Ultrawide bandgap ZnGa2O4 (ZGO) thin films were grown on sapphire (0001) substrates at various growth temperatures with a perspective to investigate the electrical and optical characteristics required for high-power electronic applications. Due to the variation in the vapor pressure of Zn and Ga, severe loss of Zn was observed during pulsed laser deposition, which was solved by using a zinc-rich Zn0.98Ga0.02O target. A pure phase single-crystalline ZGO thin film was obtained at a deposition temperature of 750 °C and an oxygen pressure of 1 × 10−2 Torr. The out-of-plane epitaxial relationship between the sapphire and ZGO thin film was obtained from φ-scan. The x-ray rocking curve of the ZGO thin film grown at 750 °C exhibits a full width at half maximum of ∼0.098°, which indicates a good crystalline phase and quality of the thin film. Core-level x-ray photoelectron spectroscopy of ZGO grown at 750 °C indicated that Zn and Ga were in the 2+ and 3+ oxidation states, respectively, and the atomic ratio of Zn/Ga was estimated to be ∼0.48 from the fitted values of Zn-2p3/2 and Ga-2p3/2. The high-resolution transmission electron microscopy images revealed a sharp interface with the thickness of the ZGO film of ∼265 nm, and the signature of minor secondary phases was observed. The bandgap of the ZGO film at different growth temperatures was calculated from the ultraviolet-diffuse reflectance spectroscopy spectra, and its value was obtained to be ∼5.08 eV for the 750 °C grown sample. The refractive index (n) and the extinction coefficient (k) were determined to be ∼1.94 and 0.023 from the ellipsometric data, respectively, and the real dielectric function (ɛr) was estimated to be ∼6.8 at energy 5 eV. The ultrawide bandgap and dielectric function of ZGO recommend its possible potential applications in deep-ultraviolet optoelectronic devices and high-power electronics.

Three generations of semiconductor materials have emerged since the invention of the transistor in 1947 and are currently empowering our modern society: first-generation semiconductors such as Si and Ge, second-generation semiconductors such as GaAs and InP, and most recently, third-generation wide bandgap (WBG) materials such as GaN and SiC.1,2 Third-generation semiconductors, such as SiC (Eg ∼ 3.2 eV) and GaN (Eg ∼ 3.4 eV), are rapidly gaining demand due to potential applications toward the advancement of small area, high power, high-frequency switching, with low energy consumption compared to conventional Si (Eg ∼ 1.1 eV).3,4 While major manufacturers are showing tremendous interest in GaN and SiC power devices, these technologies still have limitations due to their high cost, low thermal conductivity, and high melting points (for GaN ∼ 2220 °C and for SiC ∼ 2730 °C). SiC is a very hard material that is challenging to cut, grind, and polish. It grows slowly, melts at a very high energy cost, and requires very high working pressures for bulk manufacturing.3,5 Therefore, a renaissance in ultrawide bandgap (UWBG) semiconductor technology is currently under way, as evidenced by new device designs, new applications, and breakthroughs in understanding the fundamental properties of these materials. As a result, UWBG semiconductors are an emerging new area of research due to their high breakdown electric field, control of electrical conductivity, chemical composition, insensitivity to high temperature, and wide range of materials, physics, devices, and applications, such as power devices, 5G communication, fast chargers, deep-ultraviolet (DUV) optoelectronics, and Industry 4.0 technologies.6,7 Currently, Ga2O3 (gallium oxide) is gaining popularity among ultrawide bandgap semiconductors due to its critical breakdown field and high Baliga's figure of merit (BFOM), which has the potential for next-generation power device applications. Due to high material cost and purity, structural instability, low electron mobility, and device scaling and fabrication complexity, research has broadened to focus on alternative UWBG semiconductors that can resolve those issues to some extent.8–10 

Ga-based spinels with the general formula AGa2O4, where A is a divalent metal atom, are an example of ternary systems that can provide UWBG (Eg > 3.5 eV) semiconducting behavior.11 The semiconducting behavior of bulk and thin films MgGa2O4 crystals with an energy gap of 4.9 eV has been studied but possesses very low carrier mobility.12,13 Another spinel zinc gallate (ZnGa2O4) exhibits excellent optical and electrical characteristics with an extremely wide energy gap between 4.6 and 5.2 eV. The ZnGa2O4 (ZGO) has a conventional cubic spinel structure of space group F d 3 ¯ m, and the cations of Zn2+ and Ga3+ are arranged in tetrahedral and octahedral lattice positions, respectively. Because of its UWBG and proven ability to adjust electrical conductivity, ZGO may have similar potential to β-Ga2O3 and, due to its isotropic spinel structure, is considered a better option for device design than the monoclinic β-Ga2O3, which has a complex anisotropic structure–property relationship. ZGO is also a promising long-lasting brilliant phosphor that can be used for in vivo imaging by thermally driven luminescence and x-ray phosphorescence when doped with transition metal elements.14–16 ZGO is frequently studied in polycrystalline and nanopowder materials, and Galazka et al. discovered its easily adjustable electrical conductivity and reported the successful development of a single crystal high-quality bulk material.17 With a free electron concentration of ∼1020/cm3 and a Hall electron mobility of ∼107 cm2/V s at such high carrier concentrations, melt-grown ZGO crystals can be either conductive or made electrically insulating by Al doping. Studies of its inherent optical and electrical properties are limited in the literature even though these qualities are important for prospective usage for the aforementioned applications.18 Various diffuse reflectance spectroscopy (DRS) and transmittance spectroscopy techniques of ZGO samples have determined that it has a UWBG of 4.6–5 eV. The bandgap energy of a single crystal ZGO sample was observed to be 4.6 eV by Galazka et al.,18 using transmittance and reflectance spectroscopy and a calculated refractive index of 1.90 and 1.97 for wavelengths of 2 and 1 μm, respectively.19 Recently, Hilfiker et. al. have shown the variation in the dielectric function in the near-infrared to the deep-ultraviolet region with the critical point (CP) structures within the dielectric function, which are known to be caused by band-to-band transitions and associated exciton contributions.19 However, the effect of growth temperature and oxygen pressure during the fabrication of the ZGO thin film has a significant effect on the optical properties, bandgap, and dielectric function. The bandgap, dielectric constant, and carrier mobility have a significant role in Baliga's figure of merit (BFOM), which represents a very important parameter in determining the properties of (UWBG) semiconductors in optoelectronic and power electronics applications.20 Therefore, the enhancement of these parameters, i.e., bandgap and dielectric constant of UWBG ZGO, is very important for the evaluation/improvement of the performance of power semiconductor devices.

