Next Article in Journal
Direct Measurement of Sedimentation Coefficient Distributions in Multimodal Nanoparticle Mixtures
Next Article in Special Issue
Multi-Angular Colorimetric Responses of Uni- and Omni-Directional Femtosecond Laser-Induced Periodic Surface Structures on Metals
Previous Article in Journal
Graphene Nanosheet-Wrapped Mesoporous La0.8Ce0.2Fe0.5Mn0.5O3 Perovskite Oxide Composite for Improved Oxygen Reaction Electro-Kinetics and Li-O2 Battery Application
Previous Article in Special Issue
Dendritic Polymers as Promising Additives for the Manufacturing of Hybrid Organoceramic Nanocomposites with Ameliorated Properties Suitable for an Extensive Diversity of Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal Oxide Based Heterojunctions for Gas Sensors: A Review

1
Hubei Key Laboratory for Processing and Application of Catalytic Materials, School of Physics and Electronic Information, Huanggang Normal University, Huanggang 438000, China
2
Hubei Key Laboratory of Ferro & Piezoelectric Materials and Devices, Faculty of Physics and Electronic Sciences, Hubei University, Wuhan 430062, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(4), 1026; https://doi.org/10.3390/nano11041026
Submission received: 18 March 2021 / Revised: 15 April 2021 / Accepted: 16 April 2021 / Published: 17 April 2021
(This article belongs to the Special Issue Nanotechnologies and Nanomaterials: Selected Papers from CCMR)

Abstract

:
The construction of heterojunctions has been widely applied to improve the gas sensing performance of composites composed of nanostructured metal oxides. This review summarises the recent progress on assembly methods and gas sensing behaviours of sensors based on nanostructured metal oxide heterojunctions. Various methods, including the hydrothermal method, electrospinning and chemical vapour deposition, have been successfully employed to establish metal oxide heterojunctions in the sensing materials. The sensors composed with the built nanostructured heterojunctions were found to show enhanced gas sensing performance with higher sensor responses and shorter response times to the targeted reducing or oxidising gases compare with those of the pure metal oxides. Moreover, the enhanced gas sensing mechanisms of the metal oxide-based heterojunctions to the reducing or oxidising gases are also discussed, with the main emphasis on the important role of the potential barrier on the accumulation layer.

1. Introduction

Gas sensors based on nanostructured metal oxides have attracted significant interest over the last few decades due to their advantages of low cost, ease of fabrication, high sensor response and short response/recovery times [1,2,3,4]. Various metal oxides have been successfully assembled as gas sensors since Seiyama et al. reported their research on the gas sensing performance of the ZnO thin film in the 1960s [5,6]. According to the sensing behaviours of the metal oxides, sensing metal oxides are typically divided into two main groups: n-type metal oxides and p-type metal oxides. Normally, the resistances of the n-type metal oxides decrease (or increase) towards reducing gases such as H2, H2S, CO, CH4, NH3 and other volatile organic compounds (or oxidising gases such as NO2, NO, O3, SO2, etc.), while p-type metal oxides exhibit the opposite behaviour [7,8]. The n-type metal oxides of SnO2, TiO2, WO3, MoO3, Nb2O5, ZnO, etc., and the p-type metal oxides of CuO, Co3O4, Cr2O3, NiO, PdO, etc., have been widely studied for their gas sensing behaviours towards both reducing and oxidizing gases [9,10,11,12].
In recent years, the advancement of new technologies and methods has induced a boom in nanomaterials. Nanostructured metal oxides with various morphologies, such as nanoparticles, nanosheets, nanowires, nanorods, nanoribbons, nanofibres, nanoflowers and nanocages, have been successfully prepared through the routes of hydrothermal processing, thermal oxidation, sol-gel processing, atomic layer deposition, etc. Nanoscale metal oxides have been reported to exhibit promising gas sensing performances, benefiting from their high specific surface areas [13] and active surface states [14]. Specifically, the sensor response of the pure ZnO nanowires was ~15 towards 0.5 ppm NO2 at a working temperature of 225 °C [15]. Other sensors based on SnO2 nanowires [16], TiO2 nanotubes [17], WO3 nanoparticles [18] and In2O3 nanofibres [19] have also been found to respond to the gases of NO2, formaldehyde, H2S and CO, respectively. The gas sensing performances of the gas sensors mentioned above could be further improved to better meet the demands of practical applications via compositing the metal oxide with another (different) metal oxide to form a heterojunction between them. It is noteworthy that when the main phase of a metal oxide was decorated or composited with the second phase of a different semiconductor, the interface between them was known as the structure of a heterojunction in the sensing material [20]. Reports show the specific surface area of the composite is higher than that of the pure metal oxide [21], and the modulation of the potential barrier or accumulation layer in the composite effectively improves the gas sensing behaviour [22,23]. More and more researchers have focused their attention on the studies of the high-performance sensors based on nanostructured metal oxide heterojunctions, as shown in Figure 1. For example, the ordered mesoporous WO3/ZnO nanocomposites synthesised with a hydrothermal method displayed an enhanced sensor response of 168.7 to 1 ppm NO2 at a working temperature of 150 °C, over 10 times higher than that of the pure WO3 [24]. Moreover, the sensor response of CeO2 nanostructures modified with NiO was reported to be ~1570, much higher than that of the pure CeO2 (139). The response time and the recovery time of the composite was 15 s and 19 s, respectively, which is also shorter than that of the pure CeO2 (96 s/118 s) [25]. Therefore, the construction of heterojunctions could be a successful method to improve the gas sensing performances of sensors based on metal oxides.
As reported, there have been different kinds of heterojunctions assembled to improve the gas sensing performances of metal oxides, such as n-n, n-p, p-n or p-p heterojunctions [26,27,28,29]. Note that, in this paper, the type of heterojunction is defined according to the dominant material or the main phase in the composite [30,31]. Accordingly, an n-p heterojunction is formed when the main phase of an n-type metal oxide is modified with a second phase of a p-type semiconductor. Similarly, a p-n heterojunction is established through compositing a p-type metal oxide with an n-type semiconductor. An n-n (or p-p) heterojunction would also be constructed if an n-type (p-type) metal oxide is decorated with a different n-type (p-type) semiconductor in the composite. For example, CuO-decorated ZnO or ZnO-decorated WO3 are the typical n-p or n-n heterojunctions. The ZnO-decorated CuO or NiO-decorated CuO are defined as the p-n or p-p heterojunctions. With the development in the techniques to synthesise nanomaterials, it is facile to construct heterojunctions in composites composed of metal oxides. The metal oxide-based heterojunctions have been successfully established through various combined technologies, such as the thermal oxidation [28], hydrothermal method [32], electrospinning [33,34], chemical vapour deposition (CVD) [35], pulsed laser deposition (PLD) [36], the co-precipitation method [37] and the solvothermal method [38]. For example, the hydrothermal method was reported to effectively prepare the ZnO/SnO2 [39] and the NiO/SnO2 composites [32]. The difference in the Fermi levels of the two metal oxides in the obtained composite would lead to the formation of the potential barrier at their interfaces, an important factor for improving sensing property of the gas sensor based on heterojunctions. The hydrothermal method combined with CVD was also used by Li et al. to establish heterojunctions composed of vertically aligned MoS2/ZnO nanowires [40]. SnO2-CuO heterojunctions were successfully constructed via electrospinning [41]. Their results clearly indicated that the sensor response of the sensor based on the MoS2/ZnO nanowires or the SnO2-CuO heterojunctions was highly improved compared with that of the bare ZnO or SnO2. Though there have been a number of references reviewing the developments in the gas sensing performances of given metal oxides, only a few articles provide a comprehensive review of the effects of the heterojunctions on the enhanced gas sensing performances of composites based on nanostructured metal oxides or separately discuss heterojunctions with n-n, n-p, p-n or p-p structures. Moreover, the enhanced gas sensing mechanism of a given type of heterojunction to a reducing or an oxidising gas should also be studied and summarised. The synthesised methods and the gas sensing performances of the normally studied metal oxides as well as the important roles of the heterojunctions need to be systematically summarised and compared. This will allow us to fully understand the improved gas sensing properties of metal oxide heterojunctions.
In this review, the typical synthetic routes of n-n, n-p, p-n and p-p heterojunctions based on metal oxides are introduced. The gas sensing behaviours of the n-n/n-p heterojunctions (or p-n/p-p heterojunctions) are based on SnO2 and TiO2. ZnO, WO3, MoO3, In2O3, CuO, Cr2O3, NiO and Co3O4, etc., semiconductors are reviewed and compared to show the effects of the heterojunctions on the gas sensing performances of the metal oxides. The enhanced gas sensing mechanisms of the composites towards reducing and oxidising gases are also discussed in detail to systematically understand the role of the built heterojunctions in improving the gas sensing properties of the composites.

2. Nanostructured Metal Oxide Heterojunctions for High-Performance Gas Sensors

As reported, the formation of heterojunctions could be a positive effective strategy to improve the gas sensing performance of the metal oxides. Various methods such as hydrothermal [42], PLD [36], vapour-liquid-solid (VSL) [43], anodic oxidation [44], solvothermal treatment [45], sputtering [46], thermal evaporation [47], electrospinning [33], sol-gel [48] and spin-coating [49] have been successfully applied to assemble the heterostructures in the sensors, and are generally combined to form various heterojunctions (n-n, n-p, p-n or p-p types), as displayed in Figure 2. Other methods to assemble nanostructured metal oxide heterojunctions are listed in Tables S1–S4 (see Supporting Materials), along with the improved gas sensing performances of sensors based on n-p, n-n, p-n and p-p heterojunctions. Some of the typical nanostructured heterojunctions with the n-type (or p-type) metal oxides as the main phases are discussed in the following sections.

2.1. Enhanced Gas Sensing Performances of n-n Junctions or n-p Junctions

Heterojunctions with the n-p or the n-n structure in the sensing materials have been reported to be successful strategies to enhance their gas sensing properties. When the p-type metal oxide (acting as the second phase) is attached to an n-type metal oxide (acting as the main phase), an n-p heterojunction is formed between the two sensing metal oxides. Additionally, n-n heterojunctions can also be assembled in a similar way. One of the common routes to establish the n-n (or n-p) heterojunctions is to prepare the main n-type metal oxides and then decorate the prepared n-type metal oxides with the n-type (or p-type) metal oxides [50]. The modulation of the built potential barrier in the n-n or n-p heterojunction can effectively modify the resistance of the sensing material, and thus greatly improve the gas sensing properties of the sensor composed with the n-n (or n-p) heterojunctions. Meanwhile, it is also noticed that the majority of heterojunctions are assembled as a decorated structure, core-shell structure or mixed structure (one metal oxide mixed with another metal oxide), discussed in the following subsections.

2.1.1. Gas Sensors Based on n-n Junctions

Many references report the improved gas sensing performances of sensors based on n-n heterojunctions. For example, Lu et al. reported the gas sensing properties of the ZnO-decorated SnO2 hollow spheres towards the ethanol synthesised via a two-step hydrothermal method [39]. Hollow spheres of SnO2 of ~100 nm thickness were synthesised via a facile template-free hydrothermal route (see Figure 3a,b) with ZnO nanoparticles of 10–30 nm diameter (see Figure 3c,d) uniformly decorated on its surface via a solution route. The sensor response of the ZnO-decorated SnO2 hollow spheres was calculated to be 34.8 towards 30 ppm ethanol at their optimised operating temperature of 225 °C (see Figure 3e), much higher than that of the bare SnO2 (~5.7 times). Their further research indicated that the composite also exhibited promising selectivity to acetone compared with methanol (Figure 3f). The recovery time of the composite towards 30 ppm ethanol (50 s) was also much shorter than that of acetone (120 s) at the same concentrations as shown in Figure 3g. In addition, the SnO2 compositing with Co3O4 and SiO2 have been reported to be promising gas sensing materials. The SnO2/SiO2 heterojunctions were synthesised via a facile method of a magnetron sputtering process and exhibited promising H2 sensing performance at room temperature [51]. Hybrid Co3O4/SnO2 core-shell nanospheres prepared with a one-step hydrothermal method demonstrated a measured response of 13.6 to 100 ppm NH3 at 200 °C, a value two times higher than that of the solid nanospheres [52]. CeO2-decorated ZnO nanosheets were prepared by a hydrothermal process in combination with the wet impregnation method, exhibited an enhanced sensor response of 90 to 100 ppm ethanol at 310 °C [53]. Additionally, Kim et al. have fabricated ZnO-SnO2 nanofibres through an electrospinning process to effectively detect CO [54].
The α-MoO3/TiO2 core/shell nanorods have been synthesised through a hydrothermal process combined with the following annealing process in air atmosphere [55]. Uniform α-MoO3 nanorods were first prepared and then coated with a shell of TiO2 via a modified wet-chemical method. It was found that the core/shell nanorods exhibited an improved gas sensing performance to 10 ppm ethanol at 180 °C with a short response time of less than 40 s. Meanwhile, the SnO2-core/ZnO-shell nanowires [56] and the Ga2O3-core/ZnO-shell nanorods [57] were successfully synthesised through a plasma-enhanced CVD and atomic layer deposition (ALD), respectively, which also exhibited promising gas sensing performances. The MoO3 nanorods decorated with the ZnO nanoparticles were also reported to be a promising material to detect 100 ppm ethanol with a sensor response of ~30 at the working temperature of 250 °C [58]. Besides the nanocomposites discussed above, it was reported that α-MoO3 compositing with WO3 through a sol-gel method [59] or with Fe2O3 nanoparticles via a hydrothermal method [60] also showed improved gas sensing performances towards O2 or xylene, respectively.
In addition, ZnO nanorods/TiO2 nanoparticles [61] and the ZnO/La0.8Sr0.2Co0.5Ni0.5O3 heterojunction structure [62] were successfully constructed to research their improved gas sensing performances to NO2 and CO, respectively. WO3 compositing with SnO2 was reported to be a potential material to detect acetone [63], while WO3-modified ZnO nanoplates synthesised via the hydrothermal route were assembled for the detection of NH3 [26]. Other effective methods to synthesise n-n heterojunctions and their gas sensing performances are listed in Table S1.

