Next Article in Journal
Effects of W Content on Structural and Mechanical Properties of TaWN Films
Previous Article in Journal
Enhanced Anti-Tribocorrosion Performance of Ti-DLC Coatings Deposited by Filtered Cathodic Vacuum Arc with the Optimization of Bias Voltage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gas Sensors Based on Titanium Oxides (Review)

1
Department of Electrochemical Material Science, Center for Physical Sciences and Technology (FTMC), Sauletekio Avenue 3, LT-10257 Vilnius, Lithuania
2
Department of Physical Chemistry, Faculty of Chemistry and Geosciences, Institute of Chemistry, Vilnius University, Naugarduko 24, LT-03225 Vilnius, Lithuania
*
Author to whom correspondence should be addressed.
Submission received: 9 November 2021 / Revised: 13 May 2022 / Accepted: 16 May 2022 / Published: 19 May 2022

Abstract

:
Nanostructured titanium compounds have recently been applied in the design of gas sensors. Among titanium compounds, titanium oxides (TiO2) are the most frequently used in gas sensing devices. Therefore, in this review, we are paying significant attention to the variety of allotropic modifications of titanium oxides, which include anatase, rutile, brukite. Very recently, the applicability of non-stoichiometric titanium oxide (TiO2−x)-based layers for the design of gas sensors was demonstrated. For this reason, in this review, we are addressing some research related to the formation of non-stoichiometric titanium oxide (TiO2−x) and Magnéli phase (TinO2n−1)-based layers suitable for sensor design. The most promising titanium compounds and hetero- and nano-structures based on these compounds are discussed. It is also outlined that during the past decade, many new strategies for the synthesis of TiO2 and conducting polymer-based composite materials were developed, which have found some specific application areas. Therefore, in this review, we are highlighting how specific formation methods, which can be used for the formation of TiO2 and conducting polymer composites, can be applied to tune composite characteristics that are leading towards advanced applications in these specific technological fields. The possibility to tune the sensitivity and selectivity of titanium compound-based sensing layers is addressed. In this review, some other recent reviews related to the development of sensors based on titanium oxides are overviewed. Some designs of titanium-based nanomaterials used for the development of sensors are outlined.

1. Introduction

Sensors for the determination of gasses and volatile organic compounds (VOCs) (gas-sensors) find applications in many different areas. Most of these gas sensors are based on semiconducting layers, which change their electrical resistance in the presence of gases and VOCs [1,2,3,4,5,6,7]. These semiconducting layers are mostly based on semiconductors such as WO3, MoS2, ZnO, SnO2, and TiO2 TiOn [2,3,4,5,6,7,8], while Al2O3 and SiO2 are the most common dielectric substrates used for the formation of gas sensors. Some new titanium-based compounds, such as MXenes [9], and non-stoichiometric titanium oxides (TiOn) [1,2,10,11] have also been recently used in the design of gas sensors. The stoichiometric TiO2 semiconductor of n-type, which appears in three major phases (rutile, anatase, and brookite) can be rather easily converted between these phases by thermal procedures, which paves the way to tune analytical characteristics of gas sensors [11]. These phases are sensitive and could be used for the detection of oxygen alone [12,13,14] or in heterostructures with ZrO2 to improve sensing characteristics [15,16,17]. Besides stoichiometric TiO2 forms, very recently, a non-stoichiometric (TiOn) that is sometimes indicated as TiO2−x has also been applied in sensor design [1]. Titanium oxide structures based on specific Magnéli phases (TinO2n−1) are very promising [10]. Many gas sensors are very sensitive, but mostly they are not very selective and consume a lot of electrical energy, which is required for the powering of the heater that is essential for the efficient operation of most gas sensors. In order to reduce the lastly mentioned disadvantage, the ‘self-heating’ strategy was developed, which enables the exploitation of the sensing layer as a heater of the analytical system [1,2,18]. Recently, some scientists haveattempted to advance the selectivity of gas-sensors by the design of specific morphology of semiconducting structures [19,20], application of core-shell semiconducting nano-composites [21,22,23] and/or quantum dots [24,25]. Various aspects of titanium-based material are overviewed in recent reviews, addressing the properties, synthesis and modifications of TiO2-based nanostructures [10,11,26]; application of titanium-based nanomaterials for energetics and environmental purposes [27]; the design of photo-catalysts [28]; the applicability of titanium oxides [1,2] and MXenes [9] in sensor design.
In this article, we are overviewing mainly the advances in the development of gas and volatile organic compounds (VOCs) sensors based on titanium-based oxides.

