Superlattices composed of either monoclinic μ-Fe2O3 or β-(AlxGa1−x)2O3 with β-Ga2O3 spacers are grown on (010) β-Ga2O3 substrates using plasma-assisted molecular beam epitaxy. High-resolution x-ray diffraction data are quantitatively fit using commercial dynamical x-ray diffraction software (LEPTOS) to obtain layer thicknesses, strain, and compositions. The strain state of β-(AlxGa1−x)2O3 and μ-Fe2O3 superlattices as characterized using reciprocal space maps in the symmetric (020) and asymmetric (420) diffraction conditions indicates coherent growths that are strained to the (010) β-Ga2O3 lattice. β-(AlxGa1−x)2O3 and μ-Fe2O3 superlattices grown at hotter substrate temperatures result in crystal structures with better coherency and reduced defects compared to colder growths. The growth rate of μ-Fe2O3 is ∼2.6 nm/min at Tsub = 700 °C and drops to ∼1.6 nm/min at Tsub = 800 °C due to increased Fe interdiffusion at hotter substrate temperatures. Scanning transmission electron microscopy data of a μ-Fe2O3 superlattice grown at Tsub = 700 °C confirm that there is significant diffusion of Fe atoms into β-Ga2O3 layers.

β-Ga2O3 is an ultrawide bandgap (∼4.8 eV) oxide semiconductor that is being intensely studied for applications in high-power devices due to its very high breakdown voltage (Eb = 8 MV/cm).1,2 β-Ga2O3 has a monoclinic structure with a space group of C2/m and lattice parameters a = 12.23, b = 3.04, and c = 5.80 Å.1,3 High-quality β-Ga2O3 epitaxial films with smooth surfaces and uniform doping profiles can be grown using molecular beam epitaxy (MBE).1,4–8 β-Ga2O3 has attracted interest for applications in power electronics and optoelectronics.9–12 A new monoclinic μ-phase of Fe2O3, isomorphic to β-Ga2O3, was stabilized using MBE.13,14 The epitaxial stabilization of the magnetic material μ-Fe2O3 paves the way toward realizing practical magneto-electronic devices on β-Ga2O3. DFT calculations show that μ-Fe2O3 has a bandgap in the visible range of ∼2.2 eV. The smaller bandgap of μ-Fe2O3 and the excellent lattice matching to β-Ga2O3 makes μ-Fe2O3 a very useful low bandgap insert in β-Ga2O3 heterostructures.

A material that has a bandgap larger than β-Ga2O3 is attractive to form quantum wells with carrier confinement and for optoelectronic applications in the deep ultraviolet.15,16 Al2O3 is an excellent candidate since the corundum (α) phase of Al2O3 has a bandgap of 8.82 eV.15,17 The alloy β-(AlxGa1−x)2O3 with a monoclinic structure has a bandgap between ∼4.8 and ∼7.2 eV depending on the Al composition.17 With a composition of ∼80% Al, β-(AlxGa1−x)2O3 is predicted to have a breakdown voltage of Eb ∼ 16 MV/cm,18 even higher than β-Ga2O3. β-(AlxGa1−x)2O3 heterostructures have been successfully grown using MBE on (010) β-Ga2O3 substrates with Al compositions of up to x = 0.18.19,20 Heterostructures of μ-Fe2O3 and β-(AlxGa1−x)2O3 with β-Ga2O3 are of interest for applications in magneto-electronics, high-power switching applications, and UV optoelectronics.

The highly anisotropic crystal structure of β-Ga2O3 poses a challenge to epitaxial growth and characterization since thin film mechanics models typically rely on cubic or hexagonal symmetry.1,3,21–23 One of the most important quantitative tools for developing epitaxial heterostructures is high-resolution XRD, which when quantitatively fit provides high precision measurements of film thicknesses, strain state, and alloy composition. Although HRXRD has been widely reported in the literature from β-Ga2O3 heterostructures, quantitative fits of the XRD data have been lacking. Here, we provide an example of how to adapt a commercial HRXRD simulation software (LEPTOS) to enable quantitative fits of complex heterostructures derived from monoclinic β-Ga2O3. The methodology can be adapted for any crystal orientation using a simple unit cell transformation to overcome constraints in the modeling software. The results are validated by comparison to STEM and quantitative EELS chemical maps providing understanding of the issues of interdiffusion in the β-Ga2O3 heterostructure family of alloys including Fe and Al Ga2O3 alloys for bandgap engineering in UWBG oxides.

