Next Article in Journal
A Vehicle-to-Grid System for Controlling Parameters of Microgrid System
Previous Article in Journal
LoRaWAN Meets ML: A Survey on Enhancing Performance with Machine Learning
Previous Article in Special Issue
A Study on the Gas/Humidity Sensitivity of the High-Frequency SAW CO Gas Sensor Based on Noble-Metal-Modified Metal Oxide Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Road Map of Semiconductor Metal-Oxide-Based Sensors: A Review

by
Taposhree Dutta
1,
Tanzila Noushin
2,
Shawana Tabassum
3 and
Satyendra K. Mishra
4,5,*
1
Department of Chemistry, IIEST Shibpur, Howrah 711103, West Bengal, India
2
Department of Electrical and Computer Engineering, The University of Texas at Dallas, Richardson, TX 75080, USA
3
Department of Electrical Engineering, The University of Texas at Tyler, Tyler, TX 75799, USA
4
Danish Offshore Technology Center, Technical University of Denmark, 2800 Lyngby, Denmark
5
SRCOM, Centre Technologic de Telecomunicacions de Catalunya, 08860 Castelldefels, Barcelona, Spain
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(15), 6849; https://doi.org/10.3390/s23156849
Submission received: 18 April 2023 / Revised: 22 June 2023 / Accepted: 19 July 2023 / Published: 1 August 2023
(This article belongs to the Special Issue Gas Sensors for Environmental Applications)

Abstract

:
Identifying disease biomarkers and detecting hazardous, explosive, flammable, and polluting gases and chemicals with extremely sensitive and selective sensor devices remains a challenging and time-consuming research challenge. Due to their exceptional characteristics, semiconducting metal oxides (SMOxs) have received a lot of attention in terms of the development of various types of sensors in recent years. The key performance indicators of SMOx-based sensors are their sensitivity, selectivity, recovery time, and steady response over time. SMOx-based sensors are discussed in this review based on their different properties. Surface properties of the functional material, such as its (nano)structure, morphology, and crystallinity, greatly influence sensor performance. A few examples of the complicated and poorly understood processes involved in SMOx sensing systems are adsorption and chemisorption, charge transfers, and oxygen migration. The future prospects of SMOx-based gas sensors, chemical sensors, and biological sensors are also discussed.

1. Introduction

In sensing applications, metal oxides (MOxs (metal oxides), mainly II–VI semiconductors) are mostly used due to their inexpensiveness, ease of manufacture, quick response time, wide detection range, and resistance to harsh conditions [1,2]. For sensors to be effective, (i) there must be a charge transfer between the analytes and sensing materials and (ii) the measurement must have an analyte concentration dependence [3,4]. In addition, there is a need for efficient and effective methods for detecting volatile, chemical, and biological compounds and molecules. Analytical chemistry methods, such as spectrophotometry, fluorometry, gas chromatography (GC), and high-performance liquid chromatography (HPLC), were previously used to detect these molecules accurately. Furthermore, these methods were heavy, expensive, had low throughput, time-consuming pretreatment steps, required highly skilled operators and significant power consumption, and did not provide real-time information for risk reduction or decision-making [5]. As a result of these limitations, most present sensing methods rely on SMOx-based materials for sensing applications, such as 1D and 2D field effect transistors (FET) and the Internet of things (IoT) [3,4,6,7,8].
Compounds with high levels of ionic bonding or electrostatic interaction are called semiconductor metal oxides (SMOxs) [9,10]. In recent years, they have attracted a great deal of attention due to their excellent sensing capabilities, adaptability in terms of size, ease of manufacture, and low power consumption. By tuning the size and composition of the materials, the electrical, optical, mechanical, catalytic, and magnetic properties can be modulated. Because of their large surface-area-to-volume ratio, SMOxs, which have a size range between 1 and 100 nm, exhibit unique physical and chemical properties [11,12,13,14,15,16]. In addition to the size and geometry of the SMOx material, electron transport is also influenced by it [17,18]. Adding dopants, impurities, composite structures, or metal additives can further manipulate the charge transfer in a SMOx. In addition to improving the long-term stability and selectivity, these components lower the operation temperature, energy consumption, and humidity interference of pristine SMOx [19,20,21,22,23]. There are two types of SMOx materials: n-type and p-type, depending on the type of dopants. In SMOx, oxygen vacancies serve as electron donors, and hence, these compounds are n-type. Conversely, p-type SMOx, such as CuO, Co3O4, Cr2O3, and NiO, have metal ions and are electron acceptors. As a result of their tunable sensitivity, selectivity, and response time, SMOx-based materials are widely used in sensors that detect gases, chemicals, and biomolecules. With SMOx-based sensors, a wide range of gases can be detected, such as oxidizing gases and reducing gases [17,23].
Here, we present a brief review of the fundamentals and sensing properties of SMOx materials; the factors that influence their sensing performance; and their applications in gas, chemical, and biological sensing. There was a variety of literature available on SMOx gas sensors [24,25,26,27,28,29,30,31] and biosensors [32]. This article reviewed different aspects of SMOx materials and their sensing properties in one comprehensive article. This review provides researchers with a better understanding of the fundamentals and sensing applications of SMOx, enabling them to develop the next generation of SMOx-based sensors.

2. Fundamentals of Semiconductor Metal Oxides

The easy charge transfer properties of metal oxides (MOxs) make them unique among semiconducting materials. This effect is due to the large electronegativity difference, and thus, the high degree of ionic bonding, between the metal and oxygen that the MOx has. MOx has a conduction band minimum (CBM) and valence band maximum (VBM) of metal (M) ns and oxygen (O) 2p orbitals, respectively. Metals (Ms) and oxygen (O) have highly dispersed or localized orbitals (ns and 2p). Furthermore, metal oxides have a much higher dispersive valence band maximum (VBM) than n-type semiconductors. As an example, In2O3, ZnO, SnO2, and their hybrid composites function as n-type MOx, while NiO and Cu2O are p-type MOx. The first-known p-type transparent conductive oxide (TCO) was nickel oxide (NiO) [33].
The mobility (μ) of a carrier is inversely proportional to its effective mass (m*) and is given by the following equation:
μ = e τ/m*
where τ is the free carrier scattering time
Controlled physical and chemical properties of metal oxides, including structural defects, morphology, grain size, and specific surface area, enable them to be used in a range of applications, such as catalysis, sensors, energy conversion, and environmental monitoring [34,35].

3. Properties of Semiconductor-Metal-Oxide-Based Sensors

Regarding the greenhouse effect, MOx-based sensors are used for the rapid detection of harmful and toxic gases, where the low concentration (in ppm or ppb) of the target gas is converted into a measurable electrical, optical, or magnetic signal. In these sensors, metal oxide semiconductors and metal oxide–polymer composites are used to produce excellent sensitivity [36,37,38]. As a result, semiconducting metal oxides (SMOxs) can be used to detect low gas concentrations with a high sensitivity and rapid response. A SMOx-based sensor is characterized by its low cost, rapid response and recovery time, high stability, simple electronic interface, and low maintenance, making it an ideal and promising material for detecting toxic gases [5,31,39,40]. Materials made of SMOx are ionic solids, which are held together by strong ionic bonds between positive metallic and negative oxygen ions. Semiconductor metal oxides (SMOxs) have filled electronic shells, making them more thermally and chemically stable than free metal oxides. Incomplete electronic shells d endow optical properties, such as high dielectric constants [39,41,42,43]. Depending on the design, a SMOx can be flexible; porous; and can be in a zero-dimensional shape (0D), a 1D shape, a 2D shape, or a 3D shape [44,45,46]. As the temperature increases, SMOx materials’ conductivity (and hence resistance) changes. Moreover, optical, electrical, and magnetic fields affect the conductivity of SMOx.
Understanding semiconductor metal oxides (SMOxs) is crucial to developing sensors with high sensitivity. SMOx sensing properties are affected by physical factors, such as crystalline structure, defects, energy bands, impurities, charge transport, and p-n junction formation. In a SMOx, charge transfer can be controlled by doping with donor materials [47,48,49].

3.1. Crystalline Structure with Defects

Due to noble metal doping, materials with high crystalline structures have been investigated to develop sensors. SMOx can be classified into two types of crystalline structure: monocrystalline and polycrystalline. A monocrystalline structure is formed by a regular arrangement of atoms. Conversely, polycrystalline structures consist of small single crystals arranged randomly. On the other hand, non-crystals possess irregular shapes with short-range structural order [50,51,52]. As an example, the surface of a crystal of SnO2 usually lacks one or more atoms, resulting in abundant unsaturated bonds. Consequently, SnO2 exhibits high chemical activity and participates in redox reactions [53,54]. SnO2 has a tetragonal crystal structure [55,56], while ZnO has a hexagonal structure [44,57]. Semiconductor metal oxides (SMOxs) have special crystal structures that influence their physiochemical activity (Table 1).
There are various types of defects in SMOx, including point defects, line defects, plane defects, and volume defects. SMOx’s physio-chemical activity can be enhanced by partial defects caused by impurities [101,102]. Photoelectric activity can induce point defects, also called 0D defects [103,104]. Another type of defect is a line defect, which is caused by partial crystal slides. There are two types of dislocation defects: closed rings and surface defects. There are also planar defects, which can include angular grain boundaries, stack layer faults, and twin crystals. In the crystal matrix, volume defects are voids with different structures, densities, and chemical compositions [105,106].

3.2. Energy Band of SMOx in the Presence of Impurities

In addition, the sensing property of n-type semiconductors and p-type semiconductors depends on the energy band structure of semiconductor metal oxides (SMOxs). When the SMOx thickness reaches a level comparable to the depletion layer width, the energy band is no longer constrained to the surface but is affected by a significant number of grains, which, in turn, affects the electronic structure and electron–hole charge carriers [107,108,109]. In general, an electron’s conduction energy band becomes vacant when the minimum band gap energy (Eg) of the SMOx is reached (Figure 1). As a result, the valence band is left with holes. Electrons (e) and holes (h+) are easily mobilized in the presence of an external electric field, while at low energy, electron–hole pairs (e + h+) are electrostatically bound [110,111]. Since In2O3 has a small effective mass of electrons, its band structure shows a highly dispersive CBM. Its optical bandgap is 3.7 eV [112]. The presence of impurities induces intra-band electron transitions, such as electrons moving from defect states to ground states. By adjusting the size, shape, and composition of impurities, intraband gaps can be modulated. A SMOx sensor with a large Eg can work at high temperatures, which indicates that SMOx sensors are thermally stable. When the operating temperature exceeds 300 °C, gas sensors should have a band gap greater than 2.5 eV. SMOx-based gas sensors have a weakly dependent chemical activity on ambient humidity [111,113,114] (Figure 1). In some core–shell semiconductors, for instance, the conduction and valence bands of the core and shell are staggered, and electrons and holes are separated. It was found that the conduction band energy was the lowest in the shell and highest in the core. The energy band offsets in semiconductor materials segregate electrons from the shell and holes from the core, allowing carrier recombination across the interface at a lower energy than any of their constituent band gaps [110,114]. Electron-saturation velocities are high, heterojunctions are readily available, gaps are broad, and breakdown fields are large, allowing for fast and very sensitive gas detection systems to operate. The size and shape of semiconductor materials can be controlled by applying strain due to quantum confinement phenomena. It is possible to adjust the bandgap range of semiconducting nanostructures due to their high elastic limit. The band structure governs the adsorption of light, charge separation, and recombination of charge, which determines the use of a SMOx in photoelectric conversion [109,114].
Different electron–hole carriers produced by doped SMOx affect the conductivity [108,115]. There are two types of doped semiconductors: n-type and p-type [116,117,118]. The doping of materials regulates their conductivity and mass transfer, which is extremely important for gas sensors.

3.3. Carrier Transportation and Electronic Structure of SMOx

Conductivity in semiconductor metal oxides is affected by the production of free carriers, e.g., electrons in n-type semiconductors and holes in p-type semiconductors [119,120,121]. A stable concentration of conductive electrons and holes is maintained at thermal equilibrium [50,106,108]. Metal oxide’s electronic structure, temperature, applied electronic field, doping, and lack of structural order in the material can influence carrier transport mechanisms, such as drift, diffusion, and recombination [122,123]. The movement of carriers from a high concentration to a low concentration is called diffusion [124,125]. Also, the carrier recombination rate affects the carrier lifetime and gas-sensing properties [110,126].
Its d valence bond orbitals impart unique physical and chemical properties for various applications, such as gas sensing [29,46].

3.4. Formation of p-n Junctions in a SMOx

Generally, p-n junctions form between semiconductors with different electronic structures [127,128]. A hole diffuses from a p-type semiconductor to a n-type semiconductor, leaving negatively charged ions on the p-type semiconductor. The n-type semiconductor, however, loses free electrons, leaving positively charged ions. p-n junctions possess unidirectional conductivity since the ions cannot diffuse and form a zone of space charge at the interface. A p-n junction affects the electronic, optical, and magnetic properties of SMOx materials [103,129]. It is possible to control the properties of SMOx to design it for a wide range of applications, such as gas sensing, catalysis, and energy storage. C. Han and co-workers presented hollow nanofibers based on p-CuO/n-ZnO for gas sensing [130]. Using atomic layer deposition (ALD), electrospun heterostructures were fabricated to investigate the effect of the composition on gas sensing. As the concentration is increased, the response rate slowly decreases, with Rzn/cu = 15.6. Compared with pure ZnO and pure CuO, these heterostructures exhibit 6 and 45 times higher responses to H2S gas, respectively (Figure 2).
According to Dhawale and co-workers, LPG detectability changes based on resistance or barrier height with a transition metal oxide [131]. The amount of chemisorbed oxygen on the surface; charges on the surface; diffusion; and other processes, such as gas adsorption and desorption, control the electrical resistance or barrier height (Figure 3). By adsorbing oxygen from the surrounding air on the film surfaces, ionic species like O2-(ads), O-(ads), and O2-(ads) are formed. Ionic species trap electrons from the valence band (topmost) and remove them from films. In consequence, these adsorbed oxygen species reduce the conductivity of n-type TiO2. Whenever the chemisorption equilibrium is upset, the resistance or the barrier height of TiO2 changes [132]. Since LPG is a reducing gas containing components such as CH4, C3H8, and C4H10, gas-sensing mechanisms for LPG become more complicated [133]. During the exposure of LPG gas to a TiO2 sensor, chemisorbed oxygen releases trapped electrons back onto the TiO2 surface, resulting in a drastic reduction in the electrical resistance and barrier height. It is possible to increase the gas response by adding noble metals to metal oxide surfaces [134]. Painting Pd nanoparticles on TiO2 improves the response of LPG over that of pristine TiO2. The surface energy changes when Pd is added to TiO2, and a spillover effect occurs [135]. Because of the weak interaction between the Pd atom and the oxygen gas, the Pd:TiO2 sensor requires a relatively low temperature to dissolve [136]. However, a remarkably significant number of electrons are injected back into the topmost conduction band of TiO2, thereby increasing the conductivity. Active Pd nanoparticle catalysts improve the LPG response by speeding up the process and providing more active sites. Furthermore, Lee and co-workers designed a hollow cube nanostructure with ZnO and CuO cores for acetone sensing, using the CuO (a p-type material) as a catalyst [134]. The n-type ZnO and p-type CuO domains produced a consistent p-n junction. When the two materials were still connected, charge conduction took place across the p-n junction, resulting in a balance in Fermi energies. The result was the formation of charge depletion layers and a potential barrier at the contact. Finally, the ZnO–CuO core–hollow cube nanostructures at 200 °C displayed a 680 k resistance compared with the 8.8 kΩ of the CuO hollow cubes at the p-n junction. Surface-adsorbed oxygen species were consumed and the surface charge in CuO domains was reduced by acetone. Additionally, electrons were donated to the ZnO domain by removing the adsorbed oxygen species close to the interface. In response to this charge restructuring, the charge depletion area moved farther into the CuO domains. A considerable increase in resistance was observed across the CuO surface compared with the nanocubes without ZnO cores [134,135,136].

3.5. Intrinsic Physical Characteristics

Several physical characteristics affect the application of SMOx materials in numerous fields, including morphology, particle size, crystal face, and porosity. A SMOx-based sensor, for instance, has nanoscale pores that can act as barriers to small grain sizes and improve the electronic transmission of sensitive receivers, thus influencing sensitivity to gases [36,137]. In order to diffuse gas molecules into the sensitive receptor, SMOx surface pores should be large. It has excellent sensitivity characteristics because it has large pores on the surface and small pores on the bulk, and it has good grain boundary contact [138,139].

4. Gas-Sensing Applications

A gas sensor fabricated with a gas-sensing element can detect analytic gas species by converting surface interactions into electrical signals [139]. In the last few decades, metal oxide (MOx) gas sensors with a simple and cost-effective fabrication process, high sensing response, and short recovery time captivated the attention of researchers because of their excellent surface morphology, high structural stability, adaptable electrical properties, and grain size, where metal oxide is used as the chemosensory material [5,140,141,142,143]. In 1960, Seiyama et al. introduced the gas-sensing properties of ZnO thin films, which have been used and studied extensively since then [144,145]. The advancement of gas sensing technology has allowed modern gas sensors to operate at low power [146,147,148,149,150,151,152]. As a result of chemical reactions between gas molecules and semiconductor metal oxide surfaces, semiconductor-metal-oxide-based gas sensors have great potential (Figure 4) [144,153,154].
SMOx-based chemoresistors are portable devices that use a battery to operate at high temperatures. In factories, plants, and industries, these devices are used for detecting hazardous gases, for example, oxygen control in the exhaust emissions of gasoline, diesel, and gas engines, and humidity and air quality control in automobiles [155,156]. SMOx-based gas sensors are more sensitive when dopants are used [157], and the electrical properties of SMOx depend on the interactions between the sensor surface and gas molecules adsorbing on it [158].
Figure 4. Schematic illustration of the chemical reactions at the surface of an n-type gas sensor. (a) Chemisorption of oxygen (O2) traps electrons from the conduction band and forms the charged species atomic O and molecular O2. (b) Reducing gases (e.g., CO) react with the surface-bound oxygen and release electrons back into the crystal leading to changes in the electrical conductivity that are related to the CO concentration [159].
Figure 4. Schematic illustration of the chemical reactions at the surface of an n-type gas sensor. (a) Chemisorption of oxygen (O2) traps electrons from the conduction band and forms the charged species atomic O and molecular O2. (b) Reducing gases (e.g., CO) react with the surface-bound oxygen and release electrons back into the crystal leading to changes in the electrical conductivity that are related to the CO concentration [159].
Sensors 23 06849 g004

4.1. Mechanism of SMOx-Based Gas Sensors

The sensing mechanism of a SMOx is complicated by different key factors that affect the sensing attributes, including the adsorption ability, electrophysical property, catalytic and chemical activity, thermodynamic stability, and surface adsorption or desorption properties [30,160,161,162,163,164]. Two processes are involved in the sensing mechanism of SMOx-based gas sensors: reception and transduction [19,156]. In the reception process, the sensor converts the chemical reaction into energy, which is then converted into analytical signals in the transduction process [165]. Through the gas–solid interface, the target gas is detected on the SMOx surface through a change in electrical properties. In the transduction process, chemical changes are induced in the surface and transformed into electrical signals, such as resistance changes in the sensor [166]. As a result of reversible redox reactions between reactive gases and the SMOx surface, a SMOx’s electrical properties change. Finally, the electrical signals were measured and displayed using suitable circuits, such as a microprocessor unit [12,167].
The detection of gas molecules is primarily carried out by monitoring variations in the device current (IDS) or threshold voltage (VTH) caused by the adsorption of nearby molecules. In the case of Ohmic connections, such changes can occur by modulating the conductivity of FET channels, and in the case of non-Ohmic links, by modifying the Schottky barrier height. Due to the conductivity of the FET channel, the free-carrier density in 1D/2D channels is decreased or increased during this process. If the hole (electron) carrier density alters due to gas adsorption, the phenomenon is called hole (electron) doping. An adsorption-induced doping process results in a positive (electron-acceptor) or negative (electron-donor) shift in the VTH, which, in turn, causes an alteration in the IDS for a VDS value. The adsorption of molecules from the surrounding environment can increase or decrease the charge carrier surface scattering and trapping. A change in the majority carrier mobility (p for holes and n for electrons) affects the channel conductivity for both p- and n-type FETs, modulating the IDS. A Schottky contact also modifies the metal workfunction, affecting the height of the energy barrier at the semiconductor interface [6,168].
To evaluate the response and recovery times of a FET sensor as a result of gas molecule adsorption/desorption processes, the real-time measurement of the IDS for VDS and VGS values was used. Trans-characteristic analysis revealed different sensing mechanisms for a 1D and 2D FET. FETs have the advantage of simultaneously monitoring many electrical parameters, such as the IDS, VTH, SVTH, SW, and Ion/Ioff upon gas molecule adsorption, over two-terminal electrical devices (such as resistors, capacitors, and diodes). This can be used to retrieve data on the specific sensing mechanisms when target analytes interact with the FET, such as changes in the concentration of electrons/holes, the energy barrier at the semiconductor/metal interface, or the mobility of the majority of charge carriers. As a result, scientists can tweak the gadget’s architecture and materials to enhance its performance. Aside from this, the sensitivity of the sensor can also be electrically tuned for the detection of low (by lowering the amount of charge in the channel and, consequently, the background current) and high concentrations of the target gas by varying the VGS value and the charge carrier concentration in the FET channel, as opposed to two-terminal electrical components, which prevent it from being effective [6].
Reception: On the SMOx surface, the reception process involves the reactions (i) ionosorption of oxygen to form reactive oxygen species and (ii) the reaction of these reactive oxygen species with reducing gases. When SMOx-based sensors are heated without oxygen, free electrons easily flow through their boundaries. A SMOx (e.g., SnO2), on the other hand, forms a potential barrier when oxygen is adsorbed onto its surface due to its presence in the atmosphere. Through this interaction, atmospheric oxygen traps electrons from the bulk material, thereby depleting a region of electrons. As a result, an increased potential barrier is formed at the surface. The flow of electrons is impeded by this phenomenon, which increases the resistance. The surface of the SMOx sensor absorbs gas molecules when exposed to reducing gases, lowering the potential barrier and allowing electrons to flow more freely. As a result, the electrical resistance also decreases. SMOx-based sensors were demonstrated to be variable resistors.
Due to its wide band gap, SMOx has many electrophysical properties, including insulating behavior and semiconductor properties [169,170,171,172]. A change in conductivity is caused by electrons trapped in adsorbed molecules. Figure 5 [13] shows that electrons are extracted from the conduction band (Ec) when oxygen molecules (O2) adsorb on the SMOx surface [173,174,175] using ionosorption [173,174]. A space charge layer is formed when this phenomenon causes an upward band bend, resulting in an electron-depleted region. By reacting with CO, oxygen species decrease the amount of adsorbed oxygen at the surface, which reverses the band-bending process (Reaction 2). In addition, SMOx gas sensors display increased conductivity with high temperatures (300–450 °C) [25]. In n-type semiconducting metal oxides, where the depletion region is smaller than the grain size, this mechanism is crucial and well-suited for sensing gases.
1 2 O 2 + V 0 + + + 2 e     O 0
where V 0 + + —oxygen vacancy.
O 0 + CO CO 2 + V 0 + + + 2 e
SMOx gas sensors become oxidized when exposed to a reducing gas, such as CO, releasing electrons into the bulk material, resulting in a decrease in the number of O ions on the surfaces (Figure 6). As a result, the thickness of the space charge layer is reduced. By doing so, Schottky barriers between grains or particles become smaller, allowing electrons to easily pass through sensing layers.
Depending on the atmospheric composition, the surface reactions vary, which, in turn, changes the concentration of trapped charges on the surface and the associated space charge layer [177,178].
Transduction: It is necessary to convert surface charges into measurable electrical signals in order to obtain analytical signals. SMOx-based gas sensors measure surface reactions by varying the sensing layer’s resistance. The surface charge is affected by the structural and morphological properties of the sensing layer [80,178], the electrical and chemical properties of SMOx, the size and shape of the SMOx material [179,180,181], and the electrode’s geometry [12]. According to Figure 7, spherical SMOx grains with a Debye length smaller than the grain radius were not affected by changes in the surface, resulting in an unaffected bulk region. A continuous Schottky barrier is created between different grain contacts at the surface as a result of the band bending. Due to oxygen biosorption, an electron-depleted surface layer formed in n-type SMOx, whereas a hole-depleted surface layer was observed in p-type SMOx [166,178,182].
Conduction in p-type oxide semiconductors might be explained by a conflict between parallel routes spanning a broad resistive core (Rcore) and a constrained, p-semiconducting shell (Rshell). Barsan and co-workers provided a detailed explanation of the precise conduction model and energy band diagram of p-type oxide semiconductor-based gas sensors (Figure 8a) [166,182]. As shown in Figure 8a, the B region represents the electrode–semiconductor connections, while the A and C regions show the semiconductor grain-to-grain interactions. Oxygen anions on the surface of the SMOx react to form an oxidation process that injects electrons into the material, reducing the concentration of holes in the shell layer while increasing the sensor resistance (Figure 8b,c).
P-type oxide semiconductors are known to lose resistance when exposed to oxidizing gases like NO2 and O3. As a result of the ionosorption of oxidizing gas, the concentration of holes in the shell layer increases. The chemo-resistive variation of p-type oxide semiconductors to oxidizing gases also appears to be not high when considering the gas-sensing mechanism. Based on literature data, NiO and CuO sensors respond moderately to 10–100 ppm NO2 (Ra/Rg = 1.0–7.5) [183,184,185].

4.2. Sensitization Mechanism

The sensitization mechanism enhances the sensitivity of gas sensors by adding additional material or by modifying the existing material. The variation in the reception and transduction mechanisms increases the electronic and chemical interactions between the target gas and sensing material, resulting in a high sensitivity of the sensor [186,187,188]. As a result of dividing sensitization mechanisms into two main categories, namely, electronic sensitization and chemical sensitization, N. Yamazoe developed the concept of clear separation of electronic and chemical sensitization. By adding metal or metal oxide additives, both sensitizations could be achieved. During electronic sensitization, the additive accepts electrons from the analyte and changes its redox state or chemical potential (Figure 9a). Chemical sensitization, on the other hand, involves activation of the analyte, spillover, and a change in the surface oxygen concentration (Figure 9b). The concentration of charged species at the surface and space charge layers remains unchanged without an analyte. As a result of the high surface concentration of active oxygen, a different phenomenon occurs during oxygen spillover. The high concentration of charged species at the surface of SMOx causes the band bending to increase when chemical activation occurs.
Spillover effects promote the separation of molecular oxygen into an active surface. Spillover activation significantly impacted the sensor’s gas-sensing properties. As the analyte gas adsorbs on the additive phase, the activated species are transferred to the SMOx surface, where the analyte gas reacts with active oxygen species (Figure 10). The oxidation catalysis process is affected by single additive sites, i.e. dopants [189] and by separate additive phases [190,191,192]. As a result of spillover effects, molecular oxygen separates into a more active surface. Chemical sensitization involves two important aspects: (i) the change in the surface charge of SMOx caused by chemical activation and (ii) the ambivalent relationship between catalytic activity and gas sensing. As an example, CO is activated by adsorption and transfer to SMOx surfaces, while other analytes (e.g., H2) are activated by additive reactions. Analyte gas spillover activation enhances the reactivity of SMOx and accelerates oxygen vacancy formation. Despite the inverse spillover mechanism described by Korotcenkov et al. for the detection of reducing gases with Au-loaded SnO2 (Figure 11a) [191], spillover mechanisms are generally used to improve the gas-sensing properties of metal-loaded SMOx (Figure 11b) [186,193,194,195,196].
The increase in initial band bending directly affects the resistance change due to changes in the surface charge of SMOx, which shows a significant impact on the electrical and chemical properties of the SMOx surface, and consequently on transduction and reception processes. The initial band bending is described by a non-linear relationship between the surface charge (QS) and the surface potential (VS). The Schottky approximation, described by S. R. Morrison, provides a simplified explanation of this relationship [198] as follows:
V s = e 2 ε ε 0 n b × Q S 2
where
e—elementary charge;
ε—permittivity of SMOx;
ε0—permittivity of vacuum;
nb—concentration of charge carriers in the bulk.
The sensor signal (S) for the reducing gas is defined by the ratio of the resistance in the reference atmosphere (Rref) and the resistance in the presence of the analyte gas (Rgas) [178]:
S   = R r e f R g a s
The relationship between the differential surface potential Δ V s and sensor signal S is as follows:
S   =   exp   ( e Δ V s k g T )
The change in band bending at the surface, which is represented as the differential surface potential Δ V s as a function of the change in surface charge Δ Q s as follows:
Δ V S = e 2 ε ε 0 n b × Δ Q S . 2 × Q S , 0 Δ Q S
According to Equation (6), the change in the surface potential (ΔVS) is a function of the initial surface charge (QS,0) and the change in the surface charge (ΔQS) as a consequence of reactions taking place on the surface. Combining Equations (5) and (6), the impact of different QS,0 values on the sensor signal can be calculated. From Figure 10, it is evident that the change in the surface charge results in a change in the sensor signal. The slopes of the curves indicate that a higher initial band bending leads to an increased sensor signal and hence sensitivity. The increase in slopes, referred to as n-values, is experimentally observed in SnO2 materials due to the self-doping effect (Figure 12) [199]. The presence of two SMOx materials shows a major effect on the electronic coupling between them, as the second material provides additional reaction sites on the additive phase [187,188].

4.3. Factors Affecting the Sensitivity

4.3.1. Chemical Composition

The chemical composition of SMOx plays a significant role in improving the adsorption ability, catalytic activity, sensitivity, and thermodynamic stability [200]. Recently, composite materials, like SnO2-ZnO [201,202], Fe2O3-ZnO [203], and ZnO-CuO [204], were investigated for their improved performance. Furthermore, researchers are working on various ternary, quaternary, and complex metal oxides for different applications [205,206]. The combination of metal oxides and other components, for example, organic and carbon nanotubes, also showed promising results. Therefore, the chemical composition significantly influences the gas-sensing properties of composite metal oxides.
ZnO-SnO2-composite-based sensors exhibit higher sensitivity compared with sensors made solely from tin dioxide or zinc oxide [203]. As described by De Lacy Costello and colleagues, the combined effect of two components in a synergistic effect enhances the sensitivity of sensors [203].

