Sapphire is a technologically highly relevant material, but it poses many challenges when performing epitaxial thin-film deposition. We have identified and applied the conditions for adsorption-controlled homoepitaxial growth of c-plane sapphire. The films thus grown are atomically smooth, have a controlled termination, and are of outstanding crystallinity. Their chemical purity exceeds that of the substrates. The films exhibit exceptional optical properties, such as a single-crystal-like bandgap and a low density of F+ centers.

Sapphire (α-Al2O3) is a material that features highly valuable properties for fundamental research and applications. It is especially useful as a substrate and functional film for electronic applications. Sapphire combines a large bandgap (9 eV), high dielectric permittivity (9–11), low RF losses, high thermal conductivity (46 W/mK), and a high electric breakdown field with remarkable thermal and mechanical stability.1–4 The use of sapphire as a substrate material has been particularly driven by the ability to deposit high-quality silicon on sapphire for field effect transistor devices5 and to fabricate GaN-based light-emitting diodes on sapphire.6 As a result of these developments and owing to the availability of low-cost, high-quality, single-crystal sapphire wafers of practically any size, the ever-growing market share of sapphire substrates has already surpassed that of any other oxide substrate material.1 Furthermore, being a functional material itself, sapphire is a frequently used high-k gate dielectric7 and is utilized as an active material in solid-state lasers.8,9 Sapphire has also gained attention as an ultrawide-bandgap material for high-power electronics.10–12 

However, the growth of high-quality sapphire films poses several challenges for further important applications of sapphire in electronics. High-quality sapphire films are required for technologies that rely on active or passive sapphire films and, notably, for the use of sapphire as a substrate material. As substrate surfaces are usually rich in impurities and defects induced, for example, by the mechanical processing of the substrates, epitaxial layers of the substrate material are preferentially deposited on the substrate to shield devices from likely blemished substrate surfaces.13–15 Whereas the growth of sapphire single crystals has long been refined, the epitaxial deposition of high-quality sapphire films has proved to be difficult. This is in part because aluminum effusion cells are notoriously difficult to handle in the oxidizing environments needed for sapphire growth by molecular beam epitaxy (MBE). For example, the tendency of Al to creep requires the use of cold-lip effusion cells. However, aluminum oxide tends to grow at their orifices and clog them, thus leaving the effusion cell useless after a short period of usage.16 Nevertheless, a-, m-, and r-plane-oriented sapphire films have been grown by MBE, although they lack atomically flat surfaces,11,12,17,18 and atomically smooth r-plane films have been realized by pulsed laser deposition (PLD).19 Furthermore, compared to single crystals, the optical bandgap of these films is reduced, a degradation that has been attributed to impurities embedded in the films partially originating from the effusion cell crucible.11,20 Even worse, owing to the formation of the preferred γ-phase of Al2O3, the epitaxial growth of the highly important c-plane sapphire films is not even considered possible with MBE.12,18,21 Driven by the technological relevance of the c-plane sapphire surfaces, several efforts have been undertaken to resolve this impediment. Epitaxial c-plane sapphire films, for example, have been grown by electron-beam-assisted PLD. However, the films thus achievable were only several unit cells thick and partially amorphous.22 

In order to provide high-quality sapphire films, we explored the growth of c-plane sapphire by adsorption control. Adsorption-controlled growth implements the self-limiting, stoichiometric, and epitaxial growth of thin films. For film growth by adsorption control, the deposition parameters, i.e., substrate temperature and gas supply, are chosen such that the desired phase of the film to be grown is stable but that all except one of the molecular species arriving at the growing film are volatile. Therefore, when growing stoichiometric films, these species may be supplied in larger fluxes than required to match the desired film composition one-to-one because the excess molecules desorb from the growing film. Adsorption-controlled growth, used with stunning success for depositing III–V heterostructures,23–26 is typically the growth mode that yields the cleanest films of the highest crystalline quality.27,28 However, adsorption-controlled growth is limited to the growth of materials for which the required deposition parameters can be achieved.29 It remains to be explored whether the benefits of adsorption control outweigh the thermodynamic drive of increased point defect formation at the potentially high temperatures required. The growth of oxides by standard adsorption control has been mostly limited to oxides containing highly volatile species, such as ruthenates,30–32 iridates,33–35 bismuthates,36–38 plumbates,39–41 and stannates.42–44 To overcome the limitation of high volatility, organic precursors have also been used, for example, to grow SrTiO3,45–47 although these may introduce undesirable impurities at low substrate temperatures.48 Films of group-III oxides have been grown by adsorption control using suboxide-MBE. Here, the typically high volatility of the suboxide formed in the first reaction step of the oxidation of group-III elements [Eq. (1)] is utilized, with it being supplied in excess.15,49,50 However, this method is more difficult to apply to sapphire films, as both Al and its suboxide are not as volatile.51 For this reason, the adsorption-controlled growth of sapphire has long been a highly desirable epitaxial growth process, although one that has yet to be realized by
(1)
In this work, we explore and reveal the parameter space necessary for the adsorption-controlled growth of sapphire. We find these parameters to be barely achievable with standard MBE. In comparison, the extended parameter space of thermal laser epitaxy (TLE)29,52 allows us to grow atomically smooth films at high growth rates. The crystallinity and bandgap of these films are bulk-like; the film purity exceeds that of the substrate. Epitaxial growth is found to be possible up to an effective substrate temperature of 2000 °C, the temperature at which the back side of the substrate facing the heating laser beam already melts, while films can still be grown on the front side of the substrate. The two key ingredients for achieving high quality ultra-pure sapphire films are, thus, the use of high substrate temperatures and the avoidance of impurities from parasitic heat and hot crucibles.

All CrysTec GmbH c-plane sapphire substrates (miscut <0.1°) were annealed in vacuum at a background pressure of <1 × 10−7 mbar at 1700 °C for 200 s prior to growth. This results in a pure Al-rich surface termination with a double layer step height, as shown in Figs. 1(a) and 1(b).53 MBE films were grown in a Veeco Gen10, and thermal laser epitaxy (TLE) films were grown in an epiray STRATOLAS 50 with a THERMALAS heater. 5N aluminum was loaded in an p-BN crucible in a mid-temp effusion cell for MBE growth and in a sapphire crucible for TLE growth. For MBE, this means that over time, clogging of the effusion cell is expected. Thus, the flux needs to be monitored closely—an issue not present in TLE as the source laser simply breaks through any oxide crust forming. The TLE Al source is operated in a regime where an Al flux with a neglectable Al2O portion is expected.16 The oxygen source for MBE supplied a mixture of ≈80% O3 with O2, while the one for TLE supplied 100% O2 (purified). It can be considered that O3 dissociates to release O at the growth front.54 Substrate temperatures in both systems were measured by a pyrometer. Film growth was monitored in situ by reflection high-energy electron diffraction (RHEED) before and after growth because monitoring during growth produces surface defects. A Bruker DektakXT depth profilometer was utilized to measure the thickness of the resulting films along the position of the sample holder during growth [compare Fig. 2(c)]. To exclude any other effects resulting in shadowing by the substrate holder, in a control experiment, a sapphire substrate was heated to 2000 °C without deposition. In this case, no height difference was observed, as expected. An Asylum Cypher atomic force microscope (AFM) in tapping mode was used to acquire the topography images and a Panalytical Empyrean for x-ray diffraction (XRD).

FIG. 1.

