Next Article in Journal
A Novel Equivalent Method for Computing Mechanical Properties of Random and Ordered Hyperelastic Cellular Materials
Previous Article in Journal
Process Development of a Generative Method for Partial and Controlled Integration of Active Substances into Open-Porous Matrix Structures
Previous Article in Special Issue
Fluence and Temperature Dependences of Laser-Induced Ultrafast Demagnetization and Recovery Dynamics in L10-FePt Thin Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrahigh UV Responsivity Quasi-Two-Dimensional BixSn1−xO2 Films Achieved through Surface Reaction

1
School of Physics and Materials Science, Guangzhou University, Guangzhou 510006, China
2
Research Center for Advanced Information Materials (CAIM), Huangpu Research and Graduate School of Guangzhou University, Guangzhou 510006, China
3
Songshan Lake Materials Laboratory, Dongguan 523808, China
4
Key Lab of Si-Based Information Materials & Devices and Integrated Circuits Design, Department of Education of Guangdong Province, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(21), 6988; https://doi.org/10.3390/ma16216988
Submission received: 18 September 2023 / Revised: 18 October 2023 / Accepted: 26 October 2023 / Published: 31 October 2023

Abstract

:
In this study, quasi-two-dimensional BixSn1−xO2 (BTO) thin films were fabricated using a liquid metal transfer method. The ultraviolet (UV) photodetector based on BTO thin films was constructed, and the ultrahigh responsivity of 589 A/W was observed at 300 nm UV light illumination. Interestingly, by dropping ethanol during light-off period, the recovery time induced by the persistent photoconductivity (PPC) effect is reduced from 1.65 × 103 s to 5.71 s. Furthermore, the recovery time can also be reduced by dropping methanol, propylene glycol, NaNO2, and Na2SO3 after light termination. The working mechanisms are attributed to the rapid consumption of holes stored in BTO thin films by reaction with those solutions. This work demonstrates that the BTO thin films have potential applications in high-performance UV detectors and present an innovation route to weaken the PPC effects in semiconductors by introducing chemical liquids on their surface.

1. Introduction

Over the past decades, metal oxide semiconductors have attracted considerable interest in various applications, such as transparent thin-film transistors, [1,2] transparent conduction electrodes, [3] and optoelectronics [4,5]. Among the optoelectronic devices, ultraviolet (UV) photodetectors (PDs) are of particular interest due to their wide applications in plume detection, optical communications, and biomedical sensing [6,7,8,9,10,11]. UV photodetectors fall into many types, including photomultiplication (PM) type, photovoltaic type, and photoconductive type [12]. Photoconductive PDs are among the most attractive candidates due to their high responsivity, simple fabrication process, and durable performance [13,14]. The high responsivity and high internal photoconductive gain are primarily attributed to photocarriers being captured by defects within the material. However, one notable drawback associated with photoconductive PDs is the phenomenon known as the persistent photoconductivity (PPC) effect [15,16]. PPC effect refers to the photoconductivity that can persist for a long time (>1 s) after the terminating light source, which prevents the rapid recovery of the initial undisturbed state. The PPC effects prolong the response times of UV photoconductors and hinder the rapid recovery of the performance of UV photodetectors [17]. Therefore, reducing the PPC effect is an urgent need in photoconductive PDs.
To address this issue, many approaches have been proposed, such as introducing light pulses [18] and electric fields [19,20] in three-terminal devices, surface assembly of organic molecules [21], infrared irradiation [22], and heating [23,24]. For example, Sanghun et al. [18] iminate the PPC effect effectively by applying the pulsed voltage in the three-terminal device, but this device contains a stack of GIZO (25 nm)/IZO (50 nm)/GIZO (17 nm) layers and incorporates an inverted staggered switch-TFT and photo-TFT, which is challenging to fabricate. Pinto et al. [25] reported a phenyl-C61-butyric acid methyl ester (PCBM)/rubrene crystal bilayer structure with enhanced sensitivity of 4 × 104 s and a maximum photoresponsivity of 20 A/W, but the decay time in their time-resolved photosensitivity test is still in the range of a few seconds. Sun et al. [24] revealed that continuously pulsed heating can reduce the decay time of a suspended AlGaN/GaN heterostructure photodetector by 30–45% compared to the DC heating mode. However, in most cases, the decay time is just reduced from days or several hours to hundreds of seconds. There is no method that can suppress the PPC effect completely. Furthermore, the use of the aforementioned methods present challenges and is still under investigation due to the requirement of high electric fields, extra facilities, and complicated process [20,22]. The development of cheap and simple methods, along with the identification of suitable materials to eliminate the PPC effect is urgently needed.
SnO2 is a promising candidate for UV detection due to its outstanding optical and electrical properties [26,27,28]. The SnO2-based photodetectors with photoconductive mode usually exhibit ultrahigh responsivity (>103 A/W) in the UV light range. Unfortunately, the recovery process of the corresponding photodetectors is relatively slow [29,30,31] due to the PPC effect. Some efforts have been implemented to improve device performance by achieving better crystal quality and constructing low-dimension SnO2 nanostructures [4,32,33]. For example, Liu et al. [4] prepared a single SnO2 microrod photoconductor and demonstrated it has high photoconductive gain (≈1.5 × 109) and quick recovery speed (<1 s). The PPC effect in this work is eliminated by bending and straightening the microrod and subsequently applying a voltage pulse. Nevertheless, achieving such high-performance characteristics requires precise control of the radius of curvature of the bent film and the application of appropriate voltages. These requirements still present challenges in terms of complexity and cost. As an alternative, barium titanate (BTO) has attracted widespread interest in making electronic devices in recent years, including microwave electronics, actuators, piezoelectric transducers, and infrared detectors owing to its large dielectric constant, ferroelectricity, and piezoelectric characteristics [34,35,36,37]. In our previous study, a deep UV photodetector based on BixSn1−xO2 (2D-BTO) (0.017 < x < 0.041) was fabricated and it exhibits a dark current as low as 0.25 nA and a UV responsivity as high as 60 A/W [38]. The recovery speed can be reduced to approximately 1 s if this photodetector is exposed to an ethanol stream environment. This study offers a new possibility to eliminate the PPC effect by exposing the photodetectors to special chemical liquids.
In this study, we have fabricated quasi-two-dimensional BixSn1−xO2 (2D-BTO) thin films (x < 0.002) using the liquid metal transfer method. The photodetector based on 2D-BTO exhibits a high responsivity of 589 A/W and detectivity of 6.82 × 1012 Jones at 300 nm. The recovery speed of the resultant photoconductor is dramatically enhanced by introducing various reductive steam/drops on the device surface during the light-off period. The PPC effect is dramatically weakened due to consuming the photogenerated carriers with introduced chemicals, such as ethanol, propylene glycol, NaNO2, and Na2SO3.