4.3.2. Surface Modification

Controlling the catalytic activity of gas sensors is another way to enhance their performance. Several widely used SMOxs, such as TiO2, ZnO, SnO2, Cu2O, Ga2O3, and Fe2O3, have low activities [160]. Without a catalyst, pure SnO2 exhibits poor sensitivity [207]. Noble metals, such as Pt, Au, Pd, and Ag, were highly effective oxidation catalysts and were also used to enhance reactions on metal oxide surfaces [208,209,210,211,212]. In order to detect Pd upon deposition, two mechanisms are involved: (i) electronic and (ii) chemical. The electronic mechanism involves the formation of depletion zones around the modified particles (Figure 13b) and modulation of the nano-Schottky barriers, which result from changes in the oxidation state of Pd during oxygen adsorption and desorption, boosting the sensing process. In Figure 13a, oxygen is shown as ionosorbing on the surface of Pd. Pd, which catalyzes molecular oxygen dissociation, is a better oxygen dissociation catalyst than tin oxide. As a result, the atomic products diffuse to the surface of the SMOx (Figure 13a) [213]. Oxygen molecules reside on the oxide and diffuse to the catalyst particle. A back-spillover effect (Figure 13a) and effective “capture radius” (Rc) are observed around Pd particles (Figure 13b). An oxygen delivery system is developed when oxygen layers cover whole metal oxide surfaces [214]. This process enhances the oxygen ionosorption on MOx and the detection sensitivity [215].

4.3.3. Microstructure

A significant effect on sensitivity was also observed when metal oxides were synthesized with an optimal morphology and crystallographic structure. SMOx sensors are made more sensitive by using small-grain materials in this method. According to Lu et al. the sensitivity of a SnO2-based sensor to 500 ppm CO increased dramatically for particle sizes smaller than 10 nm (Figure 14a). Similarly, 20 nm particles demonstrated ten times more sensitivity compared with 25~40 nm particles (Figure 14b) [217]. In the case of small grains with narrow necks, this size is less than twice the thickness of the surface charge layer [218]. In metal oxide gas sensors, the size of the grains affects the mobility of free charge carriers, and therefore, the number of collisions that occur between them. Additional scattering centers are formed by adsorbed species, which influence carrier mobility as well [178]. As a result of this method, metal oxide gas sensors are significantly more sensitive [178,219,220,221,222]. It is worth noting that using a small crystal size did not always enhance the gas sensor’s response [223]. Sensors fabricated with SnO2 nanocrystals (50 nm) synthesized via gel combustion showed a faster response than those fabricated with SnO2 nanocrystals (12–13 nm) synthesized via the hydrothermal method. In contrast with hydrothermally synthesized SnO2 nanocrystals, which consist of small grains but tend to aggregate into large entities, gel-combustion-synthesized SnO2 nanocrystals were more porous. As a result, tuning the grain size of metal-oxide-based gas sensors could enhance their performance. In addition to modulating structural stability, grain size affected the surface changes, as well as the catalytic activity [224].

4.3.4. Humidity and Temperature

A significant role was played by humidity in modulating the activity of metal-oxide-based gas sensors. A variety of humidity sensors based on metal oxides have been developed. In contrast, the mechanisms of sensing water vapor and CO, NO2, and H2S gases differed. It was the ionic humidity sensor that was most commonly used with metal-oxide-based humidity sensors. H+ or H3O+ ions produced by the dissociation of adsorbed water on the surface are the conduction mechanism in metal-oxide-based humidity sensors. Recent works [226,227] studied and described the adsorption of water on metal oxide surfaces and the mechanism of sensing water vapor. As a result of water adsorption on metal oxide surfaces, humidity decreased the sensitivity of metal oxide sensors (Figure 15) [228,229].
Temperature is another important factor that affects the performance of metal oxide gas sensors. As shown in Figure 16, the sensor responses to different analytes show similar shapes as a function of temperature. The resultant shape depicted slow kinetics at low temperatures and increased desorption at high temperatures [216,230,231,232,233,234].

4.3.5. Control Synthesis

The ability to detect gas depends on its composition, shape, size, and distribution. Depending on the crystallographic orientation of the metal oxide semiconductor, sensitivity is improved. A higher sensitivity is observed for ZnO with the crystallographic orientation (002) [235]. For the crystallographic orientation, a high surface-to-volume ratio is also crucial to achieving greater sensitivity. Furthermore, a well-organized pore structure and particle size contribute to a high surface-to-volume ratio. ZnO was characterized using a scanning electron microscope (SEM) after it was synthesized (Figure 17a,b). Surface morphology can be modified via chemical processing and variable annealing temperatures. As a result, hexagonal-shaped ZnO was developed (Figure 17b). A hexagonal shape with a high surface-to-volume ratio produces good sensitivity.

4.3.6. Doping with a Noble Material

To assess conductivity, the effectiveness of catalytic reactions with target gas is measured on the surface of the sensing material. It is also important to improve catalytic activity in order to improve sensor performance. Undoped metal oxides exhibit significantly lower activity than doped ones [237]. Sputtering, thermal evaporation, or sol-gel doping could be used to dope semiconductors. A surface modification sometimes requires the combination of noble materials and MOxs (metal oxides). In addition to Pd, Ag, Au, and Pt nanoparticles, second-phase nanoparticles with enhanced sensitivity were applied to host metal oxides [238]. As a result of the catalytic behaviour of noble metal nanoparticles, the chemical dissociation and reactions are increased [239]. The activity of Pd-functionalized SnO2 is improved [216]. ZnO nanostructures decorated with PdO also demonstrate improved sensitivity [230].

4.4. SMOx-Based Gas Detection for Environmental Monitoring

Here, we discuss a few SMOx-based gas sensors for monitoring the toxicity of gases in the environment, industries, hospitals, health issues, etc.

4.4.1. Nitrogen Oxide Gas Detection

WO3 sensors based on SMOx have been extensively used in gas sensors [240,241,242,243,244,245]. Most of the modified WO3-based metal composites were used for NOx sensing [246,247,248,249]. Modified thermal evaporation techniques were used to obtain nanostructured WO3 films with high surface roughness [240]. This technique produces high-response sensors with high selectivity and short response times, particularly at low temperatures (minimum 100 °C). Due to a high variation in electrical resistance, sensors at this temperature exhibited high sensitivity to NO2 with a low detection limit (around 100 ppb). In contrast, low responses were observed when NH3 (10 ppm) and CO (400 ppm) were present at high concentrations [240]. WO3 sensing elements for high-temperature potentiometric NOx sensing were synthesized by Yang et al. using various synthetic methods. Mixed oxides based on WO3 are also used for sensing. SMOx materials such as WO3-Ti [246,247,248], WO3-Pd, Pt, Au [249,250,251,252], WO3-In2O3 [253], and WO3-Bi2O3 [254] are used to fabricate selective and sensitive NOx gas sensors. Furthermore, TeO2-based thin films were synthesized for NO2 gas sensing [255,256]. As the NO2 gas concentration increases, the response time decreases. A response time of about 6 min for 1 ppm of NO2 gas and about 1.2 min for 120 ppm was reported. The recovery time was found to be longer than 8 min for each gas concentration [255].
The direct printing of numerous SMOxs, including ZnO, In2O3, SnO2, and WO3, was proposed by Kim and co-workers for the creation of an aligned network of NWs as the channel of a FET for NO2 detection [257]. The diameters of the manufactured NWs, which had aggregates of grains between 5 and 15 nm, ranged from 100 to 2400 nm depending on the concentration of metallic precursor in the printing solution. It was investigated whether a ZnO FET could detect NO2 at concentrations ranging from 1 to 5 ppm using an aligned network of NWs directly printed on Pt interdigitated source/drain electrodes. The sensing capabilities of a ZnO FET at concentrations ranging from 1 to 5 ppm were tested using an aligned network of NWs printed directly on Pt sensor electrodes. A rise (fall) in NW resistance was observed after the injection of 5 ppm of NO2 (50 ppm of ethanol) due to the target gas acting as an electron acceptor (donor) on n-type NWs. The sensor calibration curve (resistance variation vs. NO2 concentration) was linear in the tested range. An extrapolated LoD of 53.5 ppt was calculated. In response to 5 ppm of NO2, the response and recovery periods were 67 and 11 s, respectively.

4.4.2. SO2 Gas Detection

Sulfur dioxide (SO2) is a common air pollutant that can be detected using sensors. Polymeric sensing films [258,259,260], as well as liquid and solid electrolytes, are used to make these sensors. SO2 gas sensors based on SnO2 [261], SnO2 doped Pd [262], WO3 doped with various metals [211,263,264], and vanadium oxide modified with TiO2 [264] were also developed and tested. During the initial detection, Berger et al. examined the interaction mechanisms between SO2 and the SnO2 sensor interface [261]. As a catalytic additive, SnO2-based gas sensors contained 0.05, 0.1, 1, and 3 mol% Pd. In recent years, thick-film technology has been used to increase the operating temperatures to 600 °C [262]. The magnetron sputtering method was used to fabricate active layers of pure and Pt-doped WO3 on a micro hotplate substrate to detect sulfur compounds (SO2 and H2S). A compact tubular SO2 sensor based on a sodium superionic conductor and V2O5-doped TiO2 sensing electrode was described by Liang et al. [265].

4.4.3. H2S Gas Detection

Due to their excellent performance in detecting hydrogen sulfide (H2S) gas, SMOx-based gas sensors caught the attention of researchers. H2S is a toxic gas with a threshold limit of 10 ppm, and concentrations over 250 ppm cause serious health problems, including death. Metal oxides are a model-sensitive material for H2S sensing [266,267] and other gases [36,268] because of their high sensitivity, quick response, and ease of integration. In contrast, SMOx-based sensors have low selectivity, are highly dependent on the relative humidity, and require high operating temperatures (above 100 °C). Metal oxides (MOx) are the most commonly used substance in chemo-resistive gas sensors, both in academic and industrial settings [36,268,269]. Hexagonal WO3 nanoparticles show the best selectivity to 10 ppm H2S at 200 °C, according to Szilagyi and co-workers [48]. If the operating temperature drops below a certain level, the response does not occur. Interlaced and condensed WO3 nanofibers can detect H2S gas at ppm levels, according to Niu and co-workers [31]. WO3 is also hindered by low selectivity or high operating temperatures when used to detect H2S gas [270].
These materials can be combined with CPs (conducting polymers) to improve their sensing capabilities while addressing some of their flaws, such as low selectivity and high temperature [271]. MOx/CP composites already demonstrated H2S detection from 0.05 to 1000 ppm. By lowering the temperature of the sensor, hybridization with MOx significantly improves the response quality and operating conditions of the H2S sensor. MOxs and CPs have a delicate synergetic effect due to the entanglement of their sensing mechanisms (from CPs, MOxs, and p-n heterojunctions), as well as numerous parameters that influence their effectiveness (e.g., synthesis, deposition, and morphology). Furthermore, environmental factors and long-term stability (>1 month) were too rarely studied, even though they are crucial for sensor applications [272]. Different SMOx-based H2S gas sensors were successfully modified using the following materials: WO3 and WO3-based materials [273,274,275,276,277], SnO2 [229,278,279,280,281], ZnO [282,283], copper oxide [280,284,285], platinum and palladium oxides [286,287], indium oxides [287,288], silver-based materials [289,290], titanium oxide [291], and cadmium oxide [292]. In dry and wet synthetic air with varying levels of humidity, WO3-based SMOx sensors responded strongly to H2S. The H2S sensitivity of some WO3 thin-film sensors is at the ppb level. A slight increase in conductance was also observed in the presence of humidity [276]. H2S can also be detected at room temperature by ZnO-based sensors down to 0.05 ppm [282].

4.4.4. Amine Gas Detection

In many fields, like food processing, fertilizers, chemical technology, medical diagnosis, and environmental protection, it is extremely important to detect any trace amount of ammonia/amine. WO3 [293,294], copper-based materials [293,295], ZnO [296], SnO2 [297], iron oxide [298], and Cr2O3 [299] are well-known materials for functionalizing ammonia/amine detecting sensors. To produce ZnO films doped with various amounts of RuO2, thick films of ZnO were immersed in an aqueous solution of 0.01 M ruthenium chloride (RuO2) [296]. At operating temperatures between 100 and 350 °C, the doped ZnO sensor was exposed to 1000 ppm NH3. With increasing operating temperature, the response increased.

4.4.5. Hydrogen Gas Detection

Among its potential uses are automobiles, electricity generation in fuel cells, medicine, space exploration, industrial chemical production, and food production. In the event hydrogen leaks into the air from storage tanks or valves, explosive mixtures can form, making hydrogen-monitoring devices necessary. Enhanced sensitivity and selectivity to H2 gas were demonstrated with a nanostructured SnO2 thin film doped with silver (Ag) and platinum (Pt). At 100 °C, nanocrystalline SnO2 shows a fast response time (around two seconds) and a quick recovery time (around ten seconds). In addition to their high sensitivity to H2 gas, porous SnO2 particles have a high surface area [300]. In today’s world, chemo-resistive gas sensors based on the IOT (Internet of things) are used as H2 sensors with low power consumption and a lower temperature [301]. A potential technique for SMOx-based gas sensors is self-heating, particularly for materials with NW shapes. Self-heating gas sensors can significantly reduce their power consumption from several W to nW levels. Power consumption reductions can significantly extend the life of sensors and save a lot of energy. This can be accomplished using single, arranged, and networked NWs. However, networked NWs are simpler to synthesize than single or ordered NWs, which makes them the most popular morphology for self-heating petrol sensors. The majority of self-heating MOx materials were reported to have NW morphology. In spite of the fact that self-heated gas sensors can display power consumption in the nW range, most of these sensors have very low response values [302].
MEMS gas sensors based on SMOxs have straightforward topologies, are easy to manufacture, and consume little energy. Using MEMS gas sensors results in both a reduction in gas sensor size and a reduction in power consumption for both gas sensors and electrical devices. In terms of power consumption, MEMS-based gas sensors trail self-heated gas sensors. As a result of gas sensing measurements still requiring high temperatures, MEMS gas sensors rely on an external heater. As a result, the problem remains difficult to solve. Despite the significant advances in the development of low-power-consumption-based gas sensors, there are still several challenges and issues to overcome in order to achieve high sensitivity, selectivity, long-term stability, and quick response/recovery times.
As a H2-sensitive medium, tungsten oxides with palladium or platinum catalysts display a color change from pale green to blue when hydrogen reduces them to tungsten bronze [295,303].

4.4.6. Volatile Organic Compound Detection

Animals and plants are both affected by volatile organic compounds (VOCs) that cause chronic diseases, such as eye irritation, throat and lung problems, and cancer in humans. Many studies were carried out on modified SMOx sensing films for the detection of atmospheric VOCs, such as ethanol, acetone, hydrocarbon, and liquefied petroleum gas (LPG). Different materials are used to modify SMOx sensors, such as SnO2 and SnO2-based materials [228,304,305,306,307], WO3 and WO3-based materials [308,309,310], titanium oxides [310,311,312], zinc oxides [310,311,312], iron oxides [313,314], cobalt oxides [315], cerium oxide sensors [316], and copper-based materials [317]. By adding basic metal oxides, such as lanthanum oxide (La2O3), ethanol gas sensors can be made more sensitive. In the presence of La2O3 and WO3, ethanol gas undergoes dehydrogenation and dehydration over SnO2-based elements, respectively [318]. SnO2 doped with cadmium oxide (CdO) shows enhanced sensitivity to C2H5OH and H2 at 300 °C with a detection limit of several ppm in air [319]. Similar to MEMS/NEMS chemo-resistive gas sensors, the IoT also enhances C2H5OH sensitivity at low temperatures or room temperatures because of the high surface area and reduced power consumption. VOCs, such as acetylene, LPG, and aldehyde, can be detected by modified tin-oxide-based films. HCHO was reported to be stable and sensitive in a SnO2–NiO composite material [320]. According to Qi and co-workers, SnO2-based sensors modified with 6 wt% Sm2O3 were 16.8 times more responsive to C2H2 than SnO2 sensors. As an excellent C2H2 sensor, the Sm2O3-doped SnO2-based sensor showed a high sensitivity under various humid conditions [228]. It was also demonstrated that SnO2-based sensors can successfully detect LPG [210,306]. In2O3 NW FETs doped with Yb were proposed by Jun and co-workers [321]. A 4 mol% Yb-doped In2O3 NW FET exhibited n = 6.67 cm2 V−1 s−1 n-type behaviour. Based on the output characteristics of VTH = 3.27 V, SVTHSW = 0.5 V dec−1, and Ion/Ioff = 107, linear and saturation zones were assessed. Undoped In2O3 NW FETs exhibited n-type behavior as well, with n = 10.82 cm2 V−1 s1, VTH = 10.26 V, SVTHSW = 2.5 V dec−1, and Ion/Ioff = 103. Over the entire test range, the calibration curve of the 4 mol% Yb-doped In2O3 NW FET displayed linear performance. In2O3 NW FETs with undoped silicon were approximately three times less sensitive [6]. The following table illustrates the gas-sensing activity of SMOx-based nanomaterials, where tres corresponds to the response time and trec corresponds to the recovery time. SMOx-based gas sensors are compared in Table 2 for comparison.

5. Chemical Sensing Applications

Chemical sensors convert chemical reactions into electrical, optical, or mechanical signals by using chemical-responsive layers. Chemical sensors are more specific than physical sensors because of their chemical-responsive layers. In order for these sensors to respond, a chemical-selective layer must interact with the target chemical, changing the transducer properties and resulting in a signal. Conductometric chemical sensors based on SMOx were previously reported. As part of this development, MOx was discovered to react with the surrounding atmosphere, and the first commercial gas sensor was developed [144,354,355]. Their low cost, simple preparation, and simple operation make SMOx-based chemical sensors a promising technology. Chemical sensors were classified by IUPAC in 1991 (Figure 18). A chemical sensor is a device that converts chemical information, such as the concentration of a specific chemical component or a composite, into useful analytical information [165].
Chemical sensors have improved the detection and quantification of various chemical substances. Medical, agricultural, industrial, and military applications are all possible with these sensors.
As illustrated in Figure 19, chemical sensors convert chemical information into quantitative or qualitative analytical signals through chemical interactions between the analyte gas or liquid and the sensor. Electric sensors produce signals through the exchange of electrons, which are electronic in nature. A chemical sensor consists of a physical transducer and a chemical-sensitive recognition layer. Stability, sensitivity, selectivity, response time, recovery, and saturation characterize them [155]. Due to their high sensitivity, compatibility with ambient conditions, and ease of fabrication, semiconductor metal oxides (SMOxs) are widely used as chemical sensing materials [163,356,357]. By exposing a metal oxide to elevated temperatures, the MOx reacts with surrounding gases, changing the surface potential and resistivity of the material.
Based on the MOx materials used for specific target species, SMOx-based chemical sensors are categorized.

5.1. SnO2-Based Chemical Sensors

SnO2 is a highly sensitive and fast-responding material that is widely used as a chemical sensor. Various morphologies are available, including nanowires, hollow spheres, nanocrystals, and others, each with its own unique sensing properties. Wang and co-workers reported that nanowires based on SnO2 were highly effective at detecting H2 at concentrations between 10 and 100 parts per million [358]. In addition, SnO2 nanowires, hollow nanospheres [359,360], nanocrystals [361], and nanocrystalline porous SnO2 [362] exhibit CO-, NO2-, and H2-sensing abilities. Hierarchical three-dimensional SnO2 nanospheres also exhibit excellent resistance to CO, methane, methanol, and ethanol [363]. Furthermore, SnO2 nano polyhedrons are highly sensitive to methanol, ethanol, and acetone, as well as highly selective to acetone, with a fast response and recovery time (only several seconds for target gas concentrations up to 200 parts per million) [364].
SnO2 is modified via doping and composite formation to enhance its sensing activity. In turn, this results in the development of highly sensitive and fast-responding SnO2-based composites, like polypyrrole-coated SnO2 hollow spheres, for the detection of ammonia [365]. Due to an increased concentration of oxygen vacancies on the surface of SnO2 nanowires at low temperatures, plasma-modified SnO2 nanowires [366] and Pt@SnO2 nanorods [367] displayed high sensitivity to ethanol gas. Adding Pt to SnO2 nanowires also enhances their chemical and electrical properties. For the detection of H2 gas (5 ppm) at 320 °C, Wang and co-workers synthesized hetero-junction p-NiO/n-SnO2 nanofiber-based sensors, which have excellent sensitivity and fast response recovery. Cu-doped SnO2 and the adsorption properties of H2S on the surface of SnO2 were investigated by Wei et al. [368]. Pd-doped SnO2-based CO gas sensors were recently described by Li et al. [369].

5.2. ZnO-Based Chemical Sensors

Since they are easy to synthesize and have unique optical, electrical, and chemical properties, ZnO-based chemical sensors have received considerable attention. Using ZnO material, Hahn and co-workers fabricated hydrazine electrochemical sensors (Figure 20) [370]. Detection limits of 0.2 mM were achieved with this sensor’s high sensitivity (8.56 mA mM−1 cm−2) and low response time (less than 5 s). ZnO nanorods and high aspect ratio ZnO nanowires were also used by these researchers to fabricate hydrazine sensors.
The electrical characteristics of aligned ZnO nanorod arrays (NRAs) were also described by Hahn and co-workers (Figure 21) [371]. H2 was detected using ZnO NRAs. ZnO NRAs became more sensitive as the H2 concentration increased. Chemical sensors based on ZnO nanowires were also used to detect multiple gases at room temperature, including H2, NH3, i-butane, and CH4.
As shown in Figure 22a [372], Li and colleagues synthesized co-doped ZnO nanorods on ITO substrates at low temperatures. As shown in Figure 22b, these nanorods responded rapidly to varying CO concentrations. In this study, co-doped ZnO sensors performed better than pristine undoped ZnO sensors. By attaching impurities to SMOx-based semiconductors, the sensing properties were greatly enhanced. Pd nanodots were incorporated into ZnO nanowires by Choi and Kim, which enhanced the CO sensitivity. A combination of electronic and chemical sensitization could be responsible for this enhanced sensitivity [373].

5.3. Other SMOx-Based Chemical Sensors

Chemical sensors rely heavily on selectivity. There is a slight deficiency of oxygen and incomplete crystallinity in gallium oxide (Ga2O3), which are preferred for high-temperature sensing applications, such as chemical, environmental, and explosive gas sensing [374]. Gallium oxide is chemically and thermally stable with low cross-sensitivity to humidity. The use of mesoporous single-crystal Ga2O3 nanoplates for CO detection was described by Yan and co-workers [375].
Using CoCl2 and urea precursors, crystalline mesoporous Co3O4 nanorod-based sensors were fabricated via facile hydrothermal methods. Benzene, acetone, and ethanol were detected using the sensor. In addition to its excellent stability, high sensitivity, rapid response, and recovery time, the SMOx-based sensor was particularly sensitive to acetone (Figure 23) [376]. The ethanol sensitivity of Fe2O3–TiO2 tube-like nanostructures was also improved [377]. In2O3-based chemical sensors were used to detect NH3, CO, H2S, NOx, ethanol, formaldehyde, and alcohol [378,379,380,381,382,383,384,385]. In Table 3, we present a comparative analysis of SMOx-based chemical sensors.

6. Biosensing Applications

Because of their exceptional electrical properties, high electron mobility, excellent chemical resistance in liquids, transparency, and ease of fabrication, semiconductor metal oxide (SMOx)-based thin-film transistors (TFTs) are widely used in liquid crystal displays (LCDs), biosensors, and photosensors [403,404,405,406,407,408]. SMOx-based biosensors are highly effective at recording and communicating biomolecule progression statistics [409]. Its versatile morphology [410], chemical stability [411], physicochemical interfacial properties [412,413], light excitation, and ability to form composite structures [412] make SMOx materials potential candidates for biosensors. Electrochemically sensitive materials, such as TiO2 [414], WO3 [415], SnO2 [416], and ZnO [417], are suitable for enzyme-based biosensors. In addition, these SMOx materials require a cost-effective synthesis procedure, including co-precipitation [416], chemical precipitation [418] thermal oxidation [419], chemical etching [420], polyol [421], hydrothermal [422], sol-gel [423], and sonochemistry [424], which allows for the formation of different architectural morphologies, including porous quasi-nanospheres [425], hollow nanospheres [426], nanorods [427], nanosheets [428], and flower-shaped particles. Additionally, SMOx materials can be combined with other materials to form heterostructures [429] hybrid structures [430], and composite structures [431]. Their advanced electrochemical properties make them ideal for specific biosensor applications. Biosensors require a sensing layer that reacts with a biomolecule, and this reaction is converted into optical, electrochemical, electrical, or other physical signals (Figure 24).
For clinical diagnostics and personalized care, biosensors must be sensitive. There are several reasons why nanostructured SMOx materials have excellent sensing capabilities: (1) their increased surface-area-to-volume ratio enhances sensitivity to small analytes, as their size becomes comparable to the SMOx materials [432,433] (2) direct electron transfer enables increased sensitivity and heightened detection limit [434]; and (3) nanostructured particles close to the Debye length, which increases their sensitivity [219]. SMOx surfaces are typically attached to biomolecules via physical adsorption, entrapment, crosslinking, covalent coupling, or encapsulation. In a biocompatible environment, such interactions form a nano–bio interface that is highly stable and preserves the biomolecules. Through the coupling of biomolecules with a bio-recognition layer, selectivity was achieved. A wearable biosensor continuously monitors physiological signals, collects sensor data, wirelessly transmits the data, and analyzes it in real time. Wearable biosensors have several advantages, including rapid continuous monitoring, detection of transient phenomena, ease of use, and accuracy. Following are a few promising nano-structured SMOx-based biosensors. Enzymes and other biomolecules are immobilized on the surfaces of these biosensors.

6.1. Enzyme-Immobilized Biosensors

Figure 25 illustrates the energy band diagram and crystal structure of different types of SMOx biosensors [435,436]. SMOx properties can be modulated by combining them with other metal nanoparticles or ions. The morphologies of SMOx materials include rods, stars, flowers, cones, and porous or dense films.

6.1.1. Glucose-Oxidase-Immobilized Biosensors

Due to a SMOx’s high isoelectric point, glucose oxidase (GOx) with a low IEP of around 4.2 could be immobilized [437]. SMOx nanostructures and biopolymer composites also improved glucose biosensor activity. ZnO nanostructures have a high IEP (around 9.5), resulting in fast electron transfer rates and high enzyme loading activity [438,439,440]. An electrospun ZnO nanofiber glucose sensor exhibited a high and reproducible sensitivity of around 70.2 mAmM−1 cm−2 for glucose at 20–85 °C (Figure 26) [438].
On a glassy carbon electrode (GCE), Fang et al. synthesized Nafion/GOx/ZnO hollow nanosphere composites. Since hollow nanosphere ZnO and GOx were adsorbed on the surface of the sensor [441,442], it exhibited high sensitivity (65.82 mA mM−1 cm−2) and fast response time (5 s). The authors reported direct electron transfer at a rate of 0.67 s−1 in glucose biosensors fabricated with GOx-immobilized ZnO/Cu nanocomposites [442]. There has been a surge in the development of wearable biosensors for the non-invasive monitoring of blood glucose. Field-effect transistors (FETs) based on SMOx gained attention in this area [443,444,445,446,447]. An electrochemical biosensor based on a FET coated with In2O3 3.5 nm thick is shown in Figure 27. To monitor glucose levels in tears, this FET-based device was decorated with glucose oxidase [448]. Blood glucose concentrations are 70–180 mg/dL in a healthy individual, while tears contain 3–15 mg/dL glucose [448,449]. Through an ultra-thin In2O3 FET, the biosensor in Figure 27 could detect ultralow glucose concentrations in tears.

6.1.2. Cholesterol-Oxidase-Immobilized Biosensors

In order to fabricate an efficient and reliable cholesterol biosensor, SMOx was immobilized and stabilized with cholesterol oxidase (ChOx). This table outlines cholesterol biosensors constructed with different types of nanostructured modified electrodes [434,450]. At a low temperature, the Hahn group immobilized cholesterol oxidase on well-crystallized ZnO nanoflowers [434] and ZnO nanoparticles (NPs) [451]. It was shown that these biosensors had high reproducible sensitivity, a low detection limit, and a very fast response time (<5 s). ZnO nanofilms [452] and nano porous ZnO thin films [453] were also employed to fabricate cholesterol biosensors, which enhanced the electron transfer between cholesterol oxidase and electrodes. Composites, including platinum–gold-functionalized ZnO nanorods [454] and Pt-incorporated ZnO nanospheres [455], improved the sensitivity with low Km values. For the first time, Hahn et al. fabricated controlled ZnO nanorods directly on a silver electrode at 90 °C (Figure 28) [456]. The loaded cholesterol oxidase surface area of Malhotra and coworkers increased the electron transport between ChOx and the electrode, resulting in improved sensitivity [457]. SnO2 nanoparticles and chitosan composite films were developed by Ansari and co-workers for enhanced cholesterol adsorption [458]. A CH-SnO2/ITO nanocomposite was synthesized to improve the electrocatalytic activity and biocompatibility. SMOx-based highly-sensitive cholesterol biosensors were also developed using co-oxidase [459] and Fe3O4 nanoparticles [450].

6.1.3. Urea- and Glutamate-Immobilized Biosensors

A urea biosensor based on ZnO nanowire arrays immobilized with urease (Urs) by Ali and co-workers had a sensitivity of about 52.8 mV per decade with linear response ranges (0.1–100 mM) [460]. Their bioelectrode also displayed a low response time (25 s), wide linear range (8 mM−3 mM), low detection limit (5.0 mM), and excellent stability [461]. Furthermore, Urs and glutamate dehydrogenase (GLDH) were combined to form a nanocomposite film on superparamagnetic Fe3O4 nanoparticles and chitosan that demonstrated a low Km value (0.56 mM) [462]. Nanostructured ZnO composites containing Urs and GLDH were reported to have a high sensitivity and a low detection limit of 13.5 mg dL−1, with a Km value of 6.1 mg dL−1 [463].

6.1.4. Lipase-Immobilized Biosensors

It is also possible to form SMOx biosensors by immobilizing lipases. Solanki et al. describe a nanostructured cerium oxide film (35 nm) for lipase immobilization. With a linear range of 50–500 mg dL−1 and a detection limit of around 32.8 mg dL−1 at a low Km value (22.27 mg dL−1), the film exhibited a high affinity for tributyrin [464].

6.1.5. Other Enzyme-Immobilized Biosensors

The TiO2 nanoneedle film immobilized with cytochrome complex (cyt c) described by Luo et al. facilitated electron transfer between redox enzymes and electrodes [465]. Their goal was to improve the enzyme activity against H2O2 released from human liver cancer cells. The detection of choline was achieved by electrochemically depositing MnO2 nanoparticles and nanowires, along with CH hydrogel and choline oxidase on GCE. MnO2 effectively trapped the target analyte on the electrode surface because of its large specific area. The linear detection range for α-MnO2 nanoparticles was 2.0–580 mM choline, while the linear detection range for β-MnO2 nanowires was 1.0–790 mM choline. In addition, an electro-chemiluminescence (ECL) lactate biosensor was made from nano-hybrids of ZnO-multiwalled carbon nanotubes (MWCNTs) modified with lactate oxidase and Nafion. Human blood plasma samples were tested using this ECL lactate biosensor.

