Routes to semi-stable phases of Ga 2O 3 are the subject of extended discussions based on the review of growth methods, growth conditions, and precursors in works that report semi-stable phases other than the thermally stable β phase. The focus here is on mist chemical vapor deposition because it has produced single-phase Ga 2O 3 of α, γ, and ε (or κ) in terms of the substrate materials, and features of this growth method for phase control are emphasized. Recent reports of phase control by other growth technology give a deeper understanding of how to determine and control the phases, increasing the opportunities to fully utilize the novel and unique properties of Ga 2O 3.

Gallium oxide (Ga 2O 3) semiconductor materials and devices are doubtless one of the hottest topics today in device-oriented applied-physics research because Ga 2O 3 is a new material whose fundamental properties have yet to be well elucidated, and Ga 2O 3-based devices are attractive for power electronics and deep ultraviolet optics that require a material, such as Ga 2 O 3, with an ultrawide bandgap ( 5 eV).1,2 Research on Ga 2O 3 was triggered by the development of highly crystalline bulk crystals grown by conventional melt-growth methods,3–5 which is similar to the history of GaAs- or InP-based devices and unlike that of SiC- and GaN-based devices. This has become motive force for the development of Ga 2O 3-based devices by homoepitaxial growth.

Ga 2O 3 has at least five polymorphous crystal phases [ α, β, γ, δ, and ε (or κ)], and the monoclinic β-gallia structure ( β-Ga 2O 3) is the most thermodynamically stable phase.6,7 Therefore, Ga 2O 3 bulk crystals are essentially composed of the β phase. At the outset, β-Ga 2O 3 bulk crystals had low-resistance n-type conductivity and optical transparency in the visible, attracting attention for use as substrates for GaN-based light-emitting diodes.8–11 

In contrast, the semiconducting properties of Ga 2O 3 became evident through the demonstration of transparent field-effect transistors (FETs) with tin-doped polycrystalline Ga 2O 3 channels fabricated on sapphire.12 Later, single-crystalline β-Ga 2O 3 thin films were grown by homoepitaxy on β-Ga 2O 3 substrates.13,14 Although the bandgap of β-Ga 2O 3 is far larger than ever reported for semiconductor-device materials, a series of publications followed that focused on its semiconducting properties and device applications followed. The earliest achievements include the formation of semi-insulating layers on n-type β-Ga 2O 3,15,16 the formation of Schottky and Ohmic contacts on β-Ga 2O 3,15,16 highly sensitive above-bandgap photoconductivity,16 and bandgap tuning with β (Al,Ga) 2O 3 alloys.17 In 2012 and 2013, the promise of β-Ga 2O 3 as the basis of new power devices was evinced by molecular-beam homoepitaxial growth of high-quality β-Ga 2O 3,18 doping-controlled conductivity,18 metal–semiconductor FETs,19 conductivity control by ion implantation,20 and metal-oxide-semiconductor FETs (MOSFETs).21 In the following years, world-wide research into β-Ga 2O 3 materials expanded significantly.22–30 

In addition to β-Ga 2O 3, other semi-stable phases of α, γ, δ, and ε (or κ) are available. We expect β-Ga 2O 3 layers homoepitaxially grown on β-Ga 2O 3 to be of the highest quality and to be the best candidates for device applications because bulk substrates of other phases do not exist. However, other phases of Ga 2O 3 are expected to have a variety of unique properties, but they must be heteroepitaxially grown on appropriate substrates. Nevertheless, heteroepitaxy is not always a fatal problem for device applications (consider, e.g., GaAs on Si, GaN on sapphire, GaN on SiC, and GaN on Si). Buffer layers and/or epitaxial layer overgrowth (ELO) contributed to overcoming various problems caused by heteroepitaxy, especially dislocation defects. Therefore, exploring Ga 2O 3 for semi-stable phases is a worthwhile challenge. Our goal is to fabricate single-crystal layers of a semi-stable phase, overcoming the natural tendency of Ga 2O 3 to prefer the β phase. Success in this endeavor should open the way to new science and offer new applications for the attractive material, that is, Ga 2O 3.

In 2008, we reported the growth on sapphire of corundum-structured α-Ga 2O 3, which has the same corundum structure without severe inclusions of other phases.31 This was blossomed in 2016 as the demonstration of a low-on-resistance Schottky barrier diode (SBD),32 which is now commercially available. α-Ga 2O 3 is advantageous because of its low cost, a wide bandgap compared to β-Ga 2O 3, flexibility of heterojunctions with other materials, and a wide range of bandgap tuning.33 Later, other semi-stable phases of γ34 and ε ( κ)35–37 were demonstrated. The evolution of a variety of semi-stable phases has widened the range of possible materials for a target application and proposes new and novel functions; here, we should explore how to control the phases and determine why a semi-stable phase predominates over the most thermodynamically stable phase.

Bosi et al.38 and Tak et al.39 published review papers on polymorphs of Ga 2O 3, emphasizing how they can be tailored in terms of growth methods and growth conditions. Our long-term research on mist chemical vapor deposition (mist CVD; details of this technology are discussed later) shows that mist CVD offers a wide growth window of semi-stable phases of Ga 2O 3 by permitting an appropriate choice of the substrate or underlying buffer layers, especially for corundum-structured α-Ga 2O 3 on sapphire.

In this paper, we briefly review the achievements of research on phase control of Ga 2O 3 and emphasize how mist CVD is suitable for this application because the Ga 2O 3 layers grown are effectively phase locked with the underlying crystal structure. Focusing on α-Ga 2O 3 on sapphire, we discuss the growth mechanism and basic properties, giving, as appropriate, perspectives of device applications of semi-stable α-Ga 2O 3.

Table I summarizes the basic parameters of Ga 2O 3 polymorphs, which are schematically illustrated in Fig. 1. Polymorphs other than the β phase have been detected in nanocrystals that rarely form under extreme environmental conditions (i.e., high temperature and high pressure), although recent research has demonstrated the possibility of growing single-phase thin films.

FIG. 1.

Polymorphs of Ga 2O 3.

FIG. 1.

Polymorphs of Ga 2O 3.

Close modal
TABLE I.

Summary of basic parameters of Ga2O3 polymorphs.26,30 For the structure ɛ-Ga2O3, the earlier research suggests a hexagonal (P63 mc) structure, but now, many researchers argue that the structure is orthorhombic (Pna21).40,41 Since the orthorhombic structure is known for Al2O3 and is called κ-Al2O3,42  ɛ-Ga2O3 is sometimes called κ-Ga2O3. The basic table structure and the data are taken with permission from Pearton et al., Appl. Phys. Rev. 5, 011301 (2018). Copyright 2018 AIP Publishing, LLC; Chen et al., Photonics Res. 7, 381 (2019). Copyright 2019 Chinese Laser Press Publishing.

Polymorph structure and space groupBandgap (eV)30 Lattice parameters (Å)
α Rhombohedral, corundum R 3 ¯ c 5.3, 5.25 a, b = 4.98 − 5.04, c = 13.4 − 13.6 
β Monoclinic C2/m 4.4–5.0 a = 12.12 − 12.34, b = 3.03 − 3.04, c = 5.80 − 5.87 
γ Cubic, defective spinel F D 3 ¯ m 4.4 (indirect), 5.0 (direct) a = 8.24 − 8.30 
δ Possibly bixbyite  a = 9.4 − 10.0 
ɛ (κOrthorhombic Pna21 4.5 (indirect), 5.0 (direct) a = 5.06 − 5.12, b = 8.69 − 8.79, c = 9.3 − 9.4 
Polymorph structure and space groupBandgap (eV)30 Lattice parameters (Å)
α Rhombohedral, corundum R 3 ¯ c 5.3, 5.25 a, b = 4.98 − 5.04, c = 13.4 − 13.6 
β Monoclinic C2/m 4.4–5.0 a = 12.12 − 12.34, b = 3.03 − 3.04, c = 5.80 − 5.87 
γ Cubic, defective spinel F D 3 ¯ m 4.4 (indirect), 5.0 (direct) a = 8.24 − 8.30 
δ Possibly bixbyite  a = 9.4 − 10.0 
ɛ (κOrthorhombic Pna21 4.5 (indirect), 5.0 (direct) a = 5.06 − 5.12, b = 8.69 − 8.79, c = 9.3 − 9.4 

Most investigations have studied the corundum α phase since well-aligned α-Ga 2O 3 films were grown epitaxially on sapphire substrates, as evinced by an x-ray 0006 diffraction ω-scan rocking curve full-width at half maximum as small as 60 arc sec.31 The growth of the semi-stable corundum α phase is attributed to the matching of the crystal structure to the sapphire substrate, which has the corundum structure ( α-Al 2O 3). Previous work shows that the matching of the crystal structure between a substrate and an epilayer is one of the basic requirements for the epitaxial growth of a semi-stable material, as in the case of cubic GaN on GaAs,43–46 cubic ZnCdS on GaAs,47 and wurzite MgZnO on ZnO.48 In contrast, on a corundum-structured sapphire substrate, the growth of both α and β-Ga 2O 3 has been reported, with more publications reporting the growth of β-Ga 2O 349–53 than α-Ga 2O 3.32,54–61 Table II lists publications reporting growth methods and conditions for α or β polymorphs for the Ga 2O 3 layer that preferably grows on sapphire substrates. α-Ga 2O 3 is grown at low temperature by mist CVD, halide vapor phase epitaxy (HVPE), and atomic layer deposition (ALD), but not by laser molecular-beam epitaxy (MBE). Actually, Ga 2O 3 grown on sapphire by MBE (Table II lists plasma-assisted MBE) or metal organic CVD (MOCVD) tends to form the β phase49–53 or mixed α and β phases.62–64 Some papers report the growth of either the β or ε phase by MOCVD depending on the growth temperature or the growth pressure.65,66 Thin layers of α-Ga 2O 3 are grown63 under limited growth conditions62,68 or for specific substrate orientations.69 It is, at this stage, difficult to grow homogeneous thick α-Ga 2O 3 layer on sapphire by MBE and MOCVD.

TABLE II.

Publications reporting the growth of α or β-Ga2O3 on sapphire substrates in terms of the growth methods and conditions. PLD, pulsed laser deposition; MBE, molecular-beam epitaxy; MOCVD, metalorganic chemical vapor deposition; HVPE, halide vapor phase epitaxy; CVD, chemical vapor deposition; ALD, atomic layer deposition.

β-Ga2O3 on sapphireα-Ga2O3 on sapphire
Growth methodTemperatureReferenceGrowth methodTemperatureReference
PLD 380–410 °C 49  Mist CVD 200–500 °C 31, 32, and 54–57  
MBE 500–1000 °C 50  HVPE 470–650 °C 58 and 59  
MOCVD 650–700 °C 51  ALD 265–475 °C 60  
Evaporation 600–800 °C 52  Laser MBE 750–850 °C 61  
HVPE 1050 °C 53     
β-Ga2O3 on sapphireα-Ga2O3 on sapphire
Growth methodTemperatureReferenceGrowth methodTemperatureReference
PLD 380–410 °C 49  Mist CVD 200–500 °C 31, 32, and 54–57  
MBE 500–1000 °C 50  HVPE 470–650 °C 58 and 59  
MOCVD 650–700 °C 51  ALD 265–475 °C 60  
Evaporation 600–800 °C 52  Laser MBE 750–850 °C 61  
HVPE 1050 °C 53     

The next question is how to grow α-Ga 2O 3 layers on sapphire or what determines the polymorph of Ga 2O 3 on sapphire. A hint appears in the papers reporting the variation of the crystal phases of Ga 2O 3 in terms of the growth methods, growth conditions, and substrate materials. Schewski et al.63 showed that very thin (three atomic layers) α-Ga 2O 3 layers were grown pseudomorphically on sapphire, and then the phase was changed to β during the growth processes of MOCVD, MBE, and PLD. Misfit dislocations were detected at the β-Ga 2O 3 α-Ga 2O 3 interface but not at the α-Ga 2O 3–sapphire interface. Akazawa70 demonstrated that α-Ga 2O 3 grows on a-plane sapphire at 600  °C and undergoes solid-phase crystallization at 800  °C, which is higher than the transition temperature from α to β,6 suggesting stable-phase locking from the corundum-structured substrate. Sun et al.62 and Yao et al.68 reported the formation of α-Ga 2O 3 rather than β-Ga 2O 3 when HCl is in the growth atmosphere. These results suggest that the essential feature for the growth of a semi-stable phase of α-Ga 2O 3 is not only the use of a corundum-structured substrate but also the growth mechanisms.