2.1.2. Gas Sensors Based on n-p Junctions

The sensors based on the n-p heterojunctions have been found to show promising gas sensing properties towards various gases. For example, PdO nanoparticles-decorated flower-like ZnO structures (see Figure 4) were prepared by Zhang et al. through a surfactant-free hydrothermal process combined with a further heat treatment [42]. The Zn(AC)2·2H2O was used in the study to synthesise the flower-like ZnO structures, a certain amount of which was dissolved in a solution of NaOH, ethanol and deionized water. The obtained precursor was kept at 150 °C for 24 h. Before decorating with PdO nanoparticles, the flower-like ZnO structures were treated by an annealing process. The annealed ZnO nanoflowers were then dispersed in methanol solvent dissolving PdCl2, and the collected products were calcined at 350 °C for 1 h to obtain the PdO-modified ZnO structures. The decorated flower-like ZnO was reported to show a gas sensor response of 35.4 to 100 ppm ethanol at 320 °C (see Figure 4e), which was much higher than that of the pure ZnO (~10 as shown in Figure 4f). Moreover, the composite presented a shorter recovery time of 7 s than that of the ZnO (14 s). ZnO/Co3O4 composite nanoparticles [64] and Al-doped ZnO/CuO nanocomposites [65] were reported to be sensitive to NO2 and ammonia, respectively.
Gao et al. synthesised CuO nanoparticles-decorated MoO3 nanorods through a hydrothermal process combined with an annealing process [66]. In the first step, MoO3 nanorods were prepared with the raw material (NH4)6Mo7O24·4H2O. Then, the obtained MoO3 nanorods were dispersed in a solution of anhydrous ethanol and copper nitrate under high intensity ultrasonication. The final collected samples were annealed at 550 °C for 2 h. The CuO nanoparticles-decorated MoO3 nanorods showed a higher H2S sensor response of 272 at 270 °C compared with that of pure MoO3, which was mainly attributed to the formation of n-p heterojunctions in the sensing material as reported in their article.
Nano-coaxial Co3O4/TiO2 heterojunctions were successfully assembled through a typical two-step process by Yang et al. [44]. The authors firstly synthesised uniform TiO2 nanotubular arrays via anodic oxidation of a Ti plate which were then decorated with Co3O4 nanoparticles by a hydrothermal process at 120 °C for 5 h. The sensor based on the nano-coaxial Co3O4/TiO2 heterojunctions showed an enhanced sensor response of 40 to 100 ppm ethanol at 260 °C with a short response/recovery time of 1.4 s/7.2 s. The SnO2-Co3O4 composite nanofibres were prepared through electrospinning combined with annealing with the working voltage of 15 kV [67]. The PdO nanoparticle-decorated WO3 nanorods were also reported to be synthesised via a modified precipitation process combined with annealing at 300 °C for 2 h [68]. The SnO2-Co3O4 composite nanofibres and the PdO nanoparticle-decorated WO3 nanorods were found to be sensitive to 10 ppm C6H6 at 350 °C and 3.0 vol% of H2 at 25 °C with enhanced sensor responses of 20 and 80.4, respectively.
CuO/ZnO heterostructural nanorods were prepared by Cao et al. via a combination of hydrothermal and wet-chemical processes [69]. The CuO nanoparticles-decorated ZnO nanorods array showed a sensor response of ~8 to 50 ppm towards triethylamine at a relatively low working temperature of 40 °C (higher than that of the bare ZnO nanorods of ~2.4), and a response time of 5 s (significantly shorter than that of the pure ZnO nanorods of ~11 s). Improved gas sensing performances were also observed in sensing materials composed of CuO-decorated SnO2 nanowires [70], CuO nanoparticles-decorated ZnO flowers [71], flower-like p-CuO/n-ZnO nanorods [72], NiO@ZnO heterostructured nanotubes [73], n-ZnO/p-NiO heterostructured nanofibres [74] and Co3O4 decorated flower-like SnO2 nanorods [75]. The various methods used to assemble the n-p heterojunctions and their gas sensing properties are provided in Table S2.
Besides the nanocomposites discussed above, sensors based on the n-n or n-p heterojunctions have also been assembled to enhance the gas sensing performances of metal oxides. TiO2 composited with ZnO, MoS2, MoO3, V2O5 and WO3 have been designed and successfully established, exhibiting improved gas sensing performances towards ethanol, NO2, alcohol and ammonia [33,55,76,77,78]. For example, the ZnO-decorated TiO2 nanotube layer (prepared by anodic oxidation combined with atomic layer deposition) [76], TiO2/V2O5 branched nanoheterostructures (synthesised by an electrospinning process followed by an annealing treatment) [33] and a TiO2-WO3 composite (obtained via plasma spraying technology using mixed feedstock suspensions) [77] have each exhibited promising gas sensing performances to 1170 ppm ethanol, 100 ppm ethanol and 100 ppm NO2, respectively. α-Fe2O3 composited with SnO2, In2O3 and CdO have also been successfully synthesised through hydrothermal, carbon sphere template and co-precipitating processes, enabling excellent gas sensitivity towards acetone, TMA and CO, respectively [37,79,80]. In2O3 composited with WO3, Fe2O3, TiO2 and SnO2 were also synthesised to assemble the high-performance gas sensors [81,82,83,84]. A series of In2O3-WO3 nanofibres were prepared via an electrostatic spinning technology, which was reported to show an enhanced gas sensing performance to acetone with the n-n semiconductor heterojunctions formed at the interface between WO3 and the In2O3 [81]. The sensor based on mixed Fe2O3-In2O3 nanotubes was also reported to show a high gas sensor response of ~33 towards 100 ppm of formaldehyde at 250 °C [82]. TiO2 nanoparticle-functionalised In2O3 nanowires [83], SnO2/In2O3 composite hetero-nanofibres [84], an octahedral-like ZnO/CuO composite [85] and a nanoporous SnO2@TiO2 heterostructure [86] were reported to show enhanced gas sensing properties towards acetone, formaldehyde and H2S, respectively.
Based on the research discussed above, it is clear that the establishment of n-n or n-p heterojunctions can effectively improve the gas sensing properties of n-type metal oxides. Typically, n-type metal oxides are decorated with zero-dimensional nanoparticles and two-dimensional nanosheets, the concentrations of which have significant effects on the performance of the main n-type phase [67,87,88,89]. More specifically, the gas sensing performance of the main phase in a sensing material improves with increasing concentration of the second phase up to an optimal value, which can be attributed to the increase in the specific surface area of the composite. However, the sensing property of the main phase always degenerates when the content of the second phase is further elevated. The interconnection of the second phase and decrease in the effective surface area was reported to be the two main factors causing a weakened sensor response of the composite. However, most of the reported sensors based on the n-n or n-p heterojunctions always worked at temperatures above 100 °C. We also found there are no clear strategies to indicate which material should be chosen to enhance the gas response of a certain n-type metal oxide, which may be paid more attention in the future by researchers.

2.2. Improved Gas Sensing Properties of p-n or p-p Junctions

Sensors based on the heterojunctions with p-n or p-p structures were also reported to exhibit promising gas sensing properties with high sensor responses and short response/recovery times. Similar to the formation of n-n or n-p heterojunctions, p-n and p-p junctions are also built with p-type metal oxides (as the main phase) decorated or coated with n-type or p-type metal oxides (as the second phase). The different Fermi levels of the metal oxides in the constructed p-n or p-p junctions induce the formation of a thick accumulation layer and thin depletion layer in the sensing composite. The modulation of the thickness of the accumulation layer (acting as the conductive channel of the carriers) significantly influences the conductivity of the sensors, further resulting in improved gas sensing properties of the p-n or p-p junctions.

2.2.1. Gas Sensors Based on p-n Junctions

The p-n heterojunctions in sensing materials have been reported to be effective in improving the gas sensing properties of metal oxides. For example, the SnO-SnO2 composite (p-n heterojunction) was successfully prepared by a facile two-step method with the raw materials of SnCl2·2H2O, NaOH and CTAB at 140 °C for 5 h. Black SnO nanopowders were synthesised via a hydrothermal method at 140 °C for 5 h, and the obtained sample was then treated with an annealing process at a high temperature of 300–500 °C in air atmosphere to obtain the SnO–SnO2 composite. The sensor response of the SnO-SnO2 composite was 2.5 towards 200 ppm NO2 at room temperature, significantly higher than that of pure SnO2 (1.27) or bare SnO (1.1) [90]. The hydrothermal method was also applied to synthesise SnO2-decorated NiO nanostructures (see Figure 5) with the raw sources of NiCl2·6H2O and SnCl4·5H2O at 160 °C for 12 h [91]. It is worth noting that NiO was modified with SnO2 nanoparticles through a one-step process without any catalysts. The SnO2-decorated NiO nanostructure was reported to show enhanced gas sensor responses to 1–200 ppm toluene (see Figure 5e,f). The calculated sensor response of the composite was measured to be 66.2 to 100 ppm toluene at 250 °C, more than 50 times higher than that of the pure NiO nanospheres (1.3). Moreover, the detection limit of this sensor was reported to be as low as 10 ppb toluene with a promising sensor response of 1.2.
Novel TiO2-decorated Co3O4 acicular nanowire arrays were also successfully synthesised by Li et al. with a hydrogen thermal method combined with pulsed laser deposition. The s acicular nanowire arrays modified with TiO2 nanoparticles were found to present a high sensor response of 65 to 100 ppm ethanol at 160 °C, much higher than that of the pure Co3O4 nanowires (~25) [92]. In2O3-decorated CuO nanowires were also prepared through thermal oxidation of Cu meshes followed by the deposition of amorphous indium hydroxide from In(AC)3 solution in ammonia [93]. The decorated CuO nanowires showed a shorter response time of 12 s to CO than that of the pure CuO nanowires (25 s). The novel rod-like α-Fe2O3/NiO heterojunction nanocomposites were synthesised with a one-step hydrothermal method, exhibiting an enhanced sensor response of 290 to 100 ppm acetone at 280 °C with a response time or a recovery time being 28 s or 40 s, respectively [94].
Additionally, CuO composited with TiO2 [95] and SnO2 [35] were constructed to investigate their improved gas-sensing properties. The nanofibres composed of SnO2-CuO heterojunctions have been reported to be successfully synthesised by an electrospinning process and exhibited an improved sensor response of ~95 compared with that of the pure CuO (<10) [35]. Co3O4 composited with In2O3 [96], SiO2 [97] and TiO2 [98] were also successfully prepared via hydrothermal, thermal conversion and facile nanoscale coordination polymer routes, respectively, which showed better gas sensing properties than those of pure Co3O4. The reported sensors based on p-n heterojunctions and their gas sensing performances are listed in Table S3.

2.2.2. Gas Sensors Based on p-p Junctions

The p-p heterojunctions have been found to enhance the gas sensing performance of metal oxides. Li et al. prepared NiO@CuO nanocomposites (a p-p junction) via a facile reflux and hydrothermal process [99]. In their work, the Ni(OH)2 was firstly synthesised with the raw material of nickel nitrate hexahydrate through a hydrothermal method at 140 °C for 5 h. Then, the obtained Ni(OH)2 and the Cu(CH3CO)2·H2O compounds were added in a solution separately with a certain amount of NaOH added during a reflux process to obtain the Ni(OH)2@Cu2(OH)3NO3. The synthesised products were finally treated by a calcination process in air atmosphere at 450 °C for 2 h. The prepared hierarchical flower-like nanostructured NiO-CuO composite exhibited an enhanced gas sensing performance to NO2 at room temperature with a higher gas sensor response compared to pure NiO. The response time of the composite to the 100 ppm NO2 was measured to be as low as 2 s, much shorter than that of the pure NiO. Moreover, the NiO/NiCr2O4 nanocomposite was also found to be more effective at detecting xylene than the pure NiO nanoparticles [100].
Co3O4 hollow nanocages (HNCs) decorated with PdO nanoparticles (see Figure 6a–d) were successfully assembled by the infiltration of metal precursors combined with a subsequent reduction process [101]. The gas sensing performance of the pure Co3O4 hollow nanocages was significantly improved when composited with PdO nanoparticles (PdO-Co3O4 HNCs), with the sensor response measured to be 2.51 towards 5 ppm acetone at 350 °C (see Figure 6e), which was higher than that of the Co3O4 powders (1.96), Co3O4 HNCs (1.45) or PdO-Co3O4 powders (1.98). Moreover, the PdO-Co3O4 HNCs also exhibited outstanding stability to 1 ppm acetone, which is shown in Figure 6f.
Lee et al. prepared TeO2/CuO core-shell nanorods by a combined method of thermal evaporation and sputter deposition [102]. In the reported study, the Te powders were used as the raw material to synthesise TeO2. The TeO2 nanorods were prepared on a substrate of p-type Si (100) by thermal evaporation of Te powders at 400 °C in air in a quartz tube furnace. Then, a thin layer of CuO was directly sputtered on the surface of the obtained TeO2 nanorods through a radio frequency magnetron sputtering process with a target of CuO. The sensor response of TeO2-core/CuO-shell nanorods was found to be 4.25 to 10 ppm NO2 at 150 °C, which was over two times higher than that of the pure TeO2. However, the relatively low sensor response of the TeO2/CuO core-shell nanorods is a drawback that limits their application. Further studies are required to further improve the gas sensing performance of the TeO2/CuO core-shell nanorods.
Meanwhile, p-NiFe2O4 nanoparticle-decorated p-NiO nanosheets were also synthesised with a solvothermal method [103]. The NiO precursor was firstly synthesised after which FeCl3·6H2O was added to prepare NiO nanosheets decorated with NiFe2O4 nanoparticles. The ratio of Fe/Ni was found to have a significant effect on the gas sensing performance of the decorated NiO nanosheets. The composite with the Fe/Ni-24.9 exhibited the optimal sensing performance to 50 ppm acetone at 280 °C, with a high response of ~23.0. The release of captured electrons back to the sensing material breaks the dynamic carrier balance between p-NiO and p-NiFe2O4. This resulted in a reduced potential barrier near the surfaces of the heterojunctions and yielded a large variation in resistance, improving the sensor response of the Fe/Ni-24.9 at%. The in situ formation of a second phase (p-NiFe2O4) on the first phase (p-NiO) was a novel and effective strategy to improve the interaction between the targeted gas and the sensing composite. Similar improvements in CuO-NiO nanotubes with controllable element content of Cu/N developed by a one-pot synthesis was also found, with a sensing capability towards 100 ppm glycol at 110 °C [104]. Based on the studies listed above, the in situ preparation of the second phase required further attention to improve the gas sensing performance of the sensor based on the metal oxide. Other sensors composed with p-p heterojunctions and their gas sensing performances are listed in Table S4.
Other types of heterojunctions based on metal oxides that improve gas sensing performances also exist. Duy et al. assembled n-p-n heterojunctions with the structure of SnO2-carbon nanotube-SnO2 by the method of CVD combined with spray coating process [105]. The obtained n-p-n heterojunctions showed a high response of 17.9 to 100 ppm NO2 at 100 °C. The n-p-n heterostructure of the ZnO-branched SnO2 nanowires decorated with Cr2O3 nanoparticles [106] or the p-n-p heterojunctions of PANI coated CuO-TiO2 nanofibres [107] were also reported to exhibit improved gas sensing performance towards hydrogen and ammonia, respectively. However, only a few references report the study of the sensor based on n-p-n or p-n-p heterojunctions. More research should be conducted to systematically investigate the gas sensing properties of metal oxide heterojunctions comprising the n-p-n or the p-n-p structures.
Based on the discussions above, many kinds of metal oxides heterojunctions have been successfully assembled to enhance the gas sensing performance towards various gases. The sensor response of sensors based on heterojunctions was much higher than that of the pure metal oxides and the response time was improved. The n-n, n-p, p-n or p-p (even the n-p-n or p-n-p) heterojunctions can be chosen to be constructed to assemble gas sensors with outstanding properties. We should point out that the enhanced gas sensing mechanisms of certain heterojunctions towards the oxidising or reducing gases need to be clearly discussed and compared to fully understand the role of the heterojunctions. Therefore, in the next section, we review the mechanisms of the improvements in the gas-sensing properties of the metal oxide heterojunctions.

