Next Article in Journal
Graphene Oxide Layer-by-Layer Films for Sensors and Devices
Previous Article in Journal
Flexible Control of Two-Channel Transmission and Group Delay in an Optomechanical System with Double Quantum Dots Driven by External Field
Previous Article in Special Issue
Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 1: Nanofibers and Features of Their Forming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 2: Gas Sensors and Their Advantages and Limitations

by
Ghenadii Korotcenkov
Department of Theoretical Physics, Moldova State University, 2009 Chisinau, Moldova
Nanomaterials 2021, 11(6), 1555; https://doi.org/10.3390/nano11061555
Submission received: 7 May 2021 / Revised: 7 June 2021 / Accepted: 8 June 2021 / Published: 12 June 2021
(This article belongs to the Special Issue Nanostructured Gas Sensors)

Abstract

:
Electrospun metal oxide nanofibers, due to their unique structural and electrical properties, are now being considered as materials with great potential for gas sensor applications. This critical review attempts to assess the feasibility of these perspectives. This article discusses approaches to the manufacture of nanofiber-based gas sensors, as well as the results of analysis of the performances of these sensors. A detailed analysis of the disadvantages that can limit the use of electrospinning technology in the development of gas sensors is also presented in this article. It also proposes some approaches to solving problems that limit the use of nanofiber-based gas sensors. Finally, the summary provides an insight into the future prospects of electrospinning technology for the development of gas sensors aimed for the gas sensor market.

1. Introduction

In recent years, noticeable interest has been shown in one-dimensional (1D) metal oxide nanomaterials [1,2,3,4,5,6,7,8,9,10] such as nanowires, nanobelts and nanotubes. Many authors also include metal oxide nanofibers (NFs) in this group of nanomaterials [11,12,13,14], although in nature, they differ from classical 1D nanomaterials. Unlike classical 1D nanomaterials, which are monocrystalline in nature, metal oxide nanofibers are amorphous or polycrystalline. Earlier in [15,16,17,18,19,20,21,22,23], it was shown that one of the most promising fields of application of metal oxide 1D nanomaterials is the development of conductometric gas sensors based on them. As it is followed from discussions presented in Part 1 of our article [24], metal oxide nanofibers are also a promising material for these applications, since a nanofiber mat, forming a gas-sensitive layer, is characterized by high porosity and a large surface-to-volume ratio [12,25,26]. In addition, the metal oxide crystallites in nanofibers can have an extremely small size [27]. According to generally accepted concepts [28,29,30,31,32], it is these parameters that should be possessed by materials capable of providing high sensitivity and a fast response of gas sensors. However, it is not clear how justified these expectations are from the use of nanofibers in conductometric gas sensors. For example, in [23], it was shown that 1D and 2D nanomaterials, despite the numerous advantages attributed to them [15,16,17,18,19,20,21,22], still do not find application in the development of gas sensors intended for the market.
In Part 1 [24], we declared that the second part of our article will be devoted to the analysis of nanofiber-based gas sensors and their advantages and limitations. This is exactly what is carried out in this article. In Part 2 of this article, approaches to the fabrication of gas sensors are considered, as well as the results of analysis of the performances of nanofiber-based conductometric gas sensors. It then provides a detailed analysis of the drawbacks that may limit the use of electrospinning technology in the development of gas sensors. Some approaches to solving these problems are also suggested in this part. Finally, the summary provides an insight into the future prospects of electrospinning applications for the development of conductometric gas sensors.
It is important to note here that the nature of gas-sensitive effects in nanofiber-based metal oxide gas sensors and conventional metal oxide gas sensors is identical, and therefore this topic will not be considered in this review. The mechanism of gas sensitivity, as well as the role of surface phenomena and structural factors in the sensor response of conductometric gas sensors, has been considered in sufficient detail in numerous reviews and books [28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64]. Most gas sensing mechanisms are based on the chemisorption of the target gas on the metal oxide surface or its interaction with chemisorbed oxygen. These processes are accompanied by a change in the surface charge, which determines the surface potential and thickness of the depleted layer, and, consequently, the conductivity of the gas-sensitive layer.

2. Fabrication of Gas Sensors Based on Metal Oxide Nanofibers

As it was indicated before, nanofibers are not 1D structures, in the classical understanding, as metal oxide nanowires or nanobelts and nanotubes. However, research has shown that the features of the nanofiber configuration and the size factor play a positive role in the development of gas sensors based on nanofibers [12,26,65,66,67,68,69,70,71]. As mentioned above, metal oxide nanofibers, which usually have a diameter in the range of 50–1000 nm and a length from several micrometers to centimeters and meters, have many unique properties of gas-sensitive materials, such as a very large surface area per unit mass, high porosity and a small size of crystallites that form nanofibers. For example, gas-sensing materials formed from nanofibers usually have a porosity of ~70–90% [72]. The presence of large pores, along with small pores, that facilitate gas diffusion can also be attributed to the advantages of these materials. It is also important to note that, under equal conditions, the size of crystallites in a nanofiber will always be smaller (see Figure 1) than in a film formed by traditional technology [27]. Therefore, many developers believe that from this point of view, metal oxide nanofibers are an ideal candidate as sensing materials for gas sensors [12,25,73], since, unlike thin and thick films, gas-sensitive materials based on nanofibers can provide high gas permeability, even with the smallest crystallite size. Thin-film and thick-film technologies are deprived of this possibility, since with a decrease in the crystallite size, the pore size in the gas-sensitive material also decreases.

2.1. Sensors with Nanofiber Mat

The simplest method of manufacturing nanofiber-based sensors is shown in Figure 2. Generally, this process has three stages. First, the nanofibers are deposited on a fixed collector in a three-dimensional nonwoven membrane. Oxidized silicon wafers are usually used as a collector. Then, the deposited nanofibers are annealed, because of which an interconnected porous metal oxide structure is formed, and only then are metal contacts of various shapes, including interdigitated electrodes, applied to the formed mat of nanofibers [74,75,76]. Of course, the formation of a nanofiber membrane on a substrate with already applied electrodes is also possible [77].
As materials for the manufacture of electrodes, developers commonly use Au [78], Ag/Pd [80], Al [81], Ni [78], Ag [11], Ti [82] and Pt [79]. It would seem that the material of the electrodes should not affect the gas sensing characteristics. However, experiments carried out by Imran et al. [78], Wang et al. [81] and Batool et al. [82], who compared Au and Ni electrodes, Ag and Al electrodes and Ti, Au and Ni electrodes, respectively, showed that the electrode material can have a significant effect on the sensor parameters. Therefore, according to Imran et al. [78] and Wang et al. [81], sensors with Au and Ag electrodes exhibited improved sensing properties in comparison with sensors using Ni and Al electrodes. At the same time, according to Batool et al. [82], sensors with Au electrodes showed the worst characteristics in comparison with devices where Ti and Ni electrodes were used. Batool et al. [82] believe that such situation takes place due to the different porosities of electrode materials. According to Imran et al. [78], the work function of electrode materials plays a more important role. However, there are other factors that affect the performance of a sensor with different electrodes [28].
As for other experiences gained in the development and manufacture of nanofiber-based gas sensors, they can be formulated as follows:
  • When manufacturing gas sensors using electrospun nanofibers suspended as bridges between contact areas, it is necessary to keep in mind that the heating rate during sintering of the metal oxide nanofibers is one of the most important factors. For example, Camargo et al. [83] showed that for manufacturing ZnO nanofiber-based bridges, the calcination of electrospun fibers at 600 °C should be carried out at a very low heating rate of ~5 °C/min. At a higher heating rate, the nanofibers were destroyed (see Figure 3). Camargo et al. [83] suggested that it is likely that the forces and mechanics involved during sintering require a slow temperature change.
  • Formation of low-resistance contacts to nanofibers is a problem of manufacturing nanofiber-based gas sensors no less important than the formation of nanofibers with desired properties. Typically, contacts have increased resistance, which negatively affects the sensor performance. Camargo et al. [83], in order to form low-resistance contacts, used a focused ion beam (FIB)-assisted metal deposition. However, they themselves admit that FIB-assisted technology is a highly time-/money-demanding technique.
  • In the absence of a proven sensor manufacturing technology, the role of the contact configuration increases significantly. Raible et al. [84] found that the contact of fibers with the top or bottom electrodes dramatically alters the response characteristics. In other words, for a better electrical contact, metal electrodes are preferably applied after nanofiber formation.
It should be noted that the problem of low-resistance contacts is an important problem for all conductometric gas sensors. If attention is not paid to this, a situation can occur where, under some conditions, the properties of the sensors will be controlled by inter-crystalline barriers, and in others, the sensor response will be controlled by the properties of the metal–nanofiber contact. In other words, sensors made even of the same gas-sensitive material can have fundamentally different characteristics. This is precisely the situation observed by Moon et al. [85] in a study of electrospun TiO2 nanofibers. They found that if In, Ag and Au form a low-resistance contact with TiO2 nanofibers, then a sputtered platinum electrode in contact with TiO2 forms a pronounced Schottky barrier with a potential barrier height of ~1.5 eV. For comparison, the height of the potential barrier at the boundary of TiO2 crystallites is ~0.7 eV. This means that the resistance of the Pt-TiO2 barrier even at T~200 °C is 109 times greater than the resistance of inter-crystalline TiO2-TiO2 barriers. This is why Moon et al. [85] concluded that back-to-back Pt-TiO2 Schottky barriers, not TiO2-TiO2 grain boundaries, are responsible for the large resistance and conductometric response of TiO2 nanofiber-based sensors with Pt contacts to NO2. It is important to note that sensors with Pt contacts had the maximum sensitivity to NO2. Sensors with In and Au contacts had significantly lower sensor responses.

2.2. Single Nanofiber-Based Sensors

Studies have shown that, as in the case of nanowires discussed in [23], individual nanofibers can also be used to develop gas sensors. For example, Nikfarjam et al. [86] fabricated individual TiO2 nanofiber-based sensors for detecting CO at concentrations as low as 30 ppb, with a rapid response and recovery time at Toper = 250 °C (tres/trec = 3/4 s) (see Figure 4). There have been other attempts to develop single nanofiber-based gas sensors [83,84,87,88,89]. However, no significant results were obtained in these studies.
It is clear that, as in the case of 1D nanowires, the formation of sensors based on individual nanofibers is associated with great difficulties due to the need to align, move and fix them in certain places [23]. However, it was found that the technology for fabricating individual nanofiber-based sensors can be significantly facilitated if nanofibers are deposited aligned and oriented during electrospinning. Studies carried out by various teams have shown that such conditions can indeed be realized (see Figure 5d). Recently, several approaches have been proposed to control the alignment of electrospun nanofibers [90,91]. One of the most common ways is to use the highly rotating drum as a collector [92,93] (Figure 5a). The properties and orientation of electrospun nanofibers in this case strongly depend on the drum rotation speed. A rotating disk with an extremely sharp edge can also be used instead of a rotating drum [94,95] (Figure 5e). However, since the edge of such a disk must be relatively sharp, this method has significant limitations if it is necessary to form well-aligned nanofibers over large areas.
There are several other methods for aligning and ordering nanofibers by the electrospinning method [89,90,98,99]. For example, there is magnetic field-assisted electrospinning [100]. However, the most progressive is still the method proposed by Li et al. [101,102]. Li et al. [101,102] demonstrated that nanofibers can be uniaxially aligned by introducing insulating gaps into conductive collectors (Figure 6). The insulating gap can be air, quartz, polystyrene or any other insulating material. The gap width can vary from hundreds of micrometers to several centimeters. According to Li et al. [102], fibers are pulled perpendicular to the edges of the gap due to electrostatic forces from two sources: a strong external field (F1) between the spinneret and the collector and repulsion between adjacent charged fibers (F2). SEM images of ceramic nanofibers made by this method are shown in Figure 6c–f. It is seen that the proposed method makes it possible to form well-aligned arrays of ceramic nanofibers. Single nanofibers deposited in controlled directions were also formed using this method. Li et al. [103] also showed that by changing the configuration of the electrodes, various structured structures of electrospun nanofibers can be obtained. The interdigital electrode can also be used to form oriented nanofibers. Ke et al. [104] found that by varying the configurations of the interdigital electrode, it is possible to obtain parallel electrospun nanofibers of different lengths. The main advantage of the methods developed by Li et al. [101,102] and Ke et al. [104] is that these methods allow the direct integration of nanofibers with controlled configurations into the electrode system used in the manufacture of gas sensors. Undoubtedly, this approach can significantly simplify the technology of manufacturing devices based on individual nanofibers. This very method was used to fabricate TiO2 nanofiber-based gas sensors, which were considered earlier in [86]. The only difference from the previously discussed methods is the use of a secondary field (Figure 7a). SEM images of a single aligned nanofiber used for gas sensor fabrication are shown in Figure 7b–e.

3. Performances of Gas Sensors Based on Metal Oxide Nanofibers

3.1. General Consideration

To date, there have been many attempts (listed in Table 1) to construct ultrasensitive gas sensors to detect NH3, H2S, CO, NO2, O2, CO2 and vapors of organic compounds (VOCs), such as CH3OH, C2H5OH, C5H10Cl2, C6H5CH3 and C4H8O, with improved detection limits using nanofibrous membranes as sensing structures [12,27,87,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137]. For example, Choi et al. [138] developed ZnO nanofiber-based NO2 sensors and compared them to conventional thin-film ZnO-based gas sensors. They found that the response of nanofiber-based NO2 sensors was higher and faster. Zhang et al. [139] reached the same conclusion when analyzing WO3 nanofiber-based NO2 gas sensors, and Du et al. [140] after considering the parameters of In2O3-based NH3 sensors.
Figure 7. (a) Schematic representation of different fields during the electrospinning process. (be) FE-SEM images of (b) aligned pure single nanofiber on a triangular-type electrode, (c) aligned pure TiO2 single nanofiber on a rectangular-type electrode, (d) aligned fiber (before thermal treatment) on the tip of a triangular electrode and (e) TEM image of a pure TiO2 nanofiber. (a) Reprinted with permission from [86]. Copyrights 2017 ACS; (be) Adapted with permission from [141]. Copyright 2015 Elsevier.
Figure 7. (a) Schematic representation of different fields during the electrospinning process. (be) FE-SEM images of (b) aligned pure single nanofiber on a triangular-type electrode, (c) aligned pure TiO2 single nanofiber on a rectangular-type electrode, (d) aligned fiber (before thermal treatment) on the tip of a triangular electrode and (e) TEM image of a pure TiO2 nanofiber. (a) Reprinted with permission from [86]. Copyrights 2017 ACS; (be) Adapted with permission from [141]. Copyright 2015 Elsevier.
Nanomaterials 11 01555 g007
It is important to note that reduced response and recovery times are a great advantage of nanofiber sensors, as a slow response and recovery were one of the major disadvantages of conventional gas sensors. Undoubtedly, this situation is due to the large surface area and high porosity of sensitive materials based on nanofibers. The extremely high porosity is the main advantage of these sensors, which show very good operating characteristics (excellent and fast response) compared to conventional sensor materials. This unique morphology facilitates the efficient penetration of the target gas into the porous ceramic layer, which is believed to be the main reason for the exceptionally high gas sensitivity of metal oxide gas sensors manufactured by this method [142]. Unlike conventional thin- and thick-film technologies, which produce mesoporous granular layers with densely packed nanoparticles that cause poor gas transfer, electrospinning sensors exhibit a bimodal pore size distribution, including both small and large pores that enhance gas transport and improve the conductometric response of these layers [143].
The explanation of the gas sensing effect in metal oxide nanofibers can be carried out within the framework of the approaches developed for traditional gas sensors based on metal oxides [144]. Moreover, it was found that the gas sensing characteristics of sensors based on nanofibers made of metal oxides obey the same regularities as conventional metal oxide gas sensors discussed in [28,145,146]. This means that, as in the case of conventional gas sensors, nanofiber-based sensors with a smaller grain size, smaller crystallite size, higher porosity and larger surface area have the maximum conductometric response [36,147,148,149,150]. This conclusion is well illustrated in Figure 8. For example, Zhang et al. [139] showed that the maximum sensitivity and minimum response time were possessed by WO3 nanofiber-based sensors, which are characterized by the maximum surface area, minimum crystallite size and maximum pore diameter in nanofibers (see Table 2). A solution of ammonium metatungstate hydrate-PVP-deionized water was used for electrospinning. The WO3 nanofibers were obtained by annealing the nanofiber precursors at 550 °C for 2 h with different heating rates. To obtain nanofibers with different porosities, Zhang et al. [139] used annealing at different rates, varying from 1 to 15 °C/min.
Table 1. Metal oxide nanofiber-based conductometric gas sensors and their performances.
Table 1. Metal oxide nanofiber-based conductometric gas sensors and their performances.
MaterialDiameter, nmAnalyte GasC., ppmT, oCResponseRes/Rec TimeRef.
TiO2120–200NO250450302–4 min/20 s[151]
WO31000.4751233 min/38 min[152]
SnO2200–40050185368400 s/200 s[153]
SnO2300–50023008155 s/5 min[154]
CeO2380O21008001.430 s/-[155]
La0.67Sr0.33MnO3~126 1008001.153 s/-[156]
ZnO250H210350109-[157]
TiO280125275~612 s/22 s[141]
WO3200NH310020061 s/5 s[153]
ZnO95–130100200~20-[147]
TiO2400–500CO25200~432–86 s/84–109 s[158]
SnO2200–400500300~4260 s/15 min[143]
ZnO35–15022001.5168–237 s/270–350 s[159]
In2O3100100300~5-[160]
WO3275Acetone5027056 6–13 s/4–9 s[161]
ZnO1451220712–17 s/11–23 s[162]
In2O3250–31053001515 s/2 s[163]
α-Fe2O3150–280Ethanol1003002 3 s/5 s[164]
SnO210010330513 s/13.9 s[105]
In2O3160–20015003003791 s/5 s[165]
In2O330–10030220~46 s/10 s[166]
Co3O4100–200100301518–23 s/59–3 s[167]
ZnO500–600DMF100RT1332 s/17 s[168]
In2O3150–200100340318 s/17 s[169]
Table 2. Parameters of electrospun WO3 nanofibers, used for fabrication of NO2 gas sensors.
Table 2. Parameters of electrospun WO3 nanofibers, used for fabrication of NO2 gas sensors.
SampleHeating Rate, °C/minSurface Area, m2/gPore Size, nmCrystallite Size, nm
WO3-1110.711.226.5
WO3-5514.417.520.9
WO3 101016.430.616.8
WO3-151512.818.516.1
Source: data extracted from [139].
As it was shown earlier, the crystallite size in nanofibers is usually controlled via adjusting the calcination temperature and time [170,171]. The higher the temperature and the longer the annealing time, the larger the crystallite size (see Figure 9). That is why, usually, the maximum sensor effect is observed at the minimum calcination temperature (see Figure 9a). However, it should be borne in mind that the lower the annealing temperature and the smaller the crystallite size, the more pronounced the temporal and thermal instability of the sensor parameters [55]. Therefore, when choosing calcination modes, a compromise between sensitivity and stability has to be found. The crystallite size can also be controlled by the solution composition and polymer content [106,172]. As a rule, an increase in the concentration of a precursor in an electrospinning solution is accompanied by an increase in the crystallite size of metal oxides (read Part 1 [24]). The diameter of nanofibers also has a strong influence on the crystallite size. The larger the nanofiber diameter, the larger the size of the metal oxide crystallites that are formed during the calcination process. To achieve the minimum diameter of nanofibers, the recommendations proposed earlier can be used.

3.2. Hollow Nanofiber-Based Gas Sensors

The use of hollow fibers is another effective solution for improving the parameters of sensors, since this approach provides significant growth in the area of the active surface. In this case, the gas molecules can interact with the inner and outer surfaces of the nanofibers. For example, it was shown that TiO2 hollow fibers (HFs) exhibited a higher response to CO at room temperature compared with solid fibers. Reducing the diameter of nanofibers and the wall thickness of hollow nanofibers is also a method to improve sensor performance. This effect is due to the fact that a decrease in the diameter of NFs contributes to a decrease in both the size of crystallites formed in the fiber (increase in sensor response) and the time of gas diffusion into the fiber (decrease in response time). Really, it was established that In2O3 and ZnO NFs with a smaller diameter (≈50–100 nm) and thinner walls (≈10 nm) exhibited an enhanced conductometric response compared with larger-diameter NFs (≈500 nm) toward formaldehyde, CO and NO2 [173,174,175,176,177]. How important it is to provide gas access to the inner surface of hollow nanofibers was shown by experiments performed by Du et al. [140]. According to [140], broken hollow nanofibers demonstrated the highest sensitivity to NH3 gas, which was about 20 times higher than the sensitivity of In2O3 nanoparticles (see Figure 10b).
As shown in Figure 10b, the formation of hollow nanofibers due to an increase in the active surface area really contributes to an increase in the sensitivity of gas sensors [178,179,180]. However, it should be borne in mind that at a certain wall thickness of such fibers, it may be difficult to access the target gas at the inner surface of these hollow fibers. As a result, we encounter a situation in which we will either not observe an increase in sensitivity, or the sensor response will be too slow. The results of this effect of the wall thickness of hollow nanofibers are shown in Figure 11. Hollow SnO2 nanofibers were formed using SnO2 thin-film deposition on electrospun PAN nanofibers by plasma-enhanced atomic layer deposition (PEALD) [181]. SnO2 thin film-coated PAN nanofibers were annealed at 700 °C for 1 h to burn out the PAN template and crystallize the SnO2. It is seen that at wall thicknesses of more than 10 nm, a decrease in the sensor response and an increase in the response time to ethanol vapor are observed.
Thus, when developing hollow nanofibers intended for use in gas sensors, their wall thickness must be optimized to achieve the required parameters. The creation of pores in the walls of hollow nanofibers can also solve the problem of poor accessibility of the test gas to the inner surface of the hollow fibers. Du et al. [140] showed that pores could be created using special template deletion modes. Choi et al. [182] developed a different approach. To form macroporous hollow WO3 NFs, they proposed adding colloidal polystyrene (PS) particles in a solution containing a W precursor and PPV. Mineral oil was used to form the core of the nanofiber. Macropores in WO3 NFs were formed as a result of the removal of PS particles incorporated in the walls of hollow WO3 nanofibers. A schematic diagram of this process is shown in Figure 12.
Another approach to pore formation was proposed by Liang et al. [163]. They used a two-step method to form porous hollow NFs. They first synthesized hollow In2O3 NFs and then exposed them in 10% HNO3. As a result of this treatment, macropores appeared in the walls. Chattopadhyay et al. [183], while developing a technology for the formation of TiO2 nanofibers, found that the addition of a structure-directing agent (Pluronic F127—nonionic triblock co-polymer) to the electrospinning solution ultimately allows the formation of TiO2 nanofibers with 3D mesoporosity and a high surface area. The choice of this additive was due to the fact that this type of surfactant, having a high hydrophobic/hydrophilic ratio (>1.5), favors the formation of ordered cubic micellar aggregates in a water–ethanol mixture [184,185]. Naturally, the presence of readily accessible macro- and mesopores in nanofibers makes them more compatible for use in gas sensors.

3.3. Surface Modification or Surface Decoration of Metal Oxide Nanofibers

Decoration of nanofibers’ surface by noble metals such as Pd, Pt, Au, Ag and Ru is also a very common method for optimizing the sensor performance of nanofiber-based gas sensors [12]. As in conventional technology [186,187,188], noble metals as effective catalysts, through a decrease in the activation energy of gas chemisorption and its catalytic oxidation, can significantly improve the selectivity (Figure 13) and sensitivity (Figure 14) of sensors to a specific gas.
As shown in Table 3, noble metals are most commonly used when developing sensors for reducing gases and vapors of organic solvents. As a rule, the maximum effect is achieved with a concentration of noble metals not exceeding 1–4 wt.% [192,193,194,195]. For example, Hu et al. [195] observed the maximum optimization effect at a Pd concentration of 3 mol% in CeO2 nanofibers. A concentration of 2.3 wt.% was the optimal concentration when doping In2O3 with Pt to achieve the maximum sensor response to H2S [196]. ZnO nanofibers doped with Ag at a concentration of 1 mol% had a maximum conductivity response to ethanol [197]. A uniform distribution of highly dispersed noble metal nanoparticles on the surface of nanofibers is also an important condition for achieving a high sensitivity and good performance of the sensors being developed [198].

