heterogeneous integration method for improved temperature reduction is flip-chip bonding to a diamond carrier, with the downside that, currently, diamond cannot be supplied in large wafers [204].

5.4. Packaging

Packaging is essentially the next step after device-level thermal management. Experimentally verified cooling methods need to translate accordingly for large-area, packaged devices. Additionally, both the device-level and package-level thermal management must be co-designed. One limitation is that device software and package software are difficult to integrate because each simplifies the results of the other [258].

5.5. Optical Effects and Remote Switching

The large, near-direct bandgap of β-Ga2O3 creates potential for solar-blind deep-UV (DUV) photodetectors, which is an ongoing area of research in β-Ga2O3-based devices [6,259,260,261]. The absorption anisotropy due to the complex crystal structure [262], as well as the strong sub-gap absorption from the conduction band to oxygen and gallium vacancies [263], remains a challenge for DUV β-Ga2O3 photodetectors. The optical characteristics of β-Ga2O3 may have potential advantages in the remote switching of high-power RF amplifier circuits. Remote switching is a cost-effective technique that can improve switching speeds while reducing or eliminating electrical noise. This has been discussed in GaN-based systems [264,265,266,267] and could likewise apply in β-Ga2O3-based systems.

5.6. Requirements in Real Applications

In high-power applications, Ebr and Ron,sp matter less than breakdown voltage and on-current. Small-area devices with high BFOMs should mainly be used as an in-between step to fabricate an equivalent large-area device to meet specific current and voltage ratings for real applications. This can provide further insight into what device optimizations are needed if the large-area device underperforms.

6. Applications and Trends

β-Ga2O3 FETs are not expected to replace their SiC and GaN counterparts because they are already commercially available. While this might occur in the future, the current trend is that high-power β-Ga2O3 FETs, with voltage and current ratings beyond those of GaN and SiC devices, will be used for ultra-high-power applications such as electric vehicles, rails, power grids, renewable energy storage, etc. High-power RF FETs can also find use in electric vehicles, power converters, data centers, and communication applications. The emergence of β-Ga2O3 RF FETs is difficult, with their performance still being below that of GaN HEMTs and newer diamond HEMTs. However, the low-cost melt growth technique, compared to the high cost of diamond, as well as the higher theoretical vsat than GaN, both show a promising future for high-frequency β-Ga2O3 devices. High-power β-Ga2O3 FETs have shown higher breakdown fields surpassing the GaN theoretical limit, and so a potentially more accessible market for RF β-Ga2O3 FETs is mid-frequency (≈tens of GHz), high-power RF FETs that can outperform high-power GaN RF FETs.

7. Conclusions and Outlook

In conclusion, considerable advancements in β-Ga2O3 FET designs have been made to push both their high-power and RF capabilities. High-power FETs have shown breakdown voltages up to 10 kV, current densities of >1 kA cm−2 and 1.5 mA mm−1, and BFOMs of 0.95 GW cm−2. RF FETs have reported high breakdown fields of 5.4 MV cm−1, operating frequencies up to 48 GHz fmax, and saturation velocities up to 3 × 106 cm s−1. While many FETs have surpassed the theoretical FOMs of Si, they are still far from those of β-Ga2O3. From the overview of β-Ga2O3 FETs, the key takeaways are as follows:
(1)
The importance of high-quality epitaxial growth and buffer layers cannot be understated. The highest BFOM FET to date also reports one of the highest mobilities of 184 cm2 V−1 s−1, realized through MOCVD varied low/high temperature layers;
(2)
The SAG is vital to both high power and RF in that it is used to scale device geometries and reduce source–gate series resistance. It should be implemented, if possible, in both lateral and vertical FETs;
(3)
For high currents, vertical transistors are preferred because the current scales with the device area as opposed to the channel thickness, as in lateral devices. FinFETs and CAVETs show the best results, with FinFETs offering more gate control and reduced leakage, but increased complexity. MacEtch FinFETs are a non-dry-etching alternative;
(4)
Normally off (E-mode) FETs are crucial for power electronics because of their reduced off-state power loss, fail-safe high-voltage operation, and simplified circuitry for power switching. The lack of p-type doping, and therefore inversion, in β-Ga2O3 requires approaches such as recessed gates (Section 3.1.3), low-doped channels and CBLs (Section 3.2.1), small-width FinFETs (Section 3.1.4), oxygen annealing (Section 3.2.2), and p-gate materials for normally off operation (Section 3.4.2);
(5)
FP structures (GFP, SFP) including T-gates are vital to any high-power device. High-k or extreme-k FP dielectrics are an attractive option to improve breakdown;
(6)
SOI FETs can be very useful in conducting studies on thermal, transport, novel gate dielectrics, etc. However, they are limited in their breakdown voltage and small sample size. SOI FETs should be considered as a proof of concept with the intent to apply successful designs into bulk devices;
(7)
Novel structures simulated through TCAD, such as vertical trench gates, GAA, air-gap FPs, HBTs, and others, should be used to evaluate the potential of a design before fabrication;
(8)
RF FETs have been realized in delta-doped MESFETs, AlGO/GO MODFETs, and HFETs, forming a 2DEG with Si-doped AlGO/UID-GO;
(9)
One commonality of RF FETs is their T-gate structure, allowing highly scaled LG while maintaining low noise figures;
(10)
RF FETs have reported high operating frequencies at ≥27 GHz with and without FP dielectrics;
(11)
Ohmic contacts should always utilize some improvements, such as regrowth, ion implantation, or interlayers;
(12)
P-NiO-gate dielectrics show promise in increasing the BFOM, while maintaining high-quality/low-defect density interfaces. A high-bandgap dielectric should be added to increase the gate swing beyond the pn turn-on voltage;
(13)
Thermal management is crucial, and the intention to use wafer-bonding techniques or flip-chips with high-thermally conductive substrates must be implemented to further enhance device performance;
(14)
For high-power applications, an appealing FET design with high FOM(s) should be fabricated as a large-area device to meet current and breakdown voltage ratings.
Defect characterization is vitally important in β-Ga2O3, and techniques need to be adapted or invented for UWBG materials. Material preparation is essential to improve peak performance and must be considered in each step of fabrication.
Device-level and package-level thermal management and modeling is critical in taking β-Ga2O3 devices to the market. Thermally conductive substrates have experimentally shown significant drops in peak temperatures, and modeling has indicated that wafer bonding via flip-chip and junction-side cooling can reduce thermal effects. The device and package must be designed and optimized simultaneously; however, co-design modeling is still limited.
Overall, the rapid progress in material quality, fabrication, defect characterization and mitigation, and thermal management indicates tremendous potential for β-Ga2O3 devices to quickly enter the power electronics application space once the presented challenges are addressed.

Author Contributions

Conceptualization, O.M. and Q.L.; methodology, O.M.; software, NA; validation, NA; formal analysis, NA; investigation, O.M.; resources, NA; data curation, NA; writing—original draft preparation, O.M.; writing—review and editing, Q.L.; visualization, O.M.; supervision, Q.L.; project administration, Q.L.; funding acquisition, Q.L. All authors have read and agreed to the published version of the manuscript.

