Next Article in Journal
Double-Antibody Sandwich Immunoassay and Plasmonic Coupling Synergistically Improved Long-Range SPR Biosensor with Low Detection Limit
Next Article in Special Issue
Novel Blue-Wavelength-Blocking Contact Lens with Er3+/TiO2 NPs: Manufacture and Characterization
Previous Article in Journal
Effect of Cu Ion Concentration on Microstructures and Mechanical Properties of Nanotwinned Cu Foils Fabricated by Rotary Electroplating
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Properties and Applications of WO3 Nanostructure−Based Optical and Electronic Devices

Shandong Key Laboratory of Optical Communication Science and Technology, School of Physics Science and Information Technology, Liaocheng University, Liaocheng 252000, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 2136; https://doi.org/10.3390/nano11082136
Submission received: 29 June 2021 / Revised: 11 August 2021 / Accepted: 20 August 2021 / Published: 22 August 2021

Abstract

:
Tungsten oxide (WO3) is a wide band gap semiconductor with unintentionally n-doping performance, excellent conductivity, and high electron hall mobility, which is considered as a candidate material for application in optoelectronics. Several reviews on WO3 and its derivatives for various applications dealing with electrochemical, photoelectrochemical, hybrid photocatalysts, electrochemical energy storage, and gas sensors have appeared recently. Moreover, the nanostructured transition metal oxides have attracted considerable attention in the past decade because of their unique chemical, photochromic, and physical properties leading to numerous other potential applications. Owing to their distinctive photoluminescence (PL), electrochromic and electrical properties, WO3 nanostructure-based optical and electronic devices application have attracted a wide range of research interests. This review mainly focuses on the up-to-date progress in different advanced strategies from fundamental analysis to improve WO3 optoelectric, electrochromic, and photochromic properties in the development of tungsten oxide-based advanced devices for optical and electronic applications including photodetectors, light-emitting diodes (LED), PL properties, electrical properties, and optical information storage. This review on the prior findings of WO3-related optical and electrical devices, as well as concluding remarks and forecasts will help researchers to advance the field of optoelectric applications of nanostructured transition metal oxides.

1. Introduction

The transition metal oxide tungsten oxide (WO3), an oxygen-deficient n-type wide band gap semiconductor material with an electronic bandgap of ~2.6–3.0 eV, has received a lot of attention [1,2,3,4]. WO3 structures include cubic, triclinic, monoclinic, orthorhombic tetragonal, and hexagonal. Because of its high melting temperature, photo electrochromic, toughness, and mechanical properties, it is regarded as a potential candidate for optical and electrical applications [5,6]. Nanostructured WO3 has a high specific surface area and good surface permeability, making it ideal for a wide range of applications. WO3 nanostructures in various morphologies (for example, instant nanorods (NRs), nanosheet, 3D nanostructured papilio paris, and thin films (TFs)) have been fabricated for a variety of applications, including gas sensors [7], efficient water splitting [8], photoelectrocatalytic activity [9], memory devices [10], photodetectors [11,12], and high temperature diodes [13,14]. At present, nanostructured WO3 are deposited on various substrates to fabricate optical and electrical devices, such as TiO2 [15], NiO [16], ZnO nanowires (NWs) [17], diamond [14], Fe2WO6 [8], and BiVO4 [18]. In the last few years, several review reports have been published based on photo catalysts [19,20], electrochromic devices [21,22], gas sensors [23,24], and oxygen-deficient WO3 [25]. However, so far there is no review focusing specifically on nanostructured WO3 optical and electrical devices.
As a result, in this review, the authors present a comprehensive overview of past developments in optical properties, such as photodetectors, light-emitting diodes (LED), photoluminescence (PL) and electrical properties, and optical information storage, as reported by various research groups, which make WO3 appealing for a variety of applications. Additionally, we offer some closing remarks as well as a forecast of the future advances in the subject. The research presented here should serve as a solid starting point for developing new nanostructured WO3 structures for emerging and future optical and electrical applications.

2. Photodetector

An ideal photodetector will display a low dark current to minimize the interference noise and high responsiveness to maximize the optical signal. The performance of photodetectors usually depends on the bandgap of the semiconductors and metal oxides. WO3 is a typical wide bandgap semiconductor with a large exciton binding energy of 0.15 eV, a high optical absorption coefficient ≥104 cm−1, and a phonon-limited electron mobility of ~12 cm2 V−1 s−1. The WO3 band gap resonates with incident ultraviolet (UV) light energy, which can generate excess electrons that contribute to the photocurrent, thereby improving the characteristics of the photodetector [26]. These physical properties show that WO3 semiconductors have great potential for the fabrication of high-performance visible light and UV detectors with reasonable performance [27].

2.1. UV Photodetector

2.1.1. Nanostructured WO3 Photodetector

In recent years, there have been a few reports on the WO3 photodetector with a single nanostructure since its response time is slow and it has a low current on/off ratio [28,29,30]. In comparison to the previously described WO3 NWs [29,30] and nanospheres [28], Liu et al. have developed a photodetector based on a few layers of WO3 nanosheets that has a faster light response, a greater on/off ratio, a higher external quantum efficiency, high sensitivity, outstanding stability, and reversibility. This distinctive performance of WO3 photodetectors lays the foundation for the fabrication of high-performance flexible and multifunctional photodetectors derived from layered semiconductor materials (Figure 1a–d) [31]. To investigate UV photoresponse, symmetrical and asymmetrical standard lithography were evaluated to differentiate the sensitivity of the Ohmic contact and Schottky contact devices to UV photodetector based on a single WO3 NWs. The linear I-V curves of the symmetrical device in dark and UV illumination conditions demonstrated that the device is Ohmic contact. For Schottky contact device, one end of a WO3 nanowire was completely covered by the gold electrode from Omhic contact, and only very small area was covered at the other end. The I-V of a single WO3 nanowire device shows a typical diode I-V curve, and the effective circuit diagram in the lower-right inset. When reverse bias, the current was completely cut-off owing to the Schottky contact (Figure 2a,b). The Schottky contact device has a faster response time than the Ohmic contact device. This is due to the oxygen adsorption and desorption in the heterojunction influence the Schottky contact device barrier height. The adsorbed oxygen molecules on the surface of WO3 nanowire can modify the density of defect states and enhances the injection barrier (Figure 2c). Under UV illumination, the generated holes can release the adsorbed oxygen ions, and reduce the height of the injection barrier (Figure 2d). The reduction of the barrier height causes more electrons to cross the barrier and then enhances the photocurrent [32].
In a recent study, Kim et al. heat-treated a large area of amorphous WO3 TFs to fabricate high quality, high-density WO3 NRs. It is used to prepare UV detectors, with a fast reaction speed (0.316 s), very sensitive to the actual UV 261 nm wavelength, and can effectively perform UV detection (Figure 3a–d) [33]. The carriers can be generated by transitions between bands caused by UV absorption. The generated carriers can also be transferred between each WO3 NRs. Finally, enough applied voltage can overcome the energy band barrier between the WO3 NRs and the electrodes.

2.1.2. WO3 Thin Films for UV Photodetector

There are several reports on WO3 TFs based UV photodetectors. Reddy et al. studied the efficiency of UV photodetector characteristics by TiO2/WO3 bilayer TFs. The enhancement of oxygen vacancies in TiO2/WO3 bilayer plays an important role in photoresponse [34]. For the first time, high response and controllable recovery UV detector derived from a WO3 gate AlGaN/GaN heterojunction integrated microheater has been investigated (Figure 4a,b). The UV response rate of the device at 240 nm is 1.67 × 104 A W−1, and the cut-off wavelength is 275 nm [11].
To improve the performance of the UV detector based on WO3 films, the sputtering parameters need to be optimized, mainly oxygen partial pressure and sputtering pressure. Yadav et al. described the effect of oxygen partial pressure on the performance of WO3 thin films UV detectors [35]. Sputtering technology is used to deposit WO3 film under different oxygen partial pressures to improve its response rate under low UV power density. The fabricated photodetector can respond to a small quantity of UV radiation. In addition, the effects of sputtering pressure on the morphology, crystallinity, and photodetector properties of WO3 films were also investigated [12]. WO3 film has high crystallinity, surface roughness, and customized grain size, which contribute to achieving high responsivity and external quantum efficiency by minimizing the overall impedance. The results show that the WO3 film deposited under a sputtering pressure of 10 m Torr has good stability and high photodetector performance.
Furthermore, the oxygen vacancies were found as doubly ionized donors and help with photodetection at the same time. In addition, Sn ions into the WO3 lattice can enhance the conductivity and reduce the resistivity by increasing the carrier concentration and oxygen vacancy. The prepared WO3 and Sn–WO3 precursor solutions of 3 mL were deposited separately on the cleaned p-Si substrates (1 cm × 1 cm) by the jet nebulizer spray pyrolysis technique with different concentrations of Sn (0, 4, 8, and 12 wt.%) the copper (Cu) metal contact was coated on the Sn–WO3/p-Si using dc sputtering (Model Name: HIND high vacuum-PS 2000) with 4 mm diameter and 500 nm thickness. The instrument details and deposition conditions of the metal contacts were mentioned in their previous work [36,37]. The Sn-WO3/p-Si diode showed a positive light response of high reverse saturation current under illumination [38]. As the concentration increases, the detection capability of the interface layer also increases. The diode measured under light conditions exhibits higher current values than under the dark conditions. This behavior outcome indicates that all the Cu/Sn–WO3/p-Si diodes are highly photo-conducting in nature. In particular, the diode fabricated with 12 wt.% of Sn shows higher current values (mA level) when compared to other diodes (Figure 5a). It is presumed that the incorporation of Sn atoms in the WO3 matrix has effectively improved the photocurrent of the Cu/Sn–WO3/p-Si diodes. The ideality factor (n) of the diodes was found to reduce steeply under dark conditions on increasing the Sn concentration from 0 to 12 wt.%. Compared to the dark conditions, the diodes measured under light conditions revealed lower n values. This is mainly due to the increase in the photo-generated charge carriers (e–h+) along with the improved conversion efficiency of the semiconductor layer and absorption of interfacial layer (Figure 5b).