2. Experimental Details

2.1. Sample Preparation

The fabrication process of the BTO films was conducted in a class 1000 clean room (see Figure 1). First, 10 g Sn is heated and melted in the heating stage at a temperature of 280 °C. Then, 1 g of Bi was added to the Sn to form a Bi-Sn alloy (Figure 1a) at a temperature of 280 °C. The formed rough oxide layer generated on the melted Bi-Sn surface was removed using a smooth glass. The smooth glass is inserted into a height-adjustable workbench and then moved in a direction parallel to the Bi-Sn alloy by applying a pushing force. Under the action of the thrust, the rough oxide layer is removed. When the rough oxide layer was removed, this sample was exposed to air for 20 s to obtain a more uniform oxide layer through a non-self-limiting oxidation reaction. Subsequently, a preheated substrate was used to touch the liquid metal for 10 s (Figure 1b). The used substrate is a p-type doped (100) crystal-oriented wafer with a SiO2 top layer of 300 nm thickness. A tweezer was used to carefully lift the substrate from liquid metals in the perpendicular direction. Since the melted Bi-Sn alloy has a lower surface energy than the SiO2 substrate, it tends to wet and adhere to the solid surface. In addition, due to van der Waals forces, the atoms and molecules on the contact area are attracted, which further contributes to the adhesion between the liquid metal and the substrate surface. Thus, a Bi-doped SnOx film was transferred to the substrate. When the substrate and the liquid metal were separated, a Bi-doped SnOx film was transferred to the substrate under van der Waals forces (Figure 1c), followed by an annealing process at 400 °C for 30 min to obtain BTO films (Figure 1d). Finally, two electrodes were mounted on the BTO thin films by an e-beams evaporation machine (TEMD 500, Beijing Technol Science Co., Ltd., Beijing, China) through a self-designed mask that retains two holes to allow the evaporated material to pass through. The used e-beam power is 20 KW, the energy is 5 MeV, and the current is 4 MA. The fabricated BTO thin films were irradiated under 5 V-biased 365 nm UV light (Figure 1e) for 20 s ~ 5 min. After the termination of the light source, 0.05 mL ethanol was dropped into the sample surface through a plastic injector (Figure 1f).

2.2. Measurement and Characterization

Keysight B1500A semiconductor device parameter analyzer was utilized to test the photocurrent and responsivity of fabricated BTO films. The used voltage is 5 V, and the illumination wavelength of UV light is 365 nm. The X-ray Photoelectron Spectroscopy (XPS, Nexsa, Thermo scientific, Waltham, MA, USA) was used to measure the element valences. The XPS uses an Ag Lα excitation source. The full spectrum, O, Sn, and Bi spectra were extracted, respectively. An atomic force microscope (AFM, dimension icon, Bruker Corporation, Billerica, MA, USA) was used to test the thickness of the prepared BTO thin films. The sample surface was scanned by AFM, and the height variations across the whole surface were collected. The average thickness is calculated using collected topographic data from the AFM scan. Absorption spectra were performed using a UV-VIS spectrophotometer (UV-Vis, UV-3600i PULS, Shimadz Corporation, Kyoto, Japan). The absorption spectra were measured from 200 to 1000 nm. Surface micromorphology was observed by a Metallographic microscope (Meta test E1-M, Metatest Corporation, Nanjing, China) and a Transmission Electron Microscopy (JEM-F2000, Jeol, Tokyo Metropolis, Japan). The used electron beam acceleration voltage is 100 kV, and the magnification is 5000× and 50,000×. The samples prepared for TEM observation were cut by a focused ion beam machine (ZEISS-AURIGA, Carl Zeiss AG, Oberkochen, Germany).