In the present work, ZGO thin-film samples are grown heteroepitaxially on the sapphire substrate by pulsed laser deposition (PLD) under varying growth temperatures at an optimized oxygen pressure. Sapphire substrates have excellent mechanical and thermal stability, chemical inertness, and higher resistance to chemical corrosion than MgO and spinel substrates. It is possible to grow thin films on sapphire substrates with low defect density and high crystal quality. The performance of electronic and optoelectronic devices depends on the epitaxial growth of superior semiconductor layers, which is dependent on the high structural perfection of sapphire crystals.21 The structural and optical properties in terms of the bandgap, refractive index, and complex dielectric function of ZnGa2O4 thin films are investigated. Furthermore, PLD represents a more efficient deposition method to build complex metal oxide thin films with different metal dopants in a relatively extensive doping range than other epitaxial thin-film development techniques such as metalorganic chemical vapor deposition (MOCVD), molecular beam epitaxy (MBE), and atomic layer deposition (ALD).22 For basic research on the characteristics of ZnGa2O4-based epitaxial thin films and heterostructures for prospective device applications, PLD is considered one of the most cost-effective techniques. Recently, heteroepitaxial ZnGa2O4 thin films have been deposited on cubic spinel MgAl2O4 (MAO) as well as on cubic MgO substrates for thin-film transistors (TFTs) applications23 and sapphire substrates for deep-ultraviolet applications24,25 employing PLD techniques. An in-depth analysis of the structural order and crystalline quality of ZGO thin films is carried out by x-ray diffraction (XRD), φ-scan, and ω-scan. X-ray photoelectron spectroscopy (XPS) allows for the determination of the elemental composition of Zn and Ga atomic ratios, cation oxidation states, etc. The cross-sectional images, interfacial characteristics, crystalline nature, and fast-Fourier transform (FFT) images are visualized using high-resolution transmission electron microscopy (HRTEM). The optical absorbance spectra of ZGO thin films are carried out by UV–vis spectroscopy, and the bandgaps are determined using the Kubelka–Munk equation and Tauc’s plot. By performing a regression analysis on the optical model and measured ellipsometry data, a multiple-sample analysis (MSA)-based method was used to assess their thickness, dielectric function, and other optical properties.

The ZGO thin films are grown on c-plane sapphire (Al2O3) substrates using the PLD technique at varying substrate temperatures from 650 to 750 °C and an oxygen pressure of 1 × 10−2 Torr. High incongruent decomposition and evaporation of the ZnO–Ga2O3 binary system, as well as the significantly higher vapor pressure of Zn compared to Ga, have resulted in a narrow growth window for single-phase and nearly stoichiometric ZnGa2O4 epitaxial thin films.26,27 Therefore, the preparation of the PLD target is more important than the deposition parameters, and three targets taking ZnO and Ga2O3 with 50%–50%, 75%–25%, and 98%–2%, respectively, were prepared. The two metal oxides are mixed thoroughly by ball milling for 1 h and pelletized at room temperature and a mechanical pressure of 1.5 × 104 psi. All three pelletized samples were calcined at 1150 °C in an ambient atmosphere for 24 h and used as the PLD target. The vapor pressure of Zn and Ga is 0.0127 and 0.0091 Torr at 750 °C, respectively.28 Despite the use of mosaic targets containing 50% ZnO–50% Ga2O3 and 75% ZnO–25% Ga2O3, the evaporation of zinc on a heated substrate due to the relatively high vapor pressure results in severe Zn losses, which are already observed above 300 °C, and leads to significant Zn deficiency in the ZnGa2O4 films deposited above 700 °C. It is expected that the higher Zn (98%) containing target can effectively offset the significant Zn loss that occurs during thin-film deposition. The ZGO thin films were grown at varying growth temperatures from 650 to 750 °C with a constant oxygen partial pressure of ∼1 × 10−2 Torr. A laser fluence of ∼3.33 J/cm2 is used to ablate and deposit the thin film since it facilitates higher kinetic energies and larger flux for the incorporation of Zn into the growing ZnGa2O4 crystalline films with the target supplying additional Zn to partially compensate for the Zn loss. The number of laser pulses and laser frequency of 7500 and 5 Hz, respectively, were used for all depositions.

The structural characterizations of ZGO thin films including XRD, φ-scan, and x-ray rocking curve (XRC) were performed using a Rigaku SmartLab x-ray diffractometer. The XRD was performed at 2θ between 10° and 90°, a scan rate of 2°/min, and a step size of 0.0002° for all samples. The φ-scan and XRC (ω-scan) were accomplished between −10° and 360° (for 104 planes for sapphire and 220 planes for ZnGa2O4) and between 17° and 20°, respectively. XPS spectra were collected from a ThermoFisher Scientific Nexsa x-ray photoelectron spectrometer (XPS) using soft x-ray radiation (from Al sources). The HRTEM and fast-Fourier transform (FFT) images were obtained using a FEI Talos f200i S/TEM with accelerating voltages between 120 and 200 kV. High-resolution scanning electron microscopy (HRSEM) images with energy-dispersive x-ray spectra (EDS) and mappings were collected using a FEI Helios NanoLab 400 DualBeam system in the field-free mode. The absorbance spectra and bandgap of ZGO thin films were examined from the data obtained from a Shimadzu UV-2600 UV–vis spectrophotometer in the range between 200 and 800 nm. The complex refractive index, extinction coefficient, and dielectric function of ZGO thin films at different growth temperatures were acquired by spectroscopic ellipsometry measurement (J.A. Woollam M2000X system).