2. Structural Features and Physicochemical Properties of Stoichiometric and Non-Stoichiometric Titanium Oxides

TiO2 belongs to n-type semiconducting materials [29,30]. Many TiO2-based heterostructures are used in the design of sensors [11,31,32,33,34] and biosensors [35,36]. However, all the most popular forms of titanium oxides are characterized by specific bandgaps, which are as follows: (i) anatase by 3.02 eV (ii) rutile by 3.23 eV; (iii) brookite by 2.96 eV [37]. The annealing procedure is mostly used for the conversion of one titanium oxide phase into another one. Besides stoichiometric titanium oxide, plenty of non-stoichiometric forms were identified, among them very attractive conducting/semiconducting characteristics and gas-sensing properties [1]. They have Magnéli phases, which are described by TinO2n−1 stoichiometry, where n = 4, …, 10. Their Magnéli phase neighbor is titanium pentoxide (Ti3O5), where n = 3 with a stoichiometry of TinO2n−1, which appears in a variety of different forms (that are indicated as α−, β−, γ−, δ−, and λ−) [38,39,40,41,42]. Stoichiometry of titanium pentoxide corresponds to that of the Magnéli phases (TinO2n−1). Titanium pentoxide forms monoclinic crystals with the following lattice constants: a = 9.9701 Å, b = 5.0747 Å, c = 7.1810 Å, β = 109.865°. Moreover, titanium pentoxide, as well as some other Magnéli phases (e.g., Ti4O7), exhibit superconductivity when cooled down below 7 K temperature [43].
The most significant difference between titanium pentoxide and the Magnéli phases is determined by their different crystal structures. Magnéli phases contains shear planes based on TiO2(rutile) [44,45], while in titanium pentoxide, such planes are absent [1,43]. A temperature of 400 °C is optimal for the appearance of TiO2(anatase) intergrowths within Ti3O5 crystals based on TiO2−x/TiO2-based heterostructures [46,47]. The incorporation of intergrowths based on TiO2(anatase) in the structure of titanium pentoxide (Ti3O5) improves their conductivity (Figure 1) and some photoluminescence-related characteristics [1]. It should be noted that such non-stoichiometric titanium oxides can be spontaneously oxidized for this reason and significant attention should be paid to the stabilization of these structures during their usage in the development of gas sensors. It should be noted that at different oxidation states, titanium oxides have different crystal structures [48], which starts from rutile for TiO2 and anatase for Ti10O19 and then it turns into the triclinic structure for many stoichiometries, ranging from Ti9O17 until Ti4O7, monoclinic for γTi3O5, tetragonal for Ti2O3 and hexagonal for TiO, Ti2O and metallic titanium.
The electrical conductivity of most non-stoichiometric titanium oxides is significantly higher than the conductivity of any allotropic form of stoichiometric titanium oxides (TiO2). The most significant increase in sensitivity is observed for titanium oxides with TinO2n−1 stoichiometry, when index ‘n’ is in the range of 4–10 [49]. Some compounds with the above-mentioned stoichiometry form Magnéli phases, which are characterized by metallic conductivity and even by superconductivity [50,51].
Non-stoichiometric titanium oxide-based layers based on Magnéli phases have well conducting intergrowths based on TinO2n−1 planar moieties that penetrate the TiO2-based matrix [1,52]. Non-stoichiometric titanium oxides, such as Ti2O3 and/or Ti3O5, which do not form real Magnéli phases anyway, are much better at conducting in comparison to stoichiometric TiO2 [1,43]. These n-type semiconducting titanium oxides have a high concentration of ‘oxygen vacancies’, which are responsible for the mobility of electrons through their structure [53] and baseline resistance [54]. It should be outlined that these ‘oxygen vacancies’ are also responsible for the sensitivity towards both oxidizing and reducing gases [1]. During the design of the sensing layer, initially formed stoichiometric titanium oxide-based layers can be chemically reduced into non-stoichiometric titanium oxide (TiO2−x) and/or Magnéli phases [2]. Magnéli phases are observed when the oxygen concentration in TiO2−x structure is decreased and ‘x’ value is between 0.1 and 0.34 [1,55,56]. Non-stoichiometric titanium oxide structures containing Magnéli phases are chemically stable and rather well conducting. For these reasons, they are often applied in wastewater treatment and the design of batteries and fuel cells [57,58]; the same characteristics are required for gas sensors.
Non-stoichiometric titanium oxide-based structures can be developed using several approaches, namely plasma-treatment [59], laser irradiation-based modification [60], reduction by metallic zinc [61], bombardment by high-energy particles [62] and thermo-chemical approaches [63]. However, the formation of large Ti3O5 monocrystals is rather challenging because titanium oxides are polymorphic [43]. In some cases, stoichiometric titanium oxide can be easily turned into a non-stoichiometric one by suitable doping and/or reduction.
During the formation of TiO2−x/TiO2-based heterostructures, the ratio between stoichiometric titanium oxide and non-stoichiometric titanium oxide can be significantly increased by thermal treatment in reducing the gas atmosphere [64]. In several works, the transition between the insulator state and metallic state of Ti3O5 (β and λ forms of Ti3O5, respectively) was performed by pulses of visible light [41] and by thermal treatment, which induced the conversion of α form into β form at 450 K [38] and the conversion of δ form into γ form at 240 K [39,40,41,42]. The phase transition of Ti3O5 at 240–450 K is the most important for the adjustment of conductivity of this semiconducting material-based layers, e.g., the switching between metal and insulator states was observed at 350 K [43].
In some cases, non-stoichiometric titanium oxide can be formed by the partial oxidation of metallic titanium layers, which is followed by thermal treatment and annealing [1]. Sensors based on such structures, which are differently thermally treated and annealed, can be used in the formation of sensors with very different selectivity and sensitivity, which are suitable for the development of sensor arrays. Non-stoichiometric titanium oxide-based sensors. despite some significant advantages related to better catalytic activity and conductivity, have some disadvantages in comparison to those based on stoichiometric titanium oxides related to their insufficient stability at ambient conditions. In addition, it should be noted that the selectivity of these sensors is not superior.

