DOI:
10.1039/D5RA06475A
(Paper)
RSC Adv., 2025,
15, 38036-38043
A high-performance solar blind photodetector based on spinel gallium oxide thin film supported by aluminum doping
Received
29th August 2025
, Accepted 27th September 2025
First published on 13th October 2025
Abstract
Metastable γ-phase gallium oxide (Ga2O3)-based semiconductor materials have recently garnered much attention for developing high-performance wide-band-gap optoelectronic devices. However, the difficulty in the synthesis of single-phase crystals and the poor thermal stability of metastable γ-Ga2O3 pose great obstacles for its potential applications. In this work, we experimentally demonstrate the epitaxial growth of γ-Ga2O3 thin films at 750 °C on α-Al2O3 (0001) substrates upon aluminum (Al) doping. A solar blind photodetector based on an Al-doped γ-Ga2O3 thin film was successfully fabricated, and its properties were investigated in detail. Benefiting from its high crystal quality and absence of oxygen vacancies in Al-doped γ-Ga2O3, the solar blind photodetector exhibited a very low dark current, fast decay time, high detectivity, and excellent wavelength selectivity, along with high stability and reproducibility. Notably, the present device exhibited no obvious performance degradation when the light intensity and bias voltage were increased, indicating good self-heating dissipation, which is beneficial for future practical applications. The γ-Ga2O3 thin film-based photodetector might open up new possibilities to obtain high-performance thin film-based solar blind ultraviolet optoelectronic devices.
1. Introduction
Solar-blind ultraviolet (UV) detection with a cutoff wavelength shorter than 280 nm has a vast and growing number of military and civil surveillance applications, such as missile tracking, short-range secure communication, UV astronomy, ozone hole monitoring, and corona detection.1–5 Commercial solar blind photodetectors (PDs) are mainly based on vacuum photomultiplier tubes, which have obvious disadvantages, such as high operation power consumption, inflexibility, and difficulty in integration, thereby limiting their practical applications.6
Recently, solar blind PDs based on wide-bandgap semiconductors, such as AlGaN, MgZnO, diamond, and Ga2O3, have attracted great attention.7–10 However, high-quality epitaxial AlGaN films are difficult to be prepared due to their crystal deterioration with increasing Al content and the breakdown of the single wurtzite phase of MgZnO with increasing Mg content, and diamonds cannot be used to detect the entire solar-blind UV region because of their non-tunable bandgap.10,11 Ga2O3 with a direct bandgap of 4.9–5.3 eV lies sharply in the solar-blind spectrum region and also exhibits a flexible tunability in bandgap by alloying with different materials, making it one of the ideal candidates to fabricate solar-blind UV PDs.12–15 Ga2O3 is known to show five crystalline phases named α, β, γ, δ and ε.16 Among these phases, the most stable monoclinic β-Ga2O3 has been widely studied. The solar-blind PDs based on Ga2O3 are concentrated on β-Ga2O3 because of its stability and ease of preparation. However, β-Ga2O3 has a monoclinic structure, which makes it difficult to grow high-quality β-Ga2O3 on hetero-substrates. It is challenging to find material lattices matching with the β-Ga2O3 lattice to grow epitaxial thin films for optoelectronic applications. Thus, it is meaningful to explore the potential application of other metastable materials with a hexagonal/cubic structure, which have the same structure as those of Al2O3, ZnO, SiC, GaN and MgAl2O4.17–19 PDs based on other metastable phases have rarely been reported because of the difficulty in synthesizing single-phase crystals with thermal stability.