6.2. Nucleic-Acid-Immobilized Biosensors

For DNA detection, SMOx materials are immobilized with nucleic acids (Table 4) [466,467,468,469,470]. Among the applications of these DNA biosensors are disease detection, genetic disorder screening, drug discovery, and forensics [471]. For the detection of acute promyelocytic leukemia, Zhang et al. immobilized single-stranded DNA sequences of 18-mer PML/RARA oligonucleotides on a carbon ionic liquid electrode modified with nanosized ZnO. The detection range was 1 × 10−8 to 1 × 10−12 M, with a detection limit of 2.5 × 10−13 M [467]. A ZnO-nanowire-based DNA biosensor made with multiwalled carbon nanotubes (MWCNTs) and gold nanoparticles was described by Wang et al. Sequence-specific target DNA can be detected using this biosensor. A single-stranded DNA probe with a thiol group at the end (HS-ssDNA) was covalently immobilized on the Au nanoparticle surface. A DNA biosensor based on [Ru(NH3)6]3+ as an intercalator [466] was capable of quantitatively detecting DNA in the range of 1.0 × 10−13 to 1.0 × 10−7 M. The 21-mer ssDNA of Mycobacterium tuberculosis was immobilized on a nano-ZrO2 film electrochemically deposited on a gold surface. With a detection limit of 65 mg mL−1 [472], this biosensor showed rapid diagnosis within 60 s. The biosensor was developed via a V2O5 nanobelt, MWCNTs, and chitosan nanocomposite material that was modified onto a carbon ionic liquid electrode (CILE) and immobilized with ssDNA for Yersinia enterocolitica detection. The biosensor detected complementary DNA at concentrations between 0.01 and 1.000 nM, with a detection limit of 1.76 pM [473].

6.3. Antibody-Immobilized Biosensors

Immunosensor sensitivity and stability were affected by the orientation, surface density, and antigen-binding efficiency of antibodies when immobilized onto functionalized surfaces. A model antibody–antigen system that represents the complex matrix immuno sensors encountered in reality was used to improve various surface functionalization processes and assess their effectiveness. Protein A/G enhanced antibody loading on surfaces substantially more than boronate ester chemistry. In spite of the fact that both enhance antigen binding by assisting in orientation-specific immobilization of antibodies, using protein A/G enhanced the antibody surface density, which is crucial to obtaining maximum antigen recognition [474,475].
In immunoassays, electrochemical immunosensors are used to detect antigens, antibodies, or other biochemical targets related to health issues, such as cancer antigens in serum and bacteria in food [476]. A CH-MnO2/MWNT-Ag composite impregnated with anti-AFP was electrodeposited on a CH-MnO2/MWNT-Ag composite developed by Che and co-workers [477]. Because of the high surface area and conductivity of the MWCNT-Ag, the detection range for AFP was 0.25–250 ng mL−1. Anti-AFP antibody immobilization on a ZnO/PAC nanowire FET allows for real-time, label-free detection of liver cancer markers (Figure 29) [478]. In addition, it was demonstrated that the biosensor could be used as a pH sensor. In an electrochemical immune sensor, Wei and his team utilized dumbbell-like Au–Fe3O4 nanoparticles for detecting prostate-specific antigen (PSA), which is a cancer biomarker. PSA detection was achieved by immobilizing primary anti-PSA antibodies on graphene and secondary antibodies on Au–Fe3O4 nanoparticles [479]. The immune sensor has a wide linear range of detection (0.01–10 ng mL−1), a low detection limit (5 pg mL−1), and excellent reproducibility and stability.

6.4. Other Biomaterial-Immobilized Biosensors

Graham and coworkers demonstrated the fabrication of a SMOx-based bioelectrode for on-chip, long-term, and noninvasive cell culture assays and label-free high-content screening. In addition to detecting the fast electrical activity of neurons, the bioelectrode detected slow changes in impedance as the cells grew and divided. The team showed that a silver- and TiO2-based biosensor could be used to analyze prokaryotic gene expression in real time [480] Table 4 compares SMOx-based biosensors.
Table 4. SMOx-based immobilized biosensors.
Table 4. SMOx-based immobilized biosensors.
Target
Biomolecule
ElectrodeSensitivityDetection
Limit (mM)
Linear Range (mM)Response Time (s)/
Potential (V)
Ref.
GlucoseGCE/ZnO NF/PVA/GOx/L-Cys70.2 mA mM−1 cm−210.25–19<4/+0.80[438]
GCE/ZnO-HNSPs/GOx/Nafion65.82 mA mM−1 cm−210.005–13.15<5/+0.8[441]
GCE/ZnO NRs/GOx/CHIT25.7 mA mM−1 cm−2100.01–0.25/0.3–0.7<2/+0.8[481]
Au/ZnO NT/GOx/Nafion21.7 mA mM−1 cm−210.05–12.03/+0.8[482]
Au/ZnO nano-tetrapods/GOx/Nafion25.3 mA mM−1 cm−240.005–6.5<6/+0.8[483]
ITO/ZnO NT arrays/GOx/Nafion30.85 mA mM−1 cm−2100.01–4.2<6/+0.80[484]
PET/Au/ZnO-NWs/GOx/Nafion19.5 mA mM−1 cm−2<500.2–2.0<5/+0.80[485]
PDDA/GOx/ZnO/MWNTs50.2 mA mM−1 cm−20.250.1–16–/–[486]
ITO/Cu/ZnO/HRP-GOx/Con A/CS-Au0.097 mA mM−1 cm−2401.0–15.0<6/−0.39[442]
GCE/porous TiO2/GOx/Nafion0.3 mA mM−1 cm−2-0.15–1.2<10/−0.45[487]
GCE/TiO2-GR/GOx6.2 mA mM−1 cm−2-0–8.0–/−0.60[488]
Au/CuO/GOx/Nafion47.19 mA mM−1 cm−21.370.01–10.0<5/-[489]
ITO/CeO2 NRs/GOx0.165 mA mM−1 cm−21002.0–26.01–2/+0.80[490]
Pt/GOx/Fe3O4/Chitosan/Nafion11.54 mA mM−1 cm−260.006–2.2–/–[491]
Au/MgO/GOx/Nafion31.6 mA mM−1 cm−20.0680.001–0.009<5/[492]
Pt/NiO doped ZnONRs/GOx61.78 mA mM−1 cm−22.50.5–8.0<5/+0.39[439]
GCE/NiO/GOx/CHIT3.43 mA mM−1 cm−2471.5–7<8/+0.[440]
CholesterolAu/flower-shaped ZnO/ChOx/Nafion61.7 mA mM−1 cm20.0121.0–15.0<5/-[434]
Au/ZnO NPs/ChOx/Nafion23.7 mA mM−1 cm20.000370.001–0.5<5/+0.355[493]
Ag/ZnO/ChOx35.2 mV per decade-0.001–10.0–/–[494]
ITO/NS-CeO2/ChOx2 mA mg dL−1 cm2-0.26–10.36~15/+0.50[495]
ITO/CH-SnO2/ChOx34.7 mA mg dL−1 cm21300.26–10.360.5/-[458]
ITO/NanoFe3O4/ChOx86 Ω mg−1 dL cm−26.50.0065–10.3625/+0.06[450]
Nucleic acidssDNA/ZnO/MWNTs/
CHIT/GCE
-2.8 × 10−12 mol L−11.0 × 10−11 − 1.0 ×
10−6 mol L−1
-[496]
ssDNA/AuNPs/MWNTs/
ZnO NWs/GCE
-3.5 × 10−14 M1.0 × 10−13 − 1.0 ×
10−7 M
-[466]
ssDNA/ZnO/CILE-2.5 × 10−13 mol L−11.0 × 10−12 − 1.0 ×
10−8 mol L−1
-[467]
ssDNA/Cu2O/CPE-1.0 × 10−10 mol L−11.0 × 10−10 − 1 × 10−6 mol L−1-[469]
ssDNA/CeO2-SWNTs BMIMPF6/GCE-2.3 × 10−13 mol L−11.0 × 10−12 − 1.0 ×
10−7 mol L−1
-[497]
ssCT-DNA/CH-Fe3O4/ITO-0.0025 ppm1–300 ppm-[462]
PNA/Fe3O4-GOPS/ITO-0.1 × 10−15 M0.1 × 10−15 − 50.0 ×
10−15 M
-[470]
AntibodyBSA/anti-AFP/CH-MnO2/
MWNT-Ag/GCE
-0.08 ng mL−10.25–250 ng mL−1-[477]
BSA/r-IgGs/Nano-ZnO/ITO189 Ω nM−1 dm−3 cm−20.006 nM dm−30.006–0.01 nM dm−3-[498]
Anti-CEA/Fe3O4 NRs/CPE-0.9 ng mL−11.5–80 ng mL−1-[499]
HRP-anti-hIgGAu/SiO2 NPs PTHGCE-0.035 ng mL−10.1–200 ng mL−1-[500]

6.5. Non-Enzymatic Biosensors

The direct electrochemistry of glucose (oxidation or reduction) was used for non-enzymatic glucose sensing, which was quick and inexpensive [501]. The direct oxidation of glucose using noble metal electrodes, however, has three significant drawbacks [502,503]: (1) limited glucose sensitivity due to the slow electrooxidation kinetics of glucose with conventional electrodes, (2) low selectivity because several sugars can be oxidized in the same potential range as glucose, and (3) decreased electrode activity due to ion contamination. An increased electrode surface area enabled more glucose to come into contact with the electrode surface, thus eliminating the sensitivity and selectivity limitations. In the non-enzymatic process of glucose oxidation, hydrogen atoms are abstracted concurrently with organic species adsorption [504]. This is the rate-determining step in the glucose electrooxidation catalytic process. IHOAM is a hypothesis put forth by Bruke et al. to explain the intricate electrocatalytic process of glucose [504].
Metals, particularly noble metals, were investigated as electrode materials for non-enzymatic glucose biosensors [501,502]. Several metal alloys and hybrid materials have been developed as a result of advances in materials science in order to improve the properties of noble metals and metal oxides alone. Xiao et al. developed a flexible electrochemical glucose sensor by incorporating the nanocomposite PtAu alloy and MnO2 into graphene paper [505]. As a result of electrodeposition on graphene paper, a PtAu-MnO2 nanocomposite was developed with tight contact between the PtAu alloy and the MnO2. This glucose sensor had a linear range of 0.1 mM to 30 mM and a high sensitivity of 58.54 μA mM−1 cm−2 [505]. Lee et al. created a disposable non-enzymatic blood glucose sensor strip using microporous Pt as an electrode material and poly(vinyl acetate) as a binding material. The mixture was then applied to a polyimide film surface with a conducting circuit screen printed on it. In whole human blood, the sensor showed acceptable stabilization for 30 days with a sensitivity of 0.0054 μA cm−2 mgdL−1 [506]. An electrode fabricated from nanoporous Cu (NPC) was used by Chen et al. to make a portable micro glucose sensor [507]. In this non-enzymatic sensor, CuO nanocoral arrays had high conductivity and high glucose catalytic activity. It had a linear range of 0.0005 to 5 mM and a sensitivity of 1621 μA mM−1 cm−2. Liu et al. combined a wet chemical process with an annealing procedure to create 3D copper oxide nanowire arrays (CuONWAs). In the end, the CuONWA/CF platform served as a glucose sensor [508]. Because the copper foam and nanowire arrays increase the surface area of the device, its sensitivity is improved. By using tellurium microtubes, Guascito et al. altered the surface of a Pt electrode using a drop-casting technique. It was found that the non-enzymatic glucose sensor was more sensitive, stable, and reproducible when compared with a Pt electrode that had not been modified [509].

7. Conclusions

A SMOx-based sensor translates a response into an electrical signal by using a receptor-transducer device. A wide range of applications, such as the detection of diseases and illnesses, environmental monitoring, water and food quality monitoring, and drug delivery, prompted scientists and researchers to develop more sensitive and selective sensors. SMOx-based sensors need to capture recognition signals efficiently and convert them into electrochemical, electrical, optical, gravimetric, or acoustic signals (transduction process). Increasing transducer performance is another challenge, as it allows for increased sensitivity, faster response times, reproducibility, and lowering detection limits, even to detect single molecules and miniaturization of the sensing devices. Combining sensing technology with nano-SMOx-based devices, like zero- to three-dimensional FETs and IoT, with high surface-to-volume ratios, good conductivities, shock-bearing properties, and colour tuning can overcome these challenges. In this review, we provide an overview of the development of SMOx-based sensors.
In this review, we discussed, the recent advancements in semiconductor metal oxides (SMOx) for gas sensing, chemical sensing, and biosensing applications. The unique intrinsic chemical, physical, optical, and electronic properties of SMOxs entail low detection limits, high sensitivity, and fast response time, making SMOx materials a popular choice for sensing applications. Over the past few decades, the synthesis of SMOx-based nanomaterials in varying sizes, structures, and crystal morphologies has enabled the detection of gases, such as H2, CO, O2, SO2, NO2, and H2S, and various chemicals and biomolecules like glucose, cholesterol, nucleic acids, and other important biomolecules. SMOx materials possess a large specific surface area, superior electron transport rate, extraordinary permeability, and active reaction sites, making them well-suited for stable sensing of specific gases, chemicals, and biomolecules at room temperature. Different architectural morphologies of nano-SMOxs are used to accelerate the rate of electron transport, resulting in improved sensitivity.
This review also highlights the improvement of sensing properties through various strategies, such as loading with nanomaterials, doping with elements, and constructing heterojunctions with other functional materials. The sensing properties of SMOx-based nanoparticles were found to be extensively improved by decorating their surfaces with nanomaterials, which change electron accumulation and enhance their catalytic effect through electronic and chemical excitation, respectively. The incorporation of a large number of dopants into the lattice of SMOx changes their crystal and electronic structure, reducing the bandwidth and increasing the active sites on the surface, which affects the sensitivity and selectivity. The formation of heterojunctions effectively rectifies electron transfer on the surface between two materials, resulting in improved sensitivity. The composites of SMOx with other functional materials also enhance the sensitivity and response rate at low temperatures due to possible synergistic effects and defect structures. These developments in SMOx composites have the potential to enable their practical applications.

8. Future Aspects

Despite the significant advancements in the development of semiconductor metal oxide (SMOx)-based sensors, there is still room for improvement in terms of selectivity, sensitivity, and working temperature. Recent studies highlighted the potential use of nanostructured SMOx-based sensors for their applications in sensing fields, specifically as gas, chemical, and biological sensors. However, new materials and heterostructure designs require special attention for further development mechanisms, including influencing the sensing activity. The future of SMOx-based sensor research can focus on (1) developing new materials and new heterojunction interfaces and (2) exploring the mechanisms that influence sensing activity and improving selectivity and sensitivity for the detection of gas, chemical, and biological molecules. The implementation of new strategies, such as doping with electronically active materials, functionalization of SMOxs with unique functional groups, and development of sensing materials, is crucial in fabricating more sensitive novel sensors. The use of SMOx-based sensors was found to offer several advantages over traditional sensors, including higher sensitivity, exceptional selectivity, quick response time, low detection limits, and compact size. The objective of this review was to aid in the development of a new generation of sensing tools that can identify a wide variety of molecules in a variety of contexts, such as environmental danger gases and therapeutic biomolecules. Rapid and precise detection of gases, chemicals, and biological materials is crucial for an effective response. Therefore, to achieve this goal, a hybrid sensor system might be highly desirable. Moreover, integrated sensors for the detection of hybridized gas; chemicals; a mixture of toxic gases; and hazardous biological substances, like pathogenic microorganisms, bacteria, and viruses, can be realized with SMOx materials. These unique SMOx-based sensors can be utilized in a wide range of applications, such as in hospitals, defense areas, or war zones; for pharmaceutical, pesticide, textile, and meat industries; and in houses. These sensors also detect the growing threat of natural infectious diseases or industrial accidents. In summary, SMOx-based sensors have the potential to uncover new opportunities for the detection, identification, and quantification of toxic gases in various settings, such as food, hospitals, and the ocean. These new and innovative approaches must be highly selective, sensitive, reliable, fast-responding, and capable of autonomous screening. Moreover, they should be able to transmit information securely and wirelessly in real time. Also, SMOxs can be used for different types of optical sensor fabrications, including fiber optics and waveguide-based sensors, and have a lot of advantages over other types of sensors [508].