3. Pristine Titanium Oxide-Based Gas Sensors and Their Sensing Mechanisms

Stoichiometric TiO2-based gas sensors show high sensitivity to different gases. It should be noted that TiO2-based gas sensors can rely on several different sensing mechanisms, which differ the most significantly for the determination of reducing gaseous compounds, such as H2, H2S, NH3, CO, CH3OH, C2H5OH, etc. and for oxidizing gaseous compounds, such as O2, NO2, CO2 [28,63,65] (Table 1). The changes in electrical resistivity of the TiO2-based layer are mostly used for the determination of analytical signals; therefore, the assessment of analytical signals generated by such sensors is rather simple. In some cases, measurement protocols were advanced by the determination of photoluminescence signals [1,2,8], which are generated by semiconducting TiO2 structures [66]. Some above-mentioned photocatalytic and photovoltaic properties can be improved by laser-based treatment [67]. However, the main disadvantage of TiO2-based sensors is poor selectivity towards gaseous materials, which significantly complicates the application of these analytical devices. Therefore, in order to improve selectivity, various heterostructures containing TiO2 hybridized with many other semiconductors have been developed, e.g., our research group has developed a TiO2−x/TiO2-based self-heating heterostructure for the determination of NH3, CH3OH and C2H5OH [1].
It was also reported that TiO2 combined with La0.8Sr0.2Co0.5Ni0.5O3 perovskite was applied for the determination of CO [77], TiO2/V2O5-based heterostructure for ozone [78], TiO2/SnO2-based heterostructure for NO2 and CO [79,80]; TiO2/graphene (TiO2/GO)-based heterostructure for toluene determination at 298 K [81]. The action of TiO2/GO-based heterostructures based on the formed n-n hetero-junction is enhanced by UV irradiation [82] and is addressed in the band diagram presented in Figure 2a,b. The bandgap of the hetero-junction in the TiO2/GO-based heterostructure is 4.7 eV [83] which is higher than that of 4.4 eV and is usually observed for GO-based structures [84]; the formed hetero-structure has both accumulation and depletion layers [85] (Figure 2b). In this structure, a Schottky barrier is formed; therefore, the depletion layer in TiO2 is thicker, and the number of electrons in the GO interface increases. The interaction of the adsorbed gases with the oxidized functional groups of GO induces variation in the resistance of the TiO2/GO-based heterostructure, which forms a sensing layer. In this heterostructure, TiO2 is contributing by increasing the number of active sites suitable for the adsorption of gas molecules [86,87]. In some particular cases, activation energy can be lowered by adsorbed gas molecules, which increases the catalytic activity of the TiO2 layer [88]. This property increases the sensitivity towards some reducing gases, such as ammonia, etc.; however, the active sites of GO are very sensitive towards moisture that increases electrical resistance, due to the interaction of water molecules with oxygen-based (carboxyl, carbonyl and hydroxyl) functional groups [89]. UV irradiation enables to solve this disadvantage by the increase in the depletion layer in TiO2 and enhancement of the accumulation layer in GO (Figure 2c). Due to this reason, the electrons are shifted towards GO and reduced carboxyl, carbonyl and hydroxyl groups, which become unsuitable for the adsorption of water molecules [90]. This reduced GO establishes a p-n junction with TiO2, which at the interface is characterized by decreased junction width (Figure 2d). UV irradiation of TiO2 induces the ‘injection’ of electrons into GO and in such a way that electron–hole pair recombination is prevented [91]. The surface concentration of adsorbed oxygen molecules decreases, which also reduces water sorption to this heterostructure [92]. The number of carboxyl, carbonyl and hydroxyl groups in GO structure can be partly restored by the UV irradiation-induced action of TiO2 [82].
The heterostructure based on TiO2/SnO2 is also very interesting because the depletion of energetic layers in the molecular orbitals of TiO2 electrons is induced when TiO2 is connected to SnO2 [79]. It should be noted that nanoparticle-based structures are preferable for the development of gas sensors [23] because the radius of nanoparticles based on some semiconducting materials that are used in the design of gas-sensors are in the same range as the Exciton Bohr radius; for this reason, such particles are very suitable for the design of gas sensors [25].
SnO2 has great charge-carrier mobility, which is the most important factor for gas sensors based on resistivity measurements [80,93]. SnO2-based gas-sensing layers are cheap and stable at ambient conditions [94] however, it should be taken into account that sensors based on stoichiometric TiO2 and SnO2 operate at a rather high temperature, which exceeds several hundred degrees [25]. However, such sensors consume a lot of energy for the heating of the sensing layer; therefore, due to energy saving related issues, sensing layers capable to operate at low temperatures are under very special interest, e.g., sensing layers based on Au/SnO2 core-shell structures can operate in the temperature interval of 25–80 °C [95]. Our investigations of non-stoichiometric titanium oxide-based sensors shows that these sensors can operate even at room temperature [1,2]. It is expected that heterostructures based on SnO2/TiO2 can be applied in low temperature gas sensors; therefore, some nanostructures were designed, which are as follows: TiO2 nanobelts covered by SnO2 [96], TiO2/SnO2-based core-shell nanofibers [97], SnO2 layers doped by TiO2 [98], structures based on SnO2/TiO2 laminates [99], SnO2 nanoparticles wrapped within TiO2-based nanofibers [79], SnO2-based quantum dots formed on a surface of TiO2 layer were reported to be sensitive towards NO2 and CO [100]. Atomic layer deposition (ALD) was used to deposit thin TiO2 layers and it was demonstrated that tuning of the selectivity can be performed by the adjustment of TiO2 layer thickness [101].
Significant energy consumptions for the heating of the sensing layer reduce the applicability of most gas sensors. Therefore, there is a demand for gas-sensing structures that can operate at low temperatures [102]. The miniaturization of sensing elements is another suitable strategy to reduce energy consumption. Low-temperature sensors based on titanium oxide-based layers were reported for O2 [99], ozone [78], formaldehyde [103,104], CO [77], ethanol [96,98,105], C7H8 [81], H2 [106,107] and other gases [108].
The ‘self-heating’ of the sensing layer can be achieved when it has rather low resistance and part of electrical energy, which passes through this layer and is converted into thermal energy [1,2]. However, stoichiometric titanium oxide-based layers are characterised by a high band gap; therefore, the conductivity of these layers is not sufficient for ‘self-heating’ because rather high voltages are required to achieve some effect. On the contrary, this operation mode is very suitable for non-stoichiometric titanium oxide-based sensors, because these layers are good at conducting at low temperatures and in the temperature interval of 72–180 °C, these layers reach very good sensitivity towards some gases [1,2]. The semiconducting properties and chemical activity of titanium oxide enable sufficient catalytic and photocatalytic activities under UV irradiation to turn into a ‘water splitting’ ability [67,109], which eventually can be exploited for sensing purposes. It should be noted that at room temperature (298 K), sensing layers are sensitive to humidity. Therefore, analytical signals generated by adsorbed water molecules interfere with an analytical signal generated by target gases.

4. The Action of Sensing Layers Based on TiO2 Heterostructures and Assessment of Analytical Signals by Titanium Oxide-Based Sensors