RHEED patterns (left column) and AFM topography images (right column) showing different growth regimes of Al2O3 on c-plane sapphire (TLE). The MBE-grown films show no significant difference and are, therefore, shown in the supplementary information. (a) and (b) show a prepared substrate, (c) and (d) represent growth at T < 900 °C, (e) and (f) at 900 °C ≤ T < 1100 °C, (g) and (h) at 1100 °C ≤ T ≤ 1300 °C , (i) and (j) at 1300 °C < Y ≤ 1800 °C, and (k) and (l) T > 1800 °C.

FIG. 1.

RHEED patterns (left column) and AFM topography images (right column) showing different growth regimes of Al2O3 on c-plane sapphire (TLE). The MBE-grown films show no significant difference and are, therefore, shown in the supplementary information. (a) and (b) show a prepared substrate, (c) and (d) represent growth at T < 900 °C, (e) and (f) at 900 °C ≤ T < 1100 °C, (g) and (h) at 1100 °C ≤ T ≤ 1300 °C , (i) and (j) at 1300 °C < Y ≤ 1800 °C, and (k) and (l) T > 1800 °C.

Close modal
FIG. 2.

(a) Micrograph of the homoepitaxial sapphire film grown at 2000 °C [Figs. 1(k) and 1(l)]. The melt pool and bubbles are frozen due to quenching. (b) The enlargement shows that growth is being precluded at the position of the holder. Step bunching results in steps visible to the naked eye. (c) Schematic of the sample depicting in red the areas that are blocked during growth; the dashed line exemplary shows where profilometry is carried out.

FIG. 2.

(a) Micrograph of the homoepitaxial sapphire film grown at 2000 °C [Figs. 1(k) and 1(l)]. The melt pool and bubbles are frozen due to quenching. (b) The enlargement shows that growth is being precluded at the position of the holder. Step bunching results in steps visible to the naked eye. (c) Schematic of the sample depicting in red the areas that are blocked during growth; the dashed line exemplary shows where profilometry is carried out.

Close modal
Considering the isobaric-isothermal nature of the sapphire growth in TLE, ΔG(T) is given by
(2)
with the change in enthalpy [ΔH(T)] and the change in entropy [ΔS(T)] determined by
(3)
and
(4)
with T0 = 295 K. The heat capacity C(T) is calculated as
(5)
ΔH0, ΔS0, a, b, c, and d are adapted from Ref. 55. For a given chemical reaction with reactants Ri and products Pj, ΔG < 0 can be determined by the sum of the Gibbs free energies of the products, Σj GPj, minus the sum of the Gibbs free energies of the reactants, Σi GRi, i.e., by
(6)
with stoichiometric coefficients pj and ri. The plot in Fig. 3(d) is derived from Eqs. (2)(6).
FIG. 3.

(a) Growth rate Γ as a function of substrate temperature T, with open symbols marking MBE growth and closed symbols marking TLE growth. Black represents amorphous films, and red represents crystalline films. The laser power PL is 300 W, and the oxygen background pressure pox is 2 × 10−3 mbar for the TLE grown films. The Al-flux is 2.2 × 1015 atoms/cm2 at pox = 7 × 10−6 mbar for MBE. The dashed curve represents a model fit. (b) TLE Γ as a function of 1000/PL and (c) a function of pox, with the dashed line marking the pressure for which the mean free path is lower than the distance between source and substrate for an unheated substrate (black) and T = 1600 °C (red). (d) Gibbs free energy (ΔG) as a function of T for the reactions expected in this parameter regime.

FIG. 3.

(a) Growth rate Γ as a function of substrate temperature T, with open symbols marking MBE growth and closed symbols marking TLE growth. Black represents amorphous films, and red represents crystalline films. The laser power PL is 300 W, and the oxygen background pressure pox is 2 × 10−3 mbar for the TLE grown films. The Al-flux is 2.2 × 1015 atoms/cm2 at pox = 7 × 10−6 mbar for MBE. The dashed curve represents a model fit. (b) TLE Γ as a function of 1000/PL and (c) a function of pox, with the dashed line marking the pressure for which the mean free path is lower than the distance between source and substrate for an unheated substrate (black) and T = 1600 °C (red). (d) Gibbs free energy (ΔG) as a function of T for the reactions expected in this parameter regime.

Close modal

TEM specimens of the films were prepared by mechanical wedge polishing followed by Ar ion beam milling at L-N2 temperature [micrometer 140 (2021) 102 979]. Scanning transmission electron microscopy (STEM) studies were performed using spherical aberration-corrected STEM (JEM-ARM200F, JEOL Co. Ltd.) equipped with a cold-field emission gun and a DCOR probe Cs corrector (CEOS GmbH) operated at 200 kV. The STEM images were obtained using an ADF detector with a convergent semi-angle of 20.4 mrad. The corresponding collection semi-angles for high-angle annular dark-field (HAADF) and low-angle ADF (LAADF) imaging were 70–300 mrad and 40–100 mrad, respectively. Eight serial images were acquired with a short dwell time (2 µs/pixel), aligned, and then added to improve the signal-to-noise ratio (SNR) and to minimize distortion of the HAADF and LAADF images. STEM images were denoised using the bandpass filter.

ToF SIMS was measured using a ToF SIMS NCS instrument (IONTOF). For mass-spectrometry probing, monoatomic Bi1+ ions were used in spectrometry mode, accelerated by a voltage of 30 keV, with currents in the range of 2.1–2.4 pA on a 50 × 50 μm2 analysis area. Ion spectra were acquired with positive and negative polarity to access a variety of different atomic species. Data analysis was performed using SurfaceLab 7.1 software. Depth profiles were acquired by a cyclic sputter-probe series. (Note that only data for positive ions are shown in the paper.) The sputter and analysis guns were operated in non-interlaced mode with an additional flood gun in order to avoid sample charging. For sputtering, an oxygen ion source was employed using a 2 keV acceleration voltage and 479 nA current on a 400 × 400 μm2 crater area. The ion intensities shown in the depth profiles are averaged ten-fold.

The real part of optical conductivity σ1 was determined from ellipsometric parameters Ψ and ∆ measured by a variable-angle spectroscopic ellipsometer VASE (J. A. Woollam Inc.) in the energy range from 1.5 to 6.0 eV. Ψ and ∆ are defined by the complex Fresnel coefficients as tan(Ψ)ei∆ = rp/rs, where rp and rs are reflection coefficients for light polarized parallel and perpendicular to the plane of incidence. The optical measurements were performed at room temperature and at various angles of incidence (from 70° to 80°). UV–Vis spectra were obtained utilizing a Cary 5000 (Agilent Technologies) capable of measuring absorptance in an integrating sphere. Transmittance spectra in the UVC spectral range were measured at room temperature (RT) using a monochromatic light in a wavelength range of 135–300 nm from a deuterium lamp (Hamamatsu L11798) dispersed by a 20-cm focal-length Czerny–Turner monochromator (Bunkoukeiki KV-200) equipped with a 1200 groove/mm grating.56 The samples were set on the rotational sample stage, which was loaded into the vacuum chamber. The mirror optical system and monochromator were connected to the chamber through MgF2 viewports. During the measurements, the entire optical path was purged with ≥99.99% pure nitrogen gas with a gas flow rate of 7 L/min to maintain a residual oxygen concentration lower than 1 ppm. The spectral resolution at 160 nm was set at 0.1 eV (2 nm) with a slit width of 0.5 mm. A 0.5 × 0.6 mm2 rectangular light spot was defined by the slit widths of the monochromator and the luminance point size of the D2 lamp. The transmitted light was detected using a photomultiplier tube (Hamamatsu R374) through a window coated with sodium salicylate. A quartz glass long-pass filter was used in the wavelength range greater than 200 nm to exclude second-order light. Lock-in detection using an optical chopper was employed for the measurements.