The μ-Fe2O3 and β-(AlxGa1−x)2O3 heterostructures used in this study are grown via plasma-assisted molecular beam epitaxy (PAMBE) in a Riber/MBE Control M7 system equipped with Ga and Fe effusion cells and a Veeco oxygen plasma source with a forward power of 300 W (with ≤1 W reflected). The flow of O2 is continuously monitored and maintained at a growth chamber pressure of 1.5 × 10−5 Torr. The substrates used for these growths are 5 × 5 × 0.5 mm3 unintentionally doped (UID) or Fe-doped (010) β-Ga2O3 single crystals (Tamura Corporation).

For the growths of μ-Fe2O3 superlattices, an ∼100-nm thick buffer layer of β-Ga2O3 is first grown followed by ten periods of β-Ga2O3 spacers and μ-Fe2O3 epilayers. The flux of Fe is held constant between samples with a beam equivalent pressure (BEP) of 3.2 × 10−8 Torr and the Fe deposition time is varied between 30 and 60 s. The substrate temperature Tsub as measured through the thermocouple is varied between 500 and 800 °C. β-Ga2O3 spacers are grown at a Ga flux between 4.0 and 1.0 × 10−7 Torr with a constant deposition time of 400 s. Some of the μ-Fe2O3 superlattices investigated in this study are described in Hettiaratchy et al.14 To investigate how the composition of β-(AlxGa1−x)2O3 changes with the Al flux, three samples were grown with a 100 nm β-Ga2O3 buffer layer and an ∼20–30 nm layer of β-(AlxGa1−x)2O3 at Al fluxes between 1.2 and 1.7 × 10−8 Torr. The Ga flux is kept fixed at 1.6 × 10−7 Torr and Tsub is kept fixed at 725 °C. Superlattices composed of ten periods of β-(AlxGa1−x)2O3 layers and β-Ga2O3 spacers were grown on a ∼40 nm buffer layer of β-Ga2O3 at Tsub = 800 °C and Tsub = 725 °C. For the superlattice grown at Tsub = 800 °C, the BEP of Al and Ga is 1.0 × 10−8 and 8.0 × 10−8 Torr, respectively. For the superlattice grown at Tsub = 725 °C, the BEP of Al and Ga is 7.0 × 10−9 and 1.4 × 10−7 Torr, respectively.

To investigate the structure and composition of superlattices based on β-Ga2O3, high-resolution x-ray diffraction (HRXRD) data are fitted with simulations based on dynamical x-ray diffraction. LEPTOS21 is a commercially available software package that is used for the analysis of structural and crystallographic parameters of thin film epitaxial heterostructures. LEPTOS uses the recursion matrix formalism for two-wave approximation of dynamical x-ray diffraction theory.22 This algorithm calculates x-ray scattering from thin film layers over the entire range of incidence and exit angles for various experimental scans. Other XRD fitting software packages, including micronova24 and CADEM,25 were also tested, and while they can measure the layer thicknesses, these software packages were currently unable to quantitatively derive the strain state of the μ-Fe2O3 superlattices; however, as they are open-source, they could be modified to include these capabilities.

In its default configuration, LEPTOS cannot simulate the XRD patterns of monoclinic structures along the (010) facet. LEPTOS is optimized for the XRD fitting of cubic and hexagonal crystal structures. As a workaround, the monoclinic unit cell of β-Ga2O3 is transformed into a triclinic unit cell while swapping the b and c axes. Following this transformation, the former (010) plane of the standard monoclinic cell is equivalent to the (001) plane of the b-c swapped triclinic cell. The triclinic space group [1] P1 has no symmetric atomic positions and all the 20 atomic sites for the transformed unit cell are defined in LEPTOS. This unit cell transformation allows for the fitting of monoclinic crystal structures with LEPTOS. For β-Ga2O3, the lattice constants used are a = 12.23, b = 3.04, and c = 5.80 Å.3 Based on DFT calculations, the lattice constants used for strained μ-Fe2O3 are a = 12.23, b = 3.16, and c = 5.80 Å,13and for relaxed μ-Fe2O3, a = 12.42, b = 3.07, c = 5.88 Å.13 To fit the XRD data for β-(AlxGa1−x)2O3 heterostructures, we defined a β-(AlxGa1−x)2O3 alloy that is fully strained to β-Ga2O3 using the triclinic, b–c swapped representations of the unit cells in LEPTOS. From Oshima et al.,26 we assume that the in-plane lattice constants of β-(AlxGa1−x)2O3 are the same as β-Ga2O3 (a = 12.23 and c = 5.80 Å) and fit for the out-of-plane lattice parameter as a function of Al composition using Vegard's law, with b = 2.849 Å for fully strained Al2O3 [from Eq. (8a) in Oshima et al.26].