A recent first-principles study on the energetics of the Ga2O3 polymorphs suggests that the differences in free energy between the β-phase and other metastable phases are small; therefore, the polymorph of Ga2O3 is selected depending upon the preparation conditions.20–22 Moreover, doping is an important method to stabilize the crystal structure against thermal processes, making it a possible approach to obtain Ga2O3 with a thermally stabilized metastable phase and preventing the transition to the β-phase.23–25 Acting as a satisfying high-quality Ga2O3 on hetero-substrates for developing high-performance wide-band-gap optoelectronic devices, metastable γ-phase Ga2O3 is considered here. Hiroyuki et al. have demonstrated the fabrication of Mn-doped γ-Ga2O3 at a temperature of 500 °C on α-Al2O3 (0001) substrates using the PLD method, which differs from the undoped film that shows the β phase.26 Sam-Dong et al. have shown the enhancement of thermal stability of α-Ga2O3 films on sapphire substrates upon Al doping. The films can be grown at a temperature as high as 800 °C and can withstand an annealing temperature as high as 850 °C without forming the β phase.23 Furthermore, Yue et al. have reported the synthesis of metastable γ-Ga2O3 nanoflowers with hexagonal nanopetals by the oxidation of metallic Ga in solutions.27 The γ-Ga2O3 nanoflowers show excellent solar-blind detection performance; however, the poor thermal stability of the metastable phase still hinders the target application. Until now, the difficulty in synthesizing single-phase crystals and the poor thermal stability of the metastable γ-phase of Ga2O3 have greatly limited its potential applications in wide-band-gap high-performance optoelectronic devices.
In this work, the Al-doped γ-Ga2O3 thin film was epitaxially grown on α-Al2O3 (0001) substrates using the radio frequency magnetron sputtering technique, with the growth temperature successfully increased to 750 °C. A solar blind PD based on Al-doped γ-Ga2O3 thin film was successfully constructed. The PD shows excellent photoelectric properties. The γ-Ga2O3 thin film-based PDs might open up new possibilities for obtaining high-performance thin film-based solar blind UV optoelectronic devices.
2. Experimental section
2.1. Preparation and characterization of Ga2O3 thin films
The film samples were prepared on top of (0001) oriented single-crystal α-Al2O3 substrates using radio frequency (RF) magnetron sputtering. The Ga2O3 ceramic (π × 302 × 8 mm3, 99.99% purity) with/without embedded Al2O3 ceramic (π × 12.52 × 4 mm3, 99.99% purity) was used as a target for Al-doped/pure Ga2O3 thin film preparations, respectively. In the experiment, the Al2O3 target was covered at the glow circle position of the Ga2O3 target, and the incorporation amount was fine-tuned by varying the size of the coverage area. The base pressure in the chamber was 1 × 10−4 Pa, and the distance between the Ga2O3 target and the substrate was 6 cm. The flow rate of Ar (99.999% purity) gas was fixed to 24 sccm by a mass flow controller. The films were deposited at a working pressure of 1 Pa and RF power of 70 W for 2 h. The substrate temperature was fixed at 750 °C. The crystal structures of the as-grown films were investigated by a PANalytical X'pert PRO diffractometer using Cu Kα (λ = 1.5405 Å) radiation. The structural quality was estimated by X-ray rocking curve (XRC) measurements using Smart Lab X-ray diffraction (XRD). The surface morphologies were characterized by a Hitachi S-4800 field emission scanning electron microscope (FE-SEM). UV-visible (UV-vis) absorption spectra were taken using a Hitachi U-3900 UV-vis spectrophotometer. The elemental content was analyzed using X-ray photoelectron spectroscopy (XPS).
2.2. Fabrication and characterization of the photodetector
To construct a metal–semiconductor-metal (MSM) PD, a three-pair interdigital Au/Ti electrode was deposited on the film by RF magnetron sputtering, as illustrated in Fig. 1. The Au/Ti electrode is chosen because the affinity of Ga2O3 is close to the work function of Ti, making it easy to form ohmic contact. The width and length of the electrode fingers were 200 μm and 2800 μm, respectively. The spacing gap was 200 μm. A small point electrode (∼0.1 mm diameter) of indium metal was also pressed onto the Au/Ti electrode as connection points with a Cu wire. A Keithley 2450 source meter was utilized to measure the current–voltage (I–V) and time-dependent photoresponse of the Ga2O3 film-based PD. A UV lamp (∼7 W) with wavelengths of 254 nm and 365 nm was used as the light source, and the light irradiation power density was tuned by adjusting the distance between the light source and the sample.