Attempts at homoepitaxial growth of c-plane sapphire usually result in the formation of the Al2O3 γ-phase, which, not being sapphire, is undesirable for electronic applications.12,18 Similarly, reflection high-energy electron diffraction (RHEED) patterns reveal amorphous growth [Fig. 1(c), Fig. S1] for both MBE and TLE growth at substrate temperatures T < 900 °C and the parameters found in Tables I and II. The surfaces of these films are rough, with a root-mean-square (RMS) value > 5 nm [Fig. 1(d), Fig. S1]. The difficulty of obtaining the α-phase (sapphire) is determined by the high temperature required for its phase formation under equilibrium conditions.57 Consequentially, growth at higher T (900 °C ≤ T < 1000 °C) yields crystalline homoepitaxial films for MBE and TLE, as confirmed by streaky and spotty RHEED patterns [Fig. 1(e), Fig. S1]. Nevertheless, the films remain rough [Fig. 1(f), Fig. S1]. As in most MBE systems with non-laser substrate heaters, the T achievable with the MBE setup we used is limited.29 Although a homoepitaxial deposition of sapphire films is thus achievable under the conditions described, their surfaces are not smooth.

TABLE I.

Growth regimes of sapphire observed in TLE and MBE for the following parameters: TLE: PL ≥ 250 W, pox ≥ 1 × 10−3 mbar O2. MBE: ΦAl = 2.2×1015 atoms/cm2, pox = 7 × 10−6 mbar, 90% O3, and 10% O2.

T (°C)Growth
RTAmorphous
900 Multilayer Epitaxial and adsorption controlled 
1300 Step flow 
1800 Step bunching 
T (°C)Growth
RTAmorphous
900 Multilayer Epitaxial and adsorption controlled 
1300 Step flow 
1800 Step bunching 
TABLE II.

Adsorption-controlled growth regimes for MBE and TLE identified in this work.

Parameter regimeIdeal range
TLE T ≥ 900 °C 1300 °C < T ≤ 1800 °C 
pox ≥ 1 × 10−3 mbar O2 
PL ≥ 250 W 
MBE 900 °C ≤ T 
pox = 7 × 10−6 mbar 80 % O3 and 20 % O2 
ΦAl = 2.2×1015 atoms/cm2 
Parameter regimeIdeal range
TLE T ≥ 900 °C 1300 °C < T ≤ 1800 °C 
pox ≥ 1 × 10−3 mbar O2 
PL ≥ 250 W 
MBE 900 °C ≤ T 
pox = 7 × 10−6 mbar 80 % O3 and 20 % O2 
ΦAl = 2.2×1015 atoms/cm2 

Enabled by a CO2-laser-based substrate heating system, substrate temperatures of ≫1000 °C are available in the TLE system used here.29,52 Indeed, deposition at 1100 °C yields much smoother films with RMS ≈0.5 nm; see Fig. 1(h). The islands visible in this figure follow the quasi-hexagonal motif of the crystal structure and are atomically smooth, with step heights equaling one monolayer. The respective RHEED pattern [Fig. 1(g)] shows a hitherto unreported complex reconstruction, probably resulting from the island structure. Thus, in the multilayer growth mode, the single-reconstruction surface of the prepared sapphire substrates is not preserved. The observed islands are likely a result of kinetic roughening, as observed for the homoepitaxial growth of other materials.58 The island size increases with increasing temperature until a shift toward the growth regime discussed below is observed.

Increasing T beyond 1300 °C results in a transition from the multilayer growth mode to step-flow growth. Independent of the film thickness, the resulting surfaces are practically indistinguishable from those of the prepared substrates. Figure 1(j) shows the corresponding surface of a film grown at 1600 °C (RMS = 0.23 nm). Each of the uniform surface steps, which border terraces with a remarkable width of ≈2 µm, has a height of two layers along [0001]. These double steps and the Al-terminated √31 × √31R + 9° reconstruction [Fig. 1(i)] are also present on the prepared substrates.53 These films thus retain the Al-termination and atomic smoothness of the substrate.

This nominally ideal surface structure is observed up to 1800 °C. At higher T, step bunching occurs. This is illustrated in Fig. 2(a), which shows a film surface after growth at T ≈ 2000 °C. As mentioned above, the back side of the substrate melts at T ≈ 2000 °C, generating the melt pool and droplets visible in the optical micrograph. Nevertheless, the sapphire films grown on the slightly cooler front side show sizable step bunching [Fig. 2(b)]. Figure 2(b) also shows the sample areas in which film growth is blocked by a shadow cast by the substrate holder, such that the film thicknesses can be measured straightforwardly by a profilometer at these places. The steps reach such a height that they are visible to the naked eye, whereas the terraces remain atomically flat, as shown by AFM with step heights of various multitudes of two monolayers along [0001] [Fig. 1(l)]. The RHEED images still represent a singly oriented √31 × √31R + 9° reconstruction [Fig. 1(k)], showing that the Al-rich surface termination is stable on the entire sample even for growth at T ≈ 2000 °C.

The different growth regimes observed above are for substrates with similar miscuts. Changes in miscut may alter the transition temperatures. In addition, the surface structure of c-plane sapphire is miscut dependent.59,60 We observed, however, that in the step-flow growth regime, the initial surface structure was maintained.

Figure 1 and Fig. S1 reveal a drastic change at T = 900 °C, where the growth for both MBE and TLE transitions from amorphous to epitaxial c-plane sapphire (summarized in Table I). Therefore, we investigate whether this transition also affects the growth rate Γ. Indeed, the growth rate Γ decreases at T = 900 °C for both MBE and TLE film growth [Fig. 3(a)], suggesting that the c-plane sapphire films grow in an adsorption-controlled manner for T ≥ 900 °C up to a hypothetical upper limit (beyond the melting point of the sapphire substrate)—which implies that each species not forming sapphire is desorbed. To explore whether this is indeed the case, we recall the two reaction steps for the formation [Eq. (1)] of sapphire by oxidation of Al,49,50,61–63
(7)

We consider O to result from the O3 dissociation and thermalization of O2, respectively. Indeed, the same reaction with O2 as a reactant may occur. Thus, in the hypothetical parameter range of adsorption-controlled growth, Al2O must be volatile (in addition, Al can be volatile as well). This Al2O volatility is expected to have three effects on the Al2O3 growth rate Γ:49,50

  1. Γ tends to be small compared to standard growth because Al2O desorbs.

  2. Γ will be independent of the Al supply because Γ is limited only by the oxygen/ozone flux.

  3. Γ will increase with the flux of oxygen/ozone impinging on the substrate.

For adsorption-controlled growth, these three effects are highly characteristic. We will use them as a fingerprint of sapphire film growth through adsorption control.

First, exploring the existence of effect (1), we find that, for the parameters listed in Table II, the growth rate of sapphire films is smaller at T ≥ 900 °C than at lower T, where the films do not exhibit the α-phase, for MBE as well as for TLE [Fig. 3(a)]. This behavior is in agreement with effect (1). Note that the Γ values achievable for growth by MBE (≈4 µm/h) and by TLE (≈15 µm/h) at T < 900 °C are considerable. Even at T ≥ 900 °C, Γ is high, namely, ≈0.5 µm/h.