Even when using bc swapped triclinic unit cells, LEPTOS is unable to model the interface roughness or the strain-relaxation properties of these heterostructures. For the μ-Fe2O3 superlattices, the strain-relaxation was modeled by defining an alloy between the strained and relaxed μ-Fe2O3 crystal structures as obtained through DFT calculations.13 Vegard's law was used to determine the c-lattice constant for the alloy in the triclinic unit cell,

(1)

where R is the composition of relaxed μ-Fe2O3 in the alloy, crelaxed=3.07Å, and cstrained=3.16Å. If the lattice constants for both the strained and relaxed structures are known, this framework can be used to quantitatively determine the strain state of a thin film.

Figure 1 shows the simulated ω–2θ XRD patterns for the (020) diffraction condition from a ten-period superlattice structure with varying degrees of relaxation. The superlattice peaks (SL0, SLn + 1) shift as R is varied, indicating that this method is a good approximation to model the strain in these superlattices. As R approaches 100% (fully relaxed), the zeroth order superlattice peak SL0 is blended with the (010) β-Ga2O3 substrate peak. LEPTOS simulations of this superlattice structure while varying the roughness of the layers did not show significant changes in the resulting XRD patterns.

FIG. 1.

Simulated XRD ω–2θ patterns of a μ-Fe2O3/β-Ga2O3 superlattice with μ-Fe2O3 layers having a relaxation R = 0%, 25%, 50%, 75%, and 100%.

FIG. 1.

Simulated XRD ω–2θ patterns of a μ-Fe2O3/β-Ga2O3 superlattice with μ-Fe2O3 layers having a relaxation R = 0%, 25%, 50%, 75%, and 100%.

Close modal

XRD is performed on the samples using either a Bede diffractometer in triple axis mode or a Bruker D8 system. Reciprocal space maps (RSMs) in the symmetrical (020) (ω=30.3449°, 2θ=60.6898°) and asymmetric glancing exit (420) diffraction conditions (ω=61.7846°, 2θ=69.3916°) were obtained to investigate the strain-relaxation properties of the superlattices. The ω–2θ XRD patterns of the μ-Fe2O3 superlattices are fit using LEPTOS with a Genetic algorithm keeping the layer thicknesses and the composition of the strained/relaxed μ-Fe2O3 alloy as free parameters. The structural model used by LEPTOS consists of a β-Ga2O3 substrate and ten identical layers of μ-Fe2O3 and β-Ga2O3 that are constrained to have the same layer thicknesses and lattice constants across the superlattice structure. Instead of fitting the relaxation parameter available in LEPTOS, the composition R of the strained/relaxed μ-Fe2O3 alloy is fitted as it can be used as an approximation of the degree of relaxation in the superlattice [Eq. (1)]. The interface roughness is kept fixed at 0 as it showed no measurable impact on the simulated patterns.

The thickness of the μ-Fe2O3 layers range from ∼0.8 to ∼1.6 nm (∼5.0–10.2 ml) and the thickness of the β-Ga2O3 spacers range from ∼8.1 to ∼33.5 nm (∼52.2–220.4 ml). The composition R of the alloyed μ-Fe2O3 layers is ∼14.4%, which corresponds to a μ-Fe2O3 lattice parameter of b = 3.147 Å. However, the RSMs of μ-Fe2O3 superlattices (Fig. 2) show that the μ-Fe2O3 layers are fully strained to the β-Ga2O3 lattice. If the μ-Fe2O3 layers are fully strained, these measurements instead suggest that the experimentally measured b-lattice parameter of μ-Fe2O3 is b = 3.15 Å, which is in good agreement with the DFT prediction of b = 3.16 Å. The slight deviation from the DFT-predicted b-lattice parameter could also arise from off-stoichiometry or Ga/Fe intermixing in the Fe2O3 layers.

FIG. 2.

(a) XRD ω–2θ pattern (black) of a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 8.1 × 10−8 Torr at Tsub = 800 °C and the fitted XRD pattern (red). (b) RSM from the symmetric [020] and (c) asymmetric [420] diffraction conditions. (d)–(f) show the same for a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 4.1 × 10−8 Torr at Tsub = 500 °C. The fully strained and fully relaxed conditions are shown as black dashed lines.

FIG. 2.

(a) XRD ω–2θ pattern (black) of a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 8.1 × 10−8 Torr at Tsub = 800 °C and the fitted XRD pattern (red). (b) RSM from the symmetric [020] and (c) asymmetric [420] diffraction conditions. (d)–(f) show the same for a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 4.1 × 10−8 Torr at Tsub = 500 °C. The fully strained and fully relaxed conditions are shown as black dashed lines.

Close modal
FIG. 3.