To confirm the single-crystalline phase of the ZGO thin film for the three different targets on sapphire substrates, the XRD spectra at a growth temperature of 750 °C and a constant oxygen pressure of 1 × 10−2 Torr are shown in Fig. 1(a). The ZGO thin film grown by the stoichiometrically balanced target (50%–50%) exhibits XRD peaks at 18.9°, 38.4°, 59.0°, and 81.9°, which are assigned to the single-crystalline XRD peaks of β-Ga2O3.29 Even 75% ZnO−25% Ga2O3 targets displayed some secondary peaks with single crystalline XRD peaks of β-Ga2O3 indicating significant re-evaporation of Zn/ZnO. Finally, the XRD spectrum for the ZGO thin film deposited using the Zn-rich (98% ZnO–2% Ga2O3) target illustrates single-crystalline XRD peaks at 18.48°, 37.44°, 57.49°, and 79.76°, which corresponds to the cubic phase of the ZGO thin film (JCPDS-381240).30 In order to view distinctly the difference between the XRD peaks of the β-Ga2O3 and ZGO thin film, an enlarged view of the (111) planes is shown in Fig. 1(b). There is a significant shift in the peaks for the (111) planes toward a lower diffraction angle due to the successful formation of a single-crystalline ZnGa2O4 at 750 °C. The peak shift and phase transition from β-Ga2O3 to ZnGa2O4 agree well with the MOCVD-grown ZnGa2O4 by Chikoidze et. al.31, Figure 1(c) represents the XRD spectra of ZGO samples grown at various substrate temperatures from 650 to 750 °C and an oxygen pressure of 1 × 10−2 Torr for the 98%–2% target. It exhibits highly oriented crystal growth with a series of reflection planes (111), (222), (333), and (444) associated with the cubic spinel structure of ZnGa2O4.32 All peaks of ZGO were calibrated and aligned using the sapphire (0006) planes at 41.72° (c = 1.297 nm). A number of secondary peaks are observed for the samples grown at temperatures lower than 700 °C due to unsuccessful incorporations of ZnO over Ga2O3. Consequently, a smaller FWHM of XRD for the sample grown at 750 °C indicates better crystalline quality and fewer imperfections, which is later confirmed by x-ray rocking curves. Figure 1(d) displays the single-crystalline XRD spectra of the ZGO sample at 750 °C in a log scale to view distinctly each peak. One very small additional peak at 2θ ∼ 85.9° is detected in the log scale, which is most likely due to the (62-4) plane and suggests the presence of a minor (<1%) polycrystalline nature of ZnGa2O4. The XRD φ-scans around the (220) out-of-plane peaks of the ZnGa2O4 thin film formed at temperature 750 °C and oxygen pressure 1 × 10−2 Torr and the (104) out-of-plane peaks of the sapphire substrate were performed to estimate the in-plane orientation and epitaxial relationship between the films and substrates. The ZGO film has six peaks that are visible at 60° intervals (represented by the red line), suggesting that it is sixfold symmetric and that in-plane twinning is present in Fig. 1(e). The φ-scans by black lines exhibit three sharp peaks between 0° and 360°, which represents the threefold rotational symmetry of the c-plane sapphire. The in-plane epitaxial matching of the (111) oriented ZnGa2O4 thin film on the (0006) sapphire substrate is confirmed by the rotation of the six peaks of ZGO films by 30° with respect to the threefold peaks from the substrate. The crystalline quality of the as-grown ZGO samples for the three different temperatures 700°, 725°, and 750° at an oxygen partial pressure of 1 × 10−2 Torr is examined by x-ray rocking curve (XRC) and shown in Fig. 1(f). The values of full width and half maxima (FWHM) of the rocking curves were determined to be 1.02°, 0.32°, and 0.098° of the (111) oriented ZGO thin film for 700, 725, and 750 °C, respectively, which indicates the excellent epitaxial crystal quality and relationship of the heteroepitaxial ZGO thin film on sapphire even with a lattice mismatch greater than 10% between the substrate and film. With the increase in the growth temperature, the FWHM of rocking curves decreases and indicates better crystalline quality for higher temperatures.33 The FWHM value of ∼0.098° for the 750 °C grown ZGO thin film (222) indicates better crystalline quality than previously reported works of ZnGa2O4.

FIG. 1.

(a) The XRD patterns of ZnGa2O4 on the sapphire substrate for three different targets, (b) comparison of (111) planes of ZGO grown by zinc-rich and stoichiometrically balanced target, (c) XRD pattern of the ZGO thin film on sapphire at different growth temperatures from 650 to 750 °C and oxygen pressure 1 × 10−2 Torr, (d) XRD pattern in the log scale of 750 °C grown samples with an additional minor peak (62-4), (e) phi (φ) scan of the 220 plane for the (111) orientated ZGO thin film (750 °C), and the 104 plane of sapphire, (f) x-ray rocking curve of the ZGO thin film grown at 700, 725, and 750 °C (arb. units illustrate the arbitrary units).

FIG. 1.

(a) The XRD patterns of ZnGa2O4 on the sapphire substrate for three different targets, (b) comparison of (111) planes of ZGO grown by zinc-rich and stoichiometrically balanced target, (c) XRD pattern of the ZGO thin film on sapphire at different growth temperatures from 650 to 750 °C and oxygen pressure 1 × 10−2 Torr, (d) XRD pattern in the log scale of 750 °C grown samples with an additional minor peak (62-4), (e) phi (φ) scan of the 220 plane for the (111) orientated ZGO thin film (750 °C), and the 104 plane of sapphire, (f) x-ray rocking curve of the ZGO thin film grown at 700, 725, and 750 °C (arb. units illustrate the arbitrary units).

Close modal

The elemental composition and bonding characteristics of the ZGO thin film grown at 750 °C were analyzed by XPS and are shown in Fig. 2. The XPS survey scan in Fig. 2(a) divulges various sharp peaks corresponding to Ga-3s, Ga-2p, Ga-3p, Ga-3d, Ga-LMM, Zn-2p, Zn-LMM, O-1s, and C-1s bonds and the obtained oxidation states +2 for Zn, +3 for Ga, and −2 for oxygen. The atomic percentages ratio of Zn/Ga was approximated at ∼0.48 from the intensity and area under the Zn-2p3/2 and Ga-2p3/2 curves. The atomic ratios of Zn:Ga:O obtained were 15.98%, 30.12%, and 53.9%, respectively, from the XPS analysis, which corresponds well to the energy dispersive x-ray (EDX) spectra shown in Fig. S1(g) in the supplementary material. The individual spectrum of Zn-2p in Fig. 2(b) reveals two broad and sharp peaks at 1021.5 and 1044.8 eV, representing Zn-2p3/2 and Zn-2p1/2, respectively. In addition, the XPS spectrum of Ga-3d in Fig. 2(c) is deconvoluted with three different contributions of O-2s, Ga3+, and Ga at 22.9, 20.2, and 18.5 eV, which originates due to an overlapping of O-2s states, Ga3+ lattice ions in Ga-3d states, and minor elemental gallium residue, respectively. The individual scan of O-1s peaks in Fig. 2(d) is deconvoluted to the two components, which are assigned to the lattice oxygen (O2−) (530.8 eV) and oxygen vacancies (532.5 eV), respectively. The broad symmetric peak at 530.8 eV corresponds to the lattice oxygen ions (O2−) in ZGO bonded with Ga and Zn, and the small peak at 532.5 eV is designated to the oxygen ions in the oxygen vacancy region. Interestingly, an oxygen vacancy in an oxygen-deficient interlayer functions as an electron trap, allowing oxide semiconductor conductivity to be regulated.34 

FIG. 2.

(a) XPS survey scan, (b) core-level spectra of Zn-2p, (c) core-level spectra of Ga-3d, (d) oxygen O-1s spectra of the ZnGa2O4 thin film grown at 750 °C and an oxygen pressure of 1 × 10−2 Torr.

FIG. 2.

(a) XPS survey scan, (b) core-level spectra of Zn-2p, (c) core-level spectra of Ga-3d, (d) oxygen O-1s spectra of the ZnGa2O4 thin film grown at 750 °C and an oxygen pressure of 1 × 10−2 Torr.