2.2. Visible Photodetectors

The nanostructures of WO3 and the UV detectors have been investigated, and device characteristics have been reported. However, there is not much research on the application of WO3 as visible light detectors. Wang et al. reported on the controlled synthesize of WO3 NWs and their application as visible photodetectors. The WO3 single NWs photodetector has shown outstanding device characteristics with a high responsivity of 19 A W−1 at 0.1 V, high detectivity of 1.06 × 1011 Jones, and a short response time of 8 ms under a 404 nm laser illumination (Figure 6a–f). Thus, the enormous possibility of WO3 NWs for manufacturing visible photodetectors has been established [27].
Sub-stoichiometric WO3−x/Si n-n homo-heterojunction with rectification properties has been fabricated by Zhang et al. (Figure 7a–d) [39]. The heterojunction shows remarkable photodetection performance involving a high specific detectivity of 3.96 × 1011 Jones, a large responsivity of 72.8 A/W, and fast response times of 5.8 μs/1.27 ms under 405 nm. The WO3−x/Si n-n homo-heterojunction has the potential for superior visible photodetector characteristics with the scope for optoelectronic applications.

3. Light-Emitting Diode

WO3 has a low emission efficiency due to the presence of a highly linked metal ion polyhedron [40], which limits its use in light-emitting diodes (LEDs). Researchers have fabricated tandem organic light-emitting diodes (OLEDs) connected to WO3 intermediate connections in the early days to fully utilise the photoelectric capabilities of WO3 and make it more useful in LEDs [41,42,43,44]. Zhang et al. used an Al/WO3/Au structure as the interconnection layer of tandem white LED. Through the microcavity effect, this connecting layer makes white light emission more stable and eliminates the angular dependence of the spectrum generated by the microcavity effect. The design of this laminated structure provides an idea for the realization of higher-efficiency organic white LEDs. Performance comparison of a serial OLED with a single LED is shown in Figure 8. Compared with traditional LEDs, serial OLEDs display higher current efficiency, brightness, and longer operating life [41]. Wei et al. investigated a pure blue OLED with a blue fluorescence emitter, in which the charge producing layer was made up of transparent WO3 TFs and the electron transport layer was Li-doped. When compared to single emitting unit devices, series devices have significantly more power and longer service life [42]. Bao et al. studied the electronic structure and the energy level arrangement of WO3-based intermediate connectors. The authors have explained the significance of using WO3 as an interlayer—it can significantly change the energy level arrangement, make the interface dipole and energy level bend, and facilitate the injection of carriers into the appropriate molecular energy level of the adjacent electroluminescent unit [43]. Bin et al. have fabricated a tandem stack to enhance the electroluminescence property of white OLED. In the tandem laminate, 1,4,5,8,9,11-hexaazatriphenylene hexacarbonitrile (HAT-CN) is used as the organic charge generation layer and WO3 is used as the inorganic charge generation layer. In double-stack OLEDs, it is observed that WO3, as the charge generating layer, has the best performance with exceptional CIE color coordinates [44]. Researchers have observed that doping nanostructured WO3 into high molecular polymers (PEDOT:PSS, PANI:PSS) [45,46] as the hole injection layer in classic OLEDs, may be a good choice. Using a simple solid-state mechanochemical approach, Zhuo et al. have produced WO3 nanoribbons with a width of 10 nm and a length of 80 nm and doped them into PEDOT:PSS as a mixed hole injection layer for QDLEDs [45]. Figure 9a,b shows the device structure and its energy level diagram. Compared with the QDLED with PEDOT:PSS as the hole injection layer alone, the QDLED based on the mixed hole injection layer shows high external quantum efficiency(EQE) and stronger current efficiency, as shown in Figure 9c–f. Zhu et al. synthesized WO3 nanocrystals hybridized with conductive polymer (PANI:PSS) and used them in the hole injection layer of OLEDs. The hybrid system reduces the surface defects of WO3 nanocrystals and improves the interface contact ability. Compared with OLED containing only WO3 nanocrystals and traditional PEDOT: PSS-WO3 devices, this device exhibits higher capacitance and stronger luminous efficiency. PANI:PSS-WO3 composite material is expected to be a candidate material as the charge injection layer in new generation OLEDs [46].

4. Photoluminescence Properties

WO3 has been extensively explored in electrochromic, photochromic, and gas sensing materials, and other applications due to its exceptional physical and chemical features. However, being an indirect bandgap semiconductor, WO3 shows lower emission efficiency, which makes its light-emitting characteristics poor. Manfredi et al. investigated light emission in WO3 TFs at liquid nitrogen temperatures in the early years. The light emission ceases while the TFs are at room temperature, indicating that studying the PL of WO3 at room temperature is not acceptable [47]. Researchers have put in a lot of effort to study WO3’s light emission at ambient temperature. For example, Niederberger et al. [48,49,50]. achieved room temperature blue emission of WO3 nanoparticles in ethanol solution. Later, Khold et al. have shown that the morphology, particle size, and quantum confinement play a critical role in the luminescence at room temperature [51], which indicates how to improve the luminescence of WO3 at room temperature. On this basis, Feng et al. prepared crystalline WO3 TFs with different nanostructures by thermal evaporation of tungsten wires and observed strong PL at room temperature [52]. Wang et al. synthesized WO3 nanosheets on a large-scale using tungsten powder as raw materials and demonstrated its blue emission at room temperature [53]. Park et al. synthesized needle-like nanostructures of WO3 by thermal evaporation in the temperature range of 590 °C to 750 °C and explained the influence of substrate temperature and the morphology of WO3 nanostructures on luminescence [54].
Except for the methods mentioned above, doping with rare-earth ions (Eu3+, Tb3+) [40,55,56] or metal ions (Li, Sn, Cu) [57,58,59,60] in WO3 is considered as another effective method to improve luminescence performance at room temperature. Luo et al. prepared Eu3+ doped WO3 TFs by hydrothermal method. They found that with the increasing of Eu ion content, the morphology of WO3 changed significantly. Moreover, Eu3+ doping significantly improved the transparency and optical contrast of WO3 [55]. Ruan et al. prepared WO3:Eu3+ inverse opal photonic crystals and studied their luminescence characteristics. The crystal generated red PL at 615 nm and showed a better red purity [56]. Kavitha and his colleagues used RF magnetron sputtering to prepare high-quality and efficient luminescent Tb3+-doped WO3 TFs, which showed strong green, blue, and red emission under UV excitation [40]. In addition, they also used the same technique to prepare Cu-doped WO3 TFs. The plasmon resonance behavior of Cu nanoparticles on the TFs has been studied, and it has been observed that they can considerably improve the quantum efficiency of various photonic devices and that Cu doping causes the TFs to emit intense blue light (Figure 10) [60]. Kovendhan et al. reported the effect of Li doping (1–5 wt.%) with different contents into the WO3 TFs. They proposed that the structure and optical characteristics of WO3 TFs changed with the increase of lithium, and observed blue PL emission that was difficult to detect at room temperature. The bandgap values of WO3 TFs doped with Li (1–5 wt.%) shifted blue, and the blue emission increased dramatically, as compared to pure WO3 TFs [57]. Mukherjee et al. investigated the preparation of Sn-doped WO3 TFs by chemical spray pyrolysis. They found that the peak intensity of near-band edge emission in doped TFs was enhanced relative to that in undoped TFs, and the spectral intensity was enhanced with the increase of Sn content [58]. Sn-doped WO3 nanosheets were produced using a simple precipitation process by Mehmood et al. and their photoelectric characteristics were examined. SEM images of WO3 nanosheets with different Sn doping levels (0–8 wt.%) are shown in Figure 11a–e. Similar to the results reported by Mukherjee et al. [58], they found that Sn doping significantly enhanced the PL intensity of WO3 nanosheets and resulted in a shift in the near-band edge emission transition (Figure 11f) [59].

5. Electrical Properties

WO3, despite its unique properties of high thermal stability, superior charge transport, tunable electrical properties, and high electron mobility, is not commonly used in the electrical device sector [61,62,63,64]. The electronic devices are used in a variety of environments, including humid, dry, and high-temperature environments, as well as in the dark or under light irradiation. In recent years, researchers are looking at the electrical features of WO3 nanostructures and TFs, including electrical transport properties, electrical conductivity, field emission mechanism, and resistance switching behavior.

5.1. Nanostructured WO3 Electrical Properties

Throughout the last decade, nanoscale electronic and optoelectronic devices involving nanometer-sized inorganic systems have been shown to have comprehensive electrical properties that are sensitive to form and size [65,66,67]. As such, researchers have paid close attention to the electrical characteristics of WO3 nanoparticles.
WO3 is one of the typically unintentionally doped n-type characteristic semiconductors [68]. WO3 often presents sub-stoichiometric oxide (WO3−x) due to the presence of several oxygen deficiencies, such as WO2.9, WO2.83, WO2.8, and WO2.72. That is to say, the lattice of WO3−x could sustain a considerable amount of oxygen vacancy and contain a number of W5+. Consequently, change of oxygen vacancies in WO3−x could effectively tune the density of electron, and then have considerable conductivity [69,70,71]. WO3 that strictly satisfy the stoichiometric ratio should be an insulator, and non-stoichiometric WO3−x exhibits n-type semiconductor behavior. A slight change in oxygen content can also greatly change the conductivity of WO3, so its electrical properties vary with its oxygen content and can be divided into exhibiting metal and semiconductor behavior. Extrinsic n-doping is therefore not required for WO3 to exhibit significant conductivity. Due to the greater bandgap of quasi-two-dimensional (Q2D) WO3, Q2D WO3 nanoflakes have more potential electrical applications. Zhuiykov et al. investigated the electrical characteristics of orthorhombic-WO3 nanoflakes with thicknesses ranging from 7 to 9 nm [72]. Sun et al. prepared high-quality WO3/CoWO4 core-shell p-n junction NWs by hydrothermal method. A physical model of Ag ions spreading along core-shell NWs to form conductive wires was proposed to interpret the bipolar resistance switching behavior. The new WO3/CoWO4 core-shell p-n junction NWs are suitable for the next generation of nonvolatile memory [73].