1 × 1 μm2 AFM images for the μ-Fe2O3/β-Ga2O3 superlattices described in Fig. 1 grown at temperatures of (a) 800 and (b) 500 °C at the same vertical scale. All scale bars are 250 nm. The AFM RMS surface roughness measurements are also shown.

FIG. 3.

1 × 1 μm2 AFM images for the μ-Fe2O3/β-Ga2O3 superlattices described in Fig. 1 grown at temperatures of (a) 800 and (b) 500 °C at the same vertical scale. All scale bars are 250 nm. The AFM RMS surface roughness measurements are also shown.

Close modal

Figures 2(a) and 2(d) show the ω–2θ XRD patterns for two different μ-Fe2O3 superlattices grown at (ΦGa = 8.1 × 10−8 Torr, Tsub = 800 °C) and (ΦGa = 4.1 × 10−8 Torr, Tsub = 500 °C), respectively. The best fits from LEPTOS are shown in red. The fitting procedure provides excellent fits to the observed XRD data. While it is not possible to determine the interface roughness directly through the XRD fitting, atomic force microscopy (AFM) root-mean-square (RMS) measurements can be used as a proxy for the interface roughness. Figure 3 shows 1 × 1 μm2 AFM images for these two superlattices. For the superlattice grown at Tsub = 800 °C, the RMS roughness is ∼0.3 nm (∼2 ml) while the superlattice grown at Tsub = 500 °C is ∼3.6× rougher with an RMS roughness of ∼1.1 nm (∼7 ml). This is consistent with the results from Hettiaratchy et al. who found that the RMS roughness of μ-Fe2O3 superlattices increases with decreasing Tsub.14 

FIG. 4.

Average μ-Fe2O3 and β-Ga2O3 growth rates normalized by the BEP flux as a function of Tsub.

FIG. 4.

Average μ-Fe2O3 and β-Ga2O3 growth rates normalized by the BEP flux as a function of Tsub.

Close modal

Figures 2(b), 2(e), 2(c), and 2(f) show the (020) and (420) RSMs for these two superlattices, respectively. Superlattice satellite peaks up to the sixth order are visible in the RSM data for the superlattice grown at Tsub = 800 °C. In the symmetrical (020) RSMs, the reciprocal lattice points (RLPs) for the superlattice peaks are vertically aligned, indicating a negligible lattice tilt within the heterostructures. The asymmetric (420) RSMs show that the RLPs for the superlattice peaks are vertically aligned to the (010) β-Ga2O3 substrate. This indicates that μ-Fe2O3 and β-Ga2O3 spacers in the superlattice are fully strained to the (010) β-Ga2O3 substrate and the superlattice growths are coherent and of good quality. The RSMs for the two superlattices differ in the amount of spread in the RLPs. The RLPs for the superlattice grown at Tsub = 800 °C are narrow, indicating few defects in the superlattice. In contrast, the RLPs for the superlattice grown at Tsub = 500 °C are significantly more spread out and indicate a greater number of defects in the superlattice structure. This is consistent with the differences in the surface roughness measurements from the AFM data. μ-Fe2O3 superlattices grown at Tsub > 700 °C are more coherent and of better quality than those grown at colder substrate temperatures.

Using the layer thicknesses determined from the LEPTOS fits, the μ-Fe2O3 growth rate can be determined as a function of Tsub. The average growth rate of μ-Fe2O3 as measured through XRD is ∼2.7 nm/min at Tsub = 500 °C and ∼2.5 nm/min at Tsub = 700 °C. At Tsub = 800 °C, the growth rate drops to ∼1.6 nm/min. Figure 4 shows the growth rate of μ-Fe2O3 and β-Ga2O3 normalized by the BEP. The reported BEPs are calibrated for N2, so for a better comparison, the BEP of the Fe and Ga cells must be normalized by the ratio of masses m(Fe/Ga)m(N2).

The normalized growth rate of μ-Fe2O3 is generally larger than that of β-Ga2O3, which may indicate that Fe atoms have a higher sticking coefficient than Ga atoms. The growth rate of μ-Fe2O3 seemingly drops at Tsub = 800 °C. The thermal decomposition of Fe2O3 happens at much higher temperatures of about 1470–1570 °C,27 so it is unlikely to cause reduced growth rates. However, Fe can diffuse substantially at interfaces and the rate of Fe diffusion is known to increase with temperature in various material systems.28–30 Increased interdiffusion of Fe into β-Ga2O3 spacers at higher Tsub can explain the reduced growth rates. This will reduce the apparent thicknesses of μ-Fe2O3 layers as probed by XRD and lead to a reduction in calculated XRD growth rates. The calibration of μ-Fe2O3 growth rates using superlattices with very thin μ-Fe2O3 layers (∼few ML) is challenging. It may be possible to improve the μ-Fe2O3 growth rates by growing heterostructures with thicker μ-Fe2O3 layers.