Author Contributions

S.K.M., T.D. and S.T. discussed the plan and agreed on it. T.D., S.K.M. and S.T. drafted designs for the manuscript. The manuscript was written by T.D., T.N., S.T. and S.K.M. They all reviewed and commented on the original draft of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This material is based upon work supported by ST’s National Science Foundation award under Grant No. ECCS-2138701 and VentureWell Grant 21716-20.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B 2018, 229, 206–217. [Google Scholar] [CrossRef]
  2. Usha, S.P.; Mishra, S.K.; Gupta, B.D. Fiber optic hydrogen sulfide gas sensors utilizing ZnO thin film/ZnO nanoparticles A comparison of surface plasmon resonance and lossy mode resonance. Sens. Actuators B Chem. 2015, 218, 196–204. [Google Scholar] [CrossRef]
  3. Amoah, P.K.; Hassan, Z.M.; Lin, P.; Redel, E.; Baumgart, H.; Obeng, Y.S. Broadband dielectric spectroscopic detection of ethanol: A side-by-side comparison of ZnO and HKUST-1 MOFs as sensing media. Chemosensors 2022, 10, 241. [Google Scholar] [CrossRef]
  4. Tabassum, R.; Mishra, S.K.; Gupta, B.D. Surface plasmon resonance-based fiber optic hydrogen sulphide gas sensor utilizing Cu–ZnO thin films. Phys. Chem. Chem. Phys. 2013, 15, 11868–11874. [Google Scholar] [CrossRef]
  5. Mirzaei, A.; Leonardi, S.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  6. Paghi, A.; Mariani, S.; Barillaro, G. 1D and 2D Field Effect Transistors in Gas Sensing: A Comprehensive Review. Small 2023, 19, 2206100. [Google Scholar] [CrossRef]
  7. Mishra, S.K.; Rani, S.; Gupta, B.D. Surface plasmon resonance based fiber optic hydrogen sulphide gas sensor utilizing nickel oxide doped ITO thin film. Sens. Actuators B Chem. 2014, 195, 215–222. [Google Scholar] [CrossRef]
  8. Usha, S.P.; Mishra, S.K.; Gupta, B.D. Fabrication and characterization of a SPR based fiber optic sensor for the detection of chlorine gas using silver and zinc oxide. Materials 2015, 8, 2204–2216. [Google Scholar] [CrossRef]
  9. He, H. Metal oxide semiconductors and conductors. In Solution Processed Metal Oxide Thin Films for Electronic Applications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 7–30. [Google Scholar]
  10. Mishra, S.K.; Gupta, B.D. Surface plasmon resonance-based fiber optic chlorine gas sensor utilizing indium-oxide-doped tin oxide film. J. Light. Technol. 2015, 33, 2770–2776. [Google Scholar] [CrossRef]
  11. Doan, T.H.P.; Hong, W.G.; Noh, J.-S. Palladium nanoparticle-decorated multi-layer Ti3C2Tx dual-functioning as a highly sensitive hydrogen gas sensor and hydrogen storage. RSC Adv. 2021, 11, 7492–7501. [Google Scholar] [CrossRef]
  12. 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]
  13. Franke, M.E.; Koplin, T.J.; Simon, U. Metal and metal oxide nanoparticles in chemiresistors: Does the nanoscale matter? Small 2006, 2, 36–50. [Google Scholar] [CrossRef] [PubMed]
  14. Usha, S.P.; Mishra, S.K.; Gupta, B.D. Zinc oxide thin film/nanorods based lossy mode resonance hydrogen sulphide gas sensor. Mater. Res. Express 2015, 2, 095003. [Google Scholar] [CrossRef]
  15. Mishra, A.K.; Mishra, S.K. MgF2 prism/rhodium/graphene: Efficient refractive index sensing structure in optical domain. J. Phys. Condens. Matter 2017, 29, 145001. [Google Scholar] [CrossRef] [PubMed]
  16. Mishra, A.K.; Mishra, S.K.; Gupta, B.D. Gas-clad two-way fiber optic SPR sensor: A novel approach for refractive index sensing. Plasmonics 2015, 10, 1071–1076. [Google Scholar] [CrossRef]
  17. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  18. Mishra, S.; Mishra, A.; Misra, K.; Verma, R. Detection of alcohol content in food products by lossy mode resonance technique. J. Electrochem. Soc. 2022, 169, 077504. [Google Scholar]
  19. Yamazoe, N.; Sakai, G.; Shimanoe, K. Oxide semiconductor gas sensors. Catal. Surv. Asia 2003, 7, 63–75. [Google Scholar] [CrossRef]
  20. 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]
  21. Moseley, P.T. Progress in the development of semiconducting metal oxide gas sensors: A review. Meas. Sci. Technol. 2017, 28, 082001. [Google Scholar] [CrossRef]
  22. Mishra, S.K.; Verma, R.K.; Mishra, A.K. Versatile Sensing Structure: GaP/Au/Graphene/Silicon. Photonics 2021, 8, 547. [Google Scholar] [CrossRef]
  23. Gaur, D.S.; Purohit, A.; Mishra, S.K.; Mishra, A.K. An Interplay between Lossy Mode Resonance and Surface Plasmon Resonance and Their Sensing Applications. Biosensors 2022, 12, 721. [Google Scholar] [CrossRef] [PubMed]
  24. Batzill, M. Surface science studies of gas sensing materials: SnO2. Sensors 2006, 6, 1345–1366. [Google Scholar] [CrossRef]
  25. Barsan, N.; Schweizer-Berberich, M.; Göpel, W. Fundamental and practical aspects in the design of nanoscaled SnO2 gas sensors: A status report. Fresenius’ J. Anal. Chem. 1999, 365, 287–304. [Google Scholar] [CrossRef]
  26. Batzill, M.; Diebold, U. The surface and materials science of tin oxide. Prog. Surf. Sci. 2005, 79, 47–154. [Google Scholar] [CrossRef]
  27. Lu, J.G.; Chang, P.; Fan, Z. Quasi-one-dimensional metal oxide materials—Synthesis, properties and applications. Mater. Sci. Eng. R Rep. 2006, 52, 49–91. [Google Scholar] [CrossRef]
  28. Rothschild, A.; Komem, Y. On the relationship between the grain size and gas-sensitivity of chemo-resistive metal-oxide gas sensors with nanosized grains. J. Electroceramics 2004, 13, 697–701. [Google Scholar] [CrossRef]
  29. Comini, E.; Baratto, C.; Faglia, G.; Ferroni, M.; Vomiero, A.; Sberveglieri, G. Quasi-one dimensional metal oxide semiconductors: Preparation, characterization and application as chemical sensors. Prog. Mater. Sci. 2009, 54, 1–67. [Google Scholar] [CrossRef]
  30. Korotcenkov, G. Gas response control through structural and chemical modification of metal oxide films: State of the art and approaches. Sens. Actuators B Chem. 2005, 107, 209–232. [Google Scholar] [CrossRef]
  31. Eranna, G.; Joshi, B.; Runthala, D.; Gupta, R. Oxide materials for development of integrated gas sensors—A comprehensive review. Crit. Rev. Solid State Mater. Sci. 2004, 29, 111–188. [Google Scholar] [CrossRef]
  32. Şerban, I.; Enesca, A. Metal oxides-based semiconductors for biosensors applications. Front. Chem. 2020, 8, 354. [Google Scholar] [CrossRef] [PubMed]
  33. Sato, H.; Minami, T.; Takata, S.; Yamada, T. Transparent conducting p-type NiO thin films prepared by magnetron sputtering. Thin Solid Film 1993, 236, 27–31. [Google Scholar] [CrossRef]
  34. Kumbhakar, P.; Chowde Gowda, C.; Mahapatra, P.L.; Mukherjee, M.; Malviya, K.D.; Chaker, M.; Chandra, A.; Lahiri, B.; Ajayan, P.M.; Jariwala, D.; et al. Emerging 2D metal oxides and their applications. Mater. Today 2021, 45, 142–168. [Google Scholar] [CrossRef]
  35. Chavali, M.S.; Nikolova, M.P. Metal oxide nanoparticles and their applications in nanotechnology. SN Appl. Sci. 2019, 1, 607. [Google Scholar] [CrossRef]
  36. Korotcenkov, G.; Cho, B. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B Chem. 2017, 244, 182–210. [Google Scholar] [CrossRef]
  37. Usha, S.P.; Shrivastav, A.M.; Gupta, B.D. Semiconductor metal oxide/polymer based fiber optic lossy mode resonance sensors: A contemporary study. Opt. Fiber Technol. 2018, 45, 146–166. [Google Scholar] [CrossRef]
  38. Espid, E.; Taghipour, F.J.C.R.i.S.S.; Sciences, M. UV-LED Photo-activated Chemical Gas Sensors: A Review. Crit. Rev. Solid State Mater. Sci. 2017, 42, 416–432. [Google Scholar] [CrossRef]
  39. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconducting metal oxides as sensors for environmentally hazardous gases. Sens. Actuators B Chem. 2011, 160, 580–591. [Google Scholar] [CrossRef]
  40. Balamurugan, C.; Song, S.-J.; Kim, H.-S.; Balamurugan, C.; Song, S.-J.; Kim, H.-S. Enhancing gas response characteristics of mixed metal oxide gas sensors. J. Korean Ceram. Soc. 2018, 55, 1–20. [Google Scholar] [CrossRef]
  41. Kanan, S.M.; El-Kadri, O.M.; Abu-Yousef, I.A.; Kanan, M.C. Semiconducting metal oxide based sensors for selective gas pollutant detection. Sensors 2009, 9, 8158–8196. [Google Scholar] [CrossRef]
  42. Tomchenko, A.A.; Harmer, G.P.; Marquis, B.T.; Allen, J.W. Semiconducting metal oxide sensor array for the selective detection of combustion gases. Sens. Actuators B Chem. 2003, 93, 126–134. [Google Scholar] [CrossRef]
  43. Afzal, A.; Cioffi, N.; Sabbatini, L.; Torsi, L. NOx sensors based on semiconducting metal oxide nanostructures: Progress and perspectives. Sens. Actuators B Chem. 2012, 171, 25–42. [Google Scholar] [CrossRef]
  44. Huang, J.; Wan, Q. Gas sensors based on semiconducting metal oxide one-dimensional nanostructures. Sensors 2009, 9, 9903–9924. [Google Scholar] [CrossRef]
  45. Pinna, N.; Neri, G.; Antonietti, M.; Niederberger, M. Nonaqueous synthesis of nanocrystalline semiconducting metal oxides for gas sensing. Angew. Chem. 2004, 116, 4445–4449. [Google Scholar] [CrossRef]
  46. Concina, I.; Ibupoto, Z.H.; Vomiero, A. Semiconducting metal oxide nanostructures for water splitting and photovoltaics. Adv. Energy Mater. 2017, 7, 1700706. [Google Scholar] [CrossRef]
  47. Wagner, T.; Waitz, T.; Roggenbuck, J.; Fröba, M.; Kohl, C.-D.; Tiemann, M. Ordered mesoporous ZnO for gas sensing. Thin Solid Film. 2007, 515, 8360–8363. [Google Scholar] [CrossRef]
  48. Szilágyi, I.M.; Saukko, S.; Mizsei, J.; Tóth, A.L.; Madarász, J.; Pokol, G. Gas sensing selectivity of hexagonal and monoclinic WO3 to H2S. Solid State Sci. 2010, 12, 1857–1860. [Google Scholar] [CrossRef]
  49. Brezesinski, T.; Fattakhova Rohlfing, D.; Sallard, S.; Antonietti, M.; Smarsly, B.M. Highly crystalline WO3 thin films with ordered 3D mesoporosity and improved electrochromic performance. Small 2006, 2, 1203–1211. [Google Scholar] [CrossRef]
  50. Cheng, J.; Wang, J.; Li, Q.; Liu, H.; Li, Y. A review of recent developments in tin dioxide composites for gas sensing application. J. Ind. Eng. Chem. 2016, 44, 1–22. [Google Scholar] [CrossRef]
  51. Cheng, J.; Liu, L.; Zhang, J.; Liu, F.; Zhang, X. Influences of anion exchange and phase transformation on the supercapacitive properties of α-Co (OH) 2. J. Electroanal. Chem. 2014, 722, 23–31. [Google Scholar] [CrossRef]
  52. Yang, X.; Cao, C.; Hohn, K.; Erickson, L.; Maghirang, R.; Hamal, D.; Klabunde, K. Highly visible-light active C-and V-doped TiO2 for degradation of acetaldehyde. J. Catal. 2007, 252, 296–302. [Google Scholar] [CrossRef]
  53. Waitz, T.; Becker, B.; Wagner, T.; Sauerwald, T.; Kohl, C.-D.; Tiemann, M. Ordered nanoporous SnO2 gas sensors with high thermal stability. Sens. Actuators B Chem. 2010, 150, 788–793. [Google Scholar] [CrossRef]
  54. Zhou, X.; Cao, Q.; Huang, H.; Yang, P.; Hu, Y. Study on sensing mechanism of CuO–SnO2 gas sensors. Mater. Sci. Eng. B 2003, 99, 44–47. [Google Scholar] [CrossRef]
  55. Choi, K.S.; Park, S.; Chang, S.-P. Enhanced ethanol sensing properties based on SnO2 nanowires coated with Fe2O3 nanoparticles. Sens. Actuators B Chem. 2017, 238, 871–879. [Google Scholar] [CrossRef]
  56. Liu, H.; Chen, S.; Wang, G.; Qiao, S.Z. Ordered Mesoporous Core/Shell SnO2/C Nanocomposite as High-Capacity Anode Material for Lithium-Ion Batteries. Chem. A Eur. J. 2013, 19, 16897–16901. [Google Scholar] [CrossRef] [PubMed]
  57. Zhao, X.; Zhou, R.; Hua, Q.; Dong, L.; Yu, R.; Pan, C. Recent progress in ohmic/Schottky-contacted ZnO nanowire sensors. J. Nanomater. 2015, 2015, 7. [Google Scholar] [CrossRef]
  58. Eto, T.; Endo, S.; Imai, M.; Katayama, Y.; Kikegawa, T. Crystal structure of NiO under high pressure. Phys. Rev. B 2000, 61, 14984. [Google Scholar] [CrossRef]
  59. Gomaa, M.M.; Sayed, M.H.; Patil, V.L.; Boshta, M.; Patil, P.S. Gas sensing performance of sprayed NiO thin films toward NO2 gas. J. Alloys Compd. 2021, 885, 160908. [Google Scholar] [CrossRef]
  60. Simion, C.E.; Ghica, C.; Mihalcea, C.G.; Ghica, D.; Mercioniu, I.; Somacescu, S.; Florea, O.G.; Stanoiu, A. Insights about CO Gas-Sensing Mechanism with NiO-Based Gas Sensors—The Influence of Humidity. Chemosensors 2021, 9, 244. [Google Scholar] [CrossRef]
  61. Li, Q.; Zeng, W.; Li, Y. NiO-Based Gas Sensors for Ethanol Detection: Recent Progress. Sensors 2022, 2022, 1855493. [Google Scholar] [CrossRef]
  62. Reddy, G.; Reddy, S.; Swamy, B.E.K.; Kumar, M.; Harish, K.N.; Naveen, C.S.; Kumar, G.R.; Aravinda, T. Electrochemical Detection of Uric Acid by using NiO Nanoparticles Anal. Bioanal. Electrochem. 2022, 14, 432–443. [Google Scholar]
  63. Amin, S.; Tahira, A.; Solangi, A.; Mazzaro, R.; Ibupoto, Z.H.; Vomiero, A. A sensitive enzyme-free lactic acid sensor based on NiO nanoparticles for practical applications. Anal. Methods 2019, 11, 3578–3583. [Google Scholar] [CrossRef]
  64. Gomaa, M.M.; RezaYazdi, G.; Rodner, M.; Greczynski, G.; Boshta, M.; Osman, M.B.S.; Khranovskyy, V.; Eriksson, J.; Yakimova, R. Exploring NiO nanosize structures for ammonia sensing. J. Mater. Sci. Mater. Electron. 2018, 29, 11870–11877. [Google Scholar] [CrossRef]
  65. Yang, H.; Hu, Y.; Yin, X.; Huang, J.; Qiao, C.; Hu, Z.; He, C.; Huo, D.; Hou, C. A disposable and sensitive non-enzymatic glucose sensor based on a 3D-Mn-doped NiO nanoflower-modified flexible electrode. Analyst 2023, 148, 153–162. [Google Scholar] [CrossRef]
  66. Jauch, W.; Reehuis, M.; Bleif, H.; Kubanek, F.; Pattison, P. Crystallographic symmetry and magnetic structure of CoO. Phys. Rev. B 2001, 64, 052102. [Google Scholar] [CrossRef]
  67. Logothetis, E.M.; Park, K.; Meitzler, A.H.; Laud, K.R. Oxygen sensors using CoO ceramics. Appl. Phys. Lett. 2008, 26, 209–211. [Google Scholar] [CrossRef]
  68. Dai, C.; Chen, M.; Lin, Y.; Qi, R.; Luo, C.; Peng, H.; Lin, H. High performance gas sensors based on layered cobaltite nanoflakes with moisture resistance. Appl. Surf. Sci. 2022, 604, 154487. [Google Scholar] [CrossRef]
  69. Ma, A.; Park, H.J.; Seo, J.H.; Jang, K.Y.; Lee, H.K.; Kim, D.Y.; Lee, J.E.; Nam, K.M.; Lee, D.-S. Phase transition of non-equilibrium wurtzite CoO: Spontaneous deposition of sensing material for ultrasensitive detection of acetone. Sens. Actuators B Chem. 2020, 308, 127698. [Google Scholar] [CrossRef]
  70. Duan, J.; Yang, S.; Liu, H.; Gong, J.; Huang, H.; Zhao, X.; Zhang, R.; Du, Y. Single crystal SnO2 zigzag nanobelts. J. Am. Chem. Soc. 2005, 127, 6180–6181. [Google Scholar] [CrossRef]
  71. Masuda, Y. Recent advances in SnO2 nanostructure based gas sensors. Sens. Actuators B Chem. 2022, 364, 131876. [Google Scholar] [CrossRef]
  72. Liu, P.; Wang, J.; Jin, H.; Ge, M.; Zhang, F.; Wang, C.; Sun, Y.; Dai, N. SnO2 mesoporous nanoparticle-based gas sensor for highly sensitive and low concentration formaldehyde detection. RSC Adv. 2023, 13, 2256–2264. [Google Scholar] [CrossRef] [PubMed]
  73. Mei, L.; Chen, Y.; Ma, J. Gas Sensing of SnO2 Nanocrystals Revisited: Developing Ultra-Sensitive Sensors for Detecting the H2S Leakage of Biogas. Sci. Rep. 2014, 4, 6028. [Google Scholar] [CrossRef] [PubMed]
  74. Choi, P.G.; Izu, N.; Shirahata, N.; Masuda, Y. SnO2 Nanosheets for Selective Alkene Gas Sensing. ACS Appl. Nano Mater. 2019, 2, 1820–1827. [Google Scholar] [CrossRef]
  75. Li, B.; Zhou, Q.; Peng, S.; Liao, Y. Recent Advances of SnO2-Based Sensors for Detecting Volatile Organic Compounds. Sec. Nanosci. 2020, 8, 321. [Google Scholar] [CrossRef]
  76. Choi, P.G.; Izu, N.; Shirahata, N.; Masuda, Y. Fabrication and H2-Sensing Properties of SnO2 Nanosheet Gas Sensors. ACS Omega 2018, 3, 14592–14596. [Google Scholar] [CrossRef]
  77. Li, M.; Gu, Y.; Zhang, Y.; Gao, X.; Ge, S.; Wei, G. Quantitative prediction of ternary mixed gases based on an SnO2 sensor array and an SSA-BP neural network model. Phys. Chem. Chem. Phys. 2023, 25, 10935–10945. [Google Scholar] [CrossRef]
  78. Khatoon, Z.; Fouad, H.; Alothman, O.Y.; Hashem, M.; Ansari, Z.A.; Ansari, S.A. Doped SnO2 Nanomaterials for E-Nose Based Electrochemical Sensing of Biomarkers of Lung Cancer. ACS Omega 2020, 5, 27645–27654. [Google Scholar] [CrossRef]
  79. Mahajan, S.; Jagtap, S. Metal-oxide semiconductors for carbon monoxide (CO) gas sensing: A review. Appl. Mater. Today 2020, 18, 100483. [Google Scholar] [CrossRef]
  80. Bârsan, N.; Weimar, U. Understanding the fundamental principles of metal oxide based gas sensors; the example of CO sensing with SnO2 sensors in the presence of humidity. J. Phys. Condens. Matter 2003, 15, R813. [Google Scholar] [CrossRef]
  81. Kim, J.-H.; Kim, J.-Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Synergistic effects of SnO2 and Au nanoparticles decorated on WS2 nanosheets for flexible, room-temperature CO gas sensing. Sens. Actuators B Chem. 2021, 332, 129493. [Google Scholar] [CrossRef]
  82. Aarik, J.; Aidla, A.; Uustare, T.; Sammelselg, V. Morphology and structure of TiO2 thin films grown by atomic layer deposition. J. Cryst. Growth 1995, 148, 268–275. [Google Scholar] [CrossRef]
  83. Tian, X.; Cui, X.; Lai, T.; Ren, J.; Yang, Z.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Gas sensors based on TiO2 nanostructured materials for the detection of hazardous gases: A review. Nano Mater. Sci. 2021, 3, 390–403. [Google Scholar] [CrossRef]
  84. Nunes, D.; Fortunato, E.; Martins, R. Flexible nanostructured TiO2-based gas and UV sensors: A review. Discov. Mater. 2022, 2, 2. [Google Scholar] [CrossRef]
  85. Shooshtari, M.; Vollebregt, S.; Vaseghi, Y.; Rajati, M.; Pahlavan, S. The sensitivity enhancement of TiO2-based VOCs sensor decorated by gold at room temperature. Nanotechnology 2023, 34, 255501. [Google Scholar] [CrossRef]
  86. Chen, Y.; Liu, B.; Chen, Z.; Zuo, X. Innovative Electrochemical Sensor Using TiO2 Nanomaterials to Detect Phosphopeptides. Anal. Chem. 2021, 93, 10635–10643. [Google Scholar] [CrossRef]
  87. Kaur, N.; Singh, M.; Moumen, A.; Duina, G.; Comini, E. 1D Titanium Dioxide: Achievements in Chemical Sensing. Materials 2020, 13, 2974. [Google Scholar] [CrossRef]
  88. Rattu, G.; Krishna, P.M. TiO2 nanoparticles reagent based nonenzymatic label-free optical sensor for the rapid detection of L-lactate in apple juice. Sens. Actuators Rep. 2021, 3, 100067. [Google Scholar] [CrossRef]
  89. Hajareh Haghighi, F.; Mercurio, M.; Cerra, S.; Salamone, T.A.; Bianymotlagh, R.; Palocci, C.; Romano Spica, V.; Fratoddi, I. Surface modification of TiO2 nanoparticles with organic molecules and their biological applications. J. Mater. Chem. B 2023, 11, 2334–2366. [Google Scholar] [CrossRef]
  90. Amri, F.; Septiani, N.L.W.; Rezki, M.; Iqbal, M.; Yamauchi, Y.; Golberg, D.; Kaneti, Y.V.; Yuliarto, B. Mesoporous TiO2-based architectures as promising sensing materials towards next-generation biosensing applications. J. Mater. Chem. B 2021, 9, 1189–1207. [Google Scholar] [CrossRef]
  91. Yoshio, K.; Onodera, A.; Satoh, H.; Sakagami, N.; Yamashita, H. Crystal structure of ZnO: Li at 293 K and 19 K by x-ray diffraction. Ferroelectrics 2001, 264, 133–138. [Google Scholar] [CrossRef]
  92. Bragg, W.L. LXII. The crystalline structure of zinc oxide. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1920, 39, 647–651. [Google Scholar] [CrossRef]
  93. Franco, M.A.; Conti, P.P.; Andre, R.S.; Correa, D.S. A review on chemiresistive ZnO gas sensors. Sens. Actuators Rep. 2022, 4, 100100. [Google Scholar] [CrossRef]
  94. Zhou, S.; Yan, W.; Ling, M.; Liang, C. High-response H2 sensing performances of ZnO nanosheets modulated by oxygen vacancies. Inorg. Chem. Front. 2023, 10, 3255–3262. [Google Scholar] [CrossRef]
  95. Sha, R.; Basak, A.; Maity, P.C.; Badhulika, S. ZnO nano-structured based devices for chemical and optical sensing applications. Sens. Actuators Rep. 2022, 4, 100098. [Google Scholar] [CrossRef]
  96. Fallatah, A.; Kuperus, N.; Almomtan, M.; Padalkar, S. Sensitive Biosensor Based on Shape-Controlled ZnO Nanostructures Grown on Flexible Porous Substrate for Pesticide Detection. Sensors 2022, 22, 3522. [Google Scholar] [CrossRef]
  97. Krishna, M.S.; Singh, S.; Batool, M.; Fahmy, H.M.; Seku, K.; Shalan, A.E.; Lanceros-Mendez, S.; Zafar, M.N. A review on 2D-ZnO nanostructure based biosensors: From materials to devices. Mater. Adv. 2023, 4, 320–354. [Google Scholar] [CrossRef]
  98. Zoubir, J.; Bakas, I.; Qourzal, S.; Tamimi, M.; Assabbane, A. Electrochemical sensor based on a ZnO-doped graphitized carbon for the electrocatalytic detection of the antibiotic hydroxychloroquine. Application: Tap water and human urine. J. Appl. Electrochem. 2023, 53, 1279–1294. [Google Scholar] [CrossRef]
  99. Jarosch, D. Crystal structure refinement and reflectance measurements of hausmannite, Mn3O4. Mineral. Petrol. 1987, 37, 15–23. [Google Scholar] [CrossRef]
  100. Bigiani, L.; Zappa, D.; Maccato, C.; Gasparotto, A.; Sada, C.; Comini, E.; Barreca, D. Hydrogen Gas Sensing Performances of p-Type Mn3O4 Nanosystems: The Role of Built-in Mn3O4/Ag and Mn3O4/SnO2 Junctions. Nanomaterials 2020, 10, 511. [Google Scholar] [CrossRef]
  101. Lupan, O.; Ursaki, V.; Chai, G.; Chow, L.; Emelchenko, G.A.; Tiginyanu, I.; Gruzintsev, A.N.; Redkin, A.J.S.; Chemical, A.B. Selective hydrogen gas nanosensor using individual ZnO nanowire with fast response at room temperature. Sens. Actuators B Chem. 2010, 144, 56–66. [Google Scholar] [CrossRef]
  102. Zakrzewska, K. Gas sensing mechanism of TiO2-based thin films. Vacuum 2004, 74, 335–338. [Google Scholar] [CrossRef]
  103. Jimenez, I.; Arbiol, J.; Dezanneau, G.; Cornet, A.; Morante, J.J.S. Crystalline structure, defects and gas sensor response to NO2 and H2S of tungsten trioxide nanopowders. Sens. Actuators B Chem. 2003, 93, 475–485. [Google Scholar] [CrossRef]
  104. Zhang, Y.-H.; Chen, Y.-B.; Zhou, K.-G.; Liu, C.-H.; Zeng, J.; Zhang, H.-L.; Peng, Y. Improving gas sensing properties of graphene by introducing dopants and defects: A first-principles study. Nanotechnology 2009, 20, 185504. [Google Scholar] [CrossRef] [PubMed]
  105. Kim, K.; Lee, H.-B.-R.; Johnson, R.W.; Tanskanen, J.T.; Liu, N.; Kim, M.-G.; Pang, C.; Ahn, C.; Bent, S.F.; Bao, Z. Selective metal deposition at graphene line defects by atomic layer deposition. Nat. Commun. 2014, 5, 4781. [Google Scholar] [CrossRef] [PubMed]
  106. Ahn, M.-W.; Park, K.-S.; Heo, J.-H.; Park, J.-G.; Kim, D.-W.; Choi, K.J.; Lee, J.-H.; Hong, S.-H. Gas sensing properties of defect-controlled ZnO-nanowire gas sensor. Appl. Phys. Lett. 2008, 93, 263103. [Google Scholar] [CrossRef]
  107. Hübner, M.; Simion, C.E.; Tomescu-Stănoiu, A.; Pokhrel, S.; Bârsan, N.; Weimar, U. Influence of humidity on CO sensing with p-type CuO thick film gas sensors. Sens. Actuators B Chem. 2011, 153, 347–353. [Google Scholar] [CrossRef]
  108. Rothschild, A.; Litzelman, S.J.; Tuller, H.L.; Menesklou, W.; Schneider, T.; Ivers-Tiffée, E. Temperature-independent resistive oxygen sensors based on SrTi1− xFexO3− δ solid solutions. Sens. Actuators B Chem. 2005, 108, 223–230. [Google Scholar] [CrossRef]
  109. Zaleska, A. Doped-TiO2: A review. Recent Pat. Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  110. Li, S.-S.; Xia, J.-B. Electronic states of a hydrogenic donor impurity in semiconductor nano-structures. Phys. Lett. A 2007, 366, 120–123. [Google Scholar] [CrossRef]
  111. Waldrop, J.; Grant, R. Semiconductor heterojunction interfaces: Nontransitivity of energy-band discontiuities. Phys. Rev. Lett. 1979, 43, 1686. [Google Scholar] [CrossRef]
  112. Walsh, A.; Da Silva, J.L.; Wei, S.-H.; Körber, C.; Klein, A.; Piper, L.; DeMasi, A.; Smith, K.E.; Panaccione, G.; Torelli, P. Nature of the Band Gap of In2O3 Revealed by First-Principles Calculations and X-Ray Spectroscopy. Phys. Rev. Lett. 2022, 100, 167402. [Google Scholar] [CrossRef] [PubMed]
  113. Anothainart, K.; Burgmair, M.; Karthigeyan, A.; Zimmer, M.; Eisele, I. Light enhanced NO2 gas sensing with tin oxide at room temperature: Conductance and work function measurements. Sens. Actuators B Chem. 2003, 93, 580–584. [Google Scholar] [CrossRef]
  114. Heyd, J.; Peralta, J.E.; Scuseria, G.E.; Martin, R.L. Energy band gaps and lattice parameters evaluated with the Heyd-Scuseria-Ernzerhof screened hybrid functional. J. Chem. Phys. 2005, 123, 174101. [Google Scholar] [CrossRef] [PubMed]
  115. White, S.; Sham, L. Electronic properties of flat-band semiconductor heterostructures. Phys. Rev. Lett. 1981, 47, 879. [Google Scholar] [CrossRef]
  116. Sadeghi, E. Impurity binding energy of excited states in spherical quantum dot. Phys. E Low Dimens. Syst. Nanostructures 2009, 41, 1319–1322. [Google Scholar] [CrossRef]
  117. Zhuravlev, M.Y.; Tsymbal, E.Y.; Vedyayev, A. Impurity-assisted interlayer exchange coupling across a tunnel barrier. Phys. Rev. Lett. 2005, 94, 026806. [Google Scholar] [CrossRef]
  118. Langer, J.M.; Heinrich, H. Deep-level impurities: A possible guide to prediction of band-edge discontinuities in semiconductor heterojunctions. Phys. Rev. Lett. 1985, 55, 1414. [Google Scholar] [CrossRef]
  119. Batzill, M.; Diebold, U. Surface studies of gas sensing metal oxides. Phys. Chem. Chem. Phys. 2007, 9, 2307–2318. [Google Scholar] [CrossRef]
  120. Basu, S.; Bhattacharyya, P. Recent developments on graphene and graphene oxide based solid state gas sensors. Sens. Actuators B Chem. 2012, 173, 1–21. [Google Scholar] [CrossRef]
  121. Bittencourt, C.; Felten, A.; Espinosa, E.; Ionescu, R.; Llobet, E.; Correig, X.; Pireaux, J.-J. WO3 films modified with functionalised multi-wall carbon nanotubes: Morphological, compositional and gas response studies. Sens. Actuators B Chem. 2006, 115, 33–41. [Google Scholar] [CrossRef]
  122. Cui, S.; Pu, H.; Wells, S.A.; Wen, Z.; Mao, S.; Chang, J.; Hersam, M.C.; Chen, J. Ultrahigh sensitivity and layer-dependent sensing performance of phosphorene-based gas sensors. Nat. Commun. 2015, 6, 8632. [Google Scholar] [CrossRef] [PubMed]
  123. de Lacy Costello, B.; Ledochowski, M.; Ratcliffe, N.M. The importance of methane breath testing: A review. J. Breath Res. 2013, 7, 024001. [Google Scholar] [CrossRef] [PubMed]
  124. Dong, C.; Liu, X.; Han, B.; Deng, S.; Xiao, X.; Wang, Y. Nonaqueous synthesis of Ag-functionalized In2O3/ZnO nanocomposites for highly sensitive formaldehyde sensor. Sens. Actuators B Chem. 2016, 224, 193–200. [Google Scholar] [CrossRef]
  125. Comini, E. Metal oxide nano-crystals for gas sensing. Anal. Chim. Acta 2006, 568, 28–40. [Google Scholar] [CrossRef] [PubMed]
  126. Artzi-Gerlitz, R.; Benkstein, K.D.; Lahr, D.L.; Hertz, J.L.; Montgomery, C.B.; Bonevich, J.E.; Semancik, S.; Tarlov, M.J. Fabrication and gas sensing performance of parallel assemblies of metal oxide nanotubes supported by porous aluminum oxide membranes. Sens. Actuators B Chem. 2009, 136, 257–264. [Google Scholar] [CrossRef]