As evinced by Table II, the method of mist CVD offers a broad opportunity to form α-Ga 2O 3 on sapphire substrates. In the remainder of this paper, we discuss the growth mechanism at work in mist CVD suitable for the growth of semi-stable phases of Ga 2O 3 due to phase matching with the substrates, not only for α-Ga 2O 3 but also for other phases.

Mist CVD is a simple, safe, and cost-effective technology for depositing oxide thin films.71  Figure 2 shows schematically an example of the growth system. The basic idea is to use mist particles containing source materials and formed by ultrasonic atomization of water or alcohol solution instead of organometallic sources as in MOCVD. The mist particles serve as a precursor for CVD. To grow Ga 2O 3 by MOCVD, trimethylgallium (TMGa) or triethylgallium (TEGa) is used as a gallium source, together with O 2 or N 2O as a oxygen source. With mist CVD, gallium acetylacetonate [Ga(III) 2,4-pentanedionate, Ga(acac) 3] or gallium chloride (GaCl 3) may serve as a starting source material. They are dissolved in water with a slight addition of hydrochloric acid (HCl), and the solution is atomized using an ultrasonic transducer. Mist particles formed by atomization are transferred by a carrier gas to a reactive area containing a heated substrate.

FIG. 2.

Schematic of a mist-CVD growth system.

FIG. 2.

Schematic of a mist-CVD growth system.

Close modal

This technology constitutes a non-vacuum-based deposition method of ferroelectric thin films72 or transparent conductors;73,74 for the latter, the technology was called “ultrasonic spray pyrolysis.” Applying this technology to epitaxial growth of device-quality single-crystalline semiconductors proved challenging, but the earlier work on ZnO achieved layer-by-layer growth75 and reasonable residual electron concentrations (on the order of 10 17 cm 3),76 which encouraged the development of Ga 2O 3 electronics based on this growth technology (which offers phase control).

α-Ga 2O 3 is a semi-stable phase of Ga 2O 3 crystals grown at high pressure (44 kbar) and high temperature (1000  °C) with a NaOH flux generated from β-Ga 2O 3 powders.77 Note that α-Ga 2O 3 thin films are grown on sapphire ( α-Al 2O 3) substrates at low temperatures and atmospheric pressure by mist CVD, overcoming the large lattice mismatch (4.8%) and thermodynamics. To elucidate the growth mechanism and determine how α-Ga 2O 3 thin films are realized, we investigate in detail the structure of the α-Ga 2O 3 α-Al 2O 3 crystal interface.

Figures 3(a) and 3(b) show cross-sectional transmission electron microscopy (TEM) images of the α-Ga 2O 3 α-Al 2O 3 interface observed along the [11 2 ¯0] and [1 1 ¯00] zone axes, respectively.33,78 A periodic structure with a period of 8.6 nm appears in Fig. 3(a) at the interface along the [1 1 ¯00] axis.

FIG. 3.

Cross-sectional TEM images of the α-Ga 2O 3/ α-Al 2O 3 interface viewed along the (a) [11 2 ¯0] and (b) [1 1 ¯00] axes.33,78 Reproduced with permission from Fujita et al., Jpn. J. Appl. Phys. 55, 1202A3 (2016). Copyright 2016 IOP Publishing; Kaneko et al., 51, 11PJ03 (2012). Copyright 2012 IOP Publishing.

FIG. 3.

Cross-sectional TEM images of the α-Ga 2O 3/ α-Al 2O 3 interface viewed along the (a) [11 2 ¯0] and (b) [1 1 ¯00] axes.33,78 Reproduced with permission from Fujita et al., Jpn. J. Appl. Phys. 55, 1202A3 (2016). Copyright 2016 IOP Publishing; Kaneko et al., 51, 11PJ03 (2012). Copyright 2012 IOP Publishing.

Close modal

Since the unit crystal cell of α-Ga 2O 3 and α-Al 2O 3 in the [1 1 ¯00] direction is 0.43 and 0.41 nm long, respectively, the length of 20 crystal cells (8.60 nm) of α-Ga 2O 3 corresponds to the length of 21 crystal cells (8.61 nm) of α-Al 2O 3.

Figure 3(b) confirms a similar periodic structure along the [11 2 ¯0] axis with a 5.0-nm-long unit crystal cell. The unit crystal cell along the [11 2 ¯0] direction of α-Ga 2O 3 and α-Al 2O 3 is 0.249 and 0.238 nm, respectively; therefore, 20 crystal cells (4.98 nm) of α-Ga 2O 3 correspond to 21 crystal cells (4.99 nm) of α-Al 2O 3. These results suggest that the α-Ga 2O 3 crystal grows on an α-Al 2O 3 substrate via semi-coherent growth or domain-matching epitaxy79 with 20 unit cells of α-Ga 2O 3 for 21 crystal cells of α-Al 2O 3.

Such semi-coherent growth is well known for zinc blende- or diamond-structured crystal systems, such as GaAs/Si,80 InAs/GaAs,81 and Ge/Si.82 Furthermore, for corundum-structured crystal systems, a similar periodic structure has been reported for an α-Fe 2O 3 crystal grown on an α-Al 2O 3 substrate,83 where the lattice mismatch is about 5%. Therefore, the semi-coherent growth or the lattice-relaxation mechanism of domain-matching epitaxy may also be characteristic of the growth of corundum-structured oxides on sapphire substrates. This is also evinced by x-ray reciprocal-lattice diffraction, which shows that α-Ga 2O 3 films are almost completely relaxed on α-Al 2O 3 substrates.33  Figure 4 shows schematically the structure at the interface.

FIG. 4.

Schematic model of domain epitaxy growth of α-Ga 2O 3 on an Al 2O 3 substrate by introducing dislocations at the interface.84 

FIG. 4.

Schematic model of domain epitaxy growth of α-Ga 2O 3 on an Al 2O 3 substrate by introducing dislocations at the interface.84 

Close modal

When using mist CVD, the crystal structure of a thin film grown on a substrate tends to well follow the crystal structure of the substrate, even if the result is a semi-stable structure for the film. Stable-phase bixbyite indium oxide ( c-In 2O 3) was grown on yttria-stabilized zirconia (YSZ) substrates (cubic crystal).85–87 Similarly, c-In 2O 3 was grown on YSZ by mist CVD.84 Conversely, indium oxide adopts the semi-stable phase of corundum ( α-In 2O 3) when deposited on α-Fe 2O 388 or α-Ga 2O 389 buffer layers or directly on α-Al 2O 3 substrates.90 In spite of being in a semi-stable phase, the thin films tend to be “phase locked” by the substrates when grown by mist CVD. In Secs. III C and III D, we discuss the mechanism for phase locking and show other semi-stable phases of Ga 2O 3 grown by mist CVD.

Figure 5 shows the transitions between crystal structures in the growth of Ga 2O 3 on c-plane sapphire. Figure 5(a) shows the result of mist CVD, where α-Ga 2O 3 is grown continuously in spite of being in the semi-stable phase. Figure 5(c) shows the results of Schewski et al. obtained by MOCVD, MBE, and PLD, where β-Ga 2O 3 is grown on an initially strained α-Ga 2O 3 layer with the epitaxial relationship β-Ga 2O 3( 2 ¯201)// α-Ga 2O 3(0001).63 This relationship stems from Ga atoms in the β-Ga 2O 3( 2 ¯201) plane forming a nonplanar hexagonal atomic arrangement, as shown in Fig. 5(b). In this case, the β-Ga 2O 3 layer only partially references the atomic arrangement during its initial growth. Figure 5(d) shows the general case of β-Ga 2O 3 grown on c-plane sapphire, where the epitaxial relationship is β-Ga 2O 3( 2 ¯01)//sapphire(0001), similar to Fig. 5(c).

FIG. 5.

Model for growth of Ga 2O 3 on sapphire. (a) α-Ga 2O 3 grows continuously. (c) The crystal phase changes from α to β on α-Ga 2O 3 initially grown on sapphire. (d) β-Ga 2O 3 grown on sapphire from the initial stage. (b) Top view of the β-Ga 2O 3 ( 2 ¯01) plane.91 Reproduced with permission from Fu et al., IEEE Trans. Electron Devices 65, 3507 (2018). Copyright 2018 IEEE.

FIG. 5.

Model for growth of Ga 2O 3 on sapphire. (a) α-Ga 2O 3 grows continuously. (c) The crystal phase changes from α to β on α-Ga 2O 3 initially grown on sapphire. (d) β-Ga 2O 3 grown on sapphire from the initial stage. (b) Top view of the β-Ga 2O 3 ( 2 ¯01) plane.91 Reproduced with permission from Fu et al., IEEE Trans. Electron Devices 65, 3507 (2018). Copyright 2018 IEEE.

Close modal

To grow the stable phase, such as with MOCVD or other vacuum growth techniques, atoms that have become supersaturated on the substrate surface naturally arrange themselves so that the most stable phase is prioritized over the atomic arrangement of the substrate. Therefore, it is reasonable that the crystal structure changes as the growth proceeds, as shown in Figs. 5(c) and 5(d). In contrast, when using mist CVD, the growth proceeds in the metastable phase, as shown in Fig. 5(a). One reason for this result is that the strong phase-locked mechanism faithfully reproduces the atomic arrangement of the surface despite the relatively low internal and/or migration energy due to the comparatively low growth temperature in mist CVD (see Table II). However, low growth temperature tends to obstruct epitaxial growth and the growth of high-quality crystals. Continuous growth of high-quality semi-stable phases requires phase-locked epitaxial growth, which is based more strongly on the atomic arrangement of the growth surface.

The source molecule used to grow Ga 2O 3 via mist CVD is currently Ga(acac) 3 or gallium halides, such as GaCl 3, gallium bromide (GaBr 3), and gallium iodide (GaI 3). Regardless of the source molecule used for source solutions, Ga ions are basically hydrated and some aqua ligands are replaced by hydroxide ions, acetylacetonate ligands, and/or halide ligands. Many investigations of Ga 2O 3 growth by mist CVD added HCl to the source solutions to enhance the solubility of source molecules. However, the HCl additive complicates not only the interpretation of the experimental results but also the discussion of the growth mechanism.

To investigate the mechanism exploited by mist CVD to produce phase-locked epitaxial growth, we prepared a Ga source solution with a molar concentration ratio of [ Ga 3 + ] : [ Cl ] = 1 : 3.92 Such a solution is safely obtained by dissolving a Ga metal in HCl. Most Ga ions in the solution form hexahydrate complexes [Ga(H 2O) 6 3 + 3Cl ].93,94 Acetylacetonation of Ga is controlled by adding acetylacetone (Hacac). The acetylacetonate (acac) anions coordinate around Ga atoms in their enolate form. The possible ligand coordination structures are Ga 3 +(H 2O) 6, Ga(acac) 2 +(H 2O) 4, Ga(acac) 2 +(H 2O) 2, and Ga(acac) 3. Ga(acac) 3 powder has very low solubility in water. Using this technique, acetylacetonated Ga aqueous solutions can be prepared with a definite concentration. Actually, to shorten the ligand-formation time, NH 4(acac) should be used instead of H(acac).