Funding

O.M. is supported through the GMU Presidential Scholarship Award. This work is also supported by the Virginia Microelectronics Consortium (VMEC) chair professorship and partially supported by NIST research grant 70NANB23H093.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Technavio. Wide-Bandgap (WBG) Power Semiconductor Devices Market Analysis 2026. Available online: https://www.technavio.com/report/wide-bandgap-wbg-power-semiconductor-devices-market-industry-analysis (accessed on 15 October 2023).
  2. Reese, S.B.; Remo, T.; Green, J.; Zakutayev, A. How Much Will Gallium Oxide Power Electronics Cost? Joule 2019, 3, 903–907. [Google Scholar] [CrossRef]
  3. Millan, J.; Godignon, P.; Perpina, X.; Perez-Tomas, A.; Rebollo, J. A Survey of Wide Bandgap Power Semiconductor Devices. IEEE Trans. Power Electron. 2014, 29, 2155–2163. [Google Scholar] [CrossRef]
  4. Chow, T.; Tyagi, R. Wide bandgap compound semiconductors for superior high-voltage unipolar power devices. IEEE Trans. Electron Devices 1994, 41, 1481–1483. [Google Scholar] [CrossRef]
  5. Pearton, S.J.; Ren, F.; Tadjer, M.; Kim, J. Perspective: Ga2O3 for ultra-high power rectifiers and MOSFETS. J. Appl. Phys. 2018, 124, 220901. [Google Scholar] [CrossRef]
  6. 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]
  7. Yadava, N.; Chauhan, R.K. Review—Recent Advances in Designing Gallium Oxide MOSFET for RF Application. ECS J. Solid State Sci. Technol. 2020, 9, 065010. [Google Scholar] [CrossRef]
  8. Reese, S.B.; Zakutayev, A. Gallium oxide techno-economic analysis for the wide bandgap semiconductor market. In Oxide-Based Materials and Devices XI; SPIE: Bellingham, WA, USA, 2020; pp. 6–10. [Google Scholar] [CrossRef]
  9. Gupta, C.; Pasayat, S.S. Vertical GaN and Vertical Ga2O3 Power Transistors: Status and Challenges. Phys. Status Solidi (Appl. Mater. Sci.) 2022, 219, 2100659. [Google Scholar] [CrossRef]
  10. Dong, H.; Xue, H.; He, Q.; Qin, Y.; Jian, G.; Long, S.; Liu, M. Progress of power field effect transistor based on ultra-wide bandgap Ga2O3 semiconductor material. J. Semicond. 2019, 40, 011802. [Google Scholar] [CrossRef]
  11. Li, B.; Zhang, X.; Zhang, L.; Ma, Y.; Tang, W.; Chen, T.; Hu, Y.; Zhou, X.; Bian, C.; Zeng, C.; et al. A comprehensive review of recent progress on enhancement-mode β-Ga2O3 FETs: Growth, devices and properties. J. Semicond. 2023, 44, 061801. [Google Scholar] [CrossRef]
  12. Liu, A.-C.; Hsieh, C.-H.; Langpoklakpam, C.; Singh, K.J.; Lee, W.-C.; Hsiao, Y.-K.; Horng, R.-H.; Kuo, H.-C.; Tu, C.-C. State-of-the-Art β-Ga2O3 Field-Effect Transistors for Power Electronics. ACS Omega 2022, 7, 36070–36091. [Google Scholar] [CrossRef]
  13. Qiao, R.; Zhang, H.; Zhao, S.; Yuan, L.; Jia, R.; Peng, B.; Zhang, Y. A state-of-art review on gallium oxide field-effect transistors. J. Phys. D: Appl. Phys. 2022, 55, 383003. [Google Scholar] [CrossRef]
  14. Roy, R.; Hill, V.G.; Osborn, E.F. Polymorphism of Ga2O3 and the System Ga2O3–H2O. J. Am. Chem. Soc. 1952, 74, 719–722. [Google Scholar] [CrossRef]
  15. Yoshioka, S.; Hayashi, H.; Kuwabara, A.; Oba, F.; Matsunaga, K.; Tanaka, I. Structures and energetics of Ga2O3 polymorphs. J. Phys. Condens. Matter 2007, 19, 346211. [Google Scholar] [CrossRef]
  16. Playford, H.Y.; Hannon, A.C.; Barney, E.R.; Walton, R.I. Structures of Uncharacterised Polymorphs of Gallium Oxide from Total Neutron Diffraction. Chem. Eur. J. 2013, 19, 2803–2813. [Google Scholar] [CrossRef] [PubMed]
  17. Tak, B.R.; Kumar, S.; Kapoor, A.K.; Wang, D.; Li, X.; Sun, H.; Singh, R. Recent advances in the growth of gallium oxide thin films employing various growth techniques—A review. J. Phys. D: Appl. Phys. 2021, 54, 453002. [Google Scholar] [CrossRef]
  18. Geller, S. Crystal Structure of β-Ga2O3. J. Chem. Phys. 1960, 33, 676–684. [Google Scholar] [CrossRef]
  19. Mock, A.; Korlacki, R.; Briley, C.; Darakchieva, V.; Monemar, B.; Kumagai, Y.; Goto, K.; Higashiwaki, M.; Schubert, M. Band-to-band transitions, selection rules, effective mass, and excitonic contributions in monoclinic β−Ga2O3. Phys. Rev. B 2017, 96, 245205. [Google Scholar] [CrossRef]
  20. Åhman, J.; Svensson, G.; Albertsson, J. A Reinvestigation of β-Gallium Oxide. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1996, 52, 1336–1338. [Google Scholar] [CrossRef]
  21. Ueda, N.; Hosono, H.; Waseda, R.; Kawazoe, H. Anisotropy of electrical and optical properties in β-Ga2O3 single crystals. Appl. Phys. Lett. 1997, 71, 933–935. [Google Scholar] [CrossRef]
  22. Higashiwaki, M.; Sasaki, K.; Murakami, H.; Kumagai, Y.; Koukitu, A.; Kuramata, A.; Masui, T.; Yamakoshi, S. Recent progress in Ga2O3 power devices. Semicond. Sci. Technol. 2016, 31, 034001. [Google Scholar] [CrossRef]
  23. Varley, J.B.; Janotti, A.; Franchini, C.; Van de Walle, C.G. Role of self-trapping in luminescence andp-type conductivity of wide-band-gap oxides. Phys. Rev. B 2012, 85, 081109. [Google Scholar] [CrossRef]
  24. Tippins, H.H. Optical Absorption and Photoconductivity in the Band Edge of β-Ga2O3. Phys. Rev. B 1965, 140, A316–A319. [Google Scholar] [CrossRef]
  25. Tsao, J.Y.; Chowdhury, S.; Hollis, M.A.; Jena, D.; Johnson, N.M.; Jones, K.A.; Kaplar, R.J.; Rajan, S.; Van de Walle, C.G.; Bellotti, E.; et al. Ultrawide-Bandgap Semiconductors: Research Opportunities and Challenges. Adv. Electron. Mater. 2018, 4, 1600501. [Google Scholar] [CrossRef]
  26. Guo, Z.; Verma, A.; Wu, X.; Sun, F.; Hickman, A.; Masui, T.; Kuramata, A.; Higashiwaki, M.; Jena, D.; Luo, T. Anisotropic thermal conductivity in single crystal β-gallium oxide. Appl. Phys. Lett. 2015, 106, 111909. [Google Scholar] [CrossRef]
  27. Boteler, L.; Lelis, A.; Berman, M.; Fish, M. Thermal Conductivity of Power Semiconductors—When Does It Matter? In Proceedings of the 2019 IEEE 7th Workshop on Wide Bandgap Power Devices and Applications (WiPDA), Raleigh, NC, USA, 29–21 October 2019; pp. 265–271. [Google Scholar]
  28. Ghosh, K.; Singisetti, U. Electron mobility in monoclinic β-Ga2O3—Effect of plasmon-phonon coupling, anisotropy, and confinement. J. Mater. Res. 2017, 32, 4142–4152. [Google Scholar] [CrossRef]
  29. Kang, Y.; Krishnaswamy, K.; Peelaers, H.; Van de Walle, C.G. Fundamental limits on the electron mobility of β-Ga2O3. J. Phys. Condens. Matter 2017, 29, 234001. [Google Scholar] [CrossRef] [PubMed]
  30. Ma, N.; Tanen, N.; Verma, A.; Guo, Z.; Luo, T.; Xing, H.; Jena, D. Intrinsic electron mobility limits in β-Ga2O3. Appl. Phys. Lett. 2016, 109, 212101. [Google Scholar] [CrossRef]
  31. Chen, X.; Jagadish, C.; Ye, J. Fundamental properties and power electronic device progress of gallium oxide. In Oxide Electronics; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2021; pp. 235–352. [Google Scholar] [CrossRef]
  32. Kuramata, A.; Koshi, K.; Watanabe, S.; Yamaoka, Y.; Masui, T.; Yamakoshi, S. High-quality β-Ga2O3 single crystals grown by edge-defined film-fed growth. Jpn. J. Appl. Phys. 2016, 55, 1202A2. [Google Scholar] [CrossRef]
  33. Aida, H.; Nishiguchi, K.; Takeda, H.; Aota, N.; Sunakawa, K.; Yaguchi, Y. Growth of β-Ga2O3 Single Crystals by the Edge-Defined, Film Fed Growth Method. Jpn. J. Appl. Phys. 2008, 47, 8506–8509. [Google Scholar] [CrossRef]
  34. Tomm, Y.; Reiche, P.; Klimm, D.; Fukuda, T. Czochralski grown Ga2O3 crystals. J. Cryst. Growth 2000, 220, 510–514. [Google Scholar] [CrossRef]
  35. Hoshikawa, K.; Kobayashi, T.; Matsuki, Y.; Ohba, E. 2-inch diameter (1 0 0) β-Ga2O3 crystal growth by the vertical Bridgman technique in a resistance heating furnace in ambient air. J. Cryst. Growth 2020, 545, 125724. [Google Scholar] [CrossRef]
  36. Hoshikawa, K.; Ohba, E.; Kobayashi, T.; Yanagisawa, J.; Miyagawa, C.; Nakamura, Y. Growth of β-Ga2O3 single crystals using vertical Bridgman method in ambient air. J. Cryst. Growth 2016, 447, 36–41. [Google Scholar] [CrossRef]
  37. Ohira, S.; Yoshioka, M.; Sugawara, T.; Nakajima, K.; Shishido, T. Fabrication of hexagonal GaN on the surface of β-Ga2O3 single crystal by nitridation with NH3. Thin Solid Films 2006, 496, 53–57. [Google Scholar] [CrossRef]
  38. Víllora, E.G.; Shimamura, K.; Yoshikawa, Y.; Aoki, K.; Ichinose, N. Large-size β-Ga2O3 single crystals and wafers. J. Cryst. Growth 2004, 270, 420–426. [Google Scholar] [CrossRef]
  39. Chase, A.O. Growth of beta-Ga2, O3 by the Verneuil Technique. J. Am. Ceram. Soc. 1964, 47, 470. [Google Scholar] [CrossRef]
  40. Lorenz, M.; Woods, J.; Gambino, R. Some electrical properties of the semiconductor β-Ga2O3. J. Phys. Chem. Solids 1967, 28, 403–404. [Google Scholar] [CrossRef]
  41. Chi, Z.; Asher, J.J.; Jennings, M.R.; Chikoidze, E.; Pérez-Tomás, A. Ga2O3 and Related Ultra-Wide Bandgap Power Semiconductor Oxides: New Energy Electronics Solutions for CO2 Emission Mitigation. Materials 2022, 15, 1164. [Google Scholar] [CrossRef]
  42. Tsai, M.-Y.; Bierwagen, O.; White, M.E.; Speck, J.S. β-Ga2O3 growth by plasma-assisted molecular beam epitaxy. J. Vac. Sci. Technol. A 2010, 28, 354–359. [Google Scholar] [CrossRef]
  43. Hadamek, T.; Posadas, A.B.; Al-Quaiti, F.; Smith, D.J.; McCartney, M.R.; Demkov, A.A. β-Ga2O3 on Si (001) grown by plasma-assisted MBE with γ-Al2O3 (111) buffer layer: Structural characterization. AIP Adv. 2021, 11, 045209. [Google Scholar] [CrossRef]
  44. Kalarickal, N.K.; Xia, Z.; McGlone, J.; Krishnamoorthy, S.; Moore, W.; Brenner, M.; Arehart, A.R.; Ringel, S.A.; Rajan, S. Mechanism of Si doping in plasma assisted MBE growth of β-Ga2O3. Appl. Phys. Lett. 2019, 115, 152106. [Google Scholar] [CrossRef]
  45. Oshima, Y.; Vίllora, E.G.; Shimamura, K. Quasi-heteroepitaxial growth of β-Ga2O3 on off-angled sapphire (0 0 0 1) substrates by halide vapor phase epitaxy. J. Cryst. Growth 2015, 410, 53–58. [Google Scholar] [CrossRef]
  46. Ranga, P.; Bhattacharyya, A.; Whittaker-Brooks, L.; Scarpulla, M.A.; Krishnamoorthy, S. N-type doping of low-pressure chemical vapor deposition grown β-Ga2O3 thin films using solid-source germanium. J. Vac. Sci. Technol. A 2021, 39, 030404. [Google Scholar] [CrossRef]
  47. Rafique, S.; Karim, R.; Johnson, J.M.; Hwang, J.; Zhao, H. LPCVD homoepitaxy of Si doped β-Ga2O3 thin films on (010) and (001) substrates. Appl. Phys. Lett. 2018, 112, 052104. [Google Scholar] [CrossRef]
  48. Rafique, S.; Han, L.; Zhao, H. (Invited) Ultrawide Bandgap β-Ga2O3 Thin Films: Growths, Properties and Devices. ECS Trans. 2017, 80, 203–216. [Google Scholar] [CrossRef]
  49. Zhang, J.; Shi, J.; Qi, D.-C.; Chen, L.; Zhang, K.H.L. Recent progress on the electronic structure, defect, and doping properties of Ga2O3. APL Mater. 2020, 8, 020906. [Google Scholar] [CrossRef]
  50. Peelaers, H.; Lyons, J.L.; Varley, J.B.; Van de Walle, C.G. Deep acceptors and their diffusion in Ga2O3. APL Mater. 2019, 7, 022519. [Google Scholar] [CrossRef]
  51. Dong, L.; Jia, R.; Xin, B.; Peng, B.; Zhang, Y. Effects of oxygen vacancies on the structural and optical properties of β-Ga2O3. Sci. Rep. 2017, 7, 40160. [Google Scholar] [CrossRef]
  52. Varley, J.B.; Weber, J.R.; Janotti, A.; Van de Walle, C.G. Oxygen vacancies and donor impurities in β-Ga2O3. Appl. Phys. Lett. 2010, 97, 142106. [Google Scholar] [CrossRef]
  53. Neal, A.T.; Mou, S.; Rafique, S.; Zhao, H.; Ahmadi, E.; Speck, J.S.; Stevens, K.T.; Blevins, J.D.; Thomson, D.B.; Moser, N.; et al. Donors and deep acceptors in β-Ga2O3. Appl. Phys. Lett. 2018, 113, 062101. [Google Scholar] [CrossRef]
  54. Wang, Y.; Su, J.; Lin, Z.; Zhang, J.; Chang, J.; Hao, Y. Recent progress on the effects of impurities and defects on the properties of Ga2O3. J. Mater. Chem. C 2022, 10, 13395–13436. [Google Scholar] [CrossRef]
  55. Ingebrigtsen, M.E.; Varley, J.B.; Kuznetsov, A.Y.; Svensson, B.G.; Alfieri, G.; Mihaila, A.; Badstübner, U.; Vines, L. Iron and intrinsic deep level states in Ga2O3. Appl. Phys. Lett. 2018, 112, 042104. [Google Scholar] [CrossRef]
  56. Chikoidze, E.; Fellous, A.; Perez-Tomas, A.; Sauthier, G.; Tchelidze, T.; Ton-That, C.; Huynh, T.T.; Phillips, M.; Russell, S.; Jennings, M.; et al. P-type β-gallium oxide: A new perspective for power and optoelectronic devices. Mater. Today Phys. 2017, 3, 118–126. [Google Scholar] [CrossRef]
  57. Chikoidze, E.; Tchelidze, T.; Sartel, C.; Chi, Z.; Kabouche, R.; Madaci, I.; Rubio, C.; Mohamed, H.; Sallet, V.; Medjdoub, F.; et al. Ultra-high critical electric field of 13.2 MV/cm for Zn-doped p-type β-Ga2O3. Mater. Today Phys. 2020, 15, 100263. [Google Scholar] [CrossRef]