 |
| | Fig. 1 Schematic of an MSM structure PD. | |
3. Results and discussion
The XRD patterns of undoped and Al-doped Ga2O3 thin films are shown in Fig. 2(a). The undoped Ga2O3 thin film formed on the α-Al2O3 substrate is a monoclinic β phase, as determined by XRD (JCPDS Card No. 43-1012). Besides the diffraction peaks of the Al2O3 substrate, diffraction peaks located at around 18.88°, 38.17°, and 58.88° are observed for the undoped film, which corresponds well to the β (
01), β (
02) and β (
03) lattice planes, respectively. The Al-doped Ga2O3 thin film formed on the α-Al2O3 substrate is a cubic γ phase, as determined by XRD (JCPDS Card No. 20-0426). Besides the diffraction peaks of the Al2O3 substrate, diffraction peaks located at around 18.67°, 37.85°, and 58.49° are observed for the Al-doped film, which corresponds well to the γ (111), γ (222) and γ (333) lattice planes, respectively. The X-ray rocking curves of γ-Ga2O3 (111) and β-Ga2O3 (
01) planes are shown in Fig. 2(b) and (c) for the Al-doped and undoped films, respectively, and correspond to the FWHM values of 3672 and 5184 arcsec. The result indicates that the crystal quality of the as-grown γ-Ga2O3 film is better than that of the β-Ga2O3 film, the preparation conditions of which are similar.
 |
| | Fig. 2 (a) XRD patterns of undoped and Al-doped Ga2O3 films. X-ray rocking curves of (b) γ-Ga2O3 (111) for the Al-doped film and (c) β-Ga2O3 ( 01)for the undoped film. | |
The top-view SEM morphologies of the Al-doped γ-Ga2O3 film and undoped β-Ga2O3 film are presented in Fig. 3(a). It is obvious that the γ-Ga2O3 and β-Ga2O3 films appear to form different surface structures on (0001) sapphire substrates. For the γ-Ga2O3 film, irregular pseudo-pyramid structures are distributed randomly on the surface, and the grain boundary is not clear; a similar surface structure has been observed for cubic-In2O3 on (0001) sapphire substrates in our previous study.28 The surface structures of the β-Ga2O3 film are columnar-like, and a clear grain boundary can be observed. Fig. 3(b) shows the UV-vis absorbance spectrum of the Al-doped γ-Ga2O3 film and the undoped β-Ga2O3 film. It is evident that all the films have significant absorption edges at ∼250 nm, near the lower edge of the solar-blind region. Moreover, the bandgap of the semiconductor film can be evaluated by extrapolating the linear region of the plots of (αhν)2 and (αhν)1/2 versus hν for direct and indirect transitions, respectively. For β-phase Ga2O3, a direct bandgap semiconductor, the bandgap value was determined to be 4.94 eV, as shown in the inset of Fig. 3(b). As the transition type of γ-phase Ga2O3 is undefined, both the direct and indirect bandgaps were estimated for the Al-doped γ-Ga2O3 film. The direct and indirect bandgap values were determined to be 5.08 and 4.43 eV, respectively, which were slightly higher than those for undoped γ-Ga2O3 (5.0 and 4.4 eV, respectively).19
 |
| | Fig. 3 (a) SEM images of the Al-doped γ-Ga2O3 and undoped β-Ga2O3 films. (b) UV-vis absorbance spectra of the Al-doped γ-Ga2O3 and undoped β-Ga2O3 films with the plot of (αhν)2 and (αhν)1/2 vs. hν in the inset. | |
XPS analysis was performed to elucidate the chemical compositions and oxygen deficiency in the as-grown films. The surface of the films was etched by Ar+ bombardment before the XPS measurement. Fig. 4(a) shows the wide survey spectrum in the binding energy range from −40 eV to 1350 eV to identify the elements in the Al-doped γ-Ga2O3 film. XPS peaks of Ga, O, Al and C were indicated in the Figure, and there were no other elements detected from the wide survey spectrum of the film. The charge-shift spectrum was calibrated using the C 1s peak at 284.8 eV. The Al 2p core level peak of the Al-doped γ-Ga2O3 film is presented in Fig. 4(b), and the energy peak center is observed at 71.1 eV. The peak position shows a significant shift