Close modal

To observe the oxygen vacancies with growth temperatures, we have shown the O-1s spectra of ZGO grown at three different temperatures 700, 725, and 750 °C with Gaussian fitting as shown in Figs. 3(a)3(c), respectively. Usually, high growth temperatures may cause the thin film to develop thermally induced minor defects and these defects can result in oxygen vacancies being created by the loss of oxygen atoms from the lattice structure.35,36 From Figs. 3(a)3(c), the peaks around 532.5 eV are associated with the oxygen vacancies and their value slightly increases with growth temperatures. The obtained area under the curves after fitting is used to calculate the ratios of oxygen vacancies (OII) with lattice oxygen (O2−), and its values were determined to be 0.09, 0. 10, and 0.11 for 700, 725, and 750 °C, respectively.

FIG. 3.

The core-level O-1s XPS spectra of the ZGO thin film grown at (a) 700, (b) 725, and (c) 750 °C. The cyan-colored region signifies the oxygen content in oxygen vacancy regions.

FIG. 3.

The core-level O-1s XPS spectra of the ZGO thin film grown at (a) 700, (b) 725, and (c) 750 °C. The cyan-colored region signifies the oxygen content in oxygen vacancy regions.

Close modal

The structure and characteristics of epitaxial thin films can be effectively characterized at the nanoscale level using transmission electron microscopy (TEM). The high-resolution TEM analysis of the ZGO thin film grown at 750 °C was performed to investigate the microstructure as well as the interface between the sapphire/ZGO thin film as shown in Fig. 4. An ultrathin cross-sectional specimen of the epitaxial thin film was prepared using a focused ion beam (FIB) to mill and thin the samples to electron transparent thickness. The cross-sectional TEM images in Fig. 4(a) are demonstrated to evaluate the film thickness, which was found to be ∼265 nm (indicated by the green line), considering the small sample volume and likely outermost surface amorphization following the focused ion beam (FIB) thinning process. The interface between the sapphire and ZGO film is clearly visible in Fig. 4(b), and it confirms the epitaxial relationship of nearly 30° obtained from XRD analysis. The interface between the substrate and film appears very sharp and distinct, and there is no sign of a minor secondary phase from unreacted residue elements. The quasi-single-crystalline structure of the transparent crystal lattice stripes of cubic ZnGa2O4 with the preferred (111) orientation perpendicular to the film surface illustrates the d-spacing value of 4.82 Å in Fig. 4(c). Minor contributions of ZnO and Ga2O3 crystal stripes were observed on a large scale near interfaces, which arises due to rapid thermal annealing during growth and bond rearrangement at the interface between the amorphous and crystalline phases.37,38 The FFT pattern from the region of Fig. 4(d) contains the film and substrate with several reflection planes of the ZGO film such as (111), (011), and (002) and gives an out-of-plane orientation relationship with the substrate. The scanning electron microscopy (SEM) micrographs with EDS and mappings of the ZGO thin film grown at 750 °C are also displayed in Fig. S1 in the supplementary material. The uniform distribution of Zn, Ga, and O with thickness ∼247 nm was confirmed by the elemental mappings and EDX spectra with stoichiometrically balanced atomic distributions of 14.37%, 28.28%, and 57.35%, respectively.

FIG. 4.

(a) Cross-sectional TEM images showing the thickness of the ZGO thin film with the interface of the sapphire substrate and top carbon layer, (b) interface between the (111) oriented ZGO thin film and the sapphire substrate (0006), (c) HRTEM images of 111 crystal planes of the ZGO thin film with crystal spacings 4.82 Å, (d) FFT pattern of the ZGO film on the sapphire c-plane substrate.

FIG. 4.

(a) Cross-sectional TEM images showing the thickness of the ZGO thin film with the interface of the sapphire substrate and top carbon layer, (b) interface between the (111) oriented ZGO thin film and the sapphire substrate (0006), (c) HRTEM images of 111 crystal planes of the ZGO thin film with crystal spacings 4.82 Å, (d) FFT pattern of the ZGO film on the sapphire c-plane substrate.

Close modal
UV–vis spectroscopy is one of the most powerful and versatile techniques in thin-film research, and it provides us with valuable information about electronic transition and bandgap of any materials. The absorbance vs wavelength of ZGO thin films grown at 700, 725, and 750 °C is shown in Fig. 5(a) in the range between 200 and 800 nm, and it exhibits very good optical transparency with minor distinct fringes in both UV and visible regions. The optical bandgap ( E g) of ZGO thin films for the three aforementioned temperatures was determined by the well-known Tauc's plot,39–41 
(1)
where α is the absorbance coefficient and it is related to absorbance by42 
(2)
FIG. 5.

(a) UV–vis absorption spectra, (b) Tauc’s plot for calculations of the band gap, (c) Refractive index (n) vs wavelength, (d) extinction coefficient (k) vs wavelength, (e) real dielectric function (ɛr) vs energy, (f) imaginary dielectric function (ɛi) vs energy of the ZGO thin film grown at 700, 725, and 750 °C.

FIG. 5.

(a) UV–vis absorption spectra, (b) Tauc’s plot for calculations of the band gap, (c) Refractive index (n) vs wavelength, (d) extinction coefficient (k) vs wavelength, (e) real dielectric function (ɛr) vs energy, (f) imaginary dielectric function (ɛi) vs energy of the ZGO thin film grown at 700, 725, and 750 °C.

Close modal

Here, h represents the Planck constant, ν is the frequency of vibrations, n = 2 for direct bandgap transitions and ½ for indirect bandgap transitions, and A and t denote the absorbance and thickness of the film. The Tauc’s plots in Fig. 5(b) show an optical bandgap of 5.08, 5.11, and 5.14 eV for growth temperatures of 700, 725, and 750 °C, respectively. The practical bandgap of the high crystalline ZGO thin film grown at 750 °C gives a real bandgap that is affected by changes in the carrier concentration at higher growth temperatures. Bandgap renormalization, a variety of many-body interactions between the carriers in the conduction band and valence band, is known to be associated with the bandgap narrowing in the ZGO thin film at lower growth temperatures.43 Additionally, the electronic energy bands may become more widely spaced as a result of these thermally assisted vibrations, which would facilitate the movement of electrons from the valence to the conduction bands and narrow the bandgap.