This method of fitting the XRD patterns of monoclinic heterostructures with LEPTOS can also be applied to the monoclinic β-(AlxGa1−x)2O3 heterostructures. The ω–2θ XRD patterns of β-(AlxGa1−x)2O3 samples with varying compositions were fit with LEPTOS keeping the thickness of the β-(AlxGa1−x)2O3 layer and its composition x as free parameters. The β-(AlxGa1−x)2O3 layers are assumed to be fully strained when fitting the XRD data. β-(AlxGa1−x)2O3 layers with thickness up to ∼60 nm were found to be fully strained to the β-Ga2O3 lattice,19 and since the thickness of the β-(AlxGa1−x)2O3 layers investigated in this study is substantially smaller than ∼60 nm, that the β-(AlxGa1−x)2O3 layers are fully strained is a valid assumption.

Figures 5(a)5(c) show the XRD data and the LEPTOS fits for β-(AlxGa1−x)2O3 sample grown with Al fluxes of 1.7 × 10−8, 1.5 × 10−8, and 1.3 × 10−8 Torr, respectively. The fits to the XRD data are excellent. The β-(AlxGa1−x)2O3 layer thicknesses are ∼24.7, ∼26.3, and ∼27.2 nm, and the Al compositions are x=19.36%, x=16.38%, and x=13.74% at 1.7 × 10−8, 1.5 × 10−8, and 1.3 × 10−8 Torr, respectively. The Al composition x can also be estimated from a numerical fit to the separation Δθ020 of the (020) β-(AlxGa1−x)2O3 peak from the β-Ga2O3 substrate peak θ020.26 This relation is given as

(2)
FIG. 5.

XRD ω–2θ pattern (black) and fit (red) for β-(AlxGa1−x)2O3 heterostructures grown at (a) 1.7 × 10−8 Torr, (b) 1.5 × 10−8 Torr, and (c) 1.3 × 10−8 Torr.

FIG. 5.

XRD ω–2θ pattern (black) and fit (red) for β-(AlxGa1−x)2O3 heterostructures grown at (a) 1.7 × 10−8 Torr, (b) 1.5 × 10−8 Torr, and (c) 1.3 × 10−8 Torr.

Close modal

Using Eq. (2), the Al compositions are x = 18.73%, x = 15.91%, and x = 12.78% at 1.7 × 10−8, 1.5 × 10−8, and 1.3 × 10−8 Torr, respectively. The Al compositions from the LEPTOS agree very well with those obtained from Eq. (2). The LEPTOS compositions are only ∼0.5%–1.0% smaller than those obtained from the analytical method. Oshima et al. validated their analytical estimate of the composition with atomic probe tomography (APT) and found that the APT results were consistent with the estimates given 2%–3% uncertainties of the APT measurements.26 Additionally, the Al and Ga beam flux ratios were found to impact the derived Al compositions. For beam flux ratios ϕAl/(ϕGa+ϕAl) of 0.061 and 0.104, the Al composition measured from APT were 12.0% and 16.8%, respectively.26 The beam flux ratio for the growth with an Al flux of 1.7 × 10−8 Torr is ϕAl/(ϕGa+ϕAl)=0.096, and a simple interpolation of the APT results gives an Al composition of ∼15.9% for this growth. The Al composition of ∼19.4% obtained from LEPTOS is ∼3.5% larger than the interpolated APT results of Oshima et al.26 Considering the uncertainties in the analytical estimate, the Al compositions derived from the LEPTOS agree well with the estimates from Oshima et al.26 

The XRD patterns of β-(AlxGa1−x)2O3 superlattices were also fit using this method. The thickness of the β-(AlxGa1−x)2O3 layers and their composition were fit along with the thickness of the β-Ga2O3 spacers. Figures 6(a) and 6(d) show the XRD data and fits for two β-(AlxGa1−x)2O3 superlattices grown at Tsub = 800 and Tsub = 725 °C, respectively. For the superlattice grown at Tsub = 725 °C, the β-Ga2O3 layers are ∼31.5 nm thick and the β-(AlxGa1−x)2O3 layers are ∼4.6 nm thick with a composition of x4.55%. For the superlattice grown at Tsub = 800 °C, the β-Ga2O3 layers are ∼13.5 nm thick and the β-(AlxGa1−x)2O3 layers are ∼2.0 nm thick with a composition of x23.0%. The β-Ga2O3 growth rates from the XRD fits are as expected for the Ga beam flux and Tsub. The growth rate of β-(AlxGa1−x)2O3 decreases with increasing Al flux. Figures 6(b), 6(c), 6(e), and 6(f) show the (020) and (420) RSMs for the two β-(AlxGa1−x)2O3 superlattices grown at Tsub = 800 and Tsub = 725 °C. Superlattice peaks up to the fourth order are visible in the RSMs. Both the β-(AlxGa1−x)2O3 superlattices have no tilt and are fully strained and coherent. For the superlattice grown at Tsub = 725 °C, the SL0 RLP is blended with that of the β-Ga2O3 substrate. The RLPs for the superlattice grown at Tsub = 725 °C are more spread out than those for the superlattice grown at Tsub = 800 °C, indicating that the superlattice quality improves at hotter growths.