Mechanisms of Surface Modification Influence on Gas Sensor Performances

An explanation of the observed behavior of sensors after decorating the surface with noble metals can be found in [35,186,188,199,200,201,202,203,204,205,206,207,208]. Typically, these explanations suggest the presence of two sensitization mechanisms proposed by Morrison [34] and Yamazoe [29]. They are electronic sensitization and chemical sensitization.
Table 3. Electrospun-based metal oxide gas sensors modified with noble metals.
Table 3. Electrospun-based metal oxide gas sensors modified with noble metals.
MaterialDopantAnalyte gasC, ppmT, (°C)ResponseDetection LimitRes./Rec. Time, sRef.
SnO2PdAcetone100275991 ppm-[209]
SnO2Formaldehyde10016019-2/7[210]
In2O3Ethanol50200181 ppm1/10[130]
ZnOCO202205.51 ppm27/15[211]
TiO2NO22.1180380.16 ppm-[212]
WO3H2S13501.41 ppm-[213]
WO3Toluene13505.520 ppb119/16[213]
WO3PtAcetone23504120 ppb-[214]
α-Fe2O3H2S10175157- [215]
NiOEthanol100400121 ppm-[216]
In2O3H2S600200149050 ppm60/120[196]
SnO2H2S203005100--[192]
SnO2Toluene10300121 ppm-[217]
SnO2AgAcetone2001601175 ppm6/10[218]
TiO2H2S13501201 ppm-[219]
In2O3Formaldehyde50115285 ppm5/10[220]
SnO2AuCO10300191 ppm-[221]
SnO2CO530084-22/235[189]
In2O3Ethanol5001401450 ppm12/24[222]
WO3n-butanol1002502301 ppm5–43/10–122[223]
In chemical sensitization, the decorated noble metal acts as a promoter for the chemical interaction between the metal oxide and the analyte. The promoter increases the sensitivity to gas as it increases the rate of chemical processes, leading to a decrease in the concentration of negatively charged adsorbed oxygen [224,225]. As a rule, chemical sensitization is accompanied by direct and back-spillover effects [188,207]. Electronic sensitization is based on the existence of a potential barrier between a semiconductor and metal nanoparticles. Differences in the Fermi levels of the metal oxide and catalyst can lead to the formation of depletion/accumulation regions at the semiconductor near the metal nanoparticles. If, in the process of interaction with gas, a change in the oxidation state of the metal occurs, then this change leads to a change in the potential barrier at the metal–metal oxide interface, and therefore to a change in the conditions of current transfer in the gas-sensitive layer [224]. As a result, a sensor signal appears. It is believed that Au and Pt, when detecting reducing gases, affect sensor performance through a chemical sensitization mechanism, and Ag, Pd and Rh through an electronic sensitization mechanism. It is known that Ag, Pd and Rh have stable oxides Ag2O, PdO and Rh2O3 in air, which are easily reduced to metal in a reducing gas atmosphere and then easily reoxidized in an oxygen atmosphere [226]. However, the roles of sensitizers and sensitization mechanisms are not always clear. For example, Barbosa et al. [208] found that both electronic and chemical sensitization effects are relevant in Pt-decorated SnO2 devices. Degler et al. [207] also believe that the separation of electronic and chemical sensitization is useful for the first assessment of the role of additives, but for a complete understanding of reality, it is necessary to take into account the mutual influence of chemical and electronic properties and processes. Thus, both the electronic and chemical contributions to sensitization should be considered, rather than strictly distinguishing between electronic and chemical sensitization.
It is important to keep in mind that the activity of noble metals depends on many factors. These include parameters such as the coating thickness, cluster size, concentration of oxygen vacancies in the metal oxide support, interaction with the support, the method used to decorate the surface of metal oxides, the temperature of post-treatment and even the surface morphology and concentration of structural defects in the metal oxide itself [202,203,204,205,206,207,208]. For example, Katoch et al. [189] believe that the high sensitivity and selectivity of the response to CO of sensors based on electrospun SnO2 nanofibers decorated with Au were achieved due to the extremely small size of both the SnO2 crystallites and gold clusters, and the high density of Au nanoparticles on the surface of the SnO2 nanofibers (see Figure 13). It is also necessary to distinguish between the behavior of noble metals incorporated in the lattice of metal oxides and those on the surface in the form of clusters [203,207]. In particular, analyzing the characteristics of SnO2-based sensors doped with noble metals, such as Pt and Pd, Korotcenkov and Cho [203] concluded that the full incorporation of doping additives in the SnO2 lattice without forming metallic clusters on the SnO2 surface is the optimal condition to achieve improved performances of SnO2:Pt, Pt-based gas sensors.

3.4. Doping of Metal Oxide Nanofibers

As we noted above, the size of the crystallites formed in the nanofiber has a significant effect on the magnitude of the sensor response. How important this is can be judged from the results shown in Figure 15. It is seen that an increase in the crystallite size in a SnO2-CuO composite from 11 to 29 nm is accompanied by a decrease in the sensor response to H2S from 3·104 to ~40, i.e., almost 103 times.
In the traditional technology used for sensor fabrication, one of the methods for reducing the size of crystallites and stabilizing their size during heat treatment is doping of metal oxides [55,203]. Experiments have shown that this approach also works successfully in the manufacture of nanofiber-based sensors. For example, Zhao et al. [227], when doping α-Fe2O3 with Ca, observed a decrease in the crystallite size from 31 to 7 nm, with an increase in the Ca content in the range of 0–15 mol%. Shan et al. [228] found that Fe2O3 doping with La (5–10 wt.%) was accompanied by a decrease in the size of Fe2O3 crystallites from 14.7 to 8.6 nm. Mohanapriya et al. [229] reported that SnO2 doping with Ce (3–9 mol%) led, in addition to a decrease in the crystallite size and nanofiber diameter, to a significant increase in the area of the active surface (Table 4), i.e., optimization of those parameters that contribute to the growth of the sensor response. It is clear that not only the size of the crystallites is responsible for the growth of the sensor response. For example, Cheng et al. [230], investigating the effect of Fe2O3 doping with Eu in the range 0–5 wt.% on the sensor response to acetone, found that the maximum decrease in the crystallite size occurs at 5 wt.% Eu, while 3 wt.% Eu is the optimal concentration to achieve the maximum sensor response to acetone [230].
As a result of numerous studies, it was found that doping of metal oxide nanofibers by rare-earth metals (Yb, Sr, Ce, Pr, Er, Sm, La) [12,113,122,125,229,231,232] and transition metals (Fe, Y, Ni, Cu, Co, Mn) [12,233,234,235,236,237] is indeed a powerful tool for enhancing the sensor response (see Figure 16), and, in some cases, for improving selectivity. For example, by increasing the concentration of Ni in the range 0–10 atom% in SnO2 NFs, Cheng et al. [234] obtained a more than 5-fold increase in the sensor response to acetone. Zhao et al. [227], due to doping of α-Fe2O3 with Ca (7 mol%), also managed to increase the response to ethanol and acetone at 200 °C by almost five times. The increase in the sensor response to other gases was significantly less.
It is important to note here that the doping concentration providing the maximum optimization effect is not the same for all cases. It depends on the metal oxide used, the operation temperature and the gas to be detected. This effect is well demonstrated in Figure 17, which shows the effect of doping SnO2 with Ce on the sensor response to various gases at two temperatures [240]. It can be seen that, at low temperatures, the optimization effect of doping manifests itself in the detection of H2S, and the most effective is doping with Ce with a concentration of 3%, while at T = 370 °C, doping has a maximum effect on the sensitivity to ethanol, and this effect occurs at a doping concentration of 7%. This means that for each test gas and operating temperature, it is necessary to conduct independent research to select the optimal doping conditions.

Influencing Mechanism of Nanofiber Doping

Our understanding of the influencing mechanisms of the processes occurring in metal oxides during bulk doping on the effects of gas sensitivity is still insufficient for their detailed description. The influence of doping on the structural, electrophysical and gas sensing properties of metal oxides is too multifactorial. It was found that doping affects all parameters of metal oxides on which the magnitude of the sensor response depends [28,186]. Therefore, the choice of dopant and its concentration must be approached very carefully.
First, when choosing the doping element, it must be kept in mind that some additives have donor or acceptor properties that affect the concentration of charge carriers, and hence the Debye length.
Second, other additives are amphoteric impurities that have a major effect on the sensor response through changes in the crystallite size and the structure of the gas-sensitive layer.
Third, additives exhibit increased catalytic activity, while, fourth, dopants, when introduced into the metal oxide, can lead to an increase in porosity and better gas permeability.
Fifth, doping with the doping element is accompanied by the generation of structural defects that improve the adsorption properties of the surface. For instance, Cr ions act as an acceptor in TiO2, decreasing the electron concentration, but Nb ions act as a donor in TiO2, increasing the electron concentration.
It should not be forgotten that some impurities hinder the growth of crystallites during heat treatment, while others stimulate this growth, as occurs when doping TiO2 with Ta and Vo, respectively [55,60].
When choosing the concentration of the dopant, one should clearly understand what structure of the metal oxide matrix is to be worked with and what consequences for the gas-sensitive material and the sensor may arise if an additional phase is introduced into the gas-sensitive material. Depending on the concentration and the type of introduced additional components, they can form doped metal oxide or solid solutions based on the main phase. They can also create a segregation of the doping element in the form of metal or metal oxide clusters on the surface of the crystallites of the main phase or form a mixture of the crystallites of two oxide phases. In the latter case, heterojunctions are formed in the gas sensing matrix. If we consider the structure of the metal oxide matrix formed by two metals, MeI and MeII, then, depending on the concentration of these metals in the composite, one can identify seven areas with radically different properties. A diagram illustrating the appearance of these areas is shown in Figure 18.
Unfortunately, we cannot say what dopant and what structure of the gas sensing matrix from those shown in Figure 18 are optimal for the gas-sensitive effects, since with appropriate optimization, it is possible to improve one or more sensor parameters when doped with a wide variety of additives. Undoubtedly, the criteria for the selection of doping additives proposed by Rumyantseva and Gaskov [51] can be used (see Figure 19).
Rumyantseva and Gaskov [51] believe that the interaction of a semiconducting oxide with the gas phase is described through the formation of the surface complexes, and the decisive role in this process belongs to the chemical nature of the modifier and its reactivity in acid–base or redox reactions. However, this approach does not explain all gas-sensitive effects observed when using nanocomposites based on doped metal oxides. Moreover, often the same gas sensing effect can be achieved using additives with radically different physicochemical properties. Undoubtedly, when developing gas sensors based on doped metal oxides, we can also follow the recommendations presented in Table 5 [60]. However, these recommendations are too general and do not reflect the specificity of the interaction of the composite with a specific gas. Based on the available information, it can be stated that in most cases, the maximum conductivity response of gas sensors based on the doped metal oxide is observed at a concentration of the second metal near its solubility limit in the basic oxides [60,203,242,243]. This information is presented in Table 6. Other useful information regarding the features of the effect of metal oxide doping on their gas-sensitive properties can be found in [199,200,207,244,245,246].

3.5. Heterostructures and Core–Shell Structures in Nanofiber-Based Gas Sensors

Using conventional technology, in a number of cases, it was possible to significantly improve the parameters of sensors due to the formation of heterostructures and the synthesis of core–shell structures [60]. The same approach has been tried with electrospinning technology [12]. As a result, various heterostructures, such as ZnO–SnO2, CuO–SnO2, CuO–TiO2, In2O3–CeO2 and Al2O3–In2O3 [247,248,249,250,251,252,253], and core–shell structures, such as ZnO–SnO2, SnO2–In2O3, ZnO–TiO2, Fe2O3–NiO, CuO–TiO2 and Co3O4–Fe2O3 [254,255,256,257,258,259,260], have been synthesized. It is important to note that such structures can be created both in the electrospinning process and using the principles of post-treatments. For example, Lu et al. [251] synthesized nanofibers of a ZnO–SnO2 composite by electrospinning. At the same time, Qi et al. [261] fabricated In2O3–SnO2 heterostructures by dipping the electrospun In2O3 NFs in the Sn(OH)4 sol solution. The same approach was used to prepare TiO2 nanofibers decorated with WO3 nanoparticles [262]. Subsequently, nanofibers with the specified structure were used in the manufacture of gas sensors (see Table 7).
Testing has shown that, under certain conditions, heterostructures and core–shell structures do improve gas sensor performances. For example, ZnO–SnO2 composite HFs, synthesized by Wan et al. [248], exhibited high sensitivity to ethanol at 260 °C, with a fast response and recovery (τres= 4–7 s, τrec = 4–5 s). Additionally, most importantly, these ZnO–SnO2 hollow NF-based sensors showed excellent selectivity to ethanol as compared with acetone, ammonia, glacial acetic acid, DMF and formaldehyde [248]. This is an important advantage of such sensors, since one of the problems with ethanol gas sensors is their similar sensitivity to acetone.
Table 6. Solubility limits for metals in metal oxides most promising for gas sensor applications.
Table 6. Solubility limits for metals in metal oxides most promising for gas sensor applications.
Metal OxideAdditionSolubility LimitRef.
SnO2In4–10%[263,264]
Mn~5–6%[265]
Fe; Ni; V; Mo<5%[241,266,267,268,269,270]
Nb~3%[271]
Co; Cr~0.5–3%[272,273,274]
Cu; Al<1%[243,275]
Si1%[276]
In2O3Fe~20%[277,278]
Ga10–12%[279]
Sn~8%[263,280,281]
Nb; Mo1–3%[282,283,284]
Co~1%[285]
Cu<<1%[286,287]
ZnOCo, Mn13–30%[288,289]
Fe2–20%[288,290]
V3–15%[289,290]
Sn4–8%[291]
Cr~6%[290]
Ni, Ti~3%[288,290]
Al0.3–2.0%[292,293]
In<1%[294]
Ga0.5%[293]
Cu<0.2%[290]
Source: data extracted from [60].
Feng et al. [252] showed that In2O3–WO3 heterojunction NFs had an increased sensitivity to acetone compared to pure WO3 NFs. In2O3–SnO2 heterojunction NFs were highly sensitive to NH3 [261] and CO [253]. It is only important to know that an improvement in gas sensor performances is observed only under optimal conditions for the formation of heterostructures. For example, as seen in Figure 20, an increase in the sensor response of In2O3–SnO2 heterostructures to NH3 was observed only under the condition that the SnO2 concentration in the heterostructure was 16 at.%.
Similarly, CuO–SnO2 NFs exhibited a very high response toward H2S compared with pure SnO2 NFs [249]. As with conventional metal oxide-based sensors, the high sensitivity to H2S of CuO-based heterostructures, such as CuO–SnO2 or CuO–In2O3, is connected with the phase transformation of CuO, a p-type semiconductor, in CuS with metallic characteristics when interacting with H2S gas. The result is a significant change in the device structure, from a p–n heterostructure to a metal–semiconductor configuration. The opposite process takes place in an oxygen atmosphere.
Core–shell α-Fe2O3/NiO nanofibers synthesized by Cao et al. [255], core–shell CuO/TiO2 nanofibers synthesized by Deng et al. [250] and core–shell In2O3/SnO2 nanofibers synthesized by Wan et al. [259] exhibited significantly improved conductivity responses to formaldehyde and selectivity performances in comparison with NiO hollow nanofibers, and α-Fe2O3, CuO, TiO2, In2O3 and SnO2 nanofibers. For example, at an operation temperature of 240 °C, Fe2O3–NiO-based sensors had fast response–recovery behavior (~2 s and ~9 s) (see Figure 21). Sensors based on α-Fe2O3/TiO2 core–shell structures had a higher response and better selectivity to a low concentration of TMA at 250 °C in comparison with pristine α-Fe2O3 and TiO2 sensors [260].
Table 7. Some sensing properties of composite and heterostructure-based electrospun metal oxide NFs.
Table 7. Some sensing properties of composite and heterostructure-based electrospun metal oxide NFs.
MaterialGasConc. (ppm)T, (°C)Response (Ra/Rg)Ref.
TiO2–ZnOO210,00030020[295]
ZnO–rGONO25400119[79]
p-In2O3–TiO2NOx972540[296]
Al2O3–In2O39725100[254]
CuO–In2O3H2S5RT9170[297]
CuO-SnO21030025799[170]
ZnO–CuO101504490[298]
SnO2–CeO22021090[240]
CuO–ZnOCO0.13007[299]
SnO2–RGO120010[300]
TiO2–ZnO0.137515[301]
SnO2–ZnO1035011[302]
SnO2–MWCNT50251.3[303]
p-NiO–n-SnO2H210032013[304]
In2-xNixO3C2H5OH10018080[305]
Cr2O3–ZnO10030024[116]
In2O3–ZnO10021025[306]
ZnO–In2O3–ZnO10021017[306]
Sn-SnO2–Carbon50024030[307]
ZnO-TiO250032051[258]
SnO2–ZnOCH3OH103508.5[308]
In2O3–WO3C3H6O0.83501.8[252]
SnO2–α-Fe2O310034031[309]
In2O3–WO30.42751.3[252]
α-Fe2O3–NiOCH2O5024013[255]
NiO–SnO2102006.3[310]
SnO2–In2O30.53752.2[311]
SnO2–In2O350300115[257]
PPy–WO3NH32010026[312]
SnO2–In2O312521[261]
PANI–TiO20.025250.4[313]
p-La0.67Sr0.33MnO3–n-CeO2C3H82080075[238]
NiO-SnO2C7H85033011[314]
La0.7Sr0.3FeO3–In2O3-SnO2C3H9N1808[315]
p-NiO–n-ZnO100260892[316]
ZnO–In2O35375119[317]
C3H6O—acetone; C2H5OH—ethanol; CH2O—formaldehyde; C3H8—propane; CH3OH—methanol; C7H8—toluene; C3H9N—trimethylamine; PPy—polypyrrole; PANI—polyaniline; rGO—reduced graphene oxide.
In accordance with the conclusions made in [65,318,319,320,321,322,323], the main reasons for improving the parameters of sensors based on heterostructures and core–shell structures are the presence of an interface between two dissimilar materials and the combination of these heterojunctions into a network. The creation of a close electrical contact at the interface between these two components, which facilitates the equilibration of Fermi levels at the interface, usually leads to charge transfer and further expansion of the charge depletion region in contacting crystallites [318,324]. Another important factor to consider for heterostructures is synergistic behavior [325,326]. When two different components in a material are in contact with the gas phase and each exhibits its own specificity when interacting with a gas, then a situation may arise where the synergistic effect of the two-component system can be greater than the effect observed in these elements separately. These two factors are the basis of unique effects that can lead to an improved performance of gas sensors based on heterostructure-based materials.
However, prediction and analysis of the gas-sensitive characteristics of sensors based on heterostructures are difficult tasks, since, depending on the properties of the contacting materials, their ratio and methods for preparing heterostructures, radically different scenarios can be realized. In the case of mixed metal oxides, either of the two metal oxides can predominate in conductivity if they offer a conduction path that minimizes electrical resistance. It is also possible that conduction occurs through both metal oxides, and thus the charge transfer must cross heterojunctions formed at grain boundaries between the two materials [58,318].
In the case of doped and loaded metal oxides, conduction occurs exclusively through the supporting metal oxide, and the sensitization effect arises from the chemical and electrical interaction between the supporting metal oxide and the additive that form the heterostructure [207]. In addition, depending on the properties of the contacting materials, both nn and pn heterojunctions can be formed, with their own specific interaction with the test gas.
Processing routes also have a strong influence on the electrophysical and gas sensing characteristics. As it was indicated above, heterostructures can be prepared directly during electrospinning or by loading a second metal oxide on the surface of an already electrospun metal oxide. Although two materials in these two formed heterostructures can have the same nominal composition, the behavior of these heterostructures can be radically different [60]. In core–shell structures, the influence of the thickness and gas permeability of shell layers is also added. Therefore, despite the large variety of heterojunction-based sensors formed and tested, there is still no clarity in the understanding of the complicated gas sensing mechanisms in such devices. As a result, it is difficult to predict the gas-sensitive effect that may occur during the formation of heterostructures and core–shell structures.

3.6. Post-Treatments of Nanofibers

It is important that, in addition to influencing the structure and composition of nanofibers, various post-modification methods can be used to control gas-sensitive properties. For instance, Du et al. [327] proposed exposing the formed In2O3 nanofibers to RF low-temperature oxygen plasma (f = 13.6 MHz, P = 450 W, t = 30 min). They found that this treatment of nanofibers was accompanied by an increase in the sensor response to acetone. Treatment in oxygen plasma, leading to an increase in the concentration of chemisorbed oxygen on the surface of metal oxides, should indeed promote an increase in the sensor response to reducing gases. However, this state of the surface cannot be stable, especially when interacting with reducing gases. At the same time, Du et al. [327] claimed that the sensors have stable parameters. This means that parameter optimization is of a different nature. Really, Du et al. [327] found that treatment in oxygen plasma has a significant effect on the structure of nanofibers, which manifests itself in a decrease in the size of In2O3 crystallites with an increase in the diameter of nanofibers, leading to an increase in their porosity (Table 8). Apparently, exactly these changes are responsible for the observed increase in the sensor response. Du et al. [327] explained this effect by etching the surface of In2O3 crystallites with oxygen plasma. However, it is not clear how etching can be accompanied by an increase in the diameter of nanofibers.
Kim et al. [328] believe that high-energy electron-beam irradiation (E = 1 MEv, 50–150 kGy) of prepared metal oxide nanofibers can also be used to improve the sensor performance of nanofiber-based devices. They established that the response of a sensor based on ZnO nanofibers to 10 ppm H2 increased after electron-beam irradiation at a dose of 100–150 kGy. However, it must be admitted that the increase in the sensor response was insignificant (see Figure 22). Kim et al. [328] assumed that this improvement took place due to surface and structural defects generated by e-beam irradiation. This manifested itself in an increase in the surface areas of the samples and a decrease in the sensor resistance after electron-beam irradiation.
Nikfarjam and Salehifar [141] found that UV irradiation of sensors during gas detection, as in the case of conventional metal oxide gas sensors, has a significant optimization effect (see Figure 23). For example, by UV irradiation (λ = 390–410 nm), the response of a TiO2 nanofiber-based sensor to H2 was increased 10-fold, and response and recovery times were reduced by three–six times (Figure 24). For TiO2 sensors modified with Au, the optimization effect was even greater. For CO, the response of Au/TiO2 sensors under the influence of UV irradiation increased by about 20 times compared to dark conditions. In addition, the operating temperature was reduced from about 290 to 170 °C. The optimization effect of UV irradiation was also observed in the detection of formaldehyde by SnO2/TiO2 [329] and SnO2/ZnO heterostructure-based sensors [330].
The mechanism of the UV irradiation influence on sensor characteristics is described in sufficient detail in [187,331,332,333,334,335,336]. In [52,187,337,338], it was suggested that the absorbed photons modulate the receptor function of the MOX by (1) excitation of the solid (in other words, enhancement of the concentration of charge carriers in the solid), (2) the formation of highly reactive surface radicals, (3) a change in both the surface density of adsorption sites of various types and the surface coverage by the gas-adsorbed species and (4) the filling of defects with charge carriers. For example, UV photoactivation can promote oxygen desorption and increase the population of target gas molecules on the metal oxide surface [339].
It is important to note that light activation is most effective at low operating temperatures of sensors, and when the concentration of free electrons in metal oxide nanostructures is low, active oxygen species are formed with difficulty, and the dynamics of gas desorption is rather slow. That is why metal oxide gas sensors operated in the dark generally exhibited poor sensitivity and long response/recovery times at room temperature. According to Comini et al. [338], photoexcitation processes can also modulate the charge transport across the grain boundaries by: (i) increasing the concentration of free charge carriers throughout the material due to the electrons remaining in the conduction band; (ii) decreasing the barrier height at the grain contacts due to the variation in the interface charge; or (iii) by increasing the probability of charge carrier tunneling through the inter-grain barriers by decreasing the depletion layer widths in the adjacent grains. However, the consistent mechanisms responsible for the photo-assisted gas sensing properties of metal oxides still need to be thoroughly studied.
Wang et al. [335] showed that a decrease in crystallite size, an increase in material porosity and the formation of heterointerfaces enhance the effect of UV irradiation on the sensor response to test gases. Considering the characteristics of metal oxide nanofibers, it can be concluded that UV photoactivation can indeed be an effective method of improving the performances of nanofiber-based sensors operated at room temperature, increasing the sensor signal and decreasing the time constants of the sensor response.

3.7. Stability of Nanofiber-Based Gas and Humidity Sensors

As for the stability of the nanofiber-based sensors being developed, the studies carried out in this area have shown that gas sensors within the tested time, 1–6 months, have an acceptable stability of sensor parameters. As it is seen in Figure 25, during the test, the sensors exhibit almost constant sensor signals. This is quite understandable, since nanofiber-based sensors, by their nature, do not differ in any way from metal oxide gas sensors that have been on the market for a long time and have confirmed their high stability.
However, in sensors in which chemical reactions occur during the detection process, changes in the morphology of the gas-sensitive layer (nanofiber mat) and the sensor response during the operation of the sensors are possible. Seitz et al. [341], investigating the behavior of CuO nanofiber-based H2S sensors, found that due to phase transformations in the interaction of CuO with H2S (CuO ↔ CuS), these sensors undergo a dramatic morphological change during their life time (see Figure 26, A → D). Due to the high mobility of ions, fibers tend to break apart and grow to bigger agglomerates with a small area of the active surface. Interestingly, percolation-induced sensing still works for these structures but results in signals with a reduced signal-to-noise ratio.
Seitz et al. [341] believe that improving the stability of the morphology of CuO-based H2S sensors, and hence the stability of their parameters, is possible through generating composite fibers with other oxides which do not undergo a chemical reaction with H2S. This approach was used in the development of H2S sensors based on In2O3 [297], SnO2 [131,170] and ZnO nanofibers [298] modified with CuO. Unlike CuO, these metal oxides do not chemically react with H2S and are stable in a H2S atmosphere. For example, Katoch et al. [299] reported that SnO2–CuO-based H2S sensors were stable for more than six months and showed slight deviations in the parameters of various fabricated samples.
Various methods can be used to modify the surface of metal oxides with CuO clusters. For instance, Katoch et al. [170] synthesized CuO–SnO2 composite nanofibers directly in the electrospinning process. For these purposes, they used a solution containing dehydrate (SnCl2·2H2O), copper chloride dihydrate (CuCl2·2H2O), DMF and ethanol as solvents, and acetate (PVAc) as a polymer. Yang et al. [131], in order to obtain CuO-modified hollow SnO2 nanofibers, immersed SnO2 hollow nanofibers into a Cu(NO3)·3H2O aqueous solution for 6 h at 95 °C. After that, SnO2 nanofibers, modified with CuO, were washed with deionized water and ethanol and dried at 70 °C for 24 h. As in CuO-based sensors, the conductometric response is determined by the reaction of the interaction between CuO and H2S, accompanied by a transformation from p-CuO to metallic CuS. However, unlike CuO-based sensors, these changes occur only in CuO clusters. After exposure to interference gases such as CO, NO2, acetone and alcohol, the reaction of the transformation from p-CuO to metallic CuS does not occur [131]. That is why these sensors have such a high selectivity for detecting H2S. For example, for sensors developed by Liang et al. [297], the ratio SH2S/Sgas reaches ~105 at Toper = 150 °C and ~102 at Toper = 300 °C (see Figure 27).
As it can be seen, the selectivity is much higher at 150 °C. However, for such low temperatures, the sensor resistance did not recover to its original value even after exposure to an air atmosphere for 100 s. The same situation was observed by Kapse et al. [342], who also reported the sluggish or incomplete recovery from H2S sensing at T < 200 °C for In2O3–CuO-based sensors. It is believed that the difficult desorption of SO2 is the cause of this phenomenon. This problem, as Liang et al. [297] suggested, can be solved by refreshing the sensor surface via pulse heating to 500 °C. Using this approach, Liang et al. [297] succeeded in achieving fully reversible gas sensing characteristics and in shortening the recovery time to <140 s.

