Abstract
This study systematically investigated the effects of oxygen plasma treatment on oxygen vacancy defects in sputtered β-gallium oxide (β-Ga2O3) films and their corresponding ultraviolet (UV) detection performance. The sputtered β-Ga2O3 film subjected to 1 min of oxygen plasma treatment exhibited optimal photodetection properties. Compared to the untreated sample, the dark current was reduced by approximately one order of magnitude to 0.378 pA at 10 V bias. It exhibited an 86% (from 2.92 s to 0.41 s) decrease in response time, a 41.6% increase in photocurrent, a very high photo-to-dark current ratio of 9.18 × 105, and a specific detectivity of 2.62 × 1010 cm·Hz1/2W−1 under 254 nm UV illumination intensity of 799 μW/cm2 at 10 V bias. Notably, appropriate oxygen plasma treatment minimizes electron capture, enhances the separation and collection of photogenerated carriers, and suppresses the persistent photoconductivity (PPC) effect, thus ultimately shortening the response time. Oxygen plasma processing thus provides an effective approach to fabricating high-performance β-Ga2O3 solar-blind photodetectors (SBPDs).
1. Introduction
With the growing demand for high-precision and anti-interference detection technologies in fields such as atmospheric environmental monitoring, public safety early warning, and national defense security, solar-blind ultraviolet (UV) band (200–280 nm) detection has emerged as a research hotspot in the optoelectronic field in recent years, owing to its unique advantage of zero interference from solar radiation [1,2]. Photodetectors operating in the solar-blind UV band (200–280 nm) exhibit low false alarm rates, high sensitivity, and low background noise [3,4], enabling their wide applications in ozone detection, fire alarms, optical communications, and missile monitoring [5,6,7]. Candidate materials for such detectors include MgZnO [8], AlGaN [9], diamond [10], Ga2O3 [11], etc. Among these, Ga2O3 is particularly promising for high-performance solar-blind photodetectors, owing to its excellent chemical stability, thermal stability, ultrahigh breakdown voltage, and a bandgap that matches the solar-blind region [12]. β-Ga2O3 is the most extensively studied crystalline phase for photodetection applications among the five polymorphs of Ga2O3, primarily due to its thermodynamic stability [13,14]. With a bandgap aligned to the solar-blind region, β-Ga2O3 possesses distinct advantages for solar-blind UV photodetector development. However, heteroepitaxially grown β-Ga2O3 films inevitably contain oxygen vacancy defects, which pose a critical bottleneck to the enhancement of photodetector performance.
Radio frequency magnetron sputtering (RFMS) is a well-established technique for β-Ga2O3 film deposition, featuring a simple operation process and easy control over film growth attributes that make it suitable for large-scale, cost-effective production [15]. Nevertheless, due to the high cost and preparation challenges of Ga2O3 single-crystal substrates, the epitaxial growth of β-Ga2O3 films on heterogeneous substrates (e.g., sapphire and silicon wafers) is preferred. Lattice mismatch between β-Ga2O3 and heterogeneous substrates can induce local structural distortions in the β-Ga2O3 crystal lattice, leading to the incomplete filling of oxygen atoms and the formation of oxygen vacancy defects. These defects act as charge carrier traps, increasing the dark current and reducing carrier mobility; consequently, they cause persistent photoconductivity (PPC) and significantly degrade the response speed of detectors [16].
The essence of the PPC effect lies in the trapping and slow release of photogenerated carriers via deep-level defects in the material (such as oxygen vacancies and impurity levels). The impact of PPC on device performance is highly scenario-dependent; in the field of photodetection where rapid response is paramount, PPC causes the current to fail to drop synchronously after light termination, manifesting as prolonged response times and degraded dynamic switching performance, thereby posing a key bottleneck to device practicality. Conversely, the presence of PPC can enable various alternative applications, such as enhancing detection responsivity and enabling its use in memory devices. Wang et al. demonstrated a feasible strategy to preserve hot electrons for efficient collection by intentionally engineering a mix of cubic zinc-blende and hexagonal wurtzite phases in III–V semiconductor nanowires. This heterostructure design creates additional energy levels that are above the conduction band minimum, which capture and store hot electrons before they thermalize to the band edges. Consequently, the lifetime of photogenerated carriers is prolonged, leading to enhanced responsivity [17]. Yang et al. report a novel negative photoconductivity (NPC) mechanism in n-type indium arsenide nanowires (NWs), revealing competing positive photoconductivity to coexist with NPC. Low temperatures slow post-illumination conductivity recovery, indicating a thermally activated detrapping mechanism. At 78 K, conductance’s spontaneous recovery is quenched, enabling a reversible memory device switchable via light and gate voltage pulses. This NPC-based platform shows promise for high-performance photodetection and low-power nonvolatile memory [18]. This study demonstrates that regulating defect concentration to reduce oxygen vacancies and surface defects can significantly suppress the PPC effect, providing a critical solution for optimizing the performance of β-Ga2O3 photodetectors.
Modulating defect concentration is critical for optimizing the performance of semiconductor devices. Current mainstream strategies include the following: (1) suppressing defect generation by optimizing process parameters (e.g., growth temperature and oxygen partial pressure) [19,20]; (2) reducing defect density via high-temperature thermal annealing, which repairs crystal structures and relieves residual stress [21]; and (3) selectively doping impurity atoms to reconstruct electronic structures and regulate defect state distributions [22].