3. Results and Discussion

Structure and Elements

Figure 2a shows the optical absorption spectra of 2D-BTO films before and after annealing at 400 °C for 30 min in an oxygen atmosphere. After thermal treatment, the absorption edge shifts towards the lower wavelength side, which indicates the optical band gap of 2D-BTO films becomes wider. And the absorbance is reduced after annealing treatment. This is because the BTO film becomes more transparent after annealing, which involves the partial conversion of SnO to SnO2 after thermal treatment in an oxygen atmosphere [39]. In addition, Bi doping can lead to changes in the bandgap energy of SnO2 when annealing, which reduces the absorbance of the prepared BTO thin films. The ( α h v ) 2 was depicted in Figure 2b. The optical band gap of 2.78 eV of as-prepared BTO films was obtained by linearly extrapolating towards zero absorption. The band gap is consistent with the reported values of SnO [40], which further illustrates that dominant oxides are SnO in the as-prepared BTO films. After annealing, the optical band gap increases to 3.10 eV, which is close to the reported values of SnO2 films [41]. The optical band gap increasing after the annealing of BTO films can be partially attributed to the crystal structure changes, since the annealing process causes the atoms in the BTO film to rearrange, forming a more ordered and perfect crystal lattice. This increased order causes changes in the material’s electronic structure, affecting its band gap. On the other hand, Bi doping and annealing can help reduce defects and vacancies in the crystal lattice, which can introduce energy levels within the band gap. Fewer defects mean a more uniform and well-defined energy band structure, which can result in a larger band gap.
Figure 3 shows the morphology and thickness of the fabricated quasi-2D-BTO films. The morphology was observed at 20× magnification of a metallurgical microscope, where the colors of the SiO2 substrate and the 2D-BTO film are purple and blue, respectively (Figure 3a). It can be seen that good uniformity is obtained in the prepared BTO film, and an average thickness of ~11 nm was realized for the BTO film (Figure 3c). In TEM images (Figure 3d,e), there are some regions where the BTO film and substrate boundaries are closely connected and some regions where cracks are observed at the boundary between the BTO film and the substrate. The occurrence of these cracks can be ascribed to the differing thermal expansion coefficients of the thin films and substrates, resulting in varying rates of expansion and, consequently, inducing stress and cracking. In addition, BTO thin film exhibits greater roughness compared to the Si substrate. This phenomenon can be attributed to the annealing process, specifically the temperature and duration to which the film is subjected. During the annealing process, the BTO thin film may undergo various surface imperfections, such as wrinkling, cracking, or increased roughness due to the induced strain and stress.
The X-ray photoelectron spectroscopy of the BTO films is presented in Figure 4. The element peaks of Sn, O, and Bi can be observed, and the atomic percentages of Sn, O, and Bi are 29.2%, 70.7%, and 0.1%, respectively. The Bi element was measured to be less than 0.2%, which can be interpreted as the Gibbs free energy (ΔGf) of BiO2 is lower than that of SnO2 [42]. According to the laws of thermodynamics, the metal oxide (SnO2) with the largest ΔGf will cover the surface of the Bi-Sn alloy, and only a small amount of the metal oxide (BiO2) will be precipitated due to a lower ΔGf. The high-resolution XPS spectrum of O1s is fitted by the amorphous oxygen at 532.52 eV and the lattice oxygen at 530.58 eV. The amorphous lattice oxygen peak is much higher than that of lattice oxygen can be partially attributed to the large amount of oxygen adsorption on the surface of the 2D-BTO film. In addition, the amorphous lattice oxygen peak may partially come from the amorphous SiO2 substrate since its oxygen atoms are not arranged in regular, repeating crystal lattices. This lack of long-range order can lead to differences in the binding energy of the oxygen atoms and result in distinct peaks in the XPS spectrum. The peaks at 164.52 eV and 159.18 eV correspond to Bi 4f5/2 and Bi 4f7/2 of BixSn1−xO2, respectively, which means that the bismuth element in the film is Bi3+. The Sn 3d3/2 and 3d5/2 peaks were located at 495.08 eV and 486.68 eV, respectively. The two Sn 3d peaks can be further fitted by the Sn2+ peak centered at 494.8 eV and 486.3 eV, and the Sn4+ peaks centered at 495.6 eV and 487.3 eV, respectively. This evidence further demonstrated the existence of both SnO and SnO2 in the resultant BTO films.
The photocurrent was tested in the wavelength ranging from 200 nm to 1000 nm under a 5 V bias. The spectral responsivity (Rλ) [43] of photodetectors is calculated by the following formula [44,45,46]:
R λ = I Ph I Dark P λ S
where Rλ is the responsivity, IPh is the photocurrent, IDark is the dark current, Pλ is the density of the incident light, and S is the optical sensitive area. As depicted in Figure 5a, the responsivity of the BTO film increases rapidly as the wavelength increases from 250 nm to 300 nm, with a peak value of 589 A/W at 300 nm. Then, the responsivity reduces to a very low value during the wavelength range of 400–1000 nm, which is well consistent with the optical band gap of 3.10 eV. Thus, the 2D-BTO film can serve as highly sensitive visible-blind photodetectors. The high responsivity is attributed to the photocarriers trapped by the internal or surface defects/impurities in the quasi-2D-BTO with high surface area. From Figure 5a, it can be seen that the full width at half maximum (FWHM) of the responsivity for the film photodetector is 49 nm, which is much narrower than reported devices based on undoped SnO2 film and can be used as a narrowband UV photodetector.
The minimum detectable level of a photodetector can be described by the detection rate D* given by the following formula [39]:
D * = R λ S 2 q I d a r k = R λ 2 q J d a r k
where Rλ is the spectral responsivity, S is the effective area (S = 4.43 × 10−3 cm2), q is the electronic charge (q = 1.602 × 10−19 C), Idark is the dark current, and Jdark is the dark current density. The calculated detection rate D* is shown in Figure 5d, from which it can be seen that in the 2D-BTO film photoconductor, the maximum value is about 6.82 × 1012 Jones.
Figure 5b shows the photocurrent generated under 300 nm light irradiation with a power density of 45.52 μW/cm2. The photocurrent initially increases dramatically and then gradually drops. The maximum photocurrent reaches ~224A. However, the response time is relatively slow, requiring 145.52 s (Figure 5b), which is not suitable for a good photoelectric detector. As shown in Figure 5c, the recovery speed is too slow, which presents a typical PPC effect. The photocurrent of the recovery process in 2D-BTO film can be fitted by the following equation [34]:
I P P C t = I P P C ( 0 ) e ( t / τ ) β
where τ represents the lifetime of the photoinduced carriers, β is the decay exponent (0 < β < 1), and IPPC(0) is the buildup photocurrent at the moment of the light source being removed. The β = 0.69, τ = 1.65 × 103 s for 2D-BTO films after 300 nm UV irradiation was calculated by fitting. The long recovery time can be attributed to the presence of internal or surface defects/impurities-related deep energy levels in the band gap of 2D-BTO films. The results of the 2D-BTO decay time fit after different solution treatments are listed in Table 1. It can be clearly seen that the solution treatment τ is significantly reduced. This indicates that the 2D-BTO film has a photocatalytic effect and the PPC effect can also be significantly reduced by this method. The lifetime of photo-induced carriers determines how long they are available to participate in chemical reactions [47]. A shorter carrier lifetime means the carriers recombine (electron and hole coming back together) more quickly. Rapid carrier recombination reduces the chances of these carriers reaching the material’s surface to participate in reactions. As a result, the photocatalytic efficiency is reduced. In addition, in BTO materials, the PPC effect is often related to the accumulation and trapping of charge carriers [48]. A reduced carrier lifetime means carriers recombine quickly instead of being trapped or accumulating in defect states. Thus, the material loses its ability to exhibit persistent photoconductivity.
The PPC effect and long recovery time of photoconductors are not beneficial for photodetector applications. The long recovery time is caused by the photocarriers trapping by the defects or impurities, and many unrecombined photocarriers still residue in the 2D-BTO after turning off the UV light. The recombination between photoelectrons and photoholes is retarded, and the reduction of photocurrent requires a long time. The rapid consumption of photocarriers, i.e., photoelectrons and photoholes, may assist in the recovery of photodetectors after turning off the UV light. It is well known that the photoholes have oxidative properties. Here, some reductive substances are employed to react with residual photoholes, and the recovery process of the photodetector is accelerated by the consumption of photocarriers through chemical reactions between photoholes and reductive substances.