FIG. 6.

(a) XRD ω–2θ pattern (black) of a β-(AlxGa1−x)2O3/β-Ga2O3 superlattice grown with ΦGa = 8 × 10−8 Torr at Tsub = 800 °C and the fitted XRD pattern (red). (b) RSM from the symmetric [020] and (c) asymmetric [420] diffraction conditions. (d)–(f) show the same for a β-(AlxGa1−x)2O3/β-Ga2O3 superlattice grown with ΦGa = 1.4 × 10−7 Torr at Tsub = 725 °C. The fully strained and fully relaxed conditions are shown as black dashed lines.

FIG. 6.

(a) XRD ω–2θ pattern (black) of a β-(AlxGa1−x)2O3/β-Ga2O3 superlattice grown with ΦGa = 8 × 10−8 Torr at Tsub = 800 °C and the fitted XRD pattern (red). (b) RSM from the symmetric [020] and (c) asymmetric [420] diffraction conditions. (d)–(f) show the same for a β-(AlxGa1−x)2O3/β-Ga2O3 superlattice grown with ΦGa = 1.4 × 10−7 Torr at Tsub = 725 °C. The fully strained and fully relaxed conditions are shown as black dashed lines.

Close modal

The interdiffusion of Al atoms in β-(AlxGa1−x)2O3 heterostructures has previously been observed.31,32 STEM data of an Al2O3/β-Ga2O3 interface showed that the interface was abrupt with a transitional region of ∼2 nm, and energy dispersive x-ray spectroscopy (EDS) found that the Al atoms are diffused within β-Ga2O3.32 Studies of Fe interdiffusion in Fe/GaAs, Fe/Si, and Fe/Al systems have shown considerable intermixing.28–30 Secondary ion mass spectrometry (SIMS) results showed that Fe originating from a Fe-doped β-Ga2O3 substrate diffused up to ∼200 nm into a β-Ga2O3 buffer layer grown at Tsub = 700 °C.33 

To investigate the interdiffusion of Fe atoms in the μ-Fe2O3 superlattices, we obtain scanning transmission electron microscopy (STEM) high-angle annular dark field (HAADF) and electron energy loss spectroscopy (EELS) data of a superlattice grown at Tsub = 700 °C and ΦGa = 6 × 10−8 Torr with ∼1.1 nm μ-Fe2O3 layers and ∼15.9 nm β-Ga2O3 spacers. The EELS data were acquired using a K2 detector on a Titan3™ G2 60–300 S/TEM at 300 kV with a pixel time of 0.25 s. At the interface between a μ-Fe2O3 layer and a β-Ga2O3 spacer shown in Fig. 7(a), Fe atoms diffuse substantially in the (010) growth direction, resulting in a dimmer contrast. Figures 7(b) and 7(c) show EELS data of the Fe-L edge signals near this interface. The concentration of Fe atoms near the interface is asymmetric indicating the asymmetric diffusion of Fe atoms into the β-Ga2O3 spacer layers. The Fe-L edge signal spans ∼8 nm, whereas the XRD thickness of the μ-Fe2O3 layer is ∼1.1 nm, suggesting the presence of Fe atoms in ∼6.9 nm of β-Ga2O3. The ordered interface phase of (FeoctGatet)2O3 with Fe occupying octahedral sites and Ga occupying tetrahedral sites is predicted to be ferromagnetic from DFT calculations.13 Hettiaratchy et al. found that increasing the interface roughness resulted in an increased ferromagnetic response from the superlattices, which supports the DFT prediction.14 Additionally, the diffusion of Fe atoms in the β-Ga2O3 spacer layers produces a paramagnetic component in the observed magnetic response of the superlattices.13,14

FIG. 7.

(a) ADF-STEM image of an interface of a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 6.0 × 10−8 Torr at Tsub = 700 °C. (b) EELS map of Fe-L edge signals marked with the white box shown in (a). (c) Line profile of a Fe-L edge signal integrated along the [010] growth direction indicating the asymmetry of Fe concentration on both sides of the μ-Fe2O3 layer.