4. Limitations of Electrospinning for Gas Sensor Design and Approaches to Resolving These Problems

There is no doubt that electrospinning is a powerful method for producing a variety of nanostructured materials and highly sensitive gas sensors. However, despite the successes achieved, essential studies are still required in this area, and many challenges have to be faced. This is due to the fact that in addition to all the advantages listed above, the electrospinning process has some limitations [343,344].
The following disadvantages of electrospinning technology are most commonly indicated:
  • First, according to Mondal and Sharma [344], some of the disadvantages of electrospinning technology are (a) the need to use a templating carrier polymer, since direct electrospinning is not possible for all metal oxides, and (b) the limited number of polymers that can be used for the production of metal oxide nanofibers by electrospinning.
However, it must be admitted that the above cannot be attributed to disadvantages, since (a) the use of a carrier polymer makes it possible to reduce the diameter of the formed nanofibers, which means reducing the size of crystallites and thereby improving the parameters of gas sensors; and (b) the electrophysical properties of polymers do not have any effect on the properties of metal oxide fibers, since the polymers are removed after electrospinning. It is also important that the removal of polymers is well combined with the temperature conditions of the processing used in the calcination of deposited metal oxide nanofibers.
  • Second, the variety of applications and performances of electrospun metal oxide nanofibers is limited due to their brittleness after calcination [345]. In particular, electrospun metal oxide nanofibers after calcination cannot be used in the development of sensors based on flexible substrates. However, conventional metal oxide conductometric gas sensors have the same limitation.
  • Third, nanofibers have poor adhesion to the substrate. Electrospun fibers are also characterized by the poor interfacial adhesion properties between the nanofibers. It is known that the mat of electrospun nanofibers consists of fibers with a weak inter-fiber interaction. As a result, such a network of nanofibers has reduced mechanical properties and a high contact resistance.
Experiments have shown that an additional hot pressing step after the polymer–metal oxide fiber has been deposited, but before calcination, makes it possible to solve the problem of poor adhesion between the fibers in the mat, as well as between the mat and the substrate [142,143,346]. Hot pressing also helped to reduce the resistance of nanofiber–metal electrode contacts [142,143]. This was due to an increase in the contact area. However, besides improving adhesion, this treatment had an effect on the microstructure of the fibers, as shown in Figure 28 for TiO2-based fibers. By introducing the hot pressing step prior to calcination, an interconnected morphology of the TiO2/polymer composite fibers was obtained, as illustrated in Figure 28b, due to the partial melting of the polymer vehicle. Subsequent calcination resulted in the structures shown in Figure 28d. As it can be seen, strong hot pressing makes the structure of the gas-sensitive layer denser in comparison with the original structure, bringing it closer in its properties to the structure of films formed by conventional thick-film technology.
The photochemical activation proposed by Meng et al. [347] also contributed to the improvement in adhesion of metal oxide nanofibers. Using this approach, Meng et al. [347] manufactured high-performance field-effect transistors based on In2O3 nanofibers. The UV treatment enabled the stable adhesion of the nanofiber network and the formation of a clean interface. Due to their improved adhesion properties, field-effect transistors have demonstrated improved device uniformity and efficient modulation of electrical characteristics. Apparently, this approach can also be used in the manufacture of gas sensors.
Figure 28. Electrospinning and hot pressing of metal oxide materials: (a) SEM image of as-spun TiO2/PVAc composite fibers fabricated by electrospinning from a DMF solution; (b) SEM image of TiO2 PVAc composite fibers after hot pressing at 120 °C for 10 min; (c) SEM image of unpressed TiO2 nanofibers after calcination at 450 °C; (d) SEM images with different magnifications of hot-pressed TiO2 nanofibers after calcination at 450 °C. Reprinted with permission from [348]. Copyright 2006 American Chemical Society.
Figure 28. Electrospinning and hot pressing of metal oxide materials: (a) SEM image of as-spun TiO2/PVAc composite fibers fabricated by electrospinning from a DMF solution; (b) SEM image of TiO2 PVAc composite fibers after hot pressing at 120 °C for 10 min; (c) SEM image of unpressed TiO2 nanofibers after calcination at 450 °C; (d) SEM images with different magnifications of hot-pressed TiO2 nanofibers after calcination at 450 °C. Reprinted with permission from [348]. Copyright 2006 American Chemical Society.
Nanomaterials 11 01555 g028
According to Seitz et al. [341], improved adhesion of the fibers to the substrate is also obtained by pretreating the substrate surface in PVP solution in EtOH. They discovered this effect when forming a CuO nanofiber mat using electrospinning of the solution, containing 7 wt.% PAN and 10 wt.% Cu-2-ethylhexanoate.
Cui et al. [349] showed that a simple and efficient nanoscale welding technology can also be used to assemble metal oxide nanofibers into a large area jointed NF network with improved inter-fiber bonds and strong interfacial adhesion properties (Figure 29). They believe that the in situ crosslinking process has several advantages over the conventional solvent vapor welding or hot pressing process. This process is more efficient, much simpler and does not require expensive reagents or equipment. Cui et al. [349] proposed using an amine-hardened epoxy resin as an adhesion agent for fabrication of an In2O3 nanofiber-based network. According to Cui et al. [349], the crosslinking welding process is controlled by the spontaneous chemical reaction between polymer chains, rather than high-energy radiation. Therefore, its application is not restricted to the limited area and flatness of the substrate. It is important to note that this technology can be effective in the development of gas sensors, since it better preserves the nanofiber mat structure, which is optimal for gas sensor applications.
  • Fourth, to date, it remains a challenge to produce nanofibers with diameters smaller than 20–50 nm by the existing electrospinning technique.
It has been found that both the response time and sensitivity improve significantly with a decreasing nanofiber diameter [237]. This effect occurs due to the fact that the diameter of the nanofibers affects the size of the crystallites formed in the nanofiber during calcination. The smaller the diameter, the smaller the crystallite size. As it is known, the size of crystallites is one of the main factors that control the response of conductometric gas sensors [28,36,55]. In addition, a smaller diameter results in a faster response due to the faster diffusion of gas molecules through the nanofiber. However, an electrospun fiber of the conventional process typically has a diameter in the range of 100–500 nm. Therefore, numerous studies focused on the development of a reliable method for producing nanofibers with an extremely small diameter in large quantities and with a uniform size. However, it is quite possible that such thin nanofibers (<20 nm) will not be needed. For example, Vuong et al. [27], investigating the gas-sensitive characteristics of WO3 nanofibers, found that the maximum sensor response to NO at 300 °C was observed with a nanofiber diameter of ~40 nm (see Figure 30a). A sharp increase in the resistance of nanofibers corresponded to this diameter (Figure 30b).
As a rule, the maximum sensor response correlates well with the ratio of the crystallite size and the Debye screening length [30,32,54]. This means that with a corresponding influence on the concentration of charge carriers in crystallites through doping of metal oxides, the maximum sensor response can be achieved even with a larger crystallite size, and therefore with a larger diameter of nanofibers. For example, Zhang et al. [350] showed that doping with Mg can be used to reduce the carrier concentration in In2O3 crystallites formed by electrospun nanofibers.
  • Fifth, electrospinning does not provide highly reproducible nanofiber parameters such as the nanofiber diameter and the response of sensors based on these nanofibers [25,351].
As it was shown earlier, too many factors affect the parameters of the formed metal oxide nanofibers [352,353]. This means that the control of all technological parameters is a necessary step in the manufacture of a gas sensor to achieve an acceptable reproducibility of the sensor characteristics. Much has been conducted in recent years to improve the reproducibility of nanofiber parameters. For example, Demir et al. [354] proposed increasing the temperature of the solution. They found that the diameters of the fibers obtained from the polymer solution at high (70 °C) temperature were much more uniform than the diameters obtained at room temperature. However, a solution to the problem of the low reproducibility of nanofiber parameters that satisfies all the requirements has not yet been found.
  • Sixth, the complexity of fabrication also creates certain difficulties in using electrospinning technology to form gas-sensitive layers on the sensor platform. This also includes the impossibility of forming gas-sensitive layers localized in certain places of the platforms of gas sensors by this method.
As it follows from the description of electrospinning technology, it is really difficult to spin the fibers in a small sensing area of the microsensor. This method is indeed more suited for the formation of a nanofiber mat on areas that significantly exceed the traditional dimensions of gas sensors, especially those manufactured in microelectronic design. For example, Kim et al. [348] fabricated TiO2 nanofiber-based gas sensors using a platform similar in appearance to that often used in the development of conventional thick-film gas sensors (see Figure 31). However, its dimensions of 10 × 15 mm are much larger than the platforms used in conventional gas sensors. The other result could not be expected, since with a standard distance between the electrode and the collector of 10–15 cm, the deposition area cannot be less than 10 cm2.
Zhang et al. [105], analyzing the reasons for these limitations, came to the conclusion that this problem can be solved only by reducing the distance between the spinneret and the collector. It was found that by reducing the distance between the collector and spinneret to 500 μm–5 cm, so-called near-field electrospinning can be realized [355,356,357]. This process, due to the absence of bending instability, allows the application of fibers with a high spatial resolution, by matching the average motion speed (SJ) of the jet with the relative speed (SR) between the collector and the spinneret. Compared with far-field electrospinning, near-field electrospinning offers a number of advantages [352]. They are as follows:
(i)
Substantial reduction in the applied voltage;
(ii)
The ability to accurately position the fibers over a relatively large area with minimal material consumption;
(iii)
The ability to manipulate the spatial positions of the fibers along all three directions, X, Y and Z, for printing of fibers [358,359,360].
Zhang et al. [105], by reducing the distance between the spinneret and the collector to 2.6, 2.1 and 1.7 cm, made it possible to reduce the deposition areas to 4, 2.8 and 1.7 cm2, respectively (see Figure 32). However, these deposition areas were still much larger than the sensing area of the microhotplate. Only when this distance was decreased to 5 mm did the deposition area decrease to about 2 mm2, which was already comparable to the active area of the sensor made on a platform with a micro-heater. Simultaneously with the decrease in the distance, the applied voltage also decreased, which at the minimum distance was 5 kV. Zhang et al. [105] reported that, as a result of the optimization of the electrospinning process of the PVA/SnCl4·5H2O solution, they were able to obtain SnO2 nanofibers with an average diameter of ~100 nm (see Figure 33) and construct, on their basis, a sensor with a large response to 10 ppm ethanol (~4.5), a low detection limit (10 ppb) and fast response/recovery processes (t < 14 s). This is a good achievement; however, the proposed approach was not further developed in the development of gas sensors. In addition, the fibers, formed using near-field methods, are generally much thicker than far-field electrospinning fibers, and the complexity of the device limits their use for mass production [358,359,360]; however, using an automated X-Y translational motion stage combined with near-field electrospinning, it would be possible to develop a technology for the formation of gas-sensitive layers similar to injection printing. For other purposes, this approach has already been implemented in [361,362,363]. However, it must be recognized that the performance of such a technology would be incomparable with the capabilities of injection printing.
There are also other approaches to the patterning of nanofibers in the process of forming a nanofiber mat [352], such as direct writing using melt electrospinning [364], selective photo-crosslinking of electrospun nanofibers containing a photoinitiator to generate a patterned mat [365] and localized removal of nanofibers using a laser, UV or a solvent [366,367,368]. However, all of these techniques are designed for the patterning of polymer nanofiber mats.
  • Seventh, the low production rate is another important disadvantage of this method. Therefore, the major task after success in laboratories is to optimize the electrospinning process in order to increase productivity.
Wang et al. [369] believe that a modification of the injection system, such as the introduction of multi-spinneret components that allows parallel multiprocessing and the development of free surface electrospinning methods, can solve the problem of increasing production volumes. A schematic diagram of this modified free surface coaxial electrospinning setup is shown in Figure 34, which contains five main components: a high-voltage direct current power supply, a stepped pyramid spinneret, a Teflon solution reservoir, two pumps and a grounded collector [370]. A stepped pyramid spinneret was used as the electrospinning generator. Indeed, the introduction of these developments made it possible to increase the rate of formation of nanofibers by more than 250 times [352,369]. However, even with the implementation of these modifications, the speed of fiber production using electrospinning technology will be much lower compared to other existing technologies for forming a sensitive layer used in thick-film technology. In addition, this technology is not suitable for forming nanofiber mats in small areas of the gas sensor platform, as required by the technology of manufacturing gas sensors.
  • Eighth, electrospun nanofibers are mostly randomly oriented (see Figure 26, Figure 28 and Figure 33), which significantly limits the ability to obtain the required repeatability of the final structures of nanofiber mats [371]. In this case, due to the arbitrary position of nanofibers on the surface of the substrates, the fewer fibers there are in the coating, the lower the repeatability of the fiber mat structure.
Researchers are trying to overcome this limitation with new methods of electrospinning [90] which give better control over the electrospun nanofiber orientation. For example, Choi et al. [138] suggested using aligned nanofibers to fabricate a gas-sensitive layer. To align the fibers parallel to each other along some axis, two strips of aluminum wires were placed along opposite edges of the substrate and connected to the ground terminal of the power supply, which created an electric field between the nozzle and the substrate during the electrospinning processes. This imposed a directional distribution of electric field lines between the nozzle and the aluminum wires, as shown in Figure 35, making it easier to align the fiber segments along the aluminum wires [101]. This technique was discussed earlier. Undoubtedly, the use of aligned nanofibers should increase the repeatability of the structure of the gas-sensitive layer and hence improve the reproducibility of the sensor parameters. Choi et al. [138] also found that NO2 sensors with aligned ZnO nanofibers had increased sensitivity compared to sensors using nonaligned ZnO nanofibers. They suggested that this effect is due to the specificity of the current flow in structures with aligned nanofibers.
  • Ninth, and most importantly, electrospinning technology is incompatible with traditional mass production processes.
Dong et al. [192] believe that this problem can be solved by eliminating the use of long fibers. They proposed fragmenting SnO2:Pt nanofibers into smaller pieces (several micrometers long) by ultrasonication in isopropanol and then using them as ink for printing on a microplatform with finger Au electrodes and a Pt micro-heater. The fabricated gas sensors operated at a power below 36 mW at 300 °C. At this temperature, the sensor response to 20 ppm H2S exceeded 5000. H2S sensors based on SnO2:CuO nanofibers were manufactured in the same way [194].
Kang et al. [372] also demonstrated the effectiveness of the proposed technology for the manufacture of gas sensors. The ink prepared on the base of fragmented nanofibers was used for electrohydrodynamic (EHD) printing. This technique was tested for four metal oxides: SnO2, WO3, In2O3 and NiO. Each metal oxide nanofiber was mixed with ethanol, and then ultrasonication process was conducted for an hour. After ultrasonication, the ethanol in the nanofiber solution was evaporated in a convection oven at 70 °C for 9 h. The dried electrospun fibers were finally dispersed in alpha-terpineol or ethyleneglycol solvents with a 15 wt% concentration. After fragmentation, the length of SnO2, WO3, In2O3 and NiO nanofibers, used to fabricate gas sensors, was ~60–70 μm, 40–60 μm, 55–70 μm and 60–70 μm, respectively. For the manufacture of gas sensors, a platform with a micro-heater was used (Figure 36). This process is illustrated in Figure 37. The active area of each gas sensor was smaller than 100 × 100 μm2. Testing of manufactured sensors for sensitivity to NO2, H2S and CO showed that all sensors could detect NO2 (0.1 ppm), H2S (1 ppm) and CO (20 ppm).
Yan et al. [373], Du et al. [327] and Zhang et al. [139] used an even simpler technique for the manufacture of sensors. After the formation of metal oxide nanofibers and their calcination at 550–600 °C, Yan et al. [373] and Du et al. [327] made a paste from these fibers using ethanol as the solvent. Subsequently, the pasts were applied on ceramic tubes with a pair of gold electrodes. Lim et al. [160] also used this approach to fabricate In2O3 nanofiber-based gas sensors. The sensor was fabricated by dropping In2O3 gel on a sapphire substrate with pre-patterned Pt electrodes. The nanofiber-based gel, in addition to In2O3 nanofibers, contained deionized water and agate mortar. Unfortunately, the authors did not report what changes occurred in the nanofibers as a result of such processing. Regarding the parameters of these sensors, the sensors fabricated by Yan et al. [373] had a high sensitivity to ethanol. At an ethanol concentration of 100 ppm, the sensor response at 280 °C reached 40.
It must be recognized that the approach suggested by Dong et al. [192], Kang et al. [372] and Yan et al. [373] allows the fabrication of nanofiber-based sensors using thick-film technology compatible with mass production. However, this technology, unfortunately, does not solve the other problems of electrospinning described earlier. In addition, a new technological operation, such as ultrasonication, which requires control, is added. Rejecting the traditional configuration of nanofiber-based membranes, we also lose the main advantage of nanofiber-based sensors, namely, good performance. For example, sensors based on fragmented In2O3 and NiO nanofibers that showed a faster response and recovery, at a temperature of 300 °C, had a response time/recovery time of 24 s/92 s and 32 s/43 s, respectively. For comparison, In2O3 nanofiber-based ethanol sensors, made from aligned nanofibers based on the standard nanofiber-based sensor manufacturing approach, had response and recovery times of 1 s and 3 s, respectively [100].

5. Summary

The analysis carried out in Part 1 and Part 2 of this article showed that electrospinning technology really presents great opportunities for the formation of gas-sensitive materials with a unique combination of parameters. Electrospinning makes it possible to form a gas-sensitive matrix from small crystallites while maintaining a very high gas permeability of the matrix due to the ultra-high porosity of the structure. This, on the one hand, provides an ultra-high sensitivity of the sensors under optimal conditions and, on the other hand, guarantees a fast response and recovery, since such a structure has no diffusion restrictions for gas penetration into the gas-sensitive matrix. However, it should be noted that this is only possible if the sensor response is limited by gas diffusion in the gas-sensitive matrix. If the kinetics of the sensor response is controlled by the kinetics of surface processes, then in this case, there will be no improvement in the kinetics of the sensor response.
However, it must be admitted that sensors manufactured using traditional technology, with appropriate optimization of the structure and composition of the gas-sensitive layer, can have the same parameters and, in some cases, even exceed the parameters reported for the best samples of nanofiber-based gas sensors [197,251,297,374,375,376]. This means that no breakthrough has been achieved in the development of gas sensors based on the use of electrospinning technology, which could manifest itself in a radical improvement in the parameters of devices in comparison with conventional technology. Unfortunately, the fulfillment of this condition is one of the prerequisites for the implementation of new technologies. If the performances of the devices being developed change insignificantly, then it makes no sense to bear the high costs associated with the implementation and development of a new technological process. This is exactly the situation in the sensor market, where there are still ceramic sensors, the technology of which was developed more than 40 years ago. If we add to this the technological difficulties that arise when organizing the mass production of nanofiber-based gas sensors, then we can confidently state that nanofiber-based sensors cannot yet compete with cheaper devices manufactured using traditional thick-film and thin-film technologies. It is important to note here that this conclusion applies only to conductometric gas sensors and does not apply to other areas where electrospun nanofibers can be used with great success [11,344,352,369,377,378,379].
Based on the above, we can conclude that electrospinning technology, despite some advantages, is unlikely to be used in the near future in the development of gas sensors designed for the market. At the same time, in the development of sensors designed for specific applications such as the security sector and the fight against terrorism, where ultra-high sensitivity is required with a quick response, this technology can find application. In this area, manufacturability and price fade into the background. Using electrospinning technology to find new gas-sensitive materials with specific multi-functional properties, combining unique structural, electrical and physical properties, is also of interest and promise in the future in the development of new-generation sensors. The ability to form solid, hollow and core–shell fibers with a controlled structure and composition offers great opportunities for this. It is only necessary to find technological solutions to introduce this technology into mass production.

Funding

This research was funded by the State Program of the Republic of Moldova project 20.80009.5007.02.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are reported in the paper directly.

Acknowledgments

The author is thankful to MDPI for supporting the publication of this article.

