Next Article in Journal
Seismic Behavior Analysis of Recycled Aggregate Concrete-Filled Square Steel Tube Frames
Previous Article in Journal
Resorbable Biomaterials Used for 3D Scaffolds in Tissue Engineering: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Spatial Correlation and Anisotropy of β-(AlxGa1−x)2O3 Single Crystal

1
Center on Nanoenergy Research, Guangxi Colleges and Universities Key Laboratory of Blue Energy and Systems Integration, Carbon Peak and Neutrality Science and Technology Development Institute, School of Physical Science & Technology, Guangxi University, Nanning 530004, China
2
State Key Laboratory of Featured Metal Materials and Life-Cycle Safety for Composite Structures, Nanning 530004, China
3
Key Laboratory of Materials for High Power Laser, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, China
4
Department of Physics, University of North Florida, Jacksonville, FL 32224, USA
5
Department of Electrical & Computer Engineering, Southern Polytechnic College of Engineering and Engineering Technology, Kennesaw State University Marietta, Marietta, GA 30060, USA
6
Hangzhou Institute of Optics and Fine Mechanics, Hangzhou 311421, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(12), 4269; https://doi.org/10.3390/ma16124269
Submission received: 25 April 2023 / Revised: 5 June 2023 / Accepted: 6 June 2023 / Published: 8 June 2023

Abstract

:
The long-range crystallographic order and anisotropy in β-(AlxGa1−x)2O3 (x = 0.0, 0.06, 0.11, 0.17, 0.26) crystals, prepared by optical floating zone method with different Al composition, is systematically studied by spatial correlation model and using an angle-resolved polarized Raman spectroscopy. Alloying with aluminum is seen as causing Raman peaks to blue shift while their full widths at half maxima broadened. As x increased, the correlation length (CL) of the Raman modes decreased. By changing x, the CL is more strongly affected for low-frequency phonons than the modes in the high-frequency region. For each Raman mode, the CL is decreased by increasing temperature. The results of angle-resolved polarized Raman spectroscopy have revealed that the intensities of β-(AlxGa1−x)2O3 peaks are highly polarization dependent, with significant effects on the anisotropy with alloying. As the Al composition increased, the anisotropy of Raman tensor elements was enhanced for the two strongest phonon modes in the low-frequency range, while the anisotropy of the sharpest Raman phonon modes in the high-frequency region decreased. Our comprehensive study has provided meaningful results for comprehending the long-range orderliness and anisotropy in technologically important β-(AlxGa1−x)2O3 crystals.

1. Introduction

Gallium oxide (Ga2O3), with a wider bandgap energy of 4.8 eV material in the deep ultraviolet (UV) spectrum, is a transparent semiconductor. Its transparency in the UV region has allowed applications such as solar-blind photodetectors and deep-UV transparent contacts. The material with a high breakdown electric field (6–8 MV/cm), large thermal conductivity, low electron effective mass, and the ability to achieve controllable n-type doping with stable chemical properties has made Ga2O3 promising for engineering the next-generation electronic devices, especially in power-electronics, viz., rectifiers, metal-semiconductor field-effect transistors (MESFETs), metal oxide semiconductor field-effect transistors (MOSFETs), deep ultraviolet detectors, Schottky barrier diodes, etc. [1,2,3,4]. In many other technological applications, it is required to have materials to serve as a dielectric with band gap energy even larger than Ga2O3 to form heterostructures for carrier confinement and to enable optoelectronics deeper into the UV region. The Al2O3 with a band gap of 8.82 eV in the corundum (sapphire) phase is considered an excellent candidate for alloying it with Ga2O3 to create novel β-(AlxGa1−x)2O3. The β-(AlxGa1−x)2O3 has a smaller bond length, which makes the atomic gap smaller and the band gap larger. By carefully increasing Al contents, the band gap of β-(AlxGa1−x)2O3 has the potential to increase from 4.8 eV to 6 eV. This material with higher critical field strength is considered suitable for power electronics devices and shorter cut-off wavelength for optoelectronic devices [5,6,7]. In a recent study [8], the drain current of (Al0.08Ga0.92)2O3/Ga2O3 MODFET (with a gate voltage of +3 V) can be 12 mA mm−1, and the on/off current ratio is about 104. The breakdown voltage of the (Al0.16Ga0.84)2O3-channel MESFETs (with a drain-to-gate spacing of 20 μm) can reach 940 V, which is higher than that of β- Ga2O3. In addition, β-(AlxGa1−x)2O3 can be used as a barrier layer for β-(AlxGa1−x)2O3/β-Ga2O3 HEMTs. The theoretical value of 2DEG mobility for β-(AlxGa1−x)2O3/β-Ga2O3 interface can reach 500 cm2/V s, which is much higher than that of the bulk β-Ga2O3 [9]. It means that (AlxGa1−x)2O3 has great potential for high-power applications.
In early research, it was found that the internal changes of materials often change their physical properties, thus affecting their application in devices [10,11]. In the process of alloying Ga2O3 with Al2O3, the microstructure and physical properties of the crystal are modified due to the change of atomic radius, lattice structure, and the formation of new compositions and defects during the crystal growth. In an earlier theoretical study, Peelaers et al. [12] have shown that the pseudo-cubic lattice constant of β-(AlxGa1−x)2O3 alloys decrease linearly by increasing Al composition x, while the bandgap increases exhibiting significant bowing. Moreover, it has been suggested that Al atoms preferentially occupy the octahedral positions to form the monoclinic structure for x < 0.71. While studying the electronic and optical properties, Ma et al. [13] have noticed that Al doping in β-Ga2O3 leads to the formation of O interstitial defects. Kim et al. [14] have suggested that the oxygen vacancies and grain boundary-related defects increase in (AlxGa1−x)2O3 with Al composition exceeding 0.6 at %. Bhuiyan et al. [15] have studied the phase stability and phase transition of (AlxGa1−x)2O3 films grown by MOCVD in the wide Al composition range. The results have shown that the strain caused by the doping of Al induces rotation of the β phase (AlxGa1-x)2O3 crystal phase domain and promotes the formation of γ phase (AlxGa1−x)2O3. Hilfiker et al. [16] reported the dielectric function tensor of β-(AlxGa1−x)2O3 (x ≤ 0.21) films exhibiting strong anisotropy and inter-band transitions in the energy range of 1.5–9.0 eV. As the Al concentration increases, the band gap of the thin films increases, and the anisotropy increases.
In our previous work [17], we have reported crystal structure distortions by studying the temperature dependence of Raman shifts in β-(AlxGa1−x)2O3 single crystals with x = 0, 0.1, 0.2. Certainly, with distortion, the lattice vibrations and anisotropy of the crystal will change. However, the changes in the long-range order and the anisotropy of lattice vibration caused by alloying with aluminum in β-(AlxGa1−x)2O3 are still unclear. The purpose of this work is to study the spatial correlation of Raman phonon modes in β-Ga2O3 single crystal and its alloyed β-(AlxGa1−x)2O3 single crystals with different Al compositions, x (≡0.0, 0.06, 0.11, 0.17, 0.26) by using a spatial correlation model. The long-range crystallographic order within the crystal and the symmetry of its sublattice structure are reflected by the correlation length (CL). In addition, the angle-resolved polarized Raman spectroscopy (ARPRS) was carried out to observe the angle dependence of β-(AlxGa1−x)2O3 alloys in different polarization configurations. Combining with the Raman selection rules, we have analyzed the Raman tensor elements of different alloyed samples and revealed the intrinsic anisotropy of differential polarizability.