Figure 6 shows the photocurrent changes of the prepared 2D-BTO film with and without dripping of ethanol. Ethanol was added 1 min later after light off. In Figure 6a, the photocurrent was measured at different illumination times using 365 nm UV light at a temperature of 40 °C. ON-IS represents the photocurrent without illumination, OFF-max shows the maximum photocurrent before the termination of the light source, ET IN-Imin represents the photocurrent when ethanol starts to drop, and IE is the final current. It can be seen that the photocurrent increases as the irradiation time increases and the maximum photocurrent reaches approximately 36.9 μA when the irradiation time reaches 5 min. After the light off, the photocurrent slowly decreases following a slow drop mode (IMax~Is). When introducing the ethanol steam to the photodetector, the photocurrent drops dramatically, and finally levels off. Herein, I is introduced, which represents the difference between the IS and IE, as follows
I Δ = I S I E
Figure 6b shows the photocurrent variation of the 2D-BTO film under different irradiation wavelengths, ranging from 300 nm to 800 nm. It is found that the 2D-BTO film can only produce a large photocurrent in the wavelength range of 300 nm to 460 nm. Under the irradiation of light at wavelengths larger than 460 nm, the photocurrent can hardly be observed. Thus, the current variation in this wavelength range is essentially the same as that in the dark conditions. With the addition of ethanol, a rapid drop trend is observed. The obtained I values are I Δ 300 n m = 15.1 μA, I Δ 365 n m = 12.7 μA, I Δ 400 n m = 5.7 μA, and I Δ 460 n m = 2.6 μA, respectively (Figure 6d), which decrease with irradiation light wavelength increasing. Notably, the current will still decrease when ethanol is added in the dark, this is due to the fact that the test was performed under the condition of 5 V bias plus and the carriers generated by electrical excitation reacted with ethanol. Therefore, the photocatalytic response of the 2D-BTO film to ethanol should be the joint result of the action of both photoexcited carriers and electrically excited carriers [38]. After introducing the ethanol, the variation of dark current I is 2.5 μA, which is much lower than the residual photocurrent after light off. Thus, the residual photocarriers mainly respond to the ethanol molecules. To check the reliability of the performance of the BTO film, a repeated experiment under 365 nm illumination was performed, as shown in Figure 6c. Similar behaviors of current curves are observed, which further demonstrate that the fabricated BTO film has excellent repetitive performance.
Figure 7 illustrates the energy level and the reaction mechanism between the 2D-BTO film and ethanol. Under the excitation of UV light, the electrons in the 2D-BTO film are initially excited from the valence band to the conduction band (①). The existence of an intermediate band (IMB) originating from defects/impurities-related deep level can store photoelectrons from the conduction band otherwise directly recombine with the photoholes in the valence band, i.e., the photoelectrons are captured by the intermediate energy level (②), and then slowly jump to the valence band, resulting in a large number of photoholes residing in the valence band that cannot be consumed immediately. When the ethanol droplet contacts with the 2D-BTO, the “dark” photocatalysis process will take place, and the unrecombined photoholes will oxidize ethanol molecular to acetaldehyde, and the photoholes are consumed rapidly by oxidizing the ethanol to acetaldehyde with high reductive property [44]. Furthermore, the photoelectrons can be depleted by reacting with adsorbed O2 on the surface of the produced H+ from the photohole oxidation reaction. In addition, the photogenerated electrons separated in the conduction band diffuse to the surface of 2D-BTO can form a strong reduction potential with strong reducing properties to react with oxidizing substances such as CO2 to form CO, CH4, and other substances [49]. Photogenerated electrons can also reduce the oxygen on the surface of the polymer film to superoxide radical ·O2. And ·O2 undergoes a series of reactions to produce ·OH radicals, thus degrading organic matter [50].
Figure 8 reports the photocurrent variation of the 2D-BTO film before and after introducing methanol, propylene glycol, Na2SO3, and NaNO2, respectively. The solutions of Na2SO3 and NaNO2 have concentrations of 0.1 mol/L each. The Na2SO3 solution was prepared by adding 12.6 g of Na2SO3 to 1L of deionized water and stirring until homogeneous. On the other hand, the NaNO2 solution was prepared by adding 5.3 g of Na2SO3 to 1L of deionized water and stirring until homogeneous. Next, the Na2SO3 and NaNO2 solutions were dropped in 1 min after the UV light off. Different UV light exposure times are explored, including 20 s, 40 s, 1 min, 3 min, and 5 min. After introducing the above substances, the decay of residual photocurrent changes from slow drop mode to rapid drop mode (Figure 8a,d,g,j). For methanol, the photocurrent initially decreased dramatically and then slowly decreased to the lowest value, but finally made a significant recovery. In contrast, the current in the samples dropped with propylene glycol always showed a downward trend. For the samples dropped with Na2SO3 and NaNO2, the current also shows a similar trend: it initially drops moderately, then decreases quickly, and finally shows a slow downward trend to the end.
The repeat experiments have been carried out (Figure 8b,e,h,k). Obviously, the sample dropped with propylene glycol shows the best reproducibility because the current variation almost maintains consistency. However, the reproducibility of the sample dropped with methanol is relatively poor, and the final current shows a large fluctuation. Figure 8c,f,i,l report the dark current change by adding these solutions without UV light illumination. A significant degree of the dark current drop is observed in the sample dropped with methanol, and the initial dark current was reduced by approximately 1.2 μA. Different degrees of the dark current drop were also observed in other cases. Notably, the dark current showed almost no change in the sample dropped with NaNO2, reducing only approximately 0.1 μA, which shows that NaNO2 is not a good candidate for UV photodetection and photocatalysis.
The photocurrent (IΔ5min) of 2D-BTO after 5 min of UV light irradiation and τ values after adding the solution are summarized in Table 1. The maximum IΔ5min is obtained in the sample dropped with ethanol, realizing an average value of approximately 12.14 μA. In comparison, the minimum IΔ5min value only has an average value of 1.22 μA, which is obtained in the sample dropped with NaNO2. Compared with other solutions, these results further suggested that ethanol is a promising candidate for UV photodetection and photocatalysis compared to other solutions.
By comparing the values of τ and IΔ in Table 1, it can be found that the sample without any chemical introduction shows a slow recovery speed with τ value of 1.65 × 103 s. The recovery times of 2D-BTO greatly decrease after introducing the chemical solution. And by fitting Equation (2), it is calculated that the τ of the 2D-BTO after the introduction of ethanol reaches 5.71 s. It is noteworthy that the τ values after sodium sulfite and sodium nitrite treatment are very small only 2.53 × 10−2 s and 4.69 × 10−2 s, but IΔ is also very small, indicating that these two solutions are well suited for application to devices with small photocurrents to eliminate the PPC benefit. A comparison of the dark current and recovery speed between the presented BTO thin films and the previous report for other micro/nanostructured SnO2 photodetectors was provided in Table 2. The results reported in this work are superior to those reported in reference [6,28,51], where the recovery time is greater than 10 s. Although references [4,5] exhibit a fast recovery time, the fabrication process of Sb-doped SnO2 nanowires is complicated and involves using many different techniques. In addition, the SnO2 microrod reported in reference [4] needs an accurate radius of curvature of the bent film and appropriate voltage, which still faces the problem of complexity and cost, while the photon responsivity in reference [38] is relatively small(i.e., 60A/W). In this case, the fabrication process is relatively simple and the used chemical liquids such as ethanol, methanol, and propylene glycol are easily available. The use of chemical liquids to eliminate PPC effects still deserves further exploration.