The spectroscopic ellipsometry in the range between 200 and 1000 nm was performed on ZGO thin films, and the experimental data furnished by the system are ellipsometric angles ψ and δ, which are the functions of the ratio of amplitude (rp) and phase (rs) by44,45
(3)
where r p and r s represent the complex Fresnel amplitude reflection coefficients for the parallel and perpendicular components. To determine the optical constant and thickness of ZGO thin films, a model was developed and fitted to ψ and δ over a range of wavelengths. One typical plot of ψ and δ for ZGO grown at 750 °C is shown in Fig. S2 in the supplementary material for three different angles 55°, 65°, and 75° between 200–1000 nm after fitting with the model based on Fresnel’s equation. The amplitude, broadening, and central energy parameters of the Herzinger–Johs (HJ) parameterized semiconductor (PSemi) oscillator function-based model were used as fitting parameters for the experimental data and determined from the regression analysis keeping the shape parameters fixed. While the optical properties were taken to be similar in accordance with the mean squared error (MSE) technique, the thickness and roughness values were also allowed to vary during the fitting procedure.46 Using this model, the measured value is compared using an estimator such as the MSE with an unknown variable of the model that is varied until the lowest MSE is obtained. The formula of MSE used for nonlinear regression is defined as44,47
(4)
where n and m are the number of spectral data points and the number of fit parameters in our optical model, respectively. The parameters N, C, and S are used for fitting, and they are described as N = cos ( 2 ψ ), C = sin ( 2 ψ ) cos δ, and S = sin ( 2 ψ ) sin δ, respectively, and their detailed calculation can be found elsewhere.48 Usually, these three parameters are measured in the precision of ∼0.001and the ideal data modeling should have an MSE value ∼1–5. In our case, the MSE value was obtained at 4.9, 4.5, and 3.2 for the ZGO thin film grown at 700, 725, and 750 °C, respectively, after fitting.
The optical properties consist of two parameters: the refractive index (n) and the extinction coefficient (k). The ratio of light speed in a vacuum to any substance is known as the refractive index (n), and the absorption loss in the material is connected to the extinction coefficient (k). These two elements combined provide the complex refractive index, N = n i k, which expresses how electromagnetic radiation interacts with different materials in terms of both the absorption loss and speed change. In this case, these two optical constants were derived from the fitting of experimental data by an extended Cauchy dispersion model,44,49
(5)
and
(6)

The A, B, and C are the fitting parameters that control the line shape of the refractive index and k 0 and E g represent the amplitude of absorption and band edges, respectively. The variation in the refractive index (n) and extinction coefficient (k) with a wavelength between 200 and 1000 nm is shown in Figs. 5(c) and 5(d), respectively, with their corresponding values. The near-infrared refractive index (n) was determined to be ∼1.94 for ZGO (750 °C) from the extrapolation, which is in very good agreement with recently published ZnGa2O4 results.19,46 Similarly, the extinction coefficient (k), which is an indication of the amount of light absorbed or transmitted in any medium, was obtained to be ∼0.023, representing potential optical applications for optical filters, coatings, fiber optics, and microscopy, among others. The dispersion of the refractive index with longer wavelengths, i.e., infrared or radio waves experiences relatively lower bending (sometimes called refraction) when they pass through the ZGO samples and, consequently, the refractive index slightly decreases.

The real (ɛr) and imaginary (ɛi) part of the dielectric function is associated with the electrical polarization anomalous dispersion and dissipation of energy through the medium, respectively. The dielectric function (ɛ) is obtained from the Kramers–Kronig relations,50,51
(7)
and solving Eq. (7), we can get the real and imaginary parts of the dielectric function,
(8)
and
(9)

The experimental real (ɛr) and imaginary (ɛi) dielectric functions are calculated using Eqs. (8) and (9), respectively, and their variation with photon energy further is emphasized in Figs. 5(e) and 5(f), respectively. The real dielectric function of two ZGO samples grown at 700 and 725 °C follows an unalike trend and decreases after a maximum near 3.5 eV. However, the ɛr of 750 °C grown ZGO increases with photon energy producing a relaxation plateau near 5 eV and giving the highest value of ∼6.8. The variation in ɛr with energy well corresponds to the theoretical data by the density functional theory (DFT) of ZnGa2O4 by Hilfiker et. al.,19 and the obtained value of the dielectric function is slightly higher than their value (∼6.2) near the relaxation anomaly region. We observed that anharmonic coupling causes the imaginary component of the dielectric function to be slightly negative from the exciton model contribution throughout a small spectral region of the 750 °C grown ZGO sample. Anharmonic coupling of two resonant processes results in a phase shift between their oscillatory characteristics because of the exchange of excitation energy. An identical observation was reported by α-phase cubic Ga2O3 by Hilfiker et.al.19 The roughness of ZGO samples grown between 700 and 750 °C was obtained between ∼18 and 25 nm, respectively. We have compared the growth and optical parameters of some recent studies of ZnGa2O4 grown on sapphire in Table I.

TABLE I.

Comparison of growth and optical parameters of our ZnGa2O4 thin film with the previous study of single-crystalline ZnGa2O4 on sapphire.

Growth techniquesCrystalline quality (FWHM)Bandgap (eV)Refractive index (n)Extinction coefficient (k)Reference
PLD 0.1° 4.9 … … 52  
MOCVD 0.06° 5.07 1.92 0.025 46  
MOCVD … 5.2 … … 53  
RF magnetron sputtering 0.29° … … … 54  
PLD 0.098° 5.14 1.94 0.023 Our study 
Growth techniquesCrystalline quality (FWHM)Bandgap (eV)Refractive index (n)Extinction coefficient (k)Reference
PLD 0.1° 4.9 … … 52  
MOCVD 0.06° 5.07 1.92 0.025 46  
MOCVD … 5.2 … … 53  
RF magnetron sputtering 0.29° … … … 54  
PLD 0.098° 5.14 1.94 0.023 Our study 

In summary, our present work demonstrates the high-quality heteroepitaxial growth of the ZnGa2O4 (111) thin film on sapphire (c-plane) by PLD techniques using a zinc-rich target Zn0.98Ga0.02O and observes their epitaxial relationship, chemical state, structural insight, and optical properties. The ZnGa2O4 thin film was grown at different temperatures ranging from 650 to 750 °C with ΔT = 25 °C after optimizing the oxygen pressure 1 × 10−2 Torr. Due to the mismatch of the vapor pressure between Zn and Ga, the zinc-rich target Zn0.98Ga0.02O gives a balanced pulsed laser ablation of both elements than the stoichiometrically balanced target. The epitaxial relationship between the ZnGa2O4 (111) and sapphire (0001) substrates was established by the φ scan, and nearly 30° out-of-plane relationship was found. The x-ray rocking curve displayed a narrow FWHM of ∼0.098° for the ZGO sample grown at 750 °C. The chemical state of Zn, Ga, and O from the XPS curve provides sound information about the elemental composition, atomic percentages, and the minor oxygen vacancies compared to lattice oxygen (∼0.098 for 750 °C), which slightly increases with the rise of growth temperatures. The HRTEM micrographs with the FFT pattern of the ZGO (111) sample confirm the quasi-single-crystalline nature with minor additional phase and crystal lattice spacings of 4.82 Å. The Tauc plot from UV–vis spectra helped us to determine the UWBG gap of ∼5.08 eV for 750 °C grown samples, and an indicative effect of growth temperature on the bandgap was observed. The utmost refractive index of ∼1.94 and low extinction coefficient of ∼0.023 were obtained for the 750 °C grown ZGO sample from the ellipsometric data using the Cauchy dispersion formula. The variation in the real dielectric function with energy exhibits one relaxation plateau near 5 eV with a maximum value of ∼6.8, which is higher than the previously reported value. The UWBG gap and obtained dielectric function of epitaxially grown ZnGa2O4 recommend its probable applications in high-power electronics and DUV sensors.