2. Material Growth and Characterization Methods

The five β-(AlxGa1−x)2O3 single-crystal samples are grown by using the optical floating zone method, with aluminum composition x < 0.3. Energy Dispersive Spectroscopy (EDS) (XMax20/C-Nano, Sigma-500 Zeiss, Oberkochen, Germany) is used to analyze the composition of aluminum fraction in these samples. The total quantity error of EDS is ≤3%. The surface scanning mode has been set with a scanning range of 2 mm × 2 mm. The samples are carefully analyzed using a high-resolution x-ray diffraction (HR-XRD) spectroscopy (model SMARTLAB 3kW, Rigaku, Tokyo, Japan) by employing a Cu target excitation source for assessing their crystalline quality. The accuracy of XRD is 0.01°.
Raman scattering spectroscopy measurements are conducted by using the Zolix Raman microscopic spectrograph (model Finder One), which is employed by using a 532 nm wavelength laser as an excitation source. The test precision of Raman spectrometer is ≤2 cm−1. The temperature-dependent studies between 80 K and 800 K are recorded with a Linkam INSTEC temperature-changing device. The temperature error is ±0.1 K. In the original Raman scattering measurement system, a polarizer is placed in front of the detector. By rotating the b-axis direction of the sample, the angular dependence from 0° to 360° is tested with an angle step of 5°. The angle error bar in ARPRS is ±0.05°. The experimental configuration is shown in Figure 1. The parallel-polarization and cross-polarization configurations are realized by adjusting the polarizer angles between 0° and 90° in the optical detecting path. The sample placement follows the notion that the polished surface of the sample is perpendicular to the direction of light propagation.

3. Results and Discussion

3.1. Crystal Structure and Composition

The results for the β-(AlxGa1−x)2O3 samples analyzed qualitatively and quantitatively by using EDS are displayed in Figure S1. For every sample, the calculated atomic percentage reported in Table 1 is conducted by considering the peak area ratio of each feature. The Al/(Al + Ga) molar ratios in five samples are 0.0, 0.06, 0.11, 0.17, and 0.26, respectively. For convenience, the β-Ga2O3, β-(Al0.06Ga0.94)2O3, β-(A0.11Ga0.89)2O3, β-(Al0.17Ga0.83)2O3, and β-(Al0.26Ga0.74)2O3 samples are marked as S0, S1, S2, S3, and S4, respectively. The results of optical transmission spectra are displayed in Figure S2 for the five samples. The bandgaps obtained for samples S0, S1, S2, S3, and S4 by linear extrapolation using Tauc’s method are 4.708, 4.761, 4.885, 4.982, and 5.63 eV, respectively (Figure S2).
The β-Ga2O3 single crystal has a monoclinic structure formed by the stacking of two GaO4 tetrahedra and two edge-sharing GaO6 octahedra, as shown in Figure 2a,b. The XRD spectroscopy results displayed in Figure 2c have clearly shown four diffraction peaks corresponding to the (400), (600), (800), and (1200) crystal planes of β-Ga2O3, referring to the β-Ga2O3 XRD JCPDS card (04-003-1858). The five samples have exhibited well-defined monoclinic structures. From the Bragg’s Equation (1)
2 d sin θ = n λ ,   ( n = 1 ,   2 ,   3 ) ,
it is possible to obtain the variation trends in the interplanar crystal spacing d with the aluminum composition, x. From Figure 2d,e, one may notice that by increasing x, the interplanar crystal spacing decreases while the FWHM broadens. As the volume size of Al atoms is smaller than that of Ga, the average bond length of Al-O is, therefore, shorter than Ga-O bonds. The replacement of Ga by Al atoms caused the (AlxGa1−x)2O3 unit cell volume to decrease [18,19]. This has resulted in the decrease of interplanar crystal spacing (d). Linear fitting of the interplanar spacing has shown that the degree of change in d with aluminum composition varied differently for different crystal planes. The absolute values of the slopes (|k|) obtained by fitting the diffraction peaks of the (400), (600), (800), and (1200) planes were 0.14, 0.09, 0.06, and 0.04, respectively. With the increase of Miller index, the influence of aluminum composition on the interplanar crystal spacing gradually decreases.