The action of TiO2-based heterostructures relies on many parallel processes, where the most important are the adsorption of gas molecules to the surface of the sensing layer and the simultaneous desorption of gas, which was initially adsorbed to the same layer. Both adsorption and desorption events and/or the establishment of new chemical bonds between sensing TiO2 layer and adsorbed gas molecules electrostatically affect the upper layer of the semiconducting TiO2 layer and due to the depletion and/or enrichment of this layer by charge carriers, the conductivity of this layer changes correspondingly. The typical analytical signal determined by the adsorption and desorption of analyte gas is represented in Figure 3.
Several different types of electrical conductivity are observed in the TiO2-layers based on nano- or micro-particles, which are as follows: (i) intrinsic charge transfer through TiO2-particles and (ii) limitations of charge transfer through the boundaries between different particles. For this reason, the volume-concentration of these boundaries and the size/shape of particles are critical in the development of such sensors [110,111] (Figure 4). The semiconducting characteristics of TiO2-based structures, such as Debay length/radius, also strongly influence charge transfer in TiO2-based sensors. It should be noted that due to numerous thermal phase transitions during the annealing-based development of non-stoichiometric titanium oxide-based sensing structures, they are formed very porous and contain many different nanostructures and phases that significantly enhance the surface/volume ratio. Therefore, such structures are very efficient at absorbing gases, which leads to enhanced sensitivity of non-stoichiometric titanium oxide-based sensors. In order to achieve great sensitivity and selectivity, the discussed properties of non-stoichiometric titanium oxide-based sensing layers should be well tailored during the design of gas sensors [63,112,113,114,115,116].
It should be noted that the adsorption of gas molecules with the surface of TiO2-based layer is a very complex process, which at a high extent is determined by Van der Waals and/or electrostatic interactions and/or the forming of new chemical bonds. Molecules of the gas that are present in the measurement cell before the introduction of analyte-gas containing aliquot are critically important for the establishment of the particular mechanism of sensing, because they are preabsorbing on the semiconducting structure and, in such a way, are responsible for the initial conductivity of the sensing layer. During the measurement, analyte gas containing aliquot appears in senor-cell molecules of preadsorbed gas, which are replaced by analyte gas molecules (and/or by other gas molecules, which are part of aliquot). During both stages, adsorbed gas molecules are electrostatically interacting with the adsorption sites. It should be noted that in TiO2, defects of the crystal structure are acting as the most efficient adsorption sites. The adsorption ability of gas molecules very often depends on the ability to accept and/or to donate electrons to active sites that can be additionally involved in photoluminescence-based light emission [35]. Hence, donor/acceptor interactions change the electrical conductivity and the efficiency of photoluminescence of TiO2-based structures that can be used for the assessment of analyte concentration. However, stoichiometric titanium oxide-based structures are characterized by high resistance, which does not significantly vary in the presence of low concentrations of gaseous materials; therefore, the dynamic range of sensors based on stoichiometric titanium oxide is not high. Conducting polymers, especially polyaniline (PANI), are well suitable for the enhancement of the sensitivity of semiconducting metal oxides towards gases [117,118,119] and volatile compounds [120].
In order to increase sensitivity, various heterostructures containing TiO2 are modified by other semiconducting materials, e.g., by conjugated polymers polyaniline (TiO2/PANI) and polypyrrole (TiO2/Ppy)), which can be involved in the modification and design of titanium oxide-based sensing structures. The formation of the n/p-junction at TiO2/(conducting polymer) interphase is the most important factor for the enhancement of sensor sensitivity [121]. TiO2/Ppy-based sensing structures for the determination of NH3 [122,123] propane/butane [124] and TiO2/PANI-based sensing structures for the detection of NH3 [123,125,126,127] were developed. In order to improve sensitivity and selectivity, titanium oxide-based semiconducting structures can be modified by some metal oxides, carbides and other materials [1,2,9,113,114,128,129,130,131].
During the development of titanium oxide-based sensors, a very important issue is the elaboration of a suitable analytical signal registration protocol. The registration can be based on many physical approaches, including the measurement of photoluminescence response and electrical signals. Electrical signals can be determined from measurements performed in potentiostatic, galvanostatic and various potentiodynamic modes. If necessary, these modes can be combined with a ‘self-heating’ based approach. The determination of electrical conductivity measurements is most frequently used in gas sensors based on semiconducting structures [1,4,5,7,11,20,22,23,24,25,33,34,65,98,104,107,112,122,125,131,132]. At room temperature, stoichiometric titanium oxide-based sensing layers have a rather low conductivity of ~10−10 S/m; for this reason, up to 200–400 °C elevated temperature is applied in order to reach optimal electrical conductivity and sensing conditions. However, elevated temperatures increase the consumption of electrical energy required for the operation of the sensor.

5. TiO2 and Conducting Polymer-Based Composites for Gas and VOC Sensors

The combination of metal oxide-based structures with conducting polymers led to the ability to tune precisely some properties of formed composite material in order to adjust it for specific applications [120]. Therefore, during the past decade, a lot of new strategies for the synthesis of different morphology TiO2 (Figure 5) and their composites with conducting polymers were developed.
Chemical formation is one of the easiest and the most popular methods for the formation of TiO2-PANI composites. Usually, for chemical formation, acidic solutions (mostly HCl [137,138,139] or H2SO4 [140,141]) are used together with reaction initiator (NH4)2S2O8 [137,138,139,140,142]. Procedure protocols in most cases are similar and involve a few steps, which are as follows: (i) aniline and acid solution in distilled water together with TiO2 is prepared in one beaker, (ii) in a second beaker, the polymerization initiator is dissolved in an acidic solution and (iii) both solutions are mixed, cooled to a near 0 °C temperature and stored for some time [137,138,143]. A visible sign of aniline polymerization reaction is a change of solution color into light blue, revealing the starting of PANI formation through an oxidation reaction. Finally, the reaction solution color becomes green, which indicates the successful formation of the PANI composite [139]. While varying in the parameters of reagent concentrations, it is possible that polymeric composite thickness, surface morphology and conductivity can be controlled by polymerization time and temperature depending on technological demand. It is important to mention that during the oxidation process, conductivity mainly depends on anions and their concentration used in the acid solution during the PANI polymerization reaction [137,138,139,140,143,144,145]. While wet chemical synthesis methods are the easiest to perform, electrochemical polymerization is one of the common methods, which provides relatively good control of the process in comparison to chemical formation. In this method, it is possible to use various modes for electrochemical PANI deposition on top of titanium dioxide substrate; however, the most popular modes are cyclic voltammetry (CV) [146,147,148,149,150] and chronoamperometry (CA) [146,150]. For CA depositions, a WE electrode potential in between 0.85–0.9 V vs. Ag/AgCl [146], while in CV mode, a potential in between −0.6–1.3 V vs. Ag/AgCl [146,147,148,149,151] and an acidic electrolyte with dissolved aniline are usually used [150,152].
Particularly, most researchers report that the modification of TiO2-based structures by conducting the polymer polyaniline (PANI) can increase their sensitivity towards some gases and/or VOC’s [153,154]. An especially high number of researchers report that the modification of TiO2-based structures by PANI increases the sensitivity of designed sensors towards ammonia (NH3) [125,126,143,155,156,157,158,159,160,161,162]. The action mechanism of the sensing layer is rather simple; the conductivity sensing layer usually increases while oxygen, which is acting as an oxidizing agent, is physico-sorbed on TiO2/PANI and decreases when NH3, which is acting as a reducing agent, replaces the adsorbed oxygen [155]. Although all common PANI forms (including leucoemeraldine and pernigraniline) [163,164,165] can be used for the modification of TiO2-based sensing layers, the most applicable seems to be an emeraldine form of PANI [154,157,159]. In addition to the most common TiO2-forms (anatase and rutile), recently a new trend is based on the application of various nonstoichiometric titanium suboxides, which are modified with conducting polymers including PANI [162,165], polypyrrole (Ppy) [123,124,166], polythiophene (PTH) [65], etc. Modifications by PANI are especially effective because PANI increases the sensitivity of titanium oxides towards NH3 [125,126,143,155,156,157,158,159,160,161,162], H2 [163,167], LPG [168,169] [170], CO2 [169], trimethylamine (TMA) [160], H2S [160], benzene [138], toluene [169], cyclohexane [169], CO [165] and ethanol [164]. Analytical signals in most (titanium oxide)/PANI-based sensors are registered by the measurement of sensing layer resistance. However, there are some reports on the application of quartz crystal microbalances (QCM) in (titanium oxide)/PANI-based sensor design. Such TiO2-PANI composite-based QCM sensors were used for the detection of trimethylamine [170], hydrazine [148] and NH3 [157].