3. Enhanced Gas Sensing Mechanisms of the Metal Oxide Heterojunctions

Compared with the pure metal oxides, sensors based on metal oxide heterojunctions show improved gas sensing performances towards the targeted gases. When in contact with each other, the transfer of carriers between the two semiconductor materials is induced due to inconsistent Fermi levels at their interfaces. In the n-n or n-p heterojunctions, the Fermi levels of the two metal oxides will move up or down to an equilibrium state, resulting in the bending of their energy bands and the formation of a potential barrier between them [28]. The gas sensing performances of the studied metal oxides are reported to be mainly attributed to the redox reactions of the adsorbed targeted gases on the surfaces of the sensing materials, which has been widely reported by researchers to explain the gas sensing mechanisms of the assembled sensors [12,108]. The variation in the concentration of the carriers induced by the redox reactions on the surfaces of the composites could be of importance to affect the height of the built-in potential barrier. This process further influences the resistance or conductivity of the sensor based on n-n or n-p heterojunctions according to Equation (1):
ΔR ∝ exp{−eΔVb/kBT}
where the ΔR is the change of the resistance of the sensor, ΔVb is the reduction of the height of the potential barrier, kB is the Boltzmann constant and T is the temperature [109]. Therefore, little change in the height of a potential barrier would make the resistance of the investigated sensor vary greatly, leading to an improved gas sensing property of the heterojunction [110]. In the case of the p-n or p-p heterojunctions, the interaction between the targeted gases and the sensing materials would also modify the carriers (especially holes) in the sensors, which would further result in the variation of the thickness of the accumulating layer in the heterojunctions, making a more effective modulation in the width of the conduction channel for the carriers. As a result, sensors based on p-n or p-p heterojunctions also show improved gas sensing properties to the reducing or oxidising gases [108]. Moreover, the composites composed of metal oxide heterojunctions always show higher specific surface areas than the pure metal oxides, which was confirmed by BET measurements of the composites. The higher specific surface area enables gas molecules to diffuse more smoothly to the surface and more easily interact with the composite as well as provide more active sites. The size of the pore volume can be increased with a higher specific surface area, facilitating the diffusion of gas molecules into the sensing material and increasing the active surface in internal parts of the composite for gas molecule adsorption. The absorption and the desorption of the gas molecules can also be accelerated during the response and recovery process of the sensor based on metal oxide heterojunctions. Therefore, the high specific surface area forms another positive factor contributing to the comprehensive improvements in the gas sensing performance of the composite [111,112,113,114,115].
Compared with the effect of the specific surface area, it is more complex to study the enhanced gas sensing mechanisms of the heterojunctions in the sensing materials. The role of the heterojunctions in enhanced gas sensing performances should be analysed in detail to fully understand their direct and significant effects on the enhancement of the gas sensing properties of the sensors based on the composites. In the following section, the gas sensing mechanisms of the metal oxides to the common reducing and oxidising gases are discussed, and the effects of various commonly studied heterojunctions on the improved gas sensing properties of the composites are systematically investigated. In order to make the discussions clear, H2 (a typical reducing gas) and NO2 (a typical oxidising gas) were selected for the discussion of the enhanced gas sensing mechanisms of the metal oxides due to their immense studies in the area of gas sensors.

3.1. Enhanced Gas Sensing Mechanisms to Reducing Gases

Gas sensors based on n-n or n-p heterojunctions always exhibit typical n-type sensing performances at relatively low working temperatures towards reducing gases such as H2, H2S, CO, NH3 and ethanol. The widely studied ZnO-based material is taken as an example to more clearly illustrate the sensing mechanism of the n-type metal oxide to a reducing gas. The resistance of ZnO-based sensors has been reported to decrease quickly when H2 (ethanol or H2S) is introduced onto their surface [116,117,118]. In air, oxygen molecules would spontaneously be adsorbed on the active sites of the surface of the ZnO to form chemisorbed oxygen molecules according to Equation (2). Then, the chemisorbed oxygen molecules can capture electrons from the conductive bands of the ZnO to become the oxygen species (O2: <150 °C, O: 150 °C~400 °C and O2−: >400 °C) based on Equation (3), which builds a depletion layer in the ZnO surface and a high resistance in air. When H2 gas is introduced, the H2 molecules will interact with the pre-adsorbed oxygen species to form H2O based on Equation (4), releasing electrons back to ZnO. This response process increases the concentration of electrons and decreases the thickness of the depletion layer in ZnO, leading to a decrease in the resistance of the sensor based on ZnO-based materials.
O2(g) + e ↔ O2(ad)
O2(ad) + e ↔ O2(ad)
2H2(g) + O2 (ad) = 2H2O(g) + e
In contrast, composites made of p-n or p-p heterojunctions show typical p-type sensing performances towards reducing gases. As reported, the resistance of CuO nanowires increased when used to detect hydrogen gas (at working temperatures between 150 °C and 400 °C) [119]. When the CuO nanowires were placed in an air atmosphere, the oxygen molecule could also adsorb on the active sites in the surface of CuO to form adsorbed oxygen species (O) based on Equation (5), releasing holes to CuO and thus increasing the concentration of holes. This forms an accumulation layer in the sensing material, which acts as the conduction channel for carriers in CuO. In an H2 atmosphere, hydrogen molecules interact with the adsorbed oxygen molecules according to Equation (6), reducing the concentration of carriers and the thickness of the accumulation layer and induces the formation of a depletion layer on the surface of CuO. Therefore, the resistance of a sensor based on CuO nanowires increases in reducing gas environments [108,119].
O2(g) ↔ O2(ad) + h+ ↔ 2O(ad) + 2h+
2H2(g) + 2O(ad) + 2h+ = 2H2O(g)
The synthesis of TiO2 nanotubes decorated with SnO2 nanoparticles and their H2 sensing performance has been reported [89] and is selected to analyse the important role of the typical n-n heterojunction in improving the sensing performance of the sensor towards reducing gases. The results showed that the H2 sensing property of the TiO2-based composite was highly improved with the help of the heterojunction between TiO2 and SnO2. It was reported that the Fermi level of TiO2 was higher than that of SnO2, resulting in the electron transfer to SnO2 from TiO2 until achieving the equilibrium states of their Fermi levels. This would make a thick depletion layer formed at the interface between TiO2 and SnO2 and induce a high potential barrier built in air due to the adsorption of oxygen molecules. The potential barrier always acts as the obstacle to the transportation of electrons in the sensing materials, resulting in the high resistance of the composites. The accumulation of electrons in the SnO2 side would induce more oxygen molecules adsorbed onto the surface of the composite. When hydrogen gas is introduced, the hydrogen gas interacts with the adsorbed oxygen species on the surfaces of TiO2 and SnO2 immediately and releases electrons back to the sensing materials. The released electrons would decrease the thickness of the depletion layers between TiO2 and SnO2, further resulting in the decrease in the height of the potential barrier. This process would increase the conductivity of the sensor and significantly enhance the H2 sensing performance of the composite. The porous MoO3/SnO2 nanoflakes with n-n junctions was also reported to show an improved gas sensing property with a higher gas sensor response being 43.5 towards 10 ppm H2S at 115 °C compared with that of the pure SnO2, which could also be attributed to the reasons mentioned previously (see Figure 7a1,a2) [120]. Moreover, the improvement in the H2S or xylene sensing performance of TiO2-decorated α-Fe2O3 nanorods [121] or Fe2O3 nanoparticles-decorated MoO3 nanobelts [122] could also be explained by the enhanced gas sensing mechanism above.
In the case of sensors based on n-p heterojunctions, the NiO-decorated Nb2O5 nanocomposites have been reported to exhibit a significant improvement in the H2 gas sensing performance compared with that of the pure Nb2O5 nanoparticles [123]. When the NiO nanoparticles are loaded onto the surface of the Nb2O5 nanoparticles, the electrons diffuse to the Nb2O5 and the holes move toward the NiO, causing the Fermi levels of the two different metal oxides to reach an equilibrium state. In air, the adsorption of the oxygen molecules on the surfaces of the NiO and the Nb2O5 also results in the formation of an accumulation layer of holes in the NiO side and a depletion layer in the Nb2O5 side. This causes the energy bands of NiO to bend upwards, increasing the potential barrier at the interfaces in the region of heterojunctions. When the NiO-decorated Nb2O5 nanocomposites is exposed to H2, the interaction with H2 and adsorbed oxygen species releases electrons to Nb2O5 but captures the holes in the NiO. This process induces the formation of a depletion layer between NiO and Nb2O5 and makes the energy bands of NiO bend downwards, dramatically decreasing the height of the potential barrier at the heterojunction (see Figure 7b1,b2). The NiO nanoparticles have also been reported to be an excellent catalyst to effectively oxidise H2, causing reactions between the adsorbed H2 and the adsorbed oxygen species to occur more sufficiently and smoothly. As a result, the NiO-decorated Nb2O5 nanocomposites exhibit an improved gas sensing property to H2. The enhanced gas sensing performances of the Co3O4-decorated WO3 nanowires [124], SnO2-Co3O4 composite nanofibres [67], CuO-loaded In2O3 nanofibres [125], hierarchical SnO/SnO2 nanocomposites [126], ZnO nanowire arrays/CuO nanospheres heterostructures [127] and p-NiS/n-In2O3 heterojunction nanocomposites [34] towards reducing gases can also be attributed to the reasons listed above.
For the p-n junction, Lee et al. reported the sensor based on Nb2O5 nanoparticles-decorated CuO nanorods to be more sensitive towards hydrogen molecules than the pure CuO nanorods [119]. The higher Fermi level of Nb2O5 makes the electrons diffuse to the CuO and the holes transfer in an opposite orientation, leading to the bending of energy bands. In air, the adsorption of oxygen molecules captures the electrons from Nb2O5 but releases the holes to the CuO, resulting in the formation of a thick depletion layer in Nb2O5 and a thick accumulation layer in the CuO. This leads to the high potential barrier in the composite in air. As reported, the accumulation layer in the CuO can act as a conduction channel for carriers in the sensing material. When exposed to H2, the hydrogen molecule can interact with the adsorbed oxygen species on the Nb2O5 and the CuO, releasing the electrons back to the Nb2O5 but capturing the holes in the CuO. Effectively, this decreases the thickness of the depletion layer in Nb2O5 and significantly thins the accumulation layer in CuO with the possible formation of a depletion layer in the CuO, attributed to more oxygen molecules adsorbed on the surface of the CuO due to the formation of heterojunctions. This dramatic decrease in the thickness of the accumulation layer greatly narrows the conduction channel width for carriers, as shown Figure 7(c1,c2). As a result, the Nb2O5 nanoparticle-decorated CuO nanorods exhibited an improved p-type sensing performance to H2. The improved sensing mechanism could also reasonably explain the enhancements in the gas sensing performances of the In2O3-decorated CuO nanowires [93], SnO2-decorated NiO nanostructure [91], hierarchical α-Fe2O3/NiO composites [128], SnO2-decorated NiO foam [129] and CuxO-modified ZnO composite [130] with a hollow structure towards H2, the toluene and the acetone.
Similarly, in the case of the p-p heterojunction, p-NiO-decorated p-CuO microspheres were prepared through a hydrothermal process and studied the enhanced gas sensing performance of the obtained composite with p-p heterojunctions [131]. The Fermi level of the CuO was higher than that of the NiO, resulting in the transfer of holes from NiO to CuO and the diffusion of electrons to NiO from CuO. As such, accumulation and depletion layers of holes build on the CuO and the NiO sides, respectively. In air atmosphere, the adsorption of the oxygen molecules on the surfaces of CuO and NiO releases holes to the sensing material as previously mentioned. As a result, the thickness of the depletion layer of the holes in the NiO decreases, but the thickness of the accumulation layer of the holes in the CuO increases. The width of the conduction channel increased in the heterojunctions between CuO and NiO, causing a low resistance of the composite. When a reducing gas was introduced on the surface of the composite, the adsorbed oxygen ions reacted with the introduced gas molecules, capturing the holes from CuO and NiO. This process decreases the concentration of the carriers in the sensing materials, further increasing the thickness of the depletion layer of the holes in NiO and decreasing the thickness of the accumulation layer of holes in CuO. Therefore, the width of the conduction channel increased in the heterojunctions was significantly narrowed, causing an increase in the resistance of the composites. Therefore, the gas sensor response of the p-NiO-decorated p-CuO microspheres was highly improved towards reducing gases. The enhanced gas sensing mechanisms of p-NiO/p-NiCr2O4 nanocomposites [132] or the Cr2O3-Co3O4 nanofibres [133] to xylene or C2H5OH can also be explained by the above discussions.
Apart from H2, there are also a number of references reporting similar improved sensing mechanisms of sensors based on metal oxide heterojunctions towards NH3, another widely investigated reducing gas. The work conducted by Shi et al. showed that the sensor response of WO3@CoWO4 (n-n type) heterojunction nanofibres was over 10 times higher than that of the bare WO3 at room temperature [134]. The authors pointed out the formation of a number of heterojunctions for WO3 composited with CoWO4 in the sensing material. The differences in the Fermi levels of WO3 and CoWO4 cause band bending and trigger the transfer of electrons and holes between them until an equilibrium in final Fermi levels is reached. In air, oxygen gas can be adsorbed on the surface of the two different sensing materials and capture electrons from their conductive bands to form chemisorbed oxygen ions (O2 at room temperature) according to Equation (3). This process results in a wider depletion layer and constructs a higher potential barrier near the surface of the heterojunction in the composite than those in the pure CoWO4. NH3 molecules could interact with the O2 according to the following equation: 4NH3(ad) + 5O2 (ad) → 4NO(g) + 6H2O(g) + 5e. Electrons were then released back to the sensing materials of the WO3@CoWO4 composite, reducing the thickness of the depletion layer and decreasing the height of the potential barrier at the heterojunction. As a result, the sensing performance of WO3@CoWO4 heterojunction nanofibres towards NH3 could be significantly enhanced at room temperature. Additionally, the specific surface area of the composite was higher, allowing more electrons to be transferred from the shallow donor levels of the WO3 nanoparticles to CoWO4 nanoparticles in the composite, thus enabling the enhanced NH3 sensing property of the sensor based on the metal oxide heterojunction. The study of Gong et al. also revealed that the enhanced NH3 sensing performance of the flower-like n-ZnO decorated with p-NiO with hierarchical structure was mainly attributed to the formation of the depletion layer and the modulation of the potential barrier height at the surface of the heterojunction [135]. A similar improved sensing mechanism was also reported in the enhanced NH3 sensing performance of the sensors based on other heterojunctions, including polyaniline/SrGe4O9 nanocomposite [136], polyaniline nanograin enchased TiO2 fibres [137], SnO2@polyaniline nanocomposites [138], V2O5/CuWO4 heterojunctions [139], Fe2O3-ZnO nanocomposites [49], rGO/WO3 nanowire nanocomposites [140], WO3@SnO2 core-shell nanosheets [141], PANI-CeO2 nanocomposite thin films [142], CuPc-loaded ZnO nanorods [143], Co3O4 nanorod-decorated MoS2 nanosheets [144], SnO2/NiO composite nanowebs [145], bilayer SnO2-WO3 nanofilms [146], Cu2O nanoparticles decorated with p-type MoS2 nanosheets [147], TiO2 and NiO nanostructured bilayer thin films [148] and mesoporous In2O3@CuO multijunction nanofibres [149].