3.2. Raman Scattering Spectroscopy

The room-temperature (RT) Raman scattering spectra of the five samples, excited by a 532 nm laser (perpendicular to the (100) plane) source, are shown in Figure 3. All five samples, S0, S1, S2, S3, and S4, have revealed 10 Raman active phonon features ( B g 2 , A g 2 , A g 3 , A g 4 , A g 5 , A g 6 , A g 7 , A g 8 , A g 9 , and A g 10 ) in the frequency range of 100~1000 cm−1. The results agreed very well with those reported in the literature [20,21,22]. One must note that the modes belonging to the B g 2 , A g 2 , and A g 3 are associated with the vibration and translation of the [GaO6]-[GaO4] chains; the A g 4 , A g 19 , A g 6 , and A g 7 modes are related to the deformation of the [GaO6] octahedron while A g 7 , A g 8 , and A g 10 phonons are linked to the symmetric stretching and bending vibrational modes of the [GaO4] tetrahedron [17,23]. The perusal of Figure 3a reveals that with an increase in Al composition, the Raman intensity of each band has weakened. The shape of every peak in the mid-frequency region is distorted due to alloying, and the features have no longer behaved as the normal Lorentz types. The XRD results indicate the lattice distortion is induced by the substitution of Al for the Ga atoms. This reduced the crystal lattice symmetry by weakening the bond strength and caused the Raman active modes to shift with declining intensity [24].
The Raman shifts and full width at half maximum (FWHM) for (AlxGa1−x)2O3 samples are displayed in Figure 4 as a function of Al composition, x. From Figure 4a, we have noticed that increasing x caused a blue shift to the Raman phonon peaks. The linear fit of Raman shifts as a function of composition indicated the slopes of B g 2 , A g 2 , A g 3 , A g 8 , A g 9 , and A g 10 modes as 48, 29, 64, 73, 99, and 77, respectively. The slope values corroborate those reported by Kranert et al. [21]. Earlier, Janzen et al. [20] calculated the contributions of different atoms to the phonon energies. By comparison, we found that the modes with a larger influence of O atoms on the total phonon energy exhibit larger k values. This is related to the force constants between the molecular bonds. As the average bond length of Al-O is shorter than those of the Ga-O bonds, alloying with Al would, therefore, cause the force constants between Al-O bonds to strengthen [18,25]. The decrease of the atomic mass and increase in the force constant are the main reasons for the observed blue shifts to the Raman peaks in the β-(AlxGa1−x)2O3 alloys [23,26]. The mass change comes from the difference in masses of the substituted Al for Ga atoms, while the change in force constant is caused by the Al-O bonds generated after the substitution of Al atoms. The larger contribution of O means higher involvement of the Al-O bonds. Obviously, the results have indicated the contribution of force constant change between Al-O bonds to the blue shift of Raman peaks is stronger than the mass change of the atoms. Interestingly, the contribution of O atoms to the B g 2 mode is similar to those of the A g 2 mode: the slope k of the B g 2 mode is nearly twice that of the A g 2 mode. This may be related to the fact that the Ag mode oscillates in the b-plane while the Bg mode vibrates perpendicular to the b-plane [26]. In our work, the light path is perpendicular to the (100) plane, and under the same comparative configuration, the component of the vibrational amplitude of Bg mode in the (100) plane is larger. In addition, it is also noticed that the Al composition x has a smaller effect on the low-frequency modes as compared to the high-frequency phonons. From the earlier studies [23,24], the low-frequency modes are suggested as mainly originating from the vibrations of the tetrahedral-octahedral chains, while the high-frequency modes involved the atomic vibrations of the tetrahedral units. As compared to the vibrations of the chains, the impact of doping on the phonon energy is more obvious in the modes related to the tetrahedral units.
Again, the perusal of Figure 3 and Figure 4b revealed that the FWHM of Raman peaks for different modes are broadened with the increase of Al composition, x, especially in the mid-frequency range, where the distortions occurred at varying degrees, and the main Lorentzian shapes are lost. This may be related to the selective substitution of Al atoms, which preferably substitute for the octahedral Ga atoms as compared to the tetrahedral Ga atoms [21,27]. Unlike the trends of blue shift, the FWHM (cf. Figure 4b) reflects the disorder in the observed phonons. It can be noticed that the low-frequency phonons are more sensitive to the Al doping concentration, with the turning point appearing near x = 0.17. After that, the FWHM begins to deviate from the linear relationship with composition, x. This may be related to the phonon vibration. When the doping amount is sufficient, the atomic substitution reaches a certain degree, and the disorder of phonons increases, which may lead to the formation of new phonon modes. Especially the mid-frequency phonons. This is mentioned in the study of β-(AlxGa1−x)2O3 ceramics by Bhattacharjee et al. [26]. The high-frequency region is only related to the vibration of the tetrahedron, so there is no turning point at x = 0.17. The low-frequency phonon is related to the vibration of the chains, so the FWHM broadening is also affected by the mid-frequency phonons, which is out of the linear relationship. As Compared to the pure β-Ga2O3 sample, when Al composition reaches close to x = 0.26 in β-(AlxGa1−x)2O3 alloys, the FWHM of B g 2 , A g 2 , and A g 3 modes broaden (nearly 5 to 10 times) with A g 3 mode showing the most significant changes. The A g 8 , A g 9 , and A g 10 modes, however, broaden only 2–3 times at x = 0.26. The B g 2 , A g 2 , and A g 3 modes involve different types of atoms; therefore, they cause a greater degree of phonon disorder and lead to a significant broadening of the FWHM.