To address the challenge of defect modulation in semiconductor materials, plasma surface treatment technology has been successfully applied in wide-bandgap semiconductor materials, such as ZnO and GaN, owing to its inherent advantages, including low-temperature operation, high efficiency, and precise controllability. Plasma surface treatment has recently emerged as a promising technology for use in semiconductor modification. Liu et al. demonstrated that oxygen plasma treatment reduces oxygen vacancy concentration in ZnO films, yielding ZnO-based photodetectors with response/recovery times below 50 µs and a responsivity of 1–10 A/W [23]. Surface passivation (via plasma or other methods) has also been shown to significantly enhance the performance of Ge-, GaAs-, and GaN-based UV photodetectors; it reduces defect density, suppresses leakage currents, and, thereby, improves photodetection sensitivity [16,24,25]. For Ga2O3-based materials, Zhang et al. fabricated β-Ga2O3 films via metal–organic chemical vapor deposition (MOCVD); through the combination of oxygen annealing and plasma treatment, these films achieved a dark current of 29 pA, a normalized detectivity of 1.3 × 1012 Jones, and a response rejection ratio of 8.6 × 106, under a 10 V bias [26]. Existing studies confirm that various plasma treatments can effectively regulate the structural and electrical properties of Ga2O3 [27,28]. These studies have consistently verified that plasma treatment enables the repair of lattice defects and the suppression of carrier traps via mechanisms, including ion bombardment and active oxygen supplementation, thereby yielding significant improvements in device performance. Nevertheless, most of the existing literature centers on performance comparisons between “plasma-treated” and “untreated” samples, with insufficient in-depth investigation into “treatment duration”—a critical process parameter. This research gap leads to a lack of quantitative insights for guiding process optimization, particularly in the context of RFMS-grown β-Ga2O3 film systems, where the correlation between plasma treatment duration and photodetector performance remains underexplored. The present study is conducting systematic work around this problem.
Building on these findings and addressing the aforementioned challenges, we deposited β-Ga2O3 films on sapphire substrates via RFMS and treated them with oxygen plasma to reduce oxygen vacancies. The results reveal substantial reductions in dark current and response time, accompanied by a notable enhancement in the overall performance of the photodetectors.
2. Materials and Methods
2.1. Fabrication of Oxygen Plasma-Treated β-Ga2O3 Photodetectors
The fabrication process of oxygen plasma-treated β-Ga2O3 photodetectors is illustrated in Figure 1. All processes were conducted following standard semiconductor manufacturing protocols, with key steps detailed as follows.
Figure 1.
Preparation process of β-Ga2O3 photodetectors with oxygen plasma treatment.
First, substrate and target preparation was performed. Double-side polished c-plane sapphire (0001) was used as the substrate, and a 99.99% pure Ga2O3 ceramic disk (60 mm in diameter, 5 mm in thickness) served as the sputtering target. Prior to deposition, the sapphire substrate was ultrasonically cleaned sequentially in acetone, ethanol, and deionized water for 15 min per step, following standard semiconductor cleaning procedures, to remove surface contaminants.
Second, β-Ga2O3 film deposition was performed. A magnetron sputtering system (Model MSP-300B, Beijing Chuangshi Weina Technology Co., Ltd., Beijing, China) was utilized for thin-film growth. Prior to deposition, the system was evacuated to a base pressure of 4.5 × 10−4 Pa. Pre-sputtering was conducted for 5 min to remove surface contaminants from the Ga2O3 target, under the following conditions: sputtering pressure of 0.3 Pa, sputtering power of 100 W, argon (Ar) flow rate of 50 sccm, and oxygen (O2) flow rate of 1 sccm. Subsequently, a 200 nm thick β-Ga2O3 film was deposited onto the cleaned sapphire substrate using the same sputtering conditions. To enhance crystallinity, the as-deposited film was then annealed in a high-temperature tube furnace at 1000 °C for 60 min under an oxygen atmosphere.
Third, oxygen plasma treatment was conducted. To investigate the effect of the treatment duration on the photodetector performance, five sample groups were prepared with the following different oxygen plasma exposure times: 0 s (pristine sample), 15 s (Plasma-15 s), 30 s (Plasma-30 s), 1 min (Plasma-1 min), and 5 min (Plasma-5 min). The specific meaning of the samples is shown in Table 1. Plasma treatment was carried out using an inductively coupled plasma cleaning machine (Model CY-P21-300W, Zhengzhou CY Scientific Instruments Co., Ltd., Zhengzhou, China) at a fixed output power of 40 W and a chamber pressure of 30 Pa.
Table 1.
Sample name definition.
Fourth, electrode fabrication and post-annealing was performed. Ti/Au (40 nm/90 nm) interdigital electrodes were deposited on all treated samples via DC magnetron sputtering, using a metal mask to ensure precise patterning. The electrodes had a width of 9000 μm, length of 9000 μm, and a finger spacing of 300 μm, and the effective light-irradiated area of each photodetector was calculated to be 2.19 × 10−5 m2. Finally, all samples were annealed in a muffle furnace at 500 °C for 10 min under ambient air to form reliable ohmic contacts.
2.2. Characterization and Performance Measurement
X-ray diffraction (XRD): Crystalline phase and structural quality were analyzed using a Smartlab XRD system (Rigaku Corporation, Tokyo, Japan) with Cu Kα1 radiation (λ = 0.15406 nm).
X-ray photoelectron spectroscopy (XPS): The chemical valence states and relative content of oxygen in the films were determined using an ESCALAB 250Xi XPS system (Thermo Scientific, Waltham, MA, USA).