compared to the Al 2p peak in pure Al2O3 film on the same substrate (74.5 eV). A similar phenomenon was also observed with a negative shift for the Al 2p peak during the formation of Al–O–Ga bonds.14 In their report, the peak position was noted at 73.9 eV in the (Al0.22Ga0.78)2O3 film; the significant shift in our result indicates that Al is present in the Al-doped γ-Ga2O3 film in the form of Al–O–Ga bonds. We have not yet identified whether the Al–O–Ga bonds are the key factors to obtain stabilized γ-Ga2O3, but they are a very significant sign to understand if the growing film is in the γ phase or not. The Ga 3d core level peaks of the Al-doped γ-Ga2O3 and undoped β-Ga2O3 films are shown in Fig. 4(c), with the energy peak centers observed at 19.7 eV and 20.1 eV, respectively. The O 1s core level peaks of the Al-doped γ-Ga2O3 and undoped β-Ga2O3 films are shown in Fig. 4(d), and the energy peaks are observed at 530.2 eV and 530.6 eV, respectively. The O 1s peaks can be divided into two components: I and II; peak I is attributed to lattice oxygen ions, and peak II is attributed to the oxygen ions in the oxygen vacancy region and surface defects of the film.5,10 The intensity of oxygen vacancies and surface defects in the film has a significant negative influence on the performance of Ga2O3 thin film-based PDs, such as dark current and response speed.29 The peak ratios of II/I were 1/5 and 1/3 for Al-doped γ-Ga2O3 and pure β-Ga2O3, respectively [Fig. 4(d)], which indicates a lower intensity of oxygen vacancies and surface defects in the Al-doped γ-Ga2O3 film than in the pure β-Ga2O3 film.
 |
| | Fig. 4 (a) XPS wide survey spectrum and (b) core level of Al 2p for the Al-doped γ-Ga2O3 film. Core levels of (c) Ga 3d and (d) O 1s for Al-doped γ-Ga2O3 and undoped β-Ga2O3 films. | |
Fig. 5 shows the room-temperature I–V characteristics of the PD based on an Al-doped γ-Ga2O3 thin film in the dark, under 365 nm (100 μW cm−2) and 254 nm (with various light intensities) light illumination. The I–V curve in the dark is linear, indicating that the contact is of an ohmic type. The corresponding resistance is 19.5 GΩ, and the dark current is about 2.57 nA at a voltage of 50 V, as shown in Fig. 5(a), which is much lower than that of undoped β-Ga2O3 films (about 100 nA at a voltage of 10 V), as per our previous reports. The I–V curve under 365 nm light illumination indicates that the device is almost insensitive to 365 nm light illumination, as shown in Fig. 5(a). Meanwhile, the device has significant sensitivity for 254 nm light illumination, as shown in Fig. 5(b).
 |
| | Fig. 5 (a) I–V characteristics of the Al-doped γ-Ga2O3 thin film-based MSM structure PD in the dark and under a 365 nm light illumination. (b) I–V characteristics of the device in the dark and under a 254 nm light illumination with various light intensities. | |
In order to further investigate the performance of the Al-doped γ-Ga2O3 thin film-based PD, the time-dependent photoresponse to solar-blind UV illuminations was investigated, as shown in Fig. 6. Fig. 6(a) and (b) show the time-dependent photoresponse of the device to 254 nm UV light illuminations with varied light intensities and bias voltages, respectively. The device shows excellent photoelectric performance, with the photocurrent gradually increasing with increasing incident light intensities and bias voltages. The larger the voltage applied, the higher the Iphoto obtained, following the basic photoconductance rule, which works as a photo-sensitive resistance. The higher the light intensity, the higher the Iphoto output, as more photo-generated carriers would be devoted to Iphoto at a certain voltage. Another important performance parameter of solar blind PDs is the response time. To study the response time of the device, the current rising and decaying process can be derived by fitting the photoresponse curve with the following relaxation equation:30