3.2. Improved Gas Sensing Mechanism towards Oxidising Gases

In contrast to the sensing behaviour of n-n or n-p heterojunctions towards reducing gases, sensors based on n-n or the n-p heterojunctions were reported to exhibit a typical p-type sensing performance towards the oxidising gases. Many researchers have studied the oxidising gas (such as NO2) sensing performance of heterojunctions based on n-type metal oxides at working temperatures within the range of 150 °C to 400 °C. The ZnO nanoparticles exhibited a typical p-type sensing performance towards 0.3–10 ppm NO2 at the working temperature of 250 °C, with the resistance of the sensor increasing quickly when exposed to an NO2 gas atmosphere [150]. In air, oxygen molecules can adsorb onto the active sites of the surface of the nanostructured ZnO according to Equation (7), capturing electrons from the conductive bands of ZnO. This process causes a decrease in carriers in ZnO and the formation of a depletion layer on the surface of ZnO. When exposed to an NO2 atmosphere, the NO2 molecules can interact with the ZnO directly and with adsorbed O on the ZnO according to Equations (8) and (9), respectively. Generally, NO2 can be adsorbed onto the active sites of the ZnO surface based on Equation (8), capturing the electrons from the ZnO to form NO2. The NO2 can then further react with adsorbed oxygen species following Equation (9), grabbing electrons from ZnO. These sensing processes decrease the concentration of the carriers in the ZnO and increase the thickness of the depletion layer in the surface of the ZnO, resulting in a significant increase in the resistance of ZnO and the p-type sensing performance to NO2 gas.
1/2O2(g) + e ↔ O(ad)
NO2(g) + e ↔ NO2(ad)
NO2(ad) + O(ad) + 2e ↔ NO(g) + 2O2
In addition, sensors based on p-n or p-p heterojunctions have been found to show typical n-type sensing performances towards oxidising gases. The resistances of the sensors assembled by heterojunctions based on p-type metal oxides decrease rapidly when exposed to oxidising gases at the working temperatures of approximately 200 °C. Taking the sensor based on Co3O4 as an example, oxygen molecules can be adsorbed onto the active sites on the surface of Co3O4 according to Equation (5), releasing holes to Co3O4 and resulting in the formation of chemisorbed oxygen species (mainly O). This process can also induce the establishment of an accumulation layer in the surface of Co3O4. In an NO2 atmosphere, NO2 molecules have also been reported to adsorb onto the active sites of a Co3O4 surface based on Equation (10), releasing holes to the sensing materials and forming NO2. The adsorbed NO2 can then interact with the adsorbed oxygen species according to Equation (11), releasing more holes to Co3O4. These processes make the accumulating layer thicker on the surface of Co3O4 and cause the resistance of the sensor to decrease, leading to an n-type sensing performance of the Co3O4-based sensor towards NO2 [64].
NO2(g) ↔ NO2(ad) + h+
NO2(ad) + O(ad) ↔ NO(g) + 2O2(ad) + 2h+
For the sensor based on an n-n heterojunction, the study of the gas sensing properties of ZnO-SnO2 hollow nanofibres showed that the composites exhibit a much higher sensor response towards NO2 than pure SnO2 [151]. In the composite, the Fermi level of the SnO2 is higher than that of the ZnO. The lower Fermi level of the ZnO can thus lead to the transfer of the holes from ZnO to SnO2 and the diffusion of electrons to ZnO from SnO2 until their Fermi levels reach an equilibrium state. This process can then form a thick accumulation layer on the ZnO side and a thin depletion layer on the SnO2 side. In air atmosphere, oxygen molecules can adsorb onto the surface of SnO2, which would capture electrons from SnO2 and increase the thickness of the built-in depletion layer. The accumulation layer of the electrons in ZnO can cause more molecules to absorb onto its surface in air, capturing electrons and significantly decreasing the thickness of the established accumulation layer and even lead to the formation of a thin depletion layer (see Figure 8(a1)). In an NO2 atmosphere, adsorbed NO2 molecules on the surfaces of the metal oxides and the reaction between NO2 and adsorbed oxygen molecules further capture electrons in ZnO and SnO2, significantly increase in the depletion layer at the interfaces between ZnO and SnO2 (see Figure 8(a2)). As a result, the height of the potential barrier increases greatly, making the ZnO-SnO2 hollow nanofibres exhibit an improved NO2 gas sensing property. The sensors based on ZnO nanorods decorated with TiO2 nanoparticles [61], Bi2O3-branched SnO2 nanowires [112], In2O3-composited SnO2 nanorods [152] and SnO2-core/ZnO-shell nanowires [153] also exhibited improved NO2 gas sensing performances according to the mechanism discussed above.
In the case of the n-p junction, Co3O4-decorated ZnO nanoparticles have been established by Lee et al. and showed a significant enhancement in the NO2 gas sensing performance [64]. The Fermi level of ZnO is higher than that of Co3O4, inducing the transfer of carriers between them and the formation of a depletion layer at the heterojunction. In air, the adsorption of the oxygen molecules on the surfaces of ZnO and Co3O4 capture the electrons from ZnO and release holes to Co3O4. This leads to the building of a depletion layer on the ZnO side and an accumulation layer on the Co3O4 side as well as a significant bending in their energy bands (see Figure 8(b1)). As a result, a potential barrier is formed at the interfaces between ZnO and Co3O4, resulting in a higher resistance of the sensor based on the composites than that of pure ZnO. In an NO2 atmosphere, the adsorption of NO2 molecules on the surfaces of ZnO and Co3O4 can lead to the capture of electrons from ZnO but the release of the holes to Co3O4. The variation in the carriers in ZnO and Co3O4 causes both the depletion layer in ZnO and the accumulation layer in Co3O4 to become thicker. This sensing process increases the height of the potential barrier in the heterojunctions and significantly increases the resistance of the Co3O4-decorated ZnO composite (see Figure 8(b1,b2)). Meanwhile, the catalytic property of Co3O4 to NO2 also acts as a positive factor for the improved NO2 gas sensing performance of the composite. Oxygen molecules are reported to be more easily adsorbed onto the surface of p-type metal oxides, which is another reason for the improved NO2 gas sensing performance of the Co3O4-decorated ZnO nanoparticles. The improvements in the NO2 gas sensing properties of the SnO-SnO2 nanocomposites [90], CuO-decorated ZnO nanowires [155], TeO2/SnO2 brush- nanowires [156] and ultra-long ZnO@Bi2O3 heterojunction nanorods [157] can also be attributed to the reasons listed above.
Sensors based on the p-n heterojunctions have also been reported to be effective at detecting oxidising gases. For example, NiO-SnO2 nanocomposites (p-n junctions) were found to exhibit an improved gas sensing performance towards NO2 compared with pure NiO [154]. The Fermi level of the n-type SnO2 was found to be higher than that of the p-type NiO. The electrons would be transferred from SnO2 to the NiO, and the holes would diffuse from the NiO to the SnO2. In an air atmosphere, the adsorption of oxygen molecules would capture electrons in the SnO2 and release the holes to NiO, resulting the formation of a thin depletion layer in SnO2 and an accumulation layer in NiO. A potential barrier is then established between NiO and SnO2, and the carriers in the sensing materials are mainly transported through the accumulation layer. In an NO2 atmosphere, the adsorption of NO2 molecules on the surface of NiO results in more holes being released to NiO, further increasing the thickness of the accumulation layer in the NiO layer. The adsorption of NO2 molecules on the surface of SnO2 would allow more electrons to be grabbed from the SnO2, further increasing the thickness of the depletion layer in SnO2. Moreover, NO2 can be adsorbed on SnO2 more easily due to its higher electron concentration. There would be more NO2 molecules adsorbed on the NiO-SnO2 nanocomposites, further resulting in the great modulation in the accumulation layer of the heterojunction nanocomposites. The increase in the thickness of the accumulation layer in NiO would widen the conduction channel for the carriers, which would result in a significant decrease in the resistance of the sensing material (see Figure 8(c1,c2)). Therefore, the sensors based on NiO-SnO2 nanocomposites exhibit a higher sensor response to NO2 than that of bare NiO.
For the p-p heterojunction, the NO2 gas sensing performance of sensors based on p-type NiO nanosheets could be successfully improved through modifying them with the p-type CuO nanoparticles [158]. In the CuO-decorated NiO nanosheets, the differences in Fermi levels of CuO and NiO lead to the transfer of carriers between the two, resulting in the formation of a hole depletion layer and a hole accumulation layer between their interfaces. In air, the adsorption of oxygen molecules on the surfaces of CuO and NiO can release holes to the sensing materials, leading to the increase in the thickness of the accumulation layer at the interfaces between CuO and NiO. When NO2 is introduced and interacts with the sensing material, more holes are released to CuO and NiO. Moreover, more NO2 molecules become adsorbed on the sensing material due to the accumulation of holes in the composite and its higher specific surface area. This sensing process would more effectively increase the carriers (holes) in the composites and widen the accumulation layer between CuO and NiO. As a result, the resistance of the p-p heterojunctions significantly decreased and the CuO-decorated NiO nanosheets presented an enhanced NO2 gas sensing performance. The enhanced NO2 sensing mechanism discussed above can also be applied to explain the improved NO2 sensing properties of the sensors based on CuO-decorated TeO2 nanorods [102] and vertically aligned Cu3Mo2O9 micro/nanorods on a CuO layer (Cu3Mo2O9@CuO nanorods) [159].
The discussions of the enhanced gas sensing mechanisms of the n-n, n-p, p-n and p-p heterojunctions reveal that modulations of the height of the potential barriers and the thickness of the accumulation layer in the heterojunctions are responsible for the improvements of the gas sensing performances of the nanocomposites. The different Fermi levels of the metal oxides induce band bending in the heterojunctions, leading to the formation of potential barriers and accumulation layers in n-type and p-type heterojunctions, respectively. The interactions between the targeted gases and the sensing materials cause variations in the height of the potential barriers in n-type heterojunctions (n-n or n-p heterojunctions) and the thickness of the accumulation layer in p-type heterojunctions (p-n or p-p heterojunctions), inducing enhancements of the sensing performance of the composites.

4. Conclusions

The gas sensing performances of metal oxides have been successfully improved through assembling heterojunctions in sensing materials. The heterojunctions are usually constructed via combined methods of electrospinning, thermal oxidation, ALD, PLD, hydrothermal process and CVD. The sensor response, response time or recovery time based on n-n, n-p, p-n or p-p heterojunctions can be effectively enhanced. Modulations in the built-in heterojunctions are mainly responsible for the enhanced gas sensing performances of the sensors based on n-n or n-p junctions. The improvement in the gas sensing behaviours of the sensors based on p-n or p-p heterojunctions can be attributed to variations in the thicknesses of the accumulation layers in the junctions. n-type or p-type nanostructured metal oxides with different morphologies can be selected to assemble heterojunctions and their concentrations can modified, indicating that more interesting gas sensors based on nanostructured metal oxide heterojunctions might be explored.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11041026/s1, Table S1: The assembled strategies of n-n heterojunctions and their gas sensing performances. Table S2: The assembled strategies of n-p heterojunctions and their gas sensing performances. Table S3: The assembled strategies of p-n heterojunctions and their gas sensing performances. Table S4: The assembled strategies of p-p heterojunctions and their gas sensing performances.