See the supplementary material for the high-resolution field emission scanning electron microscopy (FESEM) micrographs of the top surface, cross-sectional SEM images, and EDX elemental mapping of the ZGO thin film (Fig. S1), and the experimental and fitted ellipsometric angles (a) ψ and (b) δ with a wavelength between 200 and 1000 nm (Fig. S2).

We want to acknowledge the project funding from the Office of Research and Sponsored Programs (ORSP) of Texas State University awarded to PI Haque and the instrumentation facilities at Shared Research Operation (SRO). The first author, SK, would also like to thank his colleagues from the advanced UWBG materials and device lab of the Ingram School of Engineering, Texas State University.

The authors have no conflicts to disclose.

Subrata Karmakar: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Injamamul Hoque Emu: Data curation (supporting); Formal analysis (supporting); Investigation (supporting). Md Abdul Halim: Data curation (supporting); Investigation (supporting). Pallab Kumar Sarkar: Data curation (supporting). Maria Sultana: Data curation (supporting). Ayesha Tasnim: Data curation (supporting). Md Abdul Hamid: Data curation (supporting). Istiaq Firoz Shiam: Data curation (supporting). Ravi Droopad: Formal analysis (supporting); Project administration (supporting); Resources (equal); Supervision (supporting); Writing – original draft (supporting); Writing – review & editing (equal). Ariful Haque: Funding acquisition (equal); Project administration (equal); Resources (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article and its supplementary material.