Conflicts of Interest

The author declares no conflict of interest. The funder had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Zhou, Z.; Lan, C.; Wei, R.; Ho, J.C. Transparent metal-oxide nanowires and their applications in harsh electronics. J. Mater. Chem. C 2019, 7, 202–217. [Google Scholar] [CrossRef]
  2. Scheideler, W.; Subramanian, V. Printed flexible and transparent electronics: Enhancing low-temperature processed metal oxides with 0D and 1D nanomaterials. Nanotechnology 2019, 30, 272001. [Google Scholar] [CrossRef]
  3. Bagga, S.; Akhtar, J.; Mishra, S. Synthesis and applications of ZnO nanowire: A review. AIP Conf. Proc. 2018, 1989, 020004. [Google Scholar]
  4. Chen, C.; Fan, Y.; Gu, J.; Wu, L.; Passerini, S.; Mai, L. One-dimensional nanomaterials for energy storage. J. Phys. D Appl. Phys. 2018, 51, 113002. [Google Scholar] [CrossRef] [Green Version]
  5. Hahm, J.-I. Fundamental properties of one-dimensional zinc oxide nanomaterials and implementations in various detection modes of enhanced biosensing. Annu. Rev. Phys. Chem. 2016, 67, 691–717. [Google Scholar] [CrossRef] [Green Version]
  6. Weng, B.; Liu, S.; Tang, Z.-R.; Xu, Y.-J. One-dimensional nanostructure based materials for versatile photocatalytic applications. RSC Adv. 2014, 4, 12685–12700. [Google Scholar] [CrossRef]
  7. Pan, J.; Shen, H.; Mathur, S. One-dimensional SnO2 nanostructures: Synthesis and applications. J. Nanotechnol. 2012, 2012, 917320. [Google Scholar] [CrossRef] [Green Version]
  8. Devan, R.S.; Patil, R.A.; Lin, J.-H.; Ma, Y.-R. One-dimensional metal-oxide nanostructures: Recent developments in synthesis, characterization, and applications. Adv. Funct. Mater. 2012, 22, 3326–3370. [Google Scholar] [CrossRef]
  9. Zhai, T.; Fang, X.; Liao, M.; Xu, X.; Zeng, H.; Yoshio, B.; Golberg, D. A Comprehensive review of one-dimensional metal-oxide nanostructure photodetectors. Sensors 2009, 9, 6504–6529. [Google Scholar] [CrossRef]
  10. Xia, Y.; Yang, P.; Sun, Y.; Wu, Y.; Mayers, B.; Gates, B.; Yin, Y.; Kim, F.; Yan, H. One-dimensional nanostructures: Synthesis, characterization, and applications. Adv. Mater. 2003, 15, 353–389. [Google Scholar] [CrossRef]
  11. Wang, X.; Ding, B.; Yu, J.; Wang, M. Engineering biomimetic superhydrophobic surfaces of electrospun nanomaterials. Nano Today 2011, 6, 510–530. [Google Scholar] [CrossRef]
  12. Imran, M.; Motta, N.; Shafiei, M. Electrospun one-dimensional nanostructures: A new horizon for gas sensing materials. Beilstein J. Nanotechnol. 2018, 9, 2128–2170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Nasiri, N.; Clarke, C. Nanostructured chemiresistive gas sensors for medical applications. Sensors 2019, 19, 462. [Google Scholar] [CrossRef] [Green Version]
  14. Zhang, S.; Jia, Z.; Liu, T.; Wei, G.; Su, Z. Electrospinning nanoparticles-based materials interfaces for sensor applications. Sensors 2019, 19, 3977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kolmakov, A.; Moskovits, M. Chemical sensing and catalysis by one dimensional metal-oxide nanostructures. Ann. Rev. Mater. Res. 2004, 34, 151–180. [Google Scholar] [CrossRef] [Green Version]
  16. Chen, P.-C.; Shen, G.; Zhou, C. Chemical sensors and electronic noses based on 1-D metal oxide nanostructures. IEEE Trans. Nanotechnol. 2008, 7, 668–682. [Google Scholar] [CrossRef]
  17. Morante, J.R. Chemical to electrical transduction mechanisms from single metal oxide nanowire measurements: Response time constant analysis. Nanotechnology 2013, 24, 444004. [Google Scholar] [CrossRef]
  18. Chen, X.; Wong, C.K.Y.; Yuan, C.A.; Zhang, G. Nanowire-based gas sensors. Sens. Actuators B Chem. 2013, 177, 178–195. [Google Scholar] [CrossRef]
  19. Ramgir, N.; Datta, N.; Kaur, M.; Kailasaganapathi, S.; Debnath, A.K.; Aswal, D.K.; Gupta, S.K. Metal oxide nanowires for chemiresistive gas sensors: Issues, challenges and prospects. Colloids Surf. A 2013, 439, 101–116. [Google Scholar] [CrossRef]
  20. Comini, E. Metal oxide nanowire chemical sensors: Innovation and quality of life. Mater. Today 2016, 19, 559–567. [Google Scholar] [CrossRef]
  21. Lee, J.-H.; Mirzaie, A.; Kim, J.-Y.; Kim, J.-H.; Kim, H.W.; Kim, S.K. Optimization of the surface coverage of metal nanoparticles on nanowires gas sensors to achieve the optimal sensing performance. Sens. Actuators B Chem. 2020, 302, 127196. [Google Scholar] [CrossRef]
  22. Wang, Y.; Duan, L.; Deng, Z.; Liao, J. Electrically transduced gas sensors based on semiconducting metal oxide nanowires. Sensors 2020, 20, 6781. [Google Scholar] [CrossRef] [PubMed]
  23. Korotcenkov, G. Current trends in nanomaterials for metal oxide-based conductometric gas sensors: Advantages and limitations. Part 1: 1D and 2D nanostructures. Nanomaterials 2020, 10, 1392. [Google Scholar] [CrossRef] [PubMed]
  24. Korotcenkov, G. Electrospun metal oxide nanofibers and its conductometric gas sensor application. Part 1: Nanofibers and features of their forming. Nanomaterials 2021, 11, 1544. [Google Scholar] [CrossRef]
  25. Ding, B.; Wang, M.; Wang, X.; Yu, J.; Sun, G. Electrospun nanomaterials for ultrasensitive sensors. Mater. Today 2010, 13, 16–27. [Google Scholar] [CrossRef]
  26. Choi, S.-J.; Persano, L.; Camposeo, A.; Jang, J.-S.; Koo, W.-T.; Kim, S.-J.; Cho, H.-J.; Kim, I.-D.; Pisignano, D. Electrospun nanostructures for high performance chemiresistive and optical sensors. Macromol. Mater. Eng. 2017, 302, 1600569. [Google Scholar] [CrossRef] [Green Version]
  27. Vuong, N.M.; Jung, H.; Kim, D.; Kim, H.; Hong, S.-K. Realization of an open space ensemble for nanowires: A strategy for the maximum response in resistive sensors. J. Mater. Chem. 2012, 22, 6716–6725. [Google Scholar] [CrossRef]
  28. Korotcenkov, G. The role of morphology and crystallographic structure of metal oxides in response of conductometric-type gas sensors. Mater. Sci. Eng. Rep. 2008, 61, 1–39. [Google Scholar] [CrossRef]
  29. Yamazoe, N. New approaches for improving semiconductor gas sensors. Sens. Actuators B Chem. 1991, 5, 7–19. [Google Scholar] [CrossRef]
  30. Barsan, N.; Weimar, U. Conduction model of metal oxide gas sensors. J. Electroceram. 2001, 7, 143–16746. [Google Scholar] [CrossRef]
  31. Korotcenkov, G. Metal oxides for solid state gas sensors: What determines our choice? Mater. Sci. Eng. B 2007, 139, 1–23. [Google Scholar] [CrossRef]
  32. Yamazoe, N.; Shimanoe, K. Receptor function and response of semiconductor gas sensor. J. Sens. 2009, 2009, 875704. [Google Scholar] [CrossRef]
  33. Yamazoe, N.; Fuchigami, J.; Kishikawa, M.; Seiyama, T. Interactions of tin oxide surface with O2, H2O and H2. Surf. Sci. 1979, 86, 335–344. [Google Scholar] [CrossRef]
  34. Morrison, S.R. Selectivity in semiconductor gas sensor. Sens. Actuators 1987, 12, 425–440. [Google Scholar] [CrossRef]
  35. Kohl, D. The role of noble-metals in the chemistry of solid-state gas sensors. Sens. Actuators B Chem. 1990, 1, 158–165. [Google Scholar] [CrossRef]
  36. Xu, C.N.; Tamaki, J.; Miura, N.; Yamazoe, N. Grain-size effects on gas sensitivity of porous SnO2-based elements. Sens. Actuators B Chem. 1991, 3, 147–155. [Google Scholar] [CrossRef]
  37. Kupriyanov, L.Y. (Ed.) Semiconductor Sensors in Physico-Chemical Studies; Elsevier: Amsterdam, The Netherlands, 1992; p. 400. [Google Scholar]
  38. Sberveglieri, G. (Ed.) Gas Sensors; Springer: Amsterdam, The Netherlands, 1992; p. 409. [Google Scholar]
  39. Williams, D.E.; Henshaw, G.S.; Pratt, K.F.E.; Peat, R. Reaction-diffusion effects and systematic design of gas sensitive resistors based on semiconducting oxides. J. Chem. Soc. Faraday Trans. 1995, 91, 4299–4307. [Google Scholar] [CrossRef]
  40. Williams, D.E. Semiconducting oxides as gas-sensitive resistors. Sens. Actuators B Chem. 1999, 57, 1–16. [Google Scholar] [CrossRef]
  41. Brynzari, V.; Korotchenkov, G.; Dmitriev, S. Simulation of thin film gas sensor kinetics. Sens. Actuators B Chem. 1999, 61, 143–153. [Google Scholar] [CrossRef]
  42. Brynzari, V.; Korotchenkov, G.; Dmitriev, S. Theoretical study of semiconductor thin film gas sensitivity: Attempt to consistent approach. J. Electron. Technol. 2000, 33, 225–235. [Google Scholar]
  43. Hahn, S.H.; Barsan, N.; Weimar, U.; Ejakov, S.G.; Visser, J.H.; Soltis, R.E. CO sensing with SnO2 thick film sensors: Role of oxygen and water vapour. Thin Solid Film. 2003, 436, 17–24. [Google Scholar] [CrossRef]
  44. Batzill, M. Surface science studies of gas sensing materials: SnO2. Sensors 2006, 6, 1345–1366. [Google Scholar] [CrossRef] [Green Version]
  45. Gurlo, A. Interplay between O2 and SnO2: Oxygen ionosorption and spectroscopic evidence for adsorbed oxygen. ChemPhysChem 2006, 7, 2041–2052. [Google Scholar] [CrossRef] [PubMed]
  46. Korotcenkov, G.; Blinov, I.; Stetter, J.R. Kinetics of indium oxide-based thin film gas sensor response: The role of “redox” and adsorption/desorption processes in gas sensing effects. Thin Solid Film. 2007, 515, 3987–3996. [Google Scholar] [CrossRef]
  47. Korotcenkov, G.; Brinzari, V.; Stetter, J.R.; Blinov, I.; Blaja, V. The nature of processes controlling the kinetics of indium oxide-based thin film gas sensor response. Sens. Actuators B Chem. 2007, 128, 51–63. [Google Scholar] [CrossRef]
  48. Batzill, M.; Diebold, U. Surface studies of gas sensing metal oxides. Phys. Chem. Chem. Phys. 2007, 9, 2307–2318. [Google Scholar] [CrossRef] [PubMed]
  49. Barsan, N.; Koziej, D.; Weimar, U. Metal oxide-based gas sensor research: How to? Sens. Actuators B Chem. 2007, 121, 18–35. [Google Scholar] [CrossRef]
  50. Gurlo, A.; Riedel, R. In situ and operando spectroscopy for assessing mechanisms of gas sensing. Angew. Chem. Int. Ed. 2007, 46, 3826–3848. [Google Scholar] [CrossRef] [PubMed]
  51. Rumyantseva, M.N.; Gas’kov, A.M. Chemical modification of nanocrystalline metal oxides: Effect of the real structure and surface chemistry on the sensor properties. Russ. Chem. Bull. 2008, 57, 1106–1125. [Google Scholar] [CrossRef]
  52. Brinzari, V.; Ivanov, M.; Cho, B.K.; Kamei, M.; Korotcenkov, G. Photoconductivity in In2O3 nanoscale thin films: Interrelation with chemisorbed-type conductometric response towards oxygen. Sens. Actuators B Chem. 2010, 148, 427–438. [Google Scholar] [CrossRef]
  53. Barsan, N.; Simion, C.; Heine, T.; Pokhrel, S.; Weimar, U. Modeling of sensing and transduction for p-type semiconducting metal oxide based gas sensors. J. Electroceram. 2010, 25, 11–19. [Google Scholar] [CrossRef]
  54. Yamazoe, N.; Shimanoe, K. Basic approach to the transducer function of oxide semiconductor gas sensors. Sens. Actuators B Chem. 2011, 160, 1352–1362. [Google Scholar] [CrossRef]
  55. Korotcenkov, G.; Cho, B.K. The role of grain size on the thermal stability of nanostructured metal oxides used in gas sensor applications and approaches for grain-size stabilization. Prog. Crystal. Growth 2012, 58, 167–208. [Google Scholar] [CrossRef]
  56. Korotcenkov, G. (Ed.) Chemical Sensors: Simulation and Modeling. Volume 2: Conductometric Gas Sensor; Momentum Press: New York, NY, USA, 2012. [Google Scholar]
  57. Korotcenkov, G. (Ed.) Chemical Sensors: Simulation and Modeling. Volume 1: Microstructural Characterization and Modeling of Metal Oxides; Momentum Press: New York, NY, USA, 2012. [Google Scholar]
  58. Kim, H.J.; Lee, J.-H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B Chem. 2014, 192, 607–627. [Google Scholar] [CrossRef]
  59. Korotcenkov, G.; Brinzari, V.; Cho, B.K. In2O3 and SnO2 based ozone sensors: Fundamentals. J. Sens. 2016, 2016, 3816094. [Google Scholar] [CrossRef] [Green Version]
  60. Korotcenkov, G.; Cho, B.K. Metal oxide based composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B Chem. 2017, 244, 182–210. [Google Scholar] [CrossRef]
  61. Brinzari, V.; Korotcenkov, G. Kinetic approach to receptor function in chemiresistive gas sensor modeling of tin dioxide. Steady state consideration. Sens. Actuators B Chem. 2018, 259, 443–454. [Google Scholar] [CrossRef]
  62. Barsan, N.; Schierbaum, K. (Eds.) Gas Sensors Based on Conducting Metal Oxides; Elsevier Metal Oxide Series edited by Korotcenkov, G. Elsevier: Cambridge, MA, USA, 2018. [Google Scholar]
  63. Korotcenkov, G. Handbook of Humidity Measurement: Methods, Materials and Technologies. Volume 2: Electronic and Electrical Humidity Sensors; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  64. Jaaniso, R.; Tan, O.K. (Eds.) Semiconductor Gas Sensors; Woodhead Publishing: Cambridge, UK, 2020. [Google Scholar]
  65. Abideen, Z.U.; Kim, J.-H.; Lee, J.-H.; Kim, J.-Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Electrospun metal oxide composite nanofibers gas sensors: A review. J. Korean Ceram. Soc. 2017, 54, 366–379. [Google Scholar] [CrossRef] [Green Version]
  66. Wang, X.; Drew, C.; Lee, S.-H.; Senecal, K.J.; Kumar, J.; Samuelson, L.A. Electrospun nanofibrous membranes for highly sensitive optical sensors. Nano Lett. 2002, 2, 1273–1275. [Google Scholar] [CrossRef]
  67. Ding, B.; Kim, J.; Miyazaki, Y.; Shiratori, S. Electrospun nanofibrous membranes coated quartz crystal microbalance as gas sensor for NH3 detection. Sens. Actuators B Chem. 2004, 101, 373–380. [Google Scholar] [CrossRef]
  68. Ding, B.; Yamazaki, M.; Shiratori, S. Electrospun fibrous polyacrylic acid membrane-based gas sensors. Sens. Actuators B Chem. 2005, 106, 477–483. [Google Scholar] [CrossRef]
  69. Aussawasathien, D.; Dong, J.-H.; Dai, L. Electrospun polymer nanofiber sensors. Synth. Met. 2005, 154, 37–40. [Google Scholar] [CrossRef]
  70. Dan, Y.; Cao, Y.; Mallouk, T.E.; Johnson, A.T.; Evoy, S. Dielectrophorectically assembled polymer nanowires for gas sensing. Sens. Actuators B Chem. 2007, 125, 55–79. [Google Scholar] [CrossRef] [Green Version]
  71. Chen, D.; Lei, S.; Chen, Y. A single polyaniline nanofiber field effect transistor and its gas sensing mechanisms. Sensors 2011, 11, 6509–6516. [Google Scholar] [CrossRef]
  72. Haghi, A.; Zaikov, G.E. Electrospinning Process and Nanofiber Research; Nova Science Publishing: New York, NY, USA, 2011. [Google Scholar]
  73. Ding, B.; Wang, M.; Yu, J.; Sun, G. Gas sensors based on electrospun nanofibers. Sensors 2009, 9, 1609–1624. [Google Scholar] [CrossRef] [Green Version]
  74. Liu, Y.; Wang, R.; Zhang, T.; Liu, S.; Fei, T. Zeolitic imidazolate framework-8 (ZIF-8)-coated In2O3 nanofibers as an efficient sensing material for ppb-level NO2 detection. J. Colloid Interface Sci. 2019, 541, 249–257. [Google Scholar] [CrossRef]
  75. Liu, W.; Zhou, X.; Xu, L.; Zhu, S.; Yang, S.; Chen, X.; Dong, B.; Bai, X.; Lu, G.; Song, H. Graphene quantum dot-functionalized three-dimensional ordered mesoporous ZnO for acetone detection toward diagnosis of diabetes. Nanoscale 2019, 11, 11496–11504. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, W.; Sun, J.; Xu, L.; Zhu, S.; Zhou, X.; Yang, S.; Dong, B.; Bai, X.; Lu, G.; Song, H. Understanding the noble metal modifying effect on In2O3 nanowires: Highly sensitive and selective gas sensors for potential early screening of multiple diseases. Nanoscale Horiz. 2019, 4, 1361–1371. [Google Scholar] [CrossRef]
  77. Imran, M.; Haroon Rashid, S.S.A.A.; Sabri, Y.; Motta, N.; Tesfamichael, T.; Sonar, P.; Shafiei, M. Template based sintering of WO3 nanoparticles into porous tungsten oxide nanofibers for acetone sensing applications. J. Mater. Chem. C 2019, 7, 2961–2970. [Google Scholar] [CrossRef]
  78. Imran, Z.; Rasool, K.; Batool, S.S.; Ahmad, M.; Rafiq, M.A. Effect of different electrodes on the transport properties of ZnO nanofibers under humid environment. AIP Adv. 2015, 5, 11721. [Google Scholar] [CrossRef] [Green Version]
  79. Abideen, Z.U.; Katoch, A.; Kim, J.H.; Kwon, Y.J.; Kim, H.W.; Kim, S.S. Excellent gas detection of ZnO nanofibers by loading with reduced graphene oxide nanosheets. Sens. Actuators B Chem. 2015, 221, 1499–1507. [Google Scholar] [CrossRef]
  80. He, Y.; Zhang, T.; Zheng, W.; Wang, R.; Liu, X.; Xia, Y.; Zhao, J. Humidity sensing properties of BaTiO3 nanofiber prepared via electrospinning. Sens. Actuators B Chem. 2010, 146, 98–102. [Google Scholar] [CrossRef]
  81. Wang, L.; He, Y.; Hu, J.; Qi, Q.; Zhang, T. DC humidity sensing properties of BaTiO3 nanofiber sensors with different electrode materials. Sens. Actuators B Chem. 2011, 153, 460–464. [Google Scholar] [CrossRef]
  82. Batool, S.S.; Imran, Z.; Qadir, M.I.; Usman, M.; Jamil, H.; Rafiq, M.A.; Hassan, M.M.; Willander, M. Comparative analysis of Ti, Ni, and Au electrodes on characteristics of TiO2 nanofibers for humidity sensor application. J. Mater. Sci. Technol. 2013, 29, 411–414. [Google Scholar] [CrossRef]
  83. Camargo, C.J.; Melendez, A.; Robles, J.; Esteve, J.; Ramos, I. Electrospun nanobridges towards self-heated gas sensors with enhanced sensitivity. J. Phys. Conf. Ser. 2013, 421, 012002. [Google Scholar] [CrossRef]
  84. Raible, I.; Burghard, M.; Schlecht, U.; Yasuda, A.; Vossmeyer, T. V2O5 nanofibres: Novel gas sensors with extremely high sensitivity and selectivity to amines. Sens. Actuators B Chem. 2005, 106, 730–735. [Google Scholar]
  85. Moon, J.; Park, J.-A.; Lee, S.-J.; Lee, J.-L.; Zyung, T.; Shin, E.-C.; Lee, J.-S. A physicochemical mechanism of chemical gas sensors using an AC analysis. Phys. Chem. Chem. Phys. 2013, 15, 9361–9374. [Google Scholar] [CrossRef]
  86. Nikfarjam, A.; Hosseini, S.; Salehifar, N. Fabrication of a highly sensitive single aligned TiO2 and gold nanoparticle embedded TiO2 nano-fiber gas sensor. ACS Appl. Mater. Interfaces 2017, 9, 15662–15671. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, Y.; Ramos, I.; Santiago-Avilés, J.J. Detection of moisture and methanol gas using a single electrospun tin oxide nanofiber. IEEE Sens. J. 2007, 7, 1347–1348. [Google Scholar] [CrossRef] [Green Version]
  88. Ahmad, M.; Pan, C.; Luo, Z.; Zhu, J. A Single ZnO nanofiber-based highly sensitive amperometric glucose biosensor. J. Phys. Chem. C 2010, 114, 9308–9313. [Google Scholar] [CrossRef]
  89. Guo, J.; Song, Y.; Chen, D.; Jiao, X. Fabrication of ZnO nanofibers by electrospinning and electrical properties of a single nanofiber. J. Dispers. Sci. Technol. 2010, 31, 684–689. [Google Scholar] [CrossRef]
  90. Alghoraibi, I.; Alomari, S. Different methods for nanofiber design and fabrication. In Handbook of Nanofibers; Barhoum, A., Bechelany, M., Makhlouf, A.S.H., Eds.; Springer Nature: Cham, Switzerland, 2019; pp. 79–124. [Google Scholar]
  91. Wu, H.; Pan, W.; Lin, D.; Li, H. Electrospinning of ceramic nanofibers: Fabrication, assembly and applications. J. Adv. Ceram. 2012, 1, 2–23. [Google Scholar] [CrossRef] [Green Version]
  92. Matthews, J.A.; Wnek, G.E.; Simpson, D.G.; Bowlin, G.L. Electrospinning of collagen nanofibers. Biomacromolecules 2002, 3, 232–238. [Google Scholar] [CrossRef]
  93. Baniasad, M.; Huang, J.; Xu, Z.; Moreno, S.; Yang, X.; Chang, J.; Quevedo-Lopez, M.A.; Naraghi, M.; Minary-Jolandan, M. High-performance coils and yarns of polymeric piezoelectric nanofibers. ACS Appl. Mater. Interfaces 2015, 7, 5358–5366. [Google Scholar] [CrossRef] [PubMed]
  94. Zussman, E.; Theron, A.; Yarin, A.L. Formation of nanofiber crossbars in electrospinning. Appl. Phys. Lett. 2003, 82, 973–975. [Google Scholar] [CrossRef]
  95. Xu, C.Y.; Inai, R.; Kotaki, M.; Ramakrishna, S. Aligned biodegradable nanofibrous structure: A potential scaffold for blood vessel engineering. Biomaterials 2004, 25, 877–886. [Google Scholar] [CrossRef]
  96. Prabhakaran, M.P.; Vatankhah, E.; Ramakrishna, S. Electrospun aligned PHBV/collagen nanofibers as substrates for nerve tissue engineering. Biotechnol. Bioeng. 2013, 110, 2775–2784. [Google Scholar] [CrossRef]
  97. Theron, A.; Zussman, E.; Yarin, A.L. Electrostatic field-assisted alignment of electrospun nanofibers. Nanotechnology 2001, 12, 384–390. [Google Scholar] [CrossRef]
  98. Katta, P.; Alessandro, M.; Ramsier, R.D.; Chase, G.G. Continuous electrospinning of aligned polymer nanofibers onto a wire drum collector. Nano Lett. 2004, 4, 2215–2218. [Google Scholar] [CrossRef]
  99. Wu, H.; Lin, D.; Zhang, R.; Pan, W. Oriented nanofibers by a newly modified electrospinning method. J. Am. Ceram. Soc. 2007, 90, 632–634. [Google Scholar] [CrossRef]
  100. Liu, H.; Fan, H.-T.; Xu, X.-J.; Zhang, T. Uniaxially aligned In2O3 nanofibers based sensors with fast response to ethanol. Biomed. Eng. Appl. Basis Commun. 2012, 24, 105109. [Google Scholar] [CrossRef]
  101. Li, D.; Wang, Y.L.; Xia, Y.N. Electrospinning of polymeric and ceramic nanofibers as uniaxially aligned arrays. Nano Lett. 2003, 3, 1167–1171. [Google Scholar] [CrossRef]
  102. Li, D.; Wang, Y.L.; Xia, Y.N. Electrospinning nanofibers as uniaxially aligned arrays and layer-by-layer stacked films. Adv. Mater. 2004, 16, 361–366. [Google Scholar] [CrossRef]
  103. Li, D.; Ouyang, G.; Mccann, J.T.; Xia, Y.N. Collecting electrospun nanofibers with patterned electrodes. Nano Lett. 2005, 5, 913–916. [Google Scholar] [CrossRef] [PubMed]
  104. Ke, J.Y.; Chu, H.J.; Hsu, Y.H.; Lee, C.K. A highly flexible piezoelectret-fiber pressure sensor based on highly aligned P(VDF-TrFE) electrospun fibers. Proc. SPIE 2017, 10164, 101642X. [Google Scholar]
  105. Zhang, Y.; He, X.; Li, J.; Miao, Z.; Huang, F. Fabrication and ethanol-sensing properties of micro gas sensor based on electrospun SnO2 nanofibers. Sens. Actuators B Chem. 2008, 132, 67–73. [Google Scholar] [CrossRef]
  106. Landau, O.; Rothschild, A.; Zussman, E. Processing-microstructure-properties correlation of ultrasensitive gas sensors produced by electrospinning. Chem. Mater. 2009, 21, 9–11. [Google Scholar] [CrossRef]