5.1.1. Oxygen Vacancy Effect

The separate oxygen vacancies effect on WO3 electrical characteristics has been a source of debate [74,75]. The electrical characteristics of tungsten oxides with various oxygen vacancy levels are important to investigate. The thermal evaporation produced nano/microrods with the same morphology that show varying oxygen vacancy content as the WO3−x level increased [76]. The devices composed of WO3−x crystals have shown a negative to positive humidity resistance response. The device’s conductivity was boosted with more oxygen vacancies as a result of a more photogenerated carrier transit and effective separation. The ability to manufacture a range of optoelectronic devices and humidity sensors has been demonstrated by the WO3−x crystal. The current-voltage characteristics related to the Au/WO3 NW/Au devices show that with the increase of bias voltage and oxygen vacancy concentration, the conduction mechanism of the devices changes from volume-limited (space charge-limited) to electrode-limited (Schott Base launch) as have been reported by Yang et al. By adjusting the concentration of oxygen vacancies and even the scanning range of the bias voltage, the resistance switching behavior of WO3 NWs can be adjusted [77].

5.1.2. Electrical Transport Behavior

In recent years, a few applications have focused on the electrical transport characteristics of the WO3 nanostructures. Li et al. studied the high-temperature electrical transport characteristics of hydrothermally produced n-WO3 NRs/p-diamond heterojunctions. Within the temperature range of room temperature to 290 °C, the p-n heterojunction displayed excellent thermal stability and rectification properties (Figure 12a,b). With increasing temperature, the turn-on voltages decreased and the rectification ratio increased. This research broadens the design and application possibilities for heterojunctions based on BDD, particularly at high temperatures, high power, and in a variety of hostile environments [14].
Khan et al. studied the dielectric and electrical transport mechanism of multilayer flower-like WO3 microstructure by impedance spectroscopy. The equivalent circuit model is used to explain the impedance plane. The electrical transport properties of WO3 were studied in detail. The low dielectric loss at 1MHz makes WO3 a potential material for high-frequency applications. Reproduced or adapted from [78].

5.1.3. Field Emission Properties

WO3 NWs show high conductivity owing to abundant oxygen vacancies, which is contributed to the progress of its FE performance. Previous studies have also shown that WO3 NWs have excellent FE performance as a cold cathode potential.
A non-catastrophic breakdown phenomenon was found during the FE process of single defect WO3−x NWs, which can extend the life of the electron source of the NWs. The main reasons for this phenomenon are the defect-related electrical transport-induced breakdown mechanism and the Nottingham effect-induced cooling impact. The detection provided a practical method for designing a single NW point source with a long lifespan, which was critical to the advancement of high-performance semiconductor NW point sources [79]. Uniform large area and micropatterned WO3 NWs were fabricated and their FE properties were investigated by Lin et al. A high FE current up to 3.5 mA was obtained in a defect-rich WO3 sample with an effective area of 0.25 cm2 (Figure 13a–c). The high emission current was incited by the high defect density in the WO3 NWs [80].
A field emission study of well-aligned uniform WO3 nanoconifer was carried out at ~1 × 10−8 mbar pressure. The turn-on and threshold field values were achieved to be 2.43 and 3.08 V µm−1, respectively. The findings of the WO3 nanoconifer FE characteristics investigation suggest that it could be a good candidate for multifunctional applications like cold cathodes in nanoelectronic devices.

5.2. WO3 Thin Film Electrical Properties

WO3 TFs have gained importance in recent years, both in terms of basic research [81] and in terms of their application potential as energy-saving smart windows and batteries [82]. Although many studies have focused on using nanostructured WO3 TFs to increase film optical efficiency, nothing has been done to optimize their electrical properties. However, their mixed conductivity (ionic and electronic) is particularly significant. Samad et al. have studied the relationship between the nanostructure and electrical properties of amorphous WO3 TFs. The ionic conductivity and lithium chemical diffusion coefficient are explored to increase with the supplement of the grain size while the conductivity is proposed to diminish with increasing grain size (Figure 14) [83].
Shanmugasundaram et al. used spray atomizer pyrolysis technology to prepare Sn0.26WO3 TFs with Sn (0, 5, 10, 15 wt.%) and n-Sn0.26WO3/p-Si heterojunction diode 15 wt.% doping concentration at 500 °C substrate temperature. The J-V curve indicates that the 15 wt.% Sn doped WO3 TFs demonstrate a minimum conductivity of 1.429 × 10−9 S/cm. From the diode characterization under illumination and in dark, the acquired ideality factor of n values are 4.89 and 5.03 for 15 wt.% of n-Sn0.26 WO3/p-Si heterojunction diode (Figure 15) [84].
The electrical transport mechanism of p-SnS/n-WO3:Sb TFs heterojunction in the temperature range of 20–300 K was investigated. At low forward bias voltage (<0.25 V), the diode shows Ohmic conduction. At the middle voltage range (0.25 < V < 1.0 V), the carrier transport mechanism followed the thermionic emission. When the forward bias voltage is above 1.0 V, the current transport is space charge limited current (SCLC) according to the exponential trap contribution in the WO3:Sb bandgap. (Figure 16a) The temperature dependence of the ideality factor n and saturation current can be interpreted by tunneling enhanced recombination model emerging at the interface the heterojunction with E00 and Ea values about 99.5 meV and 1.565 eV, respectively (Figure 16) [13].
Pure and Al:WO3/p-Si heterojunction diode was fabricated by the sol-gel spin coating method. The electrical conductivity of the WO3 TFs increases with Al (0–9 wt.%) dopant concentration in the temperature range of 303–473 K (Figure 17). The J-V characteristic of Al:WO3/p-Si diode shows the decreasing barrier heights ΦB at the lower temperature, The decreasing barrier height with low temperature is easily understood considering that the current becomes gradually controlled by electrons that can cross the lower barrier patches, which reduces apparent barrier height. Supposing a Gaussian spatial distribution for ΦB, with mean ΦBm and the standard deviation σB, the temperature dependence of the measured barrier height ΦB at zero applied bias is expected to follow the relation [85]: Φ B = Φ Bm q σ B 2 2 KT . The results show that 9wt.% Al:WO3/p-Si diodes have better performance than other diodes [86].

6. Memory Application

Since traditional silicon-based storage technology will soon reach its limit, this brings a serious challenge to the scaling of the device size. To overcome this limitation, researchers have been trying to find new materials to develop new information storage technologies. In recent years, to solve the scaling limitations of traditional flash memory, research on non-volatile memory has become a top priority. At present, the research on non-volatile memory has become matured. The common non-volatile memory includes magnetic random access memory (MRAM), ferroelectric random access memory (FRAM), resistance random access memory (RRAM), and phase change random access memory (PCM). Compared with other memories, resistance random access memory (RRAM) is easily manufactured, a simple device structure, low operating voltage, high durability, high-density storage, long retention time, fast switching speed, and excellent performance. It has the advantages of scalability and compatibility with traditional complementary metal-oxide-semiconductor (CMOS) technology [87,88]. To exhibit high performance, the size of the memory device should be kept small, because the large devices exhibit higher noise levels and poor reproducibility. This is due to uncontrollable deviations and defects within the larger devices [89]. Therefore, building nano-level devices has become a necessity to solve this problem.
Because of its outstanding performance in three-dimensional stacking, compatibility with CMOS process [90,91,92], good durability [93], non-volatile rectification characteristics [94] with excellent electrochromic and photochromic properties, and other factors, nanostructured tungsten oxide (NWs, NRs, nanosheets, etc.) has been widely used in RRAM devices in recent years. Among these advantages, the electrochromic properties of WO3 are one of the criteria that determines the storage performance of RRAM. The degree of crystallinity of tungsten oxide determines its electrochromic performance. Compared with crystal-oriented tungsten oxide, amorphous tungsten oxide shows better electrochromic performance. However, amorphous tungsten oxide exhibits poor durability in acid electrolyte solutions [95]. To solve the above problems, researchers have studied wet chemical preparation methods to construct tungsten oxide composite systems, such as electrodeposition [96,97] and sol-gel methods [98,99]. Shim et al. prepared polycrystalline WO3 NWs on ITO substrates by electrospinning. Compared with traditional WO3 nano-films, they found that the prepared WO3 NWs showed faster charge transfer, better optical response, and better pigmentation efficiency, and the memory effect after pigmentation also significantly improved [100]. Pang et al. synthesized WO3 composite TFs modified with Ag nanoparticles by a combination of vacuum deposition and electrodeposition methods. Compared with pure WO3 TFs, the films show stronger electroactivity and electrochromic properties [101].
Kozicki et al. demonstrated that low-power RRAM devices doped with copper WO3 are achievable. This type of device exhibits a high turn-off resistance and can be switched to an on-resistance state under a low voltage state. This state is independent of the device geometry but is strongly controlled by the current. If a small reverse bias is applied, they can be restored to a high-impedance state. Schematic diagram of the device structure and its resistance switching characteristic is shown in Figure 18. Unfortunately, the manufacture of such devices requires the diffusion of copper light into WO3, which will be difficult to apply in semiconductor processing [90]. On this basis, Li and co-workers fabricated RRAM devices using Cu/WO3/Pt structures. This device exhibits better resistive switching characteristics, such as better durability, lower power consumption, and better retention. In addition, they also explained the physical mechanism of the multi-level storage characteristics in Cu/WO3/Pt storage devices. The origin of these multi-level storage characteristics can be explained as that when a higher current is applied to the device, the radial growth of conductive filaments and the formation of more conductive filaments together, lead to this characteristic [92,102]. Inspired by this phenomenon, Kozicki et al., Li et al., and Sun et al. synthesized high-quality WO3/CoWO4 core-shell NWs using the hydrothermal method. They studied the bipolar resistance switching behavior of Ag/[WO3/COWO4]/Ag devices using Ag as an electrode. They found that the device maintained excellent stability for 100 cycles and had an on/off ratio of up to 333 at room temperature [73]. Similarly, Chakrabarti et al. designed a new RRAM device with W/WO3/WOx/W structure and observed the effects of post-annealing of metal on the behavior of shapeless resistance switch in this structure, especially the F-N tunneling effect after LRS and reset. Furthermore, for all nonlinear current-voltage switching characteristics, the authors have used simulations to account for SCLC conduction in low voltage field, F-N tunneling in high voltage field, and oxygen vacancy carbon fiber with a diameter of ~34 nm. This work will contribute to comprehend the switching mechanism of other similar RRAM structures and selector-free nanoscale crossover structures [103].
Recently, Sun et al. have fabricated ITO/WO3/AZO resistance switching devices by a magnetron sputtering and observed that the devices have significantly enhanced resistance-switching memory behavior. The schematic diagram of the ITO/WO3/AZO device and I-V characteristic curve are shown in Figure 19. Subsequently, they proposed a physical model of photogenerated carriers tunneling in the Schottky barrier layer driven by electrical pulses to fully explain this phenomenon. The deployment of non-volatile RRAM devices in future development will be guided by this physical model [104]. Singh and his colleagues synthesized Ag-decorated WO3 NWs by the glancing angle deposition (GLAD) technique. The growth process of Ag-decorated WO3 NWs is shown in Figure 20a. Figure 20b shows the I-V characteristic curves of the device, exhibiting a large storage window of ~ 12.02 V at ±10 V and a low interface density trap of ~5.74 × 1010 eV−1 cm−2 at 1 MHz. It also exhibits an on/off switching time lasting up to 1500 cycles (Figure 20c) and a stable retention time of up to 103 s on/off resistance ratio (~245) [105]. Moreover, the same technology was used to synthesize WO3 NWs based capacitive memory on Si substrates. Through the measurement and analysis of the device performance, the authors found that the memory exhibited a stable retention time (103) and a good endurance cycle of up to 100 [106]. These works provide interesting ideas for the design and application of next-generation non-volatile memory.