Figure 6 shows the growth rates of α-Ga 2O 3 thin films as a function of the acetylacetone concentration.92 The growth conditions were [ Ga 3 + ] = 0.02 mol/L, and the growth temperature was 400, 450, or 500  °C. The substrates were c-sapphire, and N 2 served as a process gas. The growth rate is almost zero when using a source solution without acetylacetone and increases with increasing acetylacetone concentration,92 with a tendency to saturate at all temperatures. These results imply that hexahydrate Ga ions do not contribute to film formation, whereas acetylacetonated Ga ions promote film formation on c-sapphire substrates. The concentration of acetylacetonated Ga ions, such as Ga(acac) 2 +, Ga(acac) 2 +, and Ga(acac) 3, can be calculated using the equilibrium constants already reported: K n = [ Ga ( acac ) n ( 3 n ) + ] / ( [ Ga 3 + ] [ acac ] n ), log K 1 = 9.4, log K 2 = 17.8, and log K 3 = 23.7.95 The total-concentration profile of acetylacetonated Ga ions follows a similar trend as the growth rates shown in Fig. 6. Additionally, the root mean square examined using the dynamic force mode of atomic force microscopy (AFM) clearly depends on the growth temperature. Higher growth temperature reduces not only the root mean square but also its dependence on [NH 4acac].92 From the viewpoint of growth mechanisms, these results indicate that the growth reactions occur at the film surface.

FIG. 6.

Growth rates of α-Ga 2O 3 thin films as functions of acetylacetone concentration in mist CVD.92 Reproduced with permission from Uno et al., Appl. Phys. Lett. 117, 052106 (2020). Copyright 2020 AIP Publishing LLC.

FIG. 6.

Growth rates of α-Ga 2O 3 thin films as functions of acetylacetone concentration in mist CVD.92 Reproduced with permission from Uno et al., Appl. Phys. Lett. 117, 052106 (2020). Copyright 2020 AIP Publishing LLC.

Close modal

The use of acetylacetonate complexes to enhance the formation of oxide films by molecularly designed dispersion has been reported previously.96–98 A similar enhancement mechanism should occur when α-Ga 2O 3 is grown by mist CVD. Figure 7 shows a possible mechanism for the growth of α-Ga 2O 3 by mist CVD. According to Tamba et al., the surface of α-Ga 2O 3 is terminated by Ga atoms.99 Under the conditions used for mist-CVD growth, Lewis acid-base interactions should lead to surface hydroxyl coverage by water molecules. Next, as shown by step 1 in Fig. 7, a gallium acetylacetonate complex approaches the surface and anchors to a surface hydroxyl by hydrogen bonding. Subsequently, in steps 2–4, a Ga–O bond forms by a ligand-exchange mechanism and the ligand leaves as acetylacetone (Hacac).

FIG. 7.

A possible growth mechanism of α-Ga 2O 3 by mist CVD.92 Reproduced with permission from Uno et al., Appl. Phys. Lett. 117, 052106 (2020). Copyright 2020 AIP Publishing LLC.

FIG. 7.

A possible growth mechanism of α-Ga 2O 3 by mist CVD.92 Reproduced with permission from Uno et al., Appl. Phys. Lett. 117, 052106 (2020). Copyright 2020 AIP Publishing LLC.

Close modal

To obtain experimental support for the proposed mechanism, α-Ga 2O 3 was grown with the introduction of isotopically labeled water (H 2 18O) using N 2 and O 2 as process gases. By examining the 18O/ 16O ratio by secondary-ion mass spectrometry (SIMS), the origin of oxygen atoms in the grown films can be investigated. The resulting 18O/ 16O ratio almost equals the ratio of water in the source solution, taking into account the H 2 18O naturally present in water. These results indicate that the oxygen atoms in the grown films originate from water. This result supports the proposed mechanism for phase-locked epitaxy of α-Ga 2O 3.

Note that homogeneous Ga 2O 3 thin films of other semi-stable phases, such as γ- and ε-Ga 2O 3, have been grown using mist CVD. Cubic-defective-spinel-structured γ-Ga 2O 3 is attractive due to its possible multifunctionality, for example, by alloying with ferrimagnetic γ-Fe 2O 3. It is possible to incorporate the magnetic element Fe into Ga 2O 3 to produce ferromagnetic materials.100 Oshima et al. applied mist CVD to grow γ-Ga 2O 3 on spinel (MgAl 2O 4) substrates, which forms a cubic structure with a 2% lattice mismatch with γ-Ga 2O 3 .34 As shown in Fig. 8, the x-ray diffraction (XRD) patterns are almost completely dominated by the peaks of γ-Ga 2O 3 except for those from the substrate if the growth temperature was 390 or 400  °C. At the higher growth temperature, the most stable β phase is included; the growth window for the γ phase is very narrow. Nevertheless, the quality of the resulting films is sufficient to determine indirect and direct bandgaps of 4.4 and 5.0 eV, respectively, and the refractive index in the visible is 2.0–2.1. The α phase has been grown on sapphire substrates at similar growth temperatures, which means that, under these growth conditions, the crystal structure follows that of the underlying substrate.

FIG. 8.

X-ray diffraction patterns for undoped Ga 2O 3 films grown at various growth temperatures on (100) spinel, showing the dominant growth of γ-Ga 2O 3.34 Reproduced with permission from Oshima et al., J. Cryst. Growth 359, 60 (2012). Copyright 2012 Elsevier.

FIG. 8.

X-ray diffraction patterns for undoped Ga 2O 3 films grown at various growth temperatures on (100) spinel, showing the dominant growth of γ-Ga 2O 3.34 Reproduced with permission from Oshima et al., J. Cryst. Growth 359, 60 (2012). Copyright 2012 Elsevier.

Close modal

ε-Ga 2O 3 forms a characteristic crystal structure. After Roy et al. suggested the ε phase,6 it was determined by numerous investigators to form the hexagonal structure ( P 6 3 m c).35,36,101,102 Later, through critical investigations of the crystal structure, ε-Ga 2O 3 was suggested to form the orthorhombic ( P n a 2 1) structure.41 Because κ-Al 2O 3 was already known to form an orthorhombic structure,42 some investigators called the Ga 2O 3 of orthorhombic structure “ κ-Ga 2O 3.” Nevertheless, we use the name ε-Ga 2O 3 below. ε-Ga 2O 3 is known to spontaneously polarize,102,103 which is preferred over ferroelectric devices and high-frequency transistors with a two-dimensional electron gas (2DEG).

Oshima et al. used HVPE to grow ε-Ga 2O 3 on GaN, AlN, and β-Ga 2O 3.35 Mist CVD was used to grow ε-Ga 2O 3 on cubic (111) MgO at 400–700  °C and on (111) YSZ at 600–700  °C.37 The indirect and direct bandgap energies are approximately 4.5 and 5.0 eV, respectively. Later, Oshima et al. used mist CVD to grow ε-Ga 2O 3 on hexagonal (0001) AlN,104 (0001) GaN,105 and cubic (111) SrTiO 3.105 When grown on cubic-structured substrates, it is reasonable that Ga 2O 3 preferably forms an orthorhombic structure because of the similarity of the crystal structures. Conversely, it is interesting to discuss why Ga 2O 3 forms an orthorhombic structure when grown on a hexagonal-structured substrate. Nishinaka et al.105 showed that unit cells of ε-Ga 2O 3 are arranged so that the atomic plane of each unit cell of ε-Ga 2O 3 follows an atomic plane of the hexagonal substrate. In other words, the growth of ε-Ga 2O 3 on hexagonal substrates results from Ga 2O 3 nucleating by following the crystal structure of the substrate. The surface-stabilization mechanism may be the same as that reported for ε-Ga 2O 3 grown by MOCVD.106 The ferroelectric properties of ε-Ga 2O 3,107 the growth of ε-(Al xGa 1 x) 2O 3,108 and the type-I band lineup of their heterointerface108 are evinced in Fig. 9. These results have promoted research into 2DEG transistors.

FIG. 9.

Schematic band offset between ε-(Al xGa 1 x) 2O 3 alloy films ( x = 0.105 and 0.395) and an ε-Ga 2O 3 film.108 Reproduced with permission from Tahara et al., Appl. Phys. Lett. 112, 152102 (2018). Copyright 2018 AIP Publishing LLC.

FIG. 9.

Schematic band offset between ε-(Al xGa 1 x) 2O 3 alloy films ( x = 0.105 and 0.395) and an ε-Ga 2O 3 film.108 Reproduced with permission from Tahara et al., Appl. Phys. Lett. 112, 152102 (2018). Copyright 2018 AIP Publishing LLC.

Close modal

Mist CVD is a useful technology for growing semi-stable Ga 2O 3. When α-Ga 2O 3 was grown by mist CVD on sapphire in 2008, no other growth technology could duplicate the result. However, the knowledge that α-Ga 2O 3 is one of the key materials for new device applications has renewed efforts to use other technology to grow α-Ga 2O 3 . This is also true for other polymorphs. Not surprisingly, the most-stable β-Ga 2O 3 is grown on a variety of substrates irrespective of the growth methods and conditions, although many researchers have searched for growth windows to grow semi-stable Ga 2O 3.

HVPE can be used to grow high-purity α-Ga 2O 3 and ε-Ga 2O 3 at high growth rates. The growth of phase-pure α-Ga 2O 3 on c-plane sapphire at relatively low temperatures (typically 500–600  °C) has been reported by several groups.58,59,109–111 The growth rate increases with increasing concentration of precursors, reaching over 100  μm/h while maintaining a specular surface.110 The structural quality of HVPE-grown α-Ga 2O 3 is similar to that of mist-CVD-grown films, but the tilt angle tends to be larger ( 100 arc sec), probably because of the nonoptimized nucleation process.58 The dislocation density measured by plan-view TEM is typically 10 10 cm 2. Epitaxial lateral overgrown α-Ga 2O 3 has also been grown via HVPE.111 

Oshima et al. grew (001)-oriented ε-Ga 2O 3 on (0001) GaN and (0001) AlN at a growth temperature of about 550  °C for both polymorphs, finding that the use of c-plane GaN and AlN substrates, which have hexagonal space groups, is vitally important to obtain ε-Ga 2O 3 epitaxial layers.35 This is similar to the case of mist CVD, as suggested by Nishinaka et al.105 Yao et al. grew (001) ε-Ga 2O 3 on c-plane sapphire at 650  °C, which succeeded likely because of the increased Ar flow rate or increased film coverage.68 In the study, high-resolution TEM imagery revealed the presence of a 10-nm-thick α-Ga 2O 3 interlayer between the ε-Ga 2O 3 and the substrate, with the epitaxial relationship of [ 1 ¯100] ε-Ga 2O 3//[11 2 ¯0] α-Ga 2O 3//[11 2 ¯0] sapphire.