  127. da Silva, L.F.; M’peko, J.-C.; Catto, A.C.; Bernardini, S.; Mastelaro, V.R.; Aguir, K.; Ribeiro, C.; Longo, E. UV-enhanced ozone gas sensing response of ZnO-SnO2 heterojunctions at room temperature. Sens. Actuators B Chem. 2017, 240, 573–579. [Google Scholar] [CrossRef]
  128. Dhawale, D.; Salunkhe, R.; Patil, U.; Gurav, K.; More, A.; Lokhande, C. Room temperature liquefied petroleum gas (LPG) sensor based on p-polyaniline/n-TiO2 heterojunction. Sens. Actuators B Chem. 2008, 134, 988–992. [Google Scholar] [CrossRef]
  129. Fergus, J.W. Perovskite oxides for semiconductor-based gas sensors. Sens. Actuators B Chem. 2007, 123, 1169–1179. [Google Scholar] [CrossRef]
  130. Han, C.; Li, X.; Shao, C.; Li, X.; Ma, J.; Zhang, X.; Liu, Y. Composition-controllable p-CuO/n-ZnO hollow nanofibers for high-performance H2S detection. Sens. Actuators B Chem. 2019, 285, 495–503. [Google Scholar] [CrossRef]
  131. Dhawale, D.S.; Gujar, T.P.; Lokhande, C.D. TiO2 nanorods decorated with Pd nanoparticles for enhanced liquefied petroleum gas sensing performance. Anal. Chem. 2017, 89, 8531–8537. [Google Scholar] [CrossRef]
  132. Shinde, V.; Gujar, T.; Lokhande, C. LPG sensing properties of ZnO films prepared by spray pyrolysis method: Effect of molarity of precursor solution. Sens. Actuators B Chem. 2007, 120, 551–559. [Google Scholar] [CrossRef]
  133. Mitra, P.; Chatterjee, A.P.; Maiti, H.S. ZnO thin film sensor. Mater. Lett. 1998, 35, 33–38. [Google Scholar] [CrossRef]
  134. Lee, J.E.; Lim, C.K.; Park, H.J.; Song, H.; Choi, S.-Y.; Lee, D.-S. ZnO–CuO core-hollow cube nanostructures for highly sensitive acetone gas sensors at the ppb level. ACS Appl. Mater. Interfaces 2020, 12, 35688–35697. [Google Scholar] [CrossRef]
  135. Rai, P.; Majhi, S.; Yu, Y.; Lee, J. Noble metal oxide semiconductor shell nano-architectures as a new platform for gas sensor applications. RSC Adv. 2015, 5, 76229–76248. [Google Scholar] [CrossRef]
  136. Nakate, U.T.; Patil, P.; Ghule, B.G.; Ekar, S.; Al-Osta, A.; Jadhav, V.V.; Mane, R.S.; Kale, S.; Naushad, M.; O’Dwyer, C. Gold sensitized sprayed SnO2 nanostructured film for enhanced LPG sensing. J. Anal. Appl. Pyrolysis 2017, 124, 362–368. [Google Scholar] [CrossRef]
  137. Li, Y.; Chen, N.; Deng, D.; Xing, X.; Xiao, X.; Wang, Y. Formaldehyde detection: SnO2 microspheres for formaldehyde gas sensor with high sensitivity, fast response/recovery and good selectivity. Sens. Actuators B Chem. 2017, 238, 264–273. [Google Scholar] [CrossRef]
  138. Gutpa, J.; Shaik, H.; Kumar, K.N.; Sattar, S.A. PVD techniques proffering avenues for fabrication of porous tungsten oxide (WO3) thin films: A review. Mater. Sci. Semicond. Process. 2022, 143, 106534. [Google Scholar] [CrossRef]
  139. Saruhan, B.; Lontio Fomekong, R.; Nahirniak, S. Influences of semiconductor metal oxide properties on gas sensing characteristics. Front. Sens. 2021, 2, 657931. [Google Scholar] [CrossRef]
  140. Hussain, S.; Javed, M.S.; Ullah, N.; Shaheen, A.; Aslam, N.; Ashraf, I.; Abbas, Y.; Wang, M.; Liu, G.; Qiao, G. Unique hierarchical mesoporous LaCrO3 perovskite oxides for highly efficient electrochemical energy storage applications. Ceram. Int. 2019, 45, 15164–15170. [Google Scholar] [CrossRef]
  141. 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]
  142. Zhang, J.; Qin, Z.; Zeng, D.; Xie, C. Metal-oxide-semiconductor based gas sensors: Screening, preparation, and integration. Phys. Chem. Chem. Phys. 2017, 19, 6313–6329. [Google Scholar] [CrossRef] [PubMed]
  143. Sahay, P.J.P. Multifunctional metal oxide nanomaterials for chemical gas sensing. Procedia Eng. 2017, 215, 145–151. [Google Scholar] [CrossRef]
  144. Seiyama, T.; Kato, A.; Fujiishi, K.; Nagatani, M.J.A.C. A new detector for gaseous components using semiconductive thin films. 1962, 34, 1502–1503. Anal. Chem. 1962, 34, 1502–1503. [Google Scholar] [CrossRef]
  145. Hoa, N.D.; Duy, N.V.; El-Safty, S.A.; Hieu, N.V. Meso-/nanoporous semiconducting metal oxides for gas sensor applications. J. Nanomater. 2015, 16, 186. [Google Scholar] [CrossRef]
  146. Lewis, A.; Edwards, P. Validate personal air-pollution sensors. Nature 2016, 535, 29–31. [Google Scholar] [CrossRef]
  147. Kalantar-Zadeh, K.; Berean, K.J.; Ha, N.; Chrimes, A.F.; Xu, K.; Grando, D.; Ou, J.Z.; Pillai, N.; Campbell, J.L.; Brkljača, R. A human pilot trial of ingestible electronic capsules capable of sensing different gases in the gut. Nat. Electron. 2018, 1, 79–87. [Google Scholar] [CrossRef]
  148. Nugroho, F.A.; Darmadi, I.; Cusinato, L.; Susarrey-Arce, A.; Schreuders, H.; Bannenberg, L.J.; da Silva Fanta, A.B.; Kadkhodazadeh, S.; Wagner, J.B.; Antosiewicz, T.J. Metal–polymer hybrid nanomaterials for plasmonic ultrafast hydrogen detection. Nat. Mater. 2019, 18, 489–495. [Google Scholar] [CrossRef]
  149. van den Broek, J.; Abegg, S.; Pratsinis, S.E.; Güntner, A.T. Highly selective detection of methanol over ethanol by a handheld gas sensor. Nat. Commun. 2019, 10, 4220. [Google Scholar] [CrossRef]
  150. Tao, N. Challenges and promises of metal oxide nanosensors. ACS Sens. 2019, 4, 780. [Google Scholar] [CrossRef]
  151. Williams, D.E. Low cost sensor networks: How do we know the data are reliable? ACS Sens. 2019, 4, 2558–2565. [Google Scholar] [CrossRef]
  152. Hunter, G.W.; Akbar, S.; Bhansali, S.; Daniele, M.; Erb, P.D.; Johnson, K.; Liu, C.-C.; Miller, D.; Oralkan, O.; Hesketh, P.J. Editors’ choice—Critical review—A critical review of solid state gas sensors. J. Electrochem. Soc. 2020, 167, 037570. [Google Scholar] [CrossRef]
  153. Ihokura, K.; Watson, J. The Stannic Oxide Gas Sensorprinciples and Applications; CRC Press: Boca Raton, FL, USA, 2017. [Google Scholar]
  154. Staerz, A.; Suzuki, T.; Weimar, U.; Barsan, N. SnO2: The most important base material for semiconducting metal oxide-based materials. In Tin Oxide Materials; Elsevier: Amsterdam, The Netherlands, 2020; pp. 345–377. [Google Scholar]
  155. Stetter, J.R.; Penrose, W.R.; Yao, S. Sensors, chemical sensors, electrochemical sensors, and ECS. J. Electrochem. Soc. 2003, 150, S11. [Google Scholar] [CrossRef]
  156. Yamazoe, N. Toward innovations of gas sensor technology. Sens. Actuators B Chem. 2005, 108, 2–14. [Google Scholar] [CrossRef]
  157. Shankar, P.; Rayappan, J.B.B. Gas sensing mechanism of metal oxides: The role of ambient atmosphere, type of semiconductor and gases—A review. Sci. Lett. J 2015, 4, 126. [Google Scholar]
  158. Hussain, S.; Yang, X.; Aslam, M.K.; Shaheen, A.; Javed, M.S.; Aslam, N.; Aslam, B.; Liu, G.; Qiao, G. Robust TiN nanoparticles polysulfide anchor for Li–S storage and diffusion pathways using first principle calculations. Chem. Eng. J. 2020, 391, 123595. [Google Scholar] [CrossRef]
  159. Tiemann, M. Porous Metal Oxides as Gas Sensors. Chem. Eur. 2007, 13, 8376–8388. [Google Scholar] [CrossRef]
  160. Korotcenkov, G. Metal oxides for solid-state gas sensors: What determines our choice? Mater. Sci. Eng. B 2007, 139, 1–23. [Google Scholar] [CrossRef]
  161. Moseley, P.T.; Norris, J.O.; Williams, D.E. Techniques and Mechanisms in Gas Sensing; Adam Hilger: Bristol, UK, 1991; Volume 234. [Google Scholar]
  162. Lu, Z.; Kanan, S.M.; Tripp, C.P. Synthesis of high surface area monoclinic WO3 particles using organic ligands and emulsion based methods. J. Mater. Chem. 2002, 12, 983–989. [Google Scholar] [CrossRef]
  163. Williams, D.E. Semiconducting oxides as gas-sensitive resistors. Sens. Actuators B Chem. 1999, 57, 1–16. [Google Scholar] [CrossRef]
  164. Van de Krol, R.; Tuller, H. Electroceramics—The role of interfaces. Solid State Ion 2002, 150, 167–179. [Google Scholar] [CrossRef]
  165. Hulanicki, A.; Glab, S.; Ingman, F. Chemical sensors: Definitions and classification. Pure Appl. Chem. 1991, 63, 1247–1250. [Google Scholar] [CrossRef]
  166. 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. Electroceramics 2010, 25, 11–19. [Google Scholar] [CrossRef]
  167. Barsan, N.; Gauglitz, G.; Oprea, A.; Ostertag, E.; Proll, G.; Rebner, K.; Schierbaum, K.; Schleifenbaum, F.; Weimar, U. Chemical and biochemical sensors, 1. Fundamentals. In Ullmann’s Encyclopedia of Industrial Chemistry; John Wiley&Sons: Hoboken, NJ, USA, 2000; pp. 1–81. [Google Scholar] [CrossRef]
  168. Kuru, C.; Choi, C.; Kargar, A.; Choi, D.; Kim, Y.J.; Liu, C.H.; Yavuz, S.; Jin, S. MoS2 nanosheet–Pd nanoparticle composite for highly sensitive room temperature detection of hydrogen. Adv. Sci. 2015, 2, 1500004. [Google Scholar] [CrossRef] [PubMed]
  169. Henrich, V.E.; Cox, P.A. The Surface Science of Metal Oxides; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar]
  170. Dorenbos, P. The Eu3+ charge transfer energy and the relation with the band gap of compounds. J. Lumin. 2005, 111, 89–104. [Google Scholar] [CrossRef]
  171. Ueda, M.; Ohtsuka, T. Luminescence from band-gap photo-excitation of titanium anodic oxide films. Corros. Sci. 2002, 44, 1633–1638. [Google Scholar] [CrossRef]
  172. Palankovski, V.; Kaiblinger-Grujin, G.; Selberherr, S. Study of dopant-dependent band gap narrowing in compound semiconductor devices. Mater. Sci. Eng. B 1999, 66, 46–49. [Google Scholar] [CrossRef]
  173. Ducéré, J.M.; Hemeryck, A.; Estève, A.; Rouhani, M.D.; Landa, G.; Ménini, P.; Tropis, C.; Maisonnat, A.; Fau, P.; Chaudret, B. A computational chemist approach to gas sensors: Modeling the response of SnO2 to CO, O2, and H2O Gases. J. Comput. Chem. 2012, 33, 247–258. [Google Scholar] [CrossRef]
  174. Wang, X.; Qin, H.; Chen, Y.; Hu, J. Sensing mechanism of SnO2 surface to CO: Density functional theory calculations. J. Phys. Chem. C 2014, 118, 28548–28561. [Google Scholar] [CrossRef]
  175. Chen, Y.; Wang, X.; Shi, C.; Li, L.; Qin, H.; Hu, J. Sensing mechanism of SnO2 (1 1 0) surface to H2: Density functional theory calculations. Sens. Actuators B Chem. 2015, 220, 279–287. [Google Scholar] [CrossRef]
  176. Schierbaum, K.; Weimar, U.; Göpel, W.; Kowalkowski, R. Conductance, work function and catalytic activity of SnO2-based gas sensors. Sens. Actuators B Chem. 1991, 3, 205–214. [Google Scholar] [CrossRef]
  177. Morrison, S.R. Semiconductor gas sensors. Sens. Actuators 1981, 2, 329–341. [Google Scholar] [CrossRef]
  178. Barsan, N.; Weimar, U. Conduction model of metal oxide gas sensors. J. Electroceramics 2001, 7, 143–167. [Google Scholar] [CrossRef]
  179. Kim, I.-D.; Rothschild, A.; Tuller, H.L. Advances and new directions in gas-sensing devices. Acta Mater. 2013, 61, 974–1000. [Google Scholar] [CrossRef]
  180. Rebholz, J.; Bonanati, P.; Weimar, U.; Barsan, N. Grain shape influence on semiconducting metal oxide based gas sensor performance: Modeling versus experiment. Anal. Bioanal. Chem. 2014, 406, 3977–3983. [Google Scholar] [CrossRef] [PubMed]
  181. Yamazoe, N.; Shimanoe, K. Roles of shape and size of component crystals in semiconductor gas sensors: I. Response to oxygen. J. Electrochem. Soc. 2008, 155, J85. [Google Scholar] [CrossRef]
  182. 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]
  183. Cho, N.G.; Hwang, I.-S.; Kim, H.-G.; Lee, J.-H.; Kim, I.-D. Gas sensing properties of p-type hollow NiO hemispheres prepared by polymeric colloidal templating method. Sens. Actuators B Chem. 2011, 155, 366–371. [Google Scholar] [CrossRef]
  184. Li, D.; Hu, J.; Wu, R.; Lu, J.G. Conductometric chemical sensor based on individual CuO nanowires. Nanotechnology 2010, 21, 485502. [Google Scholar] [CrossRef]
  185. Choi, S.-W.; Park, J.Y.; Kim, S.S. Growth behavior and sensing properties of nanograins in CuO nanofibers. Chem. Eng. J. 2011, 172, 550–556. [Google Scholar] [CrossRef]
  186. Yamazoe, N. New approaches for improving semiconductor gas sensors. Sens. Actuators B Chem. 1991, 5, 7–19. [Google Scholar] [CrossRef]
  187. 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, 27, 1798. [Google Scholar] [CrossRef]
  188. Yamazoe, N.; Kurokawa, Y.; Seiyama, T. Effects of additives on semiconductor gas sensors. Sens. Actuators 1983, 4, 283–289. [Google Scholar] [CrossRef]
  189. Müller, S.A.; Degler, D.; Feldmann, C.; Türk, M.; Moos, R.; Fink, K.; Studt, F.; Gerthsen, D.; Bârsan, N.; Grunwaldt, J.D. Exploiting synergies in catalysis and gas sensing using noble metal-loaded oxide composites. ChemCatChem 2018, 10, 864–880. [Google Scholar] [CrossRef]
  190. Staerz, A.; Kim, T.-H.; Lee, J.-H.; Weimar, U.; Barsan, N. Nanolevel control of gas sensing characteristics via p–n heterojunction between Rh2O3 clusters and WO3 crystallites. J. Phys. Chem. C 2017, 121, 24701–24706. [Google Scholar] [CrossRef]
  191. Korotcenkov, G.; Brinzari, V.; Gulina, L.; Cho, B. The influence of gold nanoparticles on the conductivity response of SnO2-based thin film gas sensors. Appl. Surf. Sci. 2015, 353, 793–803. [Google Scholar] [CrossRef]
  192. Marikutsa, A.V.; Rumyantseva, M.N.; Frolov, D.D.; Morozov, I.V.; Boltalin, A.I.; Fedorova, A.A.; Petukhov, I.A.; Yashina, L.V.; Konstantinova, E.A.; Sadovskaya, E.M. Role of PdOx and RuOy clusters in oxygen exchange between nanocrystalline tin dioxide and the gas phase. J. Phys. Chem. C 2013, 117, 23858–23867. [Google Scholar] [CrossRef]
  193. Morrison, S.R. Selectivity in semiconductor gas sensors. Sens. Actuators 1987, 12, 425–440. [Google Scholar] [CrossRef]
  194. 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]
  195. Liu, C.; Kuang, Q.; Xie, Z.; Zheng, L. The effect of noble metal (Au, Pd and Pt) nanoparticles on the gas sensing performance of SnO2-based sensors: A case study on the {221} high-index faceted SnO2 octahedra. CrystEngComm 2015, 17, 6308–6313. [Google Scholar] [CrossRef]
  196. Marikutsa, A.V.; Rumyantseva, M.N.; Konstantinova, E.A.; Shatalova, T.B.; Gaskov, A.M. 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]
  197. Degler, D.; Rank, S.; Müller, S.; Pereira de Carvalho, H.W.; Grunwaldt, J.-D.; Weimar, U.; Barsan, N. Gold-loaded tin dioxide gas sensing materials: Mechanistic insights and the role of gold dispersion. ACS Sens. 2016, 1, 1322–1329. [Google Scholar] [CrossRef]
  198. Morrison, S. Roy the Chemical Physics of Surfaces; Plenum Press: New York, NY, USA; London, UK, 1977. [Google Scholar]
  199. Rebholz, J.; Dee, C.; Weimar, U.; Barsan, N. A self-doping surface effect and its influence on the sensor performance of undoped SnO2 based gas sensors. Procedia Eng. 2015, 120, 83–87. [Google Scholar] [CrossRef]
  200. Göpel, W.; Schierbaum, K.D. SnO2 sensors: Current status and future prospects. Sens. Actuators B Chem. 1995, 26, 1–12. [Google Scholar] [CrossRef]
  201. Yu, J.H.; Choi, G.M. Electrical and CO gas sensing properties of ZnO–SnO2 composites. Sens. Actuators B Chem. 1998, 52, 251–256. [Google Scholar]
  202. de Lacy Costello, B.; Ewen, R.; Jones, P.; Ratcliffe, N.; Wat, R. A study of the catalytic and vapour-sensing properties of zinc oxide and tin dioxide in relation to 1-butanol and dimethyldisulphide. Sens. Actuators B Chem. 1999, 61, 199–207. [Google Scholar] [CrossRef]
  203. Zhu, C.; Chen, Y.; Wang, R.; Wang, L.; Cao, M.; Shi, X. Synthesis and enhanced ethanol sensing properties of α-Fe2O3/ZnO heteronanostructures. Sens. Actuators B Chem. 2009, 140, 185–189. [Google Scholar] [CrossRef]
  204. Yoon, D.H.; Yu, J.H.; Choi, G.M. CO gas sensing properties of ZnO–CuO composite. Sens. Actuators B Chem. 1998, 46, 15–23. [Google Scholar] [CrossRef]
  205. Cobianu, C.; Serban, B.-C.; Dumbravescu, N.; Buiu, O.; Avramescu, V.; Pachiu, C.; Bita, B.; Bumbac, M.; Nicolescu, C.-M.; Cobianu, C. Organic–inorganic ternary nanohybrids of single-walled carbon nanohorns for room temperature chemiresistive ethanol detection. Nanomaterials 2020, 10, 2552. [Google Scholar] [CrossRef]
  206. Serban, B.-C.; Cobianu, C.; Buiu, O.; Bumbac, M.; Dumbravescu, N.; Avramescu, V.; Nicolescu, C.M.; Brezeanu, M.; Radulescu, C.; Craciun, G. Quaternary Holey Carbon Nanohorns/SnO2/ZnO/PVP Nano-Hybrid as Sensing Element for Resistive-Type Humidity Sensor. Coatings 2021, 11, 1307. [Google Scholar] [CrossRef]
  207. Haridas, D.; Gupta, V.; Sreenivas, K. Enhanced catalytic activity of nanoscale platinum islands loaded onto SnO2 thin film for sensitive LPG gas sensors. Bull. Mater. Sci. 2008, 31, 397–400. [Google Scholar] [CrossRef]
  208. Lu, Y.; Li, J.; Han, J.; Ng, H.-T.; Binder, C.; Partridge, C.; Meyyappan, M. Room temperature methane detection using palladium loaded single-walled carbon nanotube sensors. Chem. Phys. Lett. 2004, 391, 344–348. [Google Scholar] [CrossRef]
  209. Wang, D.; Ma, Z.; Dai, S.; Liu, J.; Nie, Z.; Engelhard, M.H.; Huo, Q.; Wang, C.; Kou, R. Low-temperature synthesis of tunable mesoporous crystalline transition metal oxides and applications as Au catalyst supports. J. Phys. Chem. C 2008, 112, 13499–13509. [Google Scholar] [CrossRef]
  210. Haridas, D.; Sreenivas, K.; Gupta, V. Improved response characteristics of SnO2 thin film loaded with nanoscale catalysts for LPG detection. Sens. Actuators B Chem. 2008, 133, 270–275. [Google Scholar] [CrossRef]
  211. Shimizu, Y.; Matsunaga, N.; Hyodo, T.; Egashira, M. Improvement of SO2 sensing properties of WO3 by noble metal loading. Sens. Actuators B Chem. 2001, 77, 35–40. [Google Scholar] [CrossRef]
  212. Ruiz, A.M.; Cornet, A.; Shimanoe, K.; Morante, J.R.; Yamazoe, N. Effects of various metal additives on the gas sensing performances of TiO2 nanocrystals obtained from hydrothermal treatments. Sens. Actuators B Chem. 2005, 108, 34–40. [Google Scholar] [CrossRef]
  213. Henry, C.; Chapon, C.; Duriez, C. Precursor state in the chemisorption of CO on supported palladium clusters. J. Chem. Phys. 1991, 95, 700–705. [Google Scholar] [CrossRef]
  214. Tsu, K.; Boudart, M. Recombination of atoms at the surface of thermocouple probes. Can. J. Chem. 1961, 39, 1239–1246. [Google Scholar] [CrossRef]
  215. Boudart, M. On the nature of spilt-over hydrogen’. J. Mol. Catal. A Chem. 1999, 138, 319–321. [Google Scholar] [CrossRef]
  216. Kolmakov, A.; Klenov, D.; Lilach, Y.; Stemmer, S.; Moskovits, M. Enhanced gas sensing by individual SnO2 nanowires and nanobelts functionalized with Pd catalyst particles. Nano Lett. 2005, 5, 667–673. [Google Scholar] [CrossRef]
  217. Ansari, S.; Boroojerdian, P.; Sainkar, S.; Karekar, R.; Aiyer, R.; Kulkarni, S. Grain size effects on H2 gas sensitivity of thick film resistor using SnO2 nanoparticles. Thin Solid Film. 1997, 295, 271–276. [Google Scholar] [CrossRef]
  218. Korotcenkov, G. The role of morphology and crystallographic structure of metal oxides in response of conductometric-type gas sensors. Mater. Sci. Eng. R: Rep. 2008, 61, 1–39. [Google Scholar] [CrossRef]
  219. Xu, C.; 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]
  220. Timmer, B.; Olthuis, W.; Van Den Berg, A. Ammonia sensors and their applications—A review. Sens. Actuators B Chem. 2005, 107, 666–677. [Google Scholar] [CrossRef]
  221. Wang, X.; Yee, S.S.; Carey, W.P. Transition between neck-controlled and grain-boundary-controlled sensitivity of metal-oxide gas sensors. Sens. Actuators B Chem. 1995, 25, 454–457. [Google Scholar] [CrossRef]
  222. Barsan, N. Conduction models in gas-sensing SnO2 layers: Grain-size effects and ambient atmosphere influence. Sens. Actuators B Chem. 1994, 17, 241–246. [Google Scholar] [CrossRef]
  223. Liu, H.; Gong, S.; Hu, Y.; Liu, J.; Zhou, D. Properties and mechanism study of SnO2 nanocrystals for H2S thick-film sensors. Sens. Actuators B Chem. 2009, 140, 190–195. [Google Scholar] [CrossRef]
  224. Rao, C.; Kulkarni, G.; Thomas, P.J.; Edwards, P.P. Size-dependent chemistry: Properties of nanocrystals. Chem. A Eur. J. 2002, 8, 28–35. [Google Scholar] [CrossRef]
  225. Lu, F.; Liu, Y.; Dong, M.; Wang, X. Nanosized tin oxide as the novel material with simultaneous detection towards CO, H2 and CH4. Sens. Actuators B Chem. 2000, 66, 225–227. [Google Scholar] [CrossRef]
  226. Serban, B.-C.; Cobianu, C.; Buiu, O.; Bumbac, M.; Dumbravescu, N.; Avramescu, V.; Nicolescu, C.M.; Brezeanu, M.; Pachiu, C.; Craciun, G. Ternary nanocomposites based on oxidized carbon nanohorns as sensing layers for room temperature resistive humidity sensing. Materials 2021, 14, 2705. [Google Scholar] [CrossRef]
  227. Serban, B.-C.; Buiu, O.; Bumbac, M.; Marinescu, R.; Dumbravescu, N.; Avramescu, V.; Cobianu, C.; Nicolescu, C.M.; Brezeanu, M.; Radulescu, C. Ternary Oxidized Carbon Nanohorns/TiO2/PVP Nanohybrid as Sensitive Layer for Chemoresistive Humidity Sensor. Chem. Proc. 2021, 5, 12. [Google Scholar]
  228. Qi, Q.; Zhang, T.; Zheng, X.; Fan, H.; Liu, L.; Wang, R.; Zeng, Y. Electrical response of Sm2O3-doped SnO2 to C2H2 and effect of humidity interference. Sens. Actuators B Chem. 2008, 134, 36–42. [Google Scholar] [CrossRef]
  229. Gong, J.; Chen, Q.; Lian, M.-R.; Liu, N.-C.; Stevenson, R.G.; Adami, F. Micromachined nanocrystalline silver doped SnO2 H2S sensor. Sens. Actuators B Chem. 2006, 114, 32–39. [Google Scholar] [CrossRef]
  230. Jing, Z.; Zhan, J. Fabrication and gas-sensing properties of porous ZnO nanoplates. Adv. Mater. 2008, 20, 4547–4551. [Google Scholar] [CrossRef]
  231. Malyshev, V.; Pislyakov, A. Investigation of gas-sensitivity of sensor structures to hydrogen in a wide range of temperature, concentration and humidity of gas medium. Sens. Actuators B Chem. 2008, 134, 913–921. [Google Scholar] [CrossRef]
  232. 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]
  233. Cao, M.; Wang, Y.; Chen, T.; Antonietti, M.; Niederberger, M. A highly sensitive and fast-responding ethanol sensor based on CdIn2O4 nanocrystals synthesized by a nonaqueous sol−gel route. Chem. Mater. 2008, 20, 5781–5786. [Google Scholar] [CrossRef]
  234. Van Duy, N.; Van Hieu, N.; Huy, P.T.; Chien, N.D.; Thamilselvan, M.; Yi, J. Mixed SnO2/TiO2 included with carbon nanotubes for gas-sensing application. Phys. E: Low-Dimens. Syst. Nanostruct. 2008, 41, 258–263. [Google Scholar] [CrossRef]
  235. Hasan, M.N.; Maity, S.; Sarkar, A.; Bhunia, C.T.; Acharjee, D.; Joseph, A.M. Simulation and fabrication of SAW-based gas sensor with modified surface state of active layer and electrode orientation for enhanced H2 gas sensing. J. Electron. Mater. 2017, 46, 679. [Google Scholar] [CrossRef]
  236. Sarkar, S.M.A.; Chakraborty, P.; Chakraborty, S.K. Factors Influencing on Sensitivity of the Metal Oxide Gas Sensors. Int. J. Sci. Res. 2017, 6, 372–376. [Google Scholar]
  237. Chang, C.; Hon, M.; Leu, I. Influence of size and density of Au nanoparticles on ZnO nanorod arrays for sensing reducing gases. J. Electrochem. Soc. 2013, 160, B170. [Google Scholar] [CrossRef]
  238. Liu, X.; Zhang, J.; Guo, X.; Wang, S.; Wu, S. Core–shell α–Fe2O3@ SnO2/Au hybrid structures and their enhanced gas sensing properties. RSC Adv. 2012, 2, 1650–1655. [Google Scholar] [CrossRef]
  239. Kim, H.R.; Haensch, A.; Kim, I.D.; Barsan, N.; Weimar, U.; Lee, J.H. The role of NiO doping in reducing the impact of humidity on the performance of SnO2-Based gas sensors: Synthesis strategies, and phenomenological and spectroscopic studies. Adv. Funct. Mater. 2011, 21, 4456–4463. [Google Scholar] [CrossRef]
  240. Ponzoni, A.; Comini, E.; Ferroni, M.; Sberveglieri, G. Nanostructured WO3 deposited by modified thermal evaporation for gas-sensing applications. Thin Solid Film 2005, 490, 81–85. [Google Scholar] [CrossRef]
  241. Boulova, M.; Gaskov, A.; Lucazeau, G. Tungsten oxide reactivity versus CH4, CO and NO2 molecules studied by Raman spectroscopy. Sens. Actuators B Chem. 2001, 81, 99–106. [Google Scholar] [CrossRef]
  242. Tamaki, J.; Miyaji, A.; Makinodan, J.; Ogura, S.; Konishi, S. Effect of micro-gap electrode on detection of dilute NO2 using WO3 thin film microsensors. Sens. Actuators B Chem. 2005, 108, 202–206. [Google Scholar] [CrossRef]
  243. Vallejos, S.; Khatko, V.; Calderer, J.; Gracia, I.; Cané, C.; Llobet, E.; Correig, X. Micro-machined WO3-based sensors selective to oxidizing gases. Sens. Actuators B Chem. 2008, 132, 209–215. [Google Scholar] [CrossRef]
  244. Siciliano, T.; Tepore, A.; Micocci, G.; Serra, A.; Manno, D.; Filippo, E. WO3 gas sensors prepared by thermal oxidization of tungsten. Sens. Actuators B Chem. 2008, 133, 321–326. [Google Scholar] [CrossRef]
  245. Yang, J.-C.; Dutta, P.K.J.S. Solution-based synthesis of efficient WO3 sensing electrodes for high temperature potentiometric NOx sensors. Sens. Actuators B Chem. 2009, 136, 523–529. [Google Scholar] [CrossRef]
  246. Ferroni, M.; Boscarino, D.; Comini, E.; Gnani, D.; Guidi, V.; Martinelli, G.; Nelli, P.; Rigato, V.; Sberveglieri, G. Nanosized thin films of tungsten-titanium mixed oxides as gas sensors. Sens. Actuators B Chem. 1999, 58, 289–294. [Google Scholar] [CrossRef]
  247. Guidi, V.; Boscarino, D.; Comini, E.; Faglia, G.; Ferroni, M.; Malagu, C.; Martinelli, G.; Rigato, V.; Sberveglieri, G. Preparation and characterisation of titanium–tungsten sensors. Sens. Actuators B Chem. 2000, 65, 264–266. [Google Scholar] [CrossRef]
  248. Comini, E.; Ferroni, M.; Guidi, V.; Faglia, G.; Martinelli, G.; Sberveglieri, G.J.S. Nanostructured mixed oxides compounds for gas sensing applications. Sens. Actuators B Chem. 2002, 84, 26–32. [Google Scholar] [CrossRef]
  249. Xia, H.; Wang, Y.; Kong, F.; Wang, S.; Zhu, B.; Guo, X.; Zhang, J.; Wang, Y.; Wu, S. Au-doped WO3-based sensor for NO2 detection at low operating temperature. Sens. Actuators B Chem. 2008, 134, 133–139. [Google Scholar] [CrossRef]
  250. Marquis, B.T.; Vetelino, J.F. A semiconducting metal oxide sensor array for the detection of NOx and NH3. Sens. Actuators B Chem. 2001, 77, 100–110. [Google Scholar] [CrossRef]
  251. Penza, M.; Martucci, C.; Cassano, G. NOx gas sensing characteristics of WO3 thin films activated by noble metals (Pd, Pt, Au) layers. Sens. Actuators B Chem. 1998, 50, 52–59. [Google Scholar] [CrossRef]
  252. Kawasaki, H.; Ueda, T.; Suda, Y.; Ohshima, T. Properties of metal doped tungsten oxide thin films for NOx gas sensors grown by PLD method combined with sputtering process. Sens. Actuators B Chem. 2004, 100, 266–269. [Google Scholar] [CrossRef]
  253. Filippini, D.; Fraigi, L.; Aragón, R.; Weimar, U. Thick film Au-gate field-effect devices sensitive to NO2. Sens. Actuators B Chem. 2002, 81, 296–300. [Google Scholar] [CrossRef]
  254. Yang, J.-C.; Dutta, P.K. Promoting selectivity and sensitivity for a high temperature YSZ-based electrochemical total NOx sensor by using a Pt-loaded zeolite Y filter. Sens. Actuators B Chem. 2007, 125, 30–39. [Google Scholar] [CrossRef]
  255. Siciliano, T.; Di Giulio, M.; Tepore, M.; Filippo, E.; Micocci, G.; Tepore, A. Tellurium sputtered thin films as NO2 gas sensors. Sens. Actuators B Chem. 2008, 135, 250–254. [Google Scholar] [CrossRef]
  256. Siciliano, T.; Di Giulio, M.; Tepore, M.; Filippo, E.; Micocci, G.; Tepore, A. Room temperature NO2 sensing properties of reactively sputtered TeO2 thin films. Sens. Actuators B Chem. 2009, 137, 644–648. [Google Scholar] [CrossRef]