7. Conclusions and Future Outlook

Recent research has significantly increased our knowledge of the features and uses of WO3-related nanostructures in optical and electrical devices, such as photodetectors, LEDs, PL properties, electrical properties, and optical information storage devices. Although WO3 based devices exhibit good performance in the fields of various photoelectrical applications, it is also desirable to find new fabrication routes and the types of electrodes for these devices to improve the optical and electrical properties in the future. Modifying the electrode configuration of the device, controlling the variety of morphologies of WO3 nanostructures, and optimizing the preparation process could be the effective strategy for various futuristic photoelectrical applications. In order to better understand the physical transport mechanism of the devices, it is critical to use more relevant semiconductor theory and computational models for carrier transport research toward the development of WO3 devices.
Up to present, the WO3 nanostructure related tunnel diode with negative differential resistance (NDR) investigation is scarcely researched, NDR is a non-linear carrier transport phenomenon, whereby the electrical current decreases with increasing bias voltage. N-WO3 semiconductor exerts degenerative features through heavy n-type doping and possibly displays a NDR phenomenon when combined with p-degenerative semiconductor. The NDR effect of WO3 will make an important contribution to the implementation of logic switches, oscillators, inverters, resistive switching memory, and radiation reliable device applications in the field of flexible electronics semiconductors.
Over the past decades, the development of the optoelectronic applications based on the WO3 nanostructures that can operate in harsh environments (high temperature or strong radiation environments) is still challenging. It is also necessary to put further efforts to investigate WO3 related optical and electrical devices under extreme conditions such as high temperature, high pressure, and harsh environments. Diamond is an excellent semiconductor material for manufacturing high-performance electronic devices that are used in high temperatures and a strong radiation environment. Therefore, it is expected to fabricate WO3/diamond heterojunction device for providing the possibility of photodetectors, LEDs, PL properties, and optical information storage devices application at higher temperatures.

Author Contributions

Writing—original draft preparation, Y.Y., L.Z. and D.S.; writing—review and editing, D.S., Q.W. and C.L.; Supervision, D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 62104090, 11604133 and 11874174), the Natural Science Foundation of Shandong Province (Grant No. ZR2017QA013), the Introduction and Cultivation Plan of Youth Innovation Talents for Universities of Shandong Province, the Science and Technology Plan of Youth Innovation Team for Universities of Shandong Province (Grant No. 2019KJJ019), the Open Project of State Key Laboratory of Superhard Materials (Jilin University, Grant Nos. 201503 and 201612), the Research Funding of Liaocheng University (Grant Nos. 318012016, 318051610, and 318051612), and the Special Construction Project Fund for Shandong Province Taishan Scholars.

Data Availability Statement

Data available in a publicly accessible repository.