4. Conclusions

In this work, quasi-two-dimensional BixSn1−xO2 (2D-BTO) thin films were prepared by a liquid metal transfer method. The fabricated 2D-BTO photodetectors exhibit a significant response to UV light, realizing the summit responsivity of 589 A/W at 300 nm and detectivity of 6.82 × 1012 Jones at 300 nm. After turning off the UV light, the photocurrent decreases slowly due to the PPC effect. By introducing ethanol, methanol, propylene glycol, Na2SO3, and NaNO2 on the device surface, the photocurrent of 2D-BTO film rapidly reduces in a short time. And the recovery time of 2D-BTO film is reduced from 1.65 × 103 s to 5.71 s. The rapid recovery can be attributed to the accelerated consumption of photocarriers by reacting with the introduced ethanol or other chemicals. This study demonstrates high-performance UV photodetectors based on quasi-two-dimensional oxides and provides an innovation route to reduce the recovery time of photoconductors with the PPC effect. Future work needs to be carried out to explore the long-term stability and the optimal concentration of the chemical liquids, as well as chemical solutions that can better eliminate PPC.

Author Contributions

Investigation and writing—original draft preparation, Z.X.; formal analysis and data curation, M.X.; Software and data curation, F.C.; resources and validation, R.Z.; visualization and writing—review and editing, Y.W.; supervison and conceptualization, Z.Z.; Methodolog, project administration, and funding acquisition, S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (No. 51872054), Guangdong Basic and Applied Basic Research Foundation (No. 2023A1515012700), Key Area R&D Program of Guangzhou (No. 202103030002), the Science and Technology Program of Guangzhou (No.202201020391), Key Discipline of Materials Science and Engineering, and the Bureau of Education of Guangzhou (Grant number: 202255464).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