  58. Islam, M.; Liedke, M.O.; Winarski, D.; Butterling, M.; Wagner, A.; Hosemann, P.; Wang, Y.; Uberuaga, B.; Selim, F.A. Chemical manipulation of hydrogen induced high p-type and n-type conductivity in Ga2O3. Sci. Rep. 2020, 10, 6134. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, H.; Wang, Y.; Lv, Y.; Han, S.; Han, T.; Dun, S.; Guo, H.; Bu, A.; Feng, Z. 10-kV Lateral β-Ga2O3 MESFETs with B Ion Implanted Planar Isolation. IEEE Electron Device Lett. 2023, 44, 1048–1051. [Google Scholar] [CrossRef]
  60. Joishi, C.; Xia, Z.; Jamison, J.S.; Sohel, S.H.; Myers, R.C.; Lodha, S.; Rajan, S. Deep-Recessed β-Ga2O3 Delta-Doped Field-Effect Transistors with In Situ Epitaxial Passivation. IEEE Trans. Electron. Devices 2020, 67, 4813–4819. [Google Scholar] [CrossRef]
  61. Kalarickal, N.K.; Dheenan, A.; McGlone, J.F.; Dhara, S.; Brenner, M.; Ringel, S.A.; Rajan, S. Demonstration of self-aligned β-Ga2O3 δ-doped MOSFETs with current density >550 mA/mm. Appl. Phys. Lett. 2023, 122, 113506. [Google Scholar] [CrossRef]
  62. Tetzner, K.; Treidel, E.B.; Hilt, O.; Popp, A.; Bin Anooz, S.; Wagner, G.; Thies, A.; Ickert, K.; Gargouri, H.; Wurfl, J. Lateral 1.8 kV β-Ga2O3 MOSFET with 155 MW/cm2 Power Figure of Merit. IEEE Electron. Device Lett. 2019, 40, 1503–1506. [Google Scholar] [CrossRef]
  63. Bhattacharyya, A.; Roy, S.; Ranga, P.; Peterson, C.; Krishnamoorthy, S. High-Mobility Tri-Gate β-Ga2O3 MESFETs with a Power Figure of Merit Over 0.9 GW/cm2. IEEE Electron. Device Lett. 2022, 43, 1637–1640. [Google Scholar] [CrossRef]
  64. Sharma, S.; Meng, L.; Bhuiyan, A.F.M.A.U.; Feng, Z.; Eason, D.; Zhao, H.; Singisetti, U. Vacuum Annealed β-Ga2O3 Recess Channel MOSFETs with 8.56 kV Breakdown Voltage. IEEE Electron. Device Lett. 2022, 43, 2029–2032. [Google Scholar] [CrossRef]
  65. Lv, Y.; Liu, H.; Zhou, X.; Wang, Y.; Song, X.; Cai, Y.; Yan, Q.; Wang, C.; Liang, S.; Zhang, J.; et al. Lateral β-Ga2O3 MOSFETs with High Power Figure of Merit of 277 MW/cm2. IEEE Electron. Device Lett. 2020, 41, 537–540. [Google Scholar] [CrossRef]
  66. Dryden, D.M.; Liddy, K.J.; Islam, A.E.; Williams, J.C.; Walker, D.E.; Hendricks, N.S.; Moser, N.A.; Arias-Purdue, A.; Sepelak, N.P.; DeLello, K.; et al. Scaled T-Gate β-Ga2O3 MESFETs with 2.45 kV Breakdown and High Switching Figure of Merit. IEEE Electron. Device Lett. 2022, 43, 1307–1310. [Google Scholar] [CrossRef]
  67. Feng, Z.; Cai, Y.; Yan, G.; Hu, Z.; Dang, K.; Zhang, Y.; Lu, Z.; Cheng, H.; Lian, X.; Xu, Y.; et al. A 800 V β-Ga2O3 Metal–Oxide–Semiconductor Field-Effect Transistor with High-Power Figure of Merit of over 86.3 MW cm−2. Phys. Status Solidi A 2019, 216, 1900421. [Google Scholar] [CrossRef]
  68. Zhou, H.; Maize, K.; Qiu, G.; Shakouri, A.; Ye, P.D. β-Ga2O3 on insulator field-effect transistors with drain currents exceeding 1.5 A/mm and their self-heating effect. Appl. Phys. Lett. 2017, 111, 092102. [Google Scholar] [CrossRef]
  69. Wong, M.H.; Goto, K.; Murakami, H.; Kumagai, Y.; Higashiwaki, M. Current Aperture Vertical β-Ga2O3 MOSFETs Fabricated by N- and Si-Ion Implantation Doping. IEEE Electron. Device Lett. 2019, 40, 431–434. [Google Scholar] [CrossRef]
  70. Joishi, C.; Zhang, Y.; Xia, Z.; Sun, W.; Arehart, A.R.; Ringel, S.; Lodha, S.; Rajan, S. Breakdown Characteristics of β-(Al0.22Ga0.78)2O3/Ga2O3 Field-Plated Modulation-Doped Field-Effect Transistors. IEEE Electron. Device Lett. 2019, 40, 1241–1244. [Google Scholar] [CrossRef]
  71. Lei, D.; Han, K.; Wu, Y.; Liu, Z.; Gong, X. High Performance Ga2O3 Metal-Oxide-Semiconductor Field-Effect Transistors on an AlN/Si Substrate. IEEE J. Electron. Devices Soc. 2019, 7, 596–600. [Google Scholar] [CrossRef]
  72. Song, Y.; Bhattacharyya, A.; Karim, A.; Shoemaker, D.; Huang, H.-L.; Roy, S.; McGray, C.; Leach, J.H.; Hwang, J.; Krishnamoorthy, S.; et al. Ultra-Wide Band Gap Ga2O3-on-SiC MOSFETs. ACS Appl. Mater. Interfaces 2023, 15, 7137–7147. [Google Scholar] [CrossRef]
  73. Noh, J.; Ye, P.D.; Alajlouni, S.; Tadjer, M.J.; Culbertson, J.C.; Bae, H.; Si, M.; Zhou, H.; Bermel, P.A.; Shakouri, A. High Performance β-Ga2O3 Nano-Membrane Field Effect Transistors on a High Thermal Conductivity Diamond Substrate. IEEE J. Electron. Devices Soc. 2019, 7, 914–918. [Google Scholar] [CrossRef]
  74. Wang, C.; Gong, H.; Lei, W.; Cai, Y.; Hu, Z.; Xu, S.; Liu, Z.; Feng, Q.; Zhou, H.; Ye, J.; et al. Demonstration of the p-NiOx/n-Ga2O3 Heterojunction Gate FETs and Diodes with BV2/Ron,sp Figures of Merit of 0.39 GW/cm2 and 1.38 GW/cm2. IEEE Electron. Device Lett. 2021, 42, 485–488. [Google Scholar] [CrossRef]
  75. Wang, C.; Zhou, H.; Zhang, J.; Mu, W.; Wei, J.; Jia, Z.; Zheng, X.; Luo, X.; Tao, X.; Hao, Y. Hysteresis-free and μs-switching of D/E-modes Ga2O3 hetero-junction FETs with the BV2/Ron,sp of 0.74/0.28 GW/cm2. Appl. Phys. Lett. 2022, 120, 112101. [Google Scholar] [CrossRef]
  76. Wang, C.; Yan, Q.; Su, C.; Alghamdi, S.; Ghandourah, E.; Liu, Z.; Feng, X.; Zhang, W.; Dang, K.; Wang, Y.; et al. Demonstration of the β-Ga2O3 MOS-JFETs with Suppressed Gate Leakage Current and Large Gate Swing. IEEE Electron. Device Lett. 2023, 44, 380–383. [Google Scholar] [CrossRef]
  77. Tetzner, K.; Egbo, K.; Klupsch, M.; Unger, R.-S.; Popp, A.; Chou, T.-S.; Bin Anooz, S.; Galazka, Z.; Trampert, A.; Bierwagen, O.; et al. SnO/β-Ga2O3 heterojunction field-effect transistors and vertical p–n diodes. Appl. Phys. Lett. 2022, 120, 112110. [Google Scholar] [CrossRef]
  78. Kalarickal, N.K.; Feng, Z.; Bhuiyan, A.F.M.A.U.; Xia, Z.; Moore, W.; McGlone, J.F.; Arehart, A.R.; Ringel, S.A.; Zhao, H.; Rajan, S. Electrostatic Engineering Using Extreme Permittivity Materials for Ultra-Wide Bandgap Semiconductor Transistors. IEEE Trans. Electron. Devices 2021, 68, 29–35. [Google Scholar] [CrossRef]
  79. Kalarickal, N.K.; Xia, Z.; Huang, H.-L.; Moore, W.; Liu, Y.; Brenner, M.; Hwang, J.; Rajan, S. β-(Al0.18Ga0.82)2O3/Ga2O3 Double Heterojunction Transistor with Average Field of 5.5 MV/cm. IEEE Electron. Device Lett. 2021, 42, 899–902. [Google Scholar] [CrossRef]
  80. Chabak, K.D.; McCandless, J.P.; Moser, N.A.; Green, A.J.; Mahalingam, K.; Crespo, A.; Hendricks, N.; Howe, B.M.; Tetlak, S.E.; Leedy, K.; et al. Recessed-Gate Enhancement-Mode β-Ga2O3 MOSFETs. IEEE Electron. Device Lett. 2017, 39, 67–70. [Google Scholar] [CrossRef]
  81. Li, W.; Nomoto, K.; Hu, Z.; Nakamura, T.; Jena, D.; Xing, H.G. Single and multi-fin normally-off Ga2O3 vertical transistors with a breakdown voltage over 2.6 kV. In Proceedings of the 2019 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 7–11 December 2019. [Google Scholar] [CrossRef]
  82. Zhou, H.; Si, M.; Alghamdi, S.; Qiu, G.; Yang, L.; Ye, P.D. High-Performance Depletion/Enhancement-ode β-Ga2O3 on Insulator (GOOI) Field-Effect Transistors with Record Drain Currents of 600/450 mA/mm. IEEE Electron. Device Lett. 2017, 38, 103–106. [Google Scholar] [CrossRef]
  83. Zeng, K.; Soman, R.; Bian, Z.; Jeong, S.; Chowdhury, S. Vertical Ga2O3 MOSFET with Magnesium Diffused Current Blocking Layer. IEEE Electron. Device Lett. 2022, 43, 1527–1530. [Google Scholar] [CrossRef]
  84. Zhou, X.; Ma, Y.; Xu, G.; Liu, Q.; Liu, J.; He, Q.; Zhao, X.; Long, S. Enhancement-mode β-Ga2O3 U-shaped gate trench vertical MOSFET realized by oxygen annealing. Appl. Phys. Lett. 2022, 121, 223501. [Google Scholar] [CrossRef]
  85. Wang, X.; Yan, S.; Mu, W.; Jia, Z.; Zhang, J.; Xin, Q.; Tao, X.; Song, A. Enhancement-Mode Ga2O3 FET with High Mobility Using p-Type SnO Heterojunction. IEEE Electron. Device Lett. 2022, 43, 44–47. [Google Scholar] [CrossRef]
  86. Tadjer, M.J.; Mahadik, N.A.; Wheeler, V.D.; Glaser, E.R.; Ruppalt, L.; Koehler, A.D.; Hobart, K.D.; Eddy, C.R.; Kub, F.J. Editors’ Choice Communication—A (001) β-Ga2O3MOSFET with +2.9 V Threshold Voltage and HfO2Gate Dielectric. ECS J. Solid State Sci. Technol. 2016, 5, P468–P470. [Google Scholar] [CrossRef]
  87. Feng, Z.; Cai, Y.; Li, Z.; Hu, Z.; Zhang, Y.; Lu, X.; Kang, X.; Ning, J.; Zhang, C.; Feng, Q.; et al. Design and fabrication of field-plated normally off β-Ga2O3 MOSFET with laminated-ferroelectric charge storage gate for high power application. Appl. Phys. Lett. 2020, 116, 243503. [Google Scholar] [CrossRef]
  88. Xia, Z.; Xue, H.; Joishi, C.; Mcglone, J.; Kalarickal, N.K.; Sohel, S.H.; Brenner, M.; Arehart, A.; Ringel, S.; Lodha, S.; et al. β-Ga2O3 Delta-Doped Field-Effect Transistors with Current Gain Cutoff Frequency of 27 GHz. IEEE Electron. Device Lett. 2019, 40, 1052–1055. [Google Scholar] [CrossRef]
  89. Moser, N.A.; Asel, T.; Liddy, K.J.; Lindquist, M.; Miller, N.C.; Mou, S.; Neal, A.; Walker, D.E.; Tetlak, S.; Leedy, K.D.; et al. Pulsed Power Performance of β-Ga2O3 MOSFETs at L-Band. IEEE Electron. Device Lett. 2020, 41, 989–992. [Google Scholar] [CrossRef]
  90. Green, A.J.; Chabak, K.D.; Baldini, M.; Moser, N.; Gilbert, R.; Fitch, R.C.; Wagner, G.; Galazka, Z.; McCandless, J.; Crespo, A.; et al. β-Ga2O3 MOSFETs for Radio Frequency Operation. IEEE Electron. Device Lett. 2017, 38, 790–793. [Google Scholar] [CrossRef]
  91. Yu, X.; Gong, H.; Zhou, J.; Shen, Z.; Ren, F.-F.; Chen, D.; Ou, X.; Kong, Y.; Li, Z.; Chen, T.; et al. RF performance enhancement in sub-μm scaled β-Ga2O3 tri-gate FinFETs. Appl. Phys. Lett. 2022, 121, 072102. [Google Scholar] [CrossRef]
  92. Singh, M.; Casbon, M.A.; Uren, M.J.; Pomeroy, J.W.; Dalcanale, S.; Karboyan, S.; Tasker, P.J.; Wong, M.H.; Sasaki, K.; Kuramata, A.; et al. Pulsed Large Signal RF Performance of Field-Plated Ga2O3 MOSFETs. IEEE Electron. Device Lett. 2018, 39, 1572–1575. [Google Scholar] [CrossRef]
  93. Saha, C.N.; Vaidya, A.; Bhuiyan, A.F.M.A.U.; Meng, L.; Sharma, S.; Zhao, H.; Singisetti, U. Scaled β-Ga2O3 thin channel MOSFET with 5.4 MV/cm average breakdown field and near 50 GHz fMAX. Appl. Phys. Lett. 2023, 122, 182106. [Google Scholar] [CrossRef]
  94. Yu, X.; Gong, H.; Zhou, J.; Shen, Z.; Xu, W.; You, T.; Wang, J.; Zhang, S.; Wang, Y.; Zhang, K.; et al. High-Voltage β-Ga2O3 RF MOSFETs with a Shallowly-Implanted 2DEG-Like Channel. IEEE Electron. Device Lett. 2023, 44, 1060–1063. [Google Scholar] [CrossRef]
  95. Lv, Y.; Liu, H.; Wang, Y.; Fu, X.; Ma, C.; Song, X.; Zhou, X.; Zhang, Y.; Dong, P.; Du, H.; et al. Oxygen annealing impact on β-Ga2O3 MOSFETs: Improved pinch-off characteristic and output power density. Appl. Phys. Lett. 2020, 117, 133503. [Google Scholar] [CrossRef]
  96. Saha, C.N.; Vaidya, A.; Singisetti, U. Temperature dependent pulsed IV and RF characterization of β-(AlxGa1−x)2O3/Ga2O3 hetero-structure FET with ex situ passivation. Appl. Phys. Lett. 2022, 120, 172102. [Google Scholar] [CrossRef]
  97. Vaidya, A.; Saha, C.N.; Singisetti, U. Enhancement Mode β-(AlxGa1−x)2O3/Ga2O3 Heterostructure FET (HFET) with High Transconductance and Cutoff Frequency. IEEE Electron. Device Lett. 2021, 42, 1444–1447. [Google Scholar] [CrossRef]
  98. Zhang, Y.; Huang, S.; Wei, K.; Zhang, S.; Wang, X.; Zheng, Y.; Liu, G.; Chen, X.; Li, Y.; Liu, X. Millimeter-Wave AlGaN/GaN HEMTs with 43.6% Power-Added-Efficiency at 40 GHz Fabricated by Atomic Layer Etching Gate Recess. IEEE Electron. Device Lett. 2020, 41, 701–704. [Google Scholar] [CrossRef]
  99. Moon, J.-S.; Grabar, B.; Wong, J.; Chuong, D.; Arkun, E.; Morales, D.V.; Chen, P.; Malek, C.; Fanning, D.; Venkatesan, N.; et al. Power Scaling of Graded-Channel GaN HEMTs with Mini-Field-Plate T-gate and 156 GHz fT. IEEE Electron. Device Lett. 2021, 42, 796–799. [Google Scholar] [CrossRef]
  100. Yu, C.; Zhou, C.; Guo, J.; He, Z.; Ma, M.; Yu, H.; Song, X.; Bu, A.; Feng, Z. Hydrogen-terminated diamond MOSFETs on (0 0 1) single crystal diamond with state of the art high RF power density. Funct. Diam. 2022, 2, 64–70. [Google Scholar] [CrossRef]
  101. 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]
  102. Krishnamoorthy, S.; Xia, Z.; Bajaj, S.; Brenner, M.; Rajan, S. Delta-doped β-gallium oxide field-effect transistor. Appl. Phys. Express 2017, 10, 051102. [Google Scholar] [CrossRef]
  103. Joishi, C.; Xia, Z.; McGlone, J.; Zhang, Y.; Arehart, A.R.; Ringel, S.; Lodha, S.; Rajan, S. Effect of buffer iron doping on delta-doped β-Ga2O3 metal semiconductor field effect transistors. Appl. Phys. Lett. 2018, 113, 123501. [Google Scholar] [CrossRef]