1.
M.
Xu
,
D.
Wang
,
K.
Fu
,
D. H.
Mudiyanselage
,
H.
Fu
, and
Y.
Zhao
, “
A review of ultrawide bandgap materials: Properties, synthesis and devices
,”
Oxford Open Mater. Sci.
2
(
1
),
itac004
(
2022
).
2.
S.
Choi
,
S.
Graham
,
S.
Chowdhury
,
E. R.
Heller
,
M. J.
Tadjer
,
G.
Moreno
, and
S.
Narumanchi
, “
A perspective on the electro-thermal co-design of ultra-wide bandgap lateral devices
,”
Appl. Phys. Lett.
119
(
17
),
170501
(
2021
).
3.
X.
Chen
,
C.
Zhang
,
F.
Meng
,
T.
Yu
, and
J.
Zhao
, “
Polishing mechanism analysis of silicon carbide ceramics combined ultrasonic vibration and hydroxyl
,”
Tribology International
179
,
108187
(
2023
).
4.
S.
Yuvaraja
,
V.
Khandelwal
,
X.
Tang
, and
X.
Li
, “
Wide bandgap semiconductor-based integrated circuits
,”
Chip
2
,
100072
(
2023
).
5.
Y.
Zhang
, “
Comparison between competing requirements of GaN and SiC family of power switching devices
,”
IOP Conf. Ser.: Mater. Sci. Eng.
738
(
1
),
012004
(
2020
).
6.
M.
Higashiwaki
,
R.
Kaplar
,
J.
Pernot
, and
H.
Zhao
, “
Ultrawide bandgap semiconductors
,”
Appl. Phys. Lett.
118
(
20
),
200401
(
2021
).
7.
J. Y.
Tsao
,
S.
Chowdhury
,
M. A.
Hollis
,
D.
Jena
,
N. M.
Johnson
,
K. A.
Jones
,
R. J.
Kaplar
,
S.
Rajan
,
C. G.
Van de Walle
,
E.
Bellotti
,
C. L.
Chua
,
R.
Collazo
,
M. E.
Coltrin
,
J. A.
Cooper
,
K. R.
Evans
,
S.
Graham
,
T. A.
Grotjohn
,
E. R.
Heller
,
M.
Higashiwaki
,
M. S.
Islam
,
P. W.
Juodawlkis
,
M. A.
Khan
,
A. D.
Koehler
,
J. H.
Leach
,
U. K.
Mishra
,
R. J.
Nemanich
,
R. C. N.
Pilawa-Podgurski
,
J. B.
Shealy
,
Z.
Sitar
,
M. J.
Tadjer
,
A. F.
Witulski
,
M.
Wraback
, and
J. A.
Simmons
, “
Ultrawide-bandgap semiconductors: Research opportunities and challenges
,”
Adv. Electron. Mater.
4
(
1
),
1600501
(
2018
).
8.
M.
Higashiwaki
,
K.
Sasaki
,
H.
Murakami
,
Y.
Kumagai
,
A.
Koukitu
,
A.
Kuramata
,
T.
Masui
, and
S.
Yamakoshi
, “
Recent progress in Ga2O3 power devices
,”
Semicond. Sci. Technol.
31
(
3
),
034001
(
2016
).
9.
C. J. H.
Wort
and
R. S.
Balmer
, “
Diamond as an electronic material
,”
Mater. Today
11
(
1–2
),
22
28
(
2008
).
10.
M.
Higashiwaki
and
S.
Fujita
, in
Gallium Oxide: Materials Properties, Crystal Growth, and Devices
(
Springer International Publishing
,
Cham
,
2020
).
11.
Z.
Galazka
,
Transparent Semiconducting Oxides: Bulk Crystal Growth and Fundamental Properties
, 1st ed. (
Jenny Stanford Publishing
,
2020
).
12.
Y.-N.
Kao
,
W.-L.
Huang
,
S.-P.
Chang
,
W.-C.
Lai
, and
S.-J.
Chang
, “
Investigation of different oxygen partial pressures on MgGa2O4-resistive random-access memory
,”
ACS Omega
8
(
4
),
3705
3712
(
2023
).
13.
M. A.
Hamid
,
B.
Samuels
,
S.
Karmakar
,
M. A.
Halim
,
I. H.
Emu
,
P. K.
Sarkar
,
M. F. N.
Taufique
,
A.
Haque
, and
R.
Droopad
, “
Epitaxial growth and characterization of magnesium gallate (MgGa2O4) thin films by pulsed laser deposition
,”
J. Alloys Compd.
972
,
172807
(
2024
).
14.
A.
Sood
,
F.-G.
Tarntair
,
Y.-X.
Wang
,
T.-C.
Chang
,
Y.-H.
Chen
,
P.-L.
Liu
, and
R.-H.
Horng
, “
Performance enhancement of ZnGa2O4 Schottky type deep-ultraviolet photodetectors by oxygen supercritical fluid treatment
,”
Results Phys.
29
,
104764
(
2021
).
15.
H.
Dixit
,
N.
Tandon
,
S.
Cottenier
,
R.
Saniz
,
D.
Lamoen
,
B.
Partoens
,
V.
Van Speybroeck
, and
M.
Waroquier
, “
Electronic structure and band gap of zinc spinel oxides beyond LDA: ZnAl2O4, ZnGa2O4 and ZnIn2O4
,”
New J. Phys.
13
(
6
),
063002
(
2011
).
16.
D.
Han
,
K.
Liu
,
Q.
Hou
,
X.
Chen
,
J.
Yang
,
B.
Li
,
Z.
Zhang
,
L.
Liu
, and
D.
Shen
, “
Self-powered solar-blind ZnGa2O4 UV photodetector with ultra-fast response speed
,”
Sens. Actuators, A
315
,
112354
(
2020
).
17.
Z.
Galazka
,
S.
Ganschow
,
K.
Irmscher
,
D.
Klimm
,
M.
Albrecht
,
R.
Schewski
,
M.
Pietsch
,
T.
Schulz
,
A.
Dittmar
,
A.
Kwasniewski
,
R.
Grueneberg
,
S. B.
Anooz
,
A.
Popp
,
U.
Juda
,
I. M.
Hanke
,
T.
Schroeder
, and
M.
Bickermann
, “
Bulk single crystals of β-Ga2O3 and Ga-based spinels as ultra-wide bandgap transparent semiconducting oxides
,”
Prog. Cryst. Growth Charact. Mater.
67
(
1
),
100511
(
2021
).
18.
Z.
Galazka
,
S.
Ganschow
,
R.
Schewski
,
K.
Irmscher
,
D.
Klimm
,
A.
Kwasniewski
,
M.
Pietsch
,
A.
Fiedler
,
I.
Schulze-Jonack
,
M.
Albrecht
,
T.
Schröder
, and
M.
Bickermann
, “
Ultra-wide bandgap, conductive, high mobility, and high quality melt-grown bulk ZnGa2O4 single crystals
,”
APL Mater.
7
(
2
),
022512
(
2019
).
19.
M.
Hilfiker
,
M.
Stokey
,
R.
Korlacki
,
U.
Kilic
,
Z.
Galazka
,
K.
Irmscher
,
S.
Zollner
, and
M.
Schubert
, “
Zinc gallate spinel dielectric function, band-to-band transitions, and Γ-point effective mass parameters
,”
Appl. Phys. Lett.
118
(
13
),
132102
(
2021
).
20.
L.
Cheng
,
J.-Y.
Yang
, and
W.
Zheng
, “
Bandgap, mobility, dielectric constant, and Baliga’s figure of merit of 4H-SiC, GaN, and β-Ga2O3 from 300 to 620 K
,”
ACS Appl. Electron. Mater.
4
(
8
),
4140
4145
(
2022
).
21.
Y.
Yu
,
P. W. K.
Fong
,
S.
Wang
, and
C.
Surya
, “
Fabrication of WS2/GaN p-n junction by wafer-scale WS2 thin film transfer
,”
Sci. Rep.
6
(
1
),
37833
(
2016
).
22.
M.
Lorenz
,
H.
Hochmuth
,
C.
Grüner
,
H.
Hilmer
,
A.
Lajn
,
D.
Spemann
,
M.
Brandt
,
J.
Zippel
,
R.
Schmidt-Grund
,
H.
Von Wenckstern
, and
M.
Grundmann
, “
Oxide thin film heterostructures on large area, with flexible doping, low dislocation density, and abrupt interfaces: Grown by pulsed laser deposition
,”
Laser Chem.
2010
,
1
27
(
2010
).
23.
Y.
Jang
,
S.
Hong
,
J.
Seo
,
H.
Cho
,
K.
Char
, and
Z.
Galazka
, “
Thin film transistors based on ultra-wide bandgap spinel ZnGa2O4
,”
Appl. Phys. Lett.
116
(
20
),
202104
(
2020
).
24.
C.-C.
Yen
,
A. K.
Singh
,
H.
Chang
,
K.-P.
Chang
,
P.-W.
Chen
,
P.-L.
Liu
, and
D.-S.
Wuu
, “
Pulsed laser deposition grown non-stoichiometry transferred ZnGa2O4 films for deep-ultraviolet applications
,”
Appl. Surf. Sci.
597
,
153700
(
2022
).
25.
A.
Guo
,
L.
Zhang
,
N.
Cao
,
T.
Lu
,
Y.
Zhu
,
D.
Tian
,
Z.
Zhou
,
S.
He
,
B.
Xia
, and
F.
Zhao
, “
Pulsed laser deposition of ZnGa2O4 thin films on Al2O3 and Si substrates for deep optoelectronic devices applications
,”
Appl. Phys. Express
16
(
2
),
021004
(
2023
).
26.
Y. E.
Lee
,
D. P.
Norton
,
C.
Park
, and
C. M.
Rouleau
, “