3.3. Spatial Correlation of Raman Phonons

Again, the spatial correlation model of Raman scattering spectra can be used to analyze the effects of Al composition, x, on the lattice vibrational modes order of the alloys [28,29,30,31]. As the aluminum composition is increased in our β-(AlxGa1−x)2O3 samples, each Raman peak produces different degrees of asymmetric broadening and causes a blue shift. In a perfect crystal with translational symmetry, the spatial correlation function of phonons is infinite, which follows the selection rule of q = 0. However, due to the lattice distortion and defects in β-(AlxGa1−x)2O3 alloys, the disorder of microstructures increases, the spatial correlation function of the phonon becomes limited, and the q = 0 selection rule is destroyed [30]. According to the CL of the phonons, the perfection of crystallization and disorder of lattice vibration can be analyzed quantitatively.
It is assumed that there exists a spherical region associated with a finite size of the correlation region in the alloy material. Combining with a Gaussian distribution function, the Raman intensity I(ω) at a frequency ω can be expressed as the following [32,33,34]:
I ( ω ) 0 1 exp ( q 2 L 2 4 ) d 3 q [ ω ω q ] 2 + ( Γ 0 2 ) 2
Here, q (≡2π/a) represents the magnitude of the normalized wavevector, where a is the lattice constant of the sample. The term L in Equation (2) is the CL of Raman mode which represents the range of ordered phonons described by the correlation function. It is commonly used to explain the q-vector relaxation related to finite-size effects and structural disorder. The symbol Γ0 is the FWHM of the Raman peak; ω represents the angular frequency of phonon vibration; ω(q) is the dispersion relation of optical phonons, which can be expressed as the following [35,36]:
ω q = A B q 2 ,
where A is the phonon frequency at the center of the Brillouin, and B is the correction parameter for the peak position shift of the Raman peak caused by the surrounding environment.
Figure 5 exhibits the relationship between the CLs of the phonon modes and the aluminum composition in the low and high-frequency regions. It can be observed that the CLs decrease with the increase of Al composition, indicating a decrease in the orderliness of the lattice vibrations due to the Al doping. The decrease in orderliness becomes more prominent with the increase of Al composition. The relationship between the CL and the Al composition is linearly fitted. The absolute values of the slope (|k|) of the B g 2 , A g 2 , A g 3 , A g 8 , A g 9 , and A g 10 modes are 0.42, 0.70, 0.66, 0.24, 0.30, and 0.32, respectively. The modes in the low-frequency region are more sensitive to the Al composition than those of the high-frequency modes. The value |k| of B g 2 in the low-frequency region is significantly smaller than that of the A g 2 and A g 3 modes, while in the high-frequency region, the |k| values of A g 8 , A g 9 , and A g 10 modes vary significantly. Compared to the chain structures involving multiple sites, the phonons in the high-frequency region are only affected by the atomic substitution in the sublattice related to their vibration resulting in relatively smaller changes in the spatial correlation.
Figure 6a,b shows the variable-temperature (VT) Raman spectra of S0 and S4. The VT Raman spectra of S1, S2, and S3 are shown in Figure S3. The values of CLs for A g 3 (low frequency) and A g 10 (high frequency) modes at different temperatures are obtained by spatial correlation model fitting to analyze the effects of temperature on the lattice vibration disorder. From Figure 6, one may note that the CLs of the two Raman modes decrease with increasing temperature. As the temperature increases, the disorder in the lattice vibration increases. This can be explained by the fact that as the temperature increases, the thermal motion of the molecule increases, accompanied by lattice expansion [24,37], resulting in enhanced vibrational disorder and shortening the phonon length. With the increase of Al composition, x, the spatial CL range decreases accordingly. This is consistent with the changing trend of the calculated CLs calculated (cf. Figure 6) at room temperature. From the slope of the absolute value(|k|) obtained by linear fitting, it can be noticed that with the increase of Al composition, x, the degree of sublattice order involved in A g 3 and A g 10 modes is less affected by temperature. Among them, the difference between the pure β-Ga2O3 and β-(AlxGa1−x)2O3 alloy samples is particularly obvious. The value of |k| for A g 3 decreases from 6.72 × 10−3 for the undoped sample to 3.11 × 10−3 for a sample with Al composition x = 0.17 and then begins to fluctuate slightly. The value of k for A g 10 decreases from 9.90 × 10−3 for the undoped sample to 6.40 × 10−3 with an Al composition of x = 0.11 and then begins to fluctuate slightly. This difference is attributed to the fact that the crystal distortion caused by Al doping is partly restored with increasing temperature [17], and the temperature dependence tends to stabilize after reaching a certain doping level. For bulk β-(AlxGa1−x)2O3, doping leads to changes in the microstructure of the material, and the crystal quality of the sample will seriously affect the performance of the device. This paper reveals the relationship between crystal quality, microstructure, and doping concentration in β-(AlxGa1−x)2O3, which provides a reference for the application of β-(AlxGa1−x)2O3 alloy in devices.