  107. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Reduced graphene oxide (rGO)-loaded metal-oxide nanofiber gas sensors: An overview. Sensors 2021, 21, 1352. [Google Scholar] [CrossRef]
  108. Wang, G.; Ji, Y.; Huang, X.; Yang, X.; Gouma, P.; Dudley, M. Fabrication and characterization of polycrystalline WO3 nanofibers and their application for ammonia sensing. J. Phys. Chem. B 2006, 110, 23777–23782. [Google Scholar] [CrossRef]
  109. Blachowicz, T.; Ehrmann, A. Recent developments in electrospun ZnO nanofibers: A short review. J. Eng. Fibers Fabr. 2000, 15, 1–6. [Google Scholar] [CrossRef] [Green Version]
  110. Yang, M.; Xie, T.; Peng, L.; Zhao, Y.; Wang, D. Fabrication and photoelectric oxygen sensing characteristics of electrospun Co doped ZnO nanofibers. Appl. Phys. A 2007, 89, 427–430. [Google Scholar] [CrossRef]
  111. Sahner, K.; Gouma, P.; Moos, R. Electrodeposited and sol-gel precipitated p-type SrTi1–xFexO3–δ semiconductors for gas sensing. Sensors 2007, 7, 1871–1886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Jiang, Z.; Jiang, T.; Wang, J.; Wang, Z.; Xu, X.; Wang, Z.; Zhao, R.; Li, Z.; Wang, C. Ethanol chemiresistor with enhanced discriminative ability from acetone based on Sr-doped SnO2 nanofibers. J. Colloid Interface Sci. 2015, 437, 252–258. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, T.T.; Ma, S.Y.; Cheng, L.; Luo, J.; Jiang, X.H.; Jin, W.X. Preparation of Yb-doped SnO2 hollow nanofibers with an enhanced ethanol–gas sensing performance by electrospinning. Sens. Actuators B Chem. 2015, 216, 212–220. [Google Scholar] [CrossRef]
  114. Kou, X.; Wang, C.; Ding, M.; Feng, C.; Li, X.; Ma, J.; Zhang, H.; Sun, Y.; Lu, G. Synthesis of Co-doped SnO2 nanofibers and their enhanced gas-sensing properties. Sens. Actuators B Chem. 2016, 236, 425–432. [Google Scholar] [CrossRef]
  115. Zhao, M.; Wang, X.; Cheng, J.; Zhang, L.; Jia, J.; Li, X. Synthesis and ethanol sensing properties of Al-doped ZnO nanofibers. Curr. Appl. Phys. 2013, 13, 403–407. [Google Scholar] [CrossRef]
  116. Wang, W.; Li, Z.; Zheng, W.; Huang, H.; Wang, C.; Sun, J. Cr2O3-sensitized ZnO electrospun nanofibers based ethanol detectors. Sens. Actuators B Chem. 2010, 143, 754–758. [Google Scholar] [CrossRef]
  117. Huang, B.; Zhao, C.; Zhang, M.; Zhang, Z.; Xie, E.; Zhou, J.; Han, W. Doping effect of In2O3 on structural and ethanol-sensing characteristics of ZnO nanotubes fabricated by electrospinning. Appl. Surf. Sci. 2015, 349, 615–621. [Google Scholar] [CrossRef] [Green Version]
  118. Li, Z.; Dzenis, Y. Highly efficient rapid ethanol sensing based on Co-doped In2O3 nanowires. Talanta 2011, 85, 82–85. [Google Scholar] [CrossRef]
  119. Zhao, C.; Huang, B.; Xie, E.; Zhou, J.; Zhang, Z. Improving gas-sensing properties of electrospun In2O3 nanotubes by Mg acceptor doping. Sens. Actuators B Chem. 2015, 207, 313–320. [Google Scholar] [CrossRef]
  120. Lian, H.; Feng, Y.; Wang, Z.; Liu, L.; Guo, X.; Wang, X. Porous Eu2O3-In2O3 nanotube-based ethanol gas sensor with high sensitivity and excellent selectivity. Appl. Phys. A 2017, 123, 158. [Google Scholar] [CrossRef]
  121. Wu, J.; Huang, Q.; Zeng, D.; Zhang, S.; Yang, L.; Xia, D.; Xiong, Z.; Xie, C. Al-doping induced formation of oxygen-vacancy for enhancing gas-sensing properties of SnO2 NTs by electrospinning. Sens. Actuators B Chem. 2014, 198, 62–69. [Google Scholar] [CrossRef]
  122. Wang, X.; Zhang, J.; Wang, L.; Li, S.; Liu, L.; Su, C.; Liu, L. High response gas sensors for formaldehyde based on Er-doped In2O3 nanotubes. J. Mater. Sci. Technol. 2015, 31, 1175–1180. [Google Scholar] [CrossRef]
  123. Wang, X.; Zhang, J.; He, Y.; Wang, L.; Liu, L.; Wang, H.; Guo, X.; Lian, H. Porous Nd-doped In2O3 nanotubes with excellent formaldehyde sensing properties. Chem. Phys. Lett. 2016, 658, 319–323. [Google Scholar] [CrossRef]
  124. Wang, X.; Li, H.; Ni, M.; Wang, L.; Liu, L.; Wang, H.; Guo, X. Excellent formaldehyde gas-sensing properties of ruptured Nd-doped In2O3 porous nanotubes. J. Electron. Mater. 2017, 46, 363–369. [Google Scholar] [CrossRef]
  125. Liu, C.; Wang, X.; Xie, F.; Liu, L.; Ruan, S. Fabrication of Sm-doped porous In2O3 nanotubes and their excellent formaldehyde-sensing properties. J. Mater. Sci. Mater. Electron. 2016, 27, 9870–9876. [Google Scholar] [CrossRef]
  126. He, Y.; Wang, D.; Ge, F.; Liu, L. SnO2-doped α-Fe2O3 patulous microtubes for high performance formaldehyde sensing. J. Semicond. 2015, 36, 083005. [Google Scholar] [CrossRef]
  127. Wang, X.; Zhao, M.; Liu, F.; Jia, J.; Li, X.; Cao, L. C2H2 gas sensor based on Ni-doped ZnO electrospun nanofibers. Ceram. Intern. 2013, 39, 2883–2887. [Google Scholar] [CrossRef]
  128. Feng, C.; Wang, C.; Zhang, H.; Li, X.; Wang, C.; Cheng, P.; Ma, J.; Sun, P.; Gao, Y.; Zhang, H.; et al. Enhanced sensitive and selective xylene sensors using W-doped NiO nanotubes. Sens. Actuators B Chem. 2015, 221, 1475–1482. [Google Scholar] [CrossRef]
  129. Xu, X.; Sun, J.; Zhang, H.; Wang, Z.; Dong, B.; Jiang, T.; Wang, W.; Li, Z.; Wang, C. Effects of Al doping on SnO2 nanofibers in hydrogen sensor. Sens. Actuators B Chem. 2011, 160, 858–863. [Google Scholar] [CrossRef]
  130. Liu, L.; Guo, C.; Li, S.; Wang, L.; Dong, Q.; Li, W. Improved H2 sensing properties of Co-doped SnO2 nanofiber. Sens. Actuators B Chem. 2010, 150, 806–810. [Google Scholar] [CrossRef]
  131. Yang, J.; Gao, C.; Yang, H.; Wang, X.; Jia, J. High selectivity of a CuO modified hollow SnO2 nanofiber gas sensor to H2S at low temperature. Eur. Phys. J. Appl. Phys. 2017, 79, 30101. [Google Scholar] [CrossRef]
  132. Zhao, M.; Wang, X.; Ning, L.; Jia, J.; Li, X.; Cao, L. Electrospun Cu-doped ZnO nanofibers for H2S sensing. Sens. Actuators B Chem. 2011, 156, 588–592. [Google Scholar] [CrossRef]
  133. Liu, J.; Guo, W.; Qu, F.; Feng, C.; Li, C.; Zhu, L.; Zhou, J.; Ruan, S.; Chen, W. V-doped In2O3 nanofibers for H2S detection at low temperature. Ceram. Intern. 2014, 40, 6685–6689. [Google Scholar] [CrossRef]
  134. Wang, Z.; Li, Z.; Jiang, T.; Xu, X.; Wang, C. Ultrasensitive hydrogen sensor based on Pd(0)-loaded SnO2 electrospun nanofibers at room temperature. ACS Appl. Mater. Interfaces 2013, 5, 2013–2021. [Google Scholar] [CrossRef]
  135. Fan, X.-X.; He, X.-L.; Li, J.-P.; Gao, X.-G.; Jia, J. Ethanol sensing properties of hierarchical SnO2 fibers fabricated with electrospun polyvinylpyrrolidone template. Vacuum 2016, 128, 112–117. [Google Scholar] [CrossRef]
  136. Zhang, Y.; Li, J.P.; An, G.M.; He, X.L. Highly porous SnO2 fibers by electrospinning and oxygen plasma etching and its ethanol-sensing properties. Sens. Actuators B Chem. 2010, 144, 43–48. [Google Scholar] [CrossRef]
  137. Zhao, Y.; He, X.; Li, J.; Gao, X.; Jia, J. Porous CuO/SnO2 composite nanofibers fabricated by electrospinning and their H2S sensing properties. Sens. Actuators B Chem. 2012, 165, 82–87. [Google Scholar] [CrossRef]
  138. Choi, S.-H.; Ankonina, G.; Youn, D.-Y.; Oh, S.-G.; Hong, J.-M.; Rothschild, A.; Kim, I.-D. Hollow ZnO nanofibers fabricated using electrospun polymer templates and their electronic transport properties. ACS Nano 2009, 3, 2623–2631. [Google Scholar] [CrossRef]
  139. Zhang, J.; Leng, D.; Zhang, L.; Li, G.; Ma, F.; Gao, J.; Lu, H.; Zhu, B. Porosity and oxygen vacancy engineering of mesoporous WO3 nanofibers for fast and sensitive low-temperature NO2 sensing. J. Alloys Compd. 2021, 853, 157339. [Google Scholar] [CrossRef]
  140. Du, N.; Zhang, H.; Chen, B.D.; Ma, X.Y.; Liu, Z.H.; Wu, J.B.; Yang, D.R. Porous indium oxide nanotubes: Layer-by-Layer assembly on carbon-nanotube templates and application for room-temperature NH3 gas sensors. Adv. Mater. 2007, 19, 1641–1645. [Google Scholar] [CrossRef]
  141. Nikfarjam, A.; Salehifar, N. Improvement in gas-sensing properties of TiO2 nanofiber sensor by UV irradiation. Sens. Actuators B Chem. 2015, 211, 146–156. [Google Scholar] [CrossRef]
  142. Kim, C.H.; Jung, Y.H.; Kim, H.Y.; Lee, D.R.; Nallasamy, D.; Choi, K.E. Effect of collector temperature on the porous structure of electrospun fibers. Macromol. Res. 2006, 14, 59–65. [Google Scholar] [CrossRef]
  143. Kim, I.-D.; Jeon, E.-K.; Choi, S.-H.; Choi, D.-K.; Tuller, H.L. Electrospun SnO2 nanofiber mats with thermo-compression step for gas sensing applications. J. Electroceram. 2010, 25, 159–167. [Google Scholar] [CrossRef]
  144. Korotcenkov, G.; Sysoev, V. Conductometric metal oxide gas sensors. In Chemical Sensors: Comprehensive Sensor Technologies. Volume 4: Solid State Devices; Korotcenkov, G., Ed.; Momentum Press: New York, NY, USA, 2011; pp. 53–186. [Google Scholar]
  145. Korotcenkov, G.; Han, S.D.; Cho, B.K.; Brinzari, V. Grain size effects in sensor response of nanostructured SnO2- and In2O3-based conductometric gas sensor. Crit. Rev. Solid State Mater. Sci. 2009, 34, 1–17. [Google Scholar] [CrossRef]
  146. Korotcenkov, G.; Brinzari, V.; Cho, B.K. In2O3 and SnO2-based ozone sensors: Design and characterization. Crit. Rev. Solid State Mater. Sci. 2018, 43, 83–132. [Google Scholar] [CrossRef]
  147. Senthil, T.; Anandhan, S.J. Structure-property relationship of sol-gel electrospun ZnO nanofibers developed for ammonia gas sensing. Colloid Interface Sci. 2014, 432, 285–296. [Google Scholar] [CrossRef] [PubMed]
  148. Park, J.Y.; Asokan, K.; Choi, S.-W.; Kim, S.S. Growth kinetics of nanograins in SnO2 fibers and size dependent sensing properties. Sens. Actuators B Chem. 2011, 152, 254–260. [Google Scholar] [CrossRef]
  149. Kim, I.-D.; Rothschild, A. Nanostructured metal oxide gas sensors prepared by electrospinning. Polym. Adv. Technol. 2011, 22, 318–325. [Google Scholar] [CrossRef]
  150. Jiang, C.; Zhang, G.; Wu, Y.; Li, L.; Shi, K. Facile synthesis of SnO2 nanocrystalline tubes by electrospinning and their fast response and high sensitivity to NOx at room temperature. CrystEngComm 2012, 14, 2739–2747. [Google Scholar] [CrossRef]
  151. Karimi, E.Z.; Esmaeilzadeh, J.; Marzbanrad, E. Electrospun TiO2 nanofibre-based gas sensors fabricated by AC electrophoresis deposition. Bull. Mater. Sci. 2015, 38, 209–214. [Google Scholar] [CrossRef]
  152. Giancaterini, L.; Emamjomeh, S.M.; De Marcellis, A.; Palange, E.; Resmini, A.; Anselmi-Tamburini, U.; Cantalini, C. The influence of thermal and visible light activation modes on the NO2 response of WO3 nanofibers prepared by electrospinning. Sens. Actuators B Chem. 2016, 229, 387–395. [Google Scholar] [CrossRef]
  153. Chi, X.; Liu, C.; Liu, L.; Li, Y.; Wang, Z.; Bo, X.; Liu, L.; Su, C. Tungsten trioxide nanotubes with high sensitive and selective properties to acetone. Sens. Actuators B Chem. 2014, 194, 33–37. [Google Scholar] [CrossRef]
  154. Cho, N.G.; Yang, D.J.; Jin, M.-J.; Kim, H.-G.; Tuller, H.L.; Kim, I.-D. Highly sensitive SnO2 hollow nanofiber-based NO2 gas sensors. Sens. Actuators B Chem. 2011, 160, 1468–1472. [Google Scholar] [CrossRef]
  155. Liu, Y.; Ding, Y.; Zhang, L.; Gao, P.-X.; Lei, Y. CeO2 nanofibers for in situ O2 and CO sensing in harsh environments. RSC Adv. 2012, 2, 5193–5198. [Google Scholar] [CrossRef]
  156. Liu, Y.; Ding, Y.; Gao, H.; Zhang, L.; Gao, P.; Li, B.; Lei, Y. La0.67Sr0.33MnO3 nanofibers for in situ, real-time, and stable high temperature oxygen sensing. RSC Adv. 2012, 2, 3872–3877. [Google Scholar] [CrossRef]
  157. Katoch, A.; Choi, S.-W.; Kim, H.W.; Kim, S.S. Highly sensitive and selective H2 sensing by ZnO nanofibers and the underlying sensing mechanism. J. Hazard. Mater. 2015, 286, 229–235. [Google Scholar] [CrossRef]
  158. Park, J.-A.; Moon, J.; Lee, S.-J.; Kim, S.H.; Zyung, T.; Chu, H.Y. Structure and CO gas sensing properties of electrospun TiO2 nanofibers. Mater. Lett. 2010, 64, 255–257. [Google Scholar] [CrossRef]
  159. Park, J.-A.; Moon, J.; Lee, S.-J.; Lim, S.-C.; Zyung, T. Fabrication and characterization of ZnO nanofibers by electrospinning. Curr. Appl. Phys. 2009, 9, S210–S212. [Google Scholar] [CrossRef]
  160. 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]
  161. Wei, S.; Zhao, G.; Du, W.; Tian, Q. Synthesis and excellent acetone sensing properties of porous WO3 nanofibers. Vacuum 2016, 124, 32–39. [Google Scholar] [CrossRef]
  162. Wei, S.; Zhou, M.; Du, W. Improved acetone sensing properties of ZnO hollow nanofibers by single capillary electrospinning. Sens. Actuators B Chem. 2011, 160, 753–759. [Google Scholar] [CrossRef]
  163. Liang, X.; Jin, G.; Liu, F.; Zhang, X.; An, S.; Ma, J.; Lu, G. Synthesis of In2O3 hollow nanofibers and their application in highly sensitive detection of acetone. Ceram. Int. 2015, 41, 13780–13787. [Google Scholar] [CrossRef]
  164. Zheng, W.; Li, Z.; Zhang, H.; Wang, W.; Wang, Y.; Wang, C. Electrospinning route for α-Fe2O3 ceramic nanofibers and their gas sensing properties. Mater. Res. Bull. 2009, 44, 1432–1436. [Google Scholar] [CrossRef]
  165. Zheng, W.; Lu, X.; Wang, W.; Li, Z.; Zhang, H.; Wang, Y.; Wang, Z.; Wang, C. A highly sensitive and fast-responding sensor based on electrospun In2O3 nanofibers. Sens. Actuators B Chem. 2009, 142, 61–65. [Google Scholar] [CrossRef]
  166. Liu, O.; Li, W.; Zhu, L.; Li, C.; Qu, F.; Guo, W.; Feng, C.; Ruan, S. Preparation and ethanol sensing properties of In2O3 nanotubes. J. Nanosci. Nanotechnol. 2014, 14, 3653–3657. [Google Scholar] [CrossRef]
  167. Yoon, J.-W.; Choi, J.-K.; Lee, J.-H. Design of a highly sensitive and selective C2H5OH sensor using p-type Co3O4 nanofibers. Sens. Actuators B Chem. 2012, 161, 570–577. [Google Scholar] [CrossRef]
  168. Cui, J.; Shi, L.; Xie, T.; Wang, D.; Lin, Y. UV-light illumination room temperature HCHO gas-sensing mechanism of ZnO with different nanostructures. Sens. Actuators B Chem. 2016, 227, 220–226. [Google Scholar] [CrossRef]
  169. Li, Z.; Fan, Y.; Zhan, J. In2O3 nanofibers and nanoribbons: Preparation by electrospinning and their formaldehyde gas-sensing properties. Eur. J. Inorg. Chem. 2010, 2010, 3348–3353. [Google Scholar] [CrossRef]
  170. Katoch, A.; Kim, J.H.; Kim, S.S. Significance of the nanograin size on the H2S-sensing ability of CuO-SnO2 composite nanofibers. J. Sens. 2015, 2015, 387641. [Google Scholar] [CrossRef] [Green Version]
  171. Wang, L.; Cao, J.; Qian, X.; Zhang, H. Facile synthesis and enhanced gas sensing properties of grain size-adjustable In2O3 micro/nanotubes. Mater. Lett. 2016, 171, 30–33. [Google Scholar] [CrossRef] [Green Version]
  172. Khalil, A.; Kim, J.J.; Tuller, H.L.; Rutledge, G.C.; Hashaikeh, R. Gas sensing behavior of electrospun nickel oxide nanofibers: Effect of morphology and microstructure. Sens. Actuators B Chem. 2016, 227, 54–64. [Google Scholar] [CrossRef]
  173. Katoch, A.; Abideen, Z.U.; Kim, J.-H.; Kim, S.S. Influence of hollowness variation on the gas-sensing properties of ZnO hollow nanofibers. Sens. Actuators B Chem. 2016, 232, 698–704. [Google Scholar] [CrossRef]
  174. Cho, S.; Kim, D.-H.; Lee, B.-S.; Jung, J.; Yu, W.-R.; Hong, S.-H.; Lee, S. Ethanol sensors based on ZnO nanotubes with controllable wall thickness via atomic layer deposition, an O2 plasma process and an annealing process. Sens. Actuators B Chem. 2012, 162, 300–306. [Google Scholar] [CrossRef]
  175. Yu, X.; Song, F.; Zhai, B.; Zheng, C.; Wang, Y. Electrospun ZnO nanotubes and its gas sensing applications. Phys. E Low-Dimens. Syst. Nanostructures 2013, 52, 92–96. [Google Scholar] [CrossRef]
  176. Cao, J.; Dou, H.; Zhang, H.; Mei, H.; Liu, S.; Fei, T.; Wang, R.; Wang, L.; Zhang, T. Controllable synthesis and HCHO-sensing properties of In2O3 micro/nanotubes with different diameters. Sens. Actuators B Chem. 2014, 198, 180–187. [Google Scholar] [CrossRef]
  177. Zhang, J.; Choi, S.-W.; Kim, S.S. Micro- and nano-scale hollow TiO2 fibers by coaxial electrospinning: Preparation and gas sensing. J. Solid State Chem. 2011, 184, 3008–3013. [Google Scholar] [CrossRef]
  178. Fu, J.; Zhang, J.; Peng, Y.; Zhao, C.; He, Y.; Zhang, Z.; Pan, X.; Mellors, N.J.; Xie, E. Wire-in-tube structure fabricated by single capillary electrospinning via nanoscale Kirkendall effect: The case of nickel-zinc ferrite. Nanoscale 2013, 5, 12551–12557. [Google Scholar] [CrossRef]
  179. Zhang, Z.; Yang, G.; Wei, J.; Bian, H.; Gao, J.; Li, J.; Wang, T. Morphology and magnetic properties of CoFe2O4 nanocables fabricated by electrospinning based on the Kirkendall effect. J. Cryst. Growth 2016, 445, 42–46. [Google Scholar] [CrossRef]
  180. Ji, W.; Wei, H.; Cui, Y.; Wei, Y.; Bu, J.; Feng, Z.; Wang, P.; Li, H.; Luo, L. Facile synthesis of porous forsterite nanofibres by direct electrospinning method based on the Kirkendall effect. Mater. Lett. 2018, 211, 319–322. [Google Scholar] [CrossRef]
  181. Kim, W.-S.; Lee, B.-S.; Kim, D.-H.; Kim, H.-C.; Yu, W.-R.; Hong, S.-H. SnO2 nanotubes fabricated using electrospinning and atomic layer deposition and their gas sensing performance. Nanotechnology 2010, 21, 245605. [Google Scholar] [CrossRef]
  182. Choi, S.-J.; Chattopadhyay, S.; Kim, J.J.; Kim, S.-J.; Tuller, H.L.; Rutledge, G.C.; Kim, I.-D. Coaxial electrospinning of WO3 nanotubes functionalized with bio-inspired Pd catalysts and their superior hydrogen sensing performance. Nanoscale 2016, 8, 9159–9166. [Google Scholar] [CrossRef] [Green Version]
  183. Chattopadhyay, S.; Saha, J.; De, G. Electrospun anatase TiO2 nanofibers with ordered mesoporosity. J. Mater. Chem. A 2014, 2, 19029–19035. [Google Scholar] [CrossRef]
  184. Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B.F.; Stucky, S.D. Nonionic triblock and star diblock copolymer and oligomeric surfactant syntheses of highly ordered, hydrothermally stable, mesoporous silica structures. J. Am. Chem. Soc. 1998, 120, 6024–6036. [Google Scholar] [CrossRef]
  185. Saha, J.; De, G. Highly ordered cubic mesoporous electrospun SiO2 nanofibers. Chem. Commun. 2013, 49, 6322–6324. [Google Scholar] [CrossRef] [PubMed]
  186. Korotcenkov, G. Gas response control through structural and chemical modification of metal oxides: State of the art and approaches. Sens. Actuators B Chem. 2005, 107, 209–232. [Google Scholar] [CrossRef]
  187. Korotcenkov, G.; Cho, B.K. Engineering approaches to improvement operating characteristics of conductometric gas sensors. Part 1: Improvement of sensor sensitivity and selectivity. Sens. Actuators B Chem. 2013, 188, 709–728. [Google Scholar] [CrossRef]
  188. Korotcenkov, G.; Brinzari, V.; Cho, B.K. Conductometric gas sensors based on metal oxide modified with gold nanoparticles: A review. Microchim. Acta 2016, 183, 1033–1054. [Google Scholar] [CrossRef]
  189. Katoch, A.; Byun, J.-H.; Choi, S.-W.; Kim, S.S. One-pot synthesis of Au-loaded SnO2 nanofibers and their gas sensing properties. Sens. Actuators B Chem. 2014, 202, 38–45. [Google Scholar] [CrossRef]
  190. Kou, X.; Meng, F.; Chen, K.; Wang, T.; Sun, P.; Liu, F.; Yan, X.; Sun, Y.; Liu, F.; Shimanoe, K.; et al. High-performance acetone gas sensor based on Ru-doped SnO2 nanofibers. Sens. Actuators B Chem. 2020, 320, 128292. [Google Scholar] [CrossRef]
  191. Jaroenapibal, P.; Boonma, P.; Suksilaporn, N.; Horprathum, M.; Amornkitbamrung, V.; Triroj, N. Improved NO2 sensing performance of electrospun WO3 nanofibers with silver doping. Sens. Actuators B Chem. 2018, 255, 1831–1840. [Google Scholar] [CrossRef]
  192. Dong, K.-Y.; Choi, J.-K.; Hwang, I.-S.; Lee, J.-W.; Kang, B.H.; Ham, D.-J.; Lee, J.-H.; Ju, B.-K. Enhanced H2S sensing characteristics of Pt doped SnO2 nanofibers sensors with micro heater. Sens. Actuators B Chem. 2011, 157, 154–161. [Google Scholar] [CrossRef]
  193. Kim, K.-H.; Kim, S.-J.; Cho, H.-J.; Kim, N.-H.; Jang, J.-S.; Choi, S.-J.; Kim, I.-D. WO3 nanofibers functionalized by protein-templated RuO2 nanoparticles as highly sensitive exhaled breath gas sensing layers. Sens. Actuators B Chem. 2017, 241, 1276–1282. [Google Scholar] [CrossRef]
  194. Cho, H.-J.; Kim, S.-J.; Choi, S.-J.; Jang, J.-S.; Kim, I.-D. Facile synthetic method of catalyst-loaded ZnO nanofibers composite sensor arrays using bio-inspired protein cages for pattern recognition of exhaled breath. Sens. Actuators B Chem. 2017, 243, 166–175. [Google Scholar] [CrossRef]
  195. Hu, Q.; Huang, B.; Li, Y.; Zhang, S.; Zhang, Y.; Hua, X.; Liu, G.; Li, B.; Zhou, J.; Xie, E.; et al. Methanol gas detection of electrospun CeO2 nanofibers by regulating Ce3+/Ce4+ mole ratio via Pd doping. Sens. Actuators B Chem. 2020, 307, 127638. [Google Scholar] [CrossRef]
  196. Zheng, W.; Lu, X.; Wang, W.; Li, Z.; Zhang, H.; Wang, Z.; Xu, X.; Li, S.; Wang, C. Assembly of Pt nanoparticles on electrospun In2O3 nanofibers for H2S detection. J. Colloid Interface Sci. 2009, 338, 366–370. [Google Scholar] [CrossRef]
  197. Wang, S.; Jia, F.; Wang, X.; Hu, L.; Sun, Y.; Yin, G.; Zhou, T.; Feng, Z.; Kumar, P.; Liu, B. Fabrication of ZnO nanoparticles modified by uniformly dispersed Ag nanoparticles: Enhancement of gas sensing performance. ACS Omega 2020, 5, 5209–5218. [Google Scholar] [CrossRef]
  198. Jang, J.-S.; Choi, S.-J.; Kim, S.-J.; Hakim, M.; Kim, I.-D. Rational design of highly porous SnO2 nanotubes functionalized with biomimetic nanocatalysts for direct observation of simulated diabetes. Adv. Funct. Mater. 2016, 26, 4740–4748. [Google Scholar] [CrossRef]
  199. Krivetskiy, V.; Ponzoni, A.; Comini, E.; Badalyan, S.; Rumyantseva, M.; Gaskov, A. Selectivity modification of SnO2-based materials for gas sensor arrays. Electroanalysis 2010, 22, 2809–2816. [Google Scholar] [CrossRef]
  200. Krivetskiy, V.V.; Rumyantseva, M.N.; Gaskov, A.M. Chemical modification of nanocrystalline tin dioxide for selective gas sensors. Russ. Chem. Rev. 2013, 82, 917–941. [Google Scholar] [CrossRef]