Acknowledgments

The authors would like to express their gratitude to EditSprings (https://www.editsprings.com/, 2021.6.20–2021.6.25) for the expert linguistic services provided.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yu, P.; Yang, H.; Chen, X.; Yi, Z.; Yao, W.; Chen, J.; Yi, Y.; Wu, P. Ultra-wideband solar absorber based on refractory titanium metal. Renew. Energy 2020, 158, 227–235. [Google Scholar] [CrossRef]
  2. Zhao, F.; Chen, X.; Yi, Z.; Qin, F.; Tang, Y.; Yao, W.; Zhou, Z.; Yi, Y. Study on the solar energy absorption of hybrid solar cells with trapezoid-pyramidal structure based PEDOT:PSS/c-Ge. Sol. Energy 2020, 204, 635–643. [Google Scholar] [CrossRef]
  3. Najafi-Ashtiani, H.; Bahari, A.; Ghasemi, S. A Dual electrochromic film based on nanocomposite of copolymer and WO3 nanoparticles: Enhanced electrochromic coloration efficiency and switching response. J. Electroanal. Chem. 2016, 774, 14–21. [Google Scholar] [CrossRef]
  4. Najafi-Ashtiani, H.; Bahari, A. Optical, structural and electrochromic behavior studies on nanocomposite thin film of aniline, o-toluidine and WO3. Opt. Mater. 2016, 58, 210–218. [Google Scholar] [CrossRef]
  5. Yue, S.; Pan, H.; Ning, Z.; Yin, J.; Wang, Z.; Zhang, G. Amazing ageing property and in situ comparative study of field emission from tungsten oxide nanowires. Nanotechnology 2011, 22, 115703. [Google Scholar] [CrossRef] [PubMed]
  6. Teoh, L.G.; Shieh, J.; Lai, H.W.; Hung, I.M.; Hon, M.H. Structure and optical properties of mesoporous tungsten oxide. J. Alloy. Compd. 2005, 396, 251–254. [Google Scholar] [CrossRef]
  7. Dai, J.; Li, Y.; Ruan, H.; Ye, Z.; Chai, N.; Wang, X.; Qiu, S.; Bai, W.; Yang, M. Fiber optical hydrogen sensor based on WO3-Pd2Pt-Pt nanocomposite films. Nanomaterials 2021, 11, 128. [Google Scholar] [CrossRef]
  8. Lin, H.; Long, X.; An, Y.; Yang, S. In situ growth of Fe2WO6 on WO3 nanosheets to fabricate heterojunction arrays for boosting solar water splitting. J. Chem. Phys. 2020, 152, 214704. [Google Scholar] [CrossRef]
  9. Fularz, A.; Almohammed, S.; Rice, J.H. Controlling plasmon-induced photocatalytic redox reactions on WO3 nanowire/AgNPs substrates via defect engineering. J. Phys. Chem. C 2020, 124, 25351–25360. [Google Scholar] [CrossRef]
  10. Ji, Y.; Yang, Y.; Lee, S.K.; Ruan, G.; Kim, T.W.; Fei, H.; Lee, S.H.; Kim, D.Y.; Yoon, J.; Tour, J.M. Flexible nanoporous WO3−x nonvolatile memory device. ACS Nano 2016, 10, 7598–7603. [Google Scholar] [CrossRef]
  11. 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]
  12. Karthik, Y.P.V.; Ajitha, B.; Kumar Reddy, Y.A.; Minnam Reddy, V.R.; Reddeppad, M.; Kim, M.D. Effect of sputter pressure on UV photodetector performance of WO3 thin films. Appl. Surf. Sci. 2021, 536, 147947. [Google Scholar] [CrossRef]
  13. Gaewdang, T.; Wongcharoen, N. Temperature-dependent electrical transport characteristics of p-SnS/n-WO3:Sb heterojunction diode. Mater. Sci. Eng. 2018, 383, 012006. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, L.; Cheng, S.; Wu, C.; Pei, K.; Song, Y.; Li, H.; Wang, Q.; Sang, D. Fabrication and high temperature electronic behaviors of n-WO3 nanorods/p-diamond heterojunction. Appl. Phys. Lett. 2017, 110, 052106. [Google Scholar] [CrossRef]
  15. Kim, Y.H.; Lee, S.Y.; Umh, H.N.; Song, H.D.; Han, J.W.; Choi, J.W.; Yi, J. Directional change of interfacial electric field by carbon insertion in heterojunction system TiO2/WO3. ACS Appl. Mater. Interfaces 2020, 12, 15239–15245. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, J.; Lu, H.; Liu, C.; Chen, C.; Xin, X. Porous NiO-WO3 heterojunction nanofibers fabricated by electrospinning with enhanced gas sensing properties. RSC Adv. 2017, 7, 40499–40509. [Google Scholar] [CrossRef] [Green Version]
  17. Li, N.; Fu, S.; Wu, J.; Li, X.; Zhou, J.; Wang, Y.; Zhang, X.; Liu, Y. WO3/ZnO nanowire heterojunction as hole transport channel for building up persistent holographic fringes. Appl. Phys. Lett. 2020, 116, 251606. [Google Scholar] [CrossRef]
  18. Yin, C.; Zhu, S.; Zhang, D. 3D nanostructured WO3/BiVO4 heterojunction derived from Papilio paris for efficient water splitting. RSC Adv. 2017, 7, 27354. [Google Scholar] [CrossRef] [Green Version]
  19. Tahir, M.B.; Ali, S.; Rizwan, M. A review on remediation of harmful dyes through visible light-driven WO3 photocatalytic nanomaterials. Int. J. Environ. Sci. Technol. 2019, 16, 4975–4988. [Google Scholar] [CrossRef]
  20. Dutta, V.; Sharma, S.; Raizada, P.; Thakur, V.K.; Khan, A.A.P.; Saini, V.; Asiri, A.M.; Singh, P. An overview on WO3 based photocatalyst for environmental remediation. J. Environ. Chem. Eng. 2021, 9, 105018. [Google Scholar] [CrossRef]
  21. Mehmood, A.; Long, X.; Haidry, A.A.; Zhang, X. Trends in sputter deposited tungsten oxide structures for electrochromic applications: A review. Ceram. Int. 2020, 46, 23295–23313. [Google Scholar] [CrossRef]
  22. Zhong, X.; Liu, X.; Diao, X. Electrochromic devices based on tungsten oxide and nickel oxide: A review. J. Inorg. Mater. 2021, 36, 128–139. [Google Scholar] [CrossRef]
  23. Dong, C.; Zhao, R.; Yao, L.; Ran, Y.; Zhang, X.; Wang, Y. A review on WO3 based gas sensors: Morphology control and enhanced sensing properties. J. Alloy. Compd. 2020, 15, 153194. [Google Scholar] [CrossRef]
  24. Wei, Z.; Xu, L.; Peng, S.; Zhou, Q. Application of WO3 Hierarchical structures for the detection of dissolved gases in transformer oil: A mini review. Front Chem. 2020, 8, 188. [Google Scholar] [CrossRef]
  25. Bandi, S.; Srivastav, A.K. Review: Oxygen-deficient tungsten oxides. J. Mater. Sci. 2021, 56, 6615–6644. [Google Scholar] [CrossRef]
  26. Bechinger, C.; Wirth, E.; Leiderer, P. Photochromic coloration of WO3 with visible light. Appl. Phys. Lett. 1996, 68, 2834–2836. [Google Scholar] [CrossRef] [Green Version]
  27. Wang, H.; Liu, J.L.; Wu, X.X.; Zhang, S.Q.; Zhang, Z.K.; Pan, W.W.; Yuan, G.; Yuan, C.L.; Ren, Y.L.; Lei, W. Ultra-long high quality catalyst-free WO3 nanowires for fabricating high-performance visible photodetectors. Nanotechnology 2020, 31, 274003. [Google Scholar] [CrossRef]
  28. Li, X.L.; Lou, T.J.; Sun, X.M.; Li, Y.D. Highly sensitive WO3 hollow-sphere gas sensors. Inorg. Chem. 2004, 43, 5442–5449. [Google Scholar] [CrossRef]
  29. Huo, N.; Yang, S.; Wei, Z.; Li, J. Synthesis of WO3 nanostructures and their ultraviolet photoresponse properties. J. Mater. Chem. C 2013, 1, 3999–4007. [Google Scholar] [CrossRef]
  30. Li, L.; Zhang, Y.; Fang, X.; Zhai, T.; Liao, M.; Sun, X.; Koide, Y.; Bando, Y.; Golberg, D. WO3 nanowires on carbon papers: Electronic transport, improved ultraviolet-light photodetectors and excellent field emitters. J. Mater. Chem. 2011, 21, 6525–6530. [Google Scholar] [CrossRef]
  31. Liu, J.; Zhong, M.; Li, J.; Pan, A.; Zhu, X. Few-layer WO3 nanosheets for high-performance UV-photodetectors. Mater. Lett. 2015, 148, 184–187. [Google Scholar] [CrossRef]
  32. Wan, J.H.; Huang, K.; Wan, L. Shottky contact device based on a single WO3 nanowire for ultraviolet photodetector. Appl. Mech. Mater. 2014, 556, 2097–2100. [Google Scholar] [CrossRef]
  33. Kim, Y.; Lee, S.H.; Jeong, S.; Kim, B.J.; Choi, J.Y.; Yu, H.K. Conversion of WO3 thin films into self-crosslinked nanorods for large-scale ultraviolet detection. RSC Adv. 2020, 10, 14147–14153. [Google Scholar] [CrossRef] [Green Version]
  34. Kumar Reddy, Y.A.; Ajitha, B.; Sreedhar, A.; Eswaraiah, V. Enhanced UV photodetector performance in bi-layer TiO2/WO3 sputtered films. Appl. Surf. Sci. 2019, 494, 575–582. [Google Scholar] [CrossRef]
  35. Karthik Yadava, P.V.; Kumar Reddy, Y.A.; Ajitha, B.; Minnam Reddyc, V.R. Oxygen partial pressure dependent UV photodetector performance of WO3 sputtered thin films. J. Alloy. Compd. 2020, 816, 152565. [Google Scholar] [CrossRef]
  36. Marnadu, R.; Chandrasekaran, J.; Maruthamuthu, S.; Balasubramani, V.; Vivek, P.; Suresh, R. Ultra-high photoresponse with superiorly sensitive metal-insulator-semiconductor (MIS) structured diodes for UV photodetector application. Appl. Surf. Sci. 2019, 480, 308–322. [Google Scholar] [CrossRef]
  37. Marnadu, R.; Chandrasekaran, J.; Maruthamuthu, S.; Vivek, P.; Balasubramani, V.; Balraju, P. Jet nebulizer sprayed WO3-nanoplate arrays for high-photoresponsivity based metal–insulator–semiconductor structured schottky barrier diodes. J. Inorg. Organomet. Polym. Mater. 2020, 30, 731–748. [Google Scholar] [CrossRef]
  38. Raj, M.; Joseph, C.; Subramanian, M.; Perumalsamy, V.; Elayappan, V. Superior photoresponse MIS Schottky barrier diodes with nanoporous: Sn-WO3 films for ultraviolet photodetector application. N. J. Chem. 2020, 44, 7708–7718. [Google Scholar] [CrossRef]
  39. Zhang, X.; Su, Y.; Tang, Z.; Hu, D.; Wang, Z.; Hou, Y.; Wang, X. WO3−x/Si n-n homotype heterojunction with high performance photodetection characteristics. Scr. Mater. 2020, 189, 89–94. [Google Scholar] [CrossRef]
  40. Kavithaa, V.S.; Krishnana, R.R.; Sreedharana, R.S.; Suresh, K.; Jayasankar, C.K.; Mahadevan Pillai, V.P. Tb3+-doped WO3 thin films: A potential candidate in white light emitting devices. J. Alloy. Compd. 2019, 788, 429–445. [Google Scholar] [CrossRef]
  41. Zhang, H.; Dai, Y.; Ma, D. Al/WO3/Au as the interconnecting layer for efficient tandem white organic light-emitting diodes. J. Phys. D Appl. Phys. 2008, 41, 102006. [Google Scholar] [CrossRef]
  42. Wei, F.; Fang, L. Pure-blue tandem OLEDs based on terfluorenes compounds. J. Mater. Sci. Mater. Electron. 2008, 19, 1202–1205. [Google Scholar] [CrossRef]
  43. Bao, Q.Y.; Yang, J.P.; Tang, J.X.; Li, Y.; Lee, C.S.; Lee, S.T. Interfacial electronic structures of WO3-based intermediate connectors in tandem organic light-emitting diodes. Org. Electron. 2010, 11, 1578–1583. [Google Scholar] [CrossRef]
  44. Bin, J.K.; Lee, N.Y.; Lee, S.; Seo, B.; Yang, J.H.; Kim, J.; Yoon, S.Y.; Kang, I. Two stacked tandem white organic light-emitting diodes employing WO3 as a charge generation layer. Org. Light Emit. Mater. Devices 2016, 9941, 99411T. [Google Scholar]