3.4. In-Plane Anisotropy

Angle-resolved polarized Raman spectroscopy (ARPRS) is utilized to investigate the anisotropic properties of various alloyed samples [38,39,40]. Zhang et al. [41] have successfully used ARPRS to calculate the Raman tensor of β-Ga2O3, revealing the complexity of the anisotropy of the internal structure of β-Ga2O3. However, the research on the phonon anisotropy of β-(AlxGa1−x)2O3 has not been reported. In this work, Raman spectra are acquired at different rotation angles of the samples under parallel/cross polarization configurations, where the polarization of the Raman scattered light was parallel/cross to the polarization of the incident light. By rotating the sample, the angle θ between the polarization direction of the incident light and the b-axis of the sample was altered. The experimental setup is shown in Figure 1. The generalized form of the scattering intensity (I) of the ARPRS signal can be mathematically expressed as the following [42,43]:
I A g e ^ i · R · e ^ s 2
Here, êi and ês denote the unit polarization vectors of the incident and scattered light, respectively. As the light is incident perpendicular to the sample surface, the incident and scattered light can be represented as êi = (sinθ cosθ 0), ê s = (sinθ cosθ 0), and ê s = (cosθ − sinθ 0). R represents the Raman tensor in the scattering process, and the Raman tensor elements for the Ag and Bg modes of gallium oxide are given as follows [41,44]:
R A g = b 0 d 0 c 0 d 0 a
R ( B g ) = 0 f 0 f 0 e 0 e 0
Here, a, b, c, d, e, and f are the independent Raman tensor elements. To interpret the experimental results more accurately, it is necessary to convert R into a complex second-rank Raman tensor. Therefore, the Raman tensor for β-Ga2O3 can be expressed as follows:
R A g = b e i φ b 0 d e i φ d 0 c e i φ c 0 d e i φ d 0 a e i φ a
R ( B g ) = 0 | f | e i φ f 0 | f | e i φ f 0 | e | e i φ e 0 | e | e i φ e 0
The phase information of the tensor components is represented by φa~φf. The angle dependence of the polarized Raman signalcan be described by Equation (4). Substituting the defined êi and ês, as well as the complex Raman tensor (Equations (7) and (8)), into Equation (4), the ARPRS intensity expression for the plane (100) can be obtained as follows:
I A g a 2 s i n 4 θ + c 2 c o s 4 θ + 2 | a | | c | s i n 2 θ c o s 2 θ c o s φ a c
I A g s i n 2 θ c o s 2 θ · ( a 2 + c 2 2 a c c o s φ a c )
I B g e · sin 2 θ 2
I B g e · cos 2 θ 2
where φ is the phase difference between the two distinct elements of the Raman tensor.
Figure 7. Waterfall plot of the angle-resolved polarized Raman spectra for S0, S1, S2, S3, and S4 measured in parallel-polarized and cross-polarized configurations. The green line corresponds to parallel polarization, and the red line corresponds to cross-polarization.
Figure 7. Waterfall plot of the angle-resolved polarized Raman spectra for S0, S1, S2, S3, and S4 measured in parallel-polarized and cross-polarized configurations. The green line corresponds to parallel polarization, and the red line corresponds to cross-polarization.
Materials 16 04269 g007
Figure 8. Polar diagram of ARPRS for A g 2 , A g 3 , and A g 10 phonons. Solid curves correspond to the best fit of the data using Equations (9) and (10) for each configuration and mode symmetry. The green line corresponds to parallel polarization, and the red line corresponds to cross-polarization.
Figure 8. Polar diagram of ARPRS for A g 2 , A g 3 , and A g 10 phonons. Solid curves correspond to the best fit of the data using Equations (9) and (10) for each configuration and mode symmetry. The green line corresponds to parallel polarization, and the red line corresponds to cross-polarization.
Materials 16 04269 g008
Based on the ARPRS results shown in Figure 7, distinct Raman phonon modes display different intensity variations with a periodic trend (similar to a sinusoidal wave). Voigt function is used to fit the B g 2 , A g 2 , and A g 3 peaks and investigate their polarization dependence. The Polar diagram of Raman peak intensity as a function of the angle θ in Figure 8 exhibits clear polarization dependence. In the parallel-polarization configuration, the three modes exhibit a 2-lobed shape, where the A g 2 and A g 3 modes have two main lobes and two subsidiary lobes. Their major axes are oriented at 0° (180°), and the relative intensities of the subsidiary lobes decrease with increasing Al content. A g 10 mode shows two dominant lobes with a main axis at 90° (270°), perpendicular to the main axes of the other two modes. In the cross-polarization configuration, all modes exhibit an almost equally sized 4-lobed shape, oriented at a 45° angle to the pattern observed under parallel polarization.
To figure out the differential polarization degree of the internal structure of the aluminum alloy sample, Equations (9) and (10) are used to fit the polar coordinates of each mode. The fitting results are shown by the green and red lines in Figure 8. The fitting curve is consistent with the experimental data. For some messy data, such as the A g 2 and A g 3 modes of S2, this may be related to the roughness of the sample surface. The (100) plane is a cleavage plane in β-Ga2O3 [24], and some cracks will occur on the surface and inside of the sample. When the test area contains cracks, the test results will be confused. Because of the lack of the A g 2 mode of S4, the vertical polarization data cannot be effectively fitted due to the serious peak distortion. The fitted parameters are listed in Table 2. The anisotropic properties of β-(AlxGa1−x)2O3 alloy samples on the (100) plane can be analyzed through the ratio of Raman tensor elements |a/c| and the phase difference φ. |a/c| less than 1 or greater than 1 determines whether the main axis is along the 0°/180° or 90°/270° direction, respectively. The value of |a/c| reflects the anisotropy of the Raman tensor for the Ag mode, and the closer it is to 1, the weaker the anisotropy. According to Table 2, it can be found that the Raman tensors of the A g 2 , A g 3 , and A g 10 modes in β-(AlxGa1−x)2O3 all exhibit significant anisotropy, but their trends with aluminum composition are different. For the pure β-Ga2O3 sample, the |a/c| values of the A g 2 , A g 3 , and A g 10 modes are 0.922, 0.503, and 14.740, respectively. It corroborates that reported by Zhang et al. [41]. Relatively, A g 10 exhibits very strong anisotropy, while the |a/c| value of the A g 2 mode is close to 1, which suggests that the anisotropy is not significant. The anisotropy of the Raman phonon modes of β-Ga2O3 is strongly influenced by aluminum alloying. Obviously, the anisotropy of A g 2 and A g 3 increases with the increase of aluminum content, while the incorporation of aluminum weakens the anisotropy of A g 10 . This may be caused by the large difference in the bond polarizability. The β-angle increases with the increase of Al content [26], leading to the increased asymmetry of the β-(AlxGa1−x)2O3 unit cell. Consequently, the anisotropic properties of the phonons in the low-frequency region associated with chain vibrations are strengthened. The A g 2 phonon vibrates perpendicular to the ab plane [20,24], and its anisotropy is not significant when observed on the plane (100). The A g 3 phonon is mainly affected by chain oscillation, and the vibration direction of the involved atoms is more diverse, resulting in a more significant anisotropy. The A g 10 mode involves only the stretching of the two Ga-O bonds, resulting in a very strong anisotropy on the (100) plane. The average bond length of Al-O is shorter than that of Ga-O. When Al atoms replace Ga atoms, the polarization of the original Ga-O bonds in the (100) plane is decreased, and their anisotropy is weakened. Furthermore, Figure 7 presents an interesting phenomenon, where one can see that the relative intensity of the A g 10 peak concerning the low- and medium-frequency Raman peaks undergoes significant changes with the variation of the aluminum composition, x. By taking A g 3 as a reference, the strongest peak intensity of the A g 3 and A g 10 in all angles are selected for comparison in each data group, as shown in Table 3. It can be found that the intensity of A g 10 is significantly higher than that of the A g 3 after doping in both parallel and cross configurations, and the relative ratio of A g 10 / A g 3 peak intensity becomes more and more obvious with the increase of aluminum composition. This is difficult to observe in a normal Raman spectroscopy. The ARPRS of β-(AlxGa1-x)2O3 alloy shows significant anisotropic dependence on the polarization of the incident light. The anisotropic properties of gallium oxide cannot be ignored in different phonon modes on the plane (100) as they exhibit different behaviors in the ratio of Raman tensor elements and phase differences. The anisotropic properties make β-(AlxGa1−x)2O3 have different properties and application potential in different directions. By studying its anisotropic properties, researchers can better design and optimize β-(AlxGa1−x)2O3 materials and their devices to achieve specific application requirements. Studying the phonon anisotropy of β-(AlxGa1−x)2O3 is helpful to understand the local changes in the microstructure structure of the material and is conducive to the development of potential applications of β-(AlxGa1−x)2O3 in new optoelectronic devices.

4. Conclusions

In summary, with the increase of aluminum composition in β-(AlxGa1−x)2O3, the observed Raman peaks have displayed a blue shift, and the degree of blue shift is related to the types of phonon modes. For the Ag phonons, the greater the contribution of O atoms to the phonon energy, the greater the influence of Al composition on the mode. Moreover, the FWHMs of β-(AlxGa1−x)2O3 phonons broaden with the increase of aluminum composition, x, and the phonons in the low-frequency region are influenced more by the Al composition than those in the high-frequency region. After reaching the Al composition of x = 0.17, the FWHM and Al composition begin to break away from the linear relationship. Comprehensive Raman analysis by using a spatial correlation model has revealed that the CL decreases with the increase of aluminum composition, x, and the long-range order of the lattice vibration also decreases. The CL in the low-frequency region is affected more by the Al composition than in the high-frequency region. In the range of 80 K~800 K, the CL of each Raman phonon mode decreases with the increase in temperature, and the lattice disorder becomes stronger. The temperature dependence of CL decreases with the increase of Al composition and tends to become stable after Al composition increases to a certain value. The results of angle-resolved polarized Raman spectroscopy show that the different Raman peak intensities of β-(AlxGa1−x)2O3 become highly polarization dependent. From Raman tensors, different Raman phonon modes have exhibited significant anisotropy. As the aluminum content increases, the anisotropy of phonons in the low-frequency region increases, and the anisotropy of phonons in the high-frequency region decreases. In addition, it is found that the Al doping will lead to a change in the relative ratio of the intensity of the two strongest Raman peaks in the high-frequency region and the low-frequency region of the angle-resolved polarization Raman spectrum. With the increase of aluminum content, the ratio of A g 10 / A g 3 peak intensity increases. We strongly feel that this comprehensive study has provided meaningful results for understanding the long-range orderliness and anisotropies in technologically important β-(AlxGa1−x)2O3 crystals.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16124269/s1; Figure S1: The measurement results of EDS for S0, S1, S2, S3, and S4, respectively; Figure S2 (a): The optical transmission spectra of S0, S1, S2, S3, and S4, respectively; (b) The bandgaps of the five samples calculated by using the Tauc method; Figure S3: Variable-temperature Raman spectra of S1, S2, and S3, respectively.