  201. Hubner, M.; Koziej, D.; Grunwaldt, J.-D.; Weimar, U.; Barsan, N. An Au clusters related spill-over sensitization mechanism in SnO2-based gas sensors identified by operando HERFD-XAS, work function changes, DC resistance and catalytic conversion studies. Phys. Chem. Chem. Phys. 2012, 14, 13249–13254. [Google Scholar] [CrossRef]
  202. Hübner, M.; Bârsan, N.; Weimar, U. Influences of Al, Pd and Pt additives on the conduction mechanism as well as the surface and bulk properties of SnO2 based polycrystalline thick film gas sensors. Sens. Actuators B Chem. 2012, 171, 172–180. [Google Scholar] [CrossRef]
  203. Korotcenkov, G.; Cho, B.K. Bulk doping influence on the response of conductometric SnO2 gas sensors: Understanding through cathodoluminescence study. Sens. Actuators B Chem. 2014, 196, 80–98. [Google Scholar] [CrossRef]
  204. Marikutsa, A.; Rumyantseva, M.; Konstantinova, E.; Shatalova, T.; Gaskov, A. Active sites on nanocrystalline tin dioxide surface: Effect of Palladium and Ruthenium oxides clusters. J. Phys. Chem. C 2014, 118, 21541–21549. [Google Scholar] [CrossRef]
  205. Degler, D.; Rank, R.; Müller, S.; Pereira de Carvalho, H.W.; Grunwaldt, J.D. Gold-loaded tin dioxide gas sensing materials: Mechanistic insights and the role of gold dispersion. ACS Sens. 2016, 1, 1322–1329. [Google Scholar] [CrossRef]
  206. Degler, D.; Müller, S.A.; Doronkin, D.E.; Wang, D.; Grunwaldt, J.D.; Weimar, U. Platinum loaded tin dioxide: A model system for unravelling the interplay between heterogeneous catalysis and gas sensing. J. Mater. Chem. A 2018, 6, 2034–2046. [Google Scholar] [CrossRef]
  207. Degler, D.; Weimar, U.; Barsan, N. Current understanding of the fundamental mechanisms of doped and loaded semiconducting metal-oxide-based gas sensing materials. ACS Sens. 2019, 4, 2228–2249. [Google Scholar] [CrossRef]
  208. Barbosa, M.S.; Suman, P.H.; Kim, J.J.; Tuller, H.L.; Orland, M.O. Investigation of electronic and chemical sensitization effects promoted by Pt and Pd nanoparticles on single-crystalline SnO nanobelt-based gas sensors. Sens. Actuators B Chem. 2019, 301, 127055. [Google Scholar] [CrossRef]
  209. Tang, W.; Wang, J.; Qiao, Q.; Liu, Z.; Li, X. Mechanism for acetone sensing property of Pd-loaded SnO2 nanofibers prepared by electrospinning: Fermi-level effects. J. Mater. Sci. 2015, 50, 2605–2615. [Google Scholar] [CrossRef]
  210. Lin, Y.; Wei, W.; Li, Y.; Li, F.; Zhou, J.; Sun, D.; Chen, Y.; Ruan, S. Preparation of Pd nanoparticle-decorated hollow SnO2 nanofibers and their enhanced formaldehyde sensing properties. J. Alloys Compd. 2015, 651, 690–698. [Google Scholar] [CrossRef]
  211. Wei, S.; Yu, Y.; Zhou, M. CO gas sensing of Pd-doped ZnO nanofibers synthesized by electrospinning method. Mater. Lett. 2010, 64, 2284–2286. [Google Scholar] [CrossRef]
  212. Moon, J.; Park, J.-A.; Lee, S.-J.; Zyung, T.; Kim, I.-D. Pd-doped TiO2 nanofiber networks for gas sensor applications Pd-doped TiO2 nanofiber networks for gas sensor applications. Sens. Actuators B Chem. 2010, 149, 301–305. [Google Scholar] [CrossRef]
  213. Kim, N.-H.; Choi, S.-J.; Yang, D.-J.; Bae, J.; Park, J.; Kim, I.-D. Highly sensitive and selective hydrogen sulfide and toluene sensors using Pd functionalized WO3 nanofibers for potential diagnosis of halitosis and lung cancer. Sens. Actuators B Chem. 2014, 193, 574–581. [Google Scholar] [CrossRef]
  214. Choi, S.-J.; Lee, I.; Jang, B.-H.; Youn, D.-Y.; Ryu, W.-H.; Park, C.O.; Kim, I.-D. Selective diagnosis of diabetes using Pt-functionalized WO3 hemitube networks as a sensing layer of acetone in exhaled breath. Anal. Chem. 2013, 85, 1792–1796. [Google Scholar] [CrossRef]
  215. Guo, L.; Xie, N.; Wang, C.; Kou, X.; Ding, M.; Zhang, H.; Sun, Y.; Song, H.; Wang, Y.; Lu, G. Enhanced hydrogen sulfide sensing properties of Pt-functionalized α-Fe2O3 nanowires prepared by one-step electrospinning. Sens. Actuators B Chem. 2018, 255, 1015–1023. [Google Scholar] [CrossRef]
  216. Cho, N.G.; Woo, H.-S.; Lee, J.-H.; Kim, I.-D. Thin-walled NiO tubes functionalized with catalytic Pt for highly selective C2H5OH sensors using electrospun fibers as a sacrificial template. Chem. Commun. 2011, 47, 11300–11302. [Google Scholar] [CrossRef]
  217. Kim, J.H.; Abideen, Z.U.; Zheng, Y.; Kim, S.S. Improvement of toluene-sensing performance of SnO2 nanofibers by Pt functionalization. Sensors 2016, 16, 1857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Xu, X.; Chen, Y.; Zhang, G.; Ma, S.; Lu, Y.; Bian, H.; Chen, Q. Highly sensitive VOCs-acetone sensor based on Ag-decorated SnO2 hollow nanofibers. J. Alloys Compd. 2017, 703, 572–579. [Google Scholar] [CrossRef]
  219. Ma, S.; Jia, J.; Tian, Y.; Cao, L.; Shi, S.; Li, X.; Wang, X. Improved H2S sensing properties of Ag/TiO2 nanofibers. Ceram. Intern. 2016, 42, 2041–2044. [Google Scholar] [CrossRef]
  220. Wang, J.; Zou, B.; Ruan, S.; Zhao, J.; Chen, Q.; Wu, F. HCHO sensing properties of Ag-doped In2O3 nanofibers synthesized by electrospinning. Mater. Lett. 2009, 63, 1750–1753. [Google Scholar] [CrossRef]
  221. Kim, J.H.; Zheng, Y.; Mirzaei, A.; Kim, S.S. Excellent carbon monoxide sensing performance of Au-decorated SnO2 nanofibers. Korean J. Mater. Res. 2016, 26, 741–750. [Google Scholar] [CrossRef]
  222. Xu, X.; Fan, H.; Liu, Y.; Wang, L.; Zhang, T. Au-loaded In2O3 nanofibers-based ethanol micro gas sensor with low power consumption. Sens. Actuators B Chem. 2011, 160, 713–719. [Google Scholar] [CrossRef]
  223. Yang, X.; Salles, V.; Kaneti, Y.V.; Liu, M.; Maillard, M.; Journet, C.; Jiang, X.; Brioude, A. Fabrication of highly sensitive gas sensor based on Au functionalized WO3 composite nanofibers by electrospinning. Sens. Actuators B Chem. 2015, 220, 1112–1119. [Google Scholar] [CrossRef]
  224. Matsushima, S.; Teraoka, Y.; Miura, N.; Yamazoe, N. Electronic interaction between metal additives and tin dioxide in tin dioxide-based gas sensors. Jpn. J. Appl. Phys. 1988, 51, 1798–1803. [Google Scholar] [CrossRef]
  225. Lee, Y.C.; Huang, H.; Tan, O.K.; Tse, M.S. Semiconductor gas sensor based on Pd-doped SnO2 nanorod thin films. Sens. Actuators B Chem. 2008, 132, 239–242. [Google Scholar] [CrossRef]
  226. Fryberger, T.D.; Semancik, S. Conductance response of Pd/SnO2 model gas sensors to H2 and O2. Sens. Actuators B Chem. 1990, 2, 305–319. [Google Scholar] [CrossRef]
  227. Zhao, C.; Bai, J.; Huang, B.; Wang, Y.; Zhou, J.; Xie, E. Grain refining effect of calcium dopants on gas-sensing properties of electrospun α-Fe2O3 nanotubes. Sens. Actuators B Chem. 2016, 231, 552–560. [Google Scholar] [CrossRef]
  228. Shan, H.; Liu, C.; Liu, L.; Li, S.; Wang, L.; Zhang, X.; Bo, X.; Chi, X. Highly sensitive acetone sensors based on La-doped-Fe2O3 nanotubes. Sens. Actuators B Chem. 2013, 184, 243–247. [Google Scholar] [CrossRef]
  229. Mohanapriya, P.; Segawa, H.; Watanabe, K.; Watanabe, K.; Samitsu, S.; Natarajan, T.; Jaya, N.V.; Ohashi, N. Enhanced ethanol-gas sensing performance of Ce-doped SnO2 hollow nanofibers prepared by electrospinning. Sens. Actuators B Chem. 2013, 188, 872–878. [Google Scholar] [CrossRef]
  230. Cheng, Y.; Wang, Y.; Zhang, J.; Li, H.; Liu, L.; Li, Y.; Du, L.; Duan, H. A comparison of Eu-doped α-Fe2O3 nanotubes and nanowires for acetone sensing. Nano 2017, 12, 1750138. [Google Scholar] [CrossRef]
  231. Li, W.; Ma, S.; Li, Y.; Li, X.; Wang, C.; Yang, X.; Cheng, L.; Mao, Y.; Luo, J.; Gengzang, D.; et al. Preparation of Pr-doped SnO2 hollow nanofibers by electrospinning method and their gas sensing properties. J. Alloys Compd. 2014, 605, 80–88. [Google Scholar] [CrossRef]
  232. Liu, Y.-J.; Zhang, H.-D.; Zhang, J.; Li, S.; Zhang, J.-C.; Zhu, J.-W. Effects of Ce doping and humidity on UV sensing properties of electrospun ZnO nanofibers. J. Appl. Phys. 2017, 122, 105102. [Google Scholar] [CrossRef]
  233. Cheng, L.; Ma, S.; Li, X.; Luo, J.; Li, W.; Li, F.; Mao, Y.; Wang, T.; Li, Y. Highly sensitive acetone sensors based on Y-doped SnO2 prismatic hollow nanofibers synthesized by electrospinning. Sens. Actuators B Chem. 2014, 200, 181–190. [Google Scholar] [CrossRef]
  234. Cheng, J.P.; Wang, B.B.; Zhao, M.G.; Liu, F.; Zhang, X.B. Nickel-doped tin oxide hollow nanofibers prepared by electrospinning for acetone sensing. Sens. Actuators B Chem. 2014, 190, 78–85. [Google Scholar] [CrossRef]
  235. Liu, L.; Li, S.; Zhuang, J.; Wang, L.; Zhang, J.; Li, H.; Liu, Z.; Han, Y.; Jiang, X.; Zhang, P. Improved selective acetone sensing properties of Co-doped ZnO nanofibers by electrospinning. Sens. Actuators B Chem. 2011, 155, 782–788. [Google Scholar] [CrossRef]
  236. Liu, L.; Zhang, T.; Wang, L.; Li, S. Improved ethanol sensing properties of Cu-doped SnO2 nanofibers. Mater. Lett. 2009, 63, 2041–2043. [Google Scholar] [CrossRef]
  237. Wang, Z.; Liu, L. Synthesis and ethanol sensing properties of Fe-doped SnO2 nanofibers. Mater. Lett. 2009, 63, 917–919. [Google Scholar] [CrossRef]
  238. Liu, Y.; Sun, X.; Li, B.; Lei, Y. Tunable p-n transition behaviour of a p-La0.67Sr0.33MnO3/n-CeO2 nanofibers heterojunction for the development of selective high temperature propane sensors. J. Mater. Chem. A 2014, 2, 11651–11659. [Google Scholar] [CrossRef]
  239. Bai, X.; Ji, H.; Gao, P.; Zhang, Y.; Sun, X. Morphology, phase structure and acetone sensitive properties of copper-doped tungsten oxide sensors. Sens. Actuators B Chem. 2014, 193, 100–106. [Google Scholar] [CrossRef]
  240. Qin, W.; Xu, L.; Song, J.; Xing, R.; Song, H. Highly enhanced gas sensing properties of porous SnO2-CeO2 composite nanofibers prepared by electrospinning. Sens. Actuators B Chem. 2013, 185, 231–237. [Google Scholar] [CrossRef]
  241. Gas’kov, A.; Rumyantseva, M. Metal oxide nanocomposites: Synthesis and characterization in relation with gas sensing phenomena. In Sensors for Environment, Health and Security; Baraton, M.I., Ed.; Springer Science + Business Media B.V.: Dordrecht, The Netherlands, 2009; pp. 3–29. [Google Scholar]
  242. Korotcenkov, G.; Boris, I.; Brinzari, V.; Han, S.H.; Cho, B.K. The role of doping effect on the response of SnO2-based thin film gas sensors: Analysis based on the results obtained for Co-doped SnO2 films deposited by spray pyrolysis. Sens. Actuators B Chem. 2013, 182, 112–124. [Google Scholar] [CrossRef]
  243. Korotcenkov, G.; Boris, I.; Cho, B.K. SnO2:Cu films doped during spray pyrolysis deposition: The reasons of the gas sensing properties change. Mater. Chem. Phys. 2013, 142, 124–131. [Google Scholar] [CrossRef]
  244. Rumyantseva, M.N.; Safonova, O.V.; Boulova, M.N.; Ryabova, L.I.; Gas’kov, A.M. Dopants in nanocrystalline tin dioxide. Russ. Chem. Bull. Intern. Ed. 2003, 52, 1217–1238. [Google Scholar] [CrossRef]
  245. Gerasimov, G.N.; Gromov, V.F.; Belysheva, T.V.; Ikim, M.; Trakhtenberg, L.I. Effect of the composition and structure of metal oxide nanocomposites on the sensor process when detecting reducing gases. Russ. J. Phys. Chem. 2017, 91, 1609–1620. [Google Scholar] [CrossRef]
  246. Wang, C.-N.; Li, Y.-L.; Gong, F.-L.; Zhang, Y.-H.; Fang, S.-M.; Zhang, H.-L. Advances in doped ZnO nanostructures for gas sensor. Chem. Rec. 2020, 20, 1553–1567. [Google Scholar] [CrossRef] [PubMed]
  247. Xu, L.; Xing, R.; Song, J.; Xu, W.; Song, H. ZnO-SnO2 nanotubes surface engineered by Ag nanoparticles: Synthesis, characterization, and highly enhanced HCHO gas sensing properties. J. Mater. Chem. C 2013, 1, 2174–2182. [Google Scholar] [CrossRef]
  248. Wan, G.X.; Ma, S.Y.; Sun, X.W.; Sun, A.M.; Li, X.B.; Luo, J.; Li, W.Q.; Wang, C.Y. Synthesis of wrinkled and porous ZnO-SnO2 hollow nanofibers and their gas sensing properties. Mater. Lett. 2015, 145, 48–51. [Google Scholar] [CrossRef]
  249. Katoch, A.; Kim, J.-H.; Kim, S.S. CuO/SnO2 Mixed Nanofibers for H2S Detection. J. Nanosci. Nanotechnol. 2015, 15, 8637–8641. [Google Scholar] [CrossRef]
  250. 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]
  251. Lu, Z.; Zhou, Q.; Wang, C.; Wei, Z.; Xu, L.; Gui, Y. Electrospun ZnO-SnO2 composite nanofibers and enhanced sensing properties to SF6 decomposition byproduct H2S. Front. Chem. 2018, 6, 540. [Google Scholar] [CrossRef] [Green Version]
  252. 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]
  253. Hsu, K.-C.; Fang, T.-H.; Tang, I.-T.; Hsiao, Y.-J.; Chen, C.-Y. Mechanism and characteristics of Au-functionalized SnO2/In2O3 nanofibers for highly sensitive CO detection. J. Alloys Compd. 2020, 822, 153475. [Google Scholar] [CrossRef]
  254. Gao, J.; Wang, L.; Kan, K.; Xu, S.; Jing, L.; Liu, S.; Shen, P.; Li, L.; Shi, K. One-step synthesis of mesoporous Al2O3-In2O3 nanofibres with remarkable gas-sensing performance to NOx at room temperature. J. Mater. Chem. A 2014, 2, 949–956. [Google Scholar] [CrossRef]
  255. Cao, J.; Wang, Z.Y.; Wang, R.; Liu, S.; Fei, T.; Wang, L.J. Synthesis of core-shell α-Fe2O3@NiO nanofibers with hollow structures and their enhanced HCHO sensing properties. J. Mater. Chem. A 2015, 3, 5635–5641. [Google Scholar] [CrossRef]
  256. Cao, J.; Wang, Z.; Wang, R.; Liu, S.; Fei, T.; Wang, L.; Zhang, T. Core-shell Co3O4/α-Fe2O3 heterostructure nanofibers with enhanced gas sensing properties. RSC Adv. 2015, 5, 36340–36346. [Google Scholar] [CrossRef]
  257. Liu, J.; Li, X.; Chen, X.; Niu, H.; Han, X.; Zhang, T.; Lin, H.; Qu, F. Synthesis of SnO2/In2O3 hetero-nanotubes by coaxial-electrospinning method for enhanced formaldehyde response. New J. Chem. 2016, 40, 1756–1764. [Google Scholar] [CrossRef]
  258. Deng, J.; Yu, B.; Lou, Z.; Wang, L.; Wang, R.; Zhang, T. Facile synthesis and enhanced ethanol sensing properties of the brush-like ZnO-TiO2 heterojunctions nanofibers. Sens. Actuators B Chem. 2013, 184, 21–26. [Google Scholar] [CrossRef]
  259. Wan, K.; Wang, D.; Wang, F.; Li, H.; Xu, J.; Wang, X.; Yang, J. Hierarchical In2O3@SnO2 core−shell nanofiber for high efficiency formaldehyde detection. ACS Appl. Mater. Interface 2019, 11, 45214–45225. [Google Scholar] [CrossRef]
  260. Lou, Z.; Li, F.; Deng, J.; Wang, L.; Zhang, T. Branch-like hierarchical heterostructure (α-Fe2O3/TiO2): A novel sensing material for trimethylamine gas sensors. ACS Appl. Mater. Interfaces 2013, 5, 12310–12316. [Google Scholar] [CrossRef] [PubMed]
  261. Qi, Q.; Wang, P.P.; Zhao, J.; Feng, L.L.; Zhou, L.J.; Xuan, R.F. SnO2 nanoparticle-coated In2O3 nanofibers with improved NH3 sensing properties. Sens. Actuators B Chem. 2014, 194, 440–446. [Google Scholar] [CrossRef]
  262. Zhang, N.; Li, F.; Yin, Y.; Han, J.; Li, X.; Liu, C.; Zhou, J.; Wen, S.; Adimi, S.; Chen, Y.; et al. Gas sensor based on TiO2 nanofibers decorated with monodispersed WO3 nanocubes for fast and selective xylene detection. Mater. Sci. Eng. B 2021, 263, 114901. [Google Scholar] [CrossRef]
  263. Frank, G.; Brock, L.; Bausen, H.D. The solubilities of Sn in In2O3 and of In in SnO2 crystals grown from Sn-In melts. J. Cryst. Growth 1976, 36, 179–180. [Google Scholar] [CrossRef]
  264. Enoki, H.; Echigoya, J.; Suto, H. The intermediate compound in the In2O3-SnO2 system. J. Mater. Sci. 1991, 26, 4110–4115. [Google Scholar] [CrossRef]
  265. Fitzgerald, C.B.; Venkatesan, M.; Douvalis, A.P.; Huber, S.; Coey, J.M.D. SnO2 doped with Mn, Fe or Co: Room temperature dilute magnetic semiconductors. J. Appl. Phys. 2004, 95, 7390. [Google Scholar] [CrossRef] [Green Version]
  266. Music, S. X-ray diffraction and Mössbauer spectra of the system Fe2O3-SnO2. J. Mater. Sci. Lett. 1991, 10, 197–200. [Google Scholar] [CrossRef]
  267. Lu, B.; Wang, C.; Zhang, Y. Electron beam induced crystallization in Fe-doped SnO2 nanoparticles. Appl. Phys. Lett. 1997, 70, 717–719. [Google Scholar] [CrossRef]
  268. Punnoose, A.; Hays, J.; Thurber, A.; Engelhard, M.H.; Kukkadapu, R.K.; Wang, C.; Shutthnandan, V.; Thevuthasan, S. Development of high-temperature ferromagnetism in SnO2 and paramagnetism in SnO by Fe doping. Phys. Rev. B 2005, 72, 054402. [Google Scholar] [CrossRef]
  269. Pandya, D.K.; Gopinadhan, K.; Kashyap, S.C. On the role of impurities on ferromagnetism in nanocrystalline SnO2:Ni thick films. Synth. React. Inorg. Met.-Org. Nano-Met. Chem. 2008, 38, 162–167. [Google Scholar]
  270. Aragon, F.H.; Coaquira, J.A.H.; Hidalgo, P.; da Silva, S.W.; Brito, S.L.M.; Gouvea, D.; Morais, P.C. Evidences of the evolution from solid solution to surface segregation in Ni-doped SnO2 nanoparticles using Raman spectroscopy. J. Raman Spectrosc. 2001, 42, 1081–1086. [Google Scholar] [CrossRef]
  271. Gouvea, D.; Varela, J.A.; Santilli, C.V.; Longp, E. Effect of Niobia on the sintering of SnO2. In Science of Sintering; Uskovic, D.P., Palmour, H., III, Spriggs, R.M., Eds.; Plenum Press: New York, NY, USA, 1989; pp. 529–536. [Google Scholar]
  272. Kim, B.; Jung, J.; Lee, J.; Kim, J. Precipitate concentration of Co2SnO4 in CoO-doped SnO2 ceramics at different oxygen chemical potentials. Solid State Ion. 2001, 144, 321–327. [Google Scholar] [CrossRef]
  273. Tena, M.A.; Meseguer, S.; Gargori, C.; Fores, A.; Badenes, J.A.; Monros, G. Study of Cr-SnO2 ceramic pigment and of Ti/Sn ratio on formation and coloration of these materials. J. Eur. Ceram. Soc. 2007, 27, 215–221. [Google Scholar] [CrossRef]
  274. Li, J.; Liu, S.; Pan, W. Co-doped tin oxide thin films by spin coating. Key Eng. Mater. 2008, 368–372, 524–525. [Google Scholar] [CrossRef]
  275. Cabezas, M.D.; Lamas, D.G.; Baby, R.E.; Cabanillas, E.; Walsöe de Reca, N.E. Nanostructured thick film sensors for CO(g) based on Al doped SnO2. In Anales de la Asociación Química Argentina; Asociación Química Argentina: Buenos Aires, Argentina, 2005; Volume 93, pp. 69–74. [Google Scholar]
  276. Tricoli, A.; Graf, M.; Pratsinis, S.E. Optimal doping for enhanced SnO2 sensitivity and thermal stability. Adv. Funct. Mater. 2008, 18, 1969–1976. [Google Scholar] [CrossRef]
  277. Giaquinta, D.M.; Davis, W.M.; Zur Loye, H.C. Structure of indium iron oxide. Acta Crystallogr. Sect. C Cryst. Struct. Commun. C 1994, 50, 5–7. [Google Scholar] [CrossRef]
  278. Yoo, Y.K.; Xue, Q.; Lee, H.-C.; Cheng, S.; Xiang, X.-D.; Dionne, G.F.; Xu, S.; He, J.; Chu, Y.S.; Preite, S.D.; et al. Bulk synthesis and high-temperature ferromagnetism of (In1−xFex)2O3−6 with Cu co-doping. Appl. Phys. Lett. 2005, 86, 042506. [Google Scholar] [CrossRef] [Green Version]
  279. Ratko, A.; Babushkin, O.; Baran, A.; Baran, S. Sorption and gas sensitive properties of In2O3 based ceramics doped with Ga2O3. J. Eur. Ceram. Soc. 1998, 18, 2227–2232. [Google Scholar] [CrossRef]
  280. Ohya, Y.; Ito, T.; Kaneko, M.; Ban, T.; Takahashi, Y. Solid solubility of SnO2 in In2O3. J. Ceram. Soc. Jpn. 2000, 108, 803–806. [Google Scholar] [CrossRef] [Green Version]
  281. Gonzalez, G.B.; Mason, T.O.; Okasinski, J.S.; Buslaps, T.; Honkimaki, V. Determination of the solubility of tin in indium oxide using in situ and ex situ X-ray diffraction. J. Am. Ceram. Soc. 2012, 95, 1809–1815. [Google Scholar] [CrossRef]
  282. Parthiban, S.; Gokulakrishnan, V.; Ramamurthi, K.; Elangovan, E.; Martins, R.; Fortunato, E.; Ganesan, R. High near-infrared transparent molybdenum-doped indium oxide thin films for nanocrystalline silicon solar cell applications. Sol. Energy Mater. Sol. Cells 2009, 93, 92–97. [Google Scholar] [CrossRef]
  283. Kaleemulla, S.; Rao, N.M.; Joshi, M.G.; Reddy, A.S.; Uthanna, S.; Reddy, P.S. Electrical and optical properties of In2O3:Mo thin films prepared at various Mo-doping levels. J. Alloys Compd. 2010, 504, 351–356. [Google Scholar] [CrossRef]
  284. Sunde, T.O.L.; Lindgren, M.; Mason, T.O.; Einarsrud, M.-A.; Grande, T. Solid solubility of rare earth elements (Nd, Eu, Tb) in In2−xSnxO3—effect on electrical conductivity and optical properties. Dalton Trans. 2014, 43, 9620–9632. [Google Scholar] [CrossRef] [Green Version]
  285. Berardan, A.; Guilmeau, E.; Maignan, A.; Raveau, B. In2O3:Ge: A promising n-type thermoelectric oxide composite. Solid State Commun. 2008, 146, 97–101. [Google Scholar] [CrossRef]
  286. Godzhieva, O.V.; Porotnikov, N.V.; Pikhidchuk, V.P. Physical-chemical study of triple oxides Formed in In2O3-CaO-CuO System. Zhur. Neorg. Khim. 1992, 37, 1184–1188. [Google Scholar]
  287. Horyn, R.; Bukowska, E.; Sikora, A. Studies of BaO-In2O3-CuO ternary system. Part I: Phase equilibria in the isothermal cross-section of 930 oC. J. Alloys Compd. 2000, 305, 103–108. [Google Scholar] [CrossRef]
  288. Mandal, S.K.; Das, A.K.; Nath, T.K.; Karmakar, D. Temperature dependence of solubility limits of transition metals (Co, Mn, Fe, and Ni) in ZnO nanoparticles. Appl. Phys. Lett. 2006, 89, 144105. [Google Scholar] [CrossRef]
  289. Morkoc, H.; Ozgur, U. Zinc Oxide; Wiley VCH: Weinheim, Germany, 2009. [Google Scholar]
  290. Jin, Z.; Fukumura, T.; Kawasaki, M.; Ando, K.; Saito, H.; Sekiguchi, T.; Yoo, Y.Z.; Murakami, M.; Matsumoto, Y.; Hasegawa, T.; et al. High throughput fabrication of transition-metal-doped epitaxial ZnO thin films: A series of oxide-diluted magnetic semiconductors and their properties. Appl. Phys. Lett. 2001, 78, 3824–3826. [Google Scholar] [CrossRef]
  291. Chahmat, N.; Haddad, A.; Ain-Souya, A.; Ganfoudi, R.; Attaf, N.; Salah Aida, M.; Ghers, M. Effect of Sn doping on the properties of ZnO thin films prepared by spray pyrolysis. J. Mod. Phys. 2012, 3, 1781–1785. [Google Scholar] [CrossRef]
  292. Serier, H.; Gaudon, M.; Manetrier, M. Al-doped ZnO powdered materials. Phys. Rev. B 2000, 61, 15019–15027. [Google Scholar]
  293. Yoon, M.H.; Lee, S.H.; Park, H.L.; Kim, H.K.; Jang, M.S. Solid solubility limits of Ga and Al in ZnO. J. Mater. Sci. Lett. 2002, 21, 1703–1704. [Google Scholar] [CrossRef]
  294. Park, D.H.; Son, K.Y.; Lee, J.H.; Kim, J.J.; Lee, J.S. Effect of ZnO addition in In2O3 ceramics: Defect chemistry and sintering behavior. Solid State Ion. 2004, 172, 431–434. [Google Scholar] [CrossRef]