In addition, effect (2) is present, as shown by Fig. 3(b), which displays the dependence of Γ on the source laser power PL for pox = 2 × 10−3 mbar and T = 1600 °C. The source laser power correlates with the flux as described in detail in Refs. 64 and 29, and for Al sources specifically in Ref. 16. For comparison, for the same PL, Fig. 3(b) shows Γ also for film deposition on a nominally unheated substrate (T ≈ 80 °C). For the latter case, an Arrhenius-like dependence of Γ on PL is expected,16 which is indeed observed. However, at T = 1600 °C and pox = 2 × 10−3 mbar, Γ does not depend on PL and, therefore, not on the Al supply.

To discuss the existence of effect (3), we present the dependence of Γ on pox in Fig. 3(c). For the unheated substrate, pox does not alter Γ. For pox > 9.6 × 10−3 mbar, for which the mean free path is shorter than the substrate-source distance, Γ is lower than at smaller pox. At T = 1600 °C and for 1 × 10−3 mbar ≤ pox ≤ 9.6 × 10−3 mbar, Γ is proportional to pox, in agreement with effect (3). Γ vanishes for pox < 1 × 10−3 mbar. A likely explanation is the lack of oxygen to complete the second reaction step of Eq. (7). All available oxygen goes to form Al2O, which is volatile at such values of T and, therefore, desorbs. There is also no growth observed for even lower pressures, indicating that elemental Al is also volatile at these T. For pox ≥ 1 × 10−3 mbar, however, Γ increases with pox. With more oxygen available, the oxidation of Al2O to Al2O3 increases, thus reducing Al2O desorption.

For the parameters in Table II, sapphire growth shows all three effects that together provide a fingerprint of the adsorption-controlled regime. Therefore, we conclude that the adsorption-controlled growth of c-plane sapphire is indeed achievable.

For the growth process, there may appear to be some discrepancy between the conclusion that the films grow in an adsorption-controlled manner and that the growth rate Γ is approximately independent of T for T > 900 °C; see Fig. 3(a). In fact, however, there is no discrepancy.

At first glance, the T-independence of Γ is unexpected because one would expect that Γ decreases with increasing T in the adsorption-controlled regime. This is because adsorption decreases with T, as is the case for many group-III oxides. However, a closer look at the Gibbs free energies of the reactions involved in film growth reveals that, for the growth of Al2O3 films, this saturation of Γ may be caused by Al chemically decomposing the Al2O3 to Al2O; see Eq. (8) and Fig. 3(d). The increased Al2O formation may in turn have a self-catalytic effect on the growth, ultimately resulting in a more or less stable growth rate at T > 900 °C.

Figure 3(d) provides data for the Gibbs free energy differences for all expected reactions (ΔG) as a function of T. Figure 3(d) omits reactions that are not expected for kinetic limitations.49,51,62,63,65 The data shown were calculated as described in the Methods section based on values published in Refs. 66 and 67. It is apparent that, in addition to the two reaction steps of Eq. (7), the above-mentioned etching process
(8)
exists with ΔG < 0. Note that, as shown in Fig. 3(d), ΔG becomes slightly more negative with increasing T. For this case, the standard model used to characterize the adsorption-controlled growth of group-III oxides65 applied to Eqs. (7) and (8) readily yields a good fit to the measured data [Fig. 3(a), dashed line]. Despite this success, this fit must be regarded with skepticism because, for example, the underlying model does not consider the effects of the T-dependent adhesion and surface diffusion processes on Γ. Self-catalysis as an explanation for the observed growth plateau is at this point, thus, to be considered a hypothesis. Extended experimental work and models will show if it indeed provides the underlying mechanism.

The data shown in Figs. 1 and 3 allow us to deduce the ideal growth regime for the adsorption-controlled homoepitaxial growth of c-plane sapphire, which is also summarized in Table II. The range 1300 °C < T ≤ 1800 °C results in atomically flat surfaces with a pure Al-termination under these conditions. Utilizing high oxygen background pressures allows us to maximize the growth rates, in this particular case to Γ ≈ 3 µm/h.

Having established the optimal parameters for the adsorption-controlled growth of c-plane sapphire, we now explore the properties of films grown at T = 1600 °C and pox = 8 × 10−3 mbar, i.e., adsorption controlled.

The RHEED patterns of the films, a typical example of which is shown in Fig. 1(i), already indicate an exceptional crystalline quality of the film surfaces. Moreover, the bulk is characterized by excellent crystallinity. Neither θ-2θ x-ray diffractograms nor rocking-curve measurements allowed us to observe any difference in the crystallinity between the films and the substrates. To gain further microscopic information on the film microstructures at the atomic scale, we imaged sample cross sections by STEM. A wide-field HAADF-STEM image is shown in Fig. 4(a). No crystallographic defects are visible in the film at this magnification, and the interface between the film and the substrate is not even discernible. However, as described above, the interface’s location and existence were determined by profilometry. LAADF is especially sensitive to defects,68 but microscopy in this mode also failed to reveal any defects [Fig. 4(b)]. Even imaging the interface region at higher magnification does not reveal the interface [Figs. 4(c) and 4(d)]. Indeed, we found no microscopic difference between the single-crystal sapphire substrates and the sapphire films that were grown under the conditions of adsorption-controlled growth (Table II).

FIG. 4.

(a) ADF–STEM image showing the long-range high quality of the grown sapphire film and its interface to the substrate, the position of which is marked by V symbols. (b) Same area imaged in LAADF, which is more sensitive to strain and non-stoichiometry, neither of which are observed. (c) Enlargement of the interface region in (a) that further shows the high crystal quality. (d) HAADF–STEM of the interface region, which remains indistinguishable. Red dots correspond to O and green dots correspond to Al. All images are taken along the [11̄00] direction with the growth direction from left to right.

FIG. 4.

(a) ADF–STEM image showing the long-range high quality of the grown sapphire film and its interface to the substrate, the position of which is marked by V symbols. (b) Same area imaged in LAADF, which is more sensitive to strain and non-stoichiometry, neither of which are observed. (c) Enlargement of the interface region in (a) that further shows the high crystal quality. (d) HAADF–STEM of the interface region, which remains indistinguishable. Red dots correspond to O and green dots correspond to Al. All images are taken along the [11̄00] direction with the growth direction from left to right.

Close modal

As the possible presence of chemical impurities is a key issue affecting the quality of homoepitaxial sapphire films,11,20 we also performed chemical analyses on these films by time-of-flight secondary ion mass spectroscopy (ToF SIMS). Figure 5 compares the corresponding results for several ionic species dependent on the thickness d of (a) a sample grown by MBE and (b) a sample grown by TLE.

FIG. 5.

ToF SIMS measurements of homoepitaxial sapphire films grown in the adsorption-controlled regime at substrate temperatures of (a) 900 °C in MBE and (b) 1600 °C in TLE. The position of the interface is marked by dashed lines.

FIG. 5.

ToF SIMS measurements of homoepitaxial sapphire films grown in the adsorption-controlled regime at substrate temperatures of (a) 900 °C in MBE and (b) 1600 °C in TLE. The position of the interface is marked by dashed lines.