Author Contributions

Conceptualization, L.L. and L.W.; experiment, L.L., C.X., Q.S., H.L., J.J., and P.L.; writing—original draft preparation, L.L.; writing—review and editing, L.W., D.N.T. and Z.C.F.; supervision, L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China, grant number 51972319, and the Science and Technology Commission of Shanghai Municipality, grant number 19520744400.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, D.; Guo, Q.; Chen, Z.; Wu, Z.; Li, P.; Tang, W. Review of Ga2O3-based optoelectronic devices. Mater. Today Phys. 2019, 11, 1240–1248. [Google Scholar] [CrossRef]
  2. Liao, Y.; Zhang, Z.; Gao, Z.; Qian, Q.; Hua, M. Tunable Properties of Novel Ga2O3 Monolayer for Electronic and Optoelectronic Applications. ACS Appl. Mater. Interfaces 2020, 12, 30659–30669. [Google Scholar] [CrossRef] [PubMed]
  3. Feng, Q.; Li, X.; Han, G.; Huang, L.; Li, F.; Tang, W.; Zhang, J.; Hao, Y. (AlGa)2O3 solar-blind photodetectors on sapphire with wider bandgap and improved responsivity. Opt. Mater. Express 2017, 7, 1240–1248. [Google Scholar] [CrossRef]
  4. Higashiwaki, M.; Sasaki, K.; Murakami, H.; Kumagai, Y.; Koukitu, A.; Kuramata, A.; Masui, T.; Yamakoshi, S. Recent progress in Ga2O3 power devices. Semicond. Sci. Technol. 2016, 31, 034001. [Google Scholar] [CrossRef]
  5. Li, J.; Chen, X.; Ma, T.; Cui, X.; Ren, F.-F.; Gu, S.; Zhang, R.; Zheng, Y.; Ringer, S.P.; Fu, L.; et al. Identification and modulation of electronic band structures of single-phase β-(AlxGa1-x)2O3 alloys grown by laser molecular beam epitaxy. Appl. Phys. Lett. 2018, 113, 041901. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, T.; Li, W.; Ni, C.; Janotti, A. Band Gap and Band Offset of Ga2O3 and (AlxGa1-x)2O3 Alloys. Phys. Rev. Appl. 2018, 10, 011003. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, T.; Cheng, Q.; Li, Y.; Hu, Z.; Zhang, Y.; Ma, J.; Yao, Y.; Zuo, Y.; Feng, Q.; Zhang, Y.; et al. Heterogrowth of β-(AlxGa1-x)2O3 Thin Films on Sapphire Substrates. Cryst. Growth Des. 2022, 22, 3698–3707. [Google Scholar] [CrossRef]
  8. Okumura, H.; Kato, Y.; Oshima, T.; Palacios, T. Demonstration of lateral field-effect transistors using Sn-doped β-(AlGa)2O3 (010). Jpn. J. Appl. Phys. 2019, 58, SBBD12. [Google Scholar] [CrossRef]
  9. Ghosh, K.; Singisetti, U. Electron mobility in monoclinic β-Ga2O3—Effect of plasmon-phonon coupling, anisotropy, and confinement. J. Mater. Res. 2017, 32, 4142–4152. [Google Scholar] [CrossRef] [Green Version]
  10. Huang, S.-S.; Lopez, R.; Paul, S.; Neal, A.T.; Mou, S.; Houng, M.-P.; Li, J.V. β-Ga2O3 defect study by steady-state capacitance spectroscopy. Jpn. J. Appl. Phys. 2018, 57, 091101. [Google Scholar] [CrossRef]
  11. Neal, A.T.; Mou, S.; Lopez, R.; Li, J.V.; Thomson, D.B.; Chabak, K.D.; Jessen, G.H. Incomplete Ionization of a 110 meV Unintentional Donor in beta-Ga2O3 and its Effect on Power Devices. Sci. Rep. 2017, 7, 13218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Peelaers, H.; Varley, J.B.; Speck, J.S.; Van de Walle, C.G. Structural and electronic properties of Ga2O3-Al2O3 alloys. Appl. Phys. Lett. 2018, 112, 242101. [Google Scholar] [CrossRef]
  13. Ma, X.; Zhang, Y.; Dong, L.; Jia, R. First-principles calculations of electronic and optical properties of aluminum-doped β-Ga2O3 with intrinsic defects. Results Phys. 2017, 7, 1582–1589. [Google Scholar] [CrossRef]
  14. Kim, S.; Ryou, H.; Lee, I.G.; Shin, M.; Cho, B.J.; Hwang, W.S. Impact of Al doping on a hydrothermally synthesized β-Ga2O3 nanostructure for photocatalysis applications. RSC Adv. 2021, 11, 7338–7346. [Google Scholar] [CrossRef] [PubMed]
  15. Bhuiyan, A.F.M.A.U.; Feng, Z.; Johnson, J.M.; Huang, H.-L.; Sarker, J.; Zhu, M.; Karim, M.R.; Mazumder, B.; Hwang, J.; Zhao, H. Phase transformation in MOCVD growth of (AlxGa1-x)2O3 thin films. APL Mater. 2020, 8, 031104. [Google Scholar] [CrossRef] [Green Version]
  16. Hilfiker, M.; Kilic, U.; Mock, A.; Darakchieva, V.; Knight, S.; Korlacki, R.; Mauze, A.; Zhang, Y.; Speck, J.; Schubert, M. Dielectric function tensor (1.5 eV to 9.0 eV), anisotropy, and band to band transitions of monoclinic β-(AlxGa1-x)2O3 (x ≤ 0.21) films. Appl. Phys. Lett. 2019, 114, 231901. [Google Scholar] [CrossRef]
  17. Chen, X.; Niu, W.; Wan, L.; Xia, C.; Cui, H.; Talwar, D.N.; Feng, Z.C. Microstructure and temperature-dependence of Raman scattering properties of β-(AlxGa1-x)2O3 crystals. Superlattices Microstruct. 2020, 140, 106469. [Google Scholar] [CrossRef]
  18. Kim, S.; Ryou, H.; Moon, J.; Lee, I.G.; Hwang, W.S. Codoping of Al and In atoms in β-Ga2O3 semiconductors. J. Alloys Compd. 2023, 931, 167502. [Google Scholar] [CrossRef]
  19. Sabino, F.P.; de Oliveira, L.N.; Da Silva, J.L.F. Role of atomic radius and d-states hybridization in the stability of the crystal structure of M2O3(M = Al, Ga, In) oxides. Phys. Rev. B 2014, 90, 155206. [Google Scholar] [CrossRef]