  295. Park, J.Y.; Choi, S.W.; Lee, J.W.; Lee, C.; Kim, S.S. Synthesis and gas sensing properties of TiO2-ZnO core shell nanofibers. J. Am. Ceram. Soc. 2009, 92, 2551–2554. [Google Scholar] [CrossRef]
  296. Wu, H.; Kan, K.; Wang, L.; Zhang, G.; Yang, Y.; Li, H. Electrospinning of mesoporous p-type In2O3/TiO2 composite nanofibers for enhancing NOx gas sensing properties at room temperature. CrystEngComm 2014, 16, 9116–9124. [Google Scholar] [CrossRef]
  297. 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]
  298. Katoch, A.; Choi, S.W.; Kim, J.H.; Lee, J.H.; Lee, J.S.; Kim, S.S. Importance of the nanograin size on the H2S sensing properties of ZnO-CuO composite nanofibers. Sens. Actuators B Chem. 2015, 214, 111–116. [Google Scholar] [CrossRef]
  299. Katoch, A.; Choi, S.W.; Sun, G.J.; Kim, H.W.; Kim, S.S. Mechanism and prominent enhancement of sensing ability to reducing gases in p/n core-shell nanofiber. Nanotechnology 2014, 25, 175501–175508. [Google Scholar] [CrossRef]
  300. Lee, J.H.; Katoch, A.; Choi, S.W.; Kim, J.H.; Kim, H.W.; Kim, S.S. Extraordinary improvement of gas-sensing performances in SnO2 nanofibers due to creation of local p-n heterojunctions by loading reduced graphene oxide nanosheets. ACS Appl. Mater. Interfaces 2015, 7, 3101–3109. [Google Scholar] [CrossRef] [PubMed]
  301. Katoch, A.; Kim, J.H.; Kim, S.S. TiO2/ZnO inner/outer double-layer hollow fibers for improved detection of reducing gases. ACS Appl. Mater. Interfaces 2014, 6, 21494–21499. [Google Scholar] [CrossRef] [PubMed]
  302. Katoch, A.; Choi, S.W.; Sun, G.J.; Kim, S.S. An approach to detecting a reducing gas by radial modulation of electron-depleted shells in core-shell nanofibers. J. Mater. Chem. A 2013, 1, 13588–13596. [Google Scholar] [CrossRef]
  303. Yang, A.; Tao, X.; Wang, R.; Lee, S.; Surya, C. Room temperature gas sensing properties of SnO2/multiwall carbon-nanotube composite nanofibers. Appl. Phys. Lett. 2007, 91, 133110–133113. [Google Scholar] [CrossRef] [Green Version]
  304. Wang, Z.L.; Li, Z.; Sun, J.; Zhang, H.; Wang, W.; Zheng, W.; Wang, C. Improved hydrogen monitoring properties based on p-NiO/n-SnO2 heterojunction composite nanofibers. J. Phys. Chem. C 2010, 114, 6100–6105. [Google Scholar] [CrossRef]
  305. Feng, C.; Li, W.; Li, C.; Zhu, L.; Zhang, H.; Zhang, Y.; Ruan, S.; Chen, W.; Yu, L. Highly efficient rapid ethanol sensing based on In2-xNixO3 nanofibers. Sens. Actuators B Chem. 2012, 166, 83–88. [Google Scholar] [CrossRef]
  306. Zhang, X.J.; Qiao, G.J. High performance ethanol sensing films fabricated from ZnO and In2O3 nanofibers with a double-layer structure. Appl. Surf. Sci. 2012, 258, 6643–6647. [Google Scholar] [CrossRef]
  307. Yan, S.; Wu, Q.S. Micropored Sn-SnO2/carbon heterostructure nanofibers and their highly sensitive and selective C2H5OH gas sensing performance. Sens. Actuators B Chem. 2014, 205, 329–337. [Google Scholar] [CrossRef]
  308. Tang, W.; Wang, J.; Yao, P.J.; Li, X.G. Hollow hierarchical SnO2-ZnO composite nanofibers with heterostructure based on electrospinning method for detecting methanol. Sens. Actuators B Chem. 2014, 192, 543–549. [Google Scholar] [CrossRef]
  309. Wang, B.; Fu, X.X.; Liu, F.; Shi, S.L.; Cheng, J.P.; Zhang, X.B. Fabrication and gas sensing properties of hollow core-shell SnO2/α-Fe2O3 heterogeneous structures. J. Alloys Compd. 2014, 587, 82–89. [Google Scholar] [CrossRef]
  310. Zheng, Y.; Wang, J.; Yao, P. Formaldehyde sensing properties of electrospun NiO-doped SnO2 nanofibers. Sens. Actuators B Chem. 2011, 156, 723–730. [Google Scholar] [CrossRef]
  311. Du, H.; Wang, J.; Su, M.; Yao, P.; Zheng, Y.; Yu, N. Formaldehyde gas sensor based on SnO2/In2O3 hetero-nanofibers by a modified double jets electrospinning process. Sens. Actuators B Chem. 2012, 166, 746–752. [Google Scholar] [CrossRef]
  312. Ho, T.A.; Jun, T.S.; Kim, Y.S. Material and NH3-sensing properties of polypyrrole-coated tungsten oxide nanofibers. Sens. Actuators B Chem. 2013, 185, 523–529. [Google Scholar] [CrossRef]
  313. Li, Y.H.; Gong, J.; He, G.H.; Deng, Y.L. Fabrication of polyaniline/titanium dioxide composite nanofibers for gas sensing application. Mater. Chem. Phys. 2011, 129, 477–482. [Google Scholar] [CrossRef]
  314. Liu, L.; Zhang, Y.; Wang, G.G.; Li, S.C.; Wang, L.Y.; Han, Y. High toluene sensing properties of NiO-SnO2 composite nanofiber sensors operating at 330 °C. Sens. Actuators B Chem. 2011, 160, 448–454. [Google Scholar] [CrossRef]
  315. Qi, Q.; Zou, Y.C.; Fan, M.-H.; Liu, Y.P.; Gao, S.; Wang, P.P. Trimethylamine sensors with enhanced anti humidity ability fabricated from La0.7Sr0.3FeO3 coated In2O3-SnO2 composite nanofibers. Sens. Actuators B Chem. 2014, 203, 111–117. [Google Scholar] [CrossRef]
  316. Li, C.; Feng, C.H.; Qu, F.D.; Liu, J.; Zhu, L.H.; Lin, Y. Electrospun nanofibers of p-type NiO/n-type ZnO heterojunction with different NiO content and its influence on trimethylamine sensing properties. Sens. Actuators B Chem. 2015, 207, 90–96. [Google Scholar] [CrossRef]
  317. Lee, S.; Kim, I.D.; Lee, J.H. Selective and sensitive detection of trimethylamine using ZnO-In2O3 composite nanofibers. Sens. Actuators B Chem. 2013, 181, 463–470. [Google Scholar] [CrossRef]
  318. 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]
  319. 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]
  320. Gerasimov, G.N.; Gromov, V.F.; Ilegbusi, O.J.; Trakhtenberg, L.I. The mechanisms of sensory phenomena in binary metal-oxide nanocomposites. Sens. Actuators B Chem. 2017, 240, 613–624. [Google Scholar] [CrossRef]
  321. Zappa, D.; Galstyan, V.; Kaur, N.; Arachchige, H.M.M.; Sisman, O.; Comini, E. Metal oxide -based heterostructures for gas sensors—A review. Anal. Chim. Acta 2018, 1039, 1–23. [Google Scholar] [CrossRef] [PubMed]
  322. Bag, A.; Lee, N.-E. Gas sensing with heterostructures based on two-dimensional nanostructured materials: A review. J. Mater. Chem. C 2019, 7, 13367. [Google Scholar] [CrossRef]
  323. Jian, Y.; Hu, W.; Zhao, Z.; Cheng, P.; Haick, H.; Yao, M.; Wu, W. Gas sensors based on chemi-resistive hybrid functional nanomaterials. Nano-Micro Lett. 2020, 12, 71. [Google Scholar] [CrossRef] [Green Version]
  324. Park, S.; Ko, H.; Kim, S.; Lee, C. Role of the interfaces in multiple networked one-dimensional core-shell nanostructured gas sensors. ACS Appl. Mater. Interfaces 2014, 6, 9595–60023. [Google Scholar] [CrossRef]
  325. Redel, E.; Arsenault, E.; O’Brien, P.G.; Kherani, N.P.; Ozin, G.A. Growth of metal nanocrystals on nanostructured metal oxides—Dramatic effect of antimony doping. Chem. Mater. 2011, 23, 1353–1355. [Google Scholar] [CrossRef]
  326. Walker, J.M.; Akbar, S.A.; Morris, P.A. Synergistic effects in gas sensing semiconducting oxide nanohetero structures: A review. Sens. Actuators B Chem. 2019, 286, 624–640. [Google Scholar] [CrossRef]
  327. Du, H.; Wang, H.; Yao, P.; Wang, J.; Sun, Y. In2O3 nanofibers surface modified by low-temperature plasma and their gas sensing properties. Mater. Chem. Phys. 2018, 215, 316–326. [Google Scholar] [CrossRef]
  328. Kim, J.-H.; Mirzaei, A.; Kim, H.W.; Wue, P.; Kim, S.S. Design of supersensitive and selective ZnO-nanofiber-based sensors for H2 gas sensing by electron-beam irradiation. Sens. Actuators B Chem. 2019, 293, 210–223. [Google Scholar] [CrossRef]
  329. Chen, K.; Pi, K.; Zhang, D. Ultraviolet irradiation enhanced formaldehyde-sensing performance based on SnO2@TiO2 nanofiber heteroarchitectures. J. Phys. D Appl. Phys. 2020, 53, 125301. [Google Scholar] [CrossRef]
  330. Li, J.; Gu, D.; Yang, Y.; Du, H.; Li, X. UV light activated SnO2/ZnO nanofibers for gas sensing at room temperature. Front. Mater. 2019, 6, 158. [Google Scholar] [CrossRef] [Green Version]
  331. Cho, M.; Park, I. Recent trends of light-enhanced metal oxide gas sensors: Review. J. Sens. Sci. Technol. 2016, 25, 103–109. [Google Scholar] [CrossRef] [Green Version]
  332. Espid, E.; Taghipour, F. UV-LED photo-activated chemical gas sensors: A review. Crit. Rev. Solid State 2017, 42, 416–432. [Google Scholar] [CrossRef]
  333. Xu, F.; Ho, H.-P. Light-activated metal oxide gas sensors: A review. Micromachines 2017, 8, 333. [Google Scholar] [CrossRef] [Green Version]
  334. Kumar, R.; Liu, X.; Zhang, J.; Kumar, M. Room-temperature gas sensors under photoactivation: From metal oxides to 2D materials. Nano-Micro Lett. 2020, 12, 164. [Google Scholar] [CrossRef]
  335. Wang, J.; Shen, H.; Xia, Y.; Komarneni, S. Light-activated room-temperature gas sensors based on metal oxide nanostructures: A review on recent advances. Ceram Intern. 2021, 47, 7353–7368. [Google Scholar] [CrossRef]
  336. Suh, J.M.; Eom, T.H.; Cho, S.H.; Kim, T.; Jang, H.W. Light-activated gas sensing: Prospective of integration with micro-LEDs and plasmonic nanoparticles. Mater. Adv. 2021, 2, 827–844. [Google Scholar] [CrossRef]
  337. Prades, J.D.; Jimenes-Diaz, R.; Hernandez-Ramirez, F.; Fernandez-Romero, L.; Andreu, T.; Cirera, A.; Romano-Rodriguez, A.; Cornet, A.; Morante, J.R.; Barth, S.; et al. Toward a systematic understanding of photodetectors based on individual metal oxide nanowires. J. Phys. Chem. C 2008, 112, 14639–14644. [Google Scholar] [CrossRef]
  338. Comini, E.; Faglia, G.; Sberveglieri, G. Electrical-based gas sensing. In Solid State Gas Sensing; Comini, E., Faglia, G., Sberveglieri, G., Eds.; Springer: New York, NY, USA, 2009; pp. 47–107. [Google Scholar]
  339. Prades, J.D.; Jemenes-Diaz, R.; Manzanares, M.; Hernandez-Ramirez, F.; Cirera, A.; Romano-Rodriguez, A.; Marthur, S.; Morante, J.R. A model for the response towards oxidizing gases of photoactivated sensors based on individual SnO2 nanowires. Phys. Chem. Chem. Phys. 2009, 11, 10881–10889. [Google Scholar] [CrossRef]
  340. Liu, C.; Shan, H.; Liu, L.; Li, S.; Li, H. High sensing properties of Ce-doped α-Fe2O3 nanotubes to acetone. Ceram. Intern. 2014, 40, 2395–2399. [Google Scholar] [CrossRef]
  341. Seitz, C.; Werner, S.; Marschall, R.; Smarsly, B.M. Electrospun CuO nanofibre assemblies for H2S sensing. Zeitschrift Phys. Chem. 2019, 233, 105–116. [Google Scholar] [CrossRef]
  342. Kapse, V.D.; Ghosh, S.A.; Chaudhari, G.N.; Raghuwanshi, F.C. Nanocrystalline In2O3-based H2S sensors operable at low temperatures. Talanta 2008, 76, 610–616. [Google Scholar] [CrossRef]
  343. Huang, Z.M.; Zhang, Y.Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  344. Mondal, K.; Sharma, A. Recent advances in electrospun metal-oxide nanofiber based interfaces for electrochemical biosensing. RSC Adv. 2016, 6, 94595–94616. [Google Scholar] [CrossRef]
  345. Lee, J.A.; Krogman, K.C.; Ma, M.; Hill, R.M.; Hammond, P.T.; Rutledge, G.C. Highly reactive multilayer-assembled TiO2 coating on electrospun polymer nanofibers. Adv. Mater. 2009, 21, 1252–1256. [Google Scholar] [CrossRef]
  346. Sahner, K.; Tuller, H.L. Novel deposition techniques for metal oxide: Prospects for gas sensing. J. Electroceram. 2010, 24, 177–199. [Google Scholar] [CrossRef]
  347. Meng, Y.; Liu, G.; Liu, A.; Guo, Z.; Sun, W.; Shan, F. Photochemical activation of electrospun In2O3 nanofibers for high performance electronic devices. ACS Appl. Mater. Interfaces 2017, 9, 10805–10812. [Google Scholar] [CrossRef]
  348. Kim, I.-D.; Rothschild, A.; Lee, B.H.; Kim, D.Y.; Jo, S.M.; Tuller, H.L. Ultrasensitive chemiresistors based on electrospun TiO2 nanofibers. Nano Lett. 2006, 6, 2009–2013. [Google Scholar] [CrossRef] [PubMed]
  349. Cui, Y.; Meng, Y.; Wang, Z.; Wang, C.; Liu, G.; Martins, R.; Fortunato, E.; Shan, F. High performance electronic devices based on nanofibers via crosslinking welding process. Nanoscale 2018, 10, 19427–19434. [Google Scholar] [CrossRef] [PubMed]
  350. Zhang, H.; Meng, Y.; Song, L.; Luo, L.; Qin, Y.; Han, N.; Yang, Z.; Liu, L.; Ho, J.C.; Wang, F. High-performance enhancement-mode thin-film transistors based on Mg-doped In2O3 nanofiber networks. Nano Res. 2018, 11, 1227–1237. [Google Scholar] [CrossRef]
  351. Persano, L.; Camposeo, A.; Tekmen, C.; Pisignano, D. Industrial upscaling of electrospinning and applications of polymer nanofibers: A review. Macromol. Mater. Eng. 2013, 298, 504–520. [Google Scholar] [CrossRef]
  352. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and electrospun nanofibers: Methods, materials, and applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef] [PubMed]
  353. Thompson, C.J.; Chase, G.G.; Yarin, A.L.; Reneker, D.H. Effects of parameters on nanofiber diameter determined from electrospinning model. Polymer 2007, 48, 6913–6922. [Google Scholar] [CrossRef]
  354. Demir, M.M.; Yilgor, I.; Yilgor, E.; Erman, B. Electrospinning of polyurethane fibers. Polymer 2002, 43, 3303–3309. [Google Scholar] [CrossRef]
  355. Bisht, G.S.; Canton, G.; Mirsepassi, A.; Kulinsky, L.; Oh, S.; Dunn-Rankin, D.; Madou, M.J. Controlled continuous patterning of polymeric nanofibers on three-dimensional substrates using low-voltage near-field electrospinning. Nano Lett. 2011, 11, 1831–1837. [Google Scholar] [CrossRef]
  356. Huang, Y.; Bu, N.; Duan, Y.; Pan, Y.; Liu, H.; Yin, Z.; Xiong, Y. Electrohydrodynamic direct-writing. Nanoscale 2013, 5, 12007–12017. [Google Scholar] [CrossRef] [PubMed]
  357. Lei, T.P.; Lu, X.Z.; Yang, F. Fabrication of various micro/nano structures by modified near-field electrospinning. AIP Adv. 2015, 5, 041301. [Google Scholar] [CrossRef] [Green Version]
  358. Fuh, Y.K.; Hsu, H.S. Fabrication of monolithic polymer nanofluidic channels via near-field electrospun nanofibers as sacrificial templates. J. Micro/Nanolithograph MEMS MOEMS 2011, 10, 043004. [Google Scholar] [CrossRef]
  359. Wang, X.; Zheng, G.; Xu, L.; Cheng, W.; Xu, B.; Huang, Y.; Sun, D. Fabrication of nanochannels via near-field electrospinning. Appl. Phys. A 2012, 108, 825–828. [Google Scholar] [CrossRef] [Green Version]
  360. Chang, J.; Liu, Y.M.; Heo, K.; Lee, B.Y.; Lee, S.W.; Lin, L.W. Direct-write complementary graphene field effect transistors and junctions via near-field electrospinning. Small 2014, 10, 1920–1925. [Google Scholar] [CrossRef] [PubMed]
  361. Luo, K.S.; Liu, T.Y.; Zang, X.; Wen, Z.; Lin, L. Direct-write, self-aligned electrospinning on paper for controllable fabrication of three-dimensional structures. ACS Appl. Mater. Interfaces 2015, 7, 27765–27770. [Google Scholar] [CrossRef] [Green Version]
  362. Kim, J.; Maeng, B.; Park, J. Characterization of 3D electrospinning on inkjet printed conductive pattern on paper. Micro Nano Syst. Lett. 2018, 6, 12. [Google Scholar] [CrossRef] [Green Version]
  363. Park, Y.-S.; Kim, J.; Oh, J.M.; Park, S.; Cho, S.; Ko, H.; Cho, Y.-K. Near-field electrospinning for three-dimensional stacked nanoarchitectures with high aspect ratios. Nano Lett. 2020, 20, 441–448. [Google Scholar] [CrossRef]
  364. Brown, T.D.; Dalton, P.D.; Hutmacher, D.W. Direct writing by way of melt electrospinning. Adv. Mater. 2011, 23, 5651–5657. [Google Scholar] [CrossRef] [PubMed]
  365. Carlberg, B.; Wang, T.; Liu, J. Direct photolithographic patterning of electrospun films for defined nanofibrillar microarchitectures. Langmuir 2010, 26, 2235–2239. [Google Scholar] [CrossRef]
  366. Sharma, C.S.; Sharma, A.; Madou, M. Multiscale carbon structures fabricated by direct micropatterning of electrospun mats of SU-8 photoresist nanofibers. Langmuir 2010, 26, 2218–2222. [Google Scholar] [CrossRef] [PubMed]
  367. Jia, C.; Yu, D.; Lamarre, M.; Leopold, P.L.; Teng, Y.D.; Wang, H. Patterned electrospun nanofiber matrices via localized dissolution: Potential for guided tissue formation. Adv. Mater. 2014, 26, 8192–8197. [Google Scholar] [CrossRef] [PubMed]
  368. Jun, I.; Chung, Y.-W.; Heo, Y.-H.; Han, H.-S.; Park, J.; Jeong, H.; Lee, H.; Lee, Y.B.; Kim, Y.-C.; Seok, H.-K. Creating hierarchical topographies on fibrous platforms using femtosecond laser ablation for directing myoblasts behavior. ACS Appl. Mater. Interfaces 2016, 8, 3407–3417. [Google Scholar] [CrossRef]
  369. Wang, S.-X.; Yap, C.C.; He, J.; Chen, C.; Wong, S.Y.; Li, X. Electrospinning: A facile technique for fabricating functional nanofibers for environmental applications. Nanotechnol. Rev. 2016, 5, 51–73. [Google Scholar] [CrossRef] [Green Version]
  370. Jiang, G.; Qin, X. An improved free surface electrospinning for high throughput manufacturing of core–shell nanofibers. Mater. Lett. 2014, 128, 259–262. [Google Scholar] [CrossRef]
  371. Aliheidari, N.; Aliahmad, N.; Agarwal, M.; Dalir, H. Electrospun nanofibers for label-free sensor applications. Sensors 2019, 19, 3587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  372. Kang, K.; Yang, D.; Park, J.; Kim, S.; Cho, I.; Yang, H.-H.; Cho, M.; Mousavi, S.; Choi, K.H.; Park, I. Micropatterning of metal oxide nanofibers by electrohydrodynamic (EHD) printing towards highly integrated and multiplexed gas sensor applications. Sens. Actuators B Chem. 2017, 250, 574–583. [Google Scholar] [CrossRef]
  373. Yan, S.; Zan, G.; Wu, Q. An ultrahigh-sensitivity and selective sensing material for ethanol: α-/γ-Fe2O3 mixed-phase mesoporous nanofibers. Nano Res. 2015, 8, 3673–3686. [Google Scholar] [CrossRef]
  374. 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, 2015, 972025. [Google Scholar] [CrossRef] [Green Version]
  375. Li, Z.; Li, H.; Wu, Z.; Wang, M.; Luo, J.; Torun, H.; Hu, P.A.; Yang, C.; Grundmann, M.; Liu, X.; et al. Advances in designs and mechanisms of semiconducting metal oxide nanostructures for high-precision gas sensors operated at room temperature. Mater. Horiz. 2019, 6, 470–506. [Google Scholar] [CrossRef] [Green Version]
  376. Mirzaei, A.; Kim, H.W.; Kim, S.S.; Neri, G. Nanostructured semiconducting metal oxide gas sensors for acetaldehyde detection. Chemosensors 2019, 7, 56. [Google Scholar] [CrossRef] [Green Version]
  377. Han, D.; Steckl, A.J. Coaxial electrospinning formation of complex polymer fibers and their applications. ChemPlusChem 2019, 84, 1453–1497. [Google Scholar] [CrossRef] [PubMed]
  378. Sill, T.J.; von Recum, H.A. Electrospinning: Applications in drug delivery and tissue engineering. Biomaterials 2008, 29, 1989–2006. [Google Scholar] [CrossRef]
  379. Qu, H.; Wei, S.; Guo, Z. Coaxial electrospun nanostructures and their applications. J. Mater. Chem. A 2013, 1, 11513–11528. [Google Scholar] [CrossRef]
Figure 1. Morphology of WO3 (a) nanofibers and (b) films, fabricated using similar parameters. Deposition of W for 180 s, and oxidation at 700 °C in air for 2 h. Reprinted with permission from [27]. Copyright 2012 Royal Society of Chemistry.
Figure 1. Morphology of WO3 (a) nanofibers and (b) films, fabricated using similar parameters. Deposition of W for 180 s, and oxidation at 700 °C in air for 2 h. Reprinted with permission from [27]. Copyright 2012 Royal Society of Chemistry.
Nanomaterials 11 01555 g001
Figure 2. (a,b) Schematic illustrations of the fabrication process of metal oxide nanofiber-based sensor with (a) Au and (b) Ti-Au/Pt electrodes deposited on the nanofibers. (a) Reprinted from [78].; (b) Reprinted with permission from [79]. Copyright 2015 Elsevier.
Figure 2. (a,b) Schematic illustrations of the fabrication process of metal oxide nanofiber-based sensor with (a) Au and (b) Ti-Au/Pt electrodes deposited on the nanofibers. (a) Reprinted from [78].; (b) Reprinted with permission from [79]. Copyright 2015 Elsevier.
Nanomaterials 11 01555 g002
Figure 3. Electrospun ZnO nanobridges with lengths from (a) 20 µm to (b) 150 µm. A broken fiber is depicted to show that bridge formation is not a trivial process and the temperature ramp affects sintering. Reprinted from [83].
Figure 3. Electrospun ZnO nanobridges with lengths from (a) 20 µm to (b) 150 µm. A broken fiber is depicted to show that bridge formation is not a trivial process and the temperature ramp affects sintering. Reprinted from [83].
Nanomaterials 11 01555 g003
Figure 4. (a) Response of pure TiO2 single nanofiber triangular sample in the presence of CO gas as a function of concentration at 200−350 °C. (b) The response of TiO2:Au triangular samples in the presence of CO gas as a function of concentration in ppt range at 250 °C. Reprinted with permission from [86]. Copyright 2017 ACS.
Figure 4. (a) Response of pure TiO2 single nanofiber triangular sample in the presence of CO gas as a function of concentration at 200−350 °C. (b) The response of TiO2:Au triangular samples in the presence of CO gas as a function of concentration in ppt range at 250 °C. Reprinted with permission from [86]. Copyright 2017 ACS.
Nanomaterials 11 01555 g004
Figure 5. Different approaches to prepare electrospun nanofibers: (a) conventional approach, (b,e) obtaining aligned electrospun nanofibers (b) by a high-speed rotating drum and (e) by a sharp-edged rotating disk; (c,d) SEM images of (c) random and (d) aligned nanofibers. (a-d) Reprinted with permission from [96]. Copyrights 2013 Wiley; (e) Reprinted with permission from [97]. Copyright 2001 IOP.
Figure 5. Different approaches to prepare electrospun nanofibers: (a) conventional approach, (b,e) obtaining aligned electrospun nanofibers (b) by a high-speed rotating drum and (e) by a sharp-edged rotating disk; (c,d) SEM images of (c) random and (d) aligned nanofibers. (a-d) Reprinted with permission from [96]. Copyrights 2013 Wiley; (e) Reprinted with permission from [97]. Copyright 2001 IOP.
Nanomaterials 11 01555 g005