α-Ga 2O 3 heteroepitaxial layers grown by MOCVD tend to be in mixed-phase form. In fact, α-Ga 2O 3 sometimes forms a thin interlayer between the substrate and an overlayer of a different phase. For example, Schewski et al. observed three monolayers of pseudomorphic α-Ga 2O 3 between β-Ga 2O 3 and a (0001) sapphire substrate for three different growth techniques, namely, MOCVD, PLD, and MBE.63 

Sue et al. grew three different phases of Ga 2O 3 ( α, β, and ε) films on c-plane sapphire using different flow rates of HCl in MOCVD.62 The results show a threefold increase in the growth rate of pure-phase β-Ga 2O 3 for the HCl flow rate of 5 sccm. Upon increasing the HCl flow rate to 10 sccm, the film transitions from pure-phase β-Ga 2O 3 to a mixture of β- and ε-Ga 2O 3. At a HCl flow rate of 30 sccm, which yields the highest growth rate of 1 μm/h, only ε-Ga 2O 3 forms. Furthermore, increasing the HCl flow rate produces a mixture of α and ε-Ga 2O 3. A combined growth and XRD study by Mezzadri et al. using water and trimethylgallium as reagents and palladium-purified H 2 as a carrier gas revealed the growth of ε-Ga 2O 3 (001) on c-plane sapphire with the Ga 2O 3 [10 1 ¯0] direction parallel to the sapphire direction [11 2 ¯0], yielding a lattice mismatch of about 4.1%.102 

To the best of our knowledge, no reports exist of pure γ-Ga 2O 3 films grown by MOCVD. A very thin (10–50 nm) γ-Ga 2O 3 nucleated layer was identified at the interface between an ε-Ga 2O 3 layer, grown by MOCVD, and a sapphire substrate, although the mechanism for the formation of this layer was not discussed.41 However, Bhuiyan et al. revealed the growth of pure γ-(Al xGa 1 x) 2O 3 on (010) β-Ga 2O 3 substrates with 0.37 < x < 1, which is likely due to increased strain due to Al incorporation inducing the rotation of β-phase (Al,Ga) 2O 3 domains, thereby promoting the formation of γ-phase (Al,Ga) 2O 3.113 

In addition to the thermodynamically stable β phase, the α, γ, and ε phases have been stabilized using MBE with heteroepitaxy. The MBE growth on ( c-plane) sapphire (0001) frequently results in the formation of β-Ga 2O 3 ( 2 ¯01) layers with rotational domains.52,114,115 However, a close inspection of the interface with the substrate by Schewski et al. revealed the presence of an approximately 1-nm-thick pseudomorphic α-Ga 2O 3 (0001) interfacial layer that is likely stabilized by lattice-mismatch-induced strain.63 Sapphire substrates of other orientations enable the formation of thicker α-Ga 2O 3 layers before the growth of β-Ga 2O 3 proceeds on top: a combined growth and in situ XRD study by Cheng et al. demonstrated the growth of a 143-nm-thick α-Ga 2O 3 (11 2 ¯0) layer on a-plane sapphire (11 2 ¯0) that relaxed the in-plane stain within the first 2 nm of the growth.64 Kracht et al. grew greater than 200-nm-thick α-Ga 2O 3 (10 1 ¯2) layers on ( r-plane) sapphire (10 2 ¯0) before the growth of β-Ga 2O 3 ( 2 ¯01) nucleated on the (0001)-facets that emerged by roughening of the α-Ga 2O 3 (10 1 ¯2) layer.116 Oshima et al. grew α-Ga 2O 3/ α-Al 2O 3 superlattices on r-plane sapphire, achieving a coherent interface without misfit dislocations for up to 1-nm-thick α-Ga 2O 3 (10 1 ¯2) layers.117 A combined growth and scanning transmission electron microscopy study by Jinno et al.69 revealed the growth of 51-nm-thick α-Ga 2O 3 (10 1 ¯0) layers on m-plane sapphire (10 1 ¯0) that relaxes at the α-Ga 2O 3–sapphire interface. These results indicate that the c-plane facet enhances the growth of β-Ga 2O 3 in MBE on a sapphire crystal plane perpendicular to the c-plane, such as the a- or m-planes, and allows the growth of phase-pure α-Ga 2O 3 by avoiding facets.

Another example of substrate-stabilized phase formation is that of the cubic-defective-spinel γ-Ga 2O 3 (001) on spinel MgAl 2O 4 (001) substrates.118 

Vogt et al. and Kracht et al. reported the growth of ε-Ga 2O 3 (001) on c-plane sapphire by metal-exchanged catalysis mediated through an additional In flux73 and Sn flux,116 respectively. Bi et al. reported the stabilized growth of ε-Ga 2O 3 (001) on c-plane sapphire by Mg doping.119 It is conceivable that the conditions for Ga 2O 3 growth catalyzed by In or Sn or Mg at high growth temperature thermodynamically promote the formation of the ε phase. This hypothesis is supported by the relatively small difference in formation energy between the β and ε phases, which decreases with increasing temperature.120 

ALD has been used for heteroepitaxial growth of α- and ε-Ga 2O 3 on c-plane sapphire.

Boschi et al. grew ε-Ga 2O 3 (001) on c-plane sapphire at 550  °C, obtaining uniform thickness over large surfaces.112 Roberts et al. grew α-Ga 2O 3 (0001) on c-plane sapphire by low-temperature ALD and revealed that the film consists of (0001)-oriented α-Ga 2O 3 columns originating from the substrate.121 A TEM study showed some inclusions, most likely of the ε phase and primarily at the tip of the columns, with an amorphous phase located near the surface of the film and between the α-Ga 2O 3 columns.

Wheeler et al. demonstrated the phase selectivity of β-, α-, and ε-Ga 2O 3 on c-plane sapphire as functions of the plasma gas composition, the gas flow and pressure during the plasma pulse, and the growth temperature.60 Pure α-Ga 2O 3 films with an abrupt interface were produced at low temperatures and pressures, whereas high-quality, single-phase β-Ga 2O 3 films were directly deposited on c-plane sapphire at low pressures with an Ar-O 2 plasma gas composition. In contrast, no single-phase ε-Ga 2O 3 heteroepitaxial films were obtained. ε-phase films were predominately obtained without nucleation or buffer layers by going to higher pressures and temperatures. Carbon, or other impurities, are hypothesized to help stabilize the β-Ga 2O 3 phase. Thus, to realize pure semi-stable phases, particular attention should be focused on in situ and ex situ cleaning of the sapphire substrate prior to deposition and on the initiation of growth with the metal organic precursor because ALD is conducted below the desorption temperature of most carbon species.

Table III shows examples of growth conditions used to grow semi-stable Ga 2O 3 by different growth methods and on different substrates. When using mist CVD, the crystal structure of the Ga 2O 3 epilayer is that of the underlying layer; that is, the crystal structure of the Ga 2O 3 epilayer is basically determined by the structure of the underlying layer, in spite of the lattice mismatch, rather than by the growth conditions. In other words, mist CVD introduces a strong tendency of phase-locking epitaxy, as discussed in Sec. III B.

TABLE III.

Examples of growth conditions used to grow semi-stable Ga2O3 by various growth methods.

On sapphireOn GaNOn spinel
Mist CVD α : typically <500 °C, inclusion of β at higher temperature ɛ γ 
HVPE α: low temperature (500–600 °C)58,59,109–111 ɛ35   
 ɛ: 650 °C, 10-nm-thick α-Ga2O3 interlayer68    
MOCVD α: very thin (3 ML) at the initial stage on c-sapphire, then turn to β63    
 α, β, ɛ: depending on the HCl flow rate (5 sccm: β, 30 sccm: ɛ, 60 sccm: α dominant)62    
 β, ɛ: at 500 °C, depending on pressure (100–400 mbar: β, 35 mbar: ɛ)66    
 ɛ: TMGa + H2O at 650 °C112    
 ɛ: TMGa + H2 carrier, 650 °C, 100 mbar102    
 ɛ: TEGa + O2, 500 °C, 100 mbar67    
MBE α: very thin (3 ML) at the initial stage on c-sapphire, then turn to β63   γ118  
 β: on a,13, r,64 and m69 planes   
 ɛ: with catalysis (In,114 Sn116) or doping (Mg)119    
ALD α: low temperature (250 °C), inclusion of ɛ121    
 α, β, ɛ: depending on growth conditions;60  α at low temperature (295 °C) and low pressure (≤10 mTorr)   
 ɛ: low temperature (<550 °C)112  
On sapphireOn GaNOn spinel
Mist CVD α : typically <500 °C, inclusion of β at higher temperature ɛ γ 
HVPE α: low temperature (500–600 °C)58,59,109–111 ɛ35   
 ɛ: 650 °C, 10-nm-thick α-Ga2O3 interlayer68    
MOCVD α: very thin (3 ML) at the initial stage on c-sapphire, then turn to β63    
 α, β, ɛ: depending on the HCl flow rate (5 sccm: β, 30 sccm: ɛ, 60 sccm: α dominant)62    
 β, ɛ: at 500 °C, depending on pressure (100–400 mbar: β, 35 mbar: ɛ)66    
 ɛ: TMGa + H2O at 650 °C112    
 ɛ: TMGa + H2 carrier, 650 °C, 100 mbar102    
 ɛ: TEGa + O2, 500 °C, 100 mbar67    
MBE α: very thin (3 ML) at the initial stage on c-sapphire, then turn to β63   γ118  
 β: on a,13, r,64 and m69 planes   
 ɛ: with catalysis (In,114 Sn116) or doping (Mg)119    
ALD α: low temperature (250 °C), inclusion of ɛ121    
 α, β, ɛ: depending on growth conditions;60  α at low temperature (295 °C) and low pressure (≤10 mTorr)   
 ɛ: low temperature (<550 °C)112  

Note that thick α-Ga 2O 3 was grown by HVPE for which GaCl and HCl exist in the growth atmosphere. Ga–Cl complexes and HCl are also expected in the atmosphere of mist CVD (HCl markedly influences the Ga 2O 3 phase in MOCVD).62 The similarity of the precursors as well as the existence of HCl in both mist CVD and HVPE may stabilize the growth of α-Ga 2O 3. Another important factor is the low growth temperature.

Conversely, when using other growth methods, one of the common features is the growth of very thin α-Ga 2O 3 directly on α-Al 2O 3. This is due to the nature of phase locking, which can dominate the very initial stage of growth on α-Al 2O 3, irrespective of the growth method. However, as the growth continues, Ga 2O 3 seems to voluntarily nucleate to form stable β-Ga 2O 3 rather than to follow the atomic arrangement of the underlying layer. This conclusion is supported by high kinetic or thermal energy of the precursors and by the high substrate temperature.

Since α-Ga 2O 3 is grown by heteroepitaxy on α-Al 2O 3, dislocation-originated defects form in the α-Ga 2O 3 thin films. Figure 10(a) shows a surface TEM image of an α-Ga 2O 3 thin film. The in-plane contrasting density is not homogeneous. By applying an inverse fast Fourier transform method to the image, clear dislocation lines (identified by red arrows) appear, as shown in Figs. 10(b) and 10(c), selecting the 11 2 ¯0 and 30 3 ¯0 diffraction spots, respectively. These dislocations are attributed to domain boundaries because inclusion of rotational domains has never been reported.31 

FIG. 10.

(a) Surface TEM images of the α-Ga 2O 3 thin film on α-Al 2O 3. Inverse fast Fourier transform images of (b) 11 2 ¯0 and (c) 30 3 ¯0.

FIG. 10.

(a) Surface TEM images of the α-Ga 2O 3 thin film on α-Al 2O 3. Inverse fast Fourier transform images of (b) 11 2 ¯0 and (c) 30 3 ¯0.

Close modal

For α-Ga 2O 3 thin films, the screw dislocations are rare in cross-sectional TEM images, whereas the edge dislocation density is calculated to be 7 × 10 10 cm 2 by applying Ham’s method to the images.78 Efforts are continuing to reduce the dislocation density. For example, Jinno et al. reduced the edge dislocation density to 6 × 10 8 cm 2 by introducing quasi-graded α-(Al,Ga) 2O 3 buffer layers.122 ELO produced α-Ga 2O 3 films on the mask region without noticeable dislocation, as per TEM images for HVPE111 and mist CVD.123 Kawara et al. achieved the low dislocation density of less than 5 × 10 6 cm 2 with the use of double-layered ELO and HVPE.124 

Unintentionally doped α-Ga 2O 3 has a very high resistivity, and n-type conductivity is obtained by doping with tin (Sn)125,126 or silicon (Si)127 in mist CVD. The carrier concentration was controlled in the range of 10 17–10 19 cm 3. However, the mobility was less than 24 cm 2 V 1 s 1, suggesting that crystal defects, such as dislocations, severely restrict the mobility. Improvement of the crystal quality is important to obtain better electrical properties.

Nevertheless, SBDs with low on-resistance and high breakdown voltage (e.g., 0.1 m Ω cm 2 and 531 V or 0.4 m Ω cm 2 and 855 V, respectively) were grown by FLOSFIA, Inc. using mist CVD.32 This achievement was followed by the fabrication of a normally off MOSFET with a maximum channel mobility of 72 cm 2 V 1 s 1.128 Mist CVD is a simple, safe, and cost-effective technology that can produce low residual impurities and specular surface mobility,32 promising high-performance devices at low cost. Maeda et al. showed that the α-Ga 2O 3 SBDs had nearly ideal current–voltage characteristics in spite of the inclusion of numerous dislocation defects,129 which is good news for the future evolution of α-Ga 2O 3 devices and encourages us to elucidate the structure and electronic characteristics of the defects.