Author Contributions

Writing—original draft preparation and editing, formal analysis and investigation, S.Y. and Z.W.; investigation and resources, G.L., Z.W. and Z.L.; writing—review, supervision and project administration, H.X. and H.G. All authors helped in drafting the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51802109 and 51972102), the Key Project of the Education Department of Hubei Province (Grant No. D20202903) and the ChuTian Scholars Program of Hubei Province.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, T.; Zeng, W.; Wang, Z. Quasi-one-dimensional metal-oxide-based heterostructural gas-sensing materials: A review. Sens. Actuators B Chem. 2015, 221, 1570–1585. [Google Scholar] [CrossRef]
  2. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  3. Zhang, J.; Qin, Z.; Zeng, D.; Xie, C. Metal-oxide-semiconductor based gas sensors: Screening, preparation, and integration. Phys. Chem. Chem. Phys. 2017, 19, 6313–6329. [Google Scholar] [CrossRef] [PubMed]
  4. Sahay, P.P. Multifunctional metal oxide nanomaterials for chemical gas sensing. Procedia Eng. 2017, 215, 145–151. [Google Scholar] [CrossRef]
  5. Seiyama, T.; Kato, A.; Fujiishi, K.; Nagatani, M. A new detector for gaseous components using semiconductive thin films. Anal. Chem. 1962, 34, 1502–1503. [Google Scholar] [CrossRef]
  6. Hoa, N.D.; Duy, N.V.; El-Safty, S.A.; Hieu, N.V. Meso-/Nanoporous semiconducting metal oxides for gas sensor applications. J. Nanomater. 2015, 16, 186. [Google Scholar] [CrossRef] [Green Version]
  7. Vuong, N.M.; Kim, D.; Kim, H. Surface gas sensing kinetics of a WO3 nanowire sensor: Part 2—Reducing gases. Sens. Actuators B Chem. 2016, 224, 425–433. [Google Scholar] [CrossRef]
  8. Vuong, N.M.; Kim, D.; Kim, H. Surface gas sensing kinetics of a WO3 nanowire sensor: Part 1—Oxidizing gases. Sens. Actuators B Chem. 2015, 220, 932–941. [Google Scholar] [CrossRef]
  9. Cao, P.-J.; Li, M.; Rao, C.N.; Han, S.; Xu, W.-Y.; Fang, M.; Liu, X.-K.; Zeng, Y.-X.; Liu, W.-J.; Zhu, D.-L.; et al. High sensitivity NO2 gas sensor based on 3D WO3 microflowers assembled by numerous nanoplates. J. Nanosci. Nanotechnol. 2020, 20, 1790–1798. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal oxide gas sensors: Sensitivity and influencing factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Xu, K.; Fu, C.; Gao, Z.; Wei, F.; Ying, Y.; Xu, C.; Fu, G. Nanomaterial-based gas sensors: A review. Instrum. Sci. Technol. 2018, 46, 115–145. [Google Scholar] [CrossRef]
  12. Wang, Z.; Hu, Y.; Wang, W.; Zhang, X.; Wang, B.; Tian, H.; Wang, Y.; Guan, J.; Gu, H. Fast and highly-sensitive hydrogen sensing of Nb2O5 nanowires at room temperature. Int. J. Hydrog. Energy 2012, 37, 4526–4532. [Google Scholar] [CrossRef]
  13. Yang, S.; Wang, Z.; Hu, Y.; Luo, X.; Lei, J.; Zhou, D.; Fei, L.; Wang, Y.; Gu, H. Highly responsive room-temperature hydrogen sensing of α-MoO3 nanoribbon membranes. ACS Appl. Mater. Interfaces 2015, 7, 9247–9253. [Google Scholar] [CrossRef]
  14. Lin, Y.; Wang, Y.; Wei, W.; Zhu, L.; Wen, S.; Ruan, S. Synergistically improved formaldehyde gas sensing properties of SnO2 microspheres by indium and palladium co-doping. Ceram. Int. 2015, 41, 7329–7336. [Google Scholar] [CrossRef]
  15. Ahn, M.-W.; Park, K.-S.; Heo, J.-H.; Park, J.-G.; Kim, D.-W.; Choi, K.J.; Lee, J.-H.; Hong, S.-H. Gas sensing properties of defect-controlled ZnO-nanowire gas sensor. Appl. Phys. Lett. 2008, 93, 263103. [Google Scholar] [CrossRef]
  16. Choi, Y.J.; Hwang, I.S.; Park, J.G.; Choi, K.J.; Park, J.H.; Lee, J.H. Novel fabrication of an SnO2 nanowire gas sensor with high sensitivity. Nanotechnology 2008, 19, 095508. [Google Scholar] [CrossRef] [PubMed]
  17. Lin, S.; Li, D.; Wu, J.; Li, X.; Akbar, S.A. A selective room temperature formaldehyde gas sensor using TiO2 nanotube arrays. Sens. Actuators B Chem. 2011, 156, 505–509. [Google Scholar] [CrossRef]
  18. Ionescu, R.; Hoel, A.; Granqvist, C.G.; Llobet, E.; Heszler, P. Low-level detection of ethanol and H2S with temperature-modulated WO3 nanoparticle gas sensors. Sens. Actuators B Chem. 2005, 104, 132–139. [Google Scholar] [CrossRef]
  19. Lim, S.K.; Hwang, S.-H.; Chang, D.; Kim, S. Preparation of mesoporous In2O3 nanofibers by electrospinning and their application as a CO gas sensor. Sens. Actuators B Chem. 2010, 149, 28–33. [Google Scholar] [CrossRef]
  20. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B Chem. 2014, 204, 250–272. [Google Scholar] [CrossRef]
  21. Bao, H.-F.; Yue, T.-T.; Zhang, X.-X.; Dong, Z.; Yan, Y.; Feng, W. Enhanced ethanol-sensing properties based on modified NiO-ZnO p-n heterojunction nanotubes. J. Nanosci. Nanotechnol. 2020, 20, 731–740. [Google Scholar] [CrossRef]
  22. Nakate, U.T.; Ahmad, R.; Patil, P.; Wang, Y.; Bhat, K.S.; Mahmoudi, T.; Yu, Y.T.; Suh, E.-K.; Hahn, Y.-B. Improved selectivity and low concentration hydrogen gas sensor application of Pd sensitized heterojunction n-ZnO/p-NiO nanostructures. J. Alloys Compd. 2019, 797, 456–464. [Google Scholar] [CrossRef]
  23. Sun, Q.; Wang, J.; Hao, J.; Zheng, S.; Wan, P.; Wang, T.; Fang, H.; Wang, Y. SnS2/SnS p–n heterojunctions with an accumulation layer for ultrasensitive room-temperature NO2 detection. Nanoscale 2019, 11, 13741–13749. [Google Scholar] [CrossRef]
  24. Sun, J.; Sun, L.; Han, N.; Pan, J.; Liu, W.; Bai, S.; Feng, Y.; Luo, R.; Li, D.; Chen, A. Ordered mesoporous WO3/ZnO nanocomposites with isotype heterojunctions for sensitive detection of NO2. Sens. Actuators B Chem. 2019, 285, 68–75. [Google Scholar] [CrossRef]
  25. Jayababu, N.; Poloju, M.; Shruthi, J.; Reddy, M.V.R. NiO decorated CeO2 nanostructures as room temperature isopropanol gas sensors. RSC Adv. 2019, 9, 13765–13775. [Google Scholar] [CrossRef] [Green Version]
  26. Nguyen, D.D.; Do, D.T.; Vu, X.H.; Dang, D.V.; Nguyen, D.C. ZnO nanoplates surfaced-decorated by WO3 nanorods for NH3 gas sensing application. Adv. Nat. Sci. Nanosci. Nanotechnol. 2016, 7, 015004. [Google Scholar] [CrossRef]
  27. Zhao, X.; Ji, H.; Jia, Q.; Wang, M. A nanoscale Co3O4–WO3 p-n junction sensor with enhanced acetone responsivity. J. Mater. Sci. Mater. Electron. 2015, 26, 8217–8223. [Google Scholar] [CrossRef]
  28. Park, S.; Kheel, H.; Sun, G.-J.; Ko, T.; Lee, W.I.; Lee, C. Acetone gas sensing properties of a multiple-networked Fe2O3-functionalized CuO nanorod sensor. J. Nanomater. 2015, 2015, 1–6. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, L.; Lou, Z.; Wang, R.; Fei, T.; Zhang, T. Ring-like PdO-decorated NiO with lamellar structures and their application in gas sensor. Sens. Actuators B Chem. 2012, 171–172, 1180–1185. [Google Scholar] [CrossRef]
  30. Choi, K.J.; Jang, H.W. One-dimensional oxide nanostructures as gas-sensing materials: Review and issues. Sensors 2010, 10, 4083–4099. [Google Scholar] [CrossRef] [Green Version]
  31. Karnati, P.; Akbar, S.; Morris, P.A. Conduction mechanisms in one dimensional core-shell nanostructures for gas sensing: A review. Sens. Actuators B Chem. 2019, 295, 127–143. [Google Scholar] [CrossRef]
  32. Meng, D.; Liu, D.; Wang, G.; Shen, Y.; San, X.; Li, M.; Meng, F. Low-temperature formaldehyde gas sensors based on NiO-SnO2 heterojunction microflowers assembled by thin porous nanosheets. Sens. Actuators B Chem. 2018, 273, 418–428. [Google Scholar] [CrossRef]
  33. Wang, Y.; Zhou, Y.; Meng, C.; Gao, Z.; Cao, X.; Li, X.; Xu, L.; Zhu, W.; Peng, X.; Zhang, B.; et al. A high-response ethanol gas sensor based on one-dimensional TiO2/V2O5 branched nanoheterostructures. Nanotechnology 2016, 27, 425503. [Google Scholar] [CrossRef] [PubMed]
  34. Huang, X.-Y.; Chi, Z.-T.; Liu, J.; Li, D.-H.; Sun, X.-J.; Yan, C.; Wang, Y.-C.; Li, H.; Wang, X.-D.; Xie, W.-F. Enhanced gas sensing performance based on p-NiS/n-In2O3 heterojunction nanocomposites. Sens. Actuators B Chem. 2020, 304, 127305. [Google Scholar] [CrossRef]
  35. 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. [Google Scholar] [CrossRef]
  36. Ju, D.; Xu, H.; Xu, Q.; Gong, H.; Qiu, Z.; Guo, J.; Zhang, J.; Cao, B. High triethylamine-sensing properties of NiO/SnO2 hollow sphere P-N heterojunction sensors. Sens. Actuators B Chem. 2015, 215, 39–44. [Google Scholar] [CrossRef]
  37. Chu, X.; Liang, S.; Chen, T.; Zhang, Q. Trimethylamine sensing properties of CdO–Fe2O3 nano-materials prepared using co-precipitation method in the presence of PEG400. Mater. Chem. Phys. 2010, 123, 396–400. [Google Scholar] [CrossRef]
  38. Li, S.; Xie, L.; He, M.; Hu, X.; Luo, G.; Chen, C.; Zhu, Z. Metal-Organic frameworks-derived bamboo-like CuO/In2O3 Heterostructure for high-performance H2S gas sensor with Low operating temperature. Sens. Actuators B Chem. 2020, 310, 127828. [Google Scholar] [CrossRef]
  39. Liu, J.; Wang, T.; Wang, B.; Sun, P.; Yang, Q.; Liang, X.; Song, H.; Lu, G. Highly sensitive and low detection limit of ethanol gas sensor based on hollow ZnO/SnO2 spheres composite material. Sens. Actuators B Chem. 2017, 245, 551–559. [Google Scholar] [CrossRef]
  40. Zhao, S.; Wang, G.; Liao, J.; Lv, S.; Zhu, Z.; Li, Z. Vertically aligned MoS2/ZnO nanowires nanostructures with highly enhanced NO2 sensing activities. Appl. Surf. Sci. 2018, 456, 808–816. [Google Scholar] [CrossRef]
  41. Bai, S.; Guo, W.; Sun, J.; Li, J.; Tian, Y.; Chen, A.; Luo, R.; Li, D. Synthesis of SnO2-CuO heterojunction using electrospinning and application in detecting of CO. Sens. Actuators B Chem. 2016, 226, 96–103. [Google Scholar] [CrossRef]
  42. Lou, Z.; Deng, J.; Wang, L.; Wang, L.; Fei, T.; Zhang, T. Toluene and ethanol sensing performances of pristine and PdO-decorated flower-like ZnO structures. Sens. Actuators B Chem. 2013, 176, 323–329. [Google Scholar] [CrossRef]
  43. Choi, K.S.; Park, S.; Chang, S.-P. Enhanced ethanol sensing properties based on SnO2 nanowires coated with Fe2O3 nanoparticles. Sens. Actuators B Chem. 2017, 238, 871–879. [Google Scholar] [CrossRef]
  44. Liang, Y.Q.; Cui, Z.D.; Zhu, S.L.; Li, Z.Y.; Yang, X.J.; Chen, Y.J.; Ma, J.M. Design of a highly sensitive ethanol sensor using a nano-coaxial p-Co3O4/n-TiO2 heterojunction synthesized at low temperature. Nanoscale 2013, 5, 10916–10926. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, L.; Li, J.; Wang, Y.; Yu, K.; Tang, X.; Zhang, Y.; Wang, S.; Wei, C. Construction of 1D SnO2-coated ZnO nanowire heterojunction for their improved n-butylamine sensing performances. Sci. Rep. 2016, 6, 35079. [Google Scholar] [CrossRef] [Green Version]
  46. Chowdhuri, A.; Gupta, V.; Sreenivas, K. Fast response H2S gas sensing characteristics with ultra-thin CuO islands on sputtered SnO2. Sens. Actuators B Chem. 2003, 93, 572–579. [Google Scholar] [CrossRef]
  47. Park, S.; Park, S.; Jung, J.; Hong, T.; Lee, S.; Kim, H.W.; Lee, C. H2S gas sensing properties of CuO-functionalized WO3 nanowires. Ceram. Int. 2014, 40, 11051–11056. [Google Scholar] [CrossRef]
  48. Han, T.-H.; Bak, S.-Y.; Kim, S.; Lee, S.H.; Han, Y.-J.; Yi, M. Decoration of CuO NWs Gas Sensor with ZnO NPs for Improving NO2 Sensing Characteristics. Sensors 2021, 21, 2103. [Google Scholar] [CrossRef]
  49. Tang, H.; Yan, M.; Zhang, H.; Li, S.; Ma, X.; Wang, M.; Yang, D. A selective NH3 gas sensor based on Fe2O3-ZnO nanocomposites at room temperature. Sens. Actuators B Chem. 2006, 114, 910–915. [Google Scholar] [CrossRef]