  104. Xia, Z.; Joishi, C.; Krishnamoorthy, S.; Bajaj, S.; Zhang, Y.; Brenner, M.; Lodha, S.; Rajan, S. Delta Doped β-Ga2O3 Field Effect Transistors with Regrown Ohmic Contacts. IEEE Electron. Device Lett. 2018, 39, 568–571. [Google Scholar] [CrossRef]
  105. McGlone, J.F.; Xia, Z.; Joishi, C.; Lodha, S.; Rajan, S.; Ringel, S.; Arehart, A.R. Identification of critical buffer traps in Si δ-doped β-Ga2O3 MESFETs. Appl. Phys. Lett. 2019, 115, 153501. [Google Scholar] [CrossRef]
  106. Dheenan, A.V.; Dheenan, A.V.; McGlone, J.F.; McGlone, J.F.; Kalarickal, N.K.; Kalarickal, N.K.; Huang, H.-L.; Huang, H.-L.; Brenner, M.; Brenner, M.; et al. β-Ga2O3 MESFETs with insulating Mg-doped buffer grown by plasma-assisted molecular beam epitaxy. Appl. Phys. Lett. 2022, 121, 113503. [Google Scholar] [CrossRef]
  107. Bhattacharyya, A.; Roy, S.; Ranga, P.; Shoemaker, D.; Song, Y.; Lundh, J.S.; Choi, S.; Krishnamoorthy, S. 130 mA mm−1 β-Ga2O3 metal semiconductor field effect transistor with low-temperature metalorganic vapor phase epitaxy-regrown ohmic contacts. Appl. Phys. Express 2021, 14, 076502. [Google Scholar] [CrossRef]
  108. Bhattacharyya, A.; Ranga, P.; Roy, S.; Peterson, C.; Alema, F.; Seryogin, G.; Osinsky, A.; Krishnamoorthy, S. Multi-kV Class β-Ga2O3 MESFETs with a Lateral Figure of Merit up to 355 MW/cm2. IEEE Electron. Device Lett. 2021, 42, 1272–1275. [Google Scholar] [CrossRef]
  109. Bhattacharyya, A.; Sharma, S.; Alema, F.; Ranga, P.; Roy, S.; Peterson, C.; Seryogin, G.; Osinsky, A.; Singisetti, U.; Krishnamoorthy, S. 4.4 kV β-Ga2O3 MESFETs with power figure of merit exceeding 100 MW cm−2. Appl. Phys. Express 2022, 15, 061001. [Google Scholar] [CrossRef]
  110. Liddy, K.J.; Green, A.J.; Hendricks, N.S.; Heller, E.R.; Moser, N.A.; Leedy, K.D.; Popp, A.; Lindquist, M.T.; Tetlak, S.E.; Wagner, G.; et al. Thin channel β-Ga2O3 MOSFETs with self-aligned refractory metal gates. Appl. Phys. Express 2019, 12, 126501. [Google Scholar] [CrossRef]
  111. Moser, N.; Liddy, K.; Islam, A.; Miller, N.; Leedy, K.; Asel, T.; Mou, S.; Green, A.; Chabak, K. Toward high voltage radio frequency devices in β-Ga2O3. Appl. Phys. Lett. 2020, 117, 242101. [Google Scholar] [CrossRef]
  112. Wong, M.H.; Nakata, Y.; Kuramata, A.; Yamakoshi, S.; Higashiwaki, M. Enhancement-mode Ga2O3 MOSFETs with Si-ion-implanted source and drain. Appl. Phys. Express 2017, 10, 041101. [Google Scholar] [CrossRef]
  113. Chabak, K.D.; Moser, N.; Green, A.J.; Walker, D.E.; Tetlak, S.E.; Heller, E.; Crespo, A.; Fitch, R.; McCandless, J.P.; Leedy, K.; et al. Enhancement-mode Ga2O3 wrap-gate fin field-effect transistors on native (100) β-Ga2O3 substrate with high breakdown voltage. Appl. Phys. Lett. 2016, 109, 213501. [Google Scholar] [CrossRef]
  114. Dong, H.; Long, S.; Sun, H.; Zhao, X.; He, Q.; Qin, Y.; Jian, G.; Zhou, X.; Yu, Y.; Guo, W.; et al. Fast Switching β-Ga2O3 Power MOSFET with a Trench-Gate Structure. IEEE Electron. Device Lett. 2019, 40, 1385–1388. [Google Scholar] [CrossRef]
  115. Do, H.-B.; Phan-Gia, A.-V.; Nguyen, V.Q.; De Souza, M.M. Optimization of normally-off β-Ga2O3 MOSFET with high Ion and BFOM: A TCAD study. AIP Adv. 2022, 12, 065024. [Google Scholar] [CrossRef]
  116. Sharma, R.; Patnaik, A.; Sharma, P. Impact of doping concentration and recess depth to achieve enhancement mode operation in β-Ga2O3 MOSFET. Microelectron. J. 2023, 135, 105755. [Google Scholar] [CrossRef]
  117. Jang, C.-H.; Atmaca, G.; Cha, H.-Y. Normally-off β-Ga2O3 MOSFET with an Epitaxial Drift Layer. Micromachines 2022, 13, 1185. [Google Scholar] [CrossRef] [PubMed]
  118. Hu, Z.; Nomoto, K.; Li, W.; Zhang, Z.; Tanen, N.; Thieu, Q.T.; Sasaki, K.; Kuramata, A.; Nakamura, T.; Jena, D.; et al. Breakdown mechanism in 1 kA/cm2 and 960 V E-mode β-Ga2O3 vertical transistors. Appl. Phys. Lett. 2018, 113, 122103. [Google Scholar] [CrossRef]
  119. Hu, Z.; Nomoto, K.; Li, W.; Tanen, N.; Sasaki, K.; Kuramata, A.; Nakamura, T.; Jena, D.; Xing, H.G. Enhancement-Mode Ga2O3 Vertical Transistors with Breakdown Voltage >1 kV. IEEE Electron. Device Lett. 2018, 39, 869–872. [Google Scholar] [CrossRef]
  120. 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. In Proceedings of the 2019 31st International Symposium on Power Semiconductor Devices and ICs (ISPSD), Shanghai, China, 12–23 May 2019; pp. 483–486. [Google Scholar] [CrossRef]
  121. Wakimoto, D.; Lin, C.-H.; Thieu, Q.T.; Miyamoto, H.; Sasaki, K.; Kuramata, A. Nitrogen-doped β-Ga2O3 vertical transistors with a threshold voltage of ≥1.3 V and a channel mobility of 100 cm2 V−1 s−1. Appl. Phys. Express 2023, 16, 036503. [Google Scholar] [CrossRef]
  122. Tetzner, K.; Klupsch, M.; Popp, A.; Bin Anooz, S.; Chou, T.-S.; Galazka, Z.; Ickert, K.; Matalla, M.; Unger, R.-S.; Treidel, E.B.; et al. Enhancement-mode vertical (100) β-Ga2O3 FinFETs with an average breakdown strength of 2.7 MV cm−1. Jpn. J. Appl. Phys. 2023, 62, SF1010. [Google Scholar] [CrossRef]
  123. Song, Y.; Mohseni, P.K.; Kim, S.H.; Shin, J.C.; Ishihara, T.; Adesida, I.; Li, X. Ultra-High Aspect Ratio InP Junctionless FinFETs by a Novel Wet Etching Method. IEEE Electron. Device Lett. 2016, 37, 970–973. [Google Scholar] [CrossRef]
  124. Huang, H.-C.; Ren, Z.; Chan, C.; Li, X. Wet etch, dry etch, and MacEtch of β-Ga2O3: A review of characteristics and mechanism. J. Mater. Res. 2021, 36, 4756–4770. [Google Scholar] [CrossRef]
  125. Huang, H.-C.; Ren, Z.; Bhuiyan, A.F.M.A.U.; Feng, Z.; Yang, Z.; Luo, X.; Huang, A.Q.; Green, A.; Chabak, K.; Zhao, H.; et al. β-Ga2O3 FinFETs with ultra-low hysteresis by plasma-free metal-assisted chemical etching. Appl. Phys. Lett. 2022, 121, 052102. [Google Scholar] [CrossRef]
  126. Huang, H.-C.; Kim, M.; Zhan, X.; Chabak, K.; Kim, J.D.; Kvit, A.; Liu, D.; Ma, Z.; Zuo, J.-M.; Li, X. High Aspect Ratio β-Ga2O3 Fin Arrays with Low-Interface Charge Density by Inverse Metal-Assisted Chemical Etching. ACS Nano 2019, 13, 8784–8792. [Google Scholar] [CrossRef]
  127. Ren, Z.; Huang, H.-C.; Lee, H.; Chan, C.; Roberts, H.C.; Wu, X.; Waseem, A.; Bhuiyan, A.F.M.A.U.; Zhao, H.; Zhu, W.; et al. Temperature dependent characteristics of β-Ga2O3 FinFETs by MacEtch. Appl. Phys. Lett. 2023, 123, 043505. [Google Scholar] [CrossRef]
  128. Wong, M.H.; Sasaki, K.; Kuramata, A.; Yamakoshi, S.; Higashiwaki, M. Field-Plated Ga2O3 MOSFETs with a Breakdown Voltage of over 750 V. IEEE Electron. Device Lett. 2016, 37, 212–215. [Google Scholar] [CrossRef]
  129. Zeng, K.; Vaidya, A.; Singisetti, U. 1.85 kV Breakdown Voltage in Lateral Field-Plated Ga2O3 MOSFETs. IEEE Electron. Device Lett. 2018, 39, 1385–1388. [Google Scholar] [CrossRef]
  130. Zeng, K.; Vaidya, A.; Singisetti, U. A field-plated Ga2O3 MOSFET with near 2-kV breakdown voltage and 520 mΩ · cm2 on-resistance. Appl. Phys. Express 2019, 12, 081003. [Google Scholar] [CrossRef]
  131. Sharma, S.; Zeng, K.; Saha, S.; Singisetti, U. Field-Plated Lateral Ga2O3 MOSFETs with Polymer Passivation and 8.03 kV Breakdown Voltage. IEEE Electron. Device Lett. 2020, 41, 836–839. [Google Scholar] [CrossRef]
  132. Vetury, R.; Zhang, N.Q.; Keller, S.; Mishra, U.K. The impact of surface states on the DC and RF characteristics of AlGaN/GaN HFETs. IEEE Trans. Electron. Devices 2001, 48, 560–566. [Google Scholar] [CrossRef]
  133. Vaidya, A.; Singisetti, U. Temperature-Dependent Current Dispersion Study in β-Ga2O3 FETs Using Submicrosecond Pulsed I–V Characteristics. IEEE Trans. Electron. Devices 2021, 68, 3755–3761. [Google Scholar] [CrossRef]
  134. Maimon, O.; A Moser, N.; Liddy, K.J.; Green, A.J.; Chabak, K.D.; Cheung, K.P.; Pookpanratana, S.; Li, Q. Measurement and gate-voltage dependence of channel and series resistances in lateral depletion-mode β-Ga2O3 MOSFETs. Semicond. Sci. Technol. 2023, 38, 075016. [Google Scholar] [CrossRef]
  135. Mun, J.K.; Cho, K.; Chang, W.; Jung, H.-W.; Do, J. Editors’ Choice—2.32 kV Breakdown Voltage Lateral β-Ga2O3 MOSFETs with Source-Connected Field Plate. ECS J. Solid State Sci. Technol. 2019, 8, Q3079–Q3082. [Google Scholar] [CrossRef]
  136. Lv, Y.; Zhou, X.; Long, S.; Liang, S.; Song, X.; Zhou, X.; Dong, H.; Wang, Y.; Feng, Z.; Cai, S. Lateral source field-plated β-Ga2O3 MOSFET with recorded breakdown voltage of 2360 V and low specific on-resistance of 560 mΩ cm2. Semicond. Sci. Technol. 2019, 34, 11LT02. [Google Scholar] [CrossRef]
  137. Lv, Y.; Zhou, X.; Long, S.; Song, X.; Wang, Y.; Liang, S.; He, Z.; Han, T.; Tan, X.; Feng, Z.; et al. Source-Field-Plated β-Ga2O3 MOSFET with Record Power Figure of Merit of 50.4 MW/cm2. IEEE Electron. Device Lett. 2019, 40, 83–86. [Google Scholar] [CrossRef]
  138. Zhu, M.; Xie, Y.; Shao, J.; Chen, Y. Nanofabrications of T shape gates for high electron mobility transistors in microwaves and THz waves, a review. Micro Nano Eng. 2021, 13, 100091. [Google Scholar] [CrossRef]
  139. Chabak, K.; Walker, D.; Green, A.; Crespo, A.; Lindquist, M.; Leedy, K.; Tetlak, S.; Gilbert, R.; Moser, N.A.; Jessen, G. Sub-Micron Gallium Oxide Radio Frequency Field-Effect Transistors. In Proceedings of the 2018 IEEE MTT-S International Microwave Workshop Series on Advanced Materials and Processes for RF and THz Applications (IMWS-AMP), Ann Arbor, MI, USA, 16–18 July 2018; pp. 1–3. [Google Scholar] [CrossRef]
  140. Kamimura, T.; Nakata, Y.; Higashiwaki, M. Delay-time analysis in radio-frequency β-Ga2O3 field effect transistors. Appl. Phys. Lett. 2020, 117, 253501. [Google Scholar] [CrossRef]
  141. Hwang, W.S.; Verma, A.; Peelaers, H.; Protasenko, V.; Rouvimov, S.; Xing, H.; Seabaugh, A.; Haensch, W.; Van de Walle, C.; Galazka, Z.; et al. High-voltage field effect transistors with wide-bandgap β-Ga2O3 nanomembranes. Appl. Phys. Lett. 2014, 104, 203111. [Google Scholar] [CrossRef]
  142. Ahn, S.; Ren, F.; Kim, J.; Oh, S.; Kim, J.; Mastro, M.A.; Pearton, S.J. Effect of front and back gates on β-Ga2O3 nano-belt field-effect transistors. Appl. Phys. Lett. 2016, 109, 062102. [Google Scholar] [CrossRef]
  143. Kim, J.; Oh, S.; Mastro, M.A.; Kim, J. Exfoliated β-Ga2O3 nano-belt field-effect transistors for air-stable high power and high temperature electronics. Phys. Chem. Chem. Phys. 2016, 18, 15760–15764. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, G.; Jang, S.; Ren, F.; Pearton, S.J.; Kim, J. Influence of High-Energy Proton Irradiation on β-Ga2O3 Nanobelt Field-Effect Transistors. ACS Appl. Mater. Interfaces 2017, 9, 40471–40476. [Google Scholar] [CrossRef]
  145. Son, J.; Kwon, Y.; Kim, J.; Kim, J. Tuning the Threshold Voltage of Exfoliated β-Ga2O3 Flake-Based Field-Effect Transistors by Photo-Enhanced H3PO4 Wet Etching. ECS J. Solid State Sci. Technol. 2018, 7, Q148–Q151. [Google Scholar] [CrossRef]
  146. Li, Z.; Liu, Y.; Zhang, A.; Liu, Q.; Shen, C.; Wu, F.; Xu, C.; Chen, M.; Fu, H.; Zhou, C. Quasi-two-dimensional β-Ga2O3 field effect transistors with large drain current density and low contact resistance via controlled formation of interfacial oxygen vacancies. Nano Res. 2019, 12, 143–148. [Google Scholar] [CrossRef]
  147. Kim, J.; Tadjer, M.J.; Mastro, M.A.; Kim, J. Controlling the threshold voltage of β-Ga2O3 field-effect transistors via remote fluorine plasma treatment. J. Mater. Chem. C 2019, 7, 8855–8860. [Google Scholar] [CrossRef]
  148. Polyakov, A.Y.; Smirnov, N.B.; Shchemerov, I.V.; Chernykh, S.V.; Oh, S.; Pearton, S.J.; Ren, F.; Kochkova, A.; Kim, J. Defect States Determining Dynamic Trapping-Detrapping in β-Ga2O3 Field-Effect Transistors. ECS J. Solid State Sci. Technol. 2019, 8, Q3013–Q3018. [Google Scholar] [CrossRef]
  149. Ma, J.; Lee, O.; Yoo, G. Effect of Al2O3 Passivation on Electrical Properties of β-Ga2O3 Field-Effect Transistor. IEEE J. Electron. Devices Soc. 2019, 7, 512–516. [Google Scholar] [CrossRef]
  150. Kim, S.; Kim, J. Electrical Properties of Thermally Annealed β-Ga2O3 Metal-Semiconductor Field-Effect Transistors with Pt/Au Schottky Contacts. ECS J. Solid State Sci. Technol. 2019, 8, Q3122–Q3125. [Google Scholar] [CrossRef]
  151. Chen, J.-X.; Li, X.-X.; Tao, J.-J.; Cui, H.-Y.; Huang, W.; Ji, Z.-G.; Sai, Q.-L.; Xia, C.-T.; Lu, H.-L.; Zhang, D.W. Fabrication of a Nb-Doped β-Ga2O3 Nanobelt Field-Effect Transistor and Its Low-Temperature Behavior. ACS Appl. Mater. Interfaces 2020, 12, 8437–8445. [Google Scholar] [CrossRef] [PubMed]