We are grateful to Weilong Chen for his help with the preparation of figures in this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nomura, K.; Ohta, H.; Ueda, K.; Kamiya, T.; Hirano, M.; Hosono, H. Thin-film transistor fabricated in single-crystalline transparent oxide semiconductor. Science 2003, 300, 1269–1272. [Google Scholar] [CrossRef]
  2. Nomura, K.; Ohta, H.; Takagi, A.; Kamiya, T.; Hirano, M.; Hosono, H. Room-temperature fabrication of transparent flexible thin-film transistors using amorphous oxide semiconductors. Nature 2004, 432, 488–492. [Google Scholar] [CrossRef] [PubMed]
  3. Minami, T. Transparent conducting oxide semiconductors for transparent electrodes. Semicond. Sci. Technol. 2005, 20, S35. [Google Scholar] [CrossRef]
  4. Liu, K.; Sakurai, M.; Aono, M.; Shen, D. Ultrahigh-Gain Single SnO2 Microrod Photoconductor on Flexible Substrate with Fast Recovery Speed. Adv. Funct. Mater. 2015, 25, 3157–3163. [Google Scholar] [CrossRef]
  5. Wan, Q.; Dattoli, E.; Lu, W. Doping-dependent electrical characteristics of SnO2 nanowires. Small 2008, 4, 451–454. [Google Scholar] [CrossRef]
  6. Hu, L.; Yan, J.; Liao, M.; Wu, L.; Fang, X. Ultrahigh external quantum efficiency from thin SnO2 nanowire ultraviolet photodetectors. Small 2011, 7, 1613–6810. [Google Scholar] [CrossRef]
  7. Bie, Y.Q.; Liao, Z.M.; Zhang, H.Z.; Li, G.R.; Ye, Y.; Zhou, Y.B.; Xu, J.; Qin, Z.X.; Dai, L.; Yu, D.P. Self-powered, ultrafast, visible-blind UV detection and optical logical operation based on ZnO/GaN nanoscale p-n junctions. Adv. Mater. 2011, 23, 649–653. [Google Scholar] [CrossRef]
  8. Xia, F.; Mueller, T.; Lin, Y.M.; Valdes-Garcia, A.; Avouris, P. Ultrafast graphene photodetector. Nat. Nanotechnol 2009, 4, 839–843. [Google Scholar] [CrossRef]
  9. Marimuthu, G.; Saravanakumar, K.; Jeyadheepan, K.; Mahalakshmi, K. Achieving self-powered photoresponse in mono layered SnO2 nanostructure array UV photodetector through the tailoring of electrode configuration. J. Photochem. Photobiol. A-Chem. 2022, 428, 113860. [Google Scholar] [CrossRef]
  10. Yu, H.; Qu, L.H.; Zhang, M.X.; Wang, Y.X.; Lou, C.Q.; Xu, Y.; Cui, M.Q.; Shao, Z.T.; Liu, X.; Hu, P.A.; et al. Achieving High Responsivity of Photoelectrochemical Solar-Blind Ultraviolet Photodetectors via Oxygen Vacancy Engineering. Adv. Opt. Mater. 2023, 11, 2202341. [Google Scholar] [CrossRef]
  11. Huang, Z.P.; Zhu, J.; Hu, Y.; Zhu, Y.P.; Zhu, G.H.; Hu, L.P.; Zi, Y.; Huang, W.C. Tin Oxide (SnO2) Nanoparticles: Facile Fabrication, Characterization, and Application in UV Photodetectors. Nanomaterials 2022, 12, 632. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, M.; Wang, J.; Yang, K.; Zhao, Z.; Zhou, Z.; Ma, Y.; Shen, L.; Ma, X.; Zhang, F. Highly sensitive, broad-band organic photomultiplication-type photodetectors covering UV-Vis-NIR. J. Mater. Chem. C 2021, 9, 6357–6364. [Google Scholar] [CrossRef]
  13. Jia, L.; Zheng, W.; Huang, F. Vacuum-ultraviolet photodetectors. PhotoniX 2020, 1, 22. [Google Scholar] [CrossRef]
  14. Wang, Y.; Li, S.; Cao, J.; Jiang, Y.; Zhang, Y.; Tang, W.; Wu, Z. Improved response speed of β-Ga2O3 solar-blind photodetectors by optimizing illumination and bias. Mater. Des. 2022, 221, 110917. [Google Scholar] [CrossRef]
  15. Feng, P.; Mönch, I.; Harazim, S.; Huang, G.; Mei, Y.; Schmidt, O.G. Giant persistent photoconductivity in rough silicon nanomembranes. Nano Lett. 2009, 9, 3453–3459. [Google Scholar] [CrossRef]
  16. Lany, S.; Zunger, A. Anion vacancies as a source of persistent photoconductivity in II-VI and chalcopyrite semiconductors. Phys. Rev. B 2005, 72, 035215. [Google Scholar] [CrossRef]
  17. Liu, X.; Li, S.; Li, Z.; Cao, F.; Su, L.; Shtansky, D.V.; Fang, X. Enhanced response speed in 2D perovskite oxides-based photodetectors for UV imaging through surface/interface carrier-transport modulation. ACS Appl. Mater. Interfaces 2022, 14, 48936–48947. [Google Scholar] [CrossRef]
  18. Jeon, S.; Ahn, S.-E.; Song, I.; Kim, C.J.; Chung, U.I.; Lee, E.; Yoo, I.; Nathan, A.; Lee, S.; Ghaffarzadeh, K. Gated three-terminal device architecture to eliminate persistent photoconductivity in oxide semiconductor photosensor arrays. Nat. Mater. 2012, 11, 1476–4660. [Google Scholar] [CrossRef]
  19. Hou, Q.; Wang, X.; Xiao, H.; Wang, C.; Yang, C.; Yin, H.; Deng, Q.; Li, J.; Wang, Z.; Hou, X. Influence of electric field on persistent photoconductivity in unintentionally doped n-type GaN. Appl. Phys. Lett. 2011, 98, 102104. [Google Scholar] [CrossRef]
  20. Xu, J.; You, D.; Tang, Y.; Kang, Y.; Li, X.; Li, X.; Gong, H. Electric-field effects on persistent photoconductivity in undoped n-type epitaxial GaN. Appl. Phys. Lett. 2006, 88, 072106. [Google Scholar] [CrossRef]
  21. Huang, Y.; Zhuge, F.; Hou, J.; Lv, L.; Luo, P.; Zhou, N.; Gan, L.; Zhai, T. Van der Waals Coupled Organic Molecules with Monolayer MoS2 for Fast Response Photodetectors with Gate-Tunable Responsivity. ACS Nano 2018, 12, 4062–4073. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Liao, Z.L.; She, G.W.; Mu, L.X.; Chen, D.M.; Shi, W.S. Optical modulation of persistent photoconductivity in ZnO nanowires. Appl. Phys. Lett. 2011, 98, 203108. [Google Scholar] [CrossRef]
  23. Zhou, H.; Cong, L.; Ma, J.; Li, B.; Chen, M.; Xu, H.; Liu, Y. High gain broadband photoconductor based on amorphous Ga2O3 and suppression of persistent photoconductivity. J. Mater. Chem. C 2019, 7, 13149–13155. [Google Scholar] [CrossRef]
  24. Sun, J.; Zhang, S.; Zhan, T.; Liu, Z.; Wang, J.; Yi, X.; Li, J.; Sarro, P.M.; Zhang, G. A high responsivity and controllable recovery ultraviolet detector based on a WO3 gate AlGaN/GaN heterostructure with an integrated micro-heater. J. Mater. Chem. C 2020, 8, 5409–5416. [Google Scholar] [CrossRef]
  25. Pinto, R.M.; Gouveia, W.; Neves, A.I.S.; Alves, H. Ultrasensitive organic phototransistors with multispectral response based on thin-film/single-crystal bilayer structures. Appl. Phys. Lett. 2015, 107, 223301. [Google Scholar] [CrossRef]
  26. Chen, Y.; Zhu, C.; Cao, M.; Wang, T. Photoresponse of SnO2 nanobelts grown in situ on interdigital electrodes. Nanotechnology 2007, 18, 285502. [Google Scholar] [CrossRef]
  27. Mathur, S.; Barth, S.; Shen, H.; Pyun, J.C.; Werner, U. Size-dependent photoconductance in SnO2 nanowires. Small 2005, 1, 713–717. [Google Scholar] [CrossRef]