  50. Yoo, R.; Yoo, S.; Lee, D.; Kim, J.; Cho, S.; Lee, W. Highly selective detection of dimethyl methylphosphonate (DMMP) using CuO nanoparticles/ZnO flowers heterojunction. Sens. Actuators B Chem. 2017, 240, 1099–1105. [Google Scholar] [CrossRef]
  51. Ling, C.; Xue, Q.; Han, Z.; Lu, H.; Xia, F.; Yan, Z.; Deng, L. Room temperature hydrogen sensor with ultrahigh-responsive characteristics based on Pd/SnO2/SiO2/Si heterojunctions. Sens. Actuators B Chem. 2016, 227, 438–447. [Google Scholar] [CrossRef]
  52. Wang, L.; Lou, Z.; Zhang, R.; Zhou, T.; Deng, J.; Zhang, T. Hybrid Co3O4/SnO2 core-shell nanospheres as real-time rapid-response sensors for ammonia gas. ACS Appl. Mater. Interfaces 2016, 8, 6539–6545. [Google Scholar] [CrossRef] [PubMed]
  53. Hui, G.; Zhu, M.; Yang, X.; Liu, J.; Pan, G.; Wang, Z. Highly sensitive ethanol gas sensor based on CeO2/ZnO binary heterojunction composite. Mater. Lett. 2020, 278, 128453. [Google Scholar] [CrossRef]
  54. Choi, S.-W.; Katoch, A.; Sun, G.-J.; Kim, S.S. Synthesis and gas sensing performance of ZnO-SnO2 nanofiber-nanowire stem-branch heterostructure. Sens. Actuators B Chem. 2013, 181, 787–794. [Google Scholar] [CrossRef]
  55. Chen, Y.-J.; Xiao, G.; Wang, T.-S.; Zhang, F.; Ma, Y.; Gao, P.; Zhu, C.-L.; Zhang, E.; Xu, Z.; Li, Q.-H. α-MoO3/TiO2 core/shell nanorods: Controlled-synthesis and low-temperature gas sensing properties. Sens. Actuators B Chem. 2011, 155, 270–277. [Google Scholar] [CrossRef]
  56. Huang, H.; Gong, H.; Chow, C.L.; Guo, J.; White, T.J.; Tse, M.S.; Tan, O.K. Low-temperature growth of SnO2 nanorod arrays and tunable n-p-n sensing response of a ZnO/SnO2 heterojunction for exclusive hydrogen sensors. Adv. Funct. Mater. 2011, 21, 2680–2686. [Google Scholar] [CrossRef]
  57. Jin, C.; Park, S.; Kim, H.; Lee, C. Ultrasensitive multiple networked Ga2O3-core/ZnO-shell nanorod gas sensors. Sens. Actuators B Chem. 2012, 161, 223–228. [Google Scholar] [CrossRef]
  58. Li, J.; Liu, H.; Fu, H.; Xu, L.; Jin, H.; Zhang, X.; Wang, L.; Yu, K. Synthesis of 1D α-MoO3/0D ZnO heterostructure nanobelts with enhanced gas sensing properties. J. Alloys Compd. 2019, 788, 248–256. [Google Scholar] [CrossRef]
  59. Galatsis, K.; Li, Y.X.; Wlodarski, W.; Kalantar-Zadeh, K. Sol-gel prepared MoO3-WO3 thin-films for O2 gas sensing. Sens. Actuators B Chem. 2001, 77, 478–483. [Google Scholar] [CrossRef]
  60. Jiang, D.; Wei, W.; Li, F.; Li, Y.; Liu, C.; Sun, D.; Feng, C.; Ruan, S. Xylene gas sensor based on α-MoO3/α-Fe2O3 heterostructure with high response and low operating temperature. RSC Adv. 2015, 5, 39442–39448. [Google Scholar] [CrossRef]
  61. Zou, C.W.; Wang, J.; Xie, W. Synthesis and enhanced NO2 gas sensing properties of ZnO nanorods/TiO2 nanoparticles heterojunction composites. J. Colloid Interface Sci. 2016, 478, 22–28. [Google Scholar] [CrossRef] [PubMed]
  62. Hsu, K.-C.; Fang, T.-H.; Chen, S.-H.; Kuo, E.-Y. Gas sensitivity and sensing mechanism studies on ZnO/La0.8Sr0.2Co0.5Ni0.5O3 heterojunction structure. Ceram. Int. 2019, 45, 8744–8749. [Google Scholar] [CrossRef]
  63. Zhu, L.; Zeng, W.; Li, Y. A novel cactus-like WO3-SnO2 nanocomposite and its acetone gas sensing properties. Mater. Lett. 2018, 231, 5–7. [Google Scholar] [CrossRef]
  64. Park, S.; Kim, S.; Kheel, H.; Lee, C. Oxidizing gas sensing properties of the n-ZnO/p-Co3O4 composite nanoparticle network sensor. Sens. Actuators B Chem. 2016, 222, 1193–1200. [Google Scholar] [CrossRef]
  65. Poloju, M.; Jayababu, N.; Ramana Reddy, M.V. Improved gas sensing performance of Al doped ZnO/CuO nanocomposite based ammonia gas sensor. Mater. Sci. Eng. B 2018, 227, 61–67. [Google Scholar] [CrossRef]
  66. Wang, T.-S.; Wang, Q.-S.; Zhu, C.-L.; Ouyang, Q.-Y.; Qi, L.-H.; Li, C.-Y.; Xiao, G.; Gao, P.; Chen, Y.-J. Synthesis and enhanced H2S gas sensing properties of α-MoO3/CuO p-n junction nanocomposite. Sens. Actuators B Chem. 2012, 171–172, 256–262. [Google Scholar] [CrossRef]
  67. Kim, J.-H.; Lee, J.-H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Optimization and gas sensing mechanism of n-SnO2-p-Co3O4 composite nanofibers. Sens. Actuators B Chem. 2017, 248, 500–511. [Google Scholar] [CrossRef]
  68. Kabcum, S.; Channei, D.; Tuantranont, A.; Wisitsoraat, A.; Liewhiran, C.; Phanichphant, S. Ultra-responsive hydrogen gas sensors based on PdO nanoparticle-decorated WO3 nanorods synthesized by precipitation and impregnation methods. Sens. Actuators B Chem. 2016, 226, 76–89. [Google Scholar] [CrossRef]
  69. Xu, Q.; Ju, D.; Zhang, Z.; Yuan, S.; Zhang, J.; Xu, H.; Cao, B. Near room-temperature triethylamine sensor constructed with CuO/ZnO P-N heterostructural nanorods directly on flat electrode. Sens. Actuators B Chem. 2016, 225, 16–23. [Google Scholar] [CrossRef]
  70. Sun, G.-J.; Choi, S.-W.; Katoch, A.; Wu, P.; Kim, S.S. Bi-functional mechanism of H2S detection using CuO-SnO2 nanowires. J. Mater. Chem. C 2013, 1, 5454. [Google Scholar] [CrossRef]
  71. Huang, J.; Dai, Y.; Gu, C.; Sun, Y.; Liu, J. Preparation of porous flower-like CuO/ZnO nanostructures and analysis of their gas-sensing property. J. Alloys Compd. 2013, 575, 115–122. [Google Scholar] [CrossRef]
  72. Zhang, Y.-B.; Yin, J.; Li, L.; Zhang, L.-X.; Bie, L.-J. Enhanced ethanol gas-sensing properties of flower-like p-CuO/n-ZnO heterojunction nanorods. Sens. Actuators B Chem. 2014, 202, 500–507. [Google Scholar] [CrossRef]
  73. Xu, L.; Zheng, R.; Liu, S.; Song, J.; Chen, J.; Dong, B.; Song, H. NiO@ZnO heterostructured nanotubes: Coelectrospinning fabrication, characterization, and highly enhanced gas sensing properties. Inorg. Chem. 2012, 51, 7733–7740. [Google Scholar] [CrossRef]
  74. Lokesh, K.; Kavitha, G.; Manikandan, E.; Mani, G.K.; Kaviyarasu, K.; Rayappan, J.B.B.; Ladchumananandasivam, R.; Aanand, J.S.; Jayachandran, M.; Maaza, M. Effective ammonia detection using n-ZnO/p-NiO heterostructured nanofibers. IEEE Sens. J. 2016, 16, 2477–2483. [Google Scholar] [CrossRef]
  75. Wang, H.; Chen, M.; Rong, Q.; Zhang, Y.; Hu, J.; Zhang, D.; Zhou, S.; Zhao, X.; Zhang, J.; Zhu, Z. Ultrasensitive xylene gas sensor based on flower-like SnO2/Co3O4 nanorods composites prepared by facile two-step synthesis method. Nanotechnology 2020, 31, 255501. [Google Scholar] [CrossRef]
  76. Ng, S.; Kuberský, P.; Krbal, M.; Prikryl, J.; Gärtnerová, V.; Moravcová, D.; Sopha, H.; Zazpe, R.; Yam, F.K.; Jäger, A.; et al. ZnO coated anodic 1D TiO2 nanotube layers: Efficient photo-electrochemical and gas sensing heterojunction. Adv. Eng. Mater. 2018, 20, 1700589. [Google Scholar] [CrossRef] [Green Version]
  77. Yao, Y.; Yuan, J.; Chen, X.; Tan, L.; Gu, Q.; Zhao, W.; Chen, J. In situ construction and sensing mechanism of TiO2-WO3 composite coatings based on the semiconductor heterojunctions. J. Mater. Res. Technol. 2019, 8, 3580–3588. [Google Scholar] [CrossRef]
  78. Zhao, P.X.; Tang, Y.; Mao, J.; Chen, Y.X.; Song, H.; Wang, J.W.; Song, Y.; Liang, Y.Q.; Zhang, X.M. One-dimensional MoS2-decorated TiO2 nanotube gas sensors for efficient alcohol sensing. J. Alloys Compd. 2016, 674, 252–258. [Google Scholar] [CrossRef]
  79. Sun, P.; Cai, Y.; Du, S.; Xu, X.; You, L.; Ma, J.; Liu, F.; Liang, X.; Sun, Y.; Lu, G. Hierarchical α-Fe2O3/SnO2 semiconductor composites: Hydrothermal synthesis and gas sensing properties. Sens. Actuators B Chem. 2013, 182, 336–343. [Google Scholar] [CrossRef]
  80. Zhang, S.; Song, P.; Wang, Q. Enhanced acetone sensing performance of an α-Fe2O3-In2O3 heterostructure nanocomposite sensor. J. Phys. Chem. Solids 2018, 120, 261–270. [Google Scholar] [CrossRef]
  81. Feng, C.; Li, X.; Ma, J.; Sun, Y.; Wang, C.; Sun, P.; Zheng, J.; Lu, G. Facile synthesis and gas sensing properties of In2O3-WO3 heterojunction nanofibers. Sens. Actuators B Chem. 2015, 209, 622–629. [Google Scholar] [CrossRef]
  82. Chi, X.; Liu, C.; Liu, L.; Li, S.; Li, H.; Zhang, X.; Bo, X.; Shan, H. Enhanced formaldehyde-sensing properties of mixed Fe2O3-In2O3 nanotubes. Mater. Sci. Semicond. Process. 2014, 18, 160–164. [Google Scholar] [CrossRef]
  83. Park, S. Acetone gas detection using TiO2 nanoparticles functionalized In2O3 nanowires for diagnosis of diabetes. J. Alloys Compd. 2017, 696, 655–662. [Google Scholar] [CrossRef]
  84. Du, H.; Wang, J.; Sun, Y.; Yao, P.; Li, X.; Yu, N. Investigation of gas sensing properties of SnO2/In2O3 composite hetero-nanofibers treated by oxygen plasma. Sens. Actuators B Chem. 2015, 206, 753–763. [Google Scholar] [CrossRef]
  85. Wang, X.; Li, S.; Xie, L.; Li, X.; Lin, D.; Zhu, Z. Low-temperature and highly sensitivity H2S gas sensor based on ZnO/CuO composite derived from bimetal metal-organic frameworks. Ceram. Int. 2020, 46, 15858–15866. [Google Scholar] [CrossRef]
  86. Chang, H.-K.; Ko, D.-S.; Cho, D.-H.; Kim, S.; Lee, H.-N.; Lee, H.S.; Kim, H.-J.; Park, T.J.; Park, Y.M. Enhanced response of the photoactive gas sensor on formaldehyde using porous SnO2@TiO2 heterostructure driven by gas-flow thermal evaporation and atomic layer deposition. Ceram. Int. 2021, 47, 5985–5992. [Google Scholar] [CrossRef]
  87. Katoch, A.; Kim, J.H.; Kwon, Y.J.; Kim, H.W.; Kim, S.S. Bifunctional sensing mechanism of SnO2-ZnO composite nanofibers for drastically enhancing the sensing behavior in H2 gas. ACS Appl. Mater. Interfaces 2015, 7, 11351–11358. [Google Scholar] [CrossRef]
  88. Ma, L.; Fan, H.; Tian, H.; Fang, J.; Qian, X. The n-ZnO/n-In2O3 heterojunction formed by a surface-modification and their potential barrier-control in methanal gas sensing. Sens. Actuators B Chem. 2016, 222, 508–516. [Google Scholar] [CrossRef]
  89. Xun, H.; Zhang, Z.; Yu, A.; Yi, J. Remarkably enhanced hydrogen sensing of highly-ordered SnO2-decorated TiO2 nanotubes. Sens. Actuators B Chem. 2018, 273, 983–990. [Google Scholar] [CrossRef]
  90. Yu, H.; Yang, T.; Wang, Z.; Li, Z.; Zhao, Q.; Zhang, M. p-N heterostructural sensor with SnO-SnO2 for fast NO2 sensing response properties at room temperature. Sens. Actuators B Chem. 2018, 258, 517–526. [Google Scholar] [CrossRef]
  91. Gao, H.; Zhao, L.; Wang, L.; Sun, P.; Lu, H.; Liu, F.; Chuai, X.; Lu, G. Ultrasensitive and low detection limit of toluene gas sensor based on SnO2-decorated NiO nanostructure. Sens. Actuators B Chem. 2018, 255, 3505–3515. [Google Scholar] [CrossRef]
  92. Zhang, L.; Gao, Z.; Liu, C.; Zhang, Y.; Tu, Z.; Yang, X.; Yang, F.; Wen, Z.; Zhu, L.; Liu, R.; et al. Synthesis of TiO2 decorated Co3O4 acicular nanowire arrays and their application as an ethanol sensor. J. Mater. Chem. A 2015, 3, 2794–2801. [Google Scholar] [CrossRef]
  93. Li, X.; Li, X.; Chen, N.; Li, X.; Zhang, J.; Yu, J.; Wang, J.; Tang, Z. CuO-In2O3core-shell nanowire based chemical gas sensors. J. Nanomater. 2014, 2014, 973156. [Google Scholar] [CrossRef] [Green Version]
  94. Wang, Z.; Zhang, K.; Fei, T.; Gu, F.; Han, D. α-Fe2O3/NiO heterojunction nanorods with enhanced gas sensing performance for acetone. Sens. Actuators B Chem. 2020, 318, 128191. [Google Scholar] [CrossRef]
  95. Deng, J.; Wang, L.; Lou, Z.; Zhang, T. Design of CuO-TiO2 heterostructure nanofibers and their sensing performance. J. Mater. Chem. A 2014, 2, 9030–9034. [Google Scholar] [CrossRef]
  96. Shi, S.; Zhang, F.; Lin, H.; Wang, Q.; Shi, E.; Qu, F. Enhanced triethylamine-sensing properties of P-N heterojunction Co3O4/In2O3 hollow microtubes derived from metal-organic frameworks. Sens. Actuators B Chem. 2018, 262, 739–749. [Google Scholar] [CrossRef]