  152. Ma, J.; Yoo, G. Low Subthreshold Swing Double-Gate β-Ga2O3 Field-Effect Transistors with Polycrystalline Hafnium Oxide Dielectrics. IEEE Electron. Device Lett. 2019, 40, 1317–1320. [Google Scholar] [CrossRef]
  153. Madadi, D.; Orouji, A.A. Scattering mechanisms in β-Ga2O3 junctionless SOI MOSFET: Investigation of electron mobility and short channel effects. Mater. Today Commun. 2021, 26, 102044. [Google Scholar] [CrossRef]
  154. Zhou, H.; Maize, K.; Noh, J.; Shakouri, A.; Ye, P.D. Thermodynamic Studies of β-Ga2O3 Nanomembrane Field-Effect Transistors on a Sapphire Substrate. ACS Omega 2017, 2, 7723–7729. [Google Scholar] [CrossRef] [PubMed]
  155. Ma, J.; Lee, O.; Yoo, G. Abnormal Bias-Temperature Stress and Thermal Instability of β-Ga2O3 Nanomembrane Field-Effect Transistor. IEEE J. Electron. Devices Soc. 2018, 6, 1124–1128. [Google Scholar] [CrossRef]
  156. Xu, W.; Zhang, Y.; Hao, Y.; Wang, X.; Wang, Y.; You, T.; Ou, X.; Han, G.; Hu, H.; Zhang, S.; et al. First Demonstration of Waferscale Heterogeneous Integration of Ga2O3 MOSFETs on SiC and Si Substrates by Ion-Cutting Process. In Proceedings of the 2019 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 7–11 December 2019; pp. 12.5.1–12.5.4. [Google Scholar] [CrossRef]
  157. Cheng, Z.; Yates, L.; Shi, J.; Tadjer, M.J.; Hobart, K.D.; Graham, S. Thermal conductance across β-Ga2O3-diamond van der Waals heterogeneous interfaces. APL Mater. 2019, 7, 031118. [Google Scholar] [CrossRef]
  158. Lee, D.; Kim, H.W.; Kim, J.; Moon, J.H.; Lee, G.; Kim, J. Ultra-Wide Bandgap β-Ga2O3 Heterojunction Field-Effect Transistor Using p-Type 4H-SiC Gate for Efficient Thermal Management. ECS J. Solid State Sci. Technol. 2020, 9, 065006. [Google Scholar] [CrossRef]
  159. Wang, Y.; Han, G.; Xu, W.; You, T.; Hu, H.; Liu, Y.; Zhang, X.; Huang, H.; Ou, X.; Ma, X.; et al. Recessed-Gate Ga2O3-on-SiC MOSFETs Demonstrating a Stable Power Figure of Merit of 100 mW/cm2 Up to 200 °C. IEEE Trans. Electron. Devices 2022, 69, 1945–1949. [Google Scholar] [CrossRef]
  160. Kim, K.; Jin, H.; Choi, W.; Jeong, Y.; Shin, H.G.; Lee, Y.; Kim, K.; Im, S. High Performance β-Ga2O3 Schottky Barrier Transistors with Large Work Function TMD Gate of NbS2 and TaS2. Adv. Funct. Mater. 2021, 31, 2010303. [Google Scholar] [CrossRef]
  161. Kim, J.; Mastro, M.A.; Tadjer, M.J.; Kim, J. Heterostructure WSe2−Ga2O3 Junction Field-Effect Transistor for Low-Dimensional High-Power Electronics. ACS Appl. Mater. Interfaces 2018, 10, 29724–29729. [Google Scholar] [CrossRef] [PubMed]
  162. Choi, W.; Ahn, J.; Kim, K.; Jin, H.; Hong, S.; Hwang, D.K.; Im, S. Ambipolar Channel p-TMD/n-Ga2O3 Junction Field Effect Transistors and High Speed Photo-sensing in TMD Channel. Adv. Mater. 2021, 33, 2103079. [Google Scholar] [CrossRef] [PubMed]
  163. Li, C.; Chen, C.; Chen, J.; He, T.; Li, H.; Yang, Z.; Xie, L.; Wang, Z.; Zhang, K. High-performance junction field-effect transistor based on black phosphorus/β-Ga2O3 heterostructure. J. Semicond. 2020, 41, 082002. [Google Scholar] [CrossRef]
  164. Kim, J.; Mastro, M.A.; Tadjer, M.J.; Kim, J. Quasi-Two-Dimensional h-BN/β-Ga2O3 Heterostructure Metal–Insulator–Semiconductor Field-Effect Transistor. ACS Appl. Mater. Interfaces 2017, 9, 21322–21327. [Google Scholar] [CrossRef] [PubMed]
  165. Kim, J.; Kim, J. Monolithically Integrated Enhancement-Mode and Depletion-Mode β-Ga2O3 MESFETs with Graphene-Gate Architectures and Their Logic Applications. ACS Appl. Mater. Interfaces 2020, 12, 7310–7316. [Google Scholar] [CrossRef] [PubMed]
  166. Gao, M.; Huang, H.; Yin, L.; Lu, X.; Zhang, J.; Ren, K. A Novel Field-Plated Lateral β-Ga2O3 MOSFET Featuring Self-Aligned Vertical Gate Structure. IEEE Trans. Electron. Devices 2023, 70, 4309–4314. [Google Scholar] [CrossRef]
  167. Sun, Z.; Huang, H.; Sun, N.; Tao, P.; Zhao, C.; Liang, Y.C. A Novel GaN Metal-Insulator-Semiconductor High Electron Mobility Transistor Featuring Vertical Gate Structure. Micromachines 2019, 10, 848. [Google Scholar] [CrossRef]
  168. Goyal, P.; Kaur, H. Exploring the efficacy of implementing field plate design with air gap on β-Ga2O3 MOSFET for high power & RF applications. Micro Nanostruct. 2023, 173, 207454. [Google Scholar] [CrossRef]
  169. Ranjan, R.; Kashyap, N.; Raman, A. Design and investigation of field plate-based vertical GAA—β-(AlGa)2O3/Ga2O3 high electron mobility transistor. Micro Nanostruct. 2022, 164, 107117. [Google Scholar] [CrossRef]
  170. Mehta, M.; Avasthi, S. The possibility of gallium oxide (β-Ga2O3) heterojunction bipolar transistors. Phys. Scr. 2023, 98, 025013. [Google Scholar] [CrossRef]
  171. Hu, C.; Chi, M.-H.; Patel, V. Optimum design of power MOSFET’s. IEEE Trans. Electron. Devices 1984, 31, 1693–1700. [Google Scholar] [CrossRef]
  172. Shenoy, J.; Cooper, J.; Melloch, M. High-voltage double-implanted power MOSFET’s in 6H-SiC. IEEE Electron. Device Lett. 1997, 18, 93–95. [Google Scholar] [CrossRef]
  173. Ben-Yaacov, I.; Seck, Y.-K.; Mishra, U.K.; DenBaars, S.P. AlGaN/GaN current aperture vertical electron transistors with regrown channels. J. Appl. Phys. 2004, 95, 2073–2078. [Google Scholar] [CrossRef]
  174. Wong, M.H.; Murakami, H.; Kumagai, Y.; Higashiwaki, M. Enhancement-Mode β-Ga2O3 Current Aperture Vertical MOSFETs with N-Ion-Implanted Blocker. IEEE Electron. Device Lett. 2019, 41, 296–299. [Google Scholar] [CrossRef]
  175. Wong, M.H.; Goto, K.; Morikawa, Y.; Kuramata, A.; Yamakoshi, S.; Murakami, H.; Kumagai, Y.; Higashiwaki, M. All-ion-implanted planar-gate current aperture vertical Ga2O3 MOSFETs with Mg-doped blocking layer. Appl. Phys. Express 2018, 11, 064102. [Google Scholar] [CrossRef]
  176. Wong, M.H.; Murakami, H.; Kumagai, Y.; Higashiwaki, M. Aperture-limited conduction and its possible mechanism in ion-implanted current aperture vertical β-Ga2O3 MOSFETs. Appl. Phys. Lett. 2021, 118, 012102. [Google Scholar] [CrossRef]
  177. Wong, M.H.; Lin, C.-H.; Kuramata, A.; Yamakoshi, S.; Murakami, H.; Kumagai, Y.; Higashiwaki, M. Acceptor doping of β-Ga2O3 by Mg and N ion implantations. Appl. Phys. Lett. 2018, 113, 102103. [Google Scholar] [CrossRef]
  178. Kim, I.K.; Cha, S.; Hong, S.-M. Optimization of Nitrogen Ion Implantation Condition for β-Ga2O3 Vertical MOSFETs via Process and Device Simulation. IEEE Trans. Electron. Devices 2022, 69, 6948–6955. [Google Scholar] [CrossRef]
  179. Ma, Y.; Zhou, X.; Tang, W.; Zhang, X.; Xu, G.; Zhang, L.; Chen, T.; Dai, S.; Bian, C.; Li, B.; et al. 702.3 A·cm−2/10.4 mΩ·cm2 β-Ga2O3 U-Shape Trench Gate MOSFET with N-Ion Implantation. IEEE Electron. Device Lett. 2023, 44, 384–387. [Google Scholar] [CrossRef]
  180. Hawkins, R.; Wang, X.; Moumen, N.; Wallace, R.M.; Young, C.D. Impact of process anneals on high-k/β-Ga2O3 interfaces and capacitance. J. Vac. Sci. Technol. A 2023, 41, 023203. [Google Scholar] [CrossRef]
  181. Kananen, B.E.; Halliburton, L.E.; Stevens, K.T.; Foundos, G.K.; Giles, N.C. Gallium vacancies in β-Ga2O3 crystals. Appl. Phys. Lett. 2017, 110, 202104. [Google Scholar] [CrossRef]
  182. Wu, S.; Liu, Z.; Yang, H.; Wang, Y. Effects of Annealing on Surface Residual Impurities and Intrinsic Defects of β-Ga2O3. Crystals 2023, 13, 1045. [Google Scholar] [CrossRef]
  183. Lv, Y.; Zhou, X.; Long, S.; Wang, Y.; Song, X.; Zhou, X.; Xu, G.; Liang, S.; Feng, Z.; Cai, S.; et al. Enhancement-Mode β-Ga2O3 Metal-Oxide-Semiconductor Field-Effect Transistor with High Breakdown Voltage over 3000 V Realized by Oxygen Annealing. Phys. Status Solidi (RRL) Rapid Res. Lett. 2019, 14, 1900586. [Google Scholar] [CrossRef]
  184. Oshima, T.; Kato, Y.; Kawano, N.; Kuramata, A.; Yamakoshi, S.; Fujita, S.; Oishi, T.; Kasu, M. Carrier confinement observed at modulation-doped β-(AlxGa1−x)2O3/Ga2O3 heterojunction interface. Appl. Phys. Express 2017, 10, 035701. [Google Scholar] [CrossRef]
  185. Krishnamoorthy, S.; Xia, Z.; Joishi, C.; Zhang, Y.; McGlone, J.; Johnson, J.; Brenner, M.; Arehart, A.R.; Hwang, J.; Lodha, S.; et al. Modulation-doped β-(Al0.2Ga0.8)2O3/Ga2O3 field-effect transistor. Appl. Phys. Lett. 2017, 111, 023502. [Google Scholar] [CrossRef]
  186. Ahmadi, E.; Koksaldi, O.S.; Zheng, X.; Mates, T.; Oshima, Y.; Mishra, U.K.; Speck, J.S. Demonstration of β-(AlxGa1−x)2O3/β-Ga2O3modulation doped field-effect transistors with Ge as dopant grown via plasma-assisted molecular beam epitaxy. Appl. Phys. Express 2017, 10, 071101. [Google Scholar] [CrossRef]
  187. Zhang, Y.; Neal, A.; Xia, Z.; Joishi, C.; Johnson, J.M.; Zheng, Y.; Bajaj, S.; Brenner, M.; Dorsey, D.; Chabak, K.; et al. Demonstration of high mobility and quantum transport in modulation-doped β-(AlxGa1−x)2O3/Ga2O3 heterostructures. Appl. Phys. Lett. 2018, 112, 173502. [Google Scholar] [CrossRef]
  188. Zhang, Y.; Xia, Z.; Mcglone, J.; Sun, W.; Joishi, C.; Arehart, A.R.; Ringel, S.A.; Rajan, S. Evaluation of Low-Temperature Saturation Velocity in β-(AlxGa1−x)2O3/Ga2O3 Modulation-Doped Field-Effect Transistors. IEEE Trans. Electron. Devices 2020, 66, 1574–1578. [Google Scholar] [CrossRef]
  189. Zhang, Y.; Joishi, C.; Xia, Z.; Brenner, M.; Lodha, S.; Rajan, S. Demonstration of β-(AlxGa1−x)2O3/Ga2O3 double heterostructure field effect transistors. Appl. Phys. Lett. 2018, 112, 233503. [Google Scholar] [CrossRef]
  190. Kalarickal, N.K.; Xia, Z.; McGlone, J.F.; Liu, Y.; Moore, W.; Arehart, A.R.; Ringel, S.A.; Rajan, S. High electron density β-(Al0.17Ga0.83)2O3/Ga2O3 modulation doping using an ultra-thin (1 nm) spacer layer. J. Appl. Phys. 2020, 127, 215706. [Google Scholar] [CrossRef]
  191. Atmaca, G.; Cha, H.-Y. β-(Al0.17Ga 0.83)2O3/Ga2O3 Delta-Doped Heterostructure MODFETs with an Ultrathin Spacer Layer and a Back-Barrier Layer: A Comprehensive Technology Computer-Aided Design Analysis. Phys. Status Solidi A 2022, 219, 2100732. [Google Scholar] [CrossRef]
  192. He, X.; Hu, J.; Zhang, Z.; Liu, W.; Song, K.; Meng, J. Study on the interface electronic properties of AlN(0001)/β-Ga2O3(100). Surf. Interfaces 2021, 28, 101585. [Google Scholar] [CrossRef]
  193. Singh, R.; Lenka, T.R.; Velpula, R.T.; Jain, B.; Bui, H.Q.T.; Nguyen, H.P.T. Investigation of current collapse and recovery time due to deep level defect traps in β-Ga2O3 HEMT. J. Semicond. 2020, 41, 102802. [Google Scholar] [CrossRef]
  194. Song, K.; Zhang, H.; Fu, H.; Yang, C.; Singh, R.; Zhao, Y.; Sun, H.; Long, S. Normally-off AlN/β-Ga2O3 field-effect transistors using polarization-induced doping. J. Phys. D Appl. Phys. 2020, 53, 345107. [Google Scholar] [CrossRef]
  195. Singh, R.; Lenka, T.R.; Velpula, R.T.; Jain, B.; Bui, H.Q.T.; Nguyen, H.P.T. A novel β-Ga2O3 HEMT with fT of 166 GHz and X-band POUT of 2.91 W/mm. Int. J. Numer. Model. Electron. Netw. Devices Fields 2021, 34, e2794. [Google Scholar] [CrossRef]
  196. Singh, R.; Rao, G.P.; Lenka, T.R.; Prasad, S.V.S.; Boukortt, N.E.I.; Crupi, G.; Nguyen, H.P.T. Design and simulation of T-gate AlN/β-Ga2O3 HEMT for DC, RF and high-power nanoelectronics switching applications. Int. J. Numer. Model. Electron. Netw. Devices Fields 2023, e3146. [Google Scholar] [CrossRef]
  197. Singh, R.; Lenka, T.; Panda, D.; Nguyen, H.; Boukortt, N.E.I.; Crupi, G. Analytical modeling of I–V characteristics using 2D Poisson equations in AlN/β-Ga2O3 HEMT. Mater. Sci. Semicond. Process. 2022, 145, 106627. [Google Scholar] [CrossRef]
  198. Russell, S.A.O.; Perez-Tomas, A.; McConville, C.F.; Fisher, C.A.; Hamilton, D.P.; Mawby, P.A.; Jennings, M.R. Heteroepitaxial Beta-Ga2O3 on 4H-SiC for an FET with Reduced Self Heating. IEEE J. Electron. Devices Soc. 2017, 5, 256–261. [Google Scholar] [CrossRef]
  199. Zhang, M.; Wang, L.; Yang, K.; Yao, J.; Tang, W.; Guo, Y. Breakdown Characteristics of Ga2O3-on-SiC Metal-Oxide-Semiconductor Field-Effect Transistors. Crystals 2023, 13, 917. [Google Scholar] [CrossRef]
  200. Hu, J.; Xu, B.; Zhang, Z.; He, X.; Li, L.; Cheng, H.; Wang, J.; Meng, J.; Wang, X.; Zhang, C.; et al. Step flow growth of β-Ga2O3 films on off-axis 4H-SiC substrates by LPCVD. Surf. Interfaces 2023, 37, 102732. [Google Scholar] [CrossRef]
  201. Song, Y.; Shoemaker, D.; Leach, J.H.; McGray, C.; Huang, H.-L.; Bhattacharyya, A.; Zhang, Y.; Gonzalez-Valle, C.U.; Hess, T.; Zhukovsky, S.; et al. Ga2O3-on-SiC Composite Wafer for Thermal Management of Ultrawide Bandgap Electronics. ACS Appl. Mater. Interfaces 2021, 13, 40817–40829. [Google Scholar] [CrossRef] [PubMed]