An almost-fatal problem in wide-bandgap oxide semiconductors is the difficulty to realize both n- and p-type layers from the same material. However, corundum-structured α-Ga 2O 3 is expected to form p–n junctions with other corundum-structured p-type oxides. One of the p-type oxide candidates is α-Ir 2O 3, whose p-type conductivity is attributed to the Ir 5 d orbital, which supplies holes that are located above the O 2 p orbital. Growth of p-type α-Ir 2O 3 and the rectification characteristics of a p–n junction made of α-Ir 2O 3 and α-Ga 2O 3 have been demonstrated, and it was shown that the lattice mismatches between α-Ir 2O 3 and α-Ga 2O 3 are as small as 0.3% and 0.6% along the a and c axes, respectively.130 By alloying with Ga 2O 3, the bandgap is enlarged and the lattice parameters approach those of Ga 2O 3; these characteristics are preferable for p–n junctions. We fabricated p-type α-(Ir,Ga) 2O 3 films with bandgaps up to 4.3 eV and p–n junctions from n-type Ga 2O 3.131 

Corundum-structured α-In 2O 3 is also a semi-stable phase but was grown on sapphire substrates by mist CVD.90,132 Given that α-Al 2O 3 is a stable phase, the corundum-structured alloy semiconductors with α-Al 2O 3, α-Ga 2O 3, or α-In 2O 3 [ α-(Al,Ga,In) 2O 3] enable bandgaps to be engineered over a wide range of energy. Figure 11 shows bandgap engineering by α-(Al,Ga) 2O 3 and α-(In,Ga) 2O 3 alloys from 3.7 to about 9 eV.133 Note that the growth of single-phase β-(Al xGa 1 x) 2O 3 is very limited for high-Al composition (i.e., large x) because the monoclinic phase is not a stable structure for Al 2O 3. Conversely, a variety of heterostructures are realized by α-(Al xGa 1 x) 2O 3 with a wide range of x. The band lineup of the α-(Al xGa 1 x) 2O 3/Ga 2O 3 heterostructure is shown to be type I by x-ray photoemission spectroscopy,134 which is favorable for applications involving heterostructure transistors and multiple quantum wells.

FIG. 11.

Bandgaps of α-(Al,Ga) 2O 3 and α-(In,Ga) 2O 3 alloys.133 Reproduced with permission from S. Fujita and K. Kaneko, J. Cryst. Growth 401, 588 (2014). Copyright 2014 Elsevier.

FIG. 11.

Bandgaps of α-(Al,Ga) 2O 3 and α-(In,Ga) 2O 3 alloys.133 Reproduced with permission from S. Fujita and K. Kaneko, J. Cryst. Growth 401, 588 (2014). Copyright 2014 Elsevier.

Close modal

The growth of γ-Ga 2O 3 by mist CVD has been followed by its growth by PLD and MBE. Si doping of γ-Ga 2O 3 produces highly n-doped films (electron concentration of 10 19 cm 3) with the ionization energy of the donor to be at most 7 meV,135 which is sufficiently small and similar to that in β-Ga 2O 3. γ-Al 2O 3 was grown by MBE on (100) MgAl 2O 4 substrates. A γ-Ga 2O 3/ γ-Al 2O 3 heterointerface has a type-I band lineup with conduction- and valence-band offsets of 1.6 and 0.2 eV, respectively.136 These results are encouraging for the evolution of γ-(Al xGa 1 x) 2O 3 alloys and heterostructures. Note also that the formation of phases other than the γ phase has not been reported for any growth method (i.e., mist CVD, PLD, or MBE).102,135,136 γ-Ga 2O 3 (and γ-Al 2O 3) may nucleate due to the significant influence of phase locking by (100) MgAl 2O 4 substrates.

The spontaneous piezoelectric polarization of the ferroelectric ε-Ga 2O 337,102,103 and the bandgap engineering with ε-(Al xGa 1 x) 2O 3108 may be used in applications involving 2DEG transistors. Wang et al. predicted that a 2DEG with a high density of 10 14 cm 2 forms at the interface of ε-Ga 2O 3 and m-AlN (m-GaN) without doping, which is promising for high-electron-mobility transistors.137 Ranga et al. showed theoretically that the 2DEG density at the ε-(Al,Ga) 2O 3/ ε-Ga 2O 3 heterointerface can be as high as 1.4 × 10 14 cm 2 and suggested spontaneous polarization switching in ε-Ga 2O 3/ ε-(Al,Ga) 2O 3/ ε-Ga 2O 3 heterostructures with charge contrast ratios as high as 1600.138 For future practical applications, formation of Ohmic contacts139 and Si doping140 have been investigated, and the latter shows that Si is a promising dopant for the formation of highly conductive n-type ε-Ga 2O 3. From the operation as solar-blind photodetectors, the bandgap of ε-Ga 2O 3 was determined to be 4.6–4.7 eV and deep levels were found at 0.7 and 2.3–2.75 eV below the conduction band.141 Good controllability for forming ε-(Al,Ga) 2O 3/ ε-Ga 2O 3 short-period superlattices (in that paper, the polymorph was referred to as κ rather than ε) was demonstrated by PLD, as evinced by the numerous satellite peaks in XRD spectra, as shown in Fig. 12, and the continuous variation of c-lattice constants and bandgaps of the quasialloys.142 

FIG. 12.

XRD 2 θ ω scans of κ-(Al,Ga) 2O 3/ κ-Ga 2O 3 short-period superlattices of 15 periods. (a) Satellite peaks observed for (Al 0.3Ga 0.7) 2O 3 barriers and (b) spectra for different Al compositions in the barriers.142 Reproduced with permission from Kneiß et al., APL Mater. 8, 051112 (2020). Copyright 2020 AIP Publishing LLC.

FIG. 12.

XRD 2 θ ω scans of κ-(Al,Ga) 2O 3/ κ-Ga 2O 3 short-period superlattices of 15 periods. (a) Satellite peaks observed for (Al 0.3Ga 0.7) 2O 3 barriers and (b) spectra for different Al compositions in the barriers.142 Reproduced with permission from Kneiß et al., APL Mater. 8, 051112 (2020). Copyright 2020 AIP Publishing LLC.

Close modal

Notice the substrate materials used as a function of growth methods to grow ε-Ga 2O 3 and related alloys. For mist-CVD growth, the substrate materials were, for example, MgO, YSZ, AlN, GaN, and (111) SrTiO 3,37,104,105,107,108 and sapphire was not used probably because α-Ga 2O 3 was being grown. The role of these substrates for the growth of ε-Ga 2O 3 is discussed in Sec. III D. Conversely, for MOCVD and PLD growth, the substrate was sapphire.139–142 These papers made no mention of the formation of α-Ga 2O 3 domains, which suggests a marked difference in the phase-matching force exerted by the substrate for mist CVD vs other growth technology.

Semi-stable Ga 2O 3 is transferred to the most-stable β phase after thermal annealing. This can limit the fabrication of semi-stable Ga 2O 3-based devices because some processes, such as post-ion-implantation thermal annealing, may need higher temperatures than those of the growth. Therefore, knowledge of thermal stability, the phase-transition mechanism, and a novel technology to enhance the thermal stability are important for actual device applications of semi-stable Ga 2O 3.

The α phase reportedly converts to the β phase upon heating at atmospheric pressure to 600–650  °C.6,36,143,144 However, Jinno et al.145 showed that 30-nm-thick α-Ga 2O 3 could be grown at temperatures higher than 700  °C on α-(Al 0.4Ga 0.6)O 3 buffer layers on sapphire. This means that the phase transition from α to β is not simply dominated by the temperature but depends on other factors, which encourages investigations into the phase-transition mechanism.

We focused on how the α-Ga 2O 3 layer thickness affects the thermal stability.146 Ga 2O 3 films grown on c-plane sapphire substrates via mist CVD were thermally annealed by successively increasing the annealing temperature T A in an atmospheric furnace for 30 min at each T A. After annealing at a given T A, symmetric XRD 2 θ ω measurements were made of the samples and the XRD spectra were monitored. This allowed the phase-transition temperature to be defined at the annealing temperature T A, where the β phase becomes visible in the spectra. In Fig. 13, the open and solid symbols represent samples that maintained the α phase and completely converted to β-Ga 2O 3, respectively. These results show clearly that the phase-transition temperature, which should be between the open and solid symbols, and the film thickness are strongly correlated but depend only weakly on the growth temperature and the Ga precursor (i.e., impurity concentration). In thinner films, the α-Ga 2O 3 films maintain the corundum structure at temperatures above 600  °C. An α-Ga 2O 3 film around 20 nm thick maintains the corundum structure at T A = 750 °C, which is 150  °C greater than that of a film several hundred nanometers thick. Since the critical thickness of α-Ga 2O 3 films grown on c-plane sapphire substrates is expected to be a few nm due to the large lattice mismatch between sapphire and α-Ga 2O 3 (4.6 % along the a axis),63 the α-Ga 2O 3 films used here must have been almost fully relaxed even when the films were as thin as about 20 nm. Therefore, we conclude that the dislocation density in α-Ga 2O 3 does not define the stability boundary of α-Ga 2O 3 on sapphire at the present stage.

FIG. 13.

Open symbols indicate samples that maintained the α phase, and solid symbols indicate samples that are completely converted to the β phase at the given annealing temperature and as a function of the film thickness. Samples in blue and red were grown using Ga(acac) 3 and GaCl 3, respectively, as a Ga precursor. The circular, triangular, and rhomboid symbols represent samples grown at 500, 600, and 700  °C, respectively.146 The open star depicts the selective-area-grown α-Ga 2O 3 [see Fig. 15(b)]. Reproduced with permission from Jinno et al., APL Adv. 10, 115013 (2020). Copyright 2020 AIP Publishing LLC.

FIG. 13.

Open symbols indicate samples that maintained the α phase, and solid symbols indicate samples that are completely converted to the β phase at the given annealing temperature and as a function of the film thickness. Samples in blue and red were grown using Ga(acac) 3 and GaCl 3, respectively, as a Ga precursor. The circular, triangular, and rhomboid symbols represent samples grown at 500, 600, and 700  °C, respectively.146 The open star depicts the selective-area-grown α-Ga 2O 3 [see Fig. 15(b)]. Reproduced with permission from Jinno et al., APL Adv. 10, 115013 (2020). Copyright 2020 AIP Publishing LLC.

Close modal

We also investigated the thermal stability of selective-area grown (SAG) α-Ga 2O 3 on sapphire substrates. Selective-area growth was done using a dot-patterned SiO 2 mask on a sapphire substrate, where the open dots were 3  μm in diameter. The 500-nm-thick SAG α-Ga 2O 3 maintains the corundum structure at T A = 700 °C, with the α-Ga 2O 3 films completely converting to the β phase, as shown in Fig. 13.

Figure 14(a) shows a cross-sectional TEM image of a 135-nm-thick sample after annealing at T A = 670 °C , viewed along the 11 2 ¯ 0 axis. The sample is partially converted to β-Ga 2O 3. The diffraction pattern at the position c is consistent with that at position b, suggesting that the darker regions in Fig. 14(a) are the α phase. Conversely, the diffraction patterns at positions d and e reflect those of the β phase; that is, the brighter regions in Fig. 14(a) are β-Ga 2O 3, which was converted from α-Ga 2O 3. The TEM images suggest that the phase transformation from α to β does not occur uniformly in the α-Ga 2O 3 film, but the α-Ga 2O 3 film converts to the β phase near the surface of the film.

FIG. 14.

(a) Cross-sectional bright-field TEM image of an α-Ga 2O 3 film on a sapphire substrate after annealing at 670  °C , viewed along the 11 2 ¯ 0 axis. Panels (b)–(e) show diffraction spots for the area shown in panel (a).146 Reproduced with permission from Jinno et al., APL Adv. 10, 115013 (2020). Copyright 2020 AIP Publishing LLC.