  45. Zhuo, M.P.; Liang, F.; Shi, Y.L.; Yun, H.; Wang, R.B.; Chen, W.F.; Wang, X.D.; Liao, L.S. WO3 nanobelt doped PEDOT:PSS layers for efficient hole-injection in quantum dot light-emitting diodes. J. Mater. Chem. C 2017, 5, 12343–12348. [Google Scholar] [CrossRef]
  46. Zhu, W.; Chen, X.L.; Chang, J.; Yu, R.M.; Li, H.; Liang, D.; Wu, X.Y.; Wang, Y.; Lu, C.Z. Doped polyaniline-hybridized tungsten oxide nanocrystals as hole injection layers for efficient organic light-emitting diodes. J. Mater. Chem. C 2018, 6, 7242–7248. [Google Scholar] [CrossRef]
  47. Manfredi, M.; Paracchini, C.; Salviati, G.C.; Schianchi, G. Conductive processes in transparent WO3 films irradiated with ultraviolet light. Thin Solid Film. 1981, 79, 161. [Google Scholar] [CrossRef]
  48. Niederberger, M.; Bartl, M.H.; Stucky, G.D. Benzyl alcohol and transition metal chlorides as a versatile reaction system for the nonaqueous and low-temperature synthesis of crystalline nano-objects with controlled dimensionality. J. Am. Chem. Soc. 2002, 124, 13642–13643. [Google Scholar] [CrossRef]
  49. Lee, K.; Seo, W.S.; Park, J.T. Synthesis and optical properties of colloidal tungsten oxide nanorods. J. Am. Chem. Soc. 2003, 125, 3408–3409. [Google Scholar] [CrossRef] [PubMed]
  50. Takagahara, T.; Takeda, K. Theory of the quantum confinement effect on excitons in quantum dots of indirect-gap materials. Phys. Rev. B 1992, 46, 15578. [Google Scholar] [CrossRef]
  51. Khold, A.N.; Shaposhnikov, V.L.; Sobolev, N.; Borisenko, V.E.; D’Avitaya, F.A.; Ossicini, S. Orientation effects in the electronic and optical properties of germanium quantum wires. Phys. Rev. B 2004, 70, 035317. [Google Scholar] [CrossRef]
  52. Feng, M.; Pan, A.L.; Zhang, H.R.; Li, Z.A.; Liu, F.; Liu, H.W.; Shi, D.X.; Zou, B.S.; Gao, H.J. Strong photoluminescence of nanostructured crystalline tungsten oxide thin films. Appl. Phys. Lett. 2005, 86, 141901. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, J.; Lee, P.S.; Ma, J. Synthesis, growth mechanism and room-temperature blue luminescence emission of uniform WO3 nanosheets with W as starting material. J. Cryst. Growth 2009, 31, 316–319. [Google Scholar] [CrossRef]
  54. Park, S.; Kim, H.; Jin, C.; Lee, C. Intense ultraviolet emission from needle-like WO3 nanostructures synthesized by noncatalytic thermal evaporation. Nanoscale Res. Lett. 2011, 6, 451. [Google Scholar] [CrossRef] [Green Version]
  55. Luo, G.; Shen, L.; Zheng, J.; Xu, C.A. Europium ion doped WO3 film with the bi-functionality of enhanced electrochromic switching and tunable red emission. J. Mater. Chem. C 2017, 5, 3488–3494. [Google Scholar] [CrossRef]
  56. Ruan, J.; Yang, Z.; Huang, A.; Chai, Z.; Qiu, J.; Song, Z. Red photoluminescent property and modification of WO3:Eu3+ inverse opal for blue light converted LEDs. Opt. Mater. 2018, 75, 224–229. [Google Scholar] [CrossRef]
  57. Kovendhan, M.; Joseph, D.P.; Kumar, E.S.; Sendilkumar, A.; Manimuthu, P.; Sambasivam, S.; Venkateswaran, C.; Mohan, R. Structural transition and blue emission in textured and highly transparent spray deposited Li doped WO3 thin films. Appl. Surf. Sci. 2011, 257, 8127–8133. [Google Scholar] [CrossRef]
  58. Mukherjee, R.; Prajapati, C.S.; Sahay, P.P. Tin-incorporation induced changes in the microstructural, optical, and electrical behavior of tungsten oxide nanocrystalline thin films grown via spary pyrolysis. J. Therm. Spray Technol. 2014, 23. [Google Scholar]
  59. Mehmood, F.; Iqbal, J.; Jan, T.; Ahmed, W.; Ahmed, W.; Arshad, A.; Mansoor, Q.; Ilyas, S.Z.; Ismail, M.; Ahmad, I. Effect of Sn doping on the structural, optical, electrical and anticancer properties of WO3 nanoplates. Ceram. Int. 2016, 42, 14334–14341. [Google Scholar] [CrossRef]
  60. Kavitha, V.S.; Chalana, S.R.; Krishnan, R.R.; Gopchandran, K.G.; Mahadevan Pillai, V.P. High quality, highly transparent Cu incorporated WO3 thin films suitable for blue LED application. Vacuum 2020, 172, 109044. [Google Scholar] [CrossRef]
  61. Zheng, J.Y.; Haider, Z.; Van, T.K.; Pawar, A.U.; Kang, M.J.; Kim, C.W.; Kang, Y.S. Tuning of the crystal engineering and photo-electrochemical properties of crystalline tungsten oxide for optoelectronic device applications. CrystEngComm 2015, 17, 6070–6603. [Google Scholar] [CrossRef]
  62. Papadimitropoulos, G.; Vourdas, N.; Giannakopoulos, K.; Vasilopoulou, M.; Davazoglou, D. Porous hot-wire deposited WO3 films with high optical transmission. J. Appl. Phys. 2011, 109, 103527. [Google Scholar] [CrossRef]
  63. Xin, Y.; Zhou, H.; Ni, X.; Pan, Y.; Zhang, X.; Zheng, J.Y.; Bao, S.; Jin, P. The optical properties of low infrared transmittance WO3−x nanocrystal thin films prepared by DC magnetron sputtering under different oxygen ratios. RSC Adv. 2015, 5, 57757–57763. [Google Scholar] [CrossRef]
  64. Li, W.; Sasaki, A.; Oozu, H.; Aoki, K.; Kakushima, K.; Kataoka, Y.; Nishiyama, A.; Sugii, N.; Wakabayashi, H.; Tsutsui, K.; et al. Electron transport Mechanism of tungsten trioxide powder thin film studied by investigating the effect of annealing on resistivity. Microelectron. Reliab. 2015, 55, 407–410. [Google Scholar] [CrossRef]
  65. Teo, B.K.; Sun, X.H.; Hung, T.F.; Meng, X.M.; Wong, N.B.; Lee, S.T. Precision-cut crystalline silicon nanodots and nanorods from nanowires and direct visualization of cross sections and growth orientations of silicon nanowires. Nano Lett. 2003, 3, 1735. [Google Scholar] [CrossRef]
  66. Qian, C.; Kim, F.; Ma, L.; Tsui, F.; Yang, P.; Liu, J. Solution-Phase synthesis of single-crystalline iron phosphide nanorods/nanowires. J. Am. Chem. Soc. 2004, 126, 1195. [Google Scholar] [CrossRef]
  67. Acharya, S.; Efrima, S. Two-Dimensional pressure-driven nanorod-to-nanowire reactions in langmuir monolayers at room temperature. J. Am. Chem. Soc. 2005, 127, 3486–3490. [Google Scholar] [CrossRef]
  68. Yao, Y.; Sang, D.; Duan, S.; Wang, Q.; Liu, C. Excellent optoelectronic applications and electrical transport behavior of the n-WO3 nanostructures/p-diamond heterojunction: A new perspective. Nanotechnology 2021, 32, 332501. [Google Scholar] [CrossRef]
  69. Chen, J.; Ren, Y.M.; Hu, T.Z.; Xu, T.; Xu, Q. Fabrication and application of substoichiometric tungsten oxide with tunable localized surface plasmon resonances. Appl. Surf. Sci. 2019, 465, 517–525. [Google Scholar] [CrossRef]
  70. Cheng, H.F.; Klapproth, M.; Sagaltchik, A.; Li, S.; Thomas, A. Ordered mesoporous WO2.83: Selective reduction synthesis, exceptional localized surface plasmon resonance and enhanced hydrogen evolution reaction activity. J. Mater. Chem. A 2018, 6, 2249–2256. [Google Scholar] [CrossRef]
  71. Guo, T.C.; Ling, C.C.; Zhang, T.; Li, H.; Li, X.F.; Chang, X.; Zhu, L.; Zhao, L.; Xue, Q.Z. High-performance WO3−x-WSe2/SiO2/n-Si heterojunction near-infrared photodetector via a homo-doping strategy. J. Mater. Chem. C 2018, 6, 5821–5829. [Google Scholar] [CrossRef]
  72. Zhuiykov, S.; Kats, E. Enhanced electrical properties in sub-10-nm WO3 nanoflakes prepared via a two-step sol-gel-exfoliation method. Nanoscale Res. Lett. 2014, 401, 1–10. [Google Scholar] [CrossRef] [Green Version]
  73. Sun, B.; Li, H.; Wei, L.; Chen, P. Hydrothermal synthesis and resistive switching behaviour of WO3/CoWO4 core–shell nanowires. CrystEngComm 2014, 16, 9891–9895. [Google Scholar] [CrossRef]
  74. Niklasson, G.A.; Berggren, L.; Larsson, A.L. Electrochromic tungsten oxide: The role of defects. Sol. Energy Mater. Sol. Cells 2004, 84, 315–328. [Google Scholar] [CrossRef]
  75. Gillet, M.; Lemire, C.; Gillet, E.; Aguir, K. The role of surface oxygen vacancies upon WO3 conductivity. Surf. Sci. 2003, 532, 519–525. [Google Scholar] [CrossRef]
  76. Shen, Z.; Zhao, Z.Y.; Wen, J.; Qian, J.; Peng, Z.; Fu, X. Role of oxygen vacancies in the electrical properties of WO3−x nano/microrods with identical morphology. J. Nanomater. 2018, 2018, 7802589. [Google Scholar] [CrossRef] [Green Version]
  77. Yang, W.; Yan, Y.; Peng, Y.; Zhao, G.; Liu, Y.; Li, S.; Wang, W.; Zhou, W.; Tang, D. Effect of oxygen vacancies on the resistive switching behavior of hexagonal WO3 nanowire. Mater. Res. Express 2019, 6, 085072. [Google Scholar] [CrossRef]
  78. Khan, A.A.; Khan, M.N.; Lqbal, M.; Majeed, A.; Jalil, A.; Mukhtar, S.; Javed, M. Temperature dependent charge conduction and relaxation mechanism study of nano-structure WO3 by impedance spectroscopy. Mater. Res. Express 2019, 6, 1250a5. [Google Scholar] [CrossRef]
  79. Lin, Z.; Zhan, R.; Chen, Y.; She, J.; Deng, S.; Xu, N.; Chen, J. Defective WO3−x nanowire: Possible long lifetime semiconductor nanowire point electron source. Nanoscale 2019, 11, 3370–3377. [Google Scholar] [CrossRef]
  80. Lin, Z.; Xie, P.; Zhan, R.; Chen, D.; She, J.; Deng, S.; Xu, N.; Chen, J. Defect-enhanced field emission from WO3 nanowires for flat-panel X-ray sources. ACS Appl. Nano Mater. 2019, 2, 5206–5213. [Google Scholar] [CrossRef]
  81. Garcia-Garcia, F.J.; Gil-Rostra, J.; Yubero, F.; González-Elipe, A.R. Electrochromism in WOx and WxSiyOz thin films prepared by Magnetron sputtering at glancing angles. Nanosci. Nanotechnol. Lett. 2013, 5, 89–93. [Google Scholar] [CrossRef]
  82. Garcia-Garcia, F.J.; Mosa, J.; Gonzalez-Elipe, A.R.; Aparicio, M. Sodium ion storage performance of magnetron sputtered WO3 thin films. Electrochim. Acta 2019, 321, 134669. [Google Scholar] [CrossRef]
  83. Samad, B.A.; Ashirt, P.V. Optimization of mixed conductivity through nanostructure control in WO3 thin films. Thin Solid Films 2017, 636, 717–722. [Google Scholar] [CrossRef]
  84. Shanmugasundaram, K.; Thirunavukkarasu, P.; Balaji, M. Effect of Sn doping on the structural, optical, electrical properties and diode characteristics of WO3 thin films deposited by Jet Nebulizer Spray technique. Mater. Today Proc. 2019, 18, 1648–1657. [Google Scholar] [CrossRef]