  202. Oh, J.; Ma, J.; Yoo, G. Simulation study of reduced self-heating in β-Ga2O3 MOSFET on a nano-crystalline diamond substrate. Results Phys. 2019, 13, 102151. [Google Scholar] [CrossRef]
  203. Cheng, Z.; Wheeler, V.D.; Bai, T.; Shi, J.; Tadjer, M.J.; Feygelson, T.; Hobart, K.D.; Goorsky, M.S.; Graham, S. Integration of polycrystalline Ga2O3 on diamond for thermal management. Appl. Phys. Lett. 2020, 116, 062105. [Google Scholar] [CrossRef]
  204. Chatterjee, B.; Zeng, K.; Nordquist, C.D.; Singisetti, U.; Choi, S. Device-Level Thermal Management of Gallium Oxide Field-Effect Transistors. IEEE Trans. Compon. Packag. Manuf. Technol. 2019, 9, 2352–2365. [Google Scholar] [CrossRef]
  205. Green, A.J.; Chabak, K.D.; Heller, E.R.; Fitch, R.C.; Baldini, M.; Fiedler, A.; Irmscher, K.; Wagner, G.; Galazka, Z.; Tetlak, S.E.; et al. 3.8-MV/cm Breakdown Strength of MOVPE-Grown Sn-Doped β-Ga2O3 MOSFETs. IEEE Electron Device Lett. 2016, 37, 902–905. [Google Scholar] [CrossRef]
  206. Higashiwaki, M.; Sasaki, K.; Kamimura, T.; Wong, M.H.; 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]
  207. Lee, M.-H.; Peterson, R.L. Interfacial reactions of titanium/gold ohmic contacts with Sn-doped β-Ga2O3. APL Mater. 2019, 7, 022524. [Google Scholar] [CrossRef]
  208. Lee, M.-H.; Peterson, R.L. Annealing Induced Interfacial Evolution of Titanium/Gold Metallization on Unintentionally Doped β-Ga2O3. ECS J. Solid State Sci. Technol. 2019, 8, Q3176–Q3179. [Google Scholar] [CrossRef]
  209. Lee, M.-H.; Peterson, R.L. Accelerated Aging Stability of β-Ga2O3–Titanium/Gold Ohmic Interfaces. ACS Appl. Mater. Interfaces 2020, 12, 46277–46287. [Google Scholar] [CrossRef] [PubMed]
  210. Kim, Y.; Kim, M.-K.; Baik, K.H.; Jang, S. Low-Resistance Ti/Au Ohmic Contact on (001) Plane Ga2O3 Crystal. ECS J. Solid State Sci. Technol. 2022, 11, 045003. [Google Scholar] [CrossRef]
  211. Yao, Y.; Davis, R.F.; Porter, L.M. Investigation of Different Metals as Ohmic Contacts to β-Ga2O3: Comparison and Analysis of Electrical Behavior, Morphology, and Other Physical Properties. J. Electron. Mater. 2017, 46, 2053–2060. [Google Scholar] [CrossRef]
  212. Shi, J.; Xia, X.; Liang, H.; Abbas, Q.; Liu, J.; Zhang, H.; Liu, Y. Low resistivity ohmic contacts on lightly doped n-type β-Ga2O3 using Mg/Au. J. Mater. Sci. Mater. Electron. 2019, 30, 3860–3864. [Google Scholar] [CrossRef]
  213. Tetzner, K.; Schewski, R.; Popp, A.; Bin Anooz, S.; Chou, T.-S.; Ostermay, I.; Kirmse, H.; Würfl, J. Refractory metal-based ohmic contacts on β-Ga2O3 using TiW. APL Mater. 2022, 10, 071108. [Google Scholar] [CrossRef]
  214. Sasaki, K.; Higashiwaki, M.; Kuramata, A.; Masui, T.; Yamakoshi, S. Si-Ion Implantation Doping in β-Ga2O3 and Its Application to Fabrication of Low-Resistance Ohmic Contacts. Appl. Phys. Express 2013, 6, 086502. [Google Scholar] [CrossRef]
  215. Higashiwaki, M.; Sasaki, K.; Wong, M.H.; Kamimura, T.; Krishnamurthy, D.; Kuramata, A.; Masui, T.; Yamakoshi, S. Depletion-mode Ga2O3 MOSFETs on β-Ga2O3 (010) substrates with Si-ion-implanted channel and contacts. In Proceedings of the 2013 IEEE International Electron Devices Meeting (IEDM), Washington, DC, USA, 9–11 December 2013; pp. 28.7.1–28.7.4. [Google Scholar] [CrossRef]
  216. Wong, M.H.; Sasaki, K.; Kuramata, A.; Yamakoshi, S.; Higashiwaki, M. Electron channel mobility in silicon-doped Ga2O3 MOSFETs with a resistive buffer layer. Jpn. J. Appl. Phys. 2016, 55, 1202B9. [Google Scholar] [CrossRef]
  217. Tetzner, K.; Thies, A.; Seyidov, P.; Chou, T.-S.; Rehm, J.; Ostermay, I.; Galazka, Z.; Fiedler, A.; Popp, A.; Würfl, J.; et al. Ge-ion implantation and activation in (100) β-Ga2O3 for ohmic contact improvement using pulsed rapid thermal annealing. J. Vac. Sci. Technol. A 2023, 41, 043102. [Google Scholar] [CrossRef]
  218. Zhang, Y.; Alema, F.; Mauze, A.; Koksaldi, O.S.; Miller, R.; Osinsky, A.; Speck, J.S. MOCVD grown epitaxial β-Ga2O3 thin film with an electron mobility of 176 cm2/V s at room temperature. APL Mater. 2018, 7, 022506. [Google Scholar] [CrossRef]
  219. Alema, F.; Peterson, C.; Bhattacharyya, A.; Roy, S.; Krishnamoorthy, S.; Osinsky, A. Low Resistance Ohmic Contact on Epitaxial MOVPE Grown β-Ga2O3 and β-(AlxGa1−x)2O3 Films. IEEE Electron. Device Lett. 2022, 43, 1649–1652. [Google Scholar] [CrossRef]
  220. Oshima, T.; Wakabayashi, R.; Hattori, M.; Hashiguchi, A.; Kawano, N.; Sasaki, K.; Masui, T.; Kuramata, A.; Yamakoshi, S.; Yoshimatsu, K.; et al. Formation of indium–tin oxide ohmic contacts for β-Ga2O3. Jpn. J. Appl. Phys. 2016, 55, 1202B7. [Google Scholar] [CrossRef]
  221. Carey, P.H.; Yang, J.; Ren, F.; Hays, D.C.; Pearton, S.J.; Kuramata, A.; Kravchenko, I.I. Improvement of Ohmic contacts on Ga2O3 through use of ITO-interlayers. J. Vac. Sci. Technol. B 2017, 35, 061201. [Google Scholar] [CrossRef]
  222. Carey, P.H.; Yang, J.; Ren, F.; Hays, D.C.; Pearton, S.J.; Jang, S.; Kuramata, A.; Kravchenko, I.I. Ohmic contacts on n-type β-Ga2O3 using AZO/Ti/Au. AIP Adv. 2017, 7, 095313. [Google Scholar] [CrossRef]
  223. Zeng, K.; Wallace, J.S.; Heimburger, C.; Sasaki, K.; Kuramata, A.; Masui, T.; Gardella, J.A.; Singisetti, U. Ga2O3MOSFETs Using Spin-On-Glass Source/Drain Doping Technology. IEEE Electron. Device Lett. 2017, 38, 513–516. [Google Scholar] [CrossRef]
  224. Zeng, K.; Singisetti, U. Temperature dependent characterization of Ga2O3 MOSFETs with Spin-on-Glass source/drain doping. In Proceedings of the 2017 75th Device Research Conference (DRC), South Bend, IN, USA, 25–28 June 2017; pp. 1–2. [Google Scholar] [CrossRef]
  225. Spencer, J.A.; Mock, A.L.; Jacobs, A.G.; Schubert, M.; Zhang, Y.; Tadjer, M.J. A review of band structure and material properties of transparent conducting and semiconducting oxides: Ga2O3, Al2O3, In2O3, ZnO, SnO2, CdO, NiO, CuO, and Sc2O3. Appl. Phys. Rev. 2022, 9, 011315. [Google Scholar] [CrossRef]
  226. Roy, S.; Chmielewski, A.E.; Bhattacharyya, A.; Ranga, P.; Sun, R.; Scarpulla, M.A.; Alem, N.; Krishnamoorthy, S. In Situ Dielectric Al2O3/β-Ga2O3 Interfaces Grown Using Metal–Organic Chemical Vapor Deposition. Adv. Electron. Mater. 2021, 7, 2100333. [Google Scholar] [CrossRef]
  227. Bhuiyan, A.F.M.A.U.; Meng, L.; Huang, H.-L.; Hwang, J.; Zhao, H. In situ MOCVD growth and band offsets of Al2O3 dielectric on β-Ga2O3 and β-(AlxGa1−x)2O3 thin films. J. Appl. Phys. 2022, 132, 165301. [Google Scholar] [CrossRef]
  228. Islam, A.E.; Zhang, C.; DeLello, K.; Muller, D.A.; Leedy, K.D.; Ganguli, S.; Moser, N.A.; Kahler, R.; Williams, J.C.; Dryden, D.M.; et al. Defect Engineering at the Al2O3/(010) β-Ga2O3 Interface via Surface Treatments and Forming Gas Post-Deposition Anneals. IEEE Trans. Electron. Devices 2022, 69, 5656–5663. [Google Scholar] [CrossRef]
  229. Zeng, K.; Singisetti, U. Temperature dependent quasi-static capacitance-voltage characterization of SiO2/β-Ga2O3 interface on different crystal orientations. Appl. Phys. Lett. 2017, 111, 122108. [Google Scholar] [CrossRef]
  230. Biswas, D.; Joishi, C.; Biswas, J.; Thakar, K.; Rajan, S.; Lodha, S. Enhanced n-type β-Ga2O3 (201) gate stack performance using Al2O3/SiO2 bi-layer dielectric. Appl. Phys. Lett. 2019, 114, 212106. [Google Scholar] [CrossRef]
  231. Zhang, J.; Dong, P.; Dang, K.; Zhang, Y.; Yan, Q.; Xiang, H.; Su, J.; Liu, Z.; Si, M.; Gao, J.; et al. Ultra-wide bandgap semiconductor Ga2O3 power diodes. Nat. Commun. 2022, 13, 3900. [Google Scholar] [CrossRef] [PubMed]
  232. Gong, H.H.; Chen, X.H.; Xu, Y.; Ren, F.-F.; Gu, S.L.; Ye, J.D. A 1.86-kV double-layered NiO/β-Ga2O3 vertical p–n heterojunction diode. Appl. Phys. Lett. 2020, 117, 022104. [Google Scholar] [CrossRef]
  233. Lu, X.; Zhou, X.; Jiang, H.; Ng, K.W.; Chen, Z.; Pei, Y.; Lau, K.M.; Wang, G. 1-kV Sputtered p-NiO/n-Ga2O3 Heterojunction Diodes with an Ultra-Low Leakage Current Below 1 μ A/cm2. IEEE Electron. Device Lett. 2020, 41, 449–452. [Google Scholar] [CrossRef]
  234. Zhou, X.; Liu, Q.; Hao, W.; Xu, G.; Long, S. Normally-off β-Ga2O3 Power Heterojunction Field-Effect-Transistor Realized by p-NiO and Recessed-Gate. In Proceedings of the 2022 IEEE 34th International Symposium on Power Semiconductor Devices and ICs (ISPSD), Vancouver, BC, Canada, 22–25 May 2022; pp. 101–104. [Google Scholar] [CrossRef]
  235. Lei, W.; Dang, K.; Zhou, H.; Zhang, J.; Wang, C.; Xin, Q.; Alghamdi, S.; Liu, Z.; Feng, Q.; Sun, R.; et al. Proposal and Simulation of Ga2O3 MOSFET with PN Heterojunction Structure for High-Performance E-Mode Operation. IEEE Trans. Electron. Devices 2022, 69, 3617–3622. [Google Scholar] [CrossRef]
  236. Wang, Y.; Gong, H.; Jia, X.; Han, G.; Ye, J.; Liu, Y.; Hu, H.; Ou, X.; Ma, X.; Hao, Y. First Demonstration of RESURF and Superjunction β-Ga2O3 MOSFETs with p-NiO/n- Ga2O3 Junctions. In Proceedings of the 2021 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 11–16 December 2021; pp. 36.6.1–36.6.4. [Google Scholar] [CrossRef]
  237. Wang, Y.; Gong, H.; Jia, X.; Ye, J.; Liu, Y.; Hu, H.; Ou, X.; Ma, X.; Zhang, R.; Hao, Y.; et al. Demonstration of β-Ga2O3 Superjunction-Equivalent MOSFETs. IEEE Trans. Electron. Devices 2022, 69, 2203–2209. [Google Scholar] [CrossRef]
  238. Meshram, A.D.; Sengupta, A.; Bhattacharyya, T.K.; Dutta, G. Normally-Off β-(AlxGa1−x)2O3/Ga2O3 Modulation-Doped Field-Effect Transistors with p-GaN Gate: Proposal and Investigation. IEEE Trans. Electron. Devices 2023, 70, 454–460. [Google Scholar] [CrossRef]
  239. Yi, B.; Zhang, S.; Zhang, Z.; Cheng, J.; Huang, H.; Kong, M.; Yang, H. Analytical model and simulation study of a novel enhancement-mode Ga2O3 MISFET realized by p-GaN gate. Semicond. Sci. Technol. 2023, 38, 095003. [Google Scholar] [CrossRef]
  240. Budde, M.; Splith, D.; Mazzolini, P.; Tahraoui, A.; Feldl, J.; Ramsteiner, M.; von Wenckstern, H.; Grundmann, M.; Bierwagen, O. SnO/β-Ga2O3 vertical pn heterojunction diodes. Appl. Phys. Lett. 2020, 117, 252106. [Google Scholar] [CrossRef]
  241. Xia, Z.; Wang, C.; Kalarickal, N.K.; Stemmer, S.; Rajan, S. Design of Transistors Using High-Permittivity Materials. IEEE Trans. Electron. Devices 2019, 66, 896–900. [Google Scholar] [CrossRef]
  242. Moser, N.A.; McCandless, J.P.; Crespo, A.; Leedy, K.D.; Green, A.J.; Heller, E.R.; Chabak, K.D.; Peixoto, N.; Jessen, G.H. High pulsed current density β-Ga2O3 MOSFETs verified by an analytical model corrected for interface charge. Appl. Phys. Lett. 2017, 110, 143505. [Google Scholar] [CrossRef]
  243. Xia, Z.; Chandrasekar, H.; Moore, W.; Wang, C.; Lee, A.J.; McGlone, J.; Kalarickal, N.K.; Arehart, A.; Ringel, S.; Yang, F.; et al. Metal/BaTiO3/β-Ga2O3 dielectric heterojunction diode with 5.7 MV/cm breakdown field. Appl. Phys. Lett. 2019, 115, 252104. [Google Scholar] [CrossRef]
  244. Dong, H.; Mu, W.; Hu, Y.; He, Q.; Fu, B.; Xue, H.; Qin, Y.; Jian, G.; Zhang, Y.; Long, S.; et al. C-V and J-V investigation of HfO2/Al2O3 bilayer dielectrics MOSCAPs on (100) β-Ga2O3. AIP Adv. 2018, 8, 065215. [Google Scholar] [CrossRef]
  245. Yang, J.Y.; Ma, J.; Lee, C.H.; Yoo, G. Polycrystalline/Amorphous HfO2 Bilayer Structure as a Gate Dielectric for β-Ga2O3 MOS Capacitors. IEEE Trans. Electron. Devices 2021, 68, 1011–1015. [Google Scholar] [CrossRef]
  246. Feng, Z.; Tian, X.; Li, Z.; Hu, Z.; Zhang, Y.; Kang, X.; Ning, J.; Zhang, Y.; Zhang, C.; Feng, Q.; et al. Normally-Off-β -Ga2O3 Power MOSFET with Ferroelectric Charge Storage Gate Stack Structure. IEEE Electron. Device Lett. 2020, 41, 333–336. [Google Scholar] [CrossRef]
  247. Wang, Z.; Chen, X.; Ren, F.-F.; Gu, S.; Ye, J. Deep-level defects in gallium oxide. J. Phys. D Appl. Phys. 2020, 54, 043002. [Google Scholar] [CrossRef]
  248. Jian, Z.; Mohanty, S.; Ahmadi, E. Deep UV-assisted capacitance–voltage characterization of post-deposition annealed Al2O3/β-Ga2O3 (001) MOSCAPs. Appl. Phys. Lett. 2020, 116, 242105. [Google Scholar] [CrossRef]
  249. Hirose, M.; Nabatame, T.; Irokawa, Y.; Maeda, E.; Ohi, A.; Ikeda, N.; Sang, L.; Koide, Y.; Kiyono, H. Interface characteristics of β-Ga2O3/Al2O3/Pt capacitors after postmetallization annealing. J. Vac. Sci. Technol. A 2021, 39, 012401. [Google Scholar] [CrossRef]
  250. Bae, H.; Noh, J.; Alghamdi, S.; Si, M.; Ye, P.D. Ultraviolet Light-Based Current–Voltage Method for Simultaneous Extraction of Donor- and Acceptor-like Interface Traps in β-Ga2O3 FETs. IEEE Electron. Device Lett. 2018, 39, 1708–1711. [Google Scholar] [CrossRef]