FIG. 7.

(a) ADF-STEM image of an interface of a μ-Fe2O3/β-Ga2O3 superlattice grown with ΦGa = 6.0 × 10−8 Torr at Tsub = 700 °C. (b) EELS map of Fe-L edge signals marked with the white box shown in (a). (c) Line profile of a Fe-L edge signal integrated along the [010] growth direction indicating the asymmetry of Fe concentration on both sides of the μ-Fe2O3 layer.

Close modal

Superlattices of μ-Fe2O3 or β-(AlxGa1−x)2O3 and β-Ga2O3 layers are grown on (010) β-Ga2O3 substrates using PAMBE. A method to adapt commercial dynamical x-ray diffraction software to quantitatively fit the HRXRD patterns of monoclinic heterostructures is presented. The layer thicknesses, strain-relaxation properties, and alloy compositions of the β-(AlxGa1−x)2O3 and μ-Fe2O3 heterostructures were obtained by fitting XRD data. The RSMs in the symmetric (020) and asymmetric (420) diffraction conditions for μ-Fe2O3 and β-(AlxGa1−x)2O3 superlattices indicate coherent growths that are fully strained to the (010) β-Ga2O3 lattice. Superlattices grown at hotter substrate temperatures result in fewer crystal defects and better coherency. The growth rate of μ-Fe2O3 is ∼2.6 nm/min at Tsub = 700 °C and drops to ∼1.6 nm/min at Tsub = 800 °C due to increased Fe interdiffusion at hotter substrate temperatures. STEM and EELS data of an interface in a μ-Fe2O3 superlattice grown at Tsub = 700 °C confirm that Fe atoms diffuse significantly into β-Ga2O3 spacers.

This work was supported by the Center for Emergent Materials at The Ohio State University, an NSF MRSEC (No. DMR-1420451) and by the Army Research Office MURI (No. W911NF-14-1-0016). B.W. acknowledges support from the Presidential Fellowship of the Ohio State University. Electron microscopy experiments were supported by the Center for Electron Microscopy and Analysis at the Ohio State University

The authors have no conflicts to disclose.

Elline C. Hettiaratchy: Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (lead). Binbin Wang: Investigation (supporting). Ashok Dheenan: Investigation (supporting). Joe McGlone: Investigation (supporting). Nidhin Kurian Kalarickal: Investigation (supporting). Núria Bagués: Investigation (supporting). Steven Ringel: Project administration (supporting); Supervision (supporting). David W. McComb: Supervision (supporting). Siddharth Rajan: Investigation (supporting); Project administration (supporting); Supervision (supporting). Roberto C. Myers: Conceptualization (lead); Funding acquisition (lead); Investigation (supporting); Project administration (lead); Supervision (lead); Writing – original draft (supporting); Writing – review & editing (equal).