  85. Bartolomeo, A.D.; Giubileo, F.; Luongo, G.; Iemmo, L.; Martucciello, N.; Niu, G.; Fraschke, M.; Skibitzki, O.; Schroeder, T.; Lupina, G. Tunable Schottky barrier and high responsivity ingraphene/Si-nanotip optoelectronic device. 2D Mater. 2017, 4, 015024. [Google Scholar] [CrossRef] [Green Version]
  86. Raja, M.; Chandrasekaran, J.; Balaji, M.; Kathirvel, P. Investigation of microstructural, optical and dc electrical properties of spin coated Al:WO3 thin films for n-Al:WO3/p-Si heterojunction diodes. Optik 2017, 145, 169–180. [Google Scholar] [CrossRef]
  87. Akihito, S. Resistive switching in transition metal oxides. Mater. Today 2008, 11, 28–36. [Google Scholar]
  88. Kukreja, L.M.; Das, A.K.; Misra, P. Studies on nonvolatile resistance memory switching in ZnO thin films. Bull. Mater. Sci. 2009, 32, 247–252. [Google Scholar] [CrossRef] [Green Version]
  89. Qu, B.; Younis, A.; Chu, D. Recent progress in tungsten oxides based memristors and their neuromorphological applications. Electron. Mater. Lett. 2016, 12, 715–731. [Google Scholar] [CrossRef]
  90. Kozicki, M.N.; Gopalan, C.; Balakrishnan, M.; Mitkova, M. A Low-power nonvolatile Switching element based on copper-tungsten oxide solid electrolyte. IEEE Trans. Nanotechnol. 2006, 5, 535–544. [Google Scholar] [CrossRef]
  91. Chien, W.C.; Chen, Y.C.; Lai, K.E.; Yao, Y.D.; Lin, P.; Horng, S.F.; Gong, J.; Chou, T.H.; Lin, H.M.; Chang, M.N.; et al. Unipolar switching behaviors of RTO WOx RRAM. IEEE Electron Device Lett. 2010, 31, 126–128. [Google Scholar] [CrossRef]
  92. Li, Y.T.; Long, S.B.; Lv, H.B.; Liu, Q.; Wang, Q.; Wang, Y.; Zhang, S.; Lian, W.T.; Liu, S.; Liu, M. Investigation of resistive switching behaviours in WO3-based RRAM devices. Chin. Phys. B 2011, 20, 017305. [Google Scholar] [CrossRef]
  93. Wu, W.T.; Wu, J.J.; Chen, J.S. Resistive switching behavior and multiple transmittance states in solution-processed tungsten oxide. ACS Appl. Mater. Interfaces 2011, 3, 2616–2621. [Google Scholar] [CrossRef] [PubMed]
  94. Yang, R.; Terabe, K.; Tsuruoka, T.; Hasegawa, T.; Aono, M. Oxygen migration process in the interfaces during bipolar resistance switching behavior of WO3−x-based nanoionics devices. Appl. Phys. Lett. 2012, 100, 231603. [Google Scholar] [CrossRef]
  95. Song, Y.Y.; Gao, Z.D.; Wang, J.H.; Xia, X.H.; Lynch, R. Multistage coloring electrochromic device based on TiO2 nanotube arrays modified with WO3 nanoparticles. Adv. Funct. Mater. 2011, 21, 1941–1946. [Google Scholar] [CrossRef]
  96. Habazaki, H.; Hayashi, Y.; Konno, H. Characterization of electrodeposited WO3 films and its application to electrochemical wastewater treatment. Electrochim. Acta 2002, 47, 4181–4188. [Google Scholar] [CrossRef]
  97. Deepa, M.; Kar, M.; Agnihotry, S.A. Electrodeposited tungsten oxide films: Annealing effects on structure and electrochromic performance. Thin Solid Films 2004, 468, 32–42. [Google Scholar] [CrossRef]
  98. Biswas, P.K.; Pramanik, N.C.; Mahapatra, M.K.; Ganguli, D.; Livage, J. Optical and electrochromic properties of sol-gel WO3 films on conducting glass. Mate. Lett. 2003, 57, 4429–4432. [Google Scholar] [CrossRef]
  99. Sharma, N.; Deepa, M.; Varshney, P.; Agnihotry, S.A. Influence of organic additive on the morphological, electrical and electrochromic properties of sol-gel derived WO3 coatings. J. Sol-Gel Sci. Technol. 2000, 18, 167–173. [Google Scholar] [CrossRef]
  100. Shim, H.S.; Kim, J.W.; Sung, Y.E.; Kim, W.B. Electrochromic properties of tungsten oxide nanowires fabricated by electrospinning method. Sol. Energy Mater. Sol. Cells 2009, 93, 2062–2068. [Google Scholar] [CrossRef]
  101. Pang, Y.; Chen, Q.; Shen, X.; Tang, L.; Qian, H. Size-controlled Ag nanoparticle modified WO3 composite films for adjustment of electrochromic properties. Thin Solid Films 2010, 518, 1920–1924. [Google Scholar] [CrossRef]
  102. Li, Y.; Long, S.; Liu, Q.; Wang, Q.; Zhang, M.; Lv, H.; Shao, L.; Wang, Y.; Zhang, S.; Zuo, Q.; et al. Nonvolatile multilevel memory effect in Cu/WO3/Pt device structures. Phys. Status Solidi RRL 2010, 4, 124–126. [Google Scholar] [CrossRef]
  103. Chakrabarti, S.; Samanta, S.; Maikap, S.; Rahaman, S.Z.; Cheng, H.M. Temperature-dependent non-linear resistive switching characteristics and mechanism using a new W/WO3/WOx/W structure. Nanoscale Res. Lett. 2016, 11, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Sun, B.; Guo, T.; Zhou, G.; Ranjan, S.; Hou, W.; Hou, Y.; Zhao, Y. Tunneling of photon-generated carrier in the interface barrier induced resistive switching memory behaviour. J. Colloid Interface Sci. 2019, 553, 682–687. [Google Scholar] [CrossRef] [PubMed]
  105. Rajkumari, R.; Singh, N.K. Ag Nanoparticle-decorated WO3 nanowires for nonvolatile Memory. ACS Appl. Nano Mater. 2020, 3, 12087–12094. [Google Scholar] [CrossRef]
  106. Rajkumari, R.; Singh, N.K. Presence of capacitive memory in GLAD-synthesized WO3 nanowire. J. Mater. Sci. Mater. Electron 2021, 32, 3191–3200. [Google Scholar]
Figure 1. (a) Schematic of the device operation. (b) The photocurrent responses with time under the illumination of 365 nm. (c) The time-resolved photocurrent of the photodetector in response to light on/off at an irradiance of 131 mW/cm2 with 365 nm light. (d) The photocurrent–time curve with the change of light intensity. Reproduced or adapted from [31].
Figure 1. (a) Schematic of the device operation. (b) The photocurrent responses with time under the illumination of 365 nm. (c) The time-resolved photocurrent of the photodetector in response to light on/off at an irradiance of 131 mW/cm2 with 365 nm light. (d) The photocurrent–time curve with the change of light intensity. Reproduced or adapted from [31].
Nanomaterials 11 02136 g001
Figure 2. (a) A single WO3 nanodevice symmetrical device both in the dark and under 312 nm UV illumination, and the upper left inset is an SEM image of WO3 nanowire device. (b) I-V characteristics of the nonsymmetrical contact device, upper left inset is an SEM image of WO3 nanowire device and lower right inset shows the schematic structure of the device. (c) Band diagram of the Schottky barrier of the nonsymmetrical device in dark state. (d) Under UV illumination, the change of the barrier height. Reproduced or adapted from [32].
Figure 2. (a) A single WO3 nanodevice symmetrical device both in the dark and under 312 nm UV illumination, and the upper left inset is an SEM image of WO3 nanowire device. (b) I-V characteristics of the nonsymmetrical contact device, upper left inset is an SEM image of WO3 nanowire device and lower right inset shows the schematic structure of the device. (c) Band diagram of the Schottky barrier of the nonsymmetrical device in dark state. (d) Under UV illumination, the change of the barrier height. Reproduced or adapted from [32].
Nanomaterials 11 02136 g002
Figure 3. (a) The schematic illustration of performance measurement of self-crosslinked WO3 nanorods as a UV detector. (b) The photograph of the UV detector. The photocurrent response of UV-C ray (216 nm) irradiation of self-crosslinked WO3 (c) without Ag nanoparticles and (d) with Ag nanoparticles. Reproduced with permission from [33]. Copyright Royal Society of Chemistry.
Figure 3. (a) The schematic illustration of performance measurement of self-crosslinked WO3 nanorods as a UV detector. (b) The photograph of the UV detector. The photocurrent response of UV-C ray (216 nm) irradiation of self-crosslinked WO3 (c) without Ag nanoparticles and (d) with Ag nanoparticles. Reproduced with permission from [33]. Copyright Royal Society of Chemistry.
Nanomaterials 11 02136 g003
Figure 4. (a) Schematic illustration of the WO3/AlGaN/GaN heterostructure photodetector with an integrated micro-heater. (b) Measured spectral response of the WO3/AlGaN/GaN heterostructure photodetector. Reproduced or adapted from [11].
Figure 4. (a) Schematic illustration of the WO3/AlGaN/GaN heterostructure photodetector with an integrated micro-heater. (b) Measured spectral response of the WO3/AlGaN/GaN heterostructure photodetector. Reproduced or adapted from [11].
Nanomaterials 11 02136 g004
Figure 5. (a) I-V characteristics Cu/nanoporous: Sn–WO3/p-Si SBDs fabricated with different concentrations (b) Energy band diagram of the Cu/nanoporous:Sn–WO3/p-Si (MIS) type diode. Reproduced or adapted from [38].
Figure 5. (a) I-V characteristics Cu/nanoporous: Sn–WO3/p-Si SBDs fabricated with different concentrations (b) Energy band diagram of the Cu/nanoporous:Sn–WO3/p-Si (MIS) type diode. Reproduced or adapted from [38].
Nanomaterials 11 02136 g005
Figure 6. (a) Sketch of the WO3 nanowire device structure; (b) SEM image of the WO3 nanowire device. (c) Current-voltage curves and (d) photo-switching behavior of the WO3 nanowires photodetector under the illumination of a 404 nm laser with different laser intensities. (e,f) are the rising time and decay time of the WO3 nanowire photodetector. Reproduced or adapted from [27].
Figure 6. (a) Sketch of the WO3 nanowire device structure; (b) SEM image of the WO3 nanowire device. (c) Current-voltage curves and (d) photo-switching behavior of the WO3 nanowires photodetector under the illumination of a 404 nm laser with different laser intensities. (e,f) are the rising time and decay time of the WO3 nanowire photodetector. Reproduced or adapted from [27].
Nanomaterials 11 02136 g006
Figure 7. (a) The schematic illustration of the WO3−x/Si n-n homotype heterojunction. (b) The I-V curve of a typical WO3−x/Si n-n homotype heterojunction in dark at room temperature, inset shows the semi-logarithmic I-V curve. (c) The time response under various light signals at −5 V bias. (d) The time responses at −5 V bias under pulsed 405 nm light illumination with different light intensities. Reproduced or adapted from [39].