  97. Mandal, S.; Rakibuddin, M.; Ananthakrishnan, R. Strategic synthesis of SiO2-modified porous Co3O4 nano-octahedra through the nanocoordination polymer route for enhanced and selective sensing of H2 gas over NOx. ACS Omega 2018, 3, 648–661. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, B.; Li, M.; Song, Z.; Kan, H.; Yu, H.; Liu, Q.; Zhang, G.; Liu, H. Sensitive H2S gas sensors employing colloidal zinc oxide quantum dots. Sens. Actuators B Chem. 2017, 249, 558–563. [Google Scholar] [CrossRef]
  99. Zhang, L.; Dong, B.; Xu, L.; Zhang, X.; Chen, J.; Sun, X.; Xu, H.; Zhang, T.; Bai, X.; Zhang, S.; et al. Three-dimensional ordered ZnO-Fe3O4 inverse opal gas sensor toward trace concentration acetone detection. Sens. Actuators B Chem. 2017, 252, 367–374. [Google Scholar] [CrossRef]
  100. Kim, J.-H.; Jeong, H.-M.; Na, C.W.; Yoon, J.-W.; Abdel-Hady, F.; Wazzan, A.A.; Lee, J.-H. Highly selective and sensitive xylene sensors using Cr2O3-ZnCr2O4 hetero-nanostructures prepared by galvanic replacement. Sens. Actuators B Chem. 2016, 235, 498–506. [Google Scholar] [CrossRef]
  101. Koo, W.T.; Yu, S.; Choi, S.J.; Jang, J.S.; Cheong, J.Y.; Kim, I.D. Nanoscale PdO catalyst functionalized Co3O4 hollow nanocages using MOF templates for selective detection of acetone molecules in exhaled breath. ACS Appl. Mater. Interfaces 2017, 9, 8201–8210. [Google Scholar] [CrossRef] [PubMed]
  102. Park, S.; Kim, S.; Sun, G.-J.; Lee, W.; Kim, K.K.; Lee, C. Fabrication and NO2 gas sensing performance of TeO2-core/CuO-shell heterostructure nanorod sensors. Nanoscale Res. Lett. 2014, 9, 638. [Google Scholar] [CrossRef] [Green Version]
  103. Xu, Y.; Fan, Y.; Tian, X.; Liang, Q.; Liu, X.; Sun, Y. p-p heterojunction composite of NiFe2O4 nanoparticles-decorated NiO nanosheets for acetone gas detection. Mater. Lett. 2020, 270, 127728. [Google Scholar] [CrossRef]
  104. Su, C.; Zhang, L.; Han, Y.; Ren, C.; Zeng, M.; Zhou, Z.; Su, Y.; Hu, N.; Wei, H.; Yang, Z. Controllable synthesis of heterostructured CuO-NiO nanotubes and their synergistic effect for glycol gas sensing. Sens. Actuators B Chem. 2020, 304, 127347. [Google Scholar] [CrossRef]
  105. Minh Nguyet, Q.T.; Van Duy, N.; Phuong, N.T.; Trung, N.N.; Hung, C.M.; Hoa, N.D.; Van Hieu, N. Superior enhancement of NO2 gas response using n-p-n transition of carbon nanotubes/SnO2 nanowires heterojunctions. Sens. Actuators B Chem. 2017, 238, 1120–1127. [Google Scholar] [CrossRef]
  106. Kwon, Y.J.; Kang, S.Y.; Mirzaei, A.; Choi, M.S.; Bang, J.H.; Kim, S.S.; Kim, H.W. Enhancement of gas sensing properties by the functionalization of ZnO-branched SnO 2 nanowires with Cr2O3 nanoparticles. Sens. Actuators B Chem. 2017, 249, 656–666. [Google Scholar] [CrossRef]
  107. Pang, Z.; Nie, Q.; Lv, P.; Yu, J.; Huang, F.; Wei, Q. Design of flexible PANI-coated CuO-TiO2-SiO2 heterostructure nanofibers with high ammonia sensing response values. Nanotechnology 2017, 28, 225501. [Google Scholar] [CrossRef]
  108. Volanti, D.P.; Felix, A.A.; Orlandi, M.O.; Whitfield, G.; Yang, D.-J.; Longo, E.; Tuller, H.L.; Varela, J.A. The role of hierarchical morphologies in the superior gas sensing performance of CuO-based chemiresistors. Adv. Funct. Mater. 2013, 23, 1759–1766. [Google Scholar] [CrossRef]
  109. Yang, S.; Lei, G.; Lan, Z.; Xie, W.; Yang, B.; Xu, H.; Wang, Z.; Gu, H. Enhancement of the room-temperature hydrogen sensing performance of MoO3 nanoribbons annealed in a reducing gas. Int. J. Hydrog. Energy 2019, 44, 7725–7733. [Google Scholar] [CrossRef]
  110. Park, S.; Kim, S.; Sun, G.-J.; Choi, S.; Lee, S.; Lee, C. Ethanol sensing properties of networked In2O3 nanorods decorated with Cr2O3-nanoparticles. Ceram. Int. 2015, 41, 9823–9827. [Google Scholar] [CrossRef]
  111. He, M.; Xie, L.; Zhao, X.; Hu, X.; Li, S.; Zhu, Z.-G. Highly sensitive and selective H2S gas sensors based on flower-like WO3/CuO composites operating at low/room temperature. J. Alloys Compd. 2019, 788, 36–43. [Google Scholar] [CrossRef]
  112. Bang, J.H.; Choi, M.S.; Mirzaei, A.; Kwon, Y.J.; Kim, S.S.; Kim, T.W.; Kim, H.W. Selective NO2 sensor based on Bi2O3 branched SnO2 nanowires. Sens. Actuators B Chem. 2018, 274, 356–369. [Google Scholar] [CrossRef]
  113. Tan, W.; Tan, J.; Fan, L.; Yu, Z.; Qian, J.; Huang, X. Fe2O3-loaded NiO nanosheets for fast response/recovery and high response gas sensor. Sens. Actuators B Chem. 2018, 256, 282–293. [Google Scholar] [CrossRef]
  114. Ju, D.; Xu, H.; Qiu, Z.; Guo, J.; Zhang, J.; Cao, B. Highly sensitive and selective triethylamine-sensing properties of nanosheets directly grown on ceramic tube by forming NiO/ZnO PN heterojunction. Sens. Actuators B Chem. 2014, 200, 288–296. [Google Scholar] [CrossRef]
  115. Yu, Q.; Zhu, J.; Xu, Z.; Huang, X. Facile synthesis of α-Fe2O3@SnO2 core–shell heterostructure nanotubes for high performance gas sensors. Sens. Actuators B Chem. 2015, 213, 27–34. [Google Scholar] [CrossRef]
  116. Rout, C.S.; Hari Krishna, S.; Vivekchand, S.R.C.; Govindaraj, A.; Rao, C.N.R. Hydrogen and ethanol sensors based on ZnO nanorods, nanowires and nanotubes. Chem. Phys. Lett. 2006, 418, 586–590. [Google Scholar] [CrossRef]
  117. Feng, P.; Wan, Q.; Wang, T.H. Contact-controlled sensing properties of flowerlike ZnO nanostructures. Appl. Phys. Lett. 2005, 87, 213111. [Google Scholar] [CrossRef]
  118. Yu, H.-L.; Li, L.; Gao, X.-M.; Zhang, Y.; Meng, F.; Wang, T.-S.; Xiao, G.; Chen, Y.-J.; Zhu, C.-L. Synthesis and H2S gas sensing properties of cage-like α-MoO3/ZnO composite. Sens. Actuators B Chem. 2012, 171–172, 679–685. [Google Scholar] [CrossRef]
  119. Kheel, H.; Sun, G.-J.; Lee, J.K.; Mirzaei, A.; Choi, S.; Lee, C. Hydrogen gas detection of Nb2O5 nanoparticle-decorated CuO nanorod sensors. Met. Mater. Int. 2017, 23, 214–219. [Google Scholar] [CrossRef]
  120. Gao, X.; Ouyang, Q.; Zhu, C.; Zhang, X.; Chen, Y. Porous MoO3/SnO2 nanoflakes with n-n junctions for sensing H2S. ACS Appl. Nano Mater. 2019, 2, 2418–2425. [Google Scholar] [CrossRef]
  121. Kheel, H.; Sun, G.-J.; Lee, J.K.; Lee, S.; Dwivedi, R.P.; Lee, C. Enhanced H2S sensing performance of TiO2-decorated α-Fe2O3 nanorod sensors. Ceram. Int. 2016, 42, 18597–18604. [Google Scholar] [CrossRef]
  122. Qu, F.; Zhou, X.; Zhang, B.; Zhang, S.; Jiang, C.; Ruan, S.; Yang, M. Fe2O3 nanoparticles-decorated MoO3 nanobelts for enhanced chemiresistive gas sensing. J. Alloys Compd. 2019, 782, 672–678. [Google Scholar] [CrossRef]
  123. Mirzaei, A.; Sun, G.-J.; Lee, J.K.; Lee, C.; Choi, S.; Kim, H.W. Hydrogen sensing properties and mechanism of NiO-Nb2O5 composite nanoparticle-based electrical gas sensors. Ceram. Int. 2017, 43, 5247–5254. [Google Scholar] [CrossRef]
  124. Park, S.; Sun, G.-J.; Kheel, H.; Hyun, S.K.; Jin, C.; Lee, C. Hydrogen gas sensing of Co3O4-Decorated WO3 nanowires. Met. Mater. Int. 2016, 22, 156–162. [Google Scholar] [CrossRef]
  125. Liang, X.; Kim, T.-H.; Yoon, J.-W.; Kwak, C.-H.; Lee, J.-H. Ultrasensitive and ultraselective detection of H2S using electrospun CuO-loaded In2O3 nanofiber sensors assisted by pulse heating. Sens. Actuators B Chem. 2015, 209, 934–942. [Google Scholar] [CrossRef]
  126. Shanmugasundaram, A.; Basak, P.; Satyanarayana, L.; Manorama, S.V. Hierarchical SnO/SnO2 nanocomposites: Formation of in situ p-n junctions and enhanced H2 sensing. Sens. Actuators B Chem. 2013, 185, 265–273. [Google Scholar] [CrossRef]
  127. Cai, L.; Li, H.; Zhang, H.; Fan, W.; Wang, J.; Wang, Y.; Wang, X.; Tang, Y.; Song, Y. Enhanced performance of the tangerines-like CuO-based gas sensor using ZnO nanowire arrays. Mater. Sci. Semicond. Process. 2020, 118, 105196. [Google Scholar] [CrossRef]
  128. Wang, C.; Cheng, X.; Zhou, X.; Sun, P.; Hu, X.; Shimanoe, K.; Lu, G.; Yamazoe, N. Hierarchical alpha-Fe2O3/NiO composites with a hollow structure for a gas sensor. ACS Appl. Mater. Interfaces 2014, 6, 12031–12037. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, T.; Yu, Z.; Liu, Y.; Gao, J.; Wang, X.; Suo, H.; Yang, X.; Zhao, C.; Liu, F. Gas sensor based on Ni foam: SnO2-decorated NiO for Toluene detection. Sens. Actuators B Chem. 2020, 318, 128167. [Google Scholar] [CrossRef]
  130. Xue, X.-T.; Zhu, L.-Y.; Yuan, K.-P.; Zeng, C.; Li, X.-X.; Ma, H.-P.; Lu, H.-L.; Zhang, D.W. ZnO branched p-CuxO@n-ZnO heterojunction nanowires for improving acetone gas sensing performance. Sens. Actuators B Chem. 2020, 324, 128729. [Google Scholar] [CrossRef]
  131. Wang, Y.; Qu, F.; Liu, J.; Wang, Y.; Zhou, J.; Ruan, S. Enhanced H2S sensing characteristics of CuO-NiO core-shell microspheres sensors. Sens. Actuators B Chem. 2015, 209, 515–523. [Google Scholar] [CrossRef]
  132. Gao, H.; Guo, J.; Li, Y.; Xie, C.; Li, X.; Liu, L.; Chen, Y.; Sun, P.; Liu, F.; Yan, X.; et al. Highly selective and sensitive xylene gas sensor fabricated from NiO/NiCr2O4 p-p nanoparticles. Sens. Actuators B Chem. 2019, 284, 305–315. [Google Scholar] [CrossRef]
  133. Yoon, J.-W.; Kim, H.-J.; Jeong, H.-M.; Lee, J.-H. Gas sensing characteristics of p-type Cr2O3 and Co3O4 nanofibers depending on inter-particle connectivity. Sens. Actuators B Chem. 2014, 202, 263–271. [Google Scholar] [CrossRef]
  134. Zhao, Y.; Ikram, M.; Wang, J.; Liu, Z.; Du, L.; Zhou, J.; Kan, K.; Zhang, W.; Li, L.; Shi, K. Ultrafast NH3 sensing properties of WO3@CoWO4 heterojunction nanofibres at room temperature. Aust. J. Chem. 2018, 71, 87–94. [Google Scholar] [CrossRef]
  135. Zhou, C.; Xu, L.; Song, J.; Xing, R.; Xu, S.; Liu, D.; Song, H. Ultrasensitive non-enzymatic glucose sensor based on three-dimensional network of ZnO-CuO hierarchical nanocomposites by electrospinning. Sci. Rep. 2014, 4, 7382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Zhang, Y.; Zhang, J.; Jiang, Y.; Duan, Z.; Liu, B.; Zhao, Q.; Wang, S.; Yuan, Z.; Tai, H. Ultrasensitive flexible NH3 gas sensor based on polyaniline/SrGe4O9 nanocomposite with ppt-level detection ability at room temperature. Sens. Actuators B Chem. 2020, 319, 128293. [Google Scholar] [CrossRef]
  137. Gong, J.; Li, Y.; Hu, Z.; Zhou, Z.; Deng, Y. Ultrasensitive NH3 gas sensor from polyaniline nanograin enchased TiO2 fibers. J. Phys. Chem. C 2010, 114, 9970–9974. [Google Scholar] [CrossRef]
  138. Li, S.; Liu, A.; Yang, Z.; He, J.; Wang, J.; Liu, F.; Lu, H.; Yan, X.; Sun, P.; Liang, X.; et al. Room temperature gas sensor based on tin dioxide@polyaniline nanocomposite assembled on flexible substrate: Ppb-level detection of NH3. Sens. Actuators B Chem. 2019, 299, 126970. [Google Scholar] [CrossRef]
  139. Naderi, H.; Hajati, S.; Ghaedi, M.; Dashtian, K.; Sabzehmeidani, M.M. Sensitive, selective and rapid ammonia-sensing by gold nanoparticle-sensitized V2O5/CuWO4 heterojunctions for exhaled breath analysis. Appl. Surf. Sci. 2020, 501, 144270. [Google Scholar] [CrossRef]
  140. Hung, C.M.; Dat, D.Q.; Van Duy, N.; Van Quang, V.; Van Toan, N.; Van Hieu, N.; Hoa, N.D. Facile synthesis of ultrafine rGO/WO3 nanowire nanocomposites for highly sensitive toxic NH3 gas sensors. Mater. Res. Bull. 2020, 125, 110810. [Google Scholar] [CrossRef]
  141. Yuan, K.-P.; Zhu, L.-Y.; Yang, J.-H.; Hang, C.-Z.; Tao, J.-J.; Ma, H.-P.; Jiang, A.-Q.; Zhang, D.W.; Lu, H.-L. Precise preparation of WO3@SnO2 core shell nanosheets for efficient NH3 gas sensing. J. Colloid Interface Sci. 2020, 568, 81–88. [Google Scholar] [CrossRef]
  142. Liu, C.; Tai, H.; Zhang, P.; Yuan, Z.; Du, X.; Xie, G.; Jiang, Y. A high-performance flexible gas sensor based on self-assembled PANI-CeO2 nanocomposite thin film for trace-level NH3 detection at room temperature. Sens. Actuators B Chem. 2018, 261, 587–597. [Google Scholar] [CrossRef]