Photoluminescence (PL) spectroscopy: Room-temperature (298 K) PL spectra were recorded using a steady-state/transient fluorescence spectrometer (Model FLS2000, Edinburgh Instruments, Livingston, Scotland, UK) with a 254 nm excitation wavelength.
Photodetector performance testing: The photoresponse characteristics (dark current, photocurrent, and response time) were measured at room temperature and ambient air using a Keithley 4200-SCS semiconductor parameter analyzer (Tektronix, Beaverton, OR, USA) coupled with a UV mercury lamp (254 nm, as the excitation source). The Keshengda Light Shutter-S19CA (Chengdu Keshengda Optoelectronic Technology Co., Ltd. Chengdu, China) is used to control the light-on time and light-off time. Subsequently, the optoelectronic properties of the photodetectors were discussed based on the collected data.
3. Results and Discussion
Figure 2a shows the XRD patterns of Ga2O3 films grown on (0001) sapphire substrates. For all five samples, three distinct and invariant diffraction peaks near 2θ = 18.8°, 38.4°, and 59.2°, corresponding to the (−201), (−402), and (−603) crystal planes of β-Ga2O3 (JCPDS No. 43-1012), respectively, indicate the excellent crystallographic orientation of these films [29] and signify that the oxygen plasma treatment did not induce any significant alterations to the crystal structures of the Ga2O3 thin films.
Figure 2.
Performance characterization results of all samples, as follows: (a) XRD patterns, (b) UV-Vis absorption spectra, (c) full XPS spectra, and (d) XPS spectra of Ga2p.
Figure 2b illustrates the transmission spectra of Ga2O3 films before and after oxygen plasma treatment. All samples exhibit a maximum absorption peak centered at ~254 nm, accompanied by sharp absorption edges—this spectral feature is consistent with the wavelength requirement for solar-blind UV detection, indicating highly suitability for solar-blind detection [30]. For plasma-treated samples, the average transmittance exceeds 90% in the visible wavelength range of 250–800 nm, reflecting excellent optical transparency. Furthermore, the absorption edge of the films remains unchanged after plasma treatment, with a moderate enhancement in optical transmittance; this improvement is attributed to the surface cleaning and modification induced by oxygen plasma, which eliminates surface contaminants and reduces light absorption. The optical bandgap of the films was calculated using the Tauc plot method, and the results reveal a consistent value of 4.96 eV across all samples, a finding that further confirms that oxygen plasma treatment neither alters the optical bandgap of β-Ga2O3 films, nor induces changes in their crystal structure.
To further investigate the elemental composition of β-Ga2O3 films before and after oxygen plasma treatment, X-ray photoelectron spectroscopy (XPS) characterization was conducted. All XPS spectra were calibrated using the C 1s peak at a binding energy of 284.8 eV.
Figure 2c presents the full-scan XPS spectra of the five samples (pristine and plasma 15 s–5 min). Distinct peaks corresponding to Ga 3d, Ga 3p, Ga 3s, O 1s, C 1s, Ga 2p, and Ga LMM, as well as O KLL Auger peaks, are observed for all samples. The absence of impurity-related peaks confirms that the films consist exclusively of gallium (Ga) and oxygen (O), consistent with the target composition of β-Ga2O3 [31]. Figure 2d displays the high-resolution Ga 2p X-ray photoelectron spectroscopy (XPS) spectra of β-Ga2O3 films before and after oxygen plasma treatment, showing two distinct characteristic peaks for all samples, the Ga 2p1/2 peak centered at ~1145.2 eV and the Ga 2p3/2 peak at ~1118.3 eV, with a peak separation of ~26.9 eV that confirms the presence of Ga2O3 [32]. In the study of samples (Plasma-15 s–Plasma-5 min), XPS analysis showed a redshift in Ga 2p and Ga 3d peaks. Oxygen anions adsorbed to formed chemisorbed oxygen, causing electrons to transfer to Ga, increasing its electron-cloud density and reducing the binding energy.
To identify the elemental oxygen species in Ga2O3 films, the O 1s XPS spectra were analyzed and deconvoluted into three peaks via Gaussian fitting (Figure 3). These peaks correspond to lattice oxygen (OL) at 529.8 eV, vacancy oxygen (Ov) at 530.7 eV, and adsorbed oxygen (Oa) at 532.2 eV. The Ov contents of the five samples were 66.5%, 54.9%, 51%, 46.8%, and 52.1%, while the Oa contents were 17.61%, 19.86%, 21.62%, 28.18%, and 27.68%, respectively. These results show that oxygen plasma treatment effectively modulates oxygen vacancies and recovers the crystalline surface. However, prolonged treatment (Plasma-5 min) disrupts the lattice structure, causing an increase in oxygen vacancies. All data variations observed during the characterization and testing processes are summarized in Table 2, and specific differences in relevant parameters can be referred to in this table.
Figure 3.
Fine-sweeping and split-peak fitting of the peaks of Ga2O3 thin films (a) pristine film, and (b–e) oxygen plasma-treated samples (Plasma-15 s–Plasma-5 min).
Table 2.
Device characterization and performance parameter comparison.