Close modal

First, we note that the TLE film grown under adsorption-controlled growth conditions is measured to be very pure. Indeed, the film shows fewer impurities than the underlying single-crystal substrate. The MBE-grown film shows fewer B+ impurities than previous reports.11 B originates from the p-BN crucibles used for Al sources, and the adsorption-controlled growth may minimize its incorporation. The dominant impurity detected in the MBE grown film is Ti+. Titanium is used for back-coating the substrates.

The main other impurities detected in both films are Ga+, Si+, and Fe+, which are all present in the single-crystal substrate. For the MBE-grown film, a typical enrichment of these impurities at the interface and surface is observed. Interestingly, this is not the case for the TLE-grown film, where the Ga+ and Fe+ signals fall below the detection level within a couple of nanometers. Only the Si+ signal has a small peak at the substrate film interface and does not deplete within the film. The omnipresence of Si+ is well known for numerous materials.69,70 Indeed, the high solubility of Si in sapphire over a wide temperature and pressure range makes it difficult to avoid this impurity.71 

Furthermore, owing to their high purity, the c-plane sapphire films grown by adsorption control are not subject to the bandgap reduction that vexes standard epitaxial sapphire films.11,20 Figure 6 shows the corresponding ultraviolet c (UVC) transmittance spectrum of a homoepitaxial c-plane film grown by adsorption control along with the spectra of single crystals. The band gaps of the film and the substrates are identical.

FIG. 6.

Transmittance spectra in the UVC spectral range for an as-received (blue) and annealed (green) c-plane sapphire single crystal and a homoepitaxial film (red). No reduction of the bandgap is observed for the film.

FIG. 6.

Transmittance spectra in the UVC spectral range for an as-received (blue) and annealed (green) c-plane sapphire single crystal and a homoepitaxial film (red). No reduction of the bandgap is observed for the film.

Close modal

The higher purity of the TLE-grown epitaxial film compared to the single-crystalline substrate is also revealed by ultraviolet–visible (UV–Vis) spectroscopy; see Fig. S2. This is because F+ centers in sapphire cause light absorbance at 260 nm,72 meaning that the lower absorbance at this wavelength provides evidence of fewer F+ centers in the films compared to the substrates. Consistent with this, the lower optical conductivity measured by ellipsometry in the films indicates that the defect density in the films is at least as low as in the annealed substrates;73 see Fig. S3.

We have identified the deposition conditions required to achieve adsorption-controlled homoepitaxial growth of sapphire by TLE and MBE. These conditions are summarized in Table I.

The films are characterized by excellent structural and optical properties. Compared to standard sapphire films, the films grown by adsorption control have lower defect densities despite their very high growth temperatures. The benefits of adsorption control thus outweigh the thermodynamically driven increase in defect density with increasing temperature.

The ability to grow homoepitaxial sapphire films that surpass the structural and optical properties of sapphire single crystals opens new perspectives. For example, the growth of such homoepitaxial layers allows detrimental effects resulting from defects incorporated into the surface of sapphire single crystals to be minimized, which we expect to yield enhanced mobilities, fewer losses, and enhanced quantum decoherence times.74 Furthermore, such films, with their excellent properties, can serve as functional layers in electronics, e.g., for high-power applications.10–12 

As the high quality of these films is induced by the adsorption-controlled growth mode rather than by specific properties of the Al–O material system, we anticipate that additional oxide films as well as non-oxide material systems will be obtainable in excellent quality by following the approach presented here.

See the supplementary material for the following: Fig. S1 shows RHEED and AFM data for MBE grown films. Figure S2 shows UV–Vis spectra and Fig. S3 shows optical conductivity obtained by ellipsometry.

The authors thank Bettina V. Lotsch and Alexander Boris for providing the infrastructure for UV–Vis spectroscopy and ellipsometry, respectively.

The authors have no conflicts to disclose.

Lena N. Majer: Conceptualization (equal); Data curation (lead); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting). Tolga Acartürk: Data curation (supporting); Formal analysis (supporting); Methodology (supporting); Visualization (supporting); Writing – review & editing (supporting). Peter A. van Aken: Formal analysis (supporting); Methodology (supporting); Supervision (supporting); Writing – review & editing (equal). Wolfgang Braun: Conceptualization (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (equal); Supervision (supporting); Writing – review & editing (equal). Luca Camuti: Data curation (supporting); Formal analysis (supporting); Methodology (supporting). Johan Eckl-Haese: Data curation (supporting); Investigation (supporting). Jochen Mannhart: Conceptualization (supporting); Formal analysis (supporting); Investigation (supporting); Supervision (equal); Validation (equal); Writing – review & editing (lead). Takeyoshi Onuma: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (equal); Validation (supporting); Visualization (supporting); Writing – original draft (supporting). Ksenia S. Rabinovich: Data curation (supporting); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Writing – review & editing (supporting). Darrell G. Schlom: Conceptualization (supporting); Methodology (supporting); Supervision (equal). Sander Smink: Conceptualization (equal); Data curation (equal); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Supervision (supporting); Validation (equal); Visualization (supporting); Writing – review & editing (equal). Ulrich Starke: Methodology (supporting); Supervision (supporting); Validation (supporting); Writing – review & editing (equal). Jacob Steele: Data curation (supporting); Investigation (supporting). Patrick Vogt: Conceptualization (supporting); Data curation (supporting); Formal analysis (equal); Investigation (equal); Methodology (supporting); Validation (equal); Visualization (supporting); Writing – review & editing (equal). Hongguang Wang: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Methodology (supporting); Validation (supporting); Visualization (equal); Writing – review & editing (equal). Felix V. E. Hensling: Conceptualization (lead); Data curation (equal); Formal analysis (equal); Investigation (lead); Supervision (lead); Validation (lead); Visualization (equal); Writing – original draft (lead).