| |
 | (1) |
where
I0 is the photocurrent in the steady state,
A and
B are the constant values,
t is the time, and
τ1 and
τ2 are the relaxation time constants. As shown in
Fig. 6(c) and (d), the photoresponse processes fit quite well. We note that the rising edge consists of two components (
τr1 and
τr2), and the decaying edge is steep, consisting of only one component. It has been estimated that the rise edge constants
τr1 and
τr2 are around 2.39 s/12.57 s, 1.52 s/8.16 s, 1.02 s/6.92 s and 0.88 s/5.89 s for the device under 20 μW cm
−2, 40 μW cm
−2, 60 μW cm
−2 and 80 μW cm
−2 light illuminations at a bias voltage of 50 V, respectively. In contrast, the rise edge constants
τr1 and
τr2 are estimated to be 1.22 s/8.86 s, 1.04 s/7.95 s, 0.90 s/5.84 s, 0.84 s/7.03 s and 0.71 s/6.05 s for the device under 100 μW cm
−2 light illuminations at bias voltages of 10 V, 20 V, 30 V, 40 V and 50 V, respectively. It is obvious that the rise speed increases with both light intensity and bias voltage. Generally, the fast-response component can be attributed to the rapid change in carrier concentration that occurs as soon as the light is turned on or off, while the slow-response component is caused by carrier trapping/releasing owing to the existence of oxygen vacancies and other surface defects in the Al-doped γ-Ga
2O
3 film. It is evident that, when there are plenty of photon-generated carriers under stronger intensities of light illumination, the trapping/releasing effect is easier to saturate, so the response time is shorter; when the applied bias voltage is higher, the photo-generated electrons and holes undergo a rapid separation and then transfer toward the corresponding electrodes, resulting in a faster response speed. The decay edge constants
τd are estimated to be 0.44 s, 0.42 s, 0.41 s and 0.40 s for the device under 20 μW cm
−2, 40 μW cm
−2, 60 μW cm
−2 and 80 μW cm
−2 light illuminations at a bias voltage of 50 V, respectively. The decay edge constants
τd are estimated to be 0.46 s, 0.43 s, 0.43 s, 0.41 s and 0.40 s for the device under 100 μW cm
−2 light illumination at bias voltages of 10 V, 20 V, 30 V, 40 V and 50 V, respectively. Due to the absence of oxygen vacancies in Al doped γ-Ga
2O
3 film, the trapping/releasing effect is a localization transport process without plentiful of photon-generated carriers when the light is off. It can hardly influence the photoconductance of the device, so the decaying edge is quite steep.
 |
| | Fig. 6 Time-dependent photoresponse of the Al-doped γ-Ga2O3 thin film-based MSM structure PD measured under different conditions: (a) photoresponse of the device under different light intensities at 50 V; (b) photoresponse of the device at different bias voltages under 100 μW cm2; (c) experimental curve and fitted curve of the current rising and decaying process of the device under different light intensities at 50 V; (d) experimental curve and fitted curve of the current rising and decaying process of the device at different bias voltages under 100 μW cm2; (e) photoresponse stability of the device during a long time test and after storage for one month. | |
To quantitatively assess the device performance of the present solar blind photodetector, both responsivity (Rλ) and detectivity (D*) were calculated, as shown in Fig. 7(a) and (c) as a function of light intensity and bias voltage, respectively. Rλ, which is defined as the photocurrent generated per unit power of incident light on the effective area of a PD, is calculated by the following equation:2
| | |
Rλ = (Ilight − Idark)/(Pλ·S)
| (2) |
where
Ilight is the light current,
Idark is the dark current of the device when illuminated with a light source,
Pλ is the light intensity illuminated on the device,
S is the effective illuminated area (
S = 3.24 mm
2), and
λ is the wavelength of the illuminating light.