  251. Fregolent, M.; Brusaterra, E.; De Santi, C.; Tetzner, K.; Würfl, J.; Meneghesso, G.; Zanoni, E.; Meneghini, M. Logarithmic trapping and detrapping in β-Ga2O3 MOSFETs: Experimental analysis and modeling. Appl. Phys. Lett. 2022, 120, 163502. [Google Scholar] [CrossRef]
  252. Jiang, Z.; Wei, J.; Lv, Y.; Wei, Y.; Wang, Y.; Lu, J.; Liu, H.; Feng, Z.; Zhou, H.; Zhang, J.; et al. Nonuniform Mechanism for Positive and Negative Bias Stress Instability in β-Ga2O3 MOSFET. IEEE Trans. Electron. Devices 2022, 69, 5509–5515. [Google Scholar] [CrossRef]
  253. Jiang, Z.; Li, X.; Zhou, X.; Wei, Y.; Wei, J.; Xu, G.; Long, S.; Luo, X. Experimental investigation on the instability for NiO/β-Ga2O3 heterojunction-gate FETs under negative bias stress. J. Semicond. 2023, 44, 072803. [Google Scholar] [CrossRef]
  254. Liu, C.; Wang, Y.; Xu, W.; Jia, X.; Huang, S.; Li, Y.; Li, B.; Luo, Z.; Fang, C.; Liu, Y.; et al. Unique Bias Stress Instability of Heterogeneous Ga2O3-on-SiC MOSFET. IEEE Electron. Device Lett. 2023, 44, 1256–1259. [Google Scholar] [CrossRef]
  255. Binari, S.; Klein, P.; Kazior, T. Trapping effects in GaN and SiC microwave FETs. Proc. IEEE 2002, 90, 1048–1058. [Google Scholar] [CrossRef]
  256. Zhou, H.; Alghamdi, S.; Si, M.; Qiu, G.; Ye, P.D. Al2O3/β-Ga2O3(-201) Interface Improvement Through Piranha Pretreatment and Postdeposition Annealing. IEEE Electron. Device Lett. 2016, 37, 1411–1414. [Google Scholar] [CrossRef]
  257. Feng, B.; He, T.; He, G.; Zhang, X.; Wu, Y.; Chen, X.; Li, Z.; Zhang, X.; Jia, Z.; Niu, G.; et al. Reduction of MOS interfacial states between β-Ga2O3 and Al2O3 insulator by self-reaction etching with Ga flux. Appl. Phys. Lett. 2021, 118, 181602. [Google Scholar] [CrossRef]
  258. Qin, Y.; Wang, Z.; Sasaki, K.; Ye, J.; Zhang, Y. Recent progress of Ga2O3 power technology: Large-area devices, packaging and applications. Jpn. J. Appl. Phys. 2023, 62, SF0801. [Google Scholar] [CrossRef]
  259. Galazka, Z. β-Ga2O3 for wide-bandgap electronics and optoelectronics. Semicond. Sci. Technol. 2018, 33, 113001. [Google Scholar] [CrossRef]
  260. Hou, X.; Zou, Y.; Ding, M.; Qin, Y.; Zhang, Z.; Ma, X.; Tan, P.; Yu, S.; Zhou, X.; Zhao, X.; et al. Review of polymorphous Ga2O3materials and their solar-blind photodetector applications. J. Phys. D Appl. Phys. 2020, 54, 043001. [Google Scholar] [CrossRef]
  261. Zhou, J.; Chen, H.; Fu, K.; Zhao, Y. Gallium oxide-based optical nonlinear effects and photonics devices. J. Mater. Res. 2021, 36, 4832–4845. [Google Scholar] [CrossRef]
  262. Ricci, F.; Boschi, F.; Baraldi, A.; Filippetti, A.; Higashiwaki, M.; Kuramata, A.; Fiorentini, V.; Fornari, R. Theoretical and experimental investigation of optical absorption anisotropy in β-Ga2O3. J. Phys. Condens. Matter. 2016, 28, 224005. [Google Scholar] [CrossRef]
  263. Bin Cho, J.; Jung, G.; Kim, K.; Kim, J.; Hong, S.-K.; Song, J.-H.; Jang, J.I. Highly Asymmetric Optical Properties of β-Ga2O3 as Probed by Linear and Nonlinear Optical Excitation Spectroscopy. J. Phys. Chem. C 2021, 125, 1432–1440. [Google Scholar] [CrossRef]
  264. Caddemi, A.; Cardillo, E.; Patanè, S.; Triolo, C. Light Exposure Effects on the DC Kink of AlGaN/GaN HEMTs. Electronics 2019, 8, 698. [Google Scholar] [CrossRef]
  265. Caddemi, A.; Cardillo, E.; Salvo, G.; Patanè, S. Microwave effects of UV light exposure of a GaN HEMT: Measurements and model extraction. Microelectron. Reliab. 2016, 65, 310–317. [Google Scholar] [CrossRef]
  266. Leach, J.H.; Metzger, R.; Preble, E.A.; Evans, K.R. High voltage bulk GaN-based photoconductive switches for pulsed power applications. Gallium Nitride Mater. Devices VIII 2013, 8625, 294–300. [Google Scholar]
  267. Mazumder, S.K. An Overview of Photonic Power Electronic Devices. IEEE Trans. Power Electron. 2016, 31, 6562–6574. [Google Scholar] [CrossRef]
Figure 1. (a) β-Ga2O3 unit cell. Reproduced from [22]. © IOP Publishing. Reproduced with permission. All rights reserved. (b) β-Ga2O3 band diagram. Reprinted with permission from [19]. Copyright 2017 by the American Physical Society.
Figure 1. (a) β-Ga2O3 unit cell. Reproduced from [22]. © IOP Publishing. Reproduced with permission. All rights reserved. (b) β-Ga2O3 band diagram. Reprinted with permission from [19]. Copyright 2017 by the American Physical Society.
Materials 16 07693 g001
Figure 2. Electron mobility vs. carrier concentration in β-Ga2O3 for Si, Sn, and Ge dopants in layers grown using various crystal and thin film techniques. Adapted with permission from Chen et al. [31] © 2023 John Wiley & Sons, Ltd.
Figure 2. Electron mobility vs. carrier concentration in β-Ga2O3 for Si, Sn, and Ge dopants in layers grown using various crystal and thin film techniques. Adapted with permission from Chen et al. [31] © 2023 John Wiley & Sons, Ltd.
Materials 16 07693 g002
Figure 3. (a) First MESFET reported in 2013. Reproduced from [101], with the permission of AIP Publishing. (b) Shutter pulsing scheme and doping variation showing alternating UID and uniformly doped layers. Reproduced from [102]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (c) Highly scaled T-gate delta-doped MESFET with high cutoff and maximum frequencies. © (2019) IEEE. Reprinted with permission from [88]. (d) Tri-gate MESFETs with low-temp/high-temp grown layers resulting in ultra-high mobilities and negligible I-V hysteresis. © (2022) IEEE. Reprinted with permission from [63].
Figure 3. (a) First MESFET reported in 2013. Reproduced from [101], with the permission of AIP Publishing. (b) Shutter pulsing scheme and doping variation showing alternating UID and uniformly doped layers. Reproduced from [102]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (c) Highly scaled T-gate delta-doped MESFET with high cutoff and maximum frequencies. © (2019) IEEE. Reprinted with permission from [88]. (d) Tri-gate MESFETs with low-temp/high-temp grown layers resulting in ultra-high mobilities and negligible I-V hysteresis. © (2022) IEEE. Reprinted with permission from [63].
Materials 16 07693 g003
Figure 4. (a) SAG FET using refractory metal gate W and Si ion implantation for self-alignment with an LSG of 0 µm. The RF Pout, GT, and PAE as a function of input power at 1 GHz are plotted. Reproduced from [89]. CC BY 4.0. (b) SAG process enabled by using an n++ grown cap layer as opposed to ion implantation. High gate leakage and low on/off ratio are indicative of a leaky dielectric due to its deposition or residual Ga droplets at the interface. Reproduced from [61], with the permission of AIP Publishing.
Figure 4. (a) SAG FET using refractory metal gate W and Si ion implantation for self-alignment with an LSG of 0 µm. The RF Pout, GT, and PAE as a function of input power at 1 GHz are plotted. Reproduced from [89]. CC BY 4.0. (b) SAG process enabled by using an n++ grown cap layer as opposed to ion implantation. High gate leakage and low on/off ratio are indicative of a leaky dielectric due to its deposition or residual Ga droplets at the interface. Reproduced from [61], with the permission of AIP Publishing.
Materials 16 07693 g004
Figure 6. (a) TCAD model of a recessed-gate FET studying variations in fermi level, Vth, and electron concentration with channel thickness. Reproduced from [115]; licensed under a Creative Commons Attribution (CC BY) license. (b) TCAD model of a recessed-gate FET studying variations in Vth and current density with doping and recess depth. Reprinted from [116], Copyright (2023), with permission from Elsevier. (c) A novel recessed-gate FET design with different body and drift layers recessing fully through the drift layer. Body-doping effects on E-/D-mode operation as well as a 2D cross-section of band bending through the body layer at low dopings. Reproduced from [117]. CC BY 4.0.
Figure 6. (a) TCAD model of a recessed-gate FET studying variations in fermi level, Vth, and electron concentration with channel thickness. Reproduced from [115]; licensed under a Creative Commons Attribution (CC BY) license. (b) TCAD model of a recessed-gate FET studying variations in Vth and current density with doping and recess depth. Reprinted from [116], Copyright (2023), with permission from Elsevier. (c) A novel recessed-gate FET design with different body and drift layers recessing fully through the drift layer. Body-doping effects on E-/D-mode operation as well as a 2D cross-section of band bending through the body layer at low dopings. Reproduced from [117]. CC BY 4.0.
Materials 16 07693 g006
Figure 7. (a) Lateral E-mode FinFETs and transfer curves with observed substrate conduction due to free carriers at the semi-insulating substrate. Reproduced from [113]. CC BY 4.0. (b) Cross-section of vertical multi-fin FETs, I-V curves of a single-fin FET showing significant improvement through PDA, and multi-fin FET I-V curves. © (2019) IEEE. Reprinted with permission from [81]. (c) Nitrogen doping mitigating the Vth dependence on Wfin and maintaining E-mode operation for large Wfin. Reproduced from [121]. © The Japan Society of Applied Physics. Reproduced by permission of IOP Publishing Ltd. All rights reserved.
Figure 7. (a) Lateral E-mode FinFETs and transfer curves with observed substrate conduction due to free carriers at the semi-insulating substrate. Reproduced from [113]. CC BY 4.0. (b) Cross-section of vertical multi-fin FETs, I-V curves of a single-fin FET showing significant improvement through PDA, and multi-fin FET I-V curves. © (2019) IEEE. Reprinted with permission from [81]. (c) Nitrogen doping mitigating the Vth dependence on Wfin and maintaining E-mode operation for large Wfin. Reproduced from [121]. © The Japan Society of Applied Physics. Reproduced by permission of IOP Publishing Ltd. All rights reserved.
Materials 16 07693 g007
Figure 8. (a) FinFETs fabricated via MacEtch with process and TEM image. I-V curves as well as SS and hysteresis dependence on channel angle relative to [102] are shown. Channels perpendicular to [102] show the best performance. Reproduced from [125], with the permission of AIP Publishing. (b) Temperature dependence of Vth, hysteresis, on/off ratio, and SS in MacEtch FinFETs indicating thermal degradation of the interface and/or dielectric. Reprinted from [127], with the permission of AIP Publishing.
Figure 8. (a) FinFETs fabricated via MacEtch with process and TEM image. I-V curves as well as SS and hysteresis dependence on channel angle relative to [102] are shown. Channels perpendicular to [102] show the best performance. Reproduced from [125], with the permission of AIP Publishing. (b) Temperature dependence of Vth, hysteresis, on/off ratio, and SS in MacEtch FinFETs indicating thermal degradation of the interface and/or dielectric. Reprinted from [127], with the permission of AIP Publishing.
Materials 16 07693 g008
Figure 10. (a) Lateral MOSFET with SFP and simulated TCAD electric field profiles clearly showing field spreading and reduction in overall peak field value. © (2019) IEEE. Reprinted with permission from [137]. (b) FET with SFP and T-gate structure, breakdown I-V, and benchmark plot. © (2020) IEEE. Reprinted with permission from [65]. (c) FET with SFP, T-gate, oxygen annealing (OA), and B-implantation for device isolation. The blue/red lines represent an LGD of 40/100 µm and solid/open symbols represent without/with SFP. A breakdown of 10 kV is observed. © (2023) IEEE. Reprinted with permission from [59].
Figure 10. (a) Lateral MOSFET with SFP and simulated TCAD electric field profiles clearly showing field spreading and reduction in overall peak field value. © (2019) IEEE. Reprinted with permission from [137]. (b) FET with SFP and T-gate structure, breakdown I-V, and benchmark plot. © (2020) IEEE. Reprinted with permission from [65]. (c) FET with SFP, T-gate, oxygen annealing (OA), and B-implantation for device isolation. The blue/red lines represent an LGD of 40/100 µm and solid/open symbols represent without/with SFP. A breakdown of 10 kV is observed. © (2023) IEEE. Reprinted with permission from [59].
Materials 16 07693 g010