Blue photoluminescence in ZnGa2O4 thin-film phosphors
,”
J. Appl. Phys.
89
(
3
),
1653
1656
(
2001
).
27.
Y. E.
Lee
,
D. P.
Norton
, and
J. D.
Budai
, “
Enhanced photoluminescence in epitaxial ZnGa2O4:Mn thin-film phosphors using pulsed-laser deposition
,”
Appl. Phys. Lett.
74
(
21
),
3155
3157
(
1999
).
28.
C. B.
Alcock
,
V. P.
Itkin
, and
M. K.
Horrigan
, “
Vapour pressure equations for the metallic elements: 298–2500 K
,”
Can. Metall. Q.
23
(
3
),
309
313
(
1984
).
29.
S.
Ghose
,
S.
Rahman
,
L.
Hong
,
J. S.
Rojas-Ramirez
,
H.
Jin
,
K.
Park
,
R.
Klie
, and
R.
Droopad
, “
Growth and characterization of β-Ga2O3 thin films by molecular beam epitaxy for deep-UV photodetectors
,”
J. Appl. Phys.
122
(
9
),
095302
(
2017
).
30.
A. K.
Singh
,
S.-Y.
Huang
,
P.-W.
Chen
,
J.-L.
Chiang
, and
D.-S.
Wuu
, “
The effect of annealing ambience on the material and photodetector characteristics of sputtered ZnGa2O4 films
,”
Nanomaterials
11
(
9
),
2316
(
2021
).
31.
E.
Chikoidze
,
C.
Sartel
,
I.
Madaci
,
H.
Mohamed
,
C.
Vilar
,
B.
Ballesteros
,
F.
Belarre
,
E.
del Corro
,
P.
Vales-Castro
,
G.
Sauthier
,
L.
Li
,
M.
Jennings
,
V.
Sallet
,
Y.
Dumont
, and
A.
Pérez-Tomás
, “
p-type ultrawide-band-gap spinel ZnGa2O4: New perspectives for energy electronics
,”
Cryst. Growth Des.
20
(
4
),
2535
2546
(
2020
).
32.
R.-H.
Horng
,
C.-Y.
Huang
,
S.-L.
Ou
,
T.-K.
Juang
, and
P.-L.
Liu
, “
Epitaxial growth of ZnGa2O4: A new, deep ultraviolet semiconductor candidate
,”
Cryst. Growth Des.
17
(
11
),
6071
6078
(
2017
).
33.
C. J.
Tun
,
J. K.
Sheu
,
M. L.
Lee
,
C. C.
Hu
,
C. K.
Hsieh
, and
G. C.
Chi
, “
Effects of thermal annealing on Al-doped ZnO films deposited on p-type gallium nitride
,”
J. Electrochem. Soc.
153
(
4
),
G296
(
2006
).
34.
K.-J.
Gan
,
P.-T.
Liu
,
T.-C.
Chien
,
D.-B.
Ruan
, and
S. M.
Sze
, “
Highly durable and flexible gallium-based oxide conductive-bridging random access memory
,”
Sci. Rep.
9
(
1
),
14141
(
2019
).
35.
K.
Wang
,
Y.
Chang
,
L.
Lv
, and
Y.
Long
, “
Effect of annealing temperature on oxygen vacancy concentrations of nanocrystalline CeO2 film
,”
Appl. Surf. Sci.
351
,
164
168
(
2015
).
36.
T.
Kim
,
G.
Baek
,
S.
Yang
,
J. Y.
Yang
,
K. S.
Yoon
,
S. G.
Kim
,
J. Y.
Lee
,
H. S.
Im
, and
J. P.
Hong
, “
Exploring oxygen-affinity-controlled TaN electrodes for thermally advanced TaOx bipolar resistive switching
,”
Sci. Rep.
8
(
1
),
8532
(
2018
).
37.
W.-K.
Wang
,
K.-F.
Liu
,
P.-C.
Tsai
,
Y.-J.
Xu
, and
S.-Y.
Huang
, “
Influence of annealing temperature on the properties of ZnGa2O4 thin films by magnetron sputtering
,”
Coatings
9
(
12
),
859
(
2019
).
38.
P.
Chen
,
S.
Huang
,
S.
Yuan
,
Y.
Chen
,
P.
Hsiao
, and
D.
Wuu
, “
Quasi-single-crystalline ZnGa2O4 films via solid phase epitaxy for enhancing deep-ultraviolet photoresponse
,”
Adv. Mater. Interfaces
6
(
18
),
1901075
(
2019
).
39.
P.
Makuła
,
M.
Pacia
, and
W.
Macyk
, “
How To correctly determine the band gap energy of modified semiconductor photocatalysts based on UV–Vis spectra
,”
J. Phys. Chem. Lett.
9
(
23
),
6814
6817
(
2018
).
40.
S.
Karmakar
,
B.
Panda
,
B.
Sahoo
,
K. L.
Routray
,
S.
Varma
, and
D.
Behera
, “
A study on optical and dielectric properties of Ni-ZnO nanocomposite
,”
Mater. Sci. Semicond. Process.
88
,
198
206
(
2018
).
41.
S.
Karmakar
,
K. L.
Routray
,
B.
Panda
,
B.
Sahoo
, and
D.
Behera
, “
Construction of core@shell nanostructured NiFe2O4@TiO2 ferrite NAND logic gate using fluorescence quenching mechanism for TiO2 sensing
,”
J. Alloys Compd.
765
,
527
537
(
2018
).
42.
V.
Mursyalaat
,
V. I.
Variani
,
W. O. S.
Arsyad
, and
M. Z.
Firihu
, “
The development of program for calculating the band gap energy of semiconductor material based on UV-Vis spectrum using Delphi 7.0
,”
J. Phys. Conf. Ser.
2498
(
1
),
012042
(
2023
).
43.
J.-W.
Jeon
,
D.-W.
Jeon
,
T.
Sahoo
,
M.
Kim
,
J.-H.
Baek
,
J. L.
Hoffman
,
N. S.
Kim
, and
I.-H.
Lee
, “
Effect of annealing temperature on optical band-gap of amorphous indium zinc oxide film
,”
J. Alloys Compd.
509
(
41
),
10062
10065
(
2011
).
44.
X. D.
Zhu
,
S.
Wicklein
,
F.
Gunkel
,
R.
Xiao
, and
R.
Dittmann
, “
In situ optical characterization of LaAlO3 epitaxy on SrTiO3 (001)
,”
EPL
109
(
3
),
37006
(
2015
).
45.
N.
Hale
,
M.
Hartl
,
J.
Humlíček
,
C.
Brüne
, and
M.
Kildemo
, “
Dielectric function and band gap determination of single crystal CuFeS2 using FTIR-VIS-UV spectroscopic ellipsometry
,”
Opt. Mater. Express
13
(
7
),
2020
(
2023
).
46.
S.
Bairagi
,
C.-L.
Hsiao
,
R.
Magnusson
,
J.
Birch
,
J. P.
Chu
,
F.-G.
Tarntair
,
R.-H.
Horng
, and
K.
Järrendahl
, “
Zinc gallate (ZnGa2O4) epitaxial thin films: Determination of optical properties and bandgap estimation using spectroscopic ellipsometry
,”
Opt. Mater. Express
12
(
8
),
3284
(
2022
).
47.
C.
Yim
,
M.
O’Brien
,
N.
McEvoy
,
S.
Winters
,
I.
Mirza
,
J. G.
Lunney
, and
G. S.
Duesberg
, “
Investigation of the optical properties of MoS2 thin films using spectroscopic ellipsometry
,”
Appl. Phys. Lett.
104
(
10
),
103114
(
2014
).
48.
G. E.
Jellison
, “Data analysis for spectroscopic ellipsometry,”
Thin Solid Films
234
(
1–2
),
416
422
(
1993
).
49.
W.
Rzodkiewicz
and
A.
Panas
, “
Application of spectroscopic ellipsometry for investigations of compaction and decompaction state in Si-SiO2 systems
,”
J. Phys. Conf. Ser.
181
,
012035
(
2009
).
50.
P. Y.
Yu
and
M.
Cardona
,
Fundamentals of Semiconductors: Physics and Materials Properties
(
Springer
,
Berlin
,
2010
).
51.
W.
Jones
and
N. H.
March
,
Theoretical Solid State Physics. 2: Non-Equilibrium and Disorder, 1. Publ.
(
Dover
,
New York
,
1985
).
52.
S.
Luo
,
G. F.
Harrington
,
K.-T.
Wu
, and
T.
Lippert
, “
Heteroepitaxial (111) ZnGa2O4 thin films grown on (00.1) sapphire by pulsed laser deposition
,”
Phys. Rapid Res. Lett.
14
(
9
),
2000270
(
2020
).
53.
S.-H.
Tsai
,
S.
Basu
,
C.-Y.
Huang
,
L.-C.
Hsu
,
Y.-G.
Lin
, and
R.-H.
Horng
, “
Deep-ultraviolet photodetectors based on epitaxial ZnGa2O4 thin films
,”
Sci. Rep.
8
(
1
),
14056
(
2018
).
54.
X.
Wang
,
L.
Huang
, and
Tsai
, “
Structural characteristics and photoluminescence properties of sputter-deposition ZnGa2O4 thin films on sapphire and Si(100) substrates
,”
Coatings
9
(
8
),
469
(
2019
).