 |
| | Fig. 7 (a) Responsivity and detectivity of the Al-doped γ-Ga2O3 thin film-based MSM structure PD as a function of bias voltage. (b) External quantum efficiency of the device as a function of bias voltage. (c) Responsivity and detectivity of the device as a function of light intensity. (d) External quantum efficiency of the device as a function of light intensity. | |
Detectivity, which is usually used to evaluate the smallest detectable of a PD, can be calculated as follows:31
| |
 | (3) |
where
Rλ is the responsivity,
A is the effective area of the present device channel (
A = 3.24 mm
2),
e is the electronic charge, and
Idark is the dark current. It is clear that both the
Rλ and
D* values increase with increasing light intensity and bias voltage. Specifically, the responsivity and detectivity are as high as 0.26 A W
−1 and 1.79 × 10
12 under 100 μW cm
−2 light illumination at 50 V, respectively. Notably, the present device exhibits no obvious performance degradation with increasing light intensity and bias voltage, indicating that the device has good self-heating dissipation. Moreover, the external quantum efficiency (EQE) is also the key parameter to evaluate the performance of a PD, which is defined as the number of electrons probed per incident photon and can be calculated as follows:
32| |
 | (4) |
where
Rλ is the responsivity,
h is Planck's constant,
c is the velocity of light,
e is the electronic charge, and
λ is the exciting wavelength of the solar blind light. The EQE gradually increases with increasing light intensity and bias voltage, reaching a maximum of 124.92% under 100 μW cm
−2 light illumination at 50 V, as shown in
Fig. 7(b) and (d).
In order to investigate the spectral selectivity of the present device, the spectral responsivity was measured in the range of 225 to 600 nm, as shown in Fig. 8. It can be observed that the device shows excellent wavelength selectivity: it has highest responsivity at a wavelength of about 230 nm; as the wavelength increases, the responsivity gradually decreases and reaches a minimum value at about 280 nm. The rejection ratio of responsivity values for the solar blind region versus the solar visible region was estimated by dividing the responsivity at 230 nm by that at 280 nm. The excellent rejection ratio was evaluated at 60.98, indicating an excellent solar-blind UV response. The higher rejection ratio can be attributed to the better crystal quality and the absence of oxygen vacancies in Al-doped γ-Ga2O3.33
 |
| | Fig. 8 Normalized spectral selectivity of the Al-doped γ-Ga2O3 thin film-based MSM structure PD. | |
4. Conclusion
The Al-doped γ-Ga2O3 thin film was epitaxially grown on α-Al2O3 (0001) substrates by radio frequency magnetron sputtering, and the growth temperature was successfully increased to 750 °C. The solar blind UV PD based on the as-grown thin film exhibits excellent solar-blind UV photoelectric properties. Under 100 μW cm−2 254 nm light illumination at a bias voltage of 50 V, the device exhibits a rise and decay time of 0.71 s/6.05 s and 0.40 s, responsivity of 0.26 A W−1, detectivity of 1.79 × 1012, and an EQE of 124.92%, along with high stability and reproducibility. Remarkably, the present device exhibits no obvious performance degradation with increasing light intensity and bias voltage. It indicates that the device has good self-heating dissipation, which will be beneficial for future practical applications. Moreover, the device shows excellent wavelength selectivity with a solar blind/visible rejection ratio of 60.98, which is nearly blind to photons with wavelengths longer than 280 nm. These results suggest that γ-Ga2O3 thin film-based PDs can open up new possibilities for obtaining high performance thin film-based solar blind UV optoelectronic devices.
Conflicts of interest
There are no conflicts of interest to declare.
Data availability
The data supporting this article are available from the corresponding author upon reasonable request.
Acknowledgements
National Key R&D Program of China (2023YFB2906000); National Natural Science Foundation of China (62375055); Guangdong Basic and Applied Basic Research Foundation (2023B1515020088); Basic and Applied Basic Research Project of Guangzhou Basic Research Program (2025A04J5417) and the Start-Up Funding of Guangdong Polytechnic Normal University (2022SDKYA013) are greatly acknowledged.
References
- Y. Zhao, J. Zhang and D. Jiang, et al., Ultraviolet photodetector based on a MgZnO film grown by radio-frequency magnetron sputtering, ACS Appl. Mater. Interfaces, 2009, 1(11), 2428–2430 CrossRef CAS PubMed.
- W. Cui, D. Y. Guo and X. L. Zhao, et al., Solar-blind photodetector based on Ga2O3 nanowires array film growth from inserted Al2O3 ultrathin interlayers for improving responsivity, RSC Adv., 2016, 6(103), 100683–100689 RSC.