To elucidate the impact of defects on the optical properties of pristine and oxygen plasma-treated β-Ga2O3 thin films, PL spectra of five samples were characterized under 254 nm UV excitation. As shown in Figure 4a–e, all samples exhibited asymmetric ultraviolet–green emission peaks (300–630 nm). After normalization, the peaks were assigned as follows. Peak I represents intrinsic luminescence, as the ultraviolet (UV) luminescence of gallium oxide is caused by self-trapped exciton recombination. Peak II arises from donor-acceptor pair (DAP) recombination involving oxygen vacancies (VO), gallium vacancies (VGa), and their complexes. Here, VGa/VGa-VO capture electrons as acceptor levels, while VO/Gai (gallium interstitials) supply carriers as donor levels, generating a blue emission band via radiative recombination. The origin of the green emission (Peak III) remains unclear, though V Pati et al. [33] propose that it may result from DAP transitions between VO-related donors and VGa or VGa + VO acceptors. The area ratios of Peak II/Peak I for the five samples were 3.00, 1.91, 1.51, 1.02, and 1.41. Sample Plasma-1 min, with the lowest ratio (1.02), indicated the lowest oxygen vacancy concentration, consistent with prior XPS results. This confirms that 1 min oxygen plasma treatment effectively reduces oxygen vacancies in β-Ga2O3 thin films.
Figure 4.
PL spectrum and split-peak fitting of peaks of Ga2O3 thin films (a) pristine film, and (b–e) oxygen plasma-treated samples (Plasma-15 s–Plasma-5 min) (the normalized intensity).
The I-V characteristics of the five β-Ga2O3 solar-blind UV photodetectors were measured under dark and illuminated conditions, with the results shown in Figure 5a,b, respectively. The light source utilized for these measurements has a central wavelength of 254 nm and a power density of 799 μW/cm2. Combining the results in Figure 3 and Figure 4, it can be concluded that oxygen plasma treatment significantly reduces the dark current. This effect is attributed to reducing oxygen vacancies in β-Ga2O3 films and suppressing carrier recombination induced by oxygen-related defects [34]. Notably, sample Plasma-1 min exhibits the lowest oxygen vacancy concentration, consistent with its minimal dark current among all samples. Specifically, at a bias voltage of 10 V, the dark current of sample Plasma-1 min is as low as 0.378 pA—approximately one order of magnitude lower than that of the pristine sample.
Figure 5.
Pristine and oxygen-plasma-treated Ga2O3 films (a) under UV irradiation as well as PD schematic, and (b) in dark and (c) time-dependent light response I-T curves during periodic turn-on and turn-off of 254 nm UV light at 799 μW/cm2 with 10 V bias. (d) Relation between R, D* and light intensity.
Compared to the untreated samples, the photocurrent of samples Plasma-15 s and Plasma-30 s decreased (Figure 5b,c). This trend can be attributed to the fact that energy-sufficient reactive oxygen species dissociate into oxygen anions, which overcome the energy barrier to occupy oxygen vacancy sites—reducing both the oxygen vacancy content and the corresponding carrier density, thereby leading to lower photocurrents [28]. Extending the treatment to 1 min not only further reduces oxygen vacancies concentration but also eliminates surface defects and suppresses the defect-related recombination of photogenerated carriers. Specifically, the sample treated with 1 min plasma (Plasma-1 min) exhibits a significant 41.3% increase in photocurrent. This improvement occurs because the reduction in defects, which include both oxygen vacancies and surface defects, facilitates the transport and collection of photogenerated carriers, decreases the likelihood of carrier recombination, and, thus, enhances the photocurrent. Moreover, the dark current is reduced as the defect-mediated recombination is mitigated by the 1 min oxygen plasma treatment [26]. Excessive removal of oxygen vacancies depletes free electrons from donor defects, thereby disrupting the balance of carrier concentration. This disruption reduces the density of photogenerated carriers and the corresponding photocurrent, ultimately degrading optoelectronic performance. Furthermore, the β-Ga2O3 film was subjected to prolonged treatment with oxygen plasma. For instance, the treatment used in the Plasma-5 min sample results in a sudden drop in the photocurrent; on one hand, excess chemical adsorbates act as centers for carrier scattering and trapping, and on the other hand, prolonged ion bombardment damages the crystal lattice. Together, these two factors lead to reduced photocurrent and overall electrical performance.
The β-Ga2O3 photodetector treated for 1 min achieved the highest photo-to-dark current ratio (PDCR) of 9.18 × 105. Figure 5c shows the response curves of the photocurrents over time at 10 V bias, demonstrating excellent on/off characteristics and reproducibility.
Responsivity (R) and specific detectivity (D*) are crucial parameters for evaluating the photodetector photoelectric performance [35].
R can be expressed as follows:
where Pλ is the optical power density of incident UV light and S is the effective light area of the device.
D* is used to describe the ability of the photodetector to detect weak light and can be expressed as follows:
Here, noise measurement was not conducted due to the limitations in experimental conditions, and the bandwidth was set to a default value of ∆f = 1 Hz. q denotes the charge of an electron. The calculations of R and D* at a bias voltage of 10 V are presented in Figure 5d. At a 10 V bias and 799 μW/cm2 light intensity, the untreated sample exhibited R = 1.4 mA/W and D* = 7.63 × 109 cm·Hz1/2W−1, while sample Plasma-1 min showed the significantly enhanced R = 1.95 mA/W and D* = 2.62 × 1010 cm·Hz1/2W−1. Most photodetectors exhibit decreasing responsivity as incident light intensity increases. However, in Figure 5d of this manuscript, responsivity (R) and detectivity (D*) show an upward trend with rising light intensity. Through detailed analysis, we propose two possible explanations for this phenomenon. First, the number of photogenerated charge carriers is relatively higher than that of incident photons, coupled with a lower carrier recombination rate. This allows the density of photogenerated carriers to increase continuously with light intensity, ultimately leading to the observed growth of R and D* with light intensity in Figure 5d [36,37]. Second, it may also stem from an insufficient incident light intensity, which keeps the photoresponse in an unsaturated state.
Figure 6a–e presents the response/recovery time test results for the five samples. Response time τr is defined as the time for photocurrent to rise from 10% to 90% of the peak value upon light illumination, while recovery time τd is the time for photocurrent to decrease from 90% to 10% after light cessation. The response/recovery times for the samples were 2.92 s/1.35 s, 1.17 s/0.98 s, 0.76 s/0.96 s, 0.41 s/0.17 s, and 1.18 s/1.57 s, respectively. Devices without oxygen plasma treatment have higher oxygen vacancy concentrations, which trap electrons and impede carrier migration, resulting in longer response times. After 1 min plasma treatment, τr and τd are significantly reduced due to fewer oxygen vacancies acting as recombination centers. This minimizes electron capture, enhances photogenerated carrier separation and collection, and suppresses the persistent photoconductivity (PPC) effect, thus shortening response times. Conversely, sample Plasma-5 min shows increased τr/τd due to prolonged plasma treatment causing excessive surface chemical adsorption. These adsorbates act as scattering and trapping centers, reducing electron mobility and recombination efficiency, leading to longer response times. Therefore, optimizing plasma treatment time is crucial for device performance.
Figure 6.
I-T curves of normalized rise and decay processes for (a) pristine and (b–e) plasma-treated Ga2O3 films with different treatment times (Plasma-15 s–Plasma-5 min) under 799 μW/cm2 illumination at 10 V (pink represents the time of ascent and descent).
To evaluate the long-term operational stability of the β-Ga2O3 photodetectors, the Plasma-1 min sample (optimally treated oxygen plasma) was irradiated with UV light for 100+ cycles under a 10 V bias, with each cycle lasting 10 s (total test duration ~1500 s). The corresponding transient photoresponse curves are presented in Figure 7a,b. After over 1500 s of continuous cyclic testing, the Plasma-1 min photodetector maintained a consistent and excellent photocurrent response, with no obvious degradation in signal amplitude—confirming its repeatability and long-term reliability. This result demonstrates that the optimal oxygen plasma treatment (1 min) not only effectively modulates oxygen vacancy defects to enhance photodetection performance but also preserves the photodetector’s long-term operational stability, a critical requirement for its practical application in solar-blind UV detection scenarios.
Figure 7.
(a,b) Long-term stability test of the transient optical response characteristics of the device at 10 V bias.
Oxygen vacancies (OV) are the primary intrinsic defects in Ga2O3, which originate from an insufficient oxygen source, excessively high growth temperature, or lattice stress during thin film growth. These factors cause the detachment of partial O atoms from the lattice, forming positively charged vacancies (predominantly OV2+). Oxygen plasma treatment can introduce strongly oxidizing oxygen ions (O2−) and highly reactive oxygen radicals (O) [38]. Through oxygen plasma treatment, the following occurs:
Ga2O3(OV2+) + 2O + 2e− → Ga2O3
The reactive oxygen species overcome the energy barrier to occupy OV sites, thereby reducing the oxygen vacancy concentration in the film. Short-duration treatments (15 s and 30 s) lead to a decrease in the photocurrent due to the reduced oxygen vacancy content and a corresponding drop in carrier density [28]. Extending the treatment to 1 min not only further reduces oxygen vacancies, but also eliminates surface defects and suppresses the defect-related recombination of photogenerated carriers, resulting in a slight recovery of photocurrent; meanwhile, the dark current is reduced as the defect-mediated recombination is mitigated by the oxygen plasma treatment [26]. However, excessive treatment time (e.g., sample Plasmap-5 min) results in a sudden drop in photocurrent because excessive chemical adsorbates act as carrier scattering/trapping centers, and prolonged ion bombardment damages the crystal lattice, ultimately degrading both the photocurrent and overall electrical performance.
Table 3 presents a comparison of the performances of Ga2O3 thin films treated with oxygen plasma via different preparation techniques reported in the literature, to the best of our knowledge. Our work exhibits remarkable advantages in τr/τd, along with a low dark current and relatively strong detection capability. Furthermore, by employing the RFMS process that is more conducive to industrialization, a favorable balance between performance and production feasibility has been attained.
Table 3.
Recent comparison of Ga2O3 thin film performances treated with oxygen plasma via different preparation technologies.
4. Conclusions
In summary, this study demonstrates that appropriate oxygen plasma treatment of sputtered β-Ga2O3 films effectively reduces their oxygen vacancy content, which not only weakens the electron capture capability of defects, but also significantly suppresses the persistent photoconductivity (PPC) effect, thereby enhancing the photodetection performance. Specifically, the β-Ga2O3 photodetector treated with oxygen plasma for 1 min (the optimal duration) exhibited the following exceptional optoelectronic properties: at a 10 V bias, it achieved an extremely low dark current of 0.378 pA; under 254 nm UV illumination (power intensity: 799 μW/cm2) and 10 V bias, it delivered a high photo-to-dark current ratio of 9.18 × 105, a responsivity of 1.95 mA/W, and an elevated specific detectivity of 2.62 × 1010 cm·Hz1/2W−1; and compared to the untreated (pristine) sample, its UV response performance was drastically improved, with the response rise time and recovery time shortened from 2.92 s and 1.35 s, to 0.41 s and 0.17 s, respectively, and the photocurrent intensity increased from 245 nA to 347 nA. Given its efficiency, low cost, and compatibility with β-Ga2O3 film fabrication processes, oxygen plasma treatment emerges as a highly promising strategy for advancing the performance of β-Ga2O3 solar-blind UV photodetectors, with significant potential for practical applications in fields such as ozone monitoring and missile warning. Given that the oxygen plasma treatment technique with an appropriate processing duration not only combines the advantages of having a high efficiency and low cost, but it also exhibits excellent compatibility with the magnetron sputtering-based β-Ga2O3 thin film fabrication process, which enables industrial-scale and large-volume production. Thus, this technique has emerged as a highly promising strategy for enhancing the performance of β-Ga2O3 SBPDs. Particularly in fields such as ozone monitoring and missile early warning systems, this method further demonstrates significant potential in practical applications.
Author Contributions
G.W.: Conceptualizaiton (equal); Data curation (equal); Investigation (equal); Methodology (equal); Writing—original draft (equal); Writing—review and editing (equal). R.D.: Data curation (equal); Investigation (equal); Writing—original draft (equal); L.G.: Supervision (equal); Writing—review and editing (equal); Resources (equal); Project administration (equal). Y.W.: Data curation (supporting). Y.Z.: Data curation (supporting); X.L.: Methodology (supporting). J.W.: Writing—original draft (supporting). Y.Y.: Data curation (supporting). X.Y.: Resources (equal); Supervision (equal); Writing—review and editing (equal). All authors have read and agreed to the published version of the manuscript.
Funding
This work was financially supported by the National Natural Science Foundation of China (62173128, 62303166), the Natural Science Foundation of Henan (252300421296), the Key Project of Scientific and Technology Research of Henan Province (252102210201), the Program of Henan Province Office of Education (25A510013, 24B510007), and the Outstanding Youth Project of Henan Polytechnic University (J2025-6).
Data Availability Statement
The authors declare that there are no financial interests, commercial affiliations, or other potential conflicts of interest that could have influenced the objectivity of this research or the writing of this paper. The data supporting the findings of this study are available within the paper. Any additional data connected to the study are available from the corresponding author upon reasonable request.
Conflicts of Interest
Author Xiaojie Yang director of research and development (R and D) on optoelectronic materials and devices. He is the founder of Suzhou LiangSai Technology Co., Ltd. The remaining authors declare no conflicts of interest.
Abbreviations
The following abbreviations are used in this manuscript:
| SBPDs | Solar-blind photodetectors |
| PPC | Persistent photoconductivity |
| RFMS | Radio frequency magnetron sputtering |
| UV | Ultraviolet |
References
- Ding, M.; Hao, W.; Yu, S.; Liu, Y.; Zou, Y.; Xu, G.; Zhao, X.; Hou, X.; Long, S. Self-Powered p-NiO/n-Ga2O3 Heterojunction Solar-Blind Photodetector with Record De-tectivity and Open Circuit Voltage. IEEE Electron Device Lett. 2022, 44, 277–280. [Google Scholar] [CrossRef]
- Wu, W.; Huang, H.; Wang, Y.; Yin, H.; Han, K.; Zhao, X.; Feng, X.; Zeng, Y.; Zou, Y.; Hou, X. Structure engineering of Ga2O3 photodetectors: A review. J. Phys. D Appl. Phys. 2024, 58, 063003. [Google Scholar] [CrossRef]
- Wang, G.; Wang, H.; Chen, T.; Feng, Y.; Zeng, H.; Guo, L.; Liu, X.; Yang, Y. Enhancing β-Ga2O3-film ultraviolet detectors via RF magnetron sputtering with seed layer insertion on c-plane sapphire substrate. Nanotechnology 2023, 35, 095201. [Google Scholar] [CrossRef]
- Ye, J.; Jin, S.; Cheng, Y.; Xu, H.; Wu, C.; Wu, F.; Guo, D. Photocurrent Ambipolar Behavior in Phase Junction of a Ga2O3 Porous Nanostructure for Solar-Blind Light Control Logic Devices. ACS Appl. Mater. Interfaces 2024, 16, 26512–26520. [Google Scholar] [CrossRef]
- Wu, C.; Wu, F.; Hu, H.; Wang, S.; Liu, A.; Guo, D. Review of self-powered solar-blind photodetectors based on Ga2O3. Mater. Today Phys. 2022, 28, 100883. [Google Scholar] [CrossRef]
- Kaur, D.; Kumar, M. A Strategic Review on Gallium Oxide Based Deep-Ultraviolet Photodetectors: Recent Progress and Future Prospects. Adv. Opt. Mater. 2021, 9, 2002160. [Google Scholar] [CrossRef]
- Xu, H.; Weng, Y.; Chen, K.; Wu, C.; Hu, H.; Guo, D. Ultra-Low BER Encrypted Communication Based on Self-Powered Bipolar Photoresponse Ultraviolet Photodetector. Adv. Opt. Mater. 2024, 13, 2401256. [Google Scholar] [CrossRef]
- Wang, L.K.; Ju, Z.G.; Zhang, J.Y.; Zheng, J.; Shen, D.Z.; Yao, B.; Zhao, D.X.; Zhang, Z.Z.; Li, B.H.; Shan, C.X. Single-crystalline cubic MgZnO films and their application in deep-ultraviolet optoelectronic devices. Appl. Phys. Lett. 2009, 95, 131113. [Google Scholar] [CrossRef]
- Yang, J.-L.; Liu, K.-W.; Shen, D.-Z. Recent progress of ZnMgO ultraviolet photodetector. Chin. Phys. B 2017, 26, 047308. [Google Scholar] [CrossRef]
- Lu, Y.; Lin, C.; Shan, C. Optoelectronic diamond: Growth, properties, and photodetection applications. Adv. Opt. Mater. 2018, 6, 1800359. [Google Scholar] [CrossRef]
- Su, T.; Xiao, B.; Ai, Z.; Bao, L.; Chen, W.; Shen, Y.; Cheng, Q.; Ostrikov, K. High-rate growth of gallium oxide films by plasma-enhanced thermal oxidation for solar-blind photodetectors. Appl. Surf. Sci. 2023, 624, 157162. [Google Scholar] [CrossRef]
- Tang, M.; Ma, C.; Liu, L.; Tan, X.; Li, Y.; Lee, Y.J.; Wang, G.; Jeon, D.W.; Park, J.H.; Zhang, Y.; et al. β-Ga2O3 Air-Channel Field-Emission Nanodiode with Ultrahigh Current Density and Low Turn-On Voltage. Nano Lett. 2024, 24, 1769–1775. [Google Scholar] [CrossRef]
- Kumar, S.; Dhara, S.; Agarwal, R.; Singh, R. Study of photoconduction properties of CVD grown β-Ga2O3 nanowires. J. Alloys Compd. 2016, 683, 143–148. [Google Scholar] [CrossRef]
- Wang, Z.; Zhang, G.; Zhang, X.; Wu, C.; Xia, Z.; Hu, H.; Wu, F.; Guo, D.; Wang, S. Polarization-Sensitive Artificial Optoelectronic Synapse Based on Anisotropic β-Ga2O3 Single Crystal for Neuromorphic Vision Systems and Information Encryption. Adv. Opt. Mater. 2025, 13, 2402238. [Google Scholar] [CrossRef]
- Wang, C.; Fan, W.-H.; Zhang, Y.-C.; Kang, P.-C.; Wu, W.-Y.; Wu, D.-S.; Lien, S.-Y.; Zhu, W.-Z. Effect of oxygen flow ratio on the performance of RF magnetron sputtered Sn-doped Ga2O3 films and ultraviolet photodetector. Ceram. Int. 2022, 49, 10634–10644. [Google Scholar] [CrossRef]
- Deng, L.; Hu, H.; Wang, Y.; Wu, C.; He, H.; Li, J.; Luo, X.; Zhang, F.; Guo, D. Surface plasma treatment reduces oxygen vacancies defects states to control photogenerated carriers transportation for enhanced self-powered deep UV photoelectric characteristics. Appl. Surf. Sci. 2022, 604, 154459. [Google Scholar] [CrossRef]
- Wang, H.; Wang, F.; Xu, T.; Xia, H.; Xie, R.; Zhou, X.; Ge, X.; Liu, W.; Zhu, Y.; Sun, L.; et al. Slowing Hot-Electron Relaxation in Mix-Phase Nanowires for Hot-Carrier Photovoltaics. Nano Lett. 2021, 21, 7761–7768. [Google Scholar] [CrossRef]
- Yang, Y.; Peng, X.; Kim, H.-S.; Kim, T.; Jeon, S.; Kang, H.K.; Choi, W.; Song, J.; Doh, Y.-J.; Yu, D. Hot Carrier Trapping Induced Negative Photoconductance in InAs Nanowires toward Novel Nonvolatile Memory. Nano Lett. 2015, 15, 5875–5882. [Google Scholar] [CrossRef]
- Jiang, M.; Golovynskyi, S.; Chen, J.; Yang, Z.; Lv, T.; Huang, G.; Sun, Z.; Li, L.; Wu, H.; Li, B. High-sensitive solar-blind β-Ga2O3 thin film photodetector deposited by PLD optimizing growth temperature. Vacuum 2025, 238, 114282. [Google Scholar] [CrossRef]
- Freitas, J.A.; Culbertson, J.C.; Nepal, N.; Mock, A.L.; Tadjer, M.J.; Feng, Z.; Zhao, H. Influence of oxygen partial pressure on properties of monoclinic Ga2O3 deposited on sapphire substrates. J. Vac. Sci. Technol. A 2021, 39, 033414. [Google Scholar] [CrossRef]
- Gajdics, M.; Cora, I.; Zámbó, D.; Horváth, Z.E.; Sulyok, A.; Frey, K.; Pécz, B. Evolution of structural and photoluminescent properties of sputter-deposited Ga2O3 thin films during post-deposition heat treatment. J. Alloys Compd. 2025, 1021, 179634. [Google Scholar] [CrossRef]
- Dong, Y.; Wang, Y.; Tian, X.; Feng, Q.; Zhang, J.; Hao, Y. First-principles study of Mg-Ge co-doping to realize p-type β-Ga2O3 containing divacancy-interstitial complex defects. Comput. Mater. Sci. 2025, 253, 113849. [Google Scholar] [CrossRef]
- Liu, M.; Kim, H.K. Ultraviolet detection with ultrathin ZnO epitaxial films treated with oxygen plasma. Appl. Phys. Lett. 2004, 84, 173–175. [Google Scholar] [CrossRef]
- Hu, H.; Wu, C.; Zhao, N.; Zhu, Z.; Li, P.; Wang, S.; Tang, W.; Guo, D. Epitaxial growth and solar-blind photoelectric characteristic of β-Ga2O3 film on various oriented sapphire substrates by plasma-enhanced chemical vapor deposition. Phys. Status Solidi Appl. Mater. Sci. 2021, 218, 2100076. [Google Scholar] [CrossRef]
- Le, A.H.; Kim, Y.; Lee, Y.J.; Hussain, S.Q.; Nguyen, C.P.; Lee, J.; Yi, J. On the origin of the changes in the opto-electrical properties of boron-doped zinc oxide films after plasma surface treatment for thin-film silicon solar cell applications. Appl. Surf. Sci. 2018, 433, 798–805. [Google Scholar] [CrossRef]
- Zhang, C.; Liu, K.; Ai, Q.; Huang, X.; Chen, X.; Zhu, Y.; Yang, J.; Cheng, Z.; Li, B.; Liu, L.; et al. Performance Enhancement of Ga2O3 Solar-Blind UV Photodetector by the Combination of Oxygen Annealing and Plasma Treatment. J. Phys. Chem. C 2022, 126, 21839–21846. [Google Scholar] [CrossRef]
- Qian, L.X.; Liu, H.Y.; Zhang, H.F.; Wu, Z.H.; Zhang, W.L. Simultaneously improved sensitivity and response speed of β-Ga2O3 solar-blind photodetector via localized tuning of oxygen deficiency. Appl. Phys. Lett. 2019, 114, 113506. [Google Scholar] [CrossRef]
- Wang, J.; Zhang, S.; Ji, X.; Li, M.; Zhang, J.; Liu, X.; Tian, R.; Lu, C.; Tang, W.; Li, P. High-performance Ga2O3 solar-blind UV photodetectors via film growth regulation and surface state engineering. Mater. Today Phys. 2024, 45, 101461. [Google Scholar] [CrossRef]
- Wang, Y.; Tang, Y.; Li, H.; Yang, Z.; Zhang, Q.; He, Z.; Huang, X.; Wei, X.; Tang, W.; Huang, W.; et al. p-GaSe/n-Ga2O3 van der Waals Heterostructure Photodetector at Solar-Blind Wavelengths with Ultrahigh Responsivity and Detectivity. ACS Photonics 2021, 8, 2256–2264. [Google Scholar] [CrossRef]
- Zhang, S.; Wang, J.; Ji, X.; Yan, Z.; Ye, L.; Zheng, H.; Liu, Y.; Chen, X.; Li, P. Oxygen vacancies modulating performance for Ga2O3 solar-blind photodetectors via low-cost mist chemical vapor deposition. Mater. Today Commun. 2024, 39, 108717. [Google Scholar] [CrossRef]
- Zhang, T.; Li, Y.; Feng, Q.; Zhang, Y.; Ning, J.; Zhang, C.; Zhang, J.; Hao, Y. Effects of growth pressure on the characteristics of the β-Ga2O3 thin films deposited on (0001) sapphire substrates. Mater. Sci. Semicond. Process. 2021, 123, 105572. [Google Scholar] [CrossRef]
- Mi, W.; Li, Z.; Luan, C.; Xiao, H.; Zhao, C.; Ma, J. Transparent conducting tin-doped Ga2O3 films deposited on MgAl2O4 (100) substrates by MOCVD. Ceram. Int. 2015, 41, 2572–2575. [Google Scholar] [CrossRef]
- Patil, V.; Lee, B.-T.; Jeong, S.-H. Optical and structural characterization of high crystalline β-Ga2O3 films prepared using an RF magnetron sputtering. J. Alloys Compd. 2022, 894, 162551. [Google Scholar] [CrossRef]
- Xiao, B.; Liu, B.; He, X.; Li, C.; Liang, Z.; Sun, Y.; Cheng, Q. Plasma surface treatment of amorphous Ga2O3 thin films for solar-blind ultraviolet photodetectors. Appl. Surf. Sci. 2024, 678, 161146. [Google Scholar] [CrossRef]
- Wang, L.; Xu, S.; Yang, J.; Huang, H.; Huo, Z.; Li, J.; Xu, X.; Ren, F.; He, Y.; Ma, Y.; et al. Recent Progress in Solar-Blind Photodetectors Based on Ultrawide Bandgap Semiconductors. ACS Omega 2024, 9, 25429–25447. [Google Scholar] [CrossRef]
- Veeralingam, S.; Durai, L.; Yadav, P.; Badhulika, S. Record-High Responsivity and Detectivity of a Flexible Deep-Ultraviolet Photodetector Based on Solid State-Assisted Synthesized hBN Nanosheets. ACS Appl. Electron. Mater. 2021, 3, 1162–1169. [Google Scholar] [CrossRef]
- Zeng, Y.; Huang, H.; Zhao, X.; Ding, M.; Hou, X.; Zou, Y.; Du, J.; Liu, J.; Yu, S.; Han, K.; et al. Self-Powered a-SnOX/c-Ga2O3 Pn Heterojunction Solar-Blind Photodetector with High Responsivity and Swift Response Speed. IEEE Electron Device Lett. 2023, 44, 2003–2006. [Google Scholar] [CrossRef]
- Ji, X.Q.; Liu, M.Y.; Yan, Z.Y.; Li, S.; Liu, Z.; Qi, X.H.; Yuan, J.Y.; Wang, J.J.; Zhao, Y.C.; Tang, W.H.; et al. Ultrasensitive and High-Speed Ga2O3 Solar-Blind Photodetection Based on Defect Engineering. IEEE Trans. Electron Devices 2023, 70, 4236–4242. [Google Scholar] [CrossRef]
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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).