  20. Janzen, B.M.; Mazzolini, P.; Gillen, R.; Falkenstein, A.; Martin, M.; Tornatzky, H.; Maultzsch, J.; Bierwagen, O.; Wagner, M.R. Isotopic study of Raman active phonon modes in β-Ga2O3. J. Mater. Chem. C 2021, 9, 2311–2320. [Google Scholar] [CrossRef]
  21. Kranert, C.; Jenderka, M.; Lenzner, J.; Lorenz, M.; von Wenckstern, H.; Schmidt-Grund, R.; Grundmann, M. Lattice parameters and Raman-active phonon modes of β-(AlxGa1-x)2O3. J. Appl. Phys. 2015, 117, 125703. [Google Scholar] [CrossRef] [Green Version]
  22. Janzen, B.M.; Gillen, R.; Galazka, Z.; Maultzsch, J.; Wagner, M.R. First- and second-order Raman spectroscopy of monoclinic β−Ga2O3. Phys. Rev. Mater. 2022, 6, 054601. [Google Scholar] [CrossRef]
  23. Schmidt, C.; Zahn, D.R.T. Effect of impurities on the Raman spectra of spray-coated β-Ga2O3 thin films. J. Vac. Sci. Technol. A 2022, 40, 043404. [Google Scholar] [CrossRef]
  24. Dohy, D.; Lucazeau, G.; Revcolevschi, A. Raman spectra and valence force field of single-crystalline β-Ga2O3. J. Solid State Chem. 1982, 45, 180–192. [Google Scholar] [CrossRef]
  25. Bhattacharjee, J.; Sagdeo, A.; Singh, S.D. Determination of Al occupancy and local structure for β-(AlxGa1-x)2O3 alloys across nearly full composition range from Rietveld analysis. Appl. Phys. Lett. 2022, 120, 262101. [Google Scholar] [CrossRef]
  26. Bhattacharjee, J.; Singh, S.D. Observation of mixed-mode behavior of Raman active phonon modes for β-(AlxGa1-x)2O3 alloys. Appl. Phys. Lett. 2023, 122, 112101. [Google Scholar] [CrossRef]
  27. Krueger, B.W.; Dandeneau, C.S.; Nelson, E.M.; Dunham, S.T.; Ohuchi, F.S.; Olmstead, M.A.; Jones, J. Variation of Band Gap and Lattice Parameters of β-(AlxGa1-x)2O3 Powder Produced by Solution Combustion Synthesis. J. Am. Ceram. Soc. 2016, 99, 2467–2473. [Google Scholar] [CrossRef]
  28. Khan, T.M. Into the nature of Pd-dopant induced local phonon modes and associated disorders in ZnO; based on spatial correlation model. Mater. Chem. Phys. 2015, 153, 248–255. [Google Scholar] [CrossRef]
  29. Silva, R.L.d.S.e.; Franco Jr, A. Raman spectroscopy study of structural disorder degree of ZnO ceramics. Mater. Sci. Semicond. Process. 2020, 119, 105227. [Google Scholar] [CrossRef]
  30. Parayanthal, P.; Pollak, F.H. Raman Scattering in Alloy Semiconductors: “Spatial Correlation” Model. Phys. Rev. Lett. 1984, 52, 1822–1825. [Google Scholar] [CrossRef]
  31. Dev, G.S.; Sharma, V.; Singh, A.; Baghel, V.S.; Yanagida, M.; Nagataki, A.; Tripathi, N. Raman spectroscopic study of ZnO/NiO nanocomposites based on spatial correlation model. RSC Adv. 2019, 9, 26956–26960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wang, J.; Yao, W.H.; Wang, J.B.; Lu, H.Q.; Sun, H.H.; Wang, X.; Pang, Z.L. Quality of ZnSe/GaAs epilayers studied by spatial correlation model of Raman scattering. Appl. Phys. Lett. 1993, 62, 2845–2847. [Google Scholar] [CrossRef]
  33. Tiong, K.K.; Amirtharaj, P.M.; Pollak, F.H.; Aspnes, D.E. Effects of As+ ion implantation on the Raman spectra of GaAs: ‘‘Spatial correlation’’ interpretation. Appl. Phys. Lett. 1984, 44, 122–124. [Google Scholar] [CrossRef]
  34. Farid, S.; Stroscio, M.A.; Dutta, M. Multiphonon Raman scattering and photoluminescence studies of CdS nanocrystals grown by thermal evaporation. Superlattices Microstruct. 2018, 115, 204–209. [Google Scholar] [CrossRef]
  35. Feng, Z.C. Micro-Raman scattering and microphotoluminescence of GaN thin films grown on sapphire by metal-organic chemical vapor deposition. Opt. Eng. 2002, 41, 2022–2031. [Google Scholar] [CrossRef]
  36. Liu, Y.B.; Yang, J.Y.; Xin, G.M.; Liu, L.H.; Csanyi, G.; Cao, B.Y. Machine learning interatomic potential developed for molecular simulations on thermal properties of β-Ga2O3. J. Chem. Phys. 2020, 153, 144501. [Google Scholar] [CrossRef]
  37. Zhang, K.; Xu, Z.; Zhao, J.; Wang, H.; Hao, J.; Zhang, S.; Cheng, H.; Dong, B. Temperature-dependent Raman and photoluminescence of β-Ga2O3 doped with shallow donors and deep acceptors impurities. J. Alloys Compd. 2021, 881, 160665. [Google Scholar] [CrossRef]
  38. Jin, M.; Zheng, W.; Ding, Y.; Zhu, Y.; Wang, W.; Huang, F. Raman Tensor of van der Waals MoSe2. J. Phys. Chem. Lett. 2020, 11, 4311–4316. [Google Scholar] [CrossRef]
  39. Talochkin, A.B. Circularly polarized Raman scattering in silicon. J. Raman Spectrosc. 2019, 51, 201–206. [Google Scholar] [CrossRef]
  40. Shimasaki, M.; Nishihara, T.; Matsuda, K.; Endo, T.; Takaguchi, Y.; Liu, Z.; Miyata, Y.; Miyauchi, Y. Directional Exciton-Energy Transport in a Lateral Heteromonolayer of WSe2-MoSe2. ACS Nano 2022, 16, 8205–8212. [Google Scholar] [CrossRef]
  41. Zhang, K.; Xu, Z.; Zhao, J.; Wang, H.; Hao, J.; Zhang, S.; Cheng, H.; Dong, B. Anisotropies of angle-resolved polarized Raman response identifying in low miller index β-Ga2O3 single crystal. Appl. Surf. Sci. 2022, 581, 152426. [Google Scholar] [CrossRef]