  257. Kim, T.-S.; Lee, Y.; Xu, W.; Kim, Y.H.; Kim, M.; Min, S.-Y.; Kim, T.H.; Jang, H.W.; Lee, T.-W. Direct-printed nanoscale metal-oxide-wire electronics. Nano Energy 2019, 58, 437–446. [Google Scholar] [CrossRef]
  258. Matsuguchi, M.; Tamai, K.; Sakai, Y. SO2 gas sensors using polymers with different amino groups. Sens. Actuators B Chem. 2001, 77, 363–367. [Google Scholar] [CrossRef]
  259. Dultsev, F.; Sveshnikova, L. The use of the substituted imidazoline radical as a receptor for sulphur dioxide gas sensor. Sens. Actuators B Chem. 2007, 120, 434–438. [Google Scholar] [CrossRef]
  260. Benmakroha, F.; Boudjerda, T.; Boufenar, R.; Allag, H.; Djerboua, F.; McCallum, J.J. Monitoring of sulfur dioxide using a piezoelectric crystal based controller. Analyst 1993, 118, 401–406. [Google Scholar] [CrossRef]
  261. Berger, F.; Fromm, M.; Chambaudet, A.; Planade, R. Tin dioxide-based gas sensors for SO2 detection: A chemical interpretation of the increase in sensitivity obtained after a primary detection. Sens. Actuators B Chem. 1997, 45, 175–181. [Google Scholar] [CrossRef]
  262. Torvela, H.; Huusko, J.; Lantto, V. Reduction of the interference caused by NO and SO2 in the CO response of Pd-catalysed SnO2 combustion gas sensors. Sens. Actuators B Chem. 1991, 4, 479–484. [Google Scholar] [CrossRef]
  263. Stankova, M.; Vilanova, X.; Calderer, J.; Llobet, E.; Ivanov, P.; Gràcia, I.; Cané, C.; Correig, X. Detection of SO2 and H2S in CO2 stream by means of WO3-based micro-hotplate sensors. Sens. Actuators B Chem. 2004, 102, 219–225. [Google Scholar] [CrossRef]
  264. Penza, M.; Cassano, G.; Tortorella, F. Gas recognition by activated WO3 thin-film sensors array. Sens. Actuators B Chem. 2001, 81, 115–121. [Google Scholar] [CrossRef]
  265. Liang, X.; Zhong, T.; Quan, B.; Wang, B.; Guan, H. Solid-state potentiometric SO2 sensor combining NASICON with V2O5-doped TiO2 electrode. Sens. Actuators B Chem. 2008, 134, 25–30. [Google Scholar] [CrossRef]
  266. Guo, Z.; Chen, G.; Zeng, G.; Liu, L.; Zhang, C. Metal oxides and metal salt nanostructures for hydrogen sulfide sensing: Mechanism and sensing performance. RSC Adv. 2015, 5, 54793–54805. [Google Scholar] [CrossRef]
  267. Mirzaei, A.; Kim, S.S.; Kim, H.W. Resistance-based H2S gas sensors using metal oxide nanostructures: A review of recent advances. J. Hazard. Mater. 2018, 357, 314–331. [Google Scholar] [CrossRef]
  268. Joshi, N.; Hayasaka, T.; Liu, Y.; Liu, H.; Oliveira, O.N.; Lin, L. A review on chemiresistive room temperature gas sensors based on metal oxide nanostructures, graphene and 2D transition metal dichalcogenides. Microchim. Acta 2018, 185, 213. [Google Scholar] [CrossRef] [PubMed]
  269. Llobet, E.; Brunet, J.; Pauly, A.; Ndiaye, A.; Varenne, C. Nanomaterials for the selective detection of hydrogen sulfide in air. Sensors 2017, 17, 391. [Google Scholar] [CrossRef] [PubMed]
  270. He, M.; Xie, L.; Zhao, X.; Hu, X.; Li, S.; Zhu, Z.-G. Highly sensitive and selective H2S gas sensors based on flower-like WO3/CuO composites operating at low/room temperature. J. Alloys Compd. 2019, 788, 36–43. [Google Scholar] [CrossRef]
  271. Hangarter, C.M.; Chartuprayoon, N.; Hernández, S.C.; Choa, Y.; Myung, N.V. Hybridized conducting polymer chemiresistive nano-sensors. Nano Today 2013, 8, 39–55. [Google Scholar] [CrossRef]
  272. Duc, C.; Boukhenane, M.-L.; Wojkiewicz, J.-L.; Redon, N. Hydrogen Sulfide Detection by Sensors Based on Conductive Polymers: A Review. Front. Mater. 2020, 7, 215. [Google Scholar] [CrossRef]
  273. Rout, C.S.; Hegde, M.; Rao, C.R. H2S sensors based on tungsten oxide nanostructures. Sens. Actuators B Chem. 2008, 128, 488–493. [Google Scholar] [CrossRef]
  274. Tao, W.-H.; Tsai, C.-H. H2S sensing properties of noble metal doped WO3 thin film sensor fabricated by micromachining. Sens. Actuators B Chem. 2002, 81, 237–247. [Google Scholar] [CrossRef]
  275. Solis, J.; Saukko, S.; Kish, L.; Granqvist, C.; Lantto, V. Nanocrystalline tungsten oxide thick-films with high sensitivity to H2S at room temperature. Sens. Actuators B Chem. 2001, 77, 316–321. [Google Scholar] [CrossRef]
  276. Ruokamo, I.; Kärkkäinen, T.; Huusko, J.; Ruokanen, T.; Blomberg, M.; Torvela, H.; Lantto, V. H2S response of WO3 thin-film sensors manufactured by silicon processing technology. Sens. Actuators B Chem. 1994, 19, 486–488. [Google Scholar] [CrossRef]
  277. Frühberger, B.; Grunze, M.; Dwyer, D. Surface chemistry of H2S-sensitive tungsten oxide films. Sens. Actuators B Chem. 1996, 31, 167–174. [Google Scholar] [CrossRef]
  278. Vuong, D.D.; Sakai, G.; Shimanoe, K.; Yamazoe, N. Hydrogen sulfide gas sensing properties of thin films derived from SnO2 sols different in grain size. Sens. Actuators B Chem. 2005, 105, 437–442. [Google Scholar] [CrossRef]
  279. Shuping, G.; Jing, X.; Jianqiao, L.; Dongxiang, Z.J.S.; Chemical, A.B. Highly sensitive SnO2 thin film with low operating temperature prepared by sol–gel technique. Sens. Actuators B Chem. 2008, 134, 57–61. [Google Scholar] [CrossRef]
  280. Ghimbeu, C.M.; Lumbreras, M.; Siadat, M.; van Landschoot, R.C.; Schoonman, J. Electrostatic sprayed SnO2 and Cu-doped SnO2 films for H2S detection. Sens. Actuators B Chem. 2008, 133, 694–698. [Google Scholar] [CrossRef]
  281. Liang, X.; He, Y.; Liu, F.; Wang, B.; Zhong, T.; Quan, B.; Lu, G. Solid-state potentiometric H2S sensor combining NASICON with Pr6O11-doped SnO2 electrode. Sens. Actuators B Chem. 2007, 125, 544–549. [Google Scholar] [CrossRef]
  282. Wang, C.; Chu, X.; Wu, M. Detection of H2S down to ppb levels at room temperature using sensors based on ZnO nanorods. Sens. Actuators B Chem. 2006, 113, 320–323. [Google Scholar] [CrossRef]
  283. Liu, Z.; Fan, T.; Zhang, D.; Gong, X.; Xu, J. Hierarchically porous ZnO with high sensitivity and selectivity to H2S derived from biotemplates. Sens. Actuators B Chem. 2009, 136, 499–509. [Google Scholar] [CrossRef]
  284. Wallace, K.J.; Cordero, S.R.; Tan, C.P.; Lynch, V.M.; Anslyn, E.V. A colorimetric response to hydrogen sulfide. Sens. Actuators B Chem. 2007, 120, 362–367. [Google Scholar] [CrossRef]
  285. Jain, G.; Patil, L. CuO-doped BSST thick film resistors for ppb level H2S gas sensing at room temperature. Sens. Actuators B Chem. 2007, 123, 246–253. [Google Scholar] [CrossRef]
  286. Wang, Y.; Wang, S.; Zhao, Y.; Zhu, B.; Kong, F.; Wang, D.; Wu, S.; Huang, W.; Zhang, S. H2S sensing characteristics of Pt-doped α-Fe2O3 thick film sensors. Sens. Actuators B Chem. 2007, 125, 79–84. [Google Scholar] [CrossRef]
  287. Kapse, V.; Ghosh, S.; Raghuwanshi, F.; Kapse, S. Enhanced H2S sensing characteristics of La-doped In2O3: Effect of Pd sensitization. Sens. Actuators B Chem. 2009, 137, 681–686. [Google Scholar] [CrossRef]
  288. Kaur, M.; Jain, N.; Sharma, K.; Bhattacharya, S.; Roy, M.; Tyagi, A.; Gupta, S.; Yakhmi, J. Room-temperature H2S gas sensing at ppb level by single crystal In2O3 whiskers. Sens. Actuators B Chem. 2008, 133, 456–461. [Google Scholar] [CrossRef]
  289. Wang, Y.; Wang, Y.; Cao, J.; Kong, F.; Xia, H.; Zhang, J.; Zhu, B.; Wang, S.; Wu, S. Low-temperature H2S sensors based on Ag-doped α-Fe2O3 nanoparticles. Sens. Actuators B Chem. 2008, 131, 183–189. [Google Scholar] [CrossRef]
  290. Fam, D.W.H.; Tok, A.I.Y.; Palaniappan, A.; Nopphawan, P.; Lohani, A.; Mhaisalkar, S.J.S.; Chemical, A.B. Selective sensing of hydrogen sulphide using silver nanoparticle decorated carbon nanotubes. Sens. Actuators B Chem. 2009, 138, 189–192. [Google Scholar] [CrossRef]
  291. Jain, G.; Patil, L.; Wagh, M.; Patil, D.; Patil, S.; Amalnerkar, D. Surface modified BaTiO3 thick film resistors as H2S gas sensors. Sens. Actuators B Chem. 2006, 117, 159–165. [Google Scholar] [CrossRef]
  292. Sarma, T.; Tao, S. An active core fiber optic sensor for detecting trace H2S at high temperature using a cadmium oxide doped porous silica optical fiber as a transducer. Sens. Actuators B Chem. 2007, 127, 471–479. [Google Scholar] [CrossRef]
  293. Morazzoni, F.; Scotti, R.; Origoni, L.; D’Arienzo, M.; Jimenez, I.; Cornet, A.; Morante, J. Mechanism of NH3 interaction with transition metal-added nanosized WO3 for gas sensing: In situ electron paramagnetic resonance study. Catal. Today 2007, 126, 169–176. [Google Scholar] [CrossRef]
  294. Tong, M.; Dai, G.; Gao, D. WO3 thin film sensor prepared by sol–gel technique and its low-temperature sensing properties to trimethylamine. Mater. Chem. Phys. 2001, 69, 176–179. [Google Scholar] [CrossRef]
  295. Bendahan, M.; Lauque, P.; Seguin, J.-L.; Aguir, K.; Knauth, P. Development of an ammonia gas sensor. Sens. Actuators B Chem. 2003, 95, 170–176. [Google Scholar] [CrossRef]
  296. Wagh, M.; Jain, G.; Patil, D.; Patil, S.; Patil, L. Modified zinc oxide thick film resistors as NH3 gas sensor. Sens. Actuators B Chem. 2006, 115, 128–133. [Google Scholar] [CrossRef]
  297. Wang, Y.-D.; Wu, X.-H.; Su, Q.; Li, Y.-F.; Zhou, Z.-L. Ammonia-sensing characteristics of Pt and SiO2 doped SnO2 materials. Solid-State Electron. 2001, 45, 347–350. [Google Scholar] [CrossRef]
  298. Suryawanshi, D.; Patil, D.; Patil, L. Fe2O3-activated Cr2O3 thick films as temperature dependent gas sensors. Sens. Actuators B Chem. 2008, 134, 579–584. [Google Scholar] [CrossRef]
  299. Patil, D.; Patil, L.; Patil, P. Cr2O3-activated ZnO thick film resistors for ammonia gas sensing operable at room temperature. Sens. Actuators B Chem. 2007, 126, 368–374. [Google Scholar] [CrossRef]
  300. Li, G.-J.; Kawi, S. Synthesis, characterization and sensing application of novel semiconductor oxides. Talanta 1998, 45, 759–766. [Google Scholar] [CrossRef] [PubMed]
  301. Sunny, A.I.; Zhao, A.; Li, L.; Sakiliba, S.K. Low-cost IoT-based sensor system: A case study on harsh environmental monitoring. Sensors 2020, 21, 214. [Google Scholar] [CrossRef]
  302. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Kim, T.W. Recent advances in energy-saving chemiresistive gas sensors: A review. Nano Energy 2021, 79, 105369. [Google Scholar] [CrossRef] [PubMed]
  303. Sekimoto, S.; Nakagawa, H.; Okazaki, S.; Fukuda, K.; Asakura, S.; Shigemori, T.; Takahashi, S. A fiber-optic evanescent-wave hydrogen gas sensor using palladium-supported tungsten oxide. Sens. Actuators B Chem. 2000, 66, 142–145. [Google Scholar] [CrossRef]
  304. Hellegouarch, F.; Arefi-Khonsari, F.; Planade, R.; Amouroux, J. PECVD prepared SnO2 thin films for ethanol sensors. Sens. Actuators B Chem. 2001, 73, 27–34. [Google Scholar] [CrossRef]
  305. Pourfayaz, F.; Khodadadi, A.; Mortazavi, Y.; Mohajerzadeh, S. CeO2 doped SnO2 sensor selective to ethanol in presence of CO, LPG and CH4. Sens. Actuators B Chem. 2005, 108, 172–176. [Google Scholar] [CrossRef]
  306. Jain, K.; Pant, R.; Lakshmikumar, S. Effect of Ni doping on thick film SnO2 gas sensor. Sens. Actuators B Chem. 2006, 113, 823–829. [Google Scholar] [CrossRef]
  307. Wagh, M.; Jain, G.; Patil, D.; Patil, S.; Patil, L. Surface customization of SnO2 thick films using RuO2 as a surfactant for the LPG response. Sens. Actuators B Chem. 2007, 122, 357–364. [Google Scholar] [CrossRef]
  308. Ionescu, R.; Hoel, A.; Granqvist, C.; Llobet, E.; Heszler, P. Ethanol and H2S gas detection in air and in reducing and oxidising ambience: Application of pattern recognition to analyse the output from temperature-modulated nanoparticulate WO3 gas sensors. Sens. Actuators B Chem. 2005, 104, 124–131. [Google Scholar] [CrossRef]
  309. Khadayate, R.; Sali, J.; Patil, P. Acetone vapor sensing properties of screen printed WO3 thick films. Talanta 2007, 72, 1077–1081. [Google Scholar] [CrossRef] [PubMed]
  310. Cao, X.; Wu, W.; Chen, N.; Peng, Y.; Liu, Y. An ether sensor utilizing cataluminescence on nanosized ZnWO4. Sens. Actuators B Chem. 2009, 137, 83–87. [Google Scholar] [CrossRef]
  311. de Lacy Costello, B.; Ewen, R.; Guernion, N.; Ratcliffe, N. Highly sensitive mixed oxide sensors for the detection of ethanol. Sens. Actuators B Chem. 2002, 87, 207–210. [Google Scholar] [CrossRef]
  312. Xu, J.; Han, J.; Zhang, Y.; Sun, Y.A.; Xie, B. Studies on alcohol sensing mechanism of ZnO based gas sensors. Sens. Actuators B Chem. 2008, 132, 334–339. [Google Scholar] [CrossRef]
  313. Zhao, S.; Sin, J.K.; Xu, B.; Zhao, M.; Peng, Z.; Cai, H. A high performance ethanol sensor based on field-effect transistor using a LaFeO3 nano-crystalline thin-film as a gate electrode. Sens. Actuators B Chem. 2000, 64, 83–87. [Google Scholar] [CrossRef]
  314. Liu, X.; Cheng, B.; Hu, J.; Qin, H.; Jiang, M. Preparation, structure, resistance and methane-gas sensing properties of nominal La1xMgxFeO3. Sens. Actuators B Chem. 2008, 133, 340–344. [Google Scholar] [CrossRef]
  315. Ajami, S.; Mortazavi, Y.; Khodadadi, A.; Pourfayaz, F.; Mohajerzadeh, S. Highly selective sensor to CH4 in presence of CO and ethanol using LaCoO3 perovskite filter with Pt/SnO2. Sens. Actuators B Chem. 2006, 117, 420–425. [Google Scholar] [CrossRef]
  316. Xuan, Y.; Hu, J.; Xu, K.; Hou, X.; Lv, Y. Development of sensitive carbon disulfide sensor by using its cataluminescence on nanosized-CeO2. Sens. Actuators B Chem. 2009, 136, 218–223. [Google Scholar] [CrossRef]
  317. Kida, T.; Minami, T.; Kishi, S.; Yuasa, M.; Shimanoe, K.; Yamazoe, N. Planar-type BiCuVOx solid electrolyte sensor for the detection of volatile organic compounds. Sens. Actuators B Chem. 2009, 137, 147–153. [Google Scholar] [CrossRef]
  318. Jinkawa, T.; Sakai, G.; Tamaki, J.; Miura, N.; Yamazoe, N. Relationship between ethanol gas sensitivity and surface catalytic property of tin oxide sensors modified with acidic or basic oxides. J. Mol. Catal. A: Chem. 2000, 155, 193–200. [Google Scholar] [CrossRef]
  319. Tianshu, Z.; Hing, P.; Li, Y.; Jiancheng, Z. Selective detection of ethanol vapor and hydrogen using Cd-doped SnO2-based sensors. Sens. Actuators B Chem. 1999, 60, 208–215. [Google Scholar] [CrossRef]
  320. Lv, P.; Tang, Z.A.; Yu, J.; Zhang, F.T.; Wei, G.F.; Huang, Z.X.; Hu, Y. Study on a micro-gas sensor with SnO2–NiO sensitive film for indoor formaldehyde detection. Sens. Actuators B Chem. 2008, 132, 74–80. [Google Scholar] [CrossRef]
  321. Jun, L.; Chen, Q.; Fu, W.; Yang, Y.; Zhu, W.; Zhang, J. Electrospun Yb-doped In2O3 nanofiber field-effect transistors for highly sensitive ethanol sensors. ACS Appl. Mater. Interfaces 2020, 12, 38425–38434. [Google Scholar] [CrossRef]
  322. Chen, A.; Huang, X.; Tong, Z.; Bai, S.; Luo, R.; Liu, C.C. Preparation, characterization and gas-sensing properties of SnO2–In2O3 nanocomposite oxides. Sens. Actuators B Chem. 2006, 115, 316–321. [Google Scholar] [CrossRef]
  323. Her, Y.-C.; Chiang, C.-K.; Jean, S.-T.; Huang, S.-L. Self-catalytic growth of hierarchical In2O3 nanostructures on SnO2 nanowires and their CO sensing properties. CrystEngComm 2012, 14, 1296–1300. [Google Scholar] [CrossRef]
  324. Kuang, Q.; Lao, C.-S.; Li, Z.; Liu, Y.-Z.; Xie, Z.-X.; Zheng, L.-S.; Wang, Z.L. Enhancing the photon-and gas-sensing properties of a single SnO2 nanowire based nanodevice by nanoparticle surface functionalization. J. Phys. Chem. C 2008, 112, 11539–11544. [Google Scholar] [CrossRef]
  325. Choi, U.-S.; Sakai, G.; Shimanoe, K.; Yamazoe, N. Sensing properties of Au-loaded SnO2–Co3O4 composites to CO and H2. Sens. Actuators B Chem. 2005, 107, 397–401. [Google Scholar] [CrossRef]
  326. Moon, W.J.; Yu, J.H.; Choi, G.M. The CO and H2 gas selectivity of CuO-doped SnO2–ZnO composite gas sensor. Sens. Actuators B Chem. 2002, 87, 464–470. [Google Scholar] [CrossRef]
  327. Huang, H.; Gong, H.; Chow, C.L.; Guo, J.; White, T.J.; Tse, M.S.; Tan, O.K. Low-temperature growth of SnO2 nanorod arrays and tunable n–p–n sensing response of a ZnO/SnO2 heterojunction for exclusive hydrogen sensors. Adv. Funct. Mater. 2011, 21, 2680–2686. [Google Scholar] [CrossRef]
  328. Ansari, Z.; Ansari, S.; Ko, T.; Oh, J.-H. Effect of MoO3 doping and grain size on SnO2-enhancement of sensitivity and selectivity for CO and H2 gas sensing. Sens. Actuators B Chem. 2002, 87, 105–114. [Google Scholar] [CrossRef]
  329. Kosc, I.; Hotovy, I.; Rehacek, V.; Griesseler, R.; Predanocy, M.; Wilke, M.; Spiess, L. Sputtered TiO2 thin films with NiO additives for hydrogen detection. Appl. Surf. Sci. 2013, 269, 110–115. [Google Scholar] [CrossRef]
  330. Tien, L.; Norton, D.; Gila, B.; Pearton, S.; Wang, H.-T.; Kang, B.; Ren, F. Detection of hydrogen with SnO2-coated ZnO nanorods. Appl. Surf. Sci. 2007, 253, 4748–4752. [Google Scholar] [CrossRef]
  331. Choi, S.-W.; Park, J.Y.; Kim, S.S. Synthesis of SnO2–ZnO core–shell nanofibers via a novel two-step process and their gas sensing properties. Nanotechnology 2009, 20, 465603. [Google Scholar] [CrossRef]
  332. Liangyuan, C.; Shouli, B.; Guojun, Z.; Dianqing, L.; Aifan, C.; Liu, C.C. Synthesis of ZnO–SnO2 nanocomposites by microemulsion and sensing properties for NO2. Sens. Actuators B Chem. 2008, 134, 360–366. [Google Scholar] [CrossRef]
  333. Bai, S.; Li, D.; Han, D.; Luo, R.; Chen, A.; Chung, C.L. Preparation, characterization of WO3–SnO2 nanocomposites and their sensing properties for NO2. Sens. Actuators B Chem. 2010, 150, 749–755. [Google Scholar] [CrossRef]
  334. Somacescu, S.; Dinescu, A.; Stanoiu, A.; Simion, C.E.; Moreno, J.M.C. Hydrothermal synthesis of ZnO–Eu2O3 binary oxide with straight strips morphology and sensitivity to NO2 gas. Mater. Lett. 2012, 89, 219–222. [Google Scholar] [CrossRef]
  335. Cui, G.; Zhang, M.; Zou, G. Resonant tunneling modulation in quasi-2D Cu2O/SnO2 p-n horizontal-multi-layer heterostructure for room temperature H2S sensor application. Sci. Rep. 2013, 3, 1250. [Google Scholar] [CrossRef]
  336. Verma, M.K.; Gupta, V. SnO2–CuO nanocomposite thin film sensor for fast detection of H2S gas. J. Exp. Nanosci. 2013, 8, 326–331. [Google Scholar] [CrossRef]
  337. Fang, G.; Liu, Z.; Liu, C.; Yao, K.-L. Room temperature H2S sensing properties and mechanism of CeO2–SnO2 sol–gel thin films. Sens. Actuators B Chem. 2000, 66, 46–48. [Google Scholar] [CrossRef]
  338. Wagh, M.; Patil, L.; Seth, T.; Amalnerkar, D. Surface cupricated SnO2–ZnO thick films as a H2S gas sensor. Mater. Chem. Phys. 2004, 84, 228–233. [Google Scholar] [CrossRef]
  339. Wang, L.; Kang, Y.; Wang, Y.; Zhu, B.; Zhang, S.; Huang, W.; Wang, S. CuO nanoparticle decorated ZnO nanorod sensor for low-temperature H2S detection. Mater. Sci. Eng. C 2012, 32, 2079–2085. [Google Scholar] [CrossRef] [PubMed]
  340. Tang, H.; Yan, M.; Zhang, H.; Li, S.; Ma, X.; Wang, M.; Yang, D. A selective NH3 gas sensor based on Fe2O3–ZnO nanocomposites at room temperature. Sens. Actuators B Chem. 2006, 114, 910–915. [Google Scholar] [CrossRef]
  341. Zeng, Y.; Bing, Y.-F.; Liu, C.; Zheng, W.-T.; Zou, G.-T. Self-assembly of hierarchical ZnSnO3-SnO2 nanoflakes and their gas sensing properties. Trans. Nonferrous Met. Soc. China 2012, 22, 2451–2458. [Google Scholar] [CrossRef]
  342. Kim, K.-W.; Cho, P.-S.; Kim, S.-J.; Lee, J.-H.; Kang, C.-Y.; Kim, J.-S.; Yoon, S.-J. The selective detection of C2H5OH using SnO2–ZnO thin film gas sensors prepared by combinatorial solution deposition. Sens. Actuators B Chem. 2007, 123, 318–324. [Google Scholar] [CrossRef]
  343. Rumyantseva, M.; Kovalenko, V.; Gaskov, A.; Makshina, E.; Yuschenko, V.; Ivanova, I.; Ponzoni, A.; Faglia, G.; Comini, E. Nanocomposites SnO2/Fe2O3: Sensor and catalytic properties. Sens. Actuators B Chem. 2006, 118, 208–214. [Google Scholar] [CrossRef]
  344. 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]
  345. Yu, X.; Zhang, G.; Cao, H.; An, X.; Wang, Y.; Shu, Z.; An, X.; Hua, F. ZnO@ ZnS hollow dumbbells–graphene composites as high-performance photocatalysts and alcohol sensors. New J. Chem. 2012, 36, 2593–2598. [Google Scholar] [CrossRef]
  346. Hu, Y.; Zhou, X.; Han, Q.; Cao, Q.; Huang, Y. Sensing properties of CuO–ZnO heterojunction gas sensors. Mater. Sci. Eng. B 2003, 99, 41–43. [Google Scholar] [CrossRef]
  347. Zeng, W.; Liu, T.; Wang, Z. Sensitivity improvement of TiO2-doped SnO2 to volatile organic compounds. Phys. E Low Dimens. Syst. Nanostructures 2010, 43, 633–638. [Google Scholar] [CrossRef]
  348. Liu, Y.; Zhu, G.; Chen, J.; Xu, H.; Shen, X.; Yuan, A. Co3O4/ZnO nanocomposites for gas-sensing applications. Appl. Surf. Sci. 2013, 265, 379–384. [Google Scholar] [CrossRef]
  349. Pimtong-Ngam, Y.; Jiemsirilers, S.; Supothina, S. Preparation of tungsten oxide–tin oxide nanocomposites and their ethylene sensing characteristics. Sens. Actuators A Phys. 2007, 139, 7–11. [Google Scholar] [CrossRef]
  350. Galatsis, K.; Li, Y.; Wlodarski, W.; Comini, E.; Sberveglieri, G.; Cantalini, C.; Santucci, S.; Passacantando, M.J.S.; Chemical, A.B. Comparison of single and binary oxide MoO3, TiO2 and WO3 sol–gel gas sensors. Sens. Actuators B Chem. 2002, 83, 276–280. [Google Scholar] [CrossRef]
  351. Hidalgo, P.; Castro, R.H.; Coelho, A.C.; Gouvêa, D. Surface segregation and consequent SO2 sensor response in SnO2− NiO. Chem. Mater. 2005, 17, 4149–4153. [Google Scholar] [CrossRef]
  352. Patil, D.; Patil, L. Cr2O3-modified ZnO thick film resistors as LPG sensors. Talanta 2009, 77, 1409–1414. [Google Scholar] [CrossRef] [PubMed]
  353. Zhang, W.-H.; Zhang, W.-D. Fabrication of SnO2–ZnO nanocomposite sensor for selective sensing of trimethylamine and the freshness of fishes. Sens. Actuators B Chem. 2008, 134, 403–408. [Google Scholar] [CrossRef]
  354. Heiland, G. Zum Einfluß von adsorbiertem Sauerstoff auf die elektrische Leitfähigkeit von Zinkoxydkristallen. Z. Für Phys. 1954, 138, 459–464. [Google Scholar] [CrossRef]
  355. Bielański, A.; Dereń, J.; Haber, J. Electric conductivity and catalytic activity of semiconducting oxide catalysts. Nature 1957, 179, 668–669. [Google Scholar] [CrossRef]
  356. Shimizu, Y.; Egashira, M. Basic aspects and challenges of semiconductor gas sensors. MRS Bull. 1999, 24, 18–24. [Google Scholar] [CrossRef]
  357. Strässler, S.; Reis, A. Simple models for n-type metal oxide gas sensors. Sens. Actuators 1983, 4, 465–472. [Google Scholar] [CrossRef]
  358. Wang, B.; Zhu, L.; Yang, Y.; Xu, N.; Yang, G. Fabrication of a SnO2 nanowire gas sensor and sensor performance for hydrogen. J. Phys. Chem. C 2008, 112, 6643–6647. [Google Scholar] [CrossRef]
  359. Sysoev, V.; Schneider, T.; Goschnick, J.; Kiselev, I.; Habicht, W.; Hahn, H.; Strelcov, E.; Kolmakov, A.J.S.; Chemical, A.B. Percolating SnO2 nanowire network as a stable gas sensor: Direct comparison of long-term performance versus SnO2 nanoparticle films. Sens. Actuators B Chem. 2009, 139, 699–703. [Google Scholar] [CrossRef]
  360. Gyger, F.; Hübner, M.; Feldmann, C.; Barsan, N.; Weimar, U. Nanoscale SnO2 hollow spheres and their application as a gas-sensing material. Chem. Mater. 2010, 22, 4821–4827. [Google Scholar] [CrossRef]
  361. Epifani, M.; Prades, J.D.; Comini, E.; Pellicer, E.; Avella, M.; Siciliano, P.; Faglia, G.; Cirera, A.; Scotti, R.; Morazzoni, F. The role of surface oxygen vacancies in the NO2 sensing properties of SnO2 nanocrystals. J. Phys. Chem. C 2008, 112, 19540–19546. [Google Scholar] [CrossRef]
  362. Li, L.-L.; Zhang, W.-M.; Yuan, Q.; Li, Z.-X.; Fang, C.-J.; Sun, L.-D.; Wan, L.-J.; Yan, C.-H. Room temperature ionic liquids assisted green synthesis of nanocrystalline porous SnO2 and their gas sensor behaviors. Cryst. Growth Des. 2008, 8, 4165–4172. [Google Scholar] [CrossRef]
  363. Wang, Q.; Zhang, L.-S.; Wu, J.-F.; Wang, W.D.; Song, W.-G.; Wang, W. A parallel solid-state NMR and sensor property study on flower-like nanostructured SnO2. J. Phys. Chem. C 2010, 114, 22671–22676. [Google Scholar] [CrossRef]
  364. Chen, D.; Xu, J.; Xie, Z.; Shen, G. Nanowires assembled SnO2 nanopolyhedrons with enhanced gas sensing properties. ACS Appl. Mater. Interfaces 2011, 3, 2112–2117. [Google Scholar] [CrossRef]
  365. Zhang, J.; Wang, S.; Xu, M.; Wang, Y.; Xia, H.; Zhang, S.; Guo, X.; Wu, S. Polypyrrole-coated SnO2 hollow spheres and their application for ammonia sensor. J. Phys. Chem. C 2009, 113, 1662–1665. [Google Scholar] [CrossRef]
  366. Pan, J.; Ganesan, R.; Shen, H.; Mathur, S. Plasma-modified SnO2 nanowires for enhanced gas sensing. J. Phys. Chem. C 2010, 114, 8245–8250. [Google Scholar] [CrossRef]
  367. Xue, X.; Chen, Z.; Ma, C.; Xing, L.; Chen, Y.; Wang, Y.; Wang, T. One-step synthesis and gas-sensing characteristics of uniformly loaded Pt@ SnO2 nanorods. J. Phys. Chem. C 2010, 114, 3968–3972. [Google Scholar] [CrossRef]
  368. Wei, W.; Dai, Y.; Huang, B. Role of Cu doping in SnO2 sensing properties toward H2S. J. Phys. Chem. 2011, 115, 18597–18602. [Google Scholar] [CrossRef]
  369. Li, W.; Shen, C.; Wu, G.; Ma, Y.; Gao, Z.; Xia, X.; Du, G. New model for a Pd-doped SnO2-based CO gas sensor and catalyst studied by online in-situ X-ray photoelectron spectroscopy. J. Phys. Chem. C 2011, 115, 21258–21263. [Google Scholar] [CrossRef]
  370. Umar, A.; Rahman, M.M.; Kim, S.H.; Hahn, Y.-B. Zinc oxide nanonail based chemical sensor for hydrazine detection. Chem. Commun. 2007, 2, 166–168. [Google Scholar] [CrossRef] [PubMed]
  371. Ahsanulhaq, Q.; Kim, J.; Lee, J.; Hahn, Y. Electrical and gas sensing properties of ZnO nanorod arrays directly grown on a four-probe electrode system. Electrochem. Commun. 2010, 12, 475–478. [Google Scholar] [CrossRef]
  372. Li, Y.-J.; Li, K.-M.; Wang, C.-Y.; Kuo, C.-I.; Chen, L.-J. Low-temperature electrodeposited Co-doped ZnO nanorods with enhanced ethanol and CO sensing properties. Sens. Actuators B Chem. 2012, 161, 734–739. [Google Scholar] [CrossRef]
  373. Choi, S.-W.; Kim, S.S. Room temperature CO sensing of selectively grown networked ZnO nanowires by Pd nanodot functionalization. Sens. Actuators B Chem. 2012, 168, 8–13. [Google Scholar] [CrossRef]
  374. Arnold, S.; Prokes, S.; Perkins, F.; Zaghloul, M. Design and performance of a simple, room-temperature Ga2O3 nanowire gas sensor. Appl. Phys. Lett. 2009, 95, 10. [Google Scholar] [CrossRef]
  375. Yan, S.; Wan, L.; Li, Z.; Zhou, Y.; Zou, Z. Synthesis of a mesoporous single crystal Ga 2 O 3 nanoplate with improved photoluminescence and high sensitivity in detecting CO. Chem. Commun. 2010, 46, 6388–6390. [Google Scholar] [CrossRef]
  376. Nguyen, H.; El-Safty, S.A. Meso-and macroporous Co3O4 nanorods for effective VOC gas sensors. J. Phys. Chem. C 2011, 115, 8466–8474. [Google Scholar] [CrossRef]
  377. Zhu, C.-L.; Yu, H.-L.; Zhang, Y.; Wang, T.-S.; Ouyang, Q.-Y.; Qi, L.-H.; Chen, Y.-J.; Xue, X.-Y. Fe2O3/TiO2 tube-like nanostructures: Synthesis, structural transformation and the enhanced sensing properties. ACS Appl. Mater. Interfaces 2012, 4, 665–671. [Google Scholar] [CrossRef]
  378. 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]