Figure 6. (a) Schematic illustration of the setup for collecting nanofibers as a uniaxially aligned array. The collector contains an insulating void, such as the air gap between two strips of silicon wafers. (b) Electrostatic force analysis of a charged nanofiber spanning across two silicon strips. The orientation of the nanofiber is mainly controlled by the stretching force originating from the attractive electrostatic forces. (cf) SEM images of uniaxially aligned nanofibers made of (c) carbon, (d) anatase TiO2, (e) NiFe2O4 and (f) TiO2/PVP. Reprinted with permission from [101]. Copyright 2003 American Chemical Society.
Figure 6. (a) Schematic illustration of the setup for collecting nanofibers as a uniaxially aligned array. The collector contains an insulating void, such as the air gap between two strips of silicon wafers. (b) Electrostatic force analysis of a charged nanofiber spanning across two silicon strips. The orientation of the nanofiber is mainly controlled by the stretching force originating from the attractive electrostatic forces. (cf) SEM images of uniaxially aligned nanofibers made of (c) carbon, (d) anatase TiO2, (e) NiFe2O4 and (f) TiO2/PVP. Reprinted with permission from [101]. Copyright 2003 American Chemical Society.
Nanomaterials 11 01555 g006
Figure 8. (a) Response–temperature curves of the four kinds of electrospun WO3-based sensors to 3 ppm NO2 at various temperatures; (b) the corresponding response and recovery times of all sensors to NO2 listed in Table 2. Toper = 90 °C. Reprinted with permission from [139]. Copyright 2021 Elsevier.
Figure 8. (a) Response–temperature curves of the four kinds of electrospun WO3-based sensors to 3 ppm NO2 at various temperatures; (b) the corresponding response and recovery times of all sensors to NO2 listed in Table 2. Toper = 90 °C. Reprinted with permission from [139]. Copyright 2021 Elsevier.
Nanomaterials 11 01555 g008
Figure 9. Influence of annealing (a) temperature (400–800 °C) and (b) time (0.5–48 h) on the crystallite size in (a) In2O3 and (b) SnO2:Cu nanofibers synthesized by electrospinning. (a) Data extracted from [171]; (b) Reprinted from [170].
Figure 9. Influence of annealing (a) temperature (400–800 °C) and (b) time (0.5–48 h) on the crystallite size in (a) In2O3 and (b) SnO2:Cu nanofibers synthesized by electrospinning. (a) Data extracted from [171]; (b) Reprinted from [170].
Nanomaterials 11 01555 g009
Figure 10. (a) Responses of In2O3 nanofiber-based sensors to 50 ppm formaldehyde as a function of operating temperature and the temperature of calcination (t = 3 h, Tc = 400–800 °C). Reprinted with permission from [171]. Copyright 2016 Elsevier. (b) Conductivity response to NH3 (5–25 ppm) at room temperature for four types of gas sensors based on In2O3 nanostructures: 1—broken In2O3 hollow nanofiber; 2—regular In2O3 hollow nanofiber; 3—In2O3 nanofiber; 4—In2O3 nanoparticles. Adapted with permission from [140]. Copyrights 2007 Wiley.
Figure 10. (a) Responses of In2O3 nanofiber-based sensors to 50 ppm formaldehyde as a function of operating temperature and the temperature of calcination (t = 3 h, Tc = 400–800 °C). Reprinted with permission from [171]. Copyright 2016 Elsevier. (b) Conductivity response to NH3 (5–25 ppm) at room temperature for four types of gas sensors based on In2O3 nanostructures: 1—broken In2O3 hollow nanofiber; 2—regular In2O3 hollow nanofiber; 3—In2O3 nanofiber; 4—In2O3 nanoparticles. Adapted with permission from [140]. Copyrights 2007 Wiley.
Nanomaterials 11 01555 g010
Figure 11. Effect of wall thickness of hollow SnO2 nanofibers on the (a) magnitude of the response to 100 ppm ethanol and (b) response time as a function of operation temperature; (c) dependence of the SnO2 film thickness on the number of PEALD cycles. Adapted with permission from [181]. Copyright 2010 IOP.
Figure 11. Effect of wall thickness of hollow SnO2 nanofibers on the (a) magnitude of the response to 100 ppm ethanol and (b) response time as a function of operation temperature; (c) dependence of the SnO2 film thickness on the number of PEALD cycles. Adapted with permission from [181]. Copyright 2010 IOP.
Nanomaterials 11 01555 g011
Figure 12. (ac) Schematic illustrations of coaxial electrospinning using mineral oil in the core and composite solution in the shell. Reprinted with permission from [182]. Copyright 2016 Royal Society of Chemistry.
Figure 12. (ac) Schematic illustrations of coaxial electrospinning using mineral oil in the core and composite solution in the shell. Reprinted with permission from [182]. Copyright 2016 Royal Society of Chemistry.
Nanomaterials 11 01555 g012
Figure 13. (a) Responses of 1.7 at% Au-loaded SnO2 nanofibers at Toper = 300 °C to various reducing gases (5 ppm), and (b,c) typical TEM images of SnO2:Au nanofibers. Gold clusters had sizes in the range of 5–10 nm. Adapted with permission from [189]. Copyright 2014 Elsevier.
Figure 13. (a) Responses of 1.7 at% Au-loaded SnO2 nanofibers at Toper = 300 °C to various reducing gases (5 ppm), and (b,c) typical TEM images of SnO2:Au nanofibers. Gold clusters had sizes in the range of 5–10 nm. Adapted with permission from [189]. Copyright 2014 Elsevier.
Nanomaterials 11 01555 g013
Figure 14. (a) Response of sensors based on pure and 1, 2 and 3 mol% Ru-doped SnO2 nanofibers to 100 ppm acetone as a function of the operating temperature. (b) Influence of Ag doping on the response to 5 ppm NO2 of electrospun WO3:Ag-based sensors. (a) Reprinted with permission from [190]. Copyright 2020: Elsevier; (b) Reprinted with permission from [191]. Copyright 2018 Elsevier.
Figure 14. (a) Response of sensors based on pure and 1, 2 and 3 mol% Ru-doped SnO2 nanofibers to 100 ppm acetone as a function of the operating temperature. (b) Influence of Ag doping on the response to 5 ppm NO2 of electrospun WO3:Ag-based sensors. (a) Reprinted with permission from [190]. Copyright 2020: Elsevier; (b) Reprinted with permission from [191]. Copyright 2018 Elsevier.
Nanomaterials 11 01555 g014
Figure 15. Conductometric response of CuO-SnO2 composite nanofibers with crystallites of different sizes to 10 ppm H2S gas. Adapted from [170].
Figure 15. Conductometric response of CuO-SnO2 composite nanofibers with crystallites of different sizes to 10 ppm H2S gas. Adapted from [170].
Nanomaterials 11 01555 g015
Figure 16. (a) Responses of gas sensors based on pristine and Ce-doped (3, 4 and 5 at %) α-Fe2O3 hollow nanofibers to 50 ppm acetone under different temperatures. (b) Response values of the pure, 3, 6 and 9 mol% Cu-doped WO3 hollow fibers to 20 ppm of acetone at Toper = 300 °C. (a) Adapted with permission from [238]. Copyright 2014 RSC; (b) Adapted with permission from [239]. Copyright 2014 Elsevier.
Figure 16. (a) Responses of gas sensors based on pristine and Ce-doped (3, 4 and 5 at %) α-Fe2O3 hollow nanofibers to 50 ppm acetone under different temperatures. (b) Response values of the pure, 3, 6 and 9 mol% Cu-doped WO3 hollow fibers to 20 ppm of acetone at Toper = 300 °C. (a) Adapted with permission from [238]. Copyright 2014 RSC; (b) Adapted with permission from [239]. Copyright 2014 Elsevier.
Nanomaterials 11 01555 g016
Figure 17. The selective test of SnO2:Ce nanofiber-based sensors toward (a) 20 ppm H2S and 200 ppm ethanol, CO, H2, NH3 and formaldehyde at Toper = 210 °C, and (b) 200 ppm H2S, ethanol, CO, H2, NH3 and formaldehyde at Toper = 370 °C. Reprinted with permission from [240]. Copyright 2013 Elsevier.
Figure 17. The selective test of SnO2:Ce nanofiber-based sensors toward (a) 20 ppm H2S and 200 ppm ethanol, CO, H2, NH3 and formaldehyde at Toper = 210 °C, and (b) 200 ppm H2S, ethanol, CO, H2, NH3 and formaldehyde at Toper = 370 °C. Reprinted with permission from [240]. Copyright 2013 Elsevier.
Nanomaterials 11 01555 g017
Figure 18. Scheme of mutual distribution of components in MIO−MIIO nanocomposites. Possible transformation of the crystallite size and the grain structure of the major components of the nanocomposite is also shown in the figure. x1 and x4 correspond to the solubility limits of MeII in MeIO and MeI in MeIIO. Reprinted with permission from [60]. Copyright 2017 Elsevier. Idea from Refs. [51,241].
Figure 18. Scheme of mutual distribution of components in MIO−MIIO nanocomposites. Possible transformation of the crystallite size and the grain structure of the major components of the nanocomposite is also shown in the figure. x1 and x4 correspond to the solubility limits of MeII in MeIO and MeI in MeIIO. Reprinted with permission from [60]. Copyright 2017 Elsevier. Idea from Refs. [51,241].
Nanomaterials 11 01555 g018
Figure 19. Choice of the second component of SnO2-based nanocomposite with allowance for the properties of the detected gas. Reprinted with permission from [60]. Copyright 2017 Elsevier. Idea and data from Ref. [51].
Figure 19. Choice of the second component of SnO2-based nanocomposite with allowance for the properties of the detected gas. Reprinted with permission from [60]. Copyright 2017 Elsevier. Idea and data from Ref. [51].
Nanomaterials 11 01555 g019
Figure 20. Response–time curves of In2O3 nanofiber sensors decorated with SnO2 to 1 ppm NH3 at room temperature: SnO2/In2O3—1–7.5 at.% Sn; SnO2/In2O3—2–16 at.% Sn; and SnO2/In2O3—3–21 at.% Sn. Reprinted with permission from [261]. Copyright 2014 Elsevier.
Figure 20. Response–time curves of In2O3 nanofiber sensors decorated with SnO2 to 1 ppm NH3 at room temperature: SnO2/In2O3—1–7.5 at.% Sn; SnO2/In2O3—2–16 at.% Sn; and SnO2/In2O3—3–21 at.% Sn. Reprinted with permission from [261]. Copyright 2014 Elsevier.
Nanomaterials 11 01555 g020
Figure 21. (a) Responses of the sensors based on NiO HNF, α-Fe2O3 NF and α-Fe2O3/NiO CSNF to HCHO gas (Toper = 240 °C); (b) the selectivity of α-Fe2O3/NiO CSNF, α-Fe2O3 NF and NiO HNF to 50 ppm different gases ((A) ethyne, (B) ammonia, (C) trichloromethane, (D) methylbenzene, (E) ethanol, (F) formaldehyde). Reprinted with permission from [255]. Copyright 2015 RSC.
Figure 21. (a) Responses of the sensors based on NiO HNF, α-Fe2O3 NF and α-Fe2O3/NiO CSNF to HCHO gas (Toper = 240 °C); (b) the selectivity of α-Fe2O3/NiO CSNF, α-Fe2O3 NF and NiO HNF to 50 ppm different gases ((A) ethyne, (B) ammonia, (C) trichloromethane, (D) methylbenzene, (E) ethanol, (F) formaldehyde). Reprinted with permission from [255]. Copyright 2015 RSC.
Nanomaterials 11 01555 g021
Figure 22. Response of ZnO NF sensors irradiated at different e-beam doses to 0.1, 1 and 10 ppm H2 at Toper = 350 °C. Reprinted with permission from [328]. Copyright 2019 Elsevier.
Figure 22. Response of ZnO NF sensors irradiated at different e-beam doses to 0.1, 1 and 10 ppm H2 at Toper = 350 °C. Reprinted with permission from [328]. Copyright 2019 Elsevier.
Nanomaterials 11 01555 g022
Figure 23. The response of TiO2 nanofiber sensor under dark and UV illumination conditions to several gases with concentration of 75 ppm. Reprinted with permission from [141]. Copyright 2015 Elsevier.
Figure 23. The response of TiO2 nanofiber sensor under dark and UV illumination conditions to several gases with concentration of 75 ppm. Reprinted with permission from [141]. Copyright 2015 Elsevier.
Nanomaterials 11 01555 g023
Figure 24. The response and recovery times of TiO2 nanofiber sensor in 50 ppm hydrogen gas as a function of temperature (a) with UV and (b) without UV irradiation. Reprinted with permission from [141]. Copyright 2015 Elsevier.
Figure 24. The response and recovery times of TiO2 nanofiber sensor in 50 ppm hydrogen gas as a function of temperature (a) with UV and (b) without UV irradiation. Reprinted with permission from [141]. Copyright 2015 Elsevier.
Nanomaterials 11 01555 g024
Figure 25. (a) Long-term stability of gas sensor based on 4 at.% Ce-doped α-Fe2O3 nanotubes to 10, 50, 100 and 200 ppm acetone at 240 °C. (b) Stability test of the 3% Pd-CeO2 nanofiber-based sensor to 100 ppm methanol at 200 °C. (a) Reprinted with permission from [340]. Copyright 2014 Elsevier; (b) Reprinted with permission from [195]. Copyright 2020 Elsevier.
Figure 25. (a) Long-term stability of gas sensor based on 4 at.% Ce-doped α-Fe2O3 nanotubes to 10, 50, 100 and 200 ppm acetone at 240 °C. (b) Stability test of the 3% Pd-CeO2 nanofiber-based sensor to 100 ppm methanol at 200 °C. (a) Reprinted with permission from [340]. Copyright 2014 Elsevier; (b) Reprinted with permission from [195]. Copyright 2020 Elsevier.
Nanomaterials 11 01555 g025
Figure 26. SEM images of CuO nanofibers after different numbers of H2S (10 ppm) sensing cycles (A → D). As-prepared CuO nanofibers had a diameter of 600–700 nm. Reprinted with permission from [341]. Copyright 2019 De Gruyter.
Figure 26. SEM images of CuO nanofibers after different numbers of H2S (10 ppm) sensing cycles (A → D). As-prepared CuO nanofibers had a diameter of 600–700 nm. Reprinted with permission from [341]. Copyright 2019 De Gruyter.
Nanomaterials 11 01555 g026
Figure 27. Gas responses of CuO-loaded In2O3 nanofiber sensors toward various gases at (a) 150 °C and (b) 300 °C. The response corresponds to Ra/Rg ratio for reducing gases and to Rg/Ra for NO2. Reprinted with permission from [297]. Copyright 2015 Elsevier.
Figure 27. Gas responses of CuO-loaded In2O3 nanofiber sensors toward various gases at (a) 150 °C and (b) 300 °C. The response corresponds to Ra/Rg ratio for reducing gases and to Rg/Ra for NO2. Reprinted with permission from [297]. Copyright 2015 Elsevier.
Nanomaterials 11 01555 g027
Figure 29. (ac) Schematic illustration of welding process for nanofiber networks. (d) Contact resistance variation at the nanojoint. Adapted with permission from [349]. Copyright 2018 Royal Society of Chemistry.
Figure 29. (ac) Schematic illustration of welding process for nanofiber networks. (d) Contact resistance variation at the nanojoint. Adapted with permission from [349]. Copyright 2018 Royal Society of Chemistry.
Nanomaterials 11 01555 g029
Figure 30. Influence of fiber diameter (a) on the response of WO3 NF-based sensor to 150 ppb NO gas, and (b) on the initial resistances of the sensors. Reprinted with permission from [27]. Copyright 2012 RSC.
Figure 30. Influence of fiber diameter (a) on the response of WO3 NF-based sensor to 150 ppb NO gas, and (b) on the initial resistances of the sensors. Reprinted with permission from [27]. Copyright 2012 RSC.
Nanomaterials 11 01555 g030
Figure 31. Optical micrographs of gas sensor test devices (10 × 15 mm) with TiO2 nanofiber mats after different processing steps. Reprinted with permission from [348]. Copyright 2006 ACS.
Figure 31. Optical micrographs of gas sensor test devices (10 × 15 mm) with TiO2 nanofiber mats after different processing steps. Reprinted with permission from [348]. Copyright 2006 ACS.
Nanomaterials 11 01555 g031
Figure 32. The deposition areas when electrospinning at different parameters. The applied voltage and electrode-to-collector distance are: (a) 10 kV and L = 2.6 cm; (b) 10 kV and L = 2.1 cm; (c) 10 kV and L = 1.7 cm; (d) 7 kV and L = 1.2 cm; (e) 6 kV and L = 0.8 cm; (f) 5 kV and L = 0.5 cm. Reprinted with permission from [105]. Copyright 2008 Elsevier.
Figure 32. The deposition areas when electrospinning at different parameters. The applied voltage and electrode-to-collector distance are: (a) 10 kV and L = 2.6 cm; (b) 10 kV and L = 2.1 cm; (c) 10 kV and L = 1.7 cm; (d) 7 kV and L = 1.2 cm; (e) 6 kV and L = 0.8 cm; (f) 5 kV and L = 0.5 cm. Reprinted with permission from [105]. Copyright 2008 Elsevier.
Nanomaterials 11 01555 g032
Figure 33. (a–c) SEM images of SnO2 nanofibers with different magnifications (Tan = 700 °C). Reprinted with permission from [105]. Copyright 2008 Elsevier.
Figure 33. (a–c) SEM images of SnO2 nanofibers with different magnifications (Tan = 700 °C). Reprinted with permission from [105]. Copyright 2008 Elsevier.
Nanomaterials 11 01555 g033
Figure 34. (a) Scheme of free coaxial electrospinning apparatus using a stepped pyramid spinneret; (b) a picture of coaxial jets in electrospinning process. Reprinted with permission from [370]. Copyright 2014 Elsevier.
Figure 34. (a) Scheme of free coaxial electrospinning apparatus using a stepped pyramid spinneret; (b) a picture of coaxial jets in electrospinning process. Reprinted with permission from [370]. Copyright 2014 Elsevier.
Nanomaterials 11 01555 g034
Figure 35. Schematic diagram illustrating the fabrication procedure of an array of (a) nonaligned and (b) quasi-aligned hollow ZnO fibers. Reprinted with permission from [138]. Copyright 2009 ACS.
Figure 35. Schematic diagram illustrating the fabrication procedure of an array of (a) nonaligned and (b) quasi-aligned hollow ZnO fibers. Reprinted with permission from [138]. Copyright 2009 ACS.
Nanomaterials 11 01555 g035
Figure 36. (a) MEMS gas sensor array fabricated by EHD printing of SnO2, WO3 and In2O3 nanofibers for low power consumption. The platform size is 3.5 × 3.5 mm. The configuration and SEM image of the individual element are also given here. (bd) SEM images of nanofiber materials integrated on the suspended MEMS platform by EHD printing. Reprinted with permission from [372]. Copyright 2017 Elsevier.
Figure 36. (a) MEMS gas sensor array fabricated by EHD printing of SnO2, WO3 and In2O3 nanofibers for low power consumption. The platform size is 3.5 × 3.5 mm. The configuration and SEM image of the individual element are also given here. (bd) SEM images of nanofiber materials integrated on the suspended MEMS platform by EHD printing. Reprinted with permission from [372]. Copyright 2017 Elsevier.
Nanomaterials 11 01555 g036
Figure 37. Fabrication of highly integrated gas sensor array by EHD printing of electrospun nanofibers: (a) metal oxide nanofiber fragments are prepared by electrospinning and fragmentation process (e.g., ultrasonication); (b) individual gas sensor and highly integrated gas sensor array were fabricated by micro-patterning of heterogeneous metal oxide nanofibers via sequential or parallel EHD printing process. Reprinted with permission from [372]. Copyright 2017 Elsevier.
Figure 37. Fabrication of highly integrated gas sensor array by EHD printing of electrospun nanofibers: (a) metal oxide nanofiber fragments are prepared by electrospinning and fragmentation process (e.g., ultrasonication); (b) individual gas sensor and highly integrated gas sensor array were fabricated by micro-patterning of heterogeneous metal oxide nanofibers via sequential or parallel EHD printing process. Reprinted with permission from [372]. Copyright 2017 Elsevier.
Nanomaterials 11 01555 g037
Table 4. Influence of doping with Ce (6 mol%) on the parameters of SnO2 nanofibers and their conductometric response to ethanol (50 ppm) at Toper = 250 °C.
Table 4. Influence of doping with Ce (6 mol%) on the parameters of SnO2 nanofibers and their conductometric response to ethanol (50 ppm) at Toper = 250 °C.
SampleFiber Diameter, nmCrystallite Size, nmSurface Area, m2/gResponse
SnO2~2344116.7~20
SnO2:Ce~1272735~260
Source: data extracted from [229].
Table 5. Influence of additives (in oxide form) in metal oxide matrix on structural and gas sensing characteristics of SnO2- and In2O3-based sensors. Influence on the electrophysical characteristics, i.e., acceptor or donor behavior of the additives, is not considered.
Table 5. Influence of additives (in oxide form) in metal oxide matrix on structural and gas sensing characteristics of SnO2- and In2O3-based sensors. Influence on the electrophysical characteristics, i.e., acceptor or donor behavior of the additives, is not considered.
AdditiveEffectNature
Al2O3; SiO2Increases sensor response;
improves thermal stability
Decrease in crystallite size; decrease in the area of intergrain contacts; increase in porosity
Ag (Ag2O); Cu (Cu2O)Increases response to H2S, SO2Two-phase system; phase transformations during gas detection
Fe (Fe2O3)Increases response to alcoholsChange in oxidation state
Ga(Ga2O3); Zn(ZnO)Increases sensor responseDecrease in crystallite size; increase in porosity
P, BImproves selectivityCreation of new phase
SeIncreases sensor responseIncrease in porosity
Ca; K; Rb; MgIncreases sensor response;
improves thermal stability
Decrease in crystallite size
La; Ba; Y; CeImproves thermal stability;
increases sensor response
Stabilization of crystallite size (creation of new phases); decrease in crystallite size
Transition MOXs: Co; Mn; Sr; NiIncreases sensor response;
improves selectivity
Catalytic effect; change in electron concentration;
change in A/D parameters; change in crystallite size
Source: Reprinted with permission from [60]. Copyright 2017 Elsevier.
Table 8. Influence of oxygen plasma treatment on the parameters of In2O3 fibers, and conductometric response to acetone.
Table 8. Influence of oxygen plasma treatment on the parameters of In2O3 fibers, and conductometric response to acetone.
SampleFiber Diameter, nmCrystallite Size, nmSurface Area, m2/gToper, °CS (100 ppm)τres, s
In2O3~1003518.2275~723
In2O3-O2~1702332.5275~2027
Source: data extracted from [327].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Korotcenkov, G. Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 2: Gas Sensors and Their Advantages and Limitations. Nanomaterials 2021, 11, 1555. https://doi.org/10.3390/nano11061555

AMA Style

Korotcenkov G. Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 2: Gas Sensors and Their Advantages and Limitations. Nanomaterials. 2021; 11(6):1555. https://doi.org/10.3390/nano11061555

Chicago/Turabian Style

Korotcenkov, Ghenadii. 2021. "Electrospun Metal Oxide Nanofibers and Their Conductometric Gas Sensor Application. Part 2: Gas Sensors and Their Advantages and Limitations" Nanomaterials 11, no. 6: 1555. https://doi.org/10.3390/nano11061555

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