  143. Huang, J.; Jiang, D.; Zhou, J.; Ye, J.; Sun, Y.; Li, X.; Geng, Y.; Wang, J.; Du, Y.; Qian, Z. Visible light-activated room temperature NH3 sensor base on CuPc-loaded ZnO nanorods. Sens. Actuators B Chem. 2021, 327, 128911. [Google Scholar] [CrossRef]
  144. Zhang, D.; Jiang, C.; Li, P.; Sun, Y.E. Layer-by-layer self-assembly of Co3O4 nanorod-decorated MoS2 nanosheet-based nanocomposite toward high-performance ammonia detection. ACS Appl. Mater. Interfaces 2017, 9, 6462–6471. [Google Scholar] [CrossRef]
  145. Wang, Y.; Zhang, H.; Sun, X. Electrospun nanowebs of NiO/SnO2 p-n heterojunctions for enhanced gas sensing. Appl. Surf. Sci. 2016, 389, 514–520. [Google Scholar] [CrossRef]
  146. Van Toan, N.; Hung, C.M.; van Duy, N.; Hoa, N.D.; Le, D.T.; van Hieu, N. Bilayer SnO2–WO3 nanofilms for enhanced NH3 gas sensing performance. Mater. Sci. Eng. B 2017, 224, 163–170. [Google Scholar] [CrossRef]
  147. Ding, Y.; Guo, X.; Du, B.; Hu, X.; Yang, X.; He, Y.; Zhou, Y.; Zang, Z. Low-operating temperature ammonia sensor based on Cu2O nanoparticles decorated with p-type MoS2 nanosheets. J. Mater. Chem. C 2021, 9, 4838–4846. [Google Scholar] [CrossRef]
  148. Kumar, A.; Sanger, A.; Kumar, A.; Chandra, R. Fast response ammonia sensors based on TiO2 and NiO nanostructured bilayer thin films. RSC Adv. 2016, 6, 77636–77643. [Google Scholar] [CrossRef]
  149. Zhou, J.; Ikram, M.; Rehman, A.U.; Wang, J.; Zhao, Y.; Kan, K.; Zhang, W.; Raziq, F.; Li, L.; Shi, K. Highly selective detection of NH3 and H2S using the pristine CuO and mesoporous In2O3@CuO multijunctions nanofibers at room temperature. Sens. Actuators B Chem. 2018, 255, 1819–1830. [Google Scholar] [CrossRef]
  150. Kim, S.S.; Choi, S.-W.; Na, H.G.; Kwak, D.S.; Kwon, Y.J.; Kim, H.W. ZnO–SnO2 branch–stem nanowires based on a two-step process: Synthesis and sensing capability. Curr. Appl. Phys. 2013, 13, 526–532. [Google Scholar] [CrossRef]
  151. Bai, S.; Fu, H.; Zhao, Y.; Tian, K.; Luo, R.; Li, D.; Chen, A. On the construction of hollow nanofibers of ZnO-SnO2 heterojunctions to enhance the NO2 sensing properties. Sens. Actuators B Chem. 2018, 266, 692–702. [Google Scholar] [CrossRef]
  152. Xu, S.; Gao, J.; Wang, L.; Kan, K.; Xie, Y.; Shen, P.; Li, L.; Shi, K. Role of the heterojunctions in In2O3-composite SnO2 nanorod sensors and their remarkable gas-sensing performance for NOx at room temperature. Nanoscale 2015, 7, 14643–14651. [Google Scholar] [CrossRef] [PubMed]
  153. Park, S.; An, S.; Mun, Y.; Lee, C. UV-enhanced NO2 gas sensing properties of SnO2-core/ZnO-shell nanowires at room temperature. ACS Appl. Mater. Interfaces 2013, 5, 4285–4292. [Google Scholar] [CrossRef]
  154. Zhang, J.; Zeng, D.; Zhu, Q.; Wu, J.; Huang, Q.; Zhang, W.; Xie, C. Enhanced room temperature NO2 response of NiO-SnO2 nanocomposites induced by interface bonds at the p-n heterojunction. Phys. Chem. Chem. Phys. 2016, 18, 5386–5396. [Google Scholar] [CrossRef] [PubMed]
  155. Hoa, L.T.; Hur, S.H. Highly sensitive NO2 sensors based on local p-n heterojunctions composed of 0D CuO nanoparticles and 1D ZnO nanorods. Phys. Status Solidi A 2013, 210, 1213–1216. [Google Scholar] [CrossRef]
  156. Yeh, B.Y.; Huang, P.F.; Tseng, W.J. Enhanced room-temperature NO2 gas sensing with TeO2/SnO2 brush- and bead-like nanowire hybrid structures. Nanotechnology 2017, 28, 045501. [Google Scholar] [CrossRef] [PubMed]
  157. Ramakrishnan, V.; Nair, K.G.; Dhakshinamoorthy, J.; Ravi, K.R.; Pullithadathil, B. Porous, n–p type ultra-long, ZnO@Bi2O3 heterojunction nanorods-based NO2 gas sensor: New insights towards charge transport characteristics. Phys. Chem. Chem. Phys. 2020, 22, 7524–7536. [Google Scholar] [CrossRef] [PubMed]
  158. Xu, H.; Zhang, J.; Rehman, A.U.; Gong, L.; Kan, K.; Li, L.; Shi, K. Synthesis of NiO@CuO nanocomposite as high-performance gas sensing material for NO2 at room temperature. Appl. Surf. Sci. 2017, 412, 230–237. [Google Scholar] [CrossRef]
  159. Adamu, B.I.; Falak, A.; Tian, Y.; Tan, X.; Meng, X.; Chen, P.; Wang, H.; Chu, W. p-p heterojunction sensors of p-Cu3Mo2O9 micro/nanorods vertically grown on p-CuO layers for room-temperature ultrasensitive and fast recoverable detection of NO2. ACS Appl. Mater. Interfaces 2020, 12, 8411–8421. [Google Scholar] [CrossRef]
Figure 1. The number of the published papers on nanostructured metal oxide heterojunctions for high-performance gas sensors during 2009–2020 as obtained from the Web of Science. The words “nanostructured metal oxide heterojunctions” or “high-performance gas sensors” were keyed into the “topic” search box.
Figure 1. The number of the published papers on nanostructured metal oxide heterojunctions for high-performance gas sensors during 2009–2020 as obtained from the Web of Science. The words “nanostructured metal oxide heterojunctions” or “high-performance gas sensors” were keyed into the “topic” search box.
Nanomaterials 11 01026 g001
Figure 2. Typical types of heterojunctions for high-performance gas sensors and assembly strategies.
Figure 2. Typical types of heterojunctions for high-performance gas sensors and assembly strategies.
Nanomaterials 11 01026 g002
Figure 3. SEM images of the pure SnO2 hollow spheres (a,b) and ZnO-decorated SnO2 (n-n junctions) (c,d), sensor response of pure SnO2 hollow spheres and the ZnO-decorated SnO2 hollow spheres to 30 ppm ethanol at different operating temperatures (e), the sensor response of ZnO-decorated SnO2 hollow spheres to 30 ppm ethanol, acetone and methanol at different operating temperatures (f), and the dynamic sensing performance of the decorated SnO2 hollow spheres towards 30 ppm ethanol or acetone at 225 °C (g). Copied with permission from reference [39]. Copyright 2017, Elsevier.
Figure 3. SEM images of the pure SnO2 hollow spheres (a,b) and ZnO-decorated SnO2 (n-n junctions) (c,d), sensor response of pure SnO2 hollow spheres and the ZnO-decorated SnO2 hollow spheres to 30 ppm ethanol at different operating temperatures (e), the sensor response of ZnO-decorated SnO2 hollow spheres to 30 ppm ethanol, acetone and methanol at different operating temperatures (f), and the dynamic sensing performance of the decorated SnO2 hollow spheres towards 30 ppm ethanol or acetone at 225 °C (g). Copied with permission from reference [39]. Copyright 2017, Elsevier.
Nanomaterials 11 01026 g003
Figure 4. SEM images of pure flower-like ZnO (a,b) and the PdO nanoparticle-decorated ZnO (n-p junctions) (c,d), the dynamic sensing performance of the sensor based on the PdO nanoparticle-decorated ZnO (e) or pure ZnO (f) to 100 ppm ethanol at 320 °C. Copied with permission [42]. Copyright 2013, Elsevier.
Figure 4. SEM images of pure flower-like ZnO (a,b) and the PdO nanoparticle-decorated ZnO (n-p junctions) (c,d), the dynamic sensing performance of the sensor based on the PdO nanoparticle-decorated ZnO (e) or pure ZnO (f) to 100 ppm ethanol at 320 °C. Copied with permission [42]. Copyright 2013, Elsevier.
Nanomaterials 11 01026 g004
Figure 5. The XRD pattern (a) and the SEM images (bd) of SnO2 nanoparticles-modified NiO nanostructure (p-n junctions), the dynamic sensing performances of the pure NiO and the SnO2 nanoparticles-modified NiO nanostructures towards 1–200 ppm toluene at 250 °C (c,eh). Copied with permission [91]. Copyright 2018, Elsevier.
Figure 5. The XRD pattern (a) and the SEM images (bd) of SnO2 nanoparticles-modified NiO nanostructure (p-n junctions), the dynamic sensing performances of the pure NiO and the SnO2 nanoparticles-modified NiO nanostructures towards 1–200 ppm toluene at 250 °C (c,eh). Copied with permission [91]. Copyright 2018, Elsevier.
Nanomaterials 11 01026 g005
Figure 6. SEM image (a), TEM image (b), HRTEM (c) and SEAD patterns (d) of the Co3O4 hollow nanocages (HNCs) decorated with PdO nanoparticles (p-p junctions), the dynamic sensing behaviours of the sensors based on Co3O4 powders, pure Co3O4 hollow nanocages, PdO-Co3O4 powders and PdO-Co3O4 HNCs towards 0.4–5 ppm acetone at 350 °C (e), the stability of the sensing performance of PdO-Co3O4 HNCs towards 1 ppm acetone (f). Copied with permission [101]. Copyright 2017, American Chemical Society.
Figure 6. SEM image (a), TEM image (b), HRTEM (c) and SEAD patterns (d) of the Co3O4 hollow nanocages (HNCs) decorated with PdO nanoparticles (p-p junctions), the dynamic sensing behaviours of the sensors based on Co3O4 powders, pure Co3O4 hollow nanocages, PdO-Co3O4 powders and PdO-Co3O4 HNCs towards 0.4–5 ppm acetone at 350 °C (e), the stability of the sensing performance of PdO-Co3O4 HNCs towards 1 ppm acetone (f). Copied with permission [101]. Copyright 2017, American Chemical Society.
Nanomaterials 11 01026 g006
Figure 7. The enhanced gas sensing mechanisms of (a1,a2) MoO3/SnO2 nanoflakes (n-n heterojunction) (Copied with permission [120]. Copyright 2019, American Chemical Society), (b1,b2) NiO-Nb2O5 composite nanoparticles (n-p heterojunction) (Copied with permission [123]. Copyright 2017, Elsevier) and (c1,c2) Nb2O5 nanoparticle-decorated CuO nanorods (p-n heterojunction) (Copied with permission [119]. Copyright 2017, Springer Nature) to reducing gases.
Figure 7. The enhanced gas sensing mechanisms of (a1,a2) MoO3/SnO2 nanoflakes (n-n heterojunction) (Copied with permission [120]. Copyright 2019, American Chemical Society), (b1,b2) NiO-Nb2O5 composite nanoparticles (n-p heterojunction) (Copied with permission [123]. Copyright 2017, Elsevier) and (c1,c2) Nb2O5 nanoparticle-decorated CuO nanorods (p-n heterojunction) (Copied with permission [119]. Copyright 2017, Springer Nature) to reducing gases.
Nanomaterials 11 01026 g007
Figure 8. The enhanced gas sensing mechanisms of (a1,a2) the hollow nanofibres of ZnO-SnO2 (n-n heterojunction) (Copied with permission [151]. Copyright 2018, Elsevier.), (b1,b2) the ZnO/Co3O4 composite nanoparticle (n-p heterojunction) (Copied with permission [64]. Copyright 2016, Elsevier.) and (c1,c2) NiO-SnO2 nanocomposites (p-n heterojunction) (Copied with permission [154]. Copyright 2016, Royal Society of Chemistry.) towards oxidising gases.
Figure 8. The enhanced gas sensing mechanisms of (a1,a2) the hollow nanofibres of ZnO-SnO2 (n-n heterojunction) (Copied with permission [151]. Copyright 2018, Elsevier.), (b1,b2) the ZnO/Co3O4 composite nanoparticle (n-p heterojunction) (Copied with permission [64]. Copyright 2016, Elsevier.) and (c1,c2) NiO-SnO2 nanocomposites (p-n heterojunction) (Copied with permission [154]. Copyright 2016, Royal Society of Chemistry.) towards oxidising gases.
Nanomaterials 11 01026 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, S.; Lei, G.; Xu, H.; Lan, Z.; Wang, Z.; Gu, H. Metal Oxide Based Heterojunctions for Gas Sensors: A Review. Nanomaterials 2021, 11, 1026. https://doi.org/10.3390/nano11041026

AMA Style

Yang S, Lei G, Xu H, Lan Z, Wang Z, Gu H. Metal Oxide Based Heterojunctions for Gas Sensors: A Review. Nanomaterials. 2021; 11(4):1026. https://doi.org/10.3390/nano11041026

Chicago/Turabian Style

Yang, Shulin, Gui Lei, Huoxi Xu, Zhigao Lan, Zhao Wang, and Haoshuang Gu. 2021. "Metal Oxide Based Heterojunctions for Gas Sensors: A Review" Nanomaterials 11, no. 4: 1026. https://doi.org/10.3390/nano11041026

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