FIG. 14.

(a) Cross-sectional bright-field TEM image of an α-Ga 2O 3 film on a sapphire substrate after annealing at 670  °C , viewed along the 11 2 ¯ 0 axis. Panels (b)–(e) show diffraction spots for the area shown in panel (a).146 Reproduced with permission from Jinno et al., APL Adv. 10, 115013 (2020). Copyright 2020 AIP Publishing LLC.

Close modal

McCandless et al. reported the stabilization of undoped MBE-grown α-Ga 2O 3 on m-plane sapphire at annealing temperatures up to 800  °C in a N 2 environment by masking a sample with SiO 2, Mo, and ALD AlO x.147 An uncapped α-Ga 2O 3 film grown on m-plane sapphire is stable up to 600  °C annealing, similar to what is reported for samples grown by mist CVD on c-plane sapphire. The impurity concentration of the MBE-grown samples is expected to be much less than that of the mist-CVD-grown films, indicating that the impurity concentrations in α-Ga 2O 3 are not the key factor defining the stability of α-Ga 2O 3. The α-Ga 2O 3 samples capped with SiO 2 or Mo become amorphous after 800  °C annealing, whereas the AlO x-capped α-Ga 2O 3 film is stable up to 900  °C. They conclude that some Al diffusion occurs, which promotes phase stability and/or that the AlO x mask crystallizes, which applies strain to α-Ga 2O 3 and can, therefore, aid in phase stabilization.

Based on all these results, the phase transition seems to be attributed to the surface covering and the thermal stress caused by the difference between the thermal expansion coefficients of α-Ga 2O 3 and sapphire. The thermal expansion coefficient of α-Ga 2O 3 exceeds that of sapphire and the difference between the coefficients increases with temperature,148,149 suggesting a tensile stress at the surface of the α-Ga 2O 3 film and a compressive stress at the interface between the film and the sapphire substrate, as shown in Fig. 15. The phase transition near the surface indicates that the tensile strain causes the transformation to the lower-density β-Ga 2O 3. Decreasing the film thickness or using SAG releases the stress, thereby enhancing the phase stability. It is also possible that the phase transition to the β phase is due to an inhomogeneous strain in α-Ga 2O 3. At the present stage, a small amount of β-Ga 2O 3 is included in the SAG α-Ga 2O 3 after annealing, but the inclusion is probably due to the strain on the window region where α-Ga 2O 3 is subject to strain from the sapphire substrate. Therefore, a technique to reduce the strain in α-Ga 2O 3 should enhance the thermal stability. The use of appropriate buffer layers and/or a patterned substrate may prevent α-Ga 2O 3 films from thermal stress, leading to a higher phase-transition temperature and high-temperature growth. Enhanced thermal stability by slight Al doping143 may also be due to partial strain relaxation in α-Ga 2O 3 by Al-induced deformation of the lattice.

FIG. 15.

Schematical diagram showing a relationship between an α-Ga 2O 3 film and a sapphire substrate at (a) room temperature and (b) the phase-transition temperature.

FIG. 15.

Schematical diagram showing a relationship between an α-Ga 2O 3 film and a sapphire substrate at (a) room temperature and (b) the phase-transition temperature.

Close modal

For other phases, few systematic experiments on thermal stability have been conducted. According to Roy et al., γ and ε phases transform to the β phase at 650 and 870  °C, respectively.6 Broad discussions on the thermal stability and phase-transition mechanism of semi-stable Ga 2O 3, together with how to overcome the issues, should be undertaken to stimulate further development of these attractive materials.

Although β-Ga 2O 3 is the thermally stable phase and its substrates are available, interest in the semi-stable phases of Ga 2O 3 has increased due to the desire to fully exploit the unique properties of Ga 2O 3. Numerous efforts are ongoing to control the phase by the choice of substrate materials, growth methods, and growth conditions. Note that mist CVD allows for phase control, producing an atomic arrangement at the growth surface, which is the same as that of the substrate surface, that is, α on sapphire, γ on spinel, and ε (or κ) on GaN, ZnO, or AlN. This is tentatively attributed to the low-temperature growth, which results in the incident atoms having higher interaction energy with respect to the underlying surface atoms than with respect to other incident atoms. Phase control is also attributed to the assistance of precursors, which allows the source materials to stick to the underlying surface and thereby follow its crystal structure.

Recently, other growth technologies produced the semi-stable phase; for example, MOCVD growth of α-Ga 2O 3 and α-(Al xGa 1 x)O 3 over the range 0 < x < 1 was recently demonstrated on m-plane sapphire substrates.150 However, at the present stage, the growth windows to obtain the α phase on sapphire seem to be the widest in mist CVD; in our experiments, no other phases appear for growth temperatures below 500  °C.

The crystal phase may be determined by which procedure is most energetically favorable for the incident atoms: following the underlying structure or forming the stable β phase by self-nucleation. Low-temperature growth seems favorable to promote the former over the latter. In addition, we speculate that an oxygen-terminated surface and the association of water in mist-CVD growth may strongly induce the incident atoms to follow the underlying structure. Further experimental and theoretical research is required to better understand how to tailor these phases, not only with mist CVD but also with MBE, MOCVD, and other growth technologies. We are confident that semi-stable materials, if their structure is firmly controlled, will stimulate the further evolution of electronics and optics with Ga 2O 3, as well as that of other multiphase materials.

K.K. and S.F. acknowledge the financial support by the Advanced Research Program for Energy and Environmental Technologies of the New Energy and Industrial Technology Development Organization (NEDO), the Council for Science, Technology and Innovation (CSTI), the Cross-ministerial Strategic Innovation Promotion Program (SIP), and “Energy Systems of an Internet of Energy (IoE) society” (Funding agency: Japan Science and Technology Agency), and JSPS KAKENHI (Grant Nos. 25286050, 18H01870, and 21H01811). K.U. acknowledges financial support by JSPS KAKENHI (Grant No. 18K04958).