- Z. Li, T. Yan and X. Fang, Low-dimensional wide-bandgap semiconductors for UV photodetectors, Nat. Rev. Mater., 2023, 8(9), 587–603 CrossRef.
- X. Zhu, Y. Wu and Z. Pan, et al., Advancements in Ga2O3-based heterojunction ultraviolet photodetectors: Types, fabrication techniques, and integrated materials for enhancing photoelectric conversion efficiency, J. Alloys Compd., 2025, 1010, 177757 CrossRef CAS.
- Y. H. An, Z. S. Gao and Y. Guo, et al., One ε-Ga2O3-based solar-blind Schottky photodetector emphasizing high photocurrent gain and photocurrent-intensity linearity, Chin. Phys. B, 2023, 32(5), 058502 CrossRef CAS.
- Y. Li, T. Tokizono and M. Liao, et al., Efficient Assembly of Bridged β-Ga2O3 Nanowires for Solar-Blind Photodetection, ACS Appl. Mater. Interfaces, 2010, 20(22), 3972–3978 CAS.
- Y. Gao, J. Yang and X. Ji, et al., Semipolar (1122) AlGaN-based solar-blind ultraviolet photodetectors with fast response, ACS Appl. Mater. Interfaces, 2022, 14(18), 21232–21241 CrossRef CAS PubMed.
- M. M. Fan, K. W. Liu and X. Chen, et al., Mechanism of Excellent Photoelectric Characteristics in Mixed-Phase ZnMgO Ultraviolet Photodetectors with Single Cutoff Wavelength, ACS Appl. Mater. Interfaces, 2015, 7(37), 20600–20606 CrossRef CAS PubMed.
- J. Ye, S. Jin and Y. Cheng, Photocurrent ambipolar behavior in phase junction of a Ga2O3 porous nanostructure for solar-blind light control logic devices, ACS Appl. Mater. Interfaces, 2024, 16(20), 26512–26520 CrossRef CAS PubMed.
- R. Jin, J. Liu and S. Chen, et al., Ultrasensitive Self-Powered Flexible Crystalline β-Ga2O3-Based Photodetector Obtained through Lattice Symmetry and Band Alignment Engineering, ACS Appl. Mater. Interfaces, 2024, 16(32), 42406–42414 CrossRef PubMed.
- T. Oshima, T. Okuno and S. Fujita, Ga2O3 Thin Film Growth onc-Plane Sapphire Substrates by Molecular Beam Epitaxy for Deep-Ultraviolet Photodetectors, Jpn. J. Appl. Phys., 2007, 46(11R), 7217–7220 CrossRef CAS.
- Z. Y. Wang, G. Zhang and X. J. Zhang, et al., Polarization-sensitive artificial optoelectronic synapse based on anisotropic β-Ga2O3 single crystal for neuromorphic vision systems and information encryption, Adv. Opt. Mater., 2024, 12(29), 2401256 CrossRef CAS.
- T. Oshima, Y. Kato and M. Oda, et al., Epitaxial growth of gamma-(AlxGa1-x)O3 alloy films for band-gap engineering, Appl. Phys. Express, 2017, 10(5), 051104 CrossRef.
- H. Xu, Y. Weng and K. Chen, et al., Ultra-low BER encrypted communication based on self-powered bipolar photoresponse ultraviolet photodetector, Adv. Opt. Mater., 2024, 10, 2402238 Search PubMed.
- P. Li, L. Dong and C. Li, et al., Indium doping-assisted monolayer Ga2O3 exfoliation for performance-enhanced MOSFETs, Nanoscale, 2023, 15(28), 12105–12115 RSC.
- R. Roy, V. Hill and E. Osborn, Polymorphism of Ga2O3 and the System Ga2O3-H2O, J. Am. Chem. Soc., 1952, 74(3), 719–722 CrossRef CAS.
- X. C. Xia, Y. P. Chen and Q. J. Feng, et al., Hexagonal phase-pure wide band gap epsilon-Ga2O3 films grown on 6H-SiC substrates by metal organic chemical vapor deposition, Appl. Phys. Lett., 2016, 108(20), 202103 CrossRef.