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

1.
M. S.
Akselrod
and
F. J.
Bruni
, “
Modern trends in crystal growth and new applications of sapphire
,”
J. Cryst. Growth
360
(
1
),
134
145
(
2012
).
2.
H.
Birey
, “
Dielectric properties of aluminum oxide films
,”
J. Appl. Phys.
49
(
5
),
2898
2904
(
1978
).
3.
D. G.
Cahill
,
S.-M.
Lee
, and
T. I.
Selinder
, “
Thermal conductivity of κ-Al2O3 and α-Al2O3 wear-resistant coatings
,”
J. Appl. Phys.
83
(
11
),
5783
5786
(
1998
).
4.
E. R.
Dobrovinskaya
,
L. A.
Lytvynov
, and
V.
Pishchik
, “
Properties of sapphire
,” in
Sapphire: Material, Manufacturing, Applications
(
Springer Science + Business Media
,
New York
,
2009
), pp.
55
176
.
5.
T.
Nakamura
,
H.
Matsuhashi
, and
Y.
Nagatomo
, “
Silicon on sapphire (SOS) device technology
,”
Oki Tech. Rev.
71
(
200
),
66
69
(
2004
).
6.
G.
Li
,
W.
Wang
,
W.
Yang
,
Y.
Lin
,
H.
Wang
,
Z.
Lin
, and
S.
Zhou
, “
GaN-based light-emitting diodes on various substrates: A critical review
,”
Rep. Prog. Phys.
79
(
5
),
056501
(
2016
).
7.
B.
Wang
,
W.
Huang
,
L.
Chi
,
M.
Al-Hashimi
,
T. J.
Marks
, and
A.
Facchetti
, “
High-k gate dielectrics for emerging flexible and stretchable electronics
,”
Chem. Rev.
118
(
11
),
5690
5754
(
2018
).
8.
T. H.
Maiman
, “
Stimulated optical radiation in ruby
,”
Nature
187
(
4736
),
493
494
(
1960
).
9.
P.
Moulton
, “
Ti-doped sapphire: Tunable solid-state laser
,”
Opt. News
8
(
6
),
9
(
1982
).
10.
J. B.
Varley
, “
First-principles calculations of structural, electrical, and optical properties of ultra-wide bandgap (AlxGa1-x)2O3 alloys
,”
J. Mater. Res.
36
(
23
),
4790
4803
(
2021
).
11.
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
, “
Crystal orientation dictated epitaxy of ultrawide-bandgap 5.4- to 8.6-eV α-(AlGa)2O3 on m-plane sapphire
,”
Sci. Adv.
7
(
2
),
eabd5891
(
2021
).
12.
H.
Okumura
, “
Sn and Si doping of α-Al2O3 (10-10) layers grown by plasma-assisted molecular beam epitaxy
,”
Jpn. J. Appl. Phys.
61
(
12
),
125505
(
2022
).
13.
Y.
Shiraki
, “
Silicon molecular beam epitaxy
,”
Prog. Cryst. Growth Charact.
12
(
1–4
),
45
66
(
1986
).
14.
Ü.
Özgür
,
V.
Avrutin
, and
H.
Morkoç
, “
Zinc oxide materials and devices grown by molecular beam epitaxy
,” in
Molecular Beam Epitaxy
(
Elsevier
,
2018
), pp.
343
375
.
15.
K.
Azizie
,
F. V. E.
Hensling
,
C. A.
Gorsak
,
Y.
Kim
,
N. A.
Pieczulewski
,
D. M.
Dryden
,
M. K. I.
Senevirathna
,
S.
Coye
,
S.-L.
Shang
,
J.
Steele
,
P.
Vogt
,
N. A.
Parker
,
Y. A.
Birkhölzer
,
J. P.
McCandless
,
D.
Jena
,
H. G.
Xing
,
Z.-K.
Liu
,
M. D.
Williams
,
A. J.
Green
,
K.
Chabak
,
D. A.
Muller
,
A. T.
Neal
,
S.
Mou
,
M. O.
Thompson
,
H. P.
Nair
, and
D. G.
Schlom
, “
Silicon-doped β-Ga2O3 films grown at 1 µm/h by suboxide molecular-beam epitaxy
,”
APL Mater.
11
(
4
),
041102
(
2023
).
16.
T. J.
Smart
,
F. V. E.
Hensling
,
D. Y.
Kim
,
L. N.
Majer
,
Y. E.
Suyolcu
,
D.
Dereh
,
D. G.
Schlom
,
D.
Jena
,
J.
Mannhart
, and
W.
Braun
, “
Why thermal laser epitaxy aluminum sources yield reproducible fluxes in oxidizing environments
,”
J. Vac. Sci. Technol., A
41
(
4
),
042701
(
2023
).
17.
T.
Oshima
,
Y.
Kato
,
M.
Imura
,
Y.
Nakayama
, and
M.
Takeguchi
, “
α-Al2O3/Ga2O3superlattices coherently grown on r-plane sapphire
,”
Appl. Phys. Express
11
(
6
),
065501
(
2018
).
18.
R.
Kumaran
,
T.
Tiedje
,
S. E.
Webster
,
S.
Penson
, and
W.
Li
, “
Epitaxial Nd-doped α-(Al1−xGax)2O3 films on sapphire for solid-state waveguide lasers
,”
Opt. Lett.
35
(
22
),
3793
(
2010
).
19.
M.
Lorenz
,
S.
Hohenberger
,
E.
Rose
, and
M.
Grundmann
, “
Atomically stepped, pseudomorphic, corundum-phase (Al1-xGax)2O3 thin films (0 ≤ x < 0.08) grown on R-plane sapphire
,”
Appl. Phys. Lett.
113
(
23
),
231902
(
2018
).
20.
Z.
Chen
,
M.
Arita
,
K.
Saito
,
T.
Tanaka
, and
Q.
Guo
, “
Epitaxial growth of (AlxGa1−x)2O3 thin films on sapphire substrates by plasma assisted pulsed laser deposition
,”
AIP Adv.
11
(
3
),
035319
(
2021
).
21.
Z.
Wan
and
J. R.
Schmidt
, “
Competing growth pathways of γ- and α-Al2O3 during solid-phase epitaxy: The critical role of interfacial reconstruction
,”
J. Phys. Chem. C
128
,
10634
(
2024
).
22.
T.
Maeda
,
M.
Yoshimoto
,
T.
Ohnishi
,
G. H.
Lee
, and
H.
Koinuma
, “
Orientation-defined molecular layer epitaxy of α-Al2O3 thin films
,”
J. Cryst. Growth
177
(
1–2
),
95
101
(
1997
).
23.
K. G.
Guenther
, “
Aufdampfschichten aus halbleitenden III-V Verbindungen
,”
Naturwissenschaften
45
(
17
),
415
416
(
1958
).
24.
J. R.
Arthur
, “
Interaction of Ga and As2 molecular beams with GaAs surfaces
,”
J. Appl. Phys.
39
(
8
),
4032
4034
(
1968
).
25.
R.
Heckingbottom
,
G. J.
Davies
, and
K. A.
Prior
, “
Growth and doping of gallium arsenide using molecular beam epitaxy (MBE): Thermodynamic and kinetic aspects
,”
Surf. Sci.
132
(
1–3
),
375
389
(
1983
).
26.
H.
Seki
and
A.
Koukitu
, “
Thermodynamic analysis of molecular beam epitaxy of III–V semiconductors
,”
J. Cryst. Growth
78
(
2
),
342
352
(
1986
).
27.
H.
Freller
and
K. G.
Günther
, “
Three-temperature method as an origin of molecular beam epitaxy
,”
Thin Solid Films
88
(
4
),
291
307
(
1982
).
28.
J. Y.
Tsao
,