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
Gallium Oxide, edited by S. Pearton, F. Ren, and M. Mastro (Elsevier, 2018).
2.
Gallium Oxide, edited by M. Higashiwaki and S. Fujita (Springer, 2020).
3.
A.
Kuramata
,
K.
Koshi
,
S.
Watanabe
,
Y.
Yamaoka
,
T.
Masui
, and
S.
Yamakoshi
,
Jpn. J. Appl. Phys.
55
,
1202A2
(
2016
).
4.
Z.
Galazka
,
R.
Uecker
,
D.
Klimm
,
K.
Irmscher
,
M.
Naumann
,
M.
Pietsch
,
A.
Kwasniewski
,
R.
Bertram
,
S.
Ganschow
, and
M.
Bickermann
,
ECS J. Solid State Sci. Technol.
6
,
Q3007
(
2017
).
5.
H. F.
Mohamed
,
C.
Xia
,
Q.
Sai
,
H.
Cui
,
M.
Pan
, and
H.
Qi
,
J. Semicond.
40
,
011801
(
2019
).
6.
R.
Roy
,
V. G.
Hill
, and
E. F.
Osborn
,
J. Am. Chem. Soc.
74
,
719
(
1952
).
8.
S.
Ohira
,
M.
Yoshioka
,
T.
Sugawara
,
K.
Nakajima
, and
T.
Shishido
,
Thin Solid Films
496
,
53
(
2006
).
9.
K.
Shimamura
,
E. G.
Víllora
,
K.
Domen
,
K.
Yui
,
K.
Aoki
, and
N.
Ichinose
,
Jpn. J. Appl. Phys.
44
,
L7
(
2005
).
10.
L.
Bin
,
Z.
Hong
,
J.
Ruo-Lian
,
S.
Yi
, and
Z.
You-Dou
,
Chin. Phys. Lett.
25
,
2185
(
2008
).
11.
S.
Ohira
,
N.
Suzuki
,
H.
Minami
,
K.
Takahashi
,
T.
Araki
, and
Y.
Nanishi
,
Phys. Status Solidi C
7
,
2306
(
2007
).
12.
K.
Matsuzaki
,
H.
Yanagi
,
T.
Kamiya
,
H.
Hiramatsu
,
K.
Nomura
,
M.
Hirano
, and
H.
Hosono
,
Appl. Phys. Lett.
88
,
092106
(
2006
).
13.
E. G.
Víllora
,
K.
Shimamura
,
K.
Kitamura
, and
K.
Aoki
,
Appl. Phys. Lett.
88
,
031105
(
2006
).
14.
T.
Oshima
,
N.
Arai
,
N.
Suzuki
,
S.
Ohira
, and
S.
Fujita
,
Thin Solid Films
516
,
5768
(
2008
).
15.
T.
Oshima
,
T.
Okuno
,
N.
Arai
,
N.
Suzuki
,
S.
Ohira
, and
S.
Fujita
,
Appl. Phys. Express
1
,
011202
(
2008
).
16.
T.
Oshima
,
T.
Okuno
,
N.
Arai
,
Y.
Kobayashi
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
48
,
011605
(
2009
).
17.
T.
Oshima
,
T.
Okuno
,
N.
Arai
,
Y.
Kobayashi
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
48
,
070202
(
2009
).
18.
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
,
E. G.
Víllora
,
K.
Shimamura
, and
S.
Yamakoshi
,
Appl. Phys. Express
5
,
035502
(
2012
).
19.
M.
Higashiwaki
,
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Appl. Phys. Lett.
100
,
013504
(
2012
).
20.
K.
Sasaki
,
M.
Higashiwaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Appl. Phys. Express
6
,
086502
(
2013
).
21.
M.
Higashiwaki
,
K.
Sasaki
,
T.
Kamimura
,
M. H.
Wong
,
D.
Krishnamurthy
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
Appl. Phys. Lett.
103
,
123511
(
2013
).
22.
M.
Higashiwaki
,
H.
Murakami
,
Y.
Kumagai
, and
A.
Kuramata
,
Jpn. J. Appl. Phys.
55
,
1202A1
(
2016
).
23.
M.
Higashiwaki
,
A.
Kuramata
,
H.
Murakami
, and
Y.
Kumaga
,
J. Phys. D: Appl. Phys.
50
,
333002
(
2017
).
24.
S. J.
Pearton
,
F.
Ren
,
M.
Tadjer
, and
J.
Kim
,
J. Appl. Phys.
124
,
220901
(
2018
).
25.
M.
Higashiwaki
and
G. H.
Jessen
,
Appl. Phys. Lett.
112
,
060401
(
2018
).
26.
S. J.
Pearton
,
J.
Yang
,
P. H.
Cary
,
F.
Ren
,
J.
Kim
,
M. J.
Tadjer
, and
M. A.
Mastro
,
Appl. Phys. Rev.
5
,
011301
(
2018
).
27.
M.
Baldini
,
Z.
Galazka
, and
G.
Wagner
,
Mater. Sci. Semicond. Process.
78
,
132
(
2018
).
28.
M.
Razeghi
,
J.-H.
Park
,
R.
McClintock
,
D.
Pavlidis
,
F. H.
Teherani
,
D. J.
Rogers
,
B. A.
Magill
,
G. A.
Khodaparast
,
Y.
Xu
,
J.
Wu
, and
V. P.
Dravid
,
Proc. SPIE
10533
,
105330R
(
2018
).
29.
M. J.
Tadjer
,
J. L.
Lyons
,
N.
Nepal
,
J. A.
Freitas
, Jr.
,
A. D.
Koehler
, and
G. M.
Foster
,
ECS J. Solid State Sci. Technol.
8
,
Q3187
(
2019
).
30.
X.
Chen
,
F.
Ren
,
S.
Gu
, and
J.
Ye
,
Photonics Res.
7
,
381
(
2019
).
31.
D.
Shinohara
and
S.
Fujita
,
Jpn. J. Appl. Phys.
47
,
7311
(
2008
).
32.
M.
Oda
,
R.
Tokuda
,
H.
Kambara
,
T.
Tanikawa
,
T.
Sasaki
, and
T.
Hitora
,
Appl. Phys. Express
9
,
021101
(
2016
).
33.
S.
Fujita
,
M.
Oda
,
K.
Kaneko
, and
T.
Hitora
,
Jpn. J. Appl. Phys.
55
,
1202A3
(
2016
).
34.
T.
Oshima
,
T.
Nakazono
,
A.
Mukai
, and
A.
Ohtomo
,
J. Cryst. Growth
359
,
60
(
2012
).
35.
Y.
Oshima
,
E. G.
Víllora
,
Y.
Matsushita
,
S.
Yamamoto
, and
K.
Shimamura
,
J. Appl. Phys.
118
,
085301
(
2015
).
36.
H. Y.
Playford
,
A. C.
Hannon
,
E. R.
Barney
, and
R. I.
Walton
,
Chem. Eur. J.
19
,
2803
(
2013
).
37.
H.
Nishinaka
,
D.
Tahara
, and
M.
Yoshimoto
,
Jpn. J. Appl. Phys.
55
,
1202BC
(
2016
).
38.
M.
Bosi
,
P.
Mazzolini
,
L.
Seravallim
, and
R.
Fornari
,
J. Mater. Chem. C
8
,
10975
(
2020
).
39.
B. R.
Tak
,
S.
Kumar
,
A. K.
Kapoor
,
D.
Wang
,
X.
Li
,
H.
Sun
, and
R.
Singh
,
J. Phys. D: Appl. Phys.
54
,
453002
(
2021
).
40.
J.
Lee
,
H.
Kim
,
L.
Gautam
, and
M.
Razeghi
,
Crystals
11
,
446
(
2021
).
41.
I.
Cora
,
F.
Mezzadri
,
F.
Boschi
,
M.
Bosi
,
M.
C̆aplovic̆ová
,
G.
Calestani
,
I.
Dódony
,
B.
Pécz
, and
R.
Fornari
,
CrystEngComm
19
,
1509
(
2017
).
42.
B.
Ollivier
,
R.
Retoux
,
P.
Lacorre
,
D.
Massiot
, and
G.
Férey
,
J. Mater. Chem.
7
,
1049
(
1997
).
43.
J.
Wu
,
H.
Yaguchi
,
K.
Onabe
,
Y.
Shiraki
, and
R.
Ito
,
Jpn. J. Appl. Phys.
37
,
1440
(
1998
).
44.
A.
Nakadaira
and
H.
Tanaka
,
J. Electron. Mater.
26
,
320
(
1997
).
45.
H.
Tsuchiya
,
K.
Sunaba
,
S.
Yonemura
,
T.
Suemasu
, and
F.
Hasegawa
,
Jpn. J. Appl. Phys.
36
,
L1
(
1997
).
46.
H.
Tachibana
,
T.
Ishido
,
M.
Ogawa
,
M.
Funato
,
S.
Fujita
, and
S.
Fujita
,
J. Cryst. Growth
196
,
41
(
1999
).
47.
S.
Fujita
,
S.
Hayashi
,
M.
Funato
, and
S.
Fujita
,
J. Cryst. Growth
99
,
437
(
1990
).
48.
T.
Takagi
,
H.
Tanaka
,
S.
Fujita
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
42
,
L401
(
2003
).
49.
M.
Orita
,
H.
Hiramatsu
,
H.
Ohta
,
M.
Hirano
, and
H.
Hosono
,
Thin Solid Films
411
,
134
(
2002
).
50.
T.
Oshima
,
T.
Okuno
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
46
,
7217
(
2007
).
51.
Y.
Lv
,
J.
Ma
,
W.
Mi
,
C.
Luan
,
Z.
Zhu
, and
H.
Xiao
,
Vacuum
86
,
1850
(
2012
).
52.
S.
Nakagomi
and
Y.
Kokubun
,
J. Cryst. Growth
349
,
12
(
2012
).
53.
Y.
Oshima
,
E. G.
Víllora
, and
K.
Shimamura
,
J. Cryst. Growth
410
,
53
(
2015
).
54.
R.
Cuscó
,
N.
Domènech-Amador
,
T.
Hatakeyama
,
T.
Yamaguchi
,
T.
Honda
, and
L.
Artúus
,
J. Appl. Phys.
117
,
185706
(
2015
).
55.
G. T.
Dang
,
T.
Kawaharamura
,
M.
Furuta
, and
M. W.
Allen
,
IEEE Trans. Electron Devices
62
,
3640
(
2015
).
56.
H.
Nishinaka
,
D.
Tahara
,
S.
Morimoto
, and
M.
Yoshimoto
,
Mat. Lett.
205
,
28
(
2017
).
57.
Y.
Nakabayashi
,
S.
Yamada
,
S.
Itoh
, and
T.
Kawae
,
J. Ceram. Soc. Jpn.
126
,
925
(
2018
).
58.
Y.
Oshima
,
E. G.
Vílllora
, and
K.
Shimamura
,
Appl. Phys. Express
8
,
055501
(
2015
).
59.
H.
Son
and
D.-W.
Jeon
,
J. Alloys Compd.
773
,
631
(
2019
).
60.
V. D.
Wheeler
,
N.
Nepal
,
D. R.
Boris
,
S. B.
Qadri
,
L. O.
Nyakiti
,
A.
Lang
,
A.
Koehler
,
G.
Foster
,
S. G.
Walton
,
C. R.
Eddy
, Jr.
, and
D.
Meyer
,
Chem. Mater.
32
,
1140
(
2020
).
61.
D. Y.
Guo
,
X. L.
Zhao
,
Y. S.
Zhi
,
W.
Cui
,
Y. Q.
Huang
,
Y. H.
An
,
P. G.
Li
,
Z. P.
Wu
, and
W. H.
Tang
,
Mat. Lett.
164
,
364
(
2016
).
62.
H.
Sun
,
K.-H.
Li
,
C. G. T.
Castanedo
,
S.
Okur
,
G. S.
Tompa
,
T.
Salagaj
,
S.
Lopatin
,
A.
Genovese
, and
X.
Li
,
Cryst. Growth Des.
18
,
2370
(
2018
).
63.
R.
Schewski
,
G.
Wagner
,
M.
Baldini
,
D.
Gogova
,
Z.
Galazka
,
T.
Schulz
,
T.
Remmele
,
T.
Markurt
,
H.
von Wenckstern
,
M.
Grundmann
,
O.
Bierwagen
,
P.
Vogt
, and
M.
Albrecht
,
Appl. Phys. Express
8
,
011101
(
2015
).
64.
Z.
Cheng
,
M.
Hanke
,
P.
Vogt
,
O.
Bierwagen
, and
A.
Trampert
,
Appl. Phys. Lett.
111
,
162104
(
2017
).
65.
Y.
Zhuo
,
Z.
Chen
,
W.
Tu
,
X.
Ma
,
Y.
Pei
, and
G.
Wang
,
Appl. Surf. Sci.
420
,
802
(
2017
).
66.
Y.
Chen
,
X.
Xia
,
H.
Liang
,
Q.
Abbas
,
Y.
Liu
, and
G.
Du
,
Cryst. Growth Des.
18
,
1147
(
2018
).
67.
X.
Cao
,
Y.
Xing
,
J.
Han
,
J.
Li
,
T.
He
,
X.
Zhang
,
J.
Zhao
, and
B.
Zhang
,
Mat. Sci. Semicond. Process.
123
,
105532
(
2021
).
68.
Y.
Yao
,
S.
Okur
,
L. A. M.
Lyle
,
G. S.
Tompa
,
T.
Salagaj
,
N.
Sbrockey
,
R. F.
Davis
, and
L. M.
Porter
,
Mater. Res. Lett.
6
,
268
(
2018
).
69.
R.
Jinno
,
C. S.
Chang
,
T.
Onuma
,
Y.
Cho
,
S.-T.
Ho
,
D.
Rowe
,
M. C.
Cao
,
K.
Lee
,
V.
Protasenko
,
D. G.
Schlom
,
D. A.
Muller
,
H. G.
Xing
, and
D.
Jena
,
Sci. Adv.
7
,
eabd5891
(
2021
).
71.
T.
Kawaharamura
,
Jpn. J. Appl. Phys.
53
,
05FF08
(
2014
).
72.
S.
Kawasaki
,
S.
Motoyama
,
T.
Tatsuta
,
O.
Tsuji
,
S.
Okamura
, and
T.
Shiosaki
,
Jpn. J. Appl. Phys.
43
,
6562
(
2004
).
73.
O.
Milos̆ević
,
B.
Jordović
, and
D.
Uskoković
,
Mater. Lett.
19
,
165
(
1994
).
74.
T. Y.
Ma
,
S. H.
Kim
,
H. Y.
Moon
,
G. C.
Park
,
Y. J.
Kim
, and
K. W.
Kim
,
Jpn. J. Appl. Phys.
35
,
6208
(
1996
).
75.
H.
Nishinaka
and
S.
Fujita
,
J. Cryst. Growth
310
,