  28. Chen, H.; Hu, L.F.; Fang, X.S.; Wu, L.M. General Fabrication of Monolayer SnO2 Nanonets for High-Performance Ultraviolet Photodetectors. Adv. Funct. Mater. 2012, 22, 1229–1235. [Google Scholar] [CrossRef]
  29. Lin, C.-H.; Chen, R.-S.; Chen, T.-T.; Chen, H.-Y.; Chen, Y.-F.; Chen, K.-H.; Chen, L.-C. High photocurrent gain in SnO2 nanowires. Appl. Phys. Lett. 2008, 93, 112115. [Google Scholar] [CrossRef]
  30. Muraoka, Y.; Takubo, N.; Hiroi, Z. Photoinduced conductivity in tin dioxide thin films. J. Appl. Phys. 2009, 105, 103702. [Google Scholar] [CrossRef]
  31. Viana, E.R.; González, J.C.; Ribeiro, G.M.; De Oliveira, A.G. Photoluminescence and high-temperature persistent photoconductivity experiments in SnO2 nanobelts. J. Phys. Chem. C 2013, 117, 7844–7849. [Google Scholar] [CrossRef]
  32. Tian, W.; Zhang, C.; Zhai, T.; Li, S.-L.; Wang, X.; Liao, M.; Tsukagoshi, K.; Golberg, D.; Bando, Y. Flexible SnO2 hollow nanosphere film based high-performance ultraviolet photodetector. Chem. Commun. 2013, 49, 3739–3741. [Google Scholar] [CrossRef] [PubMed]
  33. Giebelhaus, I.; Varechkina, E.; Fischer, T.; Rumyantseva, M.; Ivanov, V.; Gaskov, A.; Morante, J.R.; Arbiol, J.; Tyrra, W.; Mathur, S. One-dimensional CuO–SnO2 p–n heterojunctions for enhanced detection of H2S. J. Mater. Chem. A 2013, 1, 11261–11268. [Google Scholar] [CrossRef]
  34. Hyeon, D.Y.; Park, K.-I. Piezoelectric flexible energy harvester based on BaTiO3 thin film enabled by exfoliating the mica substrate. Energy Technol. 2019, 7, 1900638. [Google Scholar] [CrossRef]
  35. Ahmed, S.E.; Poole, V.M.; Jesenovec, J.; Dutton, B.L.; McCloy, J.S.; McCluskey, M.D. Room-Temperature Persistent Photoconductivity in Barium Calcium Titanate. J. Electron. Mater. 2023, 52, 2499–2504. [Google Scholar] [CrossRef]
  36. Zhao, J.; Niu, G.; Ren, W.; Wang, L.; Zhang, N.; Sun, Y.; Wang, Q.; Shi, P.; Liu, M.; Zhao, Y. Polarization behavior oflead-free 0.94 (Bi0.5Na0.5) TiO3-0.06 BaTiO3 thin films with enhanced ferroelectric properties. J. Eur. Ceram. Soc. 2020, 40, 3928–3935. [Google Scholar] [CrossRef]
  37. Wang, Z.; Zhao, J.; Niu, G.; Zhang, N.; Zheng, K.; Quan, Y.; Wang, L.; Zhuang, J.; Wang, G.; Li, X. Ultra-high strain responses in lead-free (Bi0.5Na0.5) TiO3-BaTiO3-NaNbO3 ferroelectric thin films. J. Eur. Ceram. Soc. 2023, 43, 5511–5520. [Google Scholar] [CrossRef]
  38. Pan, S.S.; Liu, Q.W.; Zhao, J.G.; Li, G.H. Ultrahigh Detectivity and Wide Dynamic Range Ultraviolet Photodetectors Based on BixSn1-xO2 Intermediate Band. Semiconductor. ACS Appl. Mater. Interfaces 2017, 9, 28737–28742. [Google Scholar] [CrossRef]
  39. Dai, Z.R.; Pan, Z.W.; Wang, Z.L. Growth and structure evolution of novel tin oxide diskettes. J. Am. Chem. Soc. 2002, 124, 8673–8680. [Google Scholar] [CrossRef]
  40. Wang, Y.H.; Yang, Z.B.; Li, H.R.; Li, S.; Zhi, Y.S.; Yan, Z.Y.; Huang, X.; Wei, X.H.; Tang, W.H.; Wu, Z.P. Ultrasensitive Flexible Solar-Blind Photodetectors Based on Graphene/Amorphous Ga2O3 van der Waals Heterojunctions. ACS Appl. Mater. Interfaces 2020, 12, 47714–47720. [Google Scholar] [CrossRef]
  41. Zhang, Y.C.; Yao, L.; Zhang, G.S.; Dionysiou, D.D.; Li, J.; Du, X.H. One-step hydrothermal synthesis of high-performance visible-light-driven SnS2/SnO2 nanoheterojunction photocatalyst for the reduction of aqueous Cr(VI). Appl. Catal. B-Environ. 2014, 144, 730–738. [Google Scholar] [CrossRef]
  42. Zhao, S.S.; Zhang, J.Q.; Fu, L. Liquid Metals: A Novel Possibility of Fabricating 2D Metal Oxides. Adv. Mater. 2021, 33, 2005544. [Google Scholar] [CrossRef] [PubMed]
  43. Fan, Z.H.; Zhu, M.; Pan, S.S.; Ge, J.; Hu, L. Giant photoresponse enhancement in Cr2O3 films by Ni doping-induced insulator-to-semiconductor transition. Ceram. Int. 2021, 47, 13655–13659. [Google Scholar] [CrossRef]
  44. Bai, K.L.; Fan, Z.H.; Zhao, G.C.; He, X.Y.; Zhu, Z.B.; Pan, S.S.; Ge, J.; He, C.G. Water engineering in lead free CsCu2I3 perovskite for high performance planar heterojunction photodetector applications. Ceram. Int. 2023, 49, 1970–1979. [Google Scholar] [CrossRef]
  45. Zhou, J.; Huang, J. Photodetectors Based on Organic-Inorganic Hybrid Lead Halide Perovskites. Adv. Sci. 2018, 5, 1700256. [Google Scholar] [CrossRef] [PubMed]
  46. Xie, C.; Yan, F. Flexible Photodetectors Based on Novel Functional Materials. Small 2017, 13, 1701822. [Google Scholar] [CrossRef]
  47. Huang, H.; Chen, Y.; Shi, H. Boosting Separation of Charge Carriers in 2D/0D BiOBr Nanoflower Sheets/BN Quantum Dots with the Lorentz Force via Magnetic Field. Energy Fuels 2022, 36, 11495–11502. [Google Scholar] [CrossRef]
  48. Dutta, R.; Bala, A.; Sen, A.; Spinazze, M.R.; Park, H.; Choi, W.; Yoon, Y.; Kim, S. Optical Enhancement of Indirect Bandgap Two-Dimensional Transition Metal Dichalcogenides for Multi-Functional Optoelectronic Sensors. Adv. Mater. 2023, 2303272. [Google Scholar] [CrossRef]
  49. Rao, H.; Chmidt, L.C.S.; Bonin, J.; Robert, M. Visible-light-driven methane formation from CO2 with a molecular iron catalyst. Nature 2017, 548, 74–77. [Google Scholar] [CrossRef]
  50. Al-Mamun, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic activity improvement and application of UV-TiO2 photocatalysis in textile wastewater treatment: A review. J. Environ. Chem. Eng. 2019, 7, 103248. [Google Scholar] [CrossRef]
  51. Wu, J.-M.; Kuo, C.-H. Ultraviolet photodetectors made from SnO2 nanowires. Thin Solid Film. 2009, 517, 3870–3873. [Google Scholar] [CrossRef]
  52. Kim, D.; Shin, G.; Yoon, J.; Jang, D.; Lee, S.-J.; Zi, G.; Ha, J.S. High performance stretchable UV sensor arrays of SnO2 nanowires. Nanotechnology 2013, 24, 315502. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of quasi-2D-BTO films fabrication process and UV irradiation (a) heat melted Sn-Bi alloy, (b) silicon oxide substrate in contact with liquid metal, (c) separation of silicon oxide substrates from liquid metal, (d) Bi-doped SnOX film was annealed at 400 °C for 30 min, (e) photoelectric performance test of BTO film, (f) ethanol was added to the BTO film after irradiation.
Figure 1. Schematic diagram of quasi-2D-BTO films fabrication process and UV irradiation (a) heat melted Sn-Bi alloy, (b) silicon oxide substrate in contact with liquid metal, (c) separation of silicon oxide substrates from liquid metal, (d) Bi-doped SnOX film was annealed at 400 °C for 30 min, (e) photoelectric performance test of BTO film, (f) ethanol was added to the BTO film after irradiation.