Materials Fundamentals of Molecular Beam Epitaxy
(
Academic Press
,
Boston
,
1993
).
29.
F. V. E.
Hensling
,
W.
Braun
,
D. Y.
Kim
,
L. N.
Majer
,
S.
Smink
,
B. D.
Faeth
, and
J.
Mannhart
, “
State of the art, trends, and opportunities for oxide epitaxy
,”
APL Mater.
12
(
4
),
040902
(
2024
).
30.
H. P.
Nair
,
J. P.
Ruf
,
N. J.
Schreiber
,
L.
Miao
,
M. L.
Grandon
,
D. J.
Baek
,
B. H.
Goodge
,
J. P. C.
Ruff
,
L. F.
Kourkoutis
,
K. M.
Shen
, and
D. G.
Schlom
, “
Demystifying the growth of superconducting Sr2RuO4 thin films
,”
APL Mater.
6
(
10
),
101108
(
2018
).
31.
B.
Burganov
,
C.
Adamo
,
A.
Mulder
,
M.
Uchida
,
P. D. C.
King
,
J. W.
Harter
,
D. E.
Shai
,
A. S.
Gibbs
,
A. P.
Mackenzie
,
R.
Uecker
,
M.
Bruetzam
,
M. R.
Beasley
,
C. J.
Fennie
,
D. G.
Schlom
, and
K. M.
Shen
, “
Strain control of fermiology and many-body interactions in two-dimensional ruthenates
,”
Phys. Rev. Lett.
116
(
19
),
197003
(
2016
).
32.
H. P.
Nair
,
Y.
Liu
,
J. P.
Ruf
,
N. J.
Schreiber
,
S. L.
Shang
,
D. J.
Baek
,
B. H.
Goodge
,
L. F.
Kourkoutis
,
Z. K.
Liu
,
K. M.
Shen
, and
D. G.
Schlom
, “
Synthesis science of SrRuO3 and CaRuO3 epitaxial films with high residual resistivity ratios
,”
APL Mater.
6
(
4
),
046101
(
2018
).
33.
M.
Uchida
,
Y. F.
Nie
,
P. D. C.
King
,
C. H.
Kim
,
C. J.
Fennie
,
D. G.
Schlom
, and
K. M.
Shen
, “
Correlated vs. conventional insulating behavior in the Jeff=1/2 vs. 3/2 bands in layered iridate Ba2IrO4
,”
Phys. Rev. B
90
(
7
),
075142
(
2014
).
34.
Y. F.
Nie
,
P. D. C.
King
,
C. H.
Kim
,
M.
Uchida
,
H. I.
Wei
,
B. D.
Faeth
,
J. P.
Ruf
,
J. P. C.
Ruff
,
L.
Xie
,
X.
Pan
,
C. J.
Fennie
,
D. G.
Schlom
, and
K. M.
Shen
, “
Interplay of spin-orbit interactions, dimensionality, and octahedral rotations in semimetallic SrIrO3
,”
Phys. Rev. Lett.
114
(
1
),
016401
(
2015
).
35.
J. K.
Kawasaki
,
M.
Uchida
,
H.
Paik
,
D. G.
Schlom
, and
K. M.
Shen
, “
Evolution of electronic correlations across the rutile, perovskite, and Ruddelsden-Popper iridates with octahedral connectivity
,”
Phys. Rev. B
94
(
12
),
121104
(
2016
).
36.
S.
Migita
,
Y.
Kasai
,
H.
Ota
, and
S.
Sakai
, “
Self-limiting process for the bismuth content in molecular beam epitaxial growth of Bi2Sr2CuOy thin films
,”
Appl. Phys. Lett.
71
(
25
),
3712
3714
(
1997
).
37.
C. D.
Theis
,
J.
Yeh
,
D. G.
Schlom
,
M. E.
Hawley
,
G. W.
Brown
,
J. C.
Jiang
, and
X. Q.
Pan
, “
Adsorption-controlled growth of Bi4Ti3O12 by reactive MBE
,”
Appl. Phys. Lett.
72
(
22
),
2817
2819
(
1998
).
38.
S.
Migita
,
H.
Ota
,
H.
Fujino
,
Y.
Kasai
, and
S.
Sakai
, “
Epitaxial Bi4Ti3O12 thin film growth using Bi self-limiting function
,”
J. Cryst. Growth
200
(
1–2
),
161
168
(
1999
).
39.
C. D.
Theis
and
D. G.
Schlom
, “
Epitaxial lead titanate grown by MBE
,”
J. Cryst. Growth
174
(
1–4
),
473
479
(
1997
).
40.
G.
Rispens
and
B.
Noheda
, “
Ultra-thin lead titanate films grown by molecular beam epitaxy
,”
Integr. Ferroelectr.
92
(
1
),
30
39
(
2007
).
41.
C. D.
Theis
,
J.
Yeh
,
D. G.
Schlom
,
M. E.
Hawley
, and
G. W.
Brown
, “
Adsorption-controlled growth of PbTiO3 by reactive molecular beam epitaxy
,”
Thin Solid Films
325
(
1–2
),
107
114
(
1998
).
42.
H.
Paik
,
Z.
Chen
,
E.
Lochocki
,
H.
Ariel Seidner
,
A.
Verma
,
N.
Tanen
,
J.
Park
,
M.
Uchida
,
S.
Shang
,
B.-C.
Zhou
,
M.
Brützam
,
R.
Uecker
,
Z.-K.
Liu
,
D.
Jena
,
K. M.
Shen
,
D. A.
Muller
, and
D. G.
Schlom
, “
Adsorption-controlled growth of La-doped BaSnO3 by molecular-beam epitaxy
,”
APL Mater.
5
(
11
),
116107
(
2017
).
43.
A. B.
Mei
,
L.
Miao
,
M. J.
Wahila
,
G.
Khalsa
,
Z.
Wang
,
M.
Barone
,
N. J.
Schreiber
,
L. E.
Noskin
,
H.
Paik
,
T. E.
Tiwald
,
Q.
Zheng
,
R. T.
Haasch
,
D. G.
Sangiovanni
,
L. F. J.
Piper
, and
D. G.
Schlom
, “
Adsorption-controlled growth and properties of epitaxial SnO films
,”
Phys. Rev. Mater.
3
(
10
),
105202
(
2019
).
44.
J.
Park
,
H.
Paik
,
K.
Nomoto
,
K.
Lee
,
B.-E.
Park
,
B.
Grisafe
,
L.-C.
Wang
,
S.
Salahuddin
,
S.
Datta
,
Y.
Kim
,
D.
Jena
,
H. G.
Xing
, and
D. G.
Schlom
, “
Fully transparent field-effect transistor with high drain current and on-off ratio
,”
APL Mater.
8
(
1
),
011110
(
2020
).
45.
B.
Jalan
,
R.
Engel-Herbert
,
N. J.
Wright
, and
S.
Stemmer
, “
Growth of high-quality SrTiO3 films using a hybrid molecular beam epitaxy approach
,”
J. Vac. Sci. Technol., A
27
(
3
),
461
464
(
2009
).
46.
W.
Nunn
,
T. K.
Truttmann
, and
B.
Jalan
, “
A review of molecular-beam epitaxy of wide bandgap complex oxide semiconductors
,”
J. Mater. Res.
36
(
23
),
4846
4864
(
2021
).
47.
M.
Brahlek
,
A. S.
Gupta
,
J.
Lapano
,
J.
Roth
,
H.
Zhang
,
L.
Zhang
,
R.
Haislmaier
, and
R.
Engel-Herbert
, “
Frontiers in the growth of complex oxide thin films: Past, present, and future of hybrid MBE
,”
Adv. Funct. Mater.
28
(
9
),
1702772
(
2018
).
48.
A. P.
Kajdos
,
N. G.
Combs
, and
S.
Stemmer
, “
Hybrid oxide molecular beam epitaxy
,” in
Epitaxial Growth of Complex Metal Oxides
, edited by
G.
Koster
,
M.
Huijben
, and
G.
Rijnders
(
Elsevier
,
2022
), pp.
53
74
.
49.
P.
Vogt
,
F. V. E.
Hensling
,
K.
Azizie
,
C. S.
Chang
,
D.
Turner
,
J.
Park
,
J. P.
McCandless
,
H.
Paik
,
B. J.
Bocklund
,
G.
Hoffman
,
O.
Bierwagen
,
D.
Jena
,
H. G.
Xing