5007
(
2008
).
76.
H.
Nishinaka
,
Y.
Kamada
,
N.
Kameyama
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
48
,
121103
(
2009
).
77.
J. P.
Remeika
and
M.
Marezio
,
Appl. Phys. Lett.
8
,
87
(
1966
).
78.
K.
Kaneko
,
H.
Kawanowa
,
H.
Ito
, and
S.
Fujita
,
J. Appl. Phys.
51
,
11PJ03
(
2012
).
79.
J.
Narayan
and
B. C.
Larson
,
J. Appl. Phys.
93
,
278
(
2003
).
80.
N.
Otsuka
,
C.
Choi
,
Y.
Nakamura
,
S.
Nagakura
,
R.
Fischer
,
C. K.
Peng
, and
H.
Morkoç
,
Appl. Phys. Lett.
49
,
277
(
1986
).
81.
C.
d’Anterroches
,
J. Y.
Marzin
,
G.
Leroux
, and
L.
Goldstein
,
J. Cryst. Growth
81
,
121
(
1987
).
82.
A. S.
Nandedkar
and
J.
Narayan
,
Mater. Sci. Eng. A
113
,
51
(
1989
).
83.
C.-M.
Wang
,
S.
Thevuthasan
,
F.
Gao
,
D. E.
McCready
, and
S. A.
Chambers
,
Thin Solid Films
414
,
31
(
2002
).
84.
K.
Kaneko
, Ph.D. thesis, Kyoto University, 2013.
85.
N.
Taga
,
M.
Maekawa
,
Y.
Shigesato
,
I.
Yasui
,
M.
Kakei
, and
T. E.
Haynes
,
J. Appl. Phys.
37
,
6524
(
1998
).
86.
M.
Kamei
,
T.
Yagami
,
S.
Takaki
, and
Y.
Shigesato
,
Appl. Phys. Lett.
64
,
2712
(
1994
).
87.
H.
Ohta
,
M.
Orita
,
M.
Hirano
, and
H.
Hosono
,
J. Appl. Phys.
91
,
3547
(
2002
).
88.
N.
Suzuki
,
K.
Kaneko
, and
S.
Fujita
,
J. Cryst. Growth
364
,
30
(
2013
).
89.
K.
Kaneko
,
Y.
Ito
,
T.
Uchida
, and
S.
Fujita
,
Appl. Phys. Express
8
,
095503
(
2015
).
90.
T.
Yamaguchi
,
S.
Takahashi
,
T.
Kiguchi
,
A.
Sekiguchi
,
K.
Kaneko
,
S.
Fujita
,
H.
Nagai
,
M.
Sato
,
T.
Onuma
, and
T.
Honda
,
Appl. Phys. Express
13
,
075504
(
2020
).
91.
H.
Fu
,
H.
Chen
,
X.
Huang
,
I.
Baranowski
,
J.
Montes
,
T.-H.
Yang
, and
Y.
Zhao
,
IEEE Trans. Electron Devices
65
,
3507
(
2018
).
92.
K.
Uno
,
M.
Ohta
, and
I.
Tanaka
,
Appl. Phys. Lett.
117
,
052106
(
2020
).
93.
W. W.
Rudolph
,
C. C.
Pye
, and
G.
Irmer
,
J. Raman Spectrosc.
33
,
177
(
2002
).
94.
L. E.
McInnes
,
S. E.
Rudd
, and
P. S.
Donnelly
,
Coord. Chem. Rev.
352
,
499
(
2017
).
95.
J.
Haladjian
,
P.
Bianco
, and
R.
Pilard
,
J. Chim. Phys.
6
,
1251
(
1974
).
96.
O. B.
Ajayi
,
B. S.
Akanni
,
J. N.
Lambi
,
H. D.
Burrows
,
O.
Osasona
, and
B.
Podor
,
Thin Solid Films
138
,
91
(
1986
).
97.
J. S.
Kim
,
H. A.
Marzouk
,
P. J.
Reucroft
,
J. D.
Robertson
, and
C. E.
Hamrin
, Jr.
,
Thin Solid Films
230
,
156
(
1993
).
98.
M.
Baltes
,
O.
Collart
,
P.
van der Voort
, and
E. F.
Vansant
,
Langmuir
1999
,
5841
.
99.
D.
Tamba
,
O.
Kubo
,
M.
Oda
,
S.
Osaka
,
K.
Takahashi
,
H.
Tabata
,
K.
Kaneko
,
S.
Fujita
, and
M.
Katayama
,
Appl. Phys. Lett.
108
,
251602
(
2016
).
100.
K.
Kaneko
,
I.
Kakeya
,
S.
Komori
, and
S.
Fujita
,
J. Appl. Phys.
113
,
233901
(
2013
).
101.
X.
Xia
,
Y.
Chen
,
Q.
Feng
,
H.
Liang
,
P.
Tao
,
M.
Xu
, and
G.
Du
,
Appl. Phys. Lett.
108
,
202103
(
2016
).
102.
F.
Mezzadri
,
G.
Calestani
,
F.
Boschi
,
D.
Delmonte
,
M.
Bosi
, and
R.
Fornari
,
Inorg. Chem.
55
,
12079
(
2016
).
103.
J.
Kim
,
D.
Tahara
,
Y.
Miura
, and
B. G.
Kim
,
Appl. Phys. Express
11
,
061101
(
2018
).
104.
D.
Tahara
,
H.
Nishinaka
,
S.
Morimoto
, and
M.
Yoshimoto
,
Jpn. J. Appl. Phys.
56
,
078004
(
2017
).
105.
H.
Nishinaka
,
H.
Komai
,
D.
Tahara
,
Y.
Arata
, and
M.
Yoshimoto
,
Jpn. J. Appl. Phys.
57
,
115601
(
2018
).
106.
M.
Bosi
,
L.
Seravalli
,
P.
Mazzolini
,
F.
Mezzadri
, and
R.
Fornari
,
Crys. Growth Des.
21
,
6393
(
2021
).
107.
D.
Tahara
,
H.
Nishinaka
,
M.
Noda
, and
M.
Yoshimoto
,
Mater. Lett.
232
,
47
(
2018
).
108.
D.
Tahara
,
H.
Nishinaka
,
S.
Morimoto
, and
M.
Yoshimoto
,
Appl. Phys. Lett.
112
,
152102
(
2018
).
109.
J. H.
Leach
,
K.
Udwary
,
J.
Rumsey
,
G.
Dodson
,
H.
Splawn
, and
K. R.
Evans
,
APL Mater.
7
,
022504
(
2019
).
110.
Y.
Oshima
,
K.
Kawara
,
T.
Oshima
,
M.
Okigawa
, and
T.
Shinohe
,
Semicond. Sci. Tech.
35
,
055022
(
2020
).
111.
Y.
Oshima
,
K.
Kawara
,
T.
Shinohe
,
T.
Hitora
,
M.
Kasu
, and
S.
Fujita
,
APL Mater.
7
,
022503
(
2019
).
112.
F.
Boschi
,
M.
Bosi
,
T.
Berzina
,
E.
Buffagni
,
C.
Ferrari
, and
R.
Fornari
,
J. Cryst. Growth
443
,
25
(
2016
).
113.
A. F. M. A. U.
Bhuiyan
,
Z.
Feng
,
J. M.
Johnson
,
H.-L.
Huang
,
J.
Sarker
,
M.
Zhu
,
M. R.
Karim
,
B.
Mazumder
,
J.
Hwang
, and
H.
Zhao
,
APL Mater.
8
,
031104
(
2020
).
114.
P.
Vogt
,
O.
Brandt
,
H.
Riechert
,
J.
Lähnemann
, and
O.
Bierwagen
,
Phys. Rev. Lett.
119
,
196001
(
2017
).
115.
M. Y.
Tsai
,
O.
Bierwagen
,
M. E.
White
, and
J. S.
Speck
,
J. Vac. Sci. Technol. A
28
,
354
(
2010
).
116.
M.
Kracht
,
A.
Karg
,
J.
Schörmann
,
M.
Weinhold
,
D.
Zink
,
F.
Michel
,
M.
Rohnke
,
M.
Schowalter
,
B.
Gerken
,
A.
Rosenauer
,
P. J.
Klar
,
J.
Janek
, and
M.
Eickhoff
,
Phys. Rev. Appl.
8
,
054002
(
2017
).
117.
T.
Oshima
,
Y.
Kato
,
M.
Imura
,
Y.
Nakayama
, and
M.
Takeguchi
,
Appl. Phys. Express
11
,
065501
(
2018
).
118.
T.
Oshima
,
Y.
Kato
,
M.
Oda
,
T.
Hitora
, and
M.
Kasu
,
Appl. Phys. Express
10
,
051104
(
2017
).
119.
X.
Bi
,
Z.
Wu
,
Y.
Huang
, and
W.
Tang
,
AIP Adv.
8
,
025008
(
2018
).
120.
S.
Yoshioka
,
H.
Hayashi
,
A.
Kuwabara
,
F.
Oba
,
K.
Matsunaga
, and
I.
Tanaka
,
J. Phys.: Condens. Matter
19
,
346211
(
2007
).
121.
J. W.
Roberts
,
J. C.
Jarman
,
D. N.
Johnstone
,
P. A.
Midgley
,
P. R.
Chalker
,
R. A.
Oliver
, and
F. C.-P.
Massabuau
,
J. Cryst. Growth
487
,
23
(
2018
).
122.
R.
Jinno
,
T.
Uchida
,
K.
Kaneko
, and
S.
Fujita
,
Appl. Phys. Express
9
,
071101
(
2016
).
123.
R.
Jinno
,
N.
Yoshimura
,
K.
Kaneko
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
58
,
120912
(
2019
).
124.
K.
Kawara
,
Y.
Oshima
,
M.
Okigawa
, and
T.
Shinohe
,
Appl. Phys. Express
13
,
075507
(
2020
).
125.
T.
Kawaharamura
,
G. T.
Dang
, and
M.
Furuta
,
Jpn. J. Appl. Phys.
51
,
040207
(
2012
).
126.
K.
Akaiwa
,
K.
Kaneko
,
K.
Ichino
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
55
,
1202BA
(
2016
).
127.
T.
Uchida
,
K.
Kaneko
, and
S.
Fujita
,
MRS Adv.
3
,
171
(
2018
).
128.
Press release from FLOSFIA, Inc., December 2, 2019.
129.
T.
Maeda
,
M.
Okigawa
,
Y.
Kato
,
I.
Takahashi
, and
T.
Shinohe
,
AIP Adv.
10
,
125119
(
2020
).
130.
S.
Kan
,
S.
Takemoto
,
K.
Kaneko
,
I.
Takahashi
,
M.
Sugimoto
,
T.
Shinohe
, and
S.
Fujita
,
Appl. Phys. Lett.
113
,
212104
(
2018
).
131.
K.
Kaneko
,
Y.
Masuda
,
S.
Kan
,
I.
Takahashi
,
Y.
Kato
,
T.
Shinohe
, and
S.
Fujita
,
Appl. Phys. Lett.
118
,
102104
(
2021
).
132.
K.
Kaneko
,
M.
Kitajima
, and
S.
Fujita
,
MRS Adv.
2
,
301
(
2017
).
133.
S.
Fujita
and
K.
Kaneko
,
J. Cryst. Growth
401
,
588
(
2014
).
134.
T.
Uchida
,
R.
Jinno
,
S.
Takemoto
,
K.
Kaneko
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
57
,
040314
(
2018
).
135.
T.
Oshima
,
K.
Matsuyama
,
K.
Yoshimatsu
, and
A.
Ohtomo
,
J. Cryst. Growth
421
,
23
(
2015
).
136.
T.
Oshima
,
Y.
Kato
,
E.
Magome
,
E.
Kobayashi
, and
K.
Takahashi
,
Jpn. J. Appl. Phys.
58
,
060910
(
2019
).
137.
J.
Wang
,
H.
Guo
,
C.-Z.
Zhu
,
Q.
Cai
,
G.-F.
Yang
,
J.-J.
Xue
,
D.-J.
Chen
,
Y.
Tong
,
B.
Liu
,
H.
Lu
,
R.
Zhang
, and
Y.-D.
Zheng
,
IEEE Electron Device Lett.
41
,
1052
(
2020
).
138.
P.
Ranga
,
S. B.
Cho
,
R.
Mishra
, and
S.
Krishnamoorthy
,
Appl. Phys. Express
13
,
061009
(
2020
).
139.
A.
Bosio C. Borelli
,
A.
Parisini
,
M.
Pavesi
,
S.
Vantaggio
, and
R.
Fornari
,
ECS J. Solid State Sci. Technol.
9
,
055002
(
2020
).
140.
H. J.
von Bardeleben
,
J. L.
Cantin
,
A.
Parisini
,
A.
Bosio
, and
R.
Fornari
,
Phys. Rev. Mater.
3
,
084601
(
2019
).
141.
M.
Pavesi
,
F.
Fabbri
,
F.
Boschi
,
G.
Piacentini
,
A.
Baraldi
,
M.
Bosi
,
E.
Gombia
,
A.
Parisini
, and
R.
Fornari
,
Mater. Chem. Phys.
205
,
502
(
2018
).
142.
M.
Kneiß
,
P.
Storm
,
A.
Hassa
,
D.
Splith
,
H.
von Wenckstern
,
M.
Lorenz
, and
M.
Grundmann
,
APL Mater.
8
,
051112
(
2020
).
143.
S.-D.
Lee
,
K.
Akaiwa
, and
S.
Fujita
,
Phys. Status Solidi
10
,
1592
(
2013
).
144.
S.-D.
Lee
,
Y.
Ito
,
K.
Kaneko
, and
S.
Fujita
,
Jpn. J. Appl. Phys.
54
,
030301
(
2015
).
145.
R.
Jinno
,
T.
Uchida
,
K.
Kaneko
, and
S.
Fujita
,
Phys. Status Solidi B
255
,
1700326
(
2018
).
146.
R.
Jinno
,
K.
Kaneko
, and
S.
Fujita
,
AIP Adv.
10
,
115013
(
2020
).
147.
J. P.
McCandless
,
C. S.
Chang
,
K.
Nomoto
,
J.
Casamento
,
V.
Protasenko
,
P.
Vogt
,
D.
Rowe
,
K.
Gann
,
S. T.
Ho
,
W.
Li
,
R.
Jinno
,
Y.
Cho
,
A. J.
Green
,
K. D.
Chabak
,
D. G.
Schlom
,
M. O.
Thompson
,
D. A.
Muller
,
H. G.
Xing
, and
D.
Jena
,
Appl. Phys. Lett.
119
,
062102
(
2021
).
148.
L. J.
Eckert
and
R. C.
Bradt
,
J. Am. Ceram. Soc.
56
,
229
(
1973
).
149.
W. M.
Yim
and
R. J.
Paff
,
J. Appl. Phys.
45
,
1456
(
1974
).
150.
A. F. M.
Anhar Uddin Bhuiyan
,
Z.
Feng
,
H.-L.
Huang
,
L.
Meng
,
J.
Hwang
, and
H.
Zhao
,
APL Mater.
9
,
101109
(
2021
).