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

1.
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
).
2.
S. J.
Pearton
,
F.
Ren
,
M.
Tadjer
, and
J.
Kim
,
Appl. Phys.
124
,
220901
(
2018
).
3.
S.
Geller
,
J. Chem. Phys.
33
,
676
(
1960
).
4.
N.
Nepal
,
D.
Scott Katzer
, and
D. J.
Meyer
,
Gallium Oxide Technology, Devices and Applications
(
Elsevier
,
2018
), pp.
31
46
.
5.
T.
Oshima
,
T.
Okuno
,
N.
Arai
,
N.
Suzuki
,
S.
Ohira
, and
S.
Fujita
,
Appl. Phys. Express
1
,
011202
(
2008
).
6.
K.
Sasaki
,
A.
Kuramata
,
T.
Masui
,
E. G.
Víllora
,
K.
Shimamura
, and
S.
Yamakoshi
,
Appl. Phys. Express
5
,
035502
(
2012
).
7.
K.
Sasaki
,
M.
Higashiwaki
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
,
J. Cryst. Growth
392
,
30
(
2014
).
8.
P.
Vogt
and
O.
Bierwagen
,
Appl. Phys. Lett.
108
,
072101
(
2016
).
9.
G.
Jessen
 et al, “Toward realization of Ga2O3 for power electronics applications,”
2017 75th Annual Device Research Conference (DRC)
, (IEEE, 2017), pp. 1–2.
10.
D.
Guo
,
Q.
Guo
,
Z.
Chen
,
Z.
Wu
,
P.
Li
, and
W.
Tang
,
Mater. Today Phys.
11
,
100157
(
2019
).
11.
M.
Passlack
 et al,
J. Appl. Phys.
77
,
686
(
1995
).
12.
Z.
Galazka
,
Semicond. Sci. Technol.
33
,
113001
(
2018
).
13.
J. S.
Jamison
,
B. J.
May
,
J. I.
Deitz
,
S.-C.
Chien
,
D. W.
McComb
,
T. J.
Grassman
,
W.
Windl
, and
R. C.
Myers
,
Cryst. Growth Des.
19
,
4205
(
2019
).
14.
E. C.
Hettiaratchy
,
J. S.
Jamison
,
B.
Wang
,
N.
Bagués
,
R. A.
Guest
,
D. W.
McComb
, and
R. C.
Myers
,
J. Vac. Sci. Technol. A
38
,
063413
(
2020
).
15.
A. F. M. A. U.
Bhuiyan
,
Z.
Feng
,
L.
Meng
, and
H.
Zhao
,
J. Mater. Res.
36
,
4804
(
2021
).
16.
F.
Li
,
G.
Han
,
J.
Zhang
,
L.
Huang
,
Q.
Feng
,
W.
Tang
,
X.
Li
, and
Y.
Hao
,
Opt. Mater. Express
7
,
1240
(
2017
).
17.
H.
Peelaers
,
J. B.
Varley
,
J. S.
Speck
, and
C. G.
van de Walle
,
Appl. Phys. Lett.
112
,
242101
(
2018
).
18.
J. B.
Varley
,
A.
Perron
,
V.
Lordi
,
D.
Wickramaratne
, and
J. L.
Lyons
,
Appl. Phys. Lett.
116
,
172104
(
2020
).
19.
S. W.
Kaun
,
F.
Wu
, and
J. S.
Speck
,
J. Vac. Sci. Technol. A
33
,
041508
(
2015
).
20.
N. K.
Kalarickal
,
Z.
Xia
,
J. F.
McGlone
,
Y.
Liu
,
W.
Moore
,
A. R.
Arehart
,
S. A.
Ringel
, and
S.
Rajan
,
J. Appl. Phys.
127
,
215706
(
2020
).
21.
A.
Ulyanenkov
,
Advances in Computational Methods for X-Ray and Neutron Optics
(
SPIE
,
2004
), pp.
1
15
, 5536.
22.
S. A.
Stepanov
,
E. A.
Kondrashkina
,
R.
Köhler
,
D. V.
Novikov
,
G.
Materlik
, and
S. M.
Durbin
,
Phys. Rev. B
57
,
4829
(
1998
).
23.
M.
Slomski
,
N.
Blumenschein
,
P. P.
Paskov
,
J. F.
Muth
, and
T.
Paskova
,
J. Appl. Phys.
121
,
235104
(
2017
).
24.
J.
Tiilikainen
,
V.
Bosund
,
J.-M.
Tilli
,
J.
Sormunen
,
M.
Mattila
,
T.
Hakkarainen
, and
H.
Lipsanen
,
J. Phys. D: Appl. Phys.
40
,
6000
(
2007
).
25.
P.
Komar
and
G.
Jakob
,
J. Appl. Crystallogr.
50
,
288
(
2017
).
26.
Y.
Oshima
,
E.
Ahmadi
,
S. C.
Badescu
,
F.
Wu
, and
J. S.
Speck
,
Appl. Phys. Express
9
,
061102
(
2016
).
27.
Y.
Qu
,
Y.
Yang
,
Z.
Zou
,
C.
Zeilstra
,
K.
Meijer
, and
R.
Boom
,
ISIJ Int.
54
,
2196
(
2014
).
28.
J.
Zhang
 et al,
Phys. Proc.
11
,
126
(
2011
).
29.
M. W.
Ruckman
,
J. J.
Joyce
, and
J. H.
Weaver
,
Phys. Rev. B
33
,
7029
(
1986
).
30.
L.
Yajiang
,
W.
Juan
,
Y.
Yansheng
, and
M.
Haijun
,
Bull. Mater. Sci.
28
,
69
(
2005
).
31.
C.-H.
Liao
,
K.-H.
Li
,
C. G.
Torres Castanedo
,
G.
Zhang
, and
X.
Li
, arXiv: 2005.05799 (2020).
32.
S.
Roy
,
A. E.
Chmielewski
,
A.
Bhattacharyya
,
P.
Ranga
,
R.
Sun
,
M. A.
Scarpulla
,
N.
Alem
, and
S.
Krishnamoorthy
,
Adv. Electron. Mater.
7
,
2100333
(
2021
).
33.
C.
Joishi
,
Z.
Xia
,
J.
McGlone
,
Y.
Zhang
,
A. R.
Arehart
,
S.
Ringel
,
S.
Lodha
, and
S.
Rajan
,
Appl. Phys. Lett.
113
,
123501
(
2018
).