Figure 7. (a) The schematic illustration of the WO3−x/Si n-n homotype heterojunction. (b) The I-V curve of a typical WO3−x/Si n-n homotype heterojunction in dark at room temperature, inset shows the semi-logarithmic I-V curve. (c) The time response under various light signals at −5 V bias. (d) The time responses at −5 V bias under pulsed 405 nm light illumination with different light intensities. Reproduced or adapted from [39].
Nanomaterials 11 02136 g007
Figure 8. EL performance of tandem white OLED (triangles), the control blue device (squares), and the control red device (circles): (a) current–voltage, (b) brightness–current, (c) brightness–voltage, and (d) current efficiency–current. Reproduced or adapted from [41].
Figure 8. EL performance of tandem white OLED (triangles), the control blue device (squares), and the control red device (circles): (a) current–voltage, (b) brightness–current, (c) brightness–voltage, and (d) current efficiency–current. Reproduced or adapted from [41].
Nanomaterials 11 02136 g008
Figure 9. (a) Schematic device structure. (b) Energy level diagram of QLEDs device. (cf) show the device performance of red QLEDs with different concentrations of WO3 (0–16.0 wt.%) doped PEDOT:PSS as the hole injection layer. (c) Current density–voltage characteristics. (d) Current density–luminance characteristics. Inset: the device driven at 5 mA cm−2. (e) Current efficiency–current density–EQE characteristics. (f) Electroluminescence (EL) spectra at 5 mA cm−2. Inset: Commission Internationale de l’Eclairage (CIE) coordinates of a typical red-QLED. Reproduced or adapted from [45].
Figure 9. (a) Schematic device structure. (b) Energy level diagram of QLEDs device. (cf) show the device performance of red QLEDs with different concentrations of WO3 (0–16.0 wt.%) doped PEDOT:PSS as the hole injection layer. (c) Current density–voltage characteristics. (d) Current density–luminance characteristics. Inset: the device driven at 5 mA cm−2. (e) Current efficiency–current density–EQE characteristics. (f) Electroluminescence (EL) spectra at 5 mA cm−2. Inset: Commission Internationale de l’Eclairage (CIE) coordinates of a typical red-QLED. Reproduced or adapted from [45].
Nanomaterials 11 02136 g009
Figure 10. (ae) FESEM images and (f) Photoluminescence emission spectra of the undoped and Cu doped WO3 thin films. Reproduced or adapted from [60].
Figure 10. (ae) FESEM images and (f) Photoluminescence emission spectra of the undoped and Cu doped WO3 thin films. Reproduced or adapted from [60].
Nanomaterials 11 02136 g010
Figure 11. SEM images of (a) undoped (b) 1% (c) 3% (d) 5% and (e) 8% Sn doped WO3 nanostructures. Scale bar is 500 nm. (f) PL spectra of undoped and Sn doped WO3 nanostructures. Reproduced or adapted from [59].
Figure 11. SEM images of (a) undoped (b) 1% (c) 3% (d) 5% and (e) 8% Sn doped WO3 nanostructures. Scale bar is 500 nm. (f) PL spectra of undoped and Sn doped WO3 nanostructures. Reproduced or adapted from [59].
Nanomaterials 11 02136 g011
Figure 12. (a) Schematic diagram of the n-WO3NRs/p-BDD heterojunction device. (b) I-V plots of the n-WO3 NRs/p-BDD heterojunction working at varying temperatures from 20 °C to 290 °C. The top inset is the plot of log (Current) vs. voltage, and the bottom inset is the plot of ln(Is) vs. 1/kBT to obtain the activation energy. Reproduced from [14], with the permission of AIP Publishing.
Figure 12. (a) Schematic diagram of the n-WO3NRs/p-BDD heterojunction device. (b) I-V plots of the n-WO3 NRs/p-BDD heterojunction working at varying temperatures from 20 °C to 290 °C. The top inset is the plot of log (Current) vs. voltage, and the bottom inset is the plot of ln(Is) vs. 1/kBT to obtain the activation energy. Reproduced from [14], with the permission of AIP Publishing.
Nanomaterials 11 02136 g012
Figure 13. (a) FE J–E curve of the large-area patterned defect-rich WO3 NWs film. The emission image (b) corresponding F–N plot and (c) emission current stability of the patterned WO3 NWs film. Reprinted (adapted) with permission from [80]. Copyright {2021} American Chemical Society.
Figure 13. (a) FE J–E curve of the large-area patterned defect-rich WO3 NWs film. The emission image (b) corresponding F–N plot and (c) emission current stability of the patterned WO3 NWs film. Reprinted (adapted) with permission from [80]. Copyright {2021} American Chemical Society.
Nanomaterials 11 02136 g013
Figure 14. (a) Variation of grain size and average film roughness with the working pressure. (b) Cyclic voltammetry for WO3 films deposited at different working pressures. Reproduced or adapted from [83].
Figure 14. (a) Variation of grain size and average film roughness with the working pressure. (b) Cyclic voltammetry for WO3 films deposited at different working pressures. Reproduced or adapted from [83].
Nanomaterials 11 02136 g014
Figure 15. (a) The average conductivity of 0, 5, 10, and 15 wt.% of Sn0.26WO3 thin films (b) I-V characteristics of 15 wt.% of the n-Sn0.26WO3/p-Si diode in darkness and under light illumination. Reproduced or adapted from [84].
Figure 15. (a) The average conductivity of 0, 5, 10, and 15 wt.% of Sn0.26WO3 thin films (b) I-V characteristics of 15 wt.% of the n-Sn0.26WO3/p-Si diode in darkness and under light illumination. Reproduced or adapted from [84].
Nanomaterials 11 02136 g015
Figure 16. (a) The lnI-lnV plot of p-SnS/n-WO3:Sb heterojunction at 300 K. (b) lnI-V plots of p-SnS/n-WO3:Sb heterojunction at various temperatures. Reproduced or adapted from [13].
Figure 16. (a) The lnI-lnV plot of p-SnS/n-WO3:Sb heterojunction at 300 K. (b) lnI-V plots of p-SnS/n-WO3:Sb heterojunction at various temperatures. Reproduced or adapted from [13].
Nanomaterials 11 02136 g016
Figure 17. (a) ln(J)-V characteristics of the WO3/p-Si diodes for different dopant concentrations of Al. (b) Electrical conductivity of the WO3 thin films for different dopant concentrations of Al as a function of temperature. Reproduced or adapted from [86].
Figure 17. (a) ln(J)-V characteristics of the WO3/p-Si diodes for different dopant concentrations of Al. (b) Electrical conductivity of the WO3 thin films for different dopant concentrations of Al as a function of temperature. Reproduced or adapted from [86].
Nanomaterials 11 02136 g017
Figure 18. Diagram of cross-section and electrical behavior of Cu-doped WO3 devices. (a) Diagram of a cross-section of device. The region of the active device is the W-electrolyte-Cu layer superimposed together, and its area is defined by the diameter of the through-hole in the SiO2 dielectric. (b) Data retention behavior of Cu-doped WO3 devices at 10 μA and 1 V. (c) I-V diagram of the device at 1 μA. (d) R-V plot of the device. Reproduced or adapted from [90].
Figure 18. Diagram of cross-section and electrical behavior of Cu-doped WO3 devices. (a) Diagram of a cross-section of device. The region of the active device is the W-electrolyte-Cu layer superimposed together, and its area is defined by the diameter of the through-hole in the SiO2 dielectric. (b) Data retention behavior of Cu-doped WO3 devices at 10 μA and 1 V. (c) I-V diagram of the device at 1 μA. (d) R-V plot of the device. Reproduced or adapted from [90].
Nanomaterials 11 02136 g018
Figure 19. (a) Structure diagram of ITO/WO3/AZO device. (b) The sample was annealed at 600 °C and then irradiated for 30 min, followed by 50 consecutive cycles of volt-ampere characteristic curve in logarithmic form. (c) Energy band structure of ITO/WO3/AZO. (d) Schematic diagram of carrier migration and interfacial tunnel gap formation under light irradiation. Reproduced or adapted from [104].
Figure 19. (a) Structure diagram of ITO/WO3/AZO device. (b) The sample was annealed at 600 °C and then irradiated for 30 min, followed by 50 consecutive cycles of volt-ampere characteristic curve in logarithmic form. (c) Energy band structure of ITO/WO3/AZO. (d) Schematic diagram of carrier migration and interfacial tunnel gap formation under light irradiation. Reproduced or adapted from [104].
Nanomaterials 11 02136 g019
Figure 20. (a) Schematic illustration of the growth processes of Ag-decorated WO3 NWs. (b) Semi-logarithmic scale volt-ampere characteristic curve of Ag-decorated WO3 NWs device at room temperature. The inset shows the image of Ag-decorated WO3 NWs devices. (c) Switching durability of Ag-decorated WO3 NWs at +2 V read voltage. The inset shows the data retention characteristics of Ag-decorated WO3 NWs devices. Reprinted (adapted) with permission from [105]. Copyright {2021} American Chemical Society.
Figure 20. (a) Schematic illustration of the growth processes of Ag-decorated WO3 NWs. (b) Semi-logarithmic scale volt-ampere characteristic curve of Ag-decorated WO3 NWs device at room temperature. The inset shows the image of Ag-decorated WO3 NWs devices. (c) Switching durability of Ag-decorated WO3 NWs at +2 V read voltage. The inset shows the data retention characteristics of Ag-decorated WO3 NWs devices. Reprinted (adapted) with permission from [105]. Copyright {2021} American Chemical Society.
Nanomaterials 11 02136 g020
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yao, Y.; Sang, D.; Zou, L.; Wang, Q.; Liu, C. A Review on the Properties and Applications of WO3 Nanostructure−Based Optical and Electronic Devices. Nanomaterials 2021, 11, 2136. https://doi.org/10.3390/nano11082136

AMA Style

Yao Y, Sang D, Zou L, Wang Q, Liu C. A Review on the Properties and Applications of WO3 Nanostructure−Based Optical and Electronic Devices. Nanomaterials. 2021; 11(8):2136. https://doi.org/10.3390/nano11082136

Chicago/Turabian Style

Yao, Yu, Dandan Sang, Liangrui Zou, Qinglin Wang, and Cailong Liu. 2021. "A Review on the Properties and Applications of WO3 Nanostructure−Based Optical and Electronic Devices" Nanomaterials 11, no. 8: 2136. https://doi.org/10.3390/nano11082136

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