,
S.
Mou
,
D. A.
Muller
,
S. L.
Shang
,
Z. K.
Liu
, and
D. G.
Schlom
, “
Adsorption-controlled growth of Ga2O3 by suboxide molecular-beam epitaxy
,”
APL Mater.
9
(
3
),
031101
(
2021
).
50.
P.
Vogt
,
D. G.
Schlom
,
F. V. E.
Hensling
,
K.
Azizie
,
Z.-K.
Liu
,
B. J.
Bocklund
, and
S.-L.
Shang
, “
Suboxide molecular-beam epitaxy and related structures
” U.S. Patent 11,462,402 B2 (October 4, 2022).
51.
K. M.
Adkison
,
S.-L.
Shang
,
B. J.
Bocklund
,
D.
Klimm
,
D. G.
Schlom
, and
Z.-K.
Liu
, “
Suitability of binary oxides for molecular-beam epitaxy source materials: A comprehensive thermodynamic analysis
,”
APL Mater.
8
(
8
),
081110
(
2020
).
52.
W.
Braun
and
J.
Mannhart
, “
Film deposition by thermal laser evaporation
,”
AIP Adv.
9
(
8
),
085310
(
2019
).
53.
S.
Smink
,
L. N.
Majer
,
H.
Boschker
,
J.
Mannhart
, and
W.
Braun
, “
Long-range atomic order on double-stepped Al2O3 (0001) surfaces
,”
Adv. Mater.
36
(
24
),
2312899
(
2024
).
54.
D. G.
Schlom
and
J. S.
Harris
, “
MBE growth of high Tc superconductors
,” in
Molecular Beam Epitaxy: Application to Key Materials
, edited by
R. F. C.
Farrow
(
Noyes
,
Park Ridge
,
1995
), pp.
505
622
.
55.
M.
Binnewies
and
E.
Milke
,
Thermochemical Data of Elements and Compounds
(
Wiley
,
1998
).
56.
T.
Onuma
,
K.
Kudo
,
M.
Ono
,
W.
Kosaka
,
K.
Shima
,
K.
Ishii
,
K.
Kaneko
,
Y.
Ota
,
T.
Yamaguchi
,
K.
Kojima
,
S.
Fujita
,
S. F.
Chichibu
, and
T.
Honda
, “
Steady-state and dynamic characteristics of deep UV luminescence in rock salt-structured MgxZn1−xO
,”
J. Appl. Phys.
134
(
2
),
025102
(
2023
).
57.
M.
Digne
,
P.
Sautet
,
P.
Raybaud
,
H.
Toulhoat
, and
E.
Artacho
, “
Structure and stability of aluminum hydroxides: A theoretical study
,”
J. Phys. Chem. B
106
(
20
),
5155
5162
(
2002
).
58.
B. G.
Orr
,
M. D.
Johnson
,
C.
Orme
,
J.
Sudijono
, and
A. W.
Hunt
, “
The surface evolution and kinetic roughening during homoepitaxy of GaAs (001)
,”
Solid-State Electron.
37
(
4–6
),
1057
1063
(
1994
).
59.
L.
Pham Van
,
O.
Kurnosikov
, and
J.
Cousty
, “
Evolution of steps on vicinal (0001) surfaces of α-alumina
,”
Surf. Sci.
411
(
3
),
263
271
(
1998
).
60.
Y.
Shiratsuchi
,
M.
Yamamoto
, and
Y.
Kamada
, “
Surface structure of self-organized sapphire (0001) substrates with various inclined angles
,”
Jpn. J. Appl. Phys., Part 1
41
(
9
),
5719
5725
(
2002
).
61.
L.
Brewer
and
A. W.
Searcy
, “
The gaseous species of the Al-Al2O3 System
,”
J. Am. Chem. Soc.
73
(
11
),
5308
5314
(
1951
).
62.
P.
Vogt
and
O.
Bierwagen
, “
The competing oxide and sub-oxide formation in metal-oxide molecular beam epitaxy
,”
Appl. Phys. Lett.
106
(
8
),
081910
(
2015
).
63.
P.
Vogt
and
O.
Bierwagen
, “
Quantitative subcompound-mediated reaction model for the molecular beam epitaxy of III-VI and IV-VI thin films: Applied to Ga2O3, In2O3, and SnO2
,”
Phys. Rev. Mater.
2
(
12
),
120401(R)
(
2018
).
64.
T. J.
Smart
,
J.
Mannhart
, and
W.
Braun
, “
Thermal laser evaporation of elements from across the periodic table
,”
J. Laser Appl.
33
(
2
),
022008
(
2021
).
65.
P.
Vogt
,
F. V. E.
Hensling
,
K.
Azizie
,
J. P.
McCandless
,
J.
Park
,
K.
DeLello
,
D. A.
Muller
,
H. G.
Xing
,
D.
Jena
, and
D. G.
Schlom
, “
Extending the kinetic and thermodynamic limits of molecular-beam epitaxy utilizing suboxide sources or metal-oxide-catalyzed epitaxy
,”
Phys. Rev. Appl.
17
(
3
),
034021
(
2022
).
66.
C. B.
Alcock
,
V. P.
Itkin
, and
M. K.
Horrigan
, “
Vapour pressure equations for the metallic elements: 298–2500K
,”
Can. Metall. Q.
23
(
3
),
309
313
(
1984
).
67.
Y.
Luo
, “
Bond dissociation energies
,” in
Atomic Charges, Bond Properties, and Molecular Energies
(
Wiley
,
2008
), pp.
151
166
.
68.
P. J.
Phillips
,
M.
De Graef
,
L.
Kovarik
,
A.
Agrawal
,
W.
Windl
, and
M. J.
Mills
, “
Atomic-resolution defect contrast in low angle annular dark-field STEM
,”
Ultramicroscopy
116
,
47
55
(
2012
).
69.
R.
Jalili
,
D.
Esrafilzadeh
,
S. H.
Aboutalebi
,
Y. M.
Sabri
,
A. E.
Kandjani
,
S. K.
Bhargava
,
E.
Della Gaspera
,
T. R.
Gengenbach
,
A.
Walker
,
Y.
Chao
,
C.
Wang
,
H.
Alimadadi
,
D. R. G.
Mitchell
,
D. L.
Officer
,
D. R.
MacFarlane
, and
G. G.
Wallace
, “
Silicon as a ubiquitous contaminant in graphene derivatives with significant impact on device performance
,”
Nat. Commun.
9
(
1
),
5070
(
2018
).
70.
J. P.
McCandless
,
C. A.
Gorsak
,
V.
Protasenko
,
D. G.
Schlom
,
M. O.
Thompson
,
H. G.
Xing
,
D.
Jena
, and
H. P.
Nair
, “
Accumulation and removal of Si impurities on β-Ga2O3 arising from ambient air exposure
,”
Appl. Phys. Lett.
124
(
11
),
2
7
(
2024
).
71.
H.
Worch
,
W.
Pompe
, and
W.
Schatt
,
Werkstoffwissenschaft
(
Wiley-VCH
,
2011
).
72.
B. D.
Evans
and
M.
Stapelbroek
, “
Optical properties of the F+ center in crystalline Al2O3
,”
Phys. Rev. B
18
(
12
),
7089
7098
(
1978
).
73.
F.
Gervais
, “
Optical conductivity of oxides
,”
Mater. Sci. Eng., R
39
(
2–3
),
29
92
(
2002
).
74.
N. P.
de Leon
,
K. M.
Itoh
,
D.
Kim
,
K. K.
Mehta
,
T. E.
Northup
,
H.
Paik
,
B. S.
Palmer
,
N.
Samarth
,
S.
Sangtawesin
, and
D. W.
Steuerman
, “
Materials challenges and opportunities for quantum computing hardware
,”
Science
372
(
6539
),
eabb2823
(
2021
).