6. Conclusions and Future Trends

A number of stoichiometric titanium oxide-based structures were applied in the design of gas sensors. However, these sensors still face selectivity-related problems, which can be solved by the doping of titanium oxide-based layers by dopants or the formation of hetero-structures based on a combination with some other semiconducting compounds and the adjustment of sensor-operation temperature. Among many others, sensors based on composite materials that are based on titanium oxide modified with conducting polymers are very promising, e.g., PANI. In recent investigations, it was shown that TiO2/TiO2−x-based heterostructures can be successfully applied in the design of gas sensors; it was demonstrated that the sensing properties of these sensors can be easily adjusted by changing the TiO2/TiO2−x-ratio heterostructures. Moreover, TiO2/TiO2−x-based heterostructures are good conductors; therefore, they can act in ‘self-heating’ mode and can reach temperatures that are suitable for the determination of some gaseous compounds. It is expected that TiO2/TiO2−x-based heterostructures will be more frequently applied in the design of gas sensors, but the main challenge in this research direction is still related to the control of stoichiometry and morphology of the TiO2/TiO2−x-based structure, which is very critical for the sensitivity and selectivity of the designed gas sensors.

Author Contributions

S.R. performed literature research, analysis, and drafted the paper. A.J. performed literature research, analysis, and drafted the paper. A.R. initiated and supervised the work and provided insights. All authors have read and agreed to the published version of the manuscript.

Funding

EU Horizon 2020 research and innovation programme H2020-MSCA-RISE under grant agreement № 778157 CanBioSe.