Materials 16 06988 g001
Figure 2. (a) The absorption spectra of BTO thin film before and after thermal annealing at 400 °C, (b) the ( α h v ) 2 h v of BTO thin film before and after thermal annealing at 400 °C.
Figure 2. (a) The absorption spectra of BTO thin film before and after thermal annealing at 400 °C, (b) the ( α h v ) 2 h v of BTO thin film before and after thermal annealing at 400 °C.
Materials 16 06988 g002
Figure 3. (a) Micrograph of annealed BTO film with an optical magnification of 20, (b) AFM image of BTO film, (c) variation of BTO thin film thickness along the BTO and SiO2 interface, (d,e) are the TEM images with a magnification of 5000× and 50,000×, respectively.
Figure 3. (a) Micrograph of annealed BTO film with an optical magnification of 20, (b) AFM image of BTO film, (c) variation of BTO thin film thickness along the BTO and SiO2 interface, (d,e) are the TEM images with a magnification of 5000× and 50,000×, respectively.
Materials 16 06988 g003
Figure 4. XPS spectra of quasi-2D-BTO thin film with (a) full spectrum, (b) O 1s fine spectrum, (c) Sn 3d fine spectrum, and (d) Bi 4f fine spectrum.
Figure 4. XPS spectra of quasi-2D-BTO thin film with (a) full spectrum, (b) O 1s fine spectrum, (c) Sn 3d fine spectrum, and (d) Bi 4f fine spectrum.
Materials 16 06988 g004
Figure 5. (a) Responsivity of annealed quasi-2D-BTO films to different wavelengths of light at 5 V bias, (b) photocurrent variation and response time of BTO films illuminated by 300 nm UV light with a power density of 45.5 μW/cm2, (c) recovery time function fitting of BTO films under 300 nm UV illumination, (d) detectivity (D*) of photodetectors based on BTO films.
Figure 5. (a) Responsivity of annealed quasi-2D-BTO films to different wavelengths of light at 5 V bias, (b) photocurrent variation and response time of BTO films illuminated by 300 nm UV light with a power density of 45.5 μW/cm2, (c) recovery time function fitting of BTO films under 300 nm UV illumination, (d) detectivity (D*) of photodetectors based on BTO films.
Materials 16 06988 g005
Figure 6. Response of BTO films to ethanol. (a) Current variation of BTO films under 365 nm UV light on/off and drops of ethanol added, the volume of ethanol is 5 μL. (b) Current variation of BTO films after dropwise addition of ethanol for 20s~5min after 365nm UV light illumination at applied voltage of 5 V. (c) Current variation of BTO films reacting with ethanol after irradiation at 300~800 nm light for 1 min. (d) The difference between the IS and IE of the prepared BTO films under irradiation with light of different wavelengths.
Figure 6. Response of BTO films to ethanol. (a) Current variation of BTO films under 365 nm UV light on/off and drops of ethanol added, the volume of ethanol is 5 μL. (b) Current variation of BTO films after dropwise addition of ethanol for 20s~5min after 365nm UV light illumination at applied voltage of 5 V. (c) Current variation of BTO films reacting with ethanol after irradiation at 300~800 nm light for 1 min. (d) The difference between the IS and IE of the prepared BTO films under irradiation with light of different wavelengths.
Materials 16 06988 g006
Figure 7. Schematic diagram of the electron transition process and the reaction mechanism with ethanol in BTO films.
Figure 7. Schematic diagram of the electron transition process and the reaction mechanism with ethanol in BTO films.
Materials 16 06988 g007
Figure 8. (a,d,g,j) Current changes of BTO thin film before and after adding methanol, propylene glycol, Na2SO3, and NaNO2 under different UV exposure times; (b,e,h,k) repeat current curves of dropping methanol, propylene glycol, Na2SO3, and NaNO2; (c,f,i,l) dark current changes of BTO thin film after adding methanol, propylene glycol, Na2SO3, and NaNO2.
Figure 8. (a,d,g,j) Current changes of BTO thin film before and after adding methanol, propylene glycol, Na2SO3, and NaNO2 under different UV exposure times; (b,e,h,k) repeat current curves of dropping methanol, propylene glycol, Na2SO3, and NaNO2; (c,f,i,l) dark current changes of BTO thin film after adding methanol, propylene glycol, Na2SO3, and NaNO2.
Materials 16 06988 g008
Table 1. The 2D-BTO films with PPC fitting parameters for photodetectors after solution treatment.
Table 1. The 2D-BTO films with PPC fitting parameters for photodetectors after solution treatment.
Solution τ ( s ) βIDarkIΔ5min (μA)
None 1.65 × 1036.98 × 10−115.6 μA
Ethanol (99.5%)5.711.48 × 10−117.3 μA12.2
Methanol (99.85%)1.22 × 10−13.7 × 10−215.3 μA6.2
Propylene glycol (99%)9.41 × 10−25.44 × 10−220.3 μA4.0
Sodium sulfite (99.9%)2.53 × 10−24.79 × 10−212.1 μA1.5
Sodium nitrite (99.5%)4.69 × 10−25.40 × 10−211.8 μA0.9
Table 2. Comparison of the dark current and recovery speed between the prepared BTO thin films and previous report for SnO2 photodetectors.
Table 2. Comparison of the dark current and recovery speed between the prepared BTO thin films and previous report for SnO2 photodetectors.
PhotodetectorDark CurrentResponsivityRecovery TimeRefs.
SnO2 monolayer nanofilm60–90 μA/1 VNot mentioned>50 s[28]
SnO2 nanowire array77 µA/12 VNot mentioned>150 s[51]
Sb-doped SnO2 nanowire2 pA/1 V6250 A/W≈1[5]
SnO2 nanowire19.4 nA/1 VNot mentioned>50 s[6]
SnO2 nanowire array2 pA/1 VNot mentioned10 s[52]
SnO2 microrod13 µA/1 V3 × 105 A/W<1 s[4]
BixSn1−xO2 (2D-BTO) (0.017 < x < 0.041)0.25 nA60 A/W1 s[38]
This work15.6 µA/5 V589 A/W5.71 s
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, Z.; Xu, M.; Chen, F.; Zhai, R.; Wu, Y.; Zhao, Z.; Pan, S. Ultrahigh UV Responsivity Quasi-Two-Dimensional BixSn1−xO2 Films Achieved through Surface Reaction. Materials 2023, 16, 6988. https://doi.org/10.3390/ma16216988

AMA Style

Xu Z, Xu M, Chen F, Zhai R, Wu Y, Zhao Z, Pan S. Ultrahigh UV Responsivity Quasi-Two-Dimensional BixSn1−xO2 Films Achieved through Surface Reaction. Materials. 2023; 16(21):6988. https://doi.org/10.3390/ma16216988

Chicago/Turabian Style

Xu, Zhihao, Miao Xu, Fang Chen, Rui Zhai, You Wu, Zhuan Zhao, and Shusheng Pan. 2023. "Ultrahigh UV Responsivity Quasi-Two-Dimensional BixSn1−xO2 Films Achieved through Surface Reaction" Materials 16, no. 21: 6988. https://doi.org/10.3390/ma16216988

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

Article Metrics

Back to TopTop