  379. Wang, C.Y.; Ali, M.; Kups, T.; Röhlig, C.-C.; Cimalla, V.; Stauden, T.; Ambacher, O. NOx sensing properties of In2O3 nanoparticles prepared by metal organic chemical vapor deposition. Sens. Actuators B Chem. 2008, 130, 589–593. [Google Scholar] [CrossRef]
  380. Neri, G.; Bonavita, A.; Micali, G.; Rizzo, G.; Callone, E.; Carturan, G. Resistive CO gas sensors based on In2O3 and InSnOx nanopowders synthesized via starch-aided sol–gel process for automotive applications. Sens. Actuators B Chem. 2008, 132, 224–233. [Google Scholar] [CrossRef]
  381. Kim, S.-J.; Hwang, I.-S.; Choi, J.-K.; Kang, Y.C.; Lee, J.-H. Enhanced C2H5OH sensing characteristics of nano-porous In2O3 hollow spheres prepared by sucrose-mediated hydrothermal reaction. Sens. Actuators B Chem. 2011, 155, 512–518. [Google Scholar] [CrossRef]
  382. Guo, L.; Shen, X.; Zhu, G.; Chen, K. Preparation and gas-sensing performance of In2O3 porous nanoplatelets. Sens. Actuators B Chem. 2011, 155, 752–758. [Google Scholar] [CrossRef]
  383. Zhang, Y.; Zheng, Z.; Yang, F. Highly sensitive and selective alcohol sensors based on Ag-doped In2O3 coating. Ind. Eng. Chem. Res. 2010, 49, 3539–3543. [Google Scholar] [CrossRef]
  384. Lai, X.; Wang, D.; Han, N.; Du, J.; Li, J.; Xing, C.; Chen, Y.; Li, X. Ordered arrays of bead-chain-like In2O3 nanorods and their enhanced sensing performance for formaldehyde. Chem. Mater. 2010, 22, 3033–3042. [Google Scholar] [CrossRef]
  385. Ye, Z.; Bardelli, S.; Wu, K.; Sarkar, A.; Swindle, A.; Wang, J. Synthesis, crystal growth, electronic properties and optical properties of Y6IV2. 5S14 (IV= Si, Ge). Z. Für Anorg. Und Allg. Chem. 2022, 648, e202100271. [Google Scholar]
  386. Huang, H.; Lee, Y.; Tan, O.; Zhou, W.; Peng, N.; Zhang, Q. High sensitivity SnO2 single-nanorod sensors for the detection of H2 gas at low temperature. Nanotechnology 2009, 20, 115501. [Google Scholar] [CrossRef]
  387. Liu, L.; Li, S.; Wang, L.; Guo, C.; Dong, Q.; Li, W. Enhancement ethanol sensing properties of NiO–SnO2 nanofibers. J. Am. Ceram. Soc. 2011, 94, 771–775. [Google Scholar] [CrossRef]
  388. Wang, Z.; 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]
  389. Yang, Z.; Li, L.-M.; Wan, Q.; Liu, Q.-H.; Wang, T.-H. High-performance ethanol sensing based on an aligned assembly of ZnO nanorods. Sens. Actuators B Chem. 2008, 135, 57–60. [Google Scholar] [CrossRef]
  390. Liu, J.; Li, Y.; Jiang, J.; Huang, X. C@ZnO nanorod array-based hydrazine electrochemical sensor with improved sensitivity and stability. Dalton Trans. 2010, 39, 8693–8697. [Google Scholar] [CrossRef] [PubMed]
  391. Ibrahim, A.A.; Dar, G.N.; Zaidi, S.A.; Umar, A.; Abaker, M.; Bouzid, H.; Baskoutas, S. Growth and properties of Ag-doped ZnO nanoflowers for highly sensitive phenyl hydrazine chemical sensor application. Talanta 2012, 93, 257–263. [Google Scholar] [CrossRef] [PubMed]
  392. Ahn, M.W.; Park, K.S.; Heo, J.H.; Kim, D.W.; Choi, K.J.; Park, J.G. On-chip fabrication of ZnO-nanowire gas sensor with high gas sensitivity. Sens. Actuators B Chem. 2009, 138, 168–173. [Google Scholar] [CrossRef]
  393. Chang, S.J.; Weng, W.Y.; Hsu, C.L.; Hsueh, T.J. High sensitivity of a ZnO nanowire-based ammonia gas sensor with Pt nano-particles. Nano Commun. Netw. 2010, 1, 283–288. [Google Scholar] [CrossRef]
  394. Liu, Z.; Yamazaki, T.; Shen, Y.; Kikuta, T.; Nakatani, N.; Li, Y. O2 and CO sensing of Ga2O3 multiple nanowire gas sensors. Sens. Actuators B Chem. 2008, 129, 666–670. [Google Scholar] [CrossRef]
  395. Song, X.; Gao, L.; Mathur, S. Synthesis, Characterization, and Gas Sensing Properties of Porous Nickel Oxide Nanotubes. J. Phys. Chem. C 2011, 115, 21730–21735. [Google Scholar] [CrossRef]
  396. Wu, Z.; Yu, K.; Zhang, S.; Xie, Y. Hematite Hollow Spheres with a Mesoporous Shell: Controlled Synthesis and Applications in Gas Sensor and Lithium Ion Batteries. J. Phys. Chem. C 2008, 112, 11307–11313. [Google Scholar] [CrossRef]
  397. Yu, R.; Li, Z.; Wang, D.; Lai, X.; Xing, C.; Yang, M.; Xing, X. Fe2TiO5/α-Fe2O3 nanocomposite hollow spheres with enhanced gas-sensing properties. Scr. Mater. 2010, 63, 155–158. [Google Scholar] [CrossRef]
  398. 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]
  399. Gou, X.; Wang, G.; Yang, J.; Park, J.; Wexler, D. Chemical synthesis, characterisation and gas sensing performance of copper oxide nanoribbons. J. Mater. Chem. 2008, 18, 965–969. [Google Scholar] [CrossRef]
  400. Hoa, N.D.; Van Quy, N.; Jung, H.; Kim, D.; Kim, H.; Hong, S.-K. Synthesis of porous CuO nanowires and its application to hydrogen detection. Sens. Actuators B Chem. 2010, 146, 266–272. [Google Scholar] [CrossRef]
  401. Park, J.; Shen, X.; Wang, G. Solvothermal synthesis and gas-sensing performance of Co3O4 hollow nanospheres. Sens. Actuators B Chem. 2009, 136, 494–498. [Google Scholar] [CrossRef]
  402. Ma, J.; Zhang, J.; Wang, S.; Wang, T.; Lian, J.; Duan, X.; Zheng, W. Topochemical Preparation of WO3 Nanoplates through Precursor H2WO4 and Their Gas-Sensing Performances. J. Phys. Chem. C 2011, 115, 18157–18163. [Google Scholar] [CrossRef]
  403. Fortunato, E.; Barquinha, P.; Martins, R. Oxide semiconductor thin-film transistors: A review of recent advances. Adv. Mater. 2012, 24, 2945–2986. [Google Scholar] [CrossRef]
  404. Sheng, J.; Jeong, H.-J.; Han, K.-L.; Hong, T.; Park, J.-S. Review of recent advances in flexible oxide semiconductor thin-film transistors. J. Inf. Disp. 2017, 18, 159–172. [Google Scholar] [CrossRef]
  405. Kim, S.J.; Yoon, S.; Kim, H.J. Review of solution-processed oxide thin-film transistors. Jpn. J. Appl. Phys. 2014, 53, 02BA02. [Google Scholar] [CrossRef]
  406. Du Ahn, B.; Jeon, H.-J.; Sheng, J.; Park, J.; Park, J.-S. A review on the recent developments of solution processes for oxide thin film transistors. Semicond. Sci. Technol. 2015, 30, 064001. [Google Scholar] [CrossRef]
  407. Chen, R.; Lan, L. Solution-processed metal-oxide thin-film transistors: A review of recent developments. Nanotechnology 2019, 30, 312001. [Google Scholar] [CrossRef]
  408. Park, S.P.; Kim, H.J.; Lee, J.H.; Kim, H.J. Glucose-based resistive random access memory for transient electronics. J. Inf. Disp. 2019, 20, 231–237. [Google Scholar] [CrossRef]
  409. Wang, P.; Wu, L.; Lu, Z.; Li, Q.; Yin, W.; Ding, F.; Han, H. Gecko-inspired nanotentacle surface-enhanced Raman spectroscopy substrate for sampling and reliable detection of pesticide residues in fruits and vegetables. Anal. Chem. 2017, 89, 2424–2431. [Google Scholar] [CrossRef] [PubMed]
  410. Song, H.; Zhang, Y.; Wang, S.; Huang, K.; Luo, Y.; Zhang, W.; Xu, W. Label-free polygonal-plate fluorescent-hydrogel biosensor for ultrasensitive microRNA detection. Sens. Actuators B Chem. 2020, 306, 127554. [Google Scholar] [CrossRef]
  411. Hernández-Cancel, G.; Suazo-Dávila, D.; Medina-Guzmán, J.; Rosado-González, M.; Díaz-Vázquez, L.M.; Griebenow, K. Chemically glycosylation improves the stability of an amperometric horseradish peroxidase biosensor. Anal. Chim. Acta 2015, 854, 129–139. [Google Scholar] [CrossRef]
  412. Zheng, H.; Liu, M.; Yan, Z.; Chen, J. Highly selective and stable glucose biosensor based on incorporation of platinum nanoparticles into polyaniline-montmorillonite hybrid composites. Microchem. J. 2020, 152, 104266. [Google Scholar] [CrossRef]
  413. Scognamiglio, V.; Antonacci, A.; Arduini, F.; Moscone, D.; Campos, E.V.; Fraceto, L.F.; Palleschi, G. An eco-designed paper-based algal biosensor for nanoformulated herbicide optical detection. J. Hazard. Mater. 2019, 373, 483–492. [Google Scholar] [CrossRef]
  414. Wang, M.; Yin, H.; Zhou, Y.; Sui, C.; Wang, Y.; Meng, X.; Waterhouse, G.I.; Ai, S. Photoelectrochemical biosensor for microRNA detection based on a MoS2/g-C3N4/black TiO2 heterojunction with Histostar@ AuNPs for signal amplification. Biosens. Bioelectron. 2019, 128, 137–143. [Google Scholar] [CrossRef]
  415. Liu, H.; Duan, C.; Yang, C.; Chen, X.; Shen, W.; Zhu, Z. A novel nitrite biosensor based on the direct electron transfer hemoglobin immobilized in the WO3 nanowires with high length–diameter ratio. Mater. Sci. Eng. C 2015, 53, 43–49. [Google Scholar] [CrossRef]
  416. Dong, Y.; Zheng, J. A nonenzymatic L-cysteine sensor based on SnO2-MWCNTs nanocomposites. J. Mol. Liq. 2014, 196, 280–284. [Google Scholar] [CrossRef]
  417. Zhang, X.; Li, W.; Zhou, Y.; Chai, Y.; Yuan, R. An ultrasensitive electrochemiluminescence biosensor for MicroRNA detection based on luminol-functionalized Au NPs@ ZnO nanomaterials as signal probe and dissolved O2 as coreactant. Biosens. Bioelectron. 2019, 135, 8–13. [Google Scholar] [CrossRef]
  418. Zhou, Q.; Yang, L.; Wang, G.; Yang, Y. Acetylcholinesterase biosensor based on SnO2 nanoparticles–carboxylic graphene–nafion modified electrode for detection of pesticides. Biosens. Bioelectron. 2013, 49, 25–31. [Google Scholar] [CrossRef] [PubMed]
  419. Li, L.; Huang, J.; Wang, T.; Zhang, H.; Liu, Y.; Li, J. An excellent enzyme biosensor based on Sb-doped SnO2 nanowires. Biosens. Bioelectron. 2010, 25, 2436–2441. [Google Scholar] [CrossRef] [PubMed]
  420. Liu, C.; Hong, M.; Lum, M.; Flotow, H.; Ghadessy, F.; Zhang, J. Large-area micro/nanostructures fabrication in quartz by laser interference lithography and dry etching. Appl. Phys. A 2010, 101, 237–241. [Google Scholar] [CrossRef]
  421. Elahi, N.; Kamali, M.; Baghersad, M.H.; Amini, B. A fluorescence Nano-biosensors immobilization on Iron (MNPs) and gold (AuNPs) nanoparticles for detection of Shigella spp. Mater. Sci. Eng. C 2019, 105, 110113. [Google Scholar] [CrossRef] [PubMed]
  422. Zhou, Y.; Yang, L.; Li, S.; Dang, Y. A novel electrochemical sensor for highly sensitive detection of bisphenol A based on the hydrothermal synthesized Na-doped WO3 nanorods. Sens. Actuators B Chem. 2017, 245, 238–246. [Google Scholar] [CrossRef]
  423. Rathinamala, I.; Jeyakumaran, N.; Prithivikumaran, N. Sol-gel assisted spin coated CdS/PS electrode based glucose biosensor. Vacuum 2019, 161, 291–296. [Google Scholar] [CrossRef]
  424. Cobianu, C.; Dumbravescu, N.; Serban, B.-C.; Buiu, O.; Romanitan, C.; Comanescu, F.; Danila, M.; Marinescu, R.; Avramescu, V.; Ionescu, O. Sonochemically synthetized ZnO-Graphene nanohybrids and its characterization. Rev. Adv. Mater. Sci. 2020, 59, 176–187. [Google Scholar] [CrossRef]
  425. Liu, H.; Guo, K.; Duan, C.; Dong, X.; Gao, J. Hollow TiO2 modified reduced graphene oxide microspheres encapsulating hemoglobin for a mediator-free biosensor. Biosens. Bioelectron. 2017, 87, 473–479. [Google Scholar] [CrossRef]
  426. Santos, L.; Silveira, C.M.; Elangovan, E.; Neto, J.P.; Nunes, D.; Pereira, L.; Martins, R.; Viegas, J.; Moura, J.J.; Todorovic, S. Synthesis of WO3 nanoparticles for biosensing applications. Sens. Actuators B Chem. 2016, 223, 186–194. [Google Scholar] [CrossRef]
  427. Dong, S.; Tong, M.; Zhang, D.; Huang, T. The strategy of nitrite and immunoassay human IgG biosensors based on ZnO@ ZIF-8 and ionic liquid composite film. Sens. Actuators B Chem. 2017, 251, 650–657. [Google Scholar] [CrossRef]
  428. Zhang, B.; Wang, H.; Xi, J.; Zhao, F.; Zeng, B. In situ formation of inorganic/organic heterojunction photocatalyst of WO3/Au/polydopamine for immunoassay of human epididymal protein 4. Electrochim. Acta 2020, 331, 135350. [Google Scholar] [CrossRef]
  429. Enesca, A.; Isac, L.; Duta, A. Charge carriers injection in tandem semiconductors for dyes mineralization. Appl. Catal. B Environ. 2015, 162, 352–363. [Google Scholar] [CrossRef]
  430. Mihaly, M.; Lacatusu, I.; Enesca, I.A.; Meghea, A. Hybride nanomaterials based on silica coated C60 clusters obtained by microemulsion technique. Mol. Cryst. Liq. Cryst. 2008, 483, 205–215. [Google Scholar] [CrossRef]
  431. Visa, M.; Andronic, L.; Enesca, A. Behavior of the new composites obtained from fly ash and titanium dioxide in removing of the pollutants from wastewater. Appl. Surf. Sci. 2016, 388, 359–369. [Google Scholar] [CrossRef]
  432. Korotcenkov, G.; Han, S.-D.; Cho, B.; Brinzari, V. Grain size effects in sensor response of nanostructured SnO2-and In2O3-based conductometric thin film gas sensor. Crit. Rev. Solid State Mater. Sci. 2009, 34, 1–17. [Google Scholar] [CrossRef]
  433. Vander Wal, R.; Hunter, G.; Xu, J.; Kulis, M.; Berger, G.; Ticich, T. Metal-oxide nanostructure and gas-sensing performance. Sens. Actuators B Chem. 2009, 138, 113–119. [Google Scholar] [CrossRef]
  434. Umar, A.; Rahman, M.; Al-Hajry, A.; Hahn, Y.-B. Highly-sensitive cholesterol biosensor based on well-crystallized flower-shaped ZnO nanostructures. Talanta 2009, 78, 284–289. [Google Scholar] [CrossRef]
  435. Oh, J.; Yoo, G.; Chang, Y.W.; Kim, H.J.; Jose, J.; Kim, E.; Pyun, J.-C.; Yoo, K.-H. A carbon nanotube metal semiconductor field effect transistor-based biosensor for detection of amyloid-beta in human serum. Biosens. Bioelectron. 2013, 50, 345–350. [Google Scholar] [CrossRef]
  436. Kao, C.H.; Chen, H.; Hou, F.Y.S.; Chang, S.W.; Chang, C.W.; Lai, C.S.; Chen, C.P.; He, Y.Y.; Lin, S.-R.; Hsieh, K.M. Fabrication of multianalyte CeO2 nanograin electrolyte–insulator–semiconductor biosensors by using CF4 plasma treatment. Sens. Bio-Sens. Res. 2015, 5, 71–77. [Google Scholar] [CrossRef]
  437. Zhao, J.; Wu, D.; Zhi, J. A novel tyrosinase biosensor based on biofunctional ZnO nanorod microarrays on the nanocrystalline diamond electrode for detection of phenolic compounds. Bioelectrochemistry 2009, 75, 44–49. [Google Scholar] [CrossRef]
  438. 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]
  439. Chu, X.; Zhu, X.; Dong, Y.; Chen, T.; Ye, M.; Sun, W. An amperometric glucose biosensor based on the immobilization of glucose oxidase on the platinum electrode modified with NiO doped ZnO nanorods. J. Electroanal. Chem. 2012, 676, 20–26. [Google Scholar] [CrossRef]
  440. Li, C.; Liu, Y.; Li, L.; Du, Z.; Xu, S.; Zhang, M.; Yin, X.; Wang, T. A novel amperometric biosensor based on NiO hollow nanospheres for biosensing glucose. Talanta 2008, 77, 455–459. [Google Scholar] [CrossRef] [PubMed]
  441. Fang, B.; Zhang, C.; Wang, G.; Wang, M.; Ji, Y. A glucose oxidase immobilization platform for glucose biosensor using ZnO hollow nanospheres. Sens. Actuators B Chem. 2011, 155, 304–310. [Google Scholar] [CrossRef]
  442. Yang, C.; Xu, C.; Wang, X. ZnO/Cu nanocomposite: A platform for direct electrochemistry of enzymes and biosensing applications. Langmuir 2012, 28, 4580–4585. [Google Scholar] [CrossRef] [PubMed]
  443. Chen, H.; Rim, Y.S.; Wang, I.C.; Li, C.; Zhu, B.; Sun, M.; Goorsky, M.S.; He, X.; Yang, Y. Quasi-two-dimensional metal oxide semiconductors based ultrasensitive potentiometric biosensors. ACS Nano 2017, 11, 4710–4718. [Google Scholar] [CrossRef]
  444. Rim, Y.S. Review of metal oxide semiconductors-based thin-film transistors for point-of-care sensor applications. J. Inf. Disp. 2020, 21, 203–210. [Google Scholar] [CrossRef]
  445. Sarkar, D.; Liu, W.; Xie, X.; Anselmo, A.C.; Mitragotri, S.; Banerjee, K. MoS2 field-effect transistor for next-generation label-free biosensors. ACS Nano 2014, 8, 3992–4003. [Google Scholar] [CrossRef]
  446. Smith, J.T.; Shah, S.S.; Goryll, M.; Stowell, J.R.; Allee, D.R. Flexible ISFET biosensor using IGZO metal oxide TFTs and an ITO sensing layer. IEEE Sens. J. 2013, 14, 937–938. [Google Scholar] [CrossRef]
  447. Jung, J.; Kim, S.J.; Lee, K.W.; Yoon, D.H.; Kim, Y.-g.; Kwak, H.Y.; Dugasani, S.R.; Park, S.H.; Kim, H.J. Approaches to label-free flexible DNA biosensors using low-temperature solution-processed InZnO thin-film transistors. Biosens. Bioelectron. 2014, 55, 99–105. [Google Scholar] [CrossRef]
  448. Rim, Y.S.; Bae, S.-H.; Chen, H.; Yang, J.L.; Kim, J.; Andrews, A.M.; Weiss, P.S.; Yang, Y.; Tseng, H.-R. Printable ultrathin metal oxide semiconductor-based conformal biosensors. ACS Nano 2015, 9, 12174–12181. [Google Scholar] [CrossRef]
  449. Rahman, M.M.; Saleh Ahammad, A.; Jin, J.-H.; Ahn, S.J.; Lee, J.-J. A comprehensive review of glucose biosensors based on nanostructured metal-oxides. Sensors 2010, 10, 4855–4886. [Google Scholar] [CrossRef]
  450. Kaushik, A.; Solanki, P.R.; Kaneto, K.; Kim, C.; Ahmad, S.; Malhotra, B.D. Nanostructured iron oxide platform for impedimetric cholesterol detection. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2010, 22, 1045–1055. [Google Scholar] [CrossRef]
  451. Umar, A.; Rahman, M.; Hahn, Y.-B. Ultra-sensitive hydrazine chemical sensor based on high-aspect-ratio ZnO nanowires. Talanta 2009, 77, 1376–1380. [Google Scholar] [CrossRef] [PubMed]
  452. Solanki, P.R.; Kaushik, A.; Ansari, A.A.; Malhotra, B. Nanostructured zinc oxide platform for cholesterol sensor. Appl. Phys. Lett. 2009, 94, 143901. [Google Scholar] [CrossRef]
  453. Singh, S.; Arya, S.K.; Pandey, P.; Malhotra, B.; Saha, S.; Sreenivas, K.; Gupta, V. Cholesterol biosensor based on rf sputtered zinc oxide nanoporous thin film. Appl. Phys. Lett. 2007, 91, 063901. [Google Scholar] [CrossRef]
  454. Wang, C.; Tan, X.; Chen, S.; Yuan, R.; Hu, F.; Yuan, D.; Xiang, Y. Highly-sensitive cholesterol biosensor based on platinum–gold hybrid functionalized ZnO nanorods. Talanta 2012, 94, 263–270. [Google Scholar] [CrossRef]
  455. Ahmad, M.; Pan, C.; Gan, L.; Nawaz, Z.; Zhu, J. Highly sensitive amperometric cholesterol biosensor based on Pt-incorporated fullerene-like ZnO nanospheres. J. Phys. Chem. C 2010, 114, 243–250. [Google Scholar] [CrossRef]
  456. Ahmad, R.; Tripathy, N.; Hahn, Y.-B. Wide linear-range detecting high sensitivity cholesterol biosensors based on aspect-ratio controlled ZnO nanorods grown on silver electrodes. Sens. Actuators B Chem. 2012, 169, 382–386. [Google Scholar] [CrossRef]
  457. Khan, R.; Kaushik, A.; Solanki, P.R.; Ansari, A.A.; Pandey, M.K.; Malhotra, B. Zinc oxide nanoparticles-chitosan composite film for cholesterol biosensor. Anal. Chim. Acta 2008, 616, 207–213. [Google Scholar] [CrossRef]
  458. Ansari, A.A.; Kaushik, A.; Solanki, P.R.; Malhotra, B.E.D. Electrochemical cholesterol sensor based on tin oxide-chitosan nanobiocomposite film. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2009, 21, 965–972. [Google Scholar] [CrossRef]
  459. Salimi, A.; Hallaj, R.; Soltanian, S. Fabrication of a Sensitive Cholesterol Biosensor Based on Cobalt-oxide Nanostructures Electrodeposited onto Glassy Carbon Electrode. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2009, 21, 2693–2700. [Google Scholar] [CrossRef]
  460. Ali, S.M.U.; Ibupoto, Z.H.; Salman, S.; Nur, O.; Willander, M.; Danielsson, B. Selective determination of urea using urease immobilized on ZnO nanowires. Sens. Actuators B Chem. 2011, 160, 637–643. [Google Scholar] [CrossRef]
  461. Chen, X.; Yang, Z.; Si, S. Potentiometric urea biosensor based on immobilization of urease onto molecularly imprinted TiO2 film. J. Electroanal. Chem. 2009, 635, 1–6. [Google Scholar] [CrossRef]
  462. Kaushik, A.; Solanki, P.R.; Ansari, A.A.; Sumana, G.; Ahmad, S.; Malhotra, B.D. Iron oxide-chitosan nanobiocomposite for urea sensor. Sens. Actuators B Chem. 2009, 138, 572–580. [Google Scholar] [CrossRef]
  463. Ali, A.; Ansari, A.A.; Kaushik, A.; Solanki, P.R.; Barik, A.; Pandey, M.; Malhotra, B. Nanostructured zinc oxide film for urea sensor. Mater. Lett. 2009, 63, 2473–2475. [Google Scholar] [CrossRef]
  464. Solanki, P.R.; Dhand, C.; Kaushik, A.; Ansari, A.A.; Sood, K.; Malhotra, B. Nanostructured cerium oxide film for triglyceride sensor. Sens. Actuators B Chem. 2009, 141, 551–556. [Google Scholar] [CrossRef]
  465. Luo, Y.; Liu, H.; Rui, Q.; Tian, Y. Detection of extracellular H2O2 released from human liver cancer cells based on TiO2 nanoneedles with enhanced electron transfer of cytochrome c. Anal. Chem. 2009, 81, 3035–3041. [Google Scholar] [CrossRef]
  466. Wang, J.; Li, S.; Zhang, Y. A sensitive DNA biosensor fabricated from gold nanoparticles, carbon nanotubes, and zinc oxide nanowires on a glassy carbon electrode. Electrochim. Acta 2010, 55, 4436–4440. [Google Scholar] [CrossRef]
  467. Zhang, W.; Zheng, X.; Jiao, K. Label-free and enhanced DNA sensing platform for PML/RARA fusion gene detection based on nano-ZnO functionalized carbon ionic liquid electrode. Sens. Actuators B Chem. 2012, 162, 396–399. [Google Scholar] [CrossRef]
  468. Zuo, S.-H.; Zhang, L.-F.; Yuan, H.-H.; Lan, M.-B.; Lawrance, G.A.; Wei, G. Electrochemical detection of DNA hybridization by using a zirconia modified renewable carbon paste electrode. Bioelectrochemistry 2009, 74, 223–226. [Google Scholar] [CrossRef] [PubMed]
  469. Zhu, H.; Wang, J.; Xu, G. Fast synthesis of Cu2O hollow microspheres and their application in DNA biosensor of hepatitis B virus. Cryst. Growth Des. 2009, 9, 633–638. [Google Scholar] [CrossRef]
  470. Prabhakar, N.; Solanki, P.R.; Kaushik, A.; Pandey, M.; Malhotra, B. Peptide nucleic acid immobilized biocompatible silane nanocomposite platform for Mycobacterium tuberculosis detection. Electroanalysis 2010, 22, 2672–2682. [Google Scholar] [CrossRef]
  471. Abu-Salah, K.M.; Alrokyan, S.A.; Khan, M.N.; Ansari, A.A. Nanomaterials as analytical tools for genosensors. Sensors 2010, 10, 963–993. [Google Scholar] [CrossRef] [PubMed]
  472. Das, M.; Sumana, G.; Nagarajan, R.; Malhotra, B. Zirconia based nucleic acid sensor for Mycobacterium tuberculosis detection. Appl. Phys. Lett. 2010, 96, 133703. [Google Scholar] [CrossRef]
  473. Sun, W.; Qin, P.; Gao, H.; Li, G.; Jiao, K. Electrochemical DNA biosensor based on chitosan/nano-V2O5/MWCNTs composite film modified carbon ionic liquid electrode and its application to the LAMP product of Yersinia enterocolitica gene sequence. Biosens. Bioelectron. 2010, 25, 1264–1270. [Google Scholar] [CrossRef]
  474. Tsekenis, G.; Chatzipetrou, M.; Massaouti, M.; Zergioti, I. Comparative Assessment of Affinity-Based Techniques for Oriented Antibody Immobilization towards Immunosensor Performance Optimization. J. Sens. 2019, 2019, 6754398. [Google Scholar]
  475. Gajos, K.; Szafraniec, K.; Petrou, P.; Budkowski, A. Surface density dependent orientation and immunological recognition of antibody on silicon: TOF-SIMS and surface analysis of two covalent immobilization methods. Appl. Surf. Sci. 2020, 518, 146269. [Google Scholar] [CrossRef]
  476. Wang, J. Analytical Electrochemistry, 3rd ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2006. [Google Scholar]
  477. Che, X.; Yuan, R.; Chai, Y.; Li, J.; Song, Z.; Wang, J. Amperometric immunosensor for the determination of α-1-fetoprotein based on multiwalled carbon nanotube–silver nanoparticle composite. J. Colloid Interface Sci. 2010, 345, 174–180. [Google Scholar] [CrossRef]
  478. Ra, H.-W.; Kim, J.-T.; Khan, R.; Sharma, D.; Yook, Y.-G.; Hahn, Y.-B.; Park, J.-H.; Kim, D.-G.; Im, Y.-H. Robust and multifunctional nanosheath for chemical and biological nanodevices. Nano Lett. 2012, 12, 1891–1897. [Google Scholar] [CrossRef]
  479. Wei, Q.; Xiang, Z.; He, J.; Wang, G.; Li, H.; Qian, Z.; Yang, M. Dumbbell-like Au-Fe3O4 nanoparticles as label for the preparation of electrochemical immunosensors. Biosens. Bioelectron. 2010, 26, 627–631. [Google Scholar] [CrossRef]
  480. Gou, N.; Onnis-Hayden, A.; Gu, A.Z. Mechanistic toxicity assessment of nanomaterials by whole-cell-array stress genes expression analysis. Environ. Sci. Technol. 2010, 44, 5964–5970. [Google Scholar] [CrossRef]
  481. Yang, Z.; Ye, Z.; Zhao, B.; Zong, X.; Wang, P. A rapid response time and highly sensitive amperometric glucose biosensor based on ZnO nanorod via citric acid-assisted annealing route. Phys. E Low-Dimens. Syst. Nanostructures 2010, 42, 1830–1833. [Google Scholar] [CrossRef]
  482. Kong, T.; Chen, Y.; Ye, Y.; Zhang, K.; Wang, Z.; Wang, X. An amperometric glucose biosensor based on the immobilization of glucose oxidase on the ZnO nanotubes. Sens. Actuators B Chem. 2009, 138, 344–350. [Google Scholar] [CrossRef]
  483. Lei, Y.; Yan, X.; Luo, N.; Song, Y.; Zhang, Y. ZnO nanotetrapod network as the adsorption layer for the improvement of glucose detection via multiterminal electron-exchange. Colloids Surf. A: Physicochem. Eng. Asp. 2010, 361, 169–173. [Google Scholar] [CrossRef]
  484. Yang, K.; She, G.-W.; Wang, H.; Ou, X.-M.; Zhang, X.-H.; Lee, C.-S.; Lee, S.-T. ZnO nanotube arrays as biosensors for glucose. J. Phys. Chem. 2009, 113, 20169–20172. [Google Scholar] [CrossRef]
  485. Pradhan, D.; Niroui, F.; Leung, K. High-performance, flexible enzymatic glucose biosensor based on ZnO nanowires supported on a gold-coated polyester substrate. ACS Appl. Mater. Interfaces 2010, 2, 2409–2412. [Google Scholar] [CrossRef]
  486. Wang, Y.-T.; Yu, L.; Zhu, Z.-Q.; Zhang, J.; Zhu, J.-Z.; Fan, C.-H. Improved enzyme immobilization for enhanced bioelectrocatalytic activity of glucose sensor. Sens. Actuators B Chem. 2009, 136, 332–337. [Google Scholar] [CrossRef]
  487. Bao, S.J.; Li, C.M.; Zang, J.F.; Cui, X.Q.; Qiao, Y.; Guo, J. New nanostructured TiO2 for direct electrochemistry and glucose sensor applications. Adv. Funct. Mater. 2008, 18, 591–599. [Google Scholar] [CrossRef]
  488. Jang, H.D.; Kim, S.K.; Chang, H.; Roh, K.-M.; Choi, J.-W.; Huang, J. A glucose biosensor based on TiO2–graphene composite. Biosens. Bioelectron. 2012, 38, 184–188. [Google Scholar] [CrossRef]
  489. Umar, A.; Rahman, M.; Al-Hajry, A.; Hahn, Y.-B. Enzymatic glucose biosensor based on flower-shaped copper oxide nanostructures composed of thin nanosheets. Electrochem. Commun. 2009, 11, 278–281. [Google Scholar] [CrossRef]
  490. Patil, D.; Dung, N.Q.; Jung, H.; Ahn, S.Y.; Jang, D.M.; Kim, D. Enzymatic glucose biosensor based on CeO2 nanorods synthesized by non-isothermal precipitation. Biosens. Bioelectron. 2012, 31, 176–181. [Google Scholar] [CrossRef] [PubMed]
  491. Yang, L.; Ren, X.; Tang, F.; Zhang, L. A practical glucose biosensor based on Fe3O4 nanoparticles and chitosan/nafion composite film. Biosens. Bioelectron. 2009, 25, 889–895. [Google Scholar] [CrossRef]
  492. Umar, A.; Rahman, M.; Hahn, Y.-B. MgO polyhedral nanocages and nanocrystals based glucose biosensor. Electrochem. Commun. 2009, 11, 1353–1357. [Google Scholar] [CrossRef]
  493. Umar, A.; Rahman, M.; Vaseem, M.; Hahn, Y.-B. Ultra-sensitive cholesterol biosensor based on low-temperature grown ZnO nanoparticles. Electrochem. Commun. 2009, 11, 118–121. [Google Scholar] [CrossRef]
  494. Israr, M.Q.; Sadaf, J.R.; Asif, M.; Nur, O.; Willander, M.; Danielsson, B. Potentiometric cholesterol biosensor based on ZnO nanorods chemically grown on Ag wire. Thin Solid Film 2010, 519, 1106–1109. [Google Scholar] [CrossRef]
  495. Ansari, A.A.; Kaushik, A.; Solanki, P.; Malhotra, B. Sol–gel derived nanoporous cerium oxide film for application to cholesterol biosensor. Electrochem. Commun. 2008, 10, 1246–1249. [Google Scholar] [CrossRef]
  496. Zhang, W.; Yang, T.; Huang, D.M.; Jiao, K. Electrochemical sensing of DNA immobilization and hybridization based on carbon nanotubes/nano zinc oxide/chitosan composite film. Chin. Chem. Lett. 2008, 19, 589–591. [Google Scholar] [CrossRef]
  497. Zhang, W.; Yang, T.; Zhuang, X.; Guo, Z.; Jiao, K. An ionic liquid supported CeO2 nanoshuttles–carbon nanotubes composite as a platform for impedance DNA hybridization sensing. Biosens. Bioelectron. 2009, 24, 2417–2422. [Google Scholar] [CrossRef]
  498. Ansari, A.A.; Kaushik, A.; Solanki, P.R.; Malhotra, B. Nanostructured zinc oxide platform for mycotoxin detection. Bioelectrochemistry 2010, 77, 75–81. [Google Scholar] [CrossRef]
  499. Fu, X.; Wang, J.; Li, N.; Wang, L.; Pu, L. Label-free electrochemical immunoassay of carcinoembryonic antigen in human serum using magnetic nanorods as sensing probes. Microchim. Acta 2009, 165, 437–442. [Google Scholar] [CrossRef]
  500. Wang, L.; Jia, X.; Zhou, Y.; Xie, Q.; Yao, S. Sandwich-type amperometric immunosensor for human immunoglobulin G using antibody-adsorbed Au/SiO2 nanoparticles. Microchim. Acta 2010, 168, 245–251. [Google Scholar] [CrossRef]
  501. Eryiğit, M.; Cepni, E.; Urhan, B.K.; Doğan, H.Ö.; Özer, T.Ö. Nonenzymatic glucose sensor based on poly (3, 4-ethylene dioxythiophene)/electroreduced graphene oxide modified gold electrode. Synth. Met. 2020, 268, 116488. [Google Scholar] [CrossRef]
  502. Nguyen, T.N.; Jin, X.; Nolan, J.K.; Xu, J.; Le, K.V.H.; Lam, S.; Wang, Y.; Alam, M.A.; Lee, H. Printable nonenzymatic glucose biosensors using carbon nanotube-PtNP nanocomposites modified with AuRu for improved selectivity. ACS Biomater. Sci. Eng. 2020, 6, 5315–5325. [Google Scholar] [CrossRef] [PubMed]
  503. Hwang, D.-W.; Lee, S.; Seo, M.; Chung, T.D. Recent advances in electrochemical non-enzymatic glucose sensors—A review. Anal. Chim. Acta 2018, 1033, 1–34. [Google Scholar] [CrossRef] [PubMed]
  504. Burke, L. Premonolayer oxidation and its role in electrocatalysis. Electrochim. Acta 1994, 39, 1841–1848. [Google Scholar] [CrossRef]
  505. Xiao, F.; Li, Y.; Gao, H.; Ge, S.; Duan, H. Growth of coral-like PtAu–MnO2 binary nanocomposites on free-standing graphene paper for flexible nonenzymatic glucose sensors. Biosens. Bioelectron. 2013, 41, 417–423. [Google Scholar] [CrossRef]
  506. Lee, S.; Lee, J.; Park, S.; Boo, H.; Kim, H.C.; Chung, T.D. Disposable non-enzymatic blood glucose sensing strip based on nanoporous platinum particles. Appl. Mater. Today 2018, 10, 24–29. [Google Scholar] [CrossRef]
  507. Chen, H.; Fan, G.; Zhao, J.; Qiu, M.; Sun, P.; Fu, Y.; Han, D.; Cui, G. A portable micro glucose sensor based on copper-based nanocomposite structure. New J. Chem. 2019, 43, 7806–7813. [Google Scholar] [CrossRef]
  508. Liu, X.; Yang, W.; Chen, L.; Jia, J. Three-dimensional copper foam supported CuO nanowire arrays: An efficient non-enzymatic glucose sensor. Electrochim. Acta 2017, 235, 519–526. [Google Scholar] [CrossRef]
  509. Guascito, M.R.; Chirizzi, D.; Malitesta, C.; Siciliano, M.; Siciliano, T.; Tepore, A. Amperometric non-enzymatic bimetallic glucose sensor based on platinum tellurium microtubes modified electrode. Electrochem. Commun. 2012, 22, 45–48. [Google Scholar] [CrossRef]
Figure 1. Band structure of In2O3 near the Brillouin zone. Here, a weak optical absorption is observed at 2.7 eV and a strong optical transition occurs between lower − lying valence bands [112].
Figure 1. Band structure of In2O3 near the Brillouin zone. Here, a weak optical absorption is observed at 2.7 eV and a strong optical transition occurs between lower − lying valence bands [112].
Sensors 23 06849 g001
Figure 2. Schematic illustration of (a) H2S gas detection via hollow nanofibers and (b) the formation of the depletion layer at the p−n interface [130].
Figure 2. Schematic illustration of (a) H2S gas detection via hollow nanofibers and (b) the formation of the depletion layer at the p−n interface [130].
Sensors 23 06849 g002
Figure 3. TiO2 sensing mechanism shown schematically with exposure to LPG and in air. (ac) TiO2 sensor in air, where ionic species (O2, O and O2−) form due to adsorption of oxygen from ambient air on the surface film and capture the electrons from n-type TiO2; (df) When LPG exposed and interacted with adsorbed oxygen, large number of electrons re-injected on TiO2 surface and decreased the barrier hight [131].
Figure 3. TiO2 sensing mechanism shown schematically with exposure to LPG and in air. (ac) TiO2 sensor in air, where ionic species (O2, O and O2−) form due to adsorption of oxygen from ambient air on the surface film and capture the electrons from n-type TiO2; (df) When LPG exposed and interacted with adsorbed oxygen, large number of electrons re-injected on TiO2 surface and decreased the barrier hight [131].
Sensors 23 06849 g003
Figure 5. Schematic view of band−bending after the ionosorption of oxygen (chemisorption), where EC, EV, and EF denote the energies of the conduction band, valence band, and Fermi level, respectively. “e“ represents conducting electrons and “+” represents donor sites [13].
Figure 5. Schematic view of band−bending after the ionosorption of oxygen (chemisorption), where EC, EV, and EF denote the energies of the conduction band, valence band, and Fermi level, respectively. “e“ represents conducting electrons and “+” represents donor sites [13].
Sensors 23 06849 g005
Figure 6. The structural and band models in the (a) presence of CO and (b) absence of CO [13,176].
Figure 6. The structural and band models in the (a) presence of CO and (b) absence of CO [13,176].
Sensors 23 06849 g006
Figure 7. (a) Sensing mechanism of pristine n-type SMOx materials under four conditions in the reception process: (i) formation of an accumulation layer (brown color) in the presence of reduced gas with the absence of oxygen, (ii) flat band formation in the absence of surface states due to adsorbed species, (iii) formation of a depletion layer (yellow color) in the presence of oxygen and reducing gases, and (iv) formation of a depletion layer without reducing gas in the presence of oxygen. (b) Charge transport in the sensing layer depicting tentative resistance in the transduction process [20].
Figure 7. (a) Sensing mechanism of pristine n-type SMOx materials under four conditions in the reception process: (i) formation of an accumulation layer (brown color) in the presence of reduced gas with the absence of oxygen, (ii) flat band formation in the absence of surface states due to adsorbed species, (iii) formation of a depletion layer (yellow color) in the presence of oxygen and reducing gases, and (iv) formation of a depletion layer without reducing gas in the presence of oxygen. (b) Charge transport in the sensing layer depicting tentative resistance in the transduction process [20].
Sensors 23 06849 g007
Figure 8. (a) Simplified representation of the essential sensing layer components for p-type oxide semiconductor gas sensors (low), where A and C are metal–semiconductor contacts and B is a grain-grain contact of a semiconductor. The energy band diagram described by Barsan and co-workers [166]: (b,c) p-type oxide semiconductors with a simplified gas-sensing mechanism and equivalent circuit [182].
Figure 8. (a) Simplified representation of the essential sensing layer components for p-type oxide semiconductor gas sensors (low), where A and C are metal–semiconductor contacts and B is a grain-grain contact of a semiconductor. The energy band diagram described by Barsan and co-workers [166]: (b,c) p-type oxide semiconductors with a simplified gas-sensing mechanism and equivalent circuit [182].
Sensors 23 06849 g008
Figure 9. Sensitization mechanisms of SMOx by metal or metal oxide additives. (a) Electronic sensitization via changes in the Fermi level and (b) chemical sensitization via spillover [13].
Figure 9. Sensitization mechanisms of SMOx by metal or metal oxide additives. (a) Electronic sensitization via changes in the Fermi level and (b) chemical sensitization via spillover [13].
Sensors 23 06849 g009
Figure 10. Schematic view of catalytic activation based on oxidation mechanism. (a) Direct reaction on the additive surface; (b) adsorption of oxygen on SMOx additive; (c) spillover of reactive species on the SMOx surface. O = oxygen (red colour), C = carbon (blue colour) on metal nanoparticle surface [189].
Figure 10. Schematic view of catalytic activation based on oxidation mechanism. (a) Direct reaction on the additive surface; (b) adsorption of oxygen on SMOx additive; (c) spillover of reactive species on the SMOx surface. O = oxygen (red colour), C = carbon (blue colour) on metal nanoparticle surface [189].
Sensors 23 06849 g010
Figure 11. Schematic view of spillover mechanism on a Au− loaded SnO2 gas sensor where (a) inverse oxygen spillover and (b) oxygen spillover take place [191,197].
Figure 11. Schematic view of spillover mechanism on a Au− loaded SnO2 gas sensor where (a) inverse oxygen spillover and (b) oxygen spillover take place [191,197].
Sensors 23 06849 g011
Figure 12. Sensor signal changes corresponding to changes in the surface charge for a SnO2—based sensor at 300 °C [199].
Figure 12. Sensor signal changes corresponding to changes in the surface charge for a SnO2—based sensor at 300 °C [199].
Sensors 23 06849 g012
Figure 13. (a) Schematic view of a SnO2 nanowire surface in the presence of O2 and (1) ionosorption of oxygen at defect sites of the pristine surface; (2) molecular oxygen dissociation on Pd nanoparticles followed by spillover of the atomic species onto the oxide surface; (3) capture by a Pd nanoparticle of weakly adsorbed molecular oxygen. (b) band diagram of pristine SnO2 in the vicinity of a Pd nanoparticle. The radius of the depletion region is determined by the radius of the spillover zone [216].
Figure 13. (a) Schematic view of a SnO2 nanowire surface in the presence of O2 and (1) ionosorption of oxygen at defect sites of the pristine surface; (2) molecular oxygen dissociation on Pd nanoparticles followed by spillover of the atomic species onto the oxide surface; (3) capture by a Pd nanoparticle of weakly adsorbed molecular oxygen. (b) band diagram of pristine SnO2 in the vicinity of a Pd nanoparticle. The radius of the depletion region is determined by the radius of the spillover zone [216].
Sensors 23 06849 g013
Figure 14. (a) Effect of particle size on CO gas sensitivity [225]; (b) effect of particle size on H2 gas sensitivity [217].
Figure 14. (a) Effect of particle size on CO gas sensitivity [225]; (b) effect of particle size on H2 gas sensitivity [217].
Sensors 23 06849 g014
Figure 15. Response of Sm2O3 doped SnO2 sensor for different concentrations of C2H2 at different relative humidity (RH) conditions [228].
Figure 15. Response of Sm2O3 doped SnO2 sensor for different concentrations of C2H2 at different relative humidity (RH) conditions [228].
Sensors 23 06849 g015
Figure 16. Response of a porous ZnO gas sensor as a function of operating temperatures [230].
Figure 16. Response of a porous ZnO gas sensor as a function of operating temperatures [230].
Sensors 23 06849 g016
Figure 17. SEM images of (a) ZnO before a chemical treatment and a (b) hexagonal ZnO nanorod [236].
Figure 17. SEM images of (a) ZnO before a chemical treatment and a (b) hexagonal ZnO nanorod [236].
Sensors 23 06849 g017
Figure 18. Schematic view of the classification of chemical sensors according to IUPAC [165].
Figure 18. Schematic view of the classification of chemical sensors according to IUPAC [165].
Sensors 23 06849 g018
Figure 19. Schematic view of the principle of chemical sensors.
Figure 19. Schematic view of the principle of chemical sensors.
Sensors 23 06849 g019
Figure 20. SEM images of ZnO material at (a) low magnification and (b) high magnification. (c) Cyclic voltammetry curve of a Nafion/ZnO/Au electrode in the absence of hydrazine (solid line) and presence of 1 mM N2H4 (dashed line) in 0.01 M phosphate − buffered saline (PBS) (pH = 7.4). The scan rate was 100 mV s−1. (d) Amperometric response of a Nafion/ZnO/Au electrode in the presence of hydrazine. The inset shows the 1/i versus 1/C plot [370].
Figure 20. SEM images of ZnO material at (a) low magnification and (b) high magnification. (c) Cyclic voltammetry curve of a Nafion/ZnO/Au electrode in the absence of hydrazine (solid line) and presence of 1 mM N2H4 (dashed line) in 0.01 M phosphate − buffered saline (PBS) (pH = 7.4). The scan rate was 100 mV s−1. (d) Amperometric response of a Nafion/ZnO/Au electrode in the presence of hydrazine. The inset shows the 1/i versus 1/C plot [370].
Sensors 23 06849 g020
Figure 21. (a) Schematic view of the fabrication process of the electrode; (b) (i–iv) field emission scanning electron microscopy (FESEM) images of ZnO nanorod arrays (grown without electrode) and (v–viii) ZnO nanorod arrays grown directly on the probe; (c) dynamic responses of ZnO nanorod arrays to H2 pulses at 250 °C; (d) demonstration of the sensitivity at various temperatures [371].
Figure 21. (a) Schematic view of the fabrication process of the electrode; (b) (i–iv) field emission scanning electron microscopy (FESEM) images of ZnO nanorod arrays (grown without electrode) and (v–viii) ZnO nanorod arrays grown directly on the probe; (c) dynamic responses of ZnO nanorod arrays to H2 pulses at 250 °C; (d) demonstration of the sensitivity at various temperatures [371].
Sensors 23 06849 g021
Figure 22. (a) Schematic view of the fabrication of a SMOx-based chemical sensor; (b) time-dependent resistance for continuous exposure of the sensor to CO at 350 °C; (c) sensitivity of undoped and 0.76% and 1.85% co-doped ZnO sensors exposed to different concentrations of CO [372].
Figure 22. (a) Schematic view of the fabrication of a SMOx-based chemical sensor; (b) time-dependent resistance for continuous exposure of the sensor to CO at 350 °C; (c) sensitivity of undoped and 0.76% and 1.85% co-doped ZnO sensors exposed to different concentrations of CO [372].
Sensors 23 06849 g022
Figure 23. Gas-sensing performance of meso- and macroporous Co3O4 nanorod-based sensors. (a) ethanol sensing at different temperatures using N2 as the reference. (b) Ethanol sensing at different temperatures using dry air as the reference. (c) gas-sensing property of porous Co3O4 nanorods to acetone, ethanol, and benzene at 300 °C. The sensor resistance changes in response to different concentrations of acetone, ethanol, and benzene. (d) Response of nanoparticles (NPs), meso-/macroporous nanorods (NRs), and porous plates to different concentrations of acetone, ethanol, and benzene [376].
Figure 23. Gas-sensing performance of meso- and macroporous Co3O4 nanorod-based sensors. (a) ethanol sensing at different temperatures using N2 as the reference. (b) Ethanol sensing at different temperatures using dry air as the reference. (c) gas-sensing property of porous Co3O4 nanorods to acetone, ethanol, and benzene at 300 °C. The sensor resistance changes in response to different concentrations of acetone, ethanol, and benzene. (d) Response of nanoparticles (NPs), meso-/macroporous nanorods (NRs), and porous plates to different concentrations of acetone, ethanol, and benzene [376].
Sensors 23 06849 g023
Figure 24. Schematic view of the principle of a biosensor.
Figure 24. Schematic view of the principle of a biosensor.
Sensors 23 06849 g024
Figure 25. Mechanism of SMOx—based enzyme biosensors in three steps: (i) band energies; (ii) crystalline structure; (iii) configuration of an enzyme biosensor [32].
Figure 25. Mechanism of SMOx—based enzyme biosensors in three steps: (i) band energies; (ii) crystalline structure; (iii) configuration of an enzyme biosensor [32].
Sensors 23 06849 g025
Figure 26. (a) Fabrication of zinc oxide nanoflowers; (b) SEM image of ZnO nanofibers; (c) cyclic voltammetry curves for a modified gold electrode without and with 100 mM glucose in PBS solution (pH = 7.0); (d) amperometric response of the ZnO nanoflower biosensor in different concentrations of glucose at 0.8 V in PBS solution (at pH = 7.0) [438].
Figure 26. (a) Fabrication of zinc oxide nanoflowers; (b) SEM image of ZnO nanofibers; (c) cyclic voltammetry curves for a modified gold electrode without and with 100 mM glucose in PBS solution (pH = 7.0); (d) amperometric response of the ZnO nanoflower biosensor in different concentrations of glucose at 0.8 V in PBS solution (at pH = 7.0) [438].
Sensors 23 06849 g026
Figure 27. In2O3 FET− based biosensor. (a) Images of a contacted device on an artificial eye for glucose sensing in tears. Thin-film sensor contact with the skin during tension and relaxation; (b,c) device performance of thin-film In2O3 on a rigid substrate and flexible artificial PDMS skin, with the transfer of In2O3 FETs to replicas of skin under liquid gating with PBS solution at low voltage; (d) SEM image of In2O3 FET device on artificial PDMS skin replica; (e) enzymatic oxidation of D−glucose to generate gluconic acid and H2O2; (f) representation of In2O3 sensors for the concentration of D-glucose in a low range of human diabetic tears and high range of blood, with the standard deviation shown in the inset [448].
Figure 27. In2O3 FET− based biosensor. (a) Images of a contacted device on an artificial eye for glucose sensing in tears. Thin-film sensor contact with the skin during tension and relaxation; (b,c) device performance of thin-film In2O3 on a rigid substrate and flexible artificial PDMS skin, with the transfer of In2O3 FETs to replicas of skin under liquid gating with PBS solution at low voltage; (d) SEM image of In2O3 FET device on artificial PDMS skin replica; (e) enzymatic oxidation of D−glucose to generate gluconic acid and H2O2; (f) representation of In2O3 sensors for the concentration of D-glucose in a low range of human diabetic tears and high range of blood, with the standard deviation shown in the inset [448].
Sensors 23 06849 g027
Figure 28. (a) Fabrication process flow for a cholesterol biosensor (i–vi); (b) relationship between the enzyme loading, the aspect ratio of ZnO nanorods, and growth time; (c) amperometric responses at an applied potential of +0.38 V for different aspect ratios of ZnO nanorods in the presence of cholesterol; (d) calibration curves of current response versus cholesterol concentration [456].
Figure 28. (a) Fabrication process flow for a cholesterol biosensor (i–vi); (b) relationship between the enzyme loading, the aspect ratio of ZnO nanorods, and growth time; (c) amperometric responses at an applied potential of +0.38 V for different aspect ratios of ZnO nanorods in the presence of cholesterol; (d) calibration curves of current response versus cholesterol concentration [456].
Sensors 23 06849 g028
Figure 29. Electrical characterization of amino-functionalized ZnO/PAC nanowires for AFP detection. (a) Schematic view of anti−AFP immobilization via the sandwich binding method. The anti-AFP antibody was labeled with TRITC. Fluorescence microscopy images of an (b) untreated ZnO/PAC nanowire with AFP antigen, (c) an amino-functionalized ZnO/PAC nanowire with AFP antigen, and (d) an amino-functionalized ZnO/PAC nanowire with liver carcinoma. (e) Schematic view of an electrolyte-gated ZnO/PAC nanowire-based FET. (f) Current versus potential graph of a ZnO/PAC nanowire-based FET for sequential immobilizations [478].
Figure 29. Electrical characterization of amino-functionalized ZnO/PAC nanowires for AFP detection. (a) Schematic view of anti−AFP immobilization via the sandwich binding method. The anti-AFP antibody was labeled with TRITC. Fluorescence microscopy images of an (b) untreated ZnO/PAC nanowire with AFP antigen, (c) an amino-functionalized ZnO/PAC nanowire with AFP antigen, and (d) an amino-functionalized ZnO/PAC nanowire with liver carcinoma. (e) Schematic view of an electrolyte-gated ZnO/PAC nanowire-based FET. (f) Current versus potential graph of a ZnO/PAC nanowire-based FET for sequential immobilizations [478].
Sensors 23 06849 g029
Table 1. Some semiconductor metal oxide (SMOx)-based sensors with their structural parameters.
Table 1. Some semiconductor metal oxide (SMOx)-based sensors with their structural parameters.
SMOxLattice ParametersApplicationsReferences
Nickel oxide (NiO)Cubic1. NO2, CO gas sensing[58,59,60,61,62,63,64,65]
2. Ammonia sensing
Fm3m3. Ethanol sensing
a = 2.983 Å4. Uric acid sensing
b = 2.983 Å5. Lactic acid sensing
c = 5.160 Å6. Glucose sensing
Cobalt oxide (CoO)Cubic1. Gas sensing[66,67,68,69]
2. Oxygen sensing
Fm3m3. Aceton sensing
a = 3.024 Å
b = 3.012 Å
c = 5.316 Å
Hexagonal
P63mc
a = 3.269 Å
b = 5.289 Å
c = 5.646 Å
Tin dioxide (SnO2)Cubic1. Gas sensors[70,71,72,73,74,75,76,77,78,79,80,81]
2. Formaldehyde sensing
Fm3m3. H2S sensing
a = 3.640 Å4. Alkene sensing
b = 3.640 Å5. H2 sensing
c = 3.640 Å6. Biomarker of lung cancer
7. CO sensing
Tetragonal
P4/mnm
a = 4.832 Å
b = 4.832 Å
c = 3.243 Å
Titanium dioxide (TiO2)Tetragonal1. Hazardous gas sensing[82,83,84,85,86,87,88,89,90]
2. Gas and UV sensor
I41/amd3. Phosphopeptide sensing
a = 5.566 Å4. Chemical sensing
b = 5.566 Å5. Lactate sensing
c = 5.566 Å6. Biosensors
Tetragonal
P42/mnm
a = 4.653 Å
b = 4.653 Å
c = 2.969 Å
Zinc oxide (ZnO)Cubic1. Gas sensor[91,92,93,94,95,96,97,98]
2. H2 sensing
Fm3m3. Chemical sensing
a = 3.068 Å4. Pesticide detection
b = 3.068 Å5. Biosensors
c = 3.068 Å
Hexagonal
P63mc
a = 3.289 Å
b = 3.289 Å
c = 5.307 Å
Trimanganese tetraoxide (Mn3O4)Tetragonal1. H2 gas sensing[99,100]
2. Nitrogen sensing
I41/amd
a = 5.870 Å
b = 6.348 Å
c = 5.873 Å
Table 2. Different gas-sensing activities of SMOx-based nanomaterials.
Table 2. Different gas-sensing activities of SMOx-based nanomaterials.
Analyte GasLayer CompositionMeas. Temp.ConcentrationResponsetrestrecSelective AgentsRef.
CO40% In2O3-SnO2250 °C1000 ppm16NANANA[322]
SnO2@In2O3300 °C200 ppm1.9135 s460 sNA[323]
SnO2@NiO250 °C500 ppm15.9NANACH4[324]
50 wt% Co3O4-SnO2100 °C1000 ppm175NANAH2[325]
(3 wt% ZnO-SnO2)@
CuO
235 °C200 ppm13.4NANAH2[326]
20 wt% WO3-MoO3200 °C15 ppm3002 min2 minNA[248]
H2(3 wt% ZnO-SnO2)@CuO305 °C200 ppm16NANACO[326]
SnO2@2.6mol% ZnO350 °C100 ppm18.4NANACO, NH3, CH4[327]
(0.005 mol MoO3)-SnO2240 °C1000 ppm105 s10 sNA[328]
1 wt% Co3O4-SnO2250 °C1000 ppm9100NANACO[325]
TiO2/NiO200 °C10,000
ppm
70NANANA[329]
ZnO@SnO2400 °C500 ppm70NANANA[330]
NO2SnO2@ZnORT5 ppm0.4~50 s~450 sNA[331]
40% ZnO-SnO2250 °C500 ppm34.5NANANA[332]
20% WO2-SnO2200 °C200 ppm186NANANA[333]
40% In2O3-SnO2200 °C1000 ppm7.5NANANA[322]
5% Eu2O3-ZnO300 °C3 ppm163 min3 minCO[334]
H2SCu2O/SnO2RT50 ppm45NANAToluene, LPG[335]
6% CuO/SnO2150 °C20 ppm43003 sNANA[336]
1 mol% CeO2-SnO2300 °K5 ppm340 s20 sLPG, EtOH,
NOx, CO
[337]
SnO2@ZnO350 °C500 ppm2.1NANACO, CH4[324]
(5 wt% ZnO-SnO2)@3.68 wt% CuO150 °C50 ppm60,44315 s7 minNOx, LPG, CO2, CH4[338]
SnO2350 °C50 ppm10NANANA[338]
ZnO100 °C100 ppm25NANANH3, MeOH, EtOH,butanol, acetone, ether[339]
NH3ZnO@Cr2O3RT300 ppm13.725 s75 sLPG, CO2, EtOH,
H2, Cl2
[299]
2mol%α-Fe2O3-ZnORT0.4 ppm10,00020 s20 sTMA, EtOH, MeOH[340]
EthanolZnSnO3@SnO2270 °C50 ppm27.81 s1.8 sAcetone, benzene,
chloroform,
MeOH,
formaldehyde, CO
[341]
5 wt% La2O3-SnO2300 °C1000 ppm74020 min20 minNA[318]
50 wt% SnO2-ZnO300 °C200 ppm4.6972 sNAAcetone, CO, H2,
NO2, C3H8
[342]
1.5 mol% Fe2O3-SnO2250 °C10 ppm24NANANA[343]
ZnO300 °C100 ppm7NANANA[344]
ZnO@ZnS210 °C1000 ppm2315 s15 sNA[345]
ZnO@ZnS
@Graphene
210 °C1000 ppm3815 s15 sAcetone,
formaldehyde,
benzene, cyclohexane
[345]
25 wt% SnO2-ZnO350 °C100 ppb~82NANANA[311]
2:8 mol CuO: ZnO115 °C100 ppm9613 s5 sNA[346]
20 wt% SnO2-TiO2553 °K200 ppm5110–15 s14–20 sNA[347]
ZnO-Co3O4170 °C100 ppm46NANANA[348]
Ethylene0.3 wt% WO3-SnO2300 °C6 ppm1.7~10 min~10
min
NA[349]
O3MoO3-TiO2300 °C100 ppb1.720 s2 minNA[350]
SO21 mol% NiO-SnO225 °C18 ppm0.844.5 min15 minO2, C3H8, NOx[351]
LPG[email protected] wt% Cr2O3350 °C100 ppm4618 s42 sNH3, CO2, EtOH, H2,[352]
TMA10 wt% ZnO-SnO2330 °C50 ppm1262 s5 sNH3, DMA, MA,
EtOH, MeOH,
Acetone
[353]
Table 3. Different SMOx-based chemical sensors.
Table 3. Different SMOx-based chemical sensors.
SMOx Composite MaterialsAnalyte GasDetection Limit (Conc.)Senor Response/TemperatureResponse/Recovery TimeReferences
SnO2 nanowireH210 ppm~0.4/300 °CN/A[358]
SnO2 hollow sphereCO50 ppm-/300–350 °C<1 min./30 min.[360]
SnO2 nanocrystalsNO2100 ppb-/300 °CN/A[361]
SnO2 porous NPsH2/CO160/200 ppm-/300 °CNA[362]
SnO2 flower-likeCO50 ppm~2.13/350 °C26 s/34 s[362]
SnO2 nanorodsH2100 ppm~13/150 °CN/A[386]
Plasma-modified SnO2 nanowireEthanol100 ppm-/250 °CN/A[366]
Pt@SnO2 nanorodsEthanol10 ppm3.7/300 °C2 s/20 s[367]
NiO-SnO2 nanofibersEthanol100 ppm25.5/300 °C2 s/3 s[387]
p-NiO/n-SnO2 heterojunction composite nanofibersH2100 ppm13.6/320 °C~3 S/~3 S[388]
ZnO nanorodsEthanol1 ppb~10/300 °C100 s/-[389]
ZnO nanorod arraysH2500 ppm-/250 °C6 min/17 min[390,391]
ZnO nanowireNO20.5 ppm-/225 °CN/A[392]
Pt adsorbed single crystalline ZnO nanowiresNH31000 ppm-/350 °C100 s/100 s[393]
Co-doped ZnO nanorodsCO50 ppm-/350 °CN/A[372]
Pd nanodots–functionalized ZnO nanowireCO100 ppb1.02/20 °C120 s/180 s[373]
Ga2O3 nanowireO21% O24.75/300 °CN/A[394]
NiO nanotubesEthanol200 ppm22.6/250 °CN/A[395]
α-Fe2O3 hollow spheresEthanol10 ppm~5/RTN/A[396]
Fe2TiO5/α-Fe2O3 nanocompositeEthanol10 ppm~10/320 °C28 s/21 s[397]
Fe2O3–TiO2 tube-like nanostructuresEthanol500 ppm8.2/270 °CN/A[377]
In2O3 nanofibersEthanol100 ppm~14/300 °C1 s/5 s[378]
In2O3 nanoparticlesNOx200 ppm~10,000/150 °CN/A[379]
Sn-doped In2O3 nanopowdersCO50 ppm4/250 °CN/A[380]
In2O3 hollow microspheresEthanol100 ppm137.2/400 °CN/A[381]
Ag-doped In2O3 nanoparticlesAlcohol vapors100 ppm-/150 °C42 s/34 s[383]
In2O3 nanorodsFormaldehyde32 ppm-/300 °C276 s/65 s[384]
Mesoporous In2O3 nanorodsEthanol500 ppb1.71/290 °C6 s/8 s[385]
Pt/In2O3 nanofibersH2S600 ppm1490/200 °C60 s/120 s[398]
CuO nanoribbonsMethanol5 ppm~1.4/RT2–4 s/3–7 s[399]
Porous CuO nanowireH26% H2407%/250 °C72 s/156 s[400]
Co3O4 hollow spheresButanol10 ppm3/100 °C1–3 s/4–8 s[401]
WO3 nanowiresH2S1 ppm48/250 °C-[273]
WO3 nanoplatesEthanol10 ppm~1.9/300 °C-[402]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dutta, T.; Noushin, T.; Tabassum, S.; Mishra, S.K. Road Map of Semiconductor Metal-Oxide-Based Sensors: A Review. Sensors 2023, 23, 6849. https://doi.org/10.3390/s23156849

AMA Style

Dutta T, Noushin T, Tabassum S, Mishra SK. Road Map of Semiconductor Metal-Oxide-Based Sensors: A Review. Sensors. 2023; 23(15):6849. https://doi.org/10.3390/s23156849

Chicago/Turabian Style

Dutta, Taposhree, Tanzila Noushin, Shawana Tabassum, and Satyendra K. Mishra. 2023. "Road Map of Semiconductor Metal-Oxide-Based Sensors: A Review" Sensors 23, no. 15: 6849. https://doi.org/10.3390/s23156849

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