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. Ramanavicius, S.; Tereshchenko, A.; Karpicz, R.; Ratautaite, V.; Bubniene, U.; Maneikis, A.; Jagminas, A.; Ramanavicius, A. TiO2-x/TiO2-Structure Based ‘Self-Heated’ Sensor for the Determination of Some Reducing Gases. Sensors 2020, 20, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ramanavicius, S.; Ramanavicius, A. Insights in the Application of Stoichiometric and Non-Stoichiometric Titanium Oxides for the Design of Sensors for the Determination of Gases and VOCs (TiO2−x and TinO2n−1 vs. TiO2). Sensors 2020, 20, 6833. [Google Scholar] [CrossRef] [PubMed]
  3. Ramanavičius, S.; Petrulevičienė, M.; Juodkazytė, J.; Grigucevičienė, A.; Ramanavičius, A. Selectivity of Tungsten Oxide Synthesized by Sol-Gel Method Towards Some Volatile Organic Compounds and Gaseous Materials in a Broad Range of Temperatures. Materials 2020, 13, 523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. 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] [PubMed]
  5. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of Hazardous Volatile Organic Compounds (VOCs) by Metal Oxide Nanostructures-Based Gas Sensors: A Review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  6. Mirzaei, A.; Kim, J.H.; Kim, H.W.; Kim, S.S. How Shell Thickness Can Affect the Gas Sensing Properties of Nanostructured Materials: Survey of Literature. Sens. Actuators B Chem. 2018, 258, 270–294. [Google Scholar] [CrossRef]
  7. Mirzaei, A.; Kim, J.-H.; Kim, H.W.; Kim, S.S. Resistive-Based Gas Sensors for Detection of Benzene, Toluene and Xylene (BTX) Gases: A Review. J. Mater. Chem. C 2018, 6, 4342–4370. [Google Scholar] [CrossRef]
  8. Petruleviciene, M.; Juodkazyte, J.; Parvin, M.; Tereshchenko, A.; Ramanavicius, S.; Karpicz, R.; Samukaite-Bubniene, U.; Ramanavicius, A. Tuning the Photo-Luminescence Properties of WO3 Layers by the Adjustment of Layer Formation Conditions. Materials 2020, 13, 2814. [Google Scholar] [CrossRef]
  9. Ramanavicius, S.; Ramanavicius, A. Progress and Insights in the Application of MXenes as New 2D Nano-Materials Suitable for Biosensors and Biofuel Cell Design. Int. J. Mol. Sci. 2020, 21, 9224. [Google Scholar] [CrossRef]
  10. Wang, Y.; Wu, T.; Zhou, Y.; Meng, C.; Zhu, W.; Liu, L. TiO2-Based Nanoheterostructures for Promoting Gas Sensitivity Performance: Designs, Developments, and Prospects. Sensors 2017, 17, 1971. [Google Scholar] [CrossRef]
  11. Bai, J.; Zhou, B. Titanium Dioxide Nanomaterials for Sensor Applications. Chem. Rev. 2014, 114, 10131–10176. [Google Scholar] [CrossRef] [PubMed]
  12. Mokrushin, A.S.; Gorban, Y.M.; Simonenko, N.P.; Simonenko, T.L.; Simonenko, E.P.; Sevastyanov, V.G.; Kuznetsov, N.T. Synthesis and Gas-Sensitive Chemoresistive Properties of TiO2:Cu Nanocomposite. Russ. J. Inorg. Chem. 2021, 66, 594–602. [Google Scholar] [CrossRef]
  13. Mokrushin, A.S.; Simonenko, E.P.; Simonenko, N.P.; Akkuleva, K.T.; Antipov, V.V.; Zaharova, N.V.; Malygin, A.A.; Bukunov, K.A.; Sevastyanov, V.G.; Kuznetsov, N.T. Oxygen Detection Using Nanostructured TiO2 Thin Films Obtained by the Molecular Layering Method. Appl. Surf. Sci. 2019, 463, 197–202. [Google Scholar] [CrossRef]
  14. Sevastyanov, V.G.; Simonenko, E.P.; Simonenko, N.P.; Mokrushin, A.S.; Nikolaev, V.A.; Kuznetsov, N.T. Sol-Gel Made Titanium Dioxide Nanostructured Thin Films as Gas-Sensing Materials for the Detection of Oxygen. Mendeleev Commun. 2018, 28, 164–166. [Google Scholar] [CrossRef]
  15. Mokrushin, A.S.; Simonenko, E.P.; Simonenko, N.P.; Bukunov, K.A.; Gorobtsov, P.Y.; Sevastyanov, V.G.; Kuznetsov, N.T. Gas-Sensing Properties of Nanostructured TiO2–XZrO2 Thin Films Obtained by the Sol–Gel Method. J. Sol-Gel Sci. Technol. 2019, 92, 415–426. [Google Scholar] [CrossRef]
  16. Simonenko, E.P.; Simonenko, N.P.; Kopitsa, G.P.; Mokrushin, A.S.; Khamova, T.V.; Sizova, S.V.; Khaddazh, M.; Tsvigun, N.V.; Pipich, V.; Gorshkova, Y.E.; et al. A Sol-Gel Synthesis and Gas-Sensing Properties of Finely Dispersed ZrTiO4. Mater. Chem. Phys. 2019, 225, 347–357. [Google Scholar] [CrossRef]
  17. Simonenko, E.P.; Mokrushin, A.S.; Simonenko, N.P.; Voronov, V.A.; Kim, V.P.; Tkachev, S.V.; Gubin, S.P.; Sevastyanov, V.G.; Kuznetsov, N.T. Ink-Jet Printing of a TiO2–10%ZrO2 Thin Film for Oxygen Detection Using a Solution of Metal Alkoxoacetylacetonates. Thin Solid Film. 2019, 670, 46–53. [Google Scholar] [CrossRef]
  18. Prades, J.D.; Jimenez-Diaz, R.; Hernandez-Ramirez, F.; Barth, S.; Cirera, A.; Romano-Rodriguez, A.; Mathur, S.; Morante, J.R. Ultralow Power Consumption Gas Sensors Based on Self-Heated Individual Nanowires. Appl. Phys. Lett. 2008, 93, 123110. [Google Scholar] [CrossRef]
  19. Smulko, J.M.; Trawka, M.; Granqvist, C.G.; Ionescu, R.; Annanouch, F.; Llobet, E.; Kish, L.B. New Approaches for Improving Selectivity and Sensitivity of Resistive Gas Sensors: A Review. Sens. Rev. 2015, 35, 340–347. [Google Scholar] [CrossRef]
  20. Simon, I.; Bârsan, N.; Bauer, M.; Weimar, U. Micromachined Metal Oxide Gas Sensors: Opportunities to Improve Sensor Performance. Sens. Actuators B Chem. 2001, 73, 1–26. [Google Scholar] [CrossRef]
  21. Karnati, P.; Akbar, S.; Morris, P.A. Conduction Mechanisms in One Dimensional Core-Shell Nanostructures for Gas Sensing: A Review. Sens. Actuators B Chem. 2019, 295, 127–143. [Google Scholar] [CrossRef]
  22. Kim, J.H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Low Power-Consumption CO Gas Sensors Based on Au-Functionalized SnO2-ZnO Core-Shell Nanowires. Sens. Actuators B Chem. 2018, 267, 597–607. [Google Scholar] [CrossRef]
  23. Mirzaei, A.; Janghorban, K.; Hashemi, B.; Bonavita, A.; Bonyani, M.; Leonardi, S.G.; Neri, G. Synthesis, Characterization and Gas Sensing Properties of Ag@α-Fe2O3 Core–Shell Nanocomposites. Nanomaterials 2015, 5, 737–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Deng, J.; Fu, Q.; Luo, W.; Tong, X.; Xiong, J.; Hu, Y.; Zheng, Z. Enhanced H2S Gas Sensing Properties of Undoped ZnO Nanocrystalline Films from QDs by Low-Temperature Processing. Sens. Actuators B Chem. 2016, 224, 153–158. [Google Scholar] [CrossRef]
  25. Mosadegh Sedghi, S.; Mortazavi, Y.; Khodadadi, A. Low Temperature CO and CH4 Dual Selective Gas Sensor Using SnO2 Quantum Dots Prepared by Sonochemical Method. Sens. Actuators B Chem. 2010, 145, 7–12. [Google Scholar] [CrossRef]
  26. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Jiang, Z.; Huang, J.; Lim, L.Y.; Li, W.; Deng, J.; Gong, D.; Tang, Y.; Lai, Y.; Chen, Z. Titanate and Titania Nanostructured Materials for Environmental and Energy Applications: A Review. RSC Adv. 2015, 5, 79479–79510. [Google Scholar] [CrossRef]
  28. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  29. Soni, P.; Murty, V.V.S.; Kushwaha, K.K. The Effect of Ni2+ Ions on Energy Band Gap of TiO2 Nanoparticles for Solar Cell Applications. J. Nanosci. Nanoeng. Appl 2018, 8, 69–74. [Google Scholar]
  30. Yamazoe, N.; Sakai, G.; Shimanoe, K. Oxide Semiconductor Gas Sensors. Catal. Surv. Asia 2003, 7, 63–75. [Google Scholar] [CrossRef]
  31. Wang, G.; Wang, J.; An, Y.; Wang, C. Anodization Fabrication of 3D TiO2 Photonic Crystals and Their Application for Chemical Sensors. Superlattices Microstruct. 2016, 100, 290–295. [Google Scholar] [CrossRef]
  32. Si, H.; Pan, N.; Zhang, X.; Liao, J.; Rumyantseva, M.N.; Gaskov, A.M.; Lin, S. A Real-Time on-Line Photoelectrochemical Sensor toward Chemical Oxygen Demand Determination Based on Field-Effect Transistor Using an Extended Gate with 3D TiO2 Nanotube Arrays. Sens. Actuators B Chem. 2019, 289, 106–113. [Google Scholar] [CrossRef]
  33. Qiu, J.; Zhang, S.; Zhao, H. Recent Applications of TiO2 Nanomaterials in Chemical Sensing in Aqueous Media. Sens. Actuators B: Chem. 2011, 160, 875–890. [Google Scholar] [CrossRef]
  34. Maziarz, W.; Kusior, A.; Trenczek-Zajac, A. Nanostructured TiO2-Based Gas Sensors with Enhanced Sensitivity to Reducing Gases. Beilstein J. Nanotechnol. 2016, 7, 1718–1726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Tereshchenko, A.; Smyntyna, V.; Ramanavicius, A. Interaction Mechanism between TiO2 Nanostructures and Bovine Leukemia Virus Proteins in Photoluminescence-Based Immunosensors. RSC Adv. 2018, 8, 37740–37748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Tereshchenko, A.; Viter, R.; Konup, I.; Ivanitsa, V.; Geveliuk, S.; Ishkov, Y.; Smyntyna, V. TiO2 Optical Sensor for Amino Acid Detection. In Proceedings of the Biophotonics—Riga 2013, Riga, Latvia, 26–31 August 2013; Spigulis, J., Kuzmina, I., Eds.; SPIE: Washington, DC, USA, 2013; Volume 9032, pp. 186–190. [Google Scholar]
  37. Wunderlich, W.; Oekermann, T.; Miao, L.; Hue, N.T.; Tanemura, S.; Tanemura, M. ELECTRONIC PROPERTIES OF NANO-POROUS TiO 2- AND ZnO THIN FILMS- COMPARISON OF SIMULATIONS AND EXPERIMENTS. J. Ceram. Process. Res. 2004, 5, 343–354. [Google Scholar]
  38. Åsbrink, S.; Magnéli, A. Crystal Structure Studies on Trititanium Pentoxide, Ti3O5. Acta Crystallogr. 1959, 12, 575–581. [Google Scholar] [CrossRef] [Green Version]
  39. Hong, S.-H.; Åsbrink, S. The Structure of γ-Ti3O5 at 297 K. Acta Crystallogr. Sect. B 1982, 38, 2570–2576. [Google Scholar] [CrossRef]
  40. Onoda, M. Phase Transitions of Ti3O5. J. Solid State Chem. 1998, 136, 67–73. [Google Scholar] [CrossRef]
  41. Ohkoshi, S.; Tsunobuchi, Y.; Matsuda, T.; Hashimoto, K.; Namai, A.; Hakoe, F.; Tokoro, H. Synthesis of a Metal Oxide with a Room-Temperature Photoreversible Phase Transition. Nat. Chem. 2010, 2, 539–545. [Google Scholar] [CrossRef]
  42. Tanaka, K.; Nasu, T.; Miyamoto, Y.; Ozaki, N.; Tanaka, S.; Nagata, T.; Hakoe, F.; Yoshikiyo, M.; Nakagawa, K.; Umeta, Y.; et al. Structural Phase Transition between γ-Ti3O5 and δ-Ti3O5 by Breaking of a One-Dimensionally Conducting Pathway. Cryst. Growth Des. 2015, 15, 653–657. [Google Scholar] [CrossRef]
  43. Yoshimatsu, K.; Sakata, O.; Ohtomo, A. Superconductivity in Ti4O7 and γ-Ti3O5 Films. Sci. Rep. 2017, 7, 12544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Marezio, M.; McWhan, D.B.; Dernier, P.D.; Remeika, J.P. Structural Aspects of the Metal-Insulator Transitions in Ti4O7. J. Solid State Chem. 1973, 6, 213–221. [Google Scholar] [CrossRef]
  45. Lakkis, S.; Schlenker, C.; Chakraverty, B.K.; Buder, R.; Marezio, M. Metal-Insulator Transitions in Ti4O7 Single Crystals: Crystal Characterization, Specific Heat, and Electron Paramagnetic Resonance. Phys. Rev. B 1976, 14, 1429–1440. [Google Scholar] [CrossRef]
  46. D’Angelo, A.M.; Webster, N.A.S. Evidence of Anatase Intergrowths Formed during Slow Cooling of Reduced Ilmenite. J. Appl. Crystallogr. 2018, 51, 185–192. [Google Scholar] [CrossRef] [Green Version]
  47. Grey, I.E.; Cranswick, L.M.D.; Li, C.; White, T.J.; Bursill, L.A. New M3O5-Anatase Intergrowth Structures Formed during Low-Temperature Oxidation of Anosovite. J. Solid State Chem. 2000, 150, 128–138. [Google Scholar] [CrossRef]
  48. Jayashree, S.; Ashokkumar, M. Switchable Intrinsic Defect Chemistry of Titania for Catalytic Applications. Catalysts 2018, 8, 601. [Google Scholar] [CrossRef] [Green Version]
  49. Andersson, S.; Magnéli, A. Diskrete Titanoxydphasen Im Zusammensetzungsbereich TiO_1,75-TiO_1,90. Naturwissenschaften 1956, 43, 495–496. [Google Scholar] [CrossRef]
  50. Liborio, L.; Mallia, G.; Harrison, N. Electronic Structure of the Ti4O7 Magneli Phase. Phys. Rev. B 2009, 79, 245133. [Google Scholar] [CrossRef] [Green Version]
  51. Liborio, L.; Harrison, N. Thermodynamics of Oxygen Defective Magn\’eli Phases in Rutile: A First-Principles Study. Phys. Rev. B 2008, 77, 104104. [Google Scholar] [CrossRef] [Green Version]
  52. Adamaki, V.; Clemens, F.; Ragulis, P.; Pennock, S.R.; Taylor, J.; Bowen, C.R. Manufacturing and Characterization of Magnéli Phase Conductive Fibres. J. Mater. Chem. A 2014, 2, 8328–8333. [Google Scholar] [CrossRef] [Green Version]
  53. Zhu, Q.; Peng, Y.; Lin, L.; Fan, C.-M.; Gao, G.-Q.; Wang, R.-X.; Xu, A.-W. Stable Blue TiO2−x Nanoparticles for Efficient Visible Light Photocatalysts. J. Mater. Chem. A 2014, 2, 4429–4437. [Google Scholar] [CrossRef]
  54. Al-Hashem, M.; Akbar, S.; Morris, P. Role of Oxygen Vacancies in Nanostructured Metal-Oxide Gas Sensors: A Review. Sens. Actuators B Chem. 2019, 301, 126845. [Google Scholar] [CrossRef]
  55. Seebauer, E.G.; Kratzer, M.C. Charged Point Defects in Semiconductors. Mater. Sci. Eng. R Rep. 2006, 55, 57–149. [Google Scholar] [CrossRef]
  56. Harada, S.; Tanaka, K.; Inui, H. Thermoelectric Properties and Crystallographic Shear Structures in Titanium Oxides of the Magnèli Phases Bandgap Engineering of Magnéli Phase TinO2n−1: Electron-Hole Self-Compensation. J. Appl. Phys. 2010, 108, 83703. [Google Scholar] [CrossRef] [Green Version]
  57. Smith, J.R.; Walsh, F.C.; Clarke, R.L. Electrodes Based on Magnéli Phase Titanium Oxides: The Properties and Applications of Ebonex® Materials. J. Appl. Electrochem. 1998, 28, 1021–1033. [Google Scholar] [CrossRef]
  58. Walsh, F.C.; Wills, R.G.A. The Continuing Development of Magnéli Phase Titanium Sub-Oxides and Ebonex® Electrodes. Electrochim. Acta 2010, 55, 6342–6351. [Google Scholar] [CrossRef]
  59. Nakamura, I.; Negishi, N.; Kutsuna, S.; Ihara, T.; Sugihara, S.; Takeuchi, K. Role of Oxygen Vacancy in the Plasma-Treated TiO2 Photocatalyst with Visible Light Activity for NO Removal. J. Mol. Catal. A Chem. 2000, 161, 205–212. [Google Scholar] [CrossRef]
  60. le Mercier, T.; Mariot, J.M.; Parent, P.; Fontaine, M.F.; Hague, C.F.; Quarton, M. Formation of Ti3+ Ions at the Surface of Laser-Irradiated Rutile. Appl. Surf. Sci. 1995, 86, 382–386. [Google Scholar] [CrossRef]
  61. Zheng, Z.; Huang, B.; Meng, X.; Wang, J.; Wang, S.; Lou, Z.; Wang, Z.; Qin, X.; Zhang, X.; Dai, Y. Metallic Zinc- Assisted Synthesis of Ti3+ Self-Doped TiO2 with Tunable Phase Composition and Visible-Light Photocatalytic Activity. Chem. Commun. 2013, 49, 868–870. [Google Scholar] [CrossRef]
  62. Hashimoto, S.; Tanaka, A. Alteration of Ti 2p XPS Spectrum for Titanium Oxide by Low-Energy Ar Ion Bombardment. Surf. Interface Anal. 2002, 34, 262–265. [Google Scholar] [CrossRef]
  63. Wang, Y.; Du, G.; Liu, H.; Liu, D.; Qin, S.; Wang, N.; Hu, C.; Tao, X.; Jiao, J.; Wang, J.; et al. Nanostructured Sheets of Ti—O Nanobelts for Gas Sensing and Antibacterial Applications. Adv. Funct. Mater. 2008, 18, 1131–1137. [Google Scholar] [CrossRef]
  64. Hayfield, P.C.S. Development of a New Material: Monolithic Ti4O7 Ebonex Ceramic; Royal Society of Chemistry: London, UK, 2007; ISBN 184755069X. [Google Scholar]
  65. Kimura, M.; Sakai, R.; Sato, S.; Fukawa, T.; Ikehara, T.; Maeda, R.; Mihara, T. Sensing of Vaporous Organic Compounds by TiO2 Porous Films Covered with Polythiophene Layers. Adv. Funct. Mater. 2012, 22, 469–476. [Google Scholar] [CrossRef]
  66. Viter, R.; Tereshchenko, A.; Smyntyna, V.; Ogorodniichuk, J.; Starodub, N.; Yakimova, R.; Khranovskyy, V.; Ramanavicius, A. Toward Development of Optical Biosensors Based on Photoluminescence of TiO2 Nanoparticles for the Detection of Salmonella. Sens. Actuators B Chem. 2017, 252, 95–102. [Google Scholar] [CrossRef] [Green Version]
  67. Haryński, Ł.; Grochowska, K.; Karczewski, J.; Ryl, J.; Siuzdak, K. Scalable Route toward Superior Photoresponse of UV-Laser-Treated TiO2 Nanotubes. ACS Appl. Mater. Interfaces 2020, 12, 3225–3235. [Google Scholar] [CrossRef]
  68. Gardon, M.; Monereo, O.; Dosta, S.; Vescio, G.; Cirera, A.; Guilemany, J.M. New Procedures for Building-up the Active Layer of Gas Sensors on Flexible Polymers. Surf. Coat. Technol. 2013, 235, 848–852. [Google Scholar] [CrossRef]
  69. Imawan, C.; Solzbacher, F.; Steffes, H.; Obermeier, E. TiOx-Modified NiO Thin Films for H2 Gas Sensors: Effects of TiOx-Overlayer Sputtering Parameters. Sens. Actuators B Chem. 2000, 68, 184–188. [Google Scholar] [CrossRef]
  70. Li, X.; Liu, Y.; Ma, S.; Ye, J.; Zhang, X.; Wang, G.; Qiu, Y. The Synthesis and Gas Sensitivity of the β-Ti3O5 Powder: Experimental and DFT Study. J. Alloys Compd. 2015, 649, 939–948. [Google Scholar] [CrossRef]
  71. Su, J.; Zou, X.-X.; Zou, Y.-C.; Li, G.-D.; Wang, P.-P.; Chen, J.-S. Porous Titania with Heavily Self-Doped Ti3+ for Specific Sensing of CO at Room Temperature. Inorg. Chem. 2013, 52, 5924–5930. [Google Scholar] [CrossRef]
  72. Gakhar, T.; Hazra, A. Oxygen Vacancy Modulation of Titania Nanotubes by Cathodic Polarization and Chemical Reduction Routes for Efficient Detection of Volatile Organic Compounds. Nanoscale 2020, 12, 9082–9093. [Google Scholar] [CrossRef]
  73. Navale, S.T.; Yang, Z.B.; Liu, C.; Cao, P.J.; Patil, V.B.; Ramgir, N.S.; Mane, R.S.; Stadler, F.J. Enhanced Acetone Sensing Properties of Titanium Dioxide Nanoparticles with a Sub-Ppm Detection Limit. Sens. Actuators B Chem. 2018, 255, 1701–1710. [Google Scholar] [CrossRef]
  74. Gao, X.; Li, Y.; Zeng, W.; Zhang, C.; Wei, Y. Hydrothermal Synthesis of Agglomerating TiO2 Nanoflowers and Its Gas Sensing. J. Mater. Sci. Mater. Electron. 2017, 28, 18781–18786. [Google Scholar] [CrossRef]
  75. Mintcheva, N.; Srinivasan, P.; Rayappan, J.B.B.; Kuchmizhak, A.A.; Gurbatov, S.; Kulinich, S.A. Room-Temperature Gas Sensing of Laser-Modified Anatase TiO2 Decorated with Au Nanoparticles. Appl. Surf. Sci. 2020, 507, 145169