- S. Fujita, M. Oda and K. Kaneko, et al., Evolution of corundum-structured III-oxide semiconductors: Growth, properties, and devices, Jpn. J. Appl. Phys., 2016, 55(12), 1202A3 CrossRef.
- T. Oshima, T. Nakazono and A. Mukai, et al., Epitaxial growth of gamma-Ga2O3 films by mist chemical vapor deposition, J. Cryst. Growth, 2012, 359(15), 60–63 CrossRef CAS.
- S. Yoshioka, H. Hayashi and A. Kuwabara, et al., Structures and energetics of Ga2O3 polymorphs, J. Phys.: Condens. Matter, 2007, 19(34), 346211 CrossRef.
- H. Nishinaka, D. Tahara and M. Yoshimoto, Heteroepitaxial growth of epsilon-Ga2O3 thin films on cubic (111) MgO and (111) yttria-stablized zirconia substrates by mist chemical vapor deposition, Jpn. J. Appl. Phys., 2016, 55(12), 1202BC CrossRef.
- S. X. Ge and Z. Zheng, Phase controllable synthesis of unusual epsilon-Ga2O3 single crystals with promising luminescence property, Solid State Sci., 2009, 11(9), 1592–1596 CrossRef CAS.
- S. D. Lee, Y. Ito and K. Kaneko, et al., Enhanced thermal stability of alpha gallium oxide films supported by aluminum doping, Jpn. J. Appl. Phys., 2015, 54(3), 030301 CrossRef CAS.
- R. Lorenzi, N. V. Golubev and E. S. Ignat'eva, et al., Defect-assisted photocatalytic activity of glass-embedded gallium oxide nanocrystals, J. Colloid Interface Sci., 2022, 608(3), 2830–2838 CrossRef CAS PubMed.
- G. Yang, H. A. Ni and H. D. Liu, et al., The doping effect on the crystal structure and electrochemical properties of LiMnxM1−xPO4 (M = Mg, V, Fe, Co, Gd), J. Power Sources, 2011, 196(10), 4747–4755 CrossRef CAS.
- H. Hayashi, R. Huang and H. Ikeno, et al., Room temperature ferromagnetism in Mn-doped gamma-Ga2O3 with spinel structure, Appl. Phys. Lett., 2006, 89(18), 181903 CrossRef.
- Y. Teng, L. X. Song and A. Ponchel, et al., Self-Assembled Metastable gamma-Ga2O3 Nanofl owers with Hexagonal Nanopetals for Solar-Blind Photodetection, Adv. Mater., 2014, 26(36), 6238–6243 CrossRef CAS PubMed.
- W. Cui, X. L. Zhao and Y. H. An, et al., Direct charge carrier injection into Ga2O3 thin films using an In2O3 cathode buffer layer: their optical, electrical and surface state properties, J. Phys. D Appl. Phys., 2017, 50(13), 135109 CrossRef.
- D. Guo, P. Li and Z. Wu, et al., Inhibition of unintentional extra carriers by Mn valence change for high insulating devices, Sci. Rep., 2016, 6(1), 24190 CrossRef CAS PubMed.
- D. Y. Guo, Z. P. Wu and P. G. Li, et al., Fabrication of beta-Ga2O3 thin films and solar-blind photodetectors by laser MBE technology, Opt. Mater. Express, 2014, 4(5), 1067–1076 CrossRef.
- W. Y. Kong, G. A. Wu and K. Y. Wang, et al., Graphene-beta-Ga2O3 Heterojunction for Highly Sensitive Deep UV Photodetector Application, Adv. Mater., 2016, 28(48), 10725 CrossRef CAS PubMed.
- J. M. Wu and W. E. Chang, Ultrahigh Responsivity and External Quantum Efficiency of an Ultraviolet-Light Photodetector Based on a Single VO2 Microwire, ACS Appl. Mater. Interfaces, 2014, 6(16), 14286–14292 CrossRef CAS PubMed.
- A. S. Pratiyush, S. Krishnamoorthy and S. V. Solanke, et al., High responsivity in molecular beam epitaxy grown beta-Ga2O3 metal semiconductor metal solar blind deep-UV photodetector, Appl. Phys. Lett., 2017, 110(22), 221107 CrossRef.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.