  42. Ling, X.; Huang, S.; Hasdeo, E.H.; Liang, L.; Parkin, W.M.; Tatsumi, Y.; Nugraha, A.R.T.; Puretzky, A.A.; Das, P.M.; Sumpter, B.G.; et al. Anisotropic Electron-Photon and Electron-Phonon Interactions in Black Phosphorus. Nano Lett. 2016, 16, 2260–2267. [Google Scholar] [CrossRef] [PubMed]
  43. Pezzotti, G. Raman spectroscopy of piezoelectrics. J. Appl. Phys. 2013, 113, 211301. [Google Scholar] [CrossRef]
  44. Kim, S.; Lee, J.; Lee, C.; Ryu, S. Polarized Raman Spectra and Complex Raman Tensors of Antiferromagnetic Semiconductor CrPS4. J. Phys. Chem. C 2021, 125, 2691–2698. [Google Scholar] [CrossRef]
Figure 1. Schematic diagrams of angle-resolution Raman experimental configuration; êi and ês denote the unit polarization vectors of the incident and scattered light, respectively.
Figure 1. Schematic diagrams of angle-resolution Raman experimental configuration; êi and ês denote the unit polarization vectors of the incident and scattered light, respectively.
Materials 16 04269 g001
Figure 2. (a) The crystalline structure and (b) (100) plane of β-Ga2O3; (c) XRD patterns of S0, S1, S2, S3, and S4; (d) the interplanar crystal spacing; (e) FWHM for (AlxGa1−x)2O3 samples in dependence on the Al concentration. The variable k is the slope factor. R2 is the coefficient of determination that reflects the fitting effect. The closer the value is to 1, the better the model fitting degree is, and the smaller the error is.
Figure 2. (a) The crystalline structure and (b) (100) plane of β-Ga2O3; (c) XRD patterns of S0, S1, S2, S3, and S4; (d) the interplanar crystal spacing; (e) FWHM for (AlxGa1−x)2O3 samples in dependence on the Al concentration. The variable k is the slope factor. R2 is the coefficient of determination that reflects the fitting effect. The closer the value is to 1, the better the model fitting degree is, and the smaller the error is.
Materials 16 04269 g002
Figure 3. RT Raman scattering spectra of S0, S1, S2, S3, and S4.
Figure 3. RT Raman scattering spectra of S0, S1, S2, S3, and S4.
Materials 16 04269 g003
Figure 4. The Raman shift (a) and the FWHM (b) for (AlxGa1−x)2O3 samples versus Al composition.
Figure 4. The Raman shift (a) and the FWHM (b) for (AlxGa1−x)2O3 samples versus Al composition.
Materials 16 04269 g004
Figure 5. (a) The correlation length for B g 2 , A g 2 , A g 3 phonons, (b) A g 8 , A g 9 , and A g 10 phonons in dependence on the Al composition.
Figure 5. (a) The correlation length for B g 2 , A g 2 , A g 3 phonons, (b) A g 8 , A g 9 , and A g 10 phonons in dependence on the Al composition.
Materials 16 04269 g005
Figure 6. VT Raman scattering spectra of S0 (a) and S4 (b); temperature dependence of the correlation length for A g 3 (c) and A g 10 phonon (d).
Figure 6. VT Raman scattering spectra of S0 (a) and S4 (b); temperature dependence of the correlation length for A g 3 (c) and A g 10 phonon (d).
Materials 16 04269 g006
Table 1. EDS atomic percentages for S0, S1, S2, S3, and S4, respectively.
Table 1. EDS atomic percentages for S0, S1, S2, S3, and S4, respectively.
S0 (At%)S1 (At%)S2 (At%)S3 (At%)S4 (At%)
O K57.356.656.857.565.5
Al L0.12.64.77.28.9
Ga L42.440.838.535.325.5
Al/(Al + Ga)0.000.060.110.170.26
Table 2. Values of Raman tensor for the A g 2 , A g 3 , and A g 10 modes of β-(AlxGa1−x)2O3 in the parallel polarization configuration.
Table 2. Values of Raman tensor for the A g 2 , A g 3 , and A g 10 modes of β-(AlxGa1−x)2O3 in the parallel polarization configuration.
S0S1S2S3S4
|a/c|φacR2|a/c|φacR2|a/c|φacR2|a/c|φacR2|a/c|φacR2
A g 2 0.9220.7240.8730.9190.6320.9470.9150.5570.6200.8240.5260.8490.7320.5750.915
A g 3 0.5030.7540.9640.4590.9500.9900.4650.6120.8880.3730.8380.9750.3621.0180.994
A g 10 14.7400.1290.99614.4630.1850.99813.2560.4540.99012.9030.3270.99311.9950.1280.988
Table 3. The peak intensity ratio of A g 10 / A g 3 from Figure 7.
Table 3. The peak intensity ratio of A g 10 / A g 3 from Figure 7.
S0S1S2S3S4
parallel polarization2.1292.3222.4622.8153.147
cross-polarization1.3701.4301.6271.7612.488
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, L.; Wan, L.; Xia, C.; Sai, Q.; Talwar, D.N.; Feng, Z.C.; Liu, H.; Jiang, J.; Li, P. The Spatial Correlation and Anisotropy of β-(AlxGa1−x)2O3 Single Crystal. Materials 2023, 16, 4269. https://doi.org/10.3390/ma16124269

AMA Style

Li L, Wan L, Xia C, Sai Q, Talwar DN, Feng ZC, Liu H, Jiang J, Li P. The Spatial Correlation and Anisotropy of β-(AlxGa1−x)2O3 Single Crystal. Materials. 2023; 16(12):4269. https://doi.org/10.3390/ma16124269

Chicago/Turabian Style

Li, Liuyan, Lingyu Wan, Changtai Xia, Qinglin Sai, Devki N. Talwar, Zhe Chuan Feng, Haoyue Liu, Jiang Jiang, and Ping Li. 2023. "The Spatial Correlation and Anisotropy of β-(AlxGa1−x)2O3 Single Crystal" Materials 16, no. 12: 4269. https://doi.org/10.3390/ma16124269

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop