Introduction
During the past decade, due to extensive application in modern industrial agriculture and military intelligence integration development, transistors,[ ] photodetectors,[ ] light-emitting diodes,[ ] and other power electronics have attracted widespread research.[ ] Photodetectors are photoelectronic devices that convert optical signals (UV-IR) into electrical signals. The emergence of wide bandgap semiconductors has resulted in solar-blind deep-ultraviolet (SBDU) photodetectors (PDs) with extremely low background noise receiving considerable attention.[ ] In general, SBDU PDs are constructed of semiconductor materials with a bandwidth greater than 4.4 eV, including Al x Ga x−1N,[ ] ZnMgO,[ ] SiC,[ ] Ga2O3,[ ] and diamond.[ ] A new type of ultra-wide band semiconductor material, Ga2O3 with a wide bandgap of ≈4.9 eV, does not require a complex alloying process and possesses high chemical stability, and can therefore be considered a natural material for solar-blind signals detection. To date, various types of Ga2O3-based PDs have been proposed, including p-n junctions,[ ] heterojunctions,[ ] Schottky diodes,[ ] avalanches,[ ] metal–semiconductor–metal (MSM)[ ] and photoelectrochemical[ ] PDs. Among them, MSM-type SBDU PDs have been widely studied as their preparation is simple and they are easily integrated with electronic devices. Moreover, MSM-type Ga2O3-based PDs maintain a high performance with a low dark current, large light-to-dark ratio, excellent responsivity, and a fast photoresponse time.[ ] These high-performance and easy-to-integrate MSM-type Ga2O3-based PDs present great potential in civil and military applications, including medical imaging, secure optical communication, environmental and fire monitoring, space exploration, and high-voltage leakage detection.[ ]
As early as 2018, Chen et al. prepared an MSM-type β-Ga2O3 film-based PD array (4 × 4) to obtain clear target images in solar-blind imaging.[ ] When the bias voltage is 45 V, the peak responsivity and detectivity of the PD array unit are ≈12.4 A W−1 and 1.9 × 1012 Jones, respectively. Recently, Qian et al. fabricated a large-scale high-uniformity 32 × 32 image sensor array, based on amorphous-Ga2O3 (a-Ga2O3) SBDU PDs.[ ] The MSM-type a-Ga2O3 SBDU PD exhibits a super-high sensitivity of 733 A W−1, fast photoresponse speed of 18 ms, ultra-high photo-to-dark current ratio of 3.9 × 107, and an excellent specific detectivity of 3.9 × 1016 Jones. Optical communication systems typically rely on light radiation to transmit information. Compared with traditional radio-frequency (RF) wireless communication, optical communication light signals easily bypass various terrain features by scattering and reflecting without interference, and thus can provide high-speed, ultra-low latency, green, and low-cost communication services.[ ] In 2019, Lin et al. designed a high-performance solar-blind photodetector using a diamond wafer.[ ] The device responsivity and response time are 13 A W−1 and 1.3 μs/203 μs, respectively. Based on this, the authors used the novel diamond photodetector as the signal receiver to construct a solar-blind deep-ultraviolet communication (SDUC) system. Very recently, our group built a high-performance β-Ga2O3-based solar-blind photodetector with a remarkable dark current of 82 fA and fast photoresponse speed of 11/240 μs for the purposes of solar-blind deep-ultraviolet communication.[ ] It is the first time that a signal receiver consisting of a β-Ga2O3-based PD integrated into a tailor-made SDUC system has been verified to transmit American Standard Code for Information Interchange (ASCII) codes. This proves the validity and applicability of β-Ga2O3 solar-blind deep-ultraviolet photodetectors for future use in secure communications. Among all the commonly identified polymorphs of Ga2O3, ε-Ga2O3, a metastable phase is regarded as the hexagon system belonging to the P63mc space group with high symmetry and low anisotropy crystal structure, which has unique advantages in the construction of optoelectronic devices.[ ] However, there is no known research regarding the application of ε-Ga2O3 film-based photodetector arrays for solar-blind imaging and deep-ultraviolet communication.
Generally, more stable metals with excellent electrical conductivity, such as Ti, Au, Ag, and Pt, are used as PD electrode materials. Due to the high cost and opaque nature of PDs fabricated from these metal electrodes, they cannot be applied in next-generation transparent optoelectronic technologies, such as secure communications, smart windows, artificial intelligence glasses, and light field cameras.[ ] Therefore, transparent conductive oxides (TCOs) with high electrical conductivity and transparency have attracted widespread interest.[ ] Compared to commonly used noble metals, TCOs are not only transparent, but are also an abundant earth resource, which, therefore, reduces costs. Moreover, due to the non-stoichiometric properties of the components, the optoelectronic properties of TCOs thin films can be adjusted over a wide range by doping and optimizing the preparation process.[ ] These inherent characteristics make TCOs ideal for large-scale optoelectronics industries. For example, Wu et al. prepared an organic photodetector with a remarkable external quantum efficiency of 74%, an excellent responsivity of 0.64 A W−1, and a satisfactory detectivity of 6.6 × 1012 Jones, by using indium-doped zinc oxide electrodes.[ ]
This study demonstrates three ε-Ga2O3-based SBDU photodetector arrays (PDAs), with different transparent conductive oxide electrodes, including indium tin oxide (ITO), indium gallium zinc oxide (IGZO), and aluminum zinc oxide (AZO). High-uniformity 5 × 4 arrays of fully transparent MSM-type SBDU PDs, based on post-annealed ε-Ga2O3 films, were constructed by radio frequency (rf) magnetron sputtering and masking processes. The photoelectric properties of the three fully transparent PDAs were systematically studied, and it is found that all three PDAs not only exhibit ultra-high responsivity, remarkable detectivity, and excellent stability, but also ultra-fast photoresponse time. Finally, we explored the potential applications of three ε-Ga2O3-based SBDU PDAs in solar-blind imaging and deep-ultraviolet communication.
Results and Discussion
Figure shows the X-ray diffraction (XRD) pattern of the post-annealed ε-Ga2O3 film. It can be seen from the XRD pattern that three diffraction peaks located at 19.33°, 39.2°, and 60.2° correspond with the (002), (004), and (006) crystal planes of the ε-Ga2O3, respectively.[ ] Apart from the three ε-Ga2O3 diffraction peaks (001), no other diffraction peaks are evident, indicating that the film annealed by argon at high temperature is a single-phase hexagonal gallium oxide. Figure exhibits the ε-Ga2O3 film absorption spectrum, where it is obvious that the film absorption rate in the 300–800 nm band is almost zero, which is a powerful prerequisite for the next step in the preparation of a fully transparent device. Furthermore, the film has a steep absorption edge of around 280 nm, and the ε-Ga2O3 film optical bandgap is estimated to be approximately 4.9 eV from the Tauc plot.[ ] In this study, three transparent electrodes of ITO, AZO, and IGZO are used to fabricate fully transparent MSM-type SBDU PDAs (5 × 4). The structure of a single PD is shown in the inset of Figure . At the same time, a surface scan X-ray energy dispersion diagram (EDS) of the ITO electrode single PD was obtained, as shown in Figure . Figure displays the specific element mapping, showing that the PD is composed of ε-Ga2O3 thin film photosensitive material constituting a Ga/O element and ITO electrode material consisting of In/Sn/O elements. It is apparent that the In/Sn element map possesses a good boundary shape, indicating that the electrodes deposited by magnetron sputtering have good uniformity.[ ]
[IMAGE OMITTED. SEE PDF]
To initially explore and compare the effects of different electrodes on the photoelectric properties of ε-Ga2O3 films, three different fully transparent MSM-type PDAs (5 × 4) with ITO, IGZO, and AZO electrodes were fabricated, and a single PD was selected for optoelectronic performance evaluation. Figure illustrates the current–voltage (I–V) characteristic curves of the three PDs in the dark and under 254 nm illumination with a light intensity of 1000 μw cm−2. Surprisingly, the photocurrent of the PD with ITO electrodes (ITO-PD) reached 3.3 × 10−4 A at 10 V bias, and the photocurrent of the other two PDs also achieved 1.03 × 10−4 and 3.37 × 10−5 A, respectively. Meanwhile, the photo-dark ratio of all three PDs is greater than 5 × 104 at 10 V bias, and the photo-dark current ratio (PDCR) of PD with AZO electrodes (AZO-PD) is the largest, attaining 2.6 × 105. Figure depicts the single-period transient photoresponse characteristic curves (I–T) of the three PDs. The photocurrent of the SBDU PD increases or decays rapidly when the 254 nm light is switched on or off. Generally, the ability of a PD to track this fast-changing light signal is called the photoresponse speed. Using the second-order exponential equation to fit the transient photoresponse curves, the photoresponse speed of the three PDs can be quantitatively compared. The equation is as follows[
]
[IMAGE OMITTED. SEE PDF]
To further estimate and compare the optoelectronic performance of the three fully transparent MSM-type PDs, more systematic measurement and performance calculations were performed, as shown in Figure . Figure displays the I–V curves of the PDs with ITO, AZO, and IGZO electrodes under 254 nm illumination and dark conditions, respectively. As seen, the I–V curves of AZO-PD and IGZO-PD display straight linear relationships, indicating that AZO/IGZO electrodes and ε-Ga2O3 film are good ohmic contacts, while ITO electrodes and ε-Ga2O3 film are back-to-back Schottky contacts. Meanwhile, the photocurrent of all three PDs increases with increased light intensity, which is related to the high light intensity promoting the generation of electron–hole pairs. The relationship between the specific photocurrent and light intensity at 10 V bias is shown in Figure . In general, this relationship conforms to the power-law function: ,[ ] where P is the light intensity and θ is an exponent corresponding to the photogenerated carrier recombination. After fitting, it is obvious that the ITO-PD results in an exponent of 0.83, while the fitted curves of the IGZO-PD and AZO-PD attain two different exponent values: θ = 0.77/0.53 for low optical intensity density and θ = 0.42/0.28 for high optical intensity, respectively. These are slightly smaller than the ideal value (θ = 1). These results demonstrate that all devices suffer recombination losses, among which the recombination loss of ITO-PD is slight, while the recombination losses of IGZO-PD and AZO-PD are relatively larger. This may be related to defects formed by the electrode material and ε-Ga2O3 film.[ ] Due to the increased recombination activity of photo-generated carriers as the light intensity increases, and a higher defect density existing between the electrodes and films compared to the ITO electrode, both IGZO-PD and AZO-PD have two fitting index values which are relatively smaller at high light intensities.[ ] In addition to the photocurrent, the PDs PDCR also depends on the dark current and light intensity. The PDCR of all PDs augments with increasing light intensity, which corresponds to the change law of photocurrent with light intensity, as shown in Figure . However, the AZO-PD PDCR is the largest, due to the relatively small dark current, and is as high as 2.6 × 105 under 254 nm light with an intensity of 1000 μW cm−2.
[IMAGE OMITTED. SEE PDF]
To quantitatively compare the device performance of the three fully transparent ε-Ga2O3 film-based solar-blind PDs, the PD responsivity (R) and detectivity (D
*) under different light intensities at 10 V bias were calculated, as shown in Figure . R is defined as the effective photocurrent generated by incident light, per unit of light power, over the photodetector effective area, and can be expressed as[
]
From the aforementioned, it is clear that the PD performance is basically related to photocurrent and response time. To explore the internal reasons for the small photocurrent and slow photoresponse time of the AZO-PD in the three PDs under the same conditions, the basic parameters of the three electrode materials were investigated (Table S1, Supporting Information). It is evident that the resistivity of the AZO electrode material reaches as high as 7.54 Ω cm, while the resistivity of the ITO and IGZO materials is only 0.034 and 0.068 Ω cm, respectively. In addition, the AZO electrode material carrier mobility is relatively low, at only 0.089 cm2 V s−1, while the ITO and IGZO materials carrier mobilities are 2.9 and 8.6 cm2 V s−1, respectively. Materials with low resistivity and high carrier mobility should be selected as the electrode material for collecting and transporting carriers. However, due to the AZO electrode material resistivity and carrier mobility being poor in this study, among the three PDs, the photocurrent of the AZO-PD is relatively small. Generally, the photoresponse time of a PD is not only dependent on the rapid generation and annihilation of photogenerated carriers, but is also related to the trapping/releasing of photogenerated carriers by defect states.[ ] It can be seen from Figure that the recombination loss of the AZO-PD is the largest of the PDs, which relates to a large number of defect states between the electrode and the ε-Ga2O3 film. These defect states may act as efficient photo-generated carriers trapping/releasing states, resulting in a slow PD photoresponse time.
To demonstrate the ultra-high performance of the three fully transparent solar-blind PDs, Table shows the key parameters of MSM-type Ga2O3 PDs both in previous reports, and in this research. As shown in the table, all the PDs we prepared exhibit excellent performance, including a record ultra-high responsivity of more than 250 A W−1 and ultra-fast millisecond level photoresponse time in MSM-type ε-Ga2O3 PDs. Furthermore, in contrast to previously reported PDs with noble metal electrodes, all three PDs in this work were prepared from low-cost transparent conductive oxides. Overall, the proposed fully transparent MSM-type PDs possess significant advantages of excellent performance, easy integration, and low cost, making them a promising option for next-generation fully transparent optoelectronic devices.
Table 1 Comparison of key parameters of MSM-type Ga2O3 solar-blind PDs
Materials | Array | electrode material | R [A W−1] | D* [Jones] | τ r/τ d [ms] | References |
a-Ga2O3 | No | Cr | 1430@12 V | 3.16 × 1015 | >1000 | [] |
a-Ga2O3 | Yes | Ti/Au | 733@5 V | 3.9 × 1016 | 1/18 | [] |
a-Ga2O3 | No | Cr/Au | 96.1@5 V | / | 31/78 | [] |
a-Ga2O3 | Yes | Au | 8.9@12 V | 3.3 × 1013 | 0.015/0.308 | [] |
α-Ga2O3 | No | Ti/Au | 11.5@12 V | 1.0 × 1015 | /42 | [] |
α-Ga2O3 | No | Ni/Au | 3.36@12 V | / | / | [] |
β-Ga2O3 | No | Au | 9500@5 V | 1.5 × 1015 | 10/10 | [] |
β-Ga2O3 | No | Ni/Au | 1.5@20 V | / | 3330/400 | [] |
ε-Ga2O3 | No | Ti/Au | 1.368@5 V | 9.13 × 1014 | 61.1/87.5 | [] |
ε-Ga2O3 | No | Ti/Au | 230@6 V | 1.2 × 1015 | /24 | [] |
ε-Ga2O3 | Yes | ITO | 286.2@10 V | 4.73 × 1014 | 5.6/7.2 | This work |
ε-Ga2O3 | Yes | AZO | 262.1@10 V | 2.16 × 1015 | 70/200 | This work |
ε-Ga2O3 | Yes | IGZO | 284.1@10 V | 5.06 × 1014 | 6.9/9.5 | This work |
Subsequently, we explored the potential applications of large-area fully transparent ε-Ga2O3 film-based photodetector arrays in solar-blind imaging, as shown in the Figure schematic diagram. The hollow mask engraved with “TCO” characters was irradiated under 254 nm light with a 400 μW cm−2 light intensity, and three fully transparent PD arrays were placed under the mask. The ITO-PD, AZO-PD, and IGZO-PD arrays were shielded by three masks, namely “T”, “C”, and “O” characters, respectively. When solar-blind deep-ultraviolet light passes through the mask onto the PD array, the remaining PDs are kept in the dark or under weak UV light. Keeping the test system still, the current of each PD unit was recorded cell by cell through a pair of probes connected to the semiconductor parameter analyzer. The array uniformity is a basic requirement of integrated applications, and the uniformity of the three arrays is further discussed before measuring the solar-blind imaging. Figure describes the dark current of each PD in an ITO-PD array. Obviously, the dark current of most PDs is about 10−9 A in magnitude, but the dark current at the corner of the three PD units is relatively large, which may be due to irreversible damage to the ε-Ga2O3 film during cutting in the later stages. Moreover, the dark current of PDs units in the AZO-PD and IGZO-PD arrays were also certified in dark conditions, as shown in Figure . Two fully transparent arrays exhibit excellent uniformity, and dark currents of ≈10−10 and 10−9 A magnitude, respectively. Meanwhile, the aforementioned method was used to conduct solar-blind imaging tests. Figure shows a 2D dark current comparison diagram. Surprisingly, all arrays display clearer “T”, “C”, and “O” character shapes, indicating that these fully transparent solar-blind PD arrays have high-fidelity characteristics and have excellent application prospects in solar-blind imaging and machine vision.[ ]
[IMAGE OMITTED. SEE PDF]
In recent years, with the introduction of a new generation of 5G broadband mobile communication technology, wireless spectrum resources have become increasingly scarce. Solar-blind ultraviolet communication (SBUC) not only demonstrates electromagnetic interference resistance and ultra-high signal-to-noise ratio, but also massively expands the communication spectrum. Due to the fast photoresponse speed and ultra-high responsivity of the three PDs, fully transparent ε-Ga2O3-based solar-blind PDs were prepared as optical signal receivers of an SBUC system. Figure exhibits a flowchart of the solar-blind ultraviolet communication principle. First, the text data is converted into control commands by the driver (EP4CE6F17C8) to control turning the metal oxide semicoductor field effect transistor (MOSFET) (LR7843) on/off, corresponding to the on/off of the light emitting diode (LED) light source. Subsequently, the intermittent light (255.5 nm) from the LED is detected by the solar-blind photodetector and outputs high/low-level current signals. The photocurrent signal is then converted into a voltage signal and further amplified by the differential voltage preamplifier. The differential voltage signal input to the driver (analog-to-digital converter chip) converts it to a digital signal.
[IMAGE OMITTED. SEE PDF]
To evaluate the data transmission ability and stability of the SBUC system, three PD optical signal receivers were used to try to transmit the text data of three letters “TCO”. First, “TCO” is converted into ASCII by the program. The code only contains the digits “1” and “0”. The driver (EP4CE6F17C8) then converts the “1” and “0” into the control command corresponding to the MOSFET (LR7843) on/off function. To be able to see the digital signal before conversion, the corresponding real-time images displayed on the oscilloscope, are outputted under different modulation frequencies. Figure shows the waveform signals of the three PDs SBUC systems at a transmission speed of 0.01 kbps. All the waveforms showed the same response as the digital “TCO” signal, confirming the suitability of all three devices as SBUC system receiving signals. However, at the same transmission frequency, the SDUC system based on the ITO-PD transmission waveform is the most obvious and orderly, indicating that this SDUC system transmission limit is the highest along with its late decoding accuracy. Comparing the relationship between the performance of three PDs and the SDUC system transmission, it is clear that the transmission speed of the SDUC system is directly related to the PD photoresponse speed as the signal receiving component. Therefore, the development of a PD with a faster photoresponse speed would be an effective way to increase the SDUC transmission rate. Alternatively, further optimizing the differential voltage preamplifier could also further increase the transmission rate.[ ]
Conclusion
In summary, we demonstrate for the first time that transparent conductive oxides are potential excellent electrode materials to obtain high-performance ε-Ga2O3-based solar-blind deep-ultraviolet photodetectors. A proof-of-concept of the solar-blind imaging and deep-ultraviolet communication system was verified using three fully transparent ε-Ga2O3-based arrays, composed of easily-integrated MSM-type photodetectors. The results show that all three PDAs exhibit not only ultra-high responsivity, remarkable detectivity, and excellent stability, but also an ultra-fast photoresponse time. Finally, for the first time, we explored the potential applicability of three ε-Ga2O3-based PDAs in solar-blind imaging and deep-ultraviolet communication. Our research presents a method to fabricate high-performance, fully transparent, solar-blind, deep-ultraviolet PDAs and highlights their applicability in the next generation of fully transparent optoelectronics.
Experimental Section
Material Processing and Device Fabrication
The commercially available ε-Ga2O3 thin films from Beijing Gallium Group were selected, which were deposited on sapphire substrates by metal-organic chemical vapor deposition (MOCVD). Subsequently, ε-Ga2O3 films were annealed in a tube furnace at 500 °C for 2 h in argon. Three transparent electrode materials, including ITO, IGZO, and AZO, were then sputtered on the surface of ε-Ga2O3 films by magnetron sputtering and a masking process, respectively. High-uniformity 5 × 4 arrays of fully transparent MSM-type SBDU PDs were constructed, in which a single PD consisted of 5 pairs of interdigital electrodes with a length of 800 μm and a spacing of 30 μm. During the sputtering process, the sputtering chamber base vacuum pressure and argon flow rate were 5.0 × 10−4 Pa and 40 sccm, respectively. The working pressure, sputtering power, and sputtering time were 2.5/0.8/2.5 Pa, 120/50/120 W, and 10/15/10 min, respectively.
Material and Device Characterization
The crystal structures of ε-Ga2O3 films were analyzed by an X-ray diffractometer (Bruker D8 ADVANCE A25X) using Cu Kα lines (λ = 0.154 nm). The ε-Ga2O3 films absorption spectra were tested using a UV–vis–NIR spectrophotometer (U-4100). The electrical properties of the three transparent electrode materials were obtained by a Hall effect measurement system (HMS-3000, Ecopia). The surface morphology and elemental composition were obtained by field emission scanning electron microscopy (Thermofisher Apreo 2C). The photoelectric characteristics of ε-Ga2O3-based PDs, and the corresponding solar-blind imaging analysis, were measured by a Keithley 2450 source meter. A MOS5072 oscilloscope was applied to obtain the deep-ultraviolet communication system voltage signals. All characterization was conducted at room temperature.
Acknowledgements
The authors gratefully acknowledge support from the National Natural Science Foundation of China (Grant Nos. 11904041, 12104077), the Natural Science Foundation of Chongqing (Grant No. cstc2019jcyj-msxmX0237, cstc2020jcyj-msxmX0533, cstc2020jcyj-msxmX0557), the Science and Technology Research Project of Chongqing Education Committee (Grant Nos. KJQN201900542, KJQN202000511, KJQN202100501, KJQN202100540), and the College Students Innovation and Entrepreneurship Training Program of Chongqing City (Grant Nos. S202110637123, S202110637119, S202110637130, S202210637046, 2020051101165).
Conflict of Interest
The authors declare no conflict of interest.
Data Availability Statement
Research data are not shared.
Z. Y. Han, H. L. Liang, W. X. Huo, X. S. Zhu, X. L. Du, Z. X. Mei, Adv. Opt. Mater. 2020, 8, 1901833.
W. Y. Kong, G. A. Wu, K. Y. Wang, T. F. Zhang, Y. F. Zou, D. D. Wang, L. B. Luo, Adv. Mater. 2016, 28, 10725.
J. H. Kim, C. Y. Han, K. H. Lee, K. S. An, W. Song, J. Kim, M. S. Oh, Y. R. Do, H. Yang, Chem. Mater. 2014, 27, 197.
J. X. Chen, W. X. Ouyang, W. Yang, J. H. He, X. S. Fang, Adv. Opt. Mater. 2020, 30, 1909909.
X. Chao, F. Yan, Small 2017, 13, 1701822.
X. H. Chen, F. F. Ren, S. L. Gu, J. D. Ye, Photonics Res. 2019, 7, 381.
J. J. Xu, W. Zheng, F. Huang, J. Mater. Chem. C 2019, 7, 8753.
D. H. Wang, C. Huang, X. Liu, H. C. Zhang, H. B. Yu, S. Fang, B. S. Ooi, Z. Mi, J. H. He, H. Sun, Adv. Opt. Mater. 2021, 9, 2000893.
M. M. Fan, K. W. Liu, Z. Z. Zhang, B. H. Li, X. Chen, D. X. Zhao, C. X. Shan, D. Z. Shen, Appl. Phys. Lett. 2014, 105, 011117.
S. P. Karanth, M. A. Sumesh, V. Shobha, J. Sirisha, D. Yadav, S. B. Vijay, K. V. Sriram, IEEE Trans. Electron Devices 2020, 67, 242.
Y. Xu, X. H. Chen, Y. F. Zhang, F. F. Ren, S. L. Gu, J. D. Ye, IEEE Electron Device Lett. 2020, 41, 997.
S. J. Pearton, J. C. Yang, P. Cary, F. Ren, J. Kim, M. J. Tadjer, M. A. Mastro, Appl. Phys. Rev. 2018, 5, 301.
S. Oh, C. K. Kim, J. Kim, ACS Photonics 2018, 5, 1123.
C. N. Lin, Y. J. Lu, X. Yang, Y. Z. Tian, C. J. Gao, J. L. Sun, L. Dong, F. Zhong, W. D. Hu, C. X. Shan, Adv. Opt. Mater. 2018, 6, 1800068.
H. B. Wang, H. Y. Chen, L. Li, Y. F. Wang, L. X. Su, W. P. Bian, B. S. Li, X. S. Fang, J. Phys. Chem. Lett. 2019, 10, 6850.
Y. Q. Zhang, Y. F. Wang, R. P. Fu, J. G. Ma, H. Y. Xu, B. S. Li, Y. C. Liu, J. Phys. D: Appl. Phys. 2022, 55, 324002.
R. C. Lin, W. Zheng, D. Zhang, Z. J. Zhang, Q. X. Liao, L. Yang, F. Huang, ACS Appl. Mater. Interfaces 2018, 10, 22419.
T. Zhang, Y. X. Shen, Q. Feng, X. S. Tian, Y. C. Cai, Z. Z. Hu, G. S. Yan, Z. Q. Feng, Y. C. Zhang, J. Ning, Y. Xu, X. Lian, X. Sun, C. Zhang, H. Zhou, J. Zhang, Y. Hao, Nanoscale Res. Lett. 2020, 15, 1.
X. H. Chen, Y. Xu, D. Zhou, S. Yang, F. F. Ren, H. Lu, K. Tang, S. L. Gu, R. Zhang, Y. D. Zheng, J. D. Ye, ACS Appl. Mater. Interfaces 2017, 9, 36997.
Z. X. Jiang, Z. Y. Wu, C. C. Ma, J. N. Deng, H. Zhang, Y. Xu, J. D. Ye, Z. L. Fang, G. Q. Zhang, J. Y. Kang, T. Y. Zhang, Mater. Today Phys. 2020, 14, 100226.
J. Wang, L. J. Ye, X. Wang, H. Zhang, L. Li, C. Y. Kong, W. J. Li, J. Alloys Compd. 2019, 803, 9.
L. J. Huang, Z. R. Hu, H. Zhang, Y. Q. Xiong, S. Q. Fan, C. Y. Kong, W. J. Li, L. J. Ye, H. L. Li, J. Mater. Chem. C 2021, 9, 10354.
C. Xie, X. T. Lu, Y. Liang, H. H. Chen, L. Wang, C. Y. Wu, D. Wu, W. H. Yang, L. B. Luo, J. Mater. Sci. Technol. 2021, 72, 189.
B. Zhao, F. Wang, H. Y. Chen, Y. P. Wang, M. M. Jiang, X. S. Fang, D. X. Zhao, Nano Lett. 2015, 15, 3988.
D. Kaur, M. Kumar, Adv. Opt. Mater. 2021, 9, 2002160.
H. W. Liu, S. R. Zhou, H. Zhang, L. J. Ye, Y. Q. Xiong, P. Yu, W. J. Li, X. Yang, H. L. Li, C. Y. Kong, J. Phys. D: Appl. Phys. 2022, 55, 305104.
Y. C. Chen, Y. J. Lu, Q. Liu, C. N. Lin, J. Guo, J. H. Zang, Y. Z. Tian, C. X. Shan, J. Mater. Chem. C 2019, 7, 2557.
Y. Qin, L. H. Li, Z. A. Yu, F. H. Wu, D. N. Dong, W. Guo, Z. F. Zhang, J. H. Yuan, K. H. Xue, X. S. Miao, S. B. Long, Adv. Sci. 2021, 8, 101106.
C. N. Lin, Y. J. Lu, Y. Z. Tian, C. J. Gao, M. M. Fan, X. Yang, L. Dong, C. X. Shan, Opt Express 2019, 27, 29962.
S. R. Zhou, X. Peng, H. W. Liu, Z. F. Zhang, L. J. Ye, H. L. Wen, Y. Q. Xiong, L. B. Niu, F. L. Chen, L. Fang, C. Y. Kong, W. J. Li, X. Yang, H. Zhang, Opt. Mater. Express 2022, 12, 327.
S. Li, J. Y. Yue, X. Q. Ji, C. Lu, Z. Y. Yan, P. G. Li, D. Y. Guo, Z. P. Wu, W. H. Tang, J. Mater. Chem. C 2021, 9, 5437.
S. Ju, A. Facchetti, Y. Xuan, J. Liu, F. Ishikawa, P. D. Ye, C. W. Zhou, T. J. Marks, D. B. Janes, Nat. Nanotechnol. 2007, 2, 378.
H. Kim, K. R. Chauhan, J. Kim, E. H. Choi, Appl. Phys. Lett. 2017, 110, 101907.
J. Vidal, F. Trani, F. Bruneval, M. Marques, S. Botti, Phys. Rev. Lett. 2010, 104, 136401.
J. L. Wu, L. H. Lai, Y. T. Hsiao, K. W. Tsai, C. M. Yang, Z. W. Sun, J. C. Hsieh, Y. M. Chang, Adv. Opt. Mater. 2021, 10, 202101723.
Y. C. Oshima, E. Víllora, Y. Matsushita, S. Yamamoto, K. Shimamura, J. Appl. Phys. 2015, 118, 085301.
T. Wang, H. L. Liang, Z. Y. Han, Y. X. Sui, Z. X. Mei, Adv. Mater. Technol. 2021, 6, 2000945.
H. Kan, W. Zheng, H. Chen, R. C. Lin, J. T. Luo, F. Huang, ACS Appl. Mater. Interfaces 2020, 12, 6030.
D. Y. Guo, Y. L. Su, H. Z. Shi, P. G. Li, N. Zhao, J. H. Ye, S. L. Wang, A. P. Liu, Z. W. Chen, C. R. Li, W. H. Tang, ACS Nano 2018, 12, 12827.
L. X. Qian, Z. H. Wu, Y. Y. Zhang, P. T. Lai, X. Z. Liu, Y. R. Li, ACS Photonics 2017, 4, 2203.
D. Y. Guo, Z. P. Wu, Y. H. An, X. C. Guo, X. L. Chu, C. L. Sun, L. H. Li, P. G. Li, W. H. Tang, Appl. Phys. Lett. 2014, 105, 023507.
C. Xie, X. T. Lu, M. R. Ma, X. W. Tong, Z. X. Zhang, L. Wang, C. Y. Wu, W. H. Yang, L. B. Luo, Adv. Opt. Mater. 2019, 7, 1901257.
Z. X. Zhang, L. H. Zeng, X. W. Tong, Y. Gao, C. Xie, Y. H. Tsang, L. B. Luo, Y. C. Wu, J. Phys. Chem. Lett. 2018, 9, 1185.
L. B. Luo, G. A. Wu, Y. Gao, L. Liang, C. Xie, Z. X. Zhang, X. W. Tong, T. Wang, F. X. Liang, Adv. Opt. Mater. 2019, 7, 1900272.
K. Arora, N. Goel, M. Kumar, M. Kumar, ACS Photonics 2018, 5, 2391.
X. H. Hou, H. D. Sun, S. B. Long, G. Tompa, T. Salagaj, Y. Qin, Z. F. Zhang, P. Tan, S. J. Yu, M. Liu, IEEE Electron Device Lett. 2019, 40, 1483.
S. Li, D. Y. Guo, P. G. Li, X. Wang, Y. H. Wang, Z. Y. Yan, Z. Liu, Y. S. Zhi, Y. Q. Huang, Z. P. Wu, W. H. Tang, ACS Appl. Mater. Interfaces 2019, 11, 35105.
Y. C. Chen, Y. J. Lu, M. Y. Liao, Y. Z. Tian, Q. Liu, C. J. Gao, X. Yang, C. X. Shan, Adv. Funct. Mater. 2019, 29, 1906040.
G. Qiao, Q. Cai, T. C. Ma, J. Wang, X. H. Chen, Y. Xu, Z. G. Shao, J. D. Ye, D. J. Chen, ACS Appl. Mater. Interfaces 2019, 11, 40283.
A. S. Pratiyush, S. Krishnamoorthy, S. V. Solanke, Z. B. Xia, R. Muralidharan, S. Rajan, D. Nath, Appl. Phys. Lett. 2017, 110, 221107.
Y. Qin, L. H. Li, X. L. Zhao, G. Tompa, H. Dong, G. Z. Jian, Q. M. He, P. J. Tan, X. H. Hou, Z. F. Zhang, S. J. Yu, H. D. Sun, G. W. Xu, X. S. Miao, K. H. Xue, S. B. Long, M. Liu, ACS Photonics 2020, 7, 812.
G. B. Cen, Y. J. Liu, C. X. Zhao, G. H. Wang, Y. Fu, G. H. Yan, Y. Yuan, C. H. Su, Z. J. Zhao, W. J. Mai, Small 2019, 15, 1902135.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Copyright John Wiley & Sons, Inc. 2022
Abstract
With the development of new broadband semiconductor materials and the expansion of solar‐blind deep‐ultraviolet (SBDU) optoelectronics, the demand for high‐performance fully transparent array integration has also increased. Herein, metal–semiconductor–metal (MSM)‐type fully transparent SBDU photodetector arrays (PDAs), based on commercially available ε‐Ga2O3 films, are constructed using indium tin oxide (ITO), indium gallium zinc oxide (IGZO), and aluminum zinc oxide (AZO) transparent conductive oxides as electrodes. The experimental results show that the overall performance of the fully transparent SBDU PDAs used in this study places them at the forefront of ε‐Ga2O3‐based photodetectors (PDs). All three PDs not only exhibit ultra‐high responsivity (286.2, 284.1, and 262.1 A W−1), remarkable detectivity (4.73 × 1014, 5.06 × 1014, and 2.16 × 1015 Jones), and excellent stability, but also fast photoresponse time (5.6/7.2, 6.9/9.5, and 70/200 ms). It is found that the resistivity and carrier concentration of transparent conductive oxides significantly influence device performance. Consequently, it is demonstrated, for the first time, that fully transparent ε‐Ga2O3‐based SBDU PDAs can be used as ultraviolet imagers and signal receivers integrated into a homemade solar‐blind ultraviolet communication (SBUC) system to transmit text data. These results indicate great potential for future applications of fully transparent ε‐Ga2O3 thin‐film solar‐blind photodetectors in next‐generation artificial intelligence eyes, smart windows, light field cameras, and solar‐blind imaging, among others.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Chongqing Key Laboratory of Photo-Electric Functional Materials, College of Physics and Electronic Engineering, Chongqing Normal University, Chongqing, P. R. China
2 School of Automation, Chongqing University of Posts and Telecommunications, Chongqing, P. R. China