Figure 11. A variety of RF T-gate FETs are shown. (a) FET incorporates a recessed-gate architecture. Reprinted with permission from [139]. (b) FET uses air as the FP dielectric. Reprinted from [140], with the permission of AIP Publishing. (c) FET uses both an air FP dielectric and an ultra-thin implanted channel [94]. (d) FET incorporating Al2O3 surface and gate metal passivation. Reprinted from [66]. CC BY-NC-ND 4.0. (e) T-gate RF FET with SiNx passivation with highest-reported fmax and high breakdown field of 48 GHz and 5.4 MV cm−1, respectively. Reproduced from [93], with the permission of AIP Publishing.
Figure 11. A variety of RF T-gate FETs are shown. (a) FET incorporates a recessed-gate architecture. Reprinted with permission from [139]. (b) FET uses air as the FP dielectric. Reprinted from [140], with the permission of AIP Publishing. (c) FET uses both an air FP dielectric and an ultra-thin implanted channel [94]. (d) FET incorporating Al2O3 surface and gate metal passivation. Reprinted from [66]. CC BY-NC-ND 4.0. (e) T-gate RF FET with SiNx passivation with highest-reported fmax and high breakdown field of 48 GHz and 5.4 MV cm−1, respectively. Reproduced from [93], with the permission of AIP Publishing.
Materials 16 07693 g011
Figure 12. (a) SOI FET with Vth modulation using constant back-gate voltage to accumulate or deplete the channel, while the top gate controls the device. © (2019) IEEE. Reprinted with permission from [152]. (b) SOI FET obtaining a record mobility of 191 cm2 V−1 s−1 using a floating p-SnO layer in the channel [85]. (c) SOI FET with a p++ back gate and doping of 8 × 1018 cm−3, measuring record currents of 1.5 A mm−1. Reprinted from [68], with the permission of AIP Publishing. (d) TMD high Schottky barrier gate with near-ideal SS of 61 mV dec−1 when using TaS2. Reprinted with permission from Kim et al. [160] © 2023 Wiley-VCH GmbH.
Figure 12. (a) SOI FET with Vth modulation using constant back-gate voltage to accumulate or deplete the channel, while the top gate controls the device. © (2019) IEEE. Reprinted with permission from [152]. (b) SOI FET obtaining a record mobility of 191 cm2 V−1 s−1 using a floating p-SnO layer in the channel [85]. (c) SOI FET with a p++ back gate and doping of 8 × 1018 cm−3, measuring record currents of 1.5 A mm−1. Reprinted from [68], with the permission of AIP Publishing. (d) TMD high Schottky barrier gate with near-ideal SS of 61 mV dec−1 when using TaS2. Reprinted with permission from Kim et al. [160] © 2023 Wiley-VCH GmbH.
Materials 16 07693 g012
Figure 13. TCAD simulations of novel, not yet realized β-Ga2O3 FETs. (a) An FP self-aligned trench vertical gate with variable Vth based on gate trench thickness into the UID layer, tUID. © (2023) IEEE. Reprinted with permission from [166]. (b) SFP with air gap dielectric to better mitigate electric fields at device edges. Reprinted with permission from [168], Copyright Elsevier (2022). (c) GAA FET with 2DEG to improve Pout and fT. Reprinted with permission from [169], Copyright Elsevier (2021). (d) Band diagram and current gain of npn HBT using p-CuO2, but limited by p-oxide bandgap, interface traps, and CBO between emitter and base. Reproduced from [170]. © IOP Publishing. Reproduced with permission. All rights reserved.
Figure 13. TCAD simulations of novel, not yet realized β-Ga2O3 FETs. (a) An FP self-aligned trench vertical gate with variable Vth based on gate trench thickness into the UID layer, tUID. © (2023) IEEE. Reprinted with permission from [166]. (b) SFP with air gap dielectric to better mitigate electric fields at device edges. Reprinted with permission from [168], Copyright Elsevier (2022). (c) GAA FET with 2DEG to improve Pout and fT. Reprinted with permission from [169], Copyright Elsevier (2021). (d) Band diagram and current gain of npn HBT using p-CuO2, but limited by p-oxide bandgap, interface traps, and CBO between emitter and base. Reproduced from [170]. © IOP Publishing. Reproduced with permission. All rights reserved.
Materials 16 07693 g013
Figure 14. (a) Cross-section and transfer I-V curve of E-mode CAVET with CBL surrounding source. Copyright (2022) IEEE. Reprinted with permission from [83]. (b) E-/D-mode CAVETs implemented via nch variation [174]. (c) I-V output curves of CAVETs with different Lap showing diode-like turn-on behavior. Reproduced from [176], with the permission of AIP Publishing. (d) U-MOSFET with CBL. Reproduced from [84], with the permission of AIP Publishing.
Figure 14. (a) Cross-section and transfer I-V curve of E-mode CAVET with CBL surrounding source. Copyright (2022) IEEE. Reprinted with permission from [83]. (b) E-/D-mode CAVETs implemented via nch variation [174]. (c) I-V output curves of CAVETs with different Lap showing diode-like turn-on behavior. Reproduced from [176], with the permission of AIP Publishing. (d) U-MOSFET with CBL. Reproduced from [84], with the permission of AIP Publishing.
Materials 16 07693 g014
Figure 15. (a) Cross-section band diagram showing 2DEG and measured mobility up to 180 cm2 V−1 s−1 at room temperature of an AlGO/GO MODFET using delta doping. Reproduced from [187], with the permission of AIP Publishing. (b) Cross-section of a double-heterostructure MODFET. Reproduced from [189], with the permission of AIP Publishing. (c) Cross-section of a heterostructure FET. Reproduced from [96], with the permission of AIP Publishing.
Figure 15. (a) Cross-section band diagram showing 2DEG and measured mobility up to 180 cm2 V−1 s−1 at room temperature of an AlGO/GO MODFET using delta doping. Reproduced from [187], with the permission of AIP Publishing. (b) Cross-section of a double-heterostructure MODFET. Reproduced from [189], with the permission of AIP Publishing. (c) Cross-section of a heterostructure FET. Reproduced from [96], with the permission of AIP Publishing.
Materials 16 07693 g015
Figure 21. (a) Si concentration peak at the substrate/epi interface. Reproduced from [216]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (b) Negative effects of impurities from semi-insulating substrates and back depletion on mobility, charge density, and current dispersion. Reproduced from [103], with the permission of AIP Publishing. (c) Mg delta doping at substrate/epi interface to compensate Si impurities and reduce leakage current. Reproduced from [106], with the permission of AIP Publishing. (d) Buffer traps at EC—0.77 eV only for a 100 nm buffer, and at EC—0.70 eV for both a 100 nm and 600 nm buffer using CID-DLTS. Reproduced from [105], with the permission of AIP Publishing.
Figure 21. (a) Si concentration peak at the substrate/epi interface. Reproduced from [216]. © The Japan Society of Applied Physics. Reproduced with the permission of IOP Publishing Ltd. All rights reserved. (b) Negative effects of impurities from semi-insulating substrates and back depletion on mobility, charge density, and current dispersion. Reproduced from [103], with the permission of AIP Publishing. (c) Mg delta doping at substrate/epi interface to compensate Si impurities and reduce leakage current. Reproduced from [106], with the permission of AIP Publishing. (d) Buffer traps at EC—0.77 eV only for a 100 nm buffer, and at EC—0.70 eV for both a 100 nm and 600 nm buffer using CID-DLTS. Reproduced from [105], with the permission of AIP Publishing.
Materials 16 07693 g021
Figure 22. (a) Effects of PDA and PMA on Vfb and Dit. PMA shows improvement in both reducing fixed charge and shallow Dit, while having little effect on deep Dit. Reproduced from [249], with the permission of AVS: Science & Technology of Materials, Interfaces, and Processing. (b) First and second C-V sweeps of MOSCAPs with different surface cleans. SPB with FG-PDA introduces the least interface defects relative to the others. Reprinted with permission from [228]. (c) PCV comparing SRE to TMAH for removing ICP damage. Reproduced from [257], with the permission of AIP Publishing.
Figure 22. (a) Effects of PDA and PMA on Vfb and Dit. PMA shows improvement in both reducing fixed charge and shallow Dit, while having little effect on deep Dit. Reproduced from [249], with the permission of AVS: Science & Technology of Materials, Interfaces, and Processing. (b) First and second C-V sweeps of MOSCAPs with different surface cleans. SPB with FG-PDA introduces the least interface defects relative to the others. Reprinted with permission from [228]. (c) PCV comparing SRE to TMAH for removing ICP damage. Reproduced from [257], with the permission of AIP Publishing.
Materials 16 07693 g022
Table 2. Performance comparison of D-mode high-power FETs.
Table 2. Performance comparison of D-mode high-power FETs.
Ref.FET DesignOn/OffID,max (mA mm−1)Vbr (V)Ebr (MV cm−1)Ron,sp (Ω cm2)µ (cm2 V−1 s−1)BFOM (MW cm−2)
[59]MESFET, T-gate + SFP, OA1063.310 k12.92NR>34.2
[60]Delta-doped MESFET w/GFP1071803152.3NR73118
[61]Delta-doped SAG103560NRNRNR65NR
[62]Recessed and T-gate109491.80 k1.820.9 m128155
[63]Tri-gate lateral FinFET10101871.13 k4.21.34 m184950
[64]Composite + SU8 GFP109407.16 k1.798.98NR5.71
[65]SFP, T-gate, Al2O3/HfO2 gate oxide1092301.40 k2.907.08 mNR277
[66]Scaled T-gate MESFET104602.45 k2.0817.3 m84347
[67]SOI on sapphire108232800NR7.41 m13786.3
[68]Back-gate SOI on SiO2/Si 10101500NRNRNRNRNR
[69]CAVET, N++ ion implant108420 A cm−225NR31.5 m140NR
[70]AlGO/GO w/GFP 108NR1.37 k0.86120 m10115.6
[71]SOI on AlN/Si1095801181.041.44 m82.99.70
[72]SiC/GO composite wafer108NR2.37 k1.2318.4 m94303
[73]SOI on DiamondNR980NRNRNRNRNR
[74]p-NiO gate oxide10104501.12 k2.483.19 mNR390
[75]p-NiO gate oxide10102822.15 k3.56.24 m130740
[76]p-NiO/SiO2 gate oxide1093001.32 k1.474.30 mNR405
[77]p-SnO gate oxide1061007501.93.15 m100178
[78]BTO (ε≈235) gate oxide105359 6401.51.08 m72376
[79]Al2O3/BTO gate oxide107220 8404.101.72 m85408
Table 3. Performance comparison of E-mode high-power FETs.
Table 3. Performance comparison of E-mode high-power FETs.
Ref.FET DesignOn/OffID,max (mA mm−1)Vbr (V)Ebr (MV cm−1)Ron,sp (Ω cm2)µ (cm2 V−1 s−1)BFOM (MW cm−2)
[80]Recessed gate109405050.8417.2 m10614.8
[81]Multi-fin vertical FET108230 A cm−22.66 kNR25.2 m40280
[82]SOI on SiO2/Si 1010450185220 Ω mm55.2NR
[83]Mg-diffused CAVET109150 A cm−272NRNR7.5NR
[84]Vertical U-trench w/CBL6.4 × 10411 A cm−2102NR1.48NR0.007
[76]p-NiO/SiO2 gate oxide108NR2.96 k0.985115 mNR76
[75]p-NiO gate oxide10743.21.98 k3.313.8 m140284
[85]Back-gate SOI on SiO2/Si
p-SnO on top
2.26 × 10614.1NRNRNR191NR
[86]SOI on SiO2/Si
HfO2 gate oxide
10511.1800.1682 m810.078
[87]Multi-stack gate: HZO/Al2O3/HfO2/Al2O310823.22.14 k3.4524 m97193
Table 4. Performance comparison of D-/E-mode RF FETs.
Table 4. Performance comparison of D-/E-mode RF FETs.
Ref.TypeStructureOn/OffID,max (mA mm−1)Vbr (V)Ebr (MV cm−1)µ (cm2 V−1 s−1)fT (GHz)fmax (GHz)Gp (dB)GT (dB)Pout (mW mm−1)PAE (%)fT Vbr (THz V)vsat (fT Lg 2π) (cm s−1)
[88]D-MT-gate delta-doped MESFET1082601501.07702716NRNRNRNR4.052.01 × 106
[89]D-MSAG108Pulsed ≈ 300NRNR74NRNRNR1371523.4NRNR
[90]D-MRecessed gate
SiO2 passivation
106150NRNR963.312.95.11.82306.3NR1.45 × 106
[91]D-MTri-gate FinFETNR88NRNRNR5.411.4NRNRNRNRNR1.19 × 106
[92]D-MSiO2 GFPNR58NRNRNRNRNR4.81NR13022.4NRNR
[93]D-MT-gate, SiNx passivation
SiO2 gate oxide
1.23 × 1052851925.4801148NRNRNRNR2.1122.45 × 106
[94]D-MT-gate, shallow ion-implanted channel1081651932.092329357NR11.2 dBm11.65.5972.73 × 106
[95]D-MOA, SiNx T-gate
Multi-stack gate oxide: Al2O3/HfO2
109200NRNR751.84.23.6NR4306.42NR1.13 × 106
[96]D-MAlGO/GO HFETNRPulsed ≈ 80NRNRNR1422NRNRNRNRNR1.76 × 106
[97]E-MAlGO/GO HFET1.55 × 10574231.35NR3037NRNRNRNRNR3.02 × 106
GaN and Diamond RF FETs
[98]D-MGaN HEMT1031000600.419001042058NR510043.66.249.80 × 106
[99]D-MGaN HEMT3 × 105Pulsed 130050NR142315630815NR2500707.85.89 × 106
[100]D-MDiamond HEMTNR5001210.811016.21712.2NR420021.50.753.51 × 106
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

Maimon, O.; Li, Q. Progress in Gallium Oxide Field-Effect Transistors for High-Power and RF Applications. Materials 2023, 16, 7693. https://doi.org/10.3390/ma16247693

AMA Style

Maimon O, Li Q. Progress in Gallium Oxide Field-Effect Transistors for High-Power and RF Applications. Materials. 2023; 16(24):7693. https://doi.org/10.3390/ma16247693

Chicago/Turabian Style

Maimon, Ory, and Qiliang Li. 2023. "Progress in Gallium Oxide Field-Effect Transistors for High-Power and RF Applications" Materials 16, no. 24: 7693. https://doi.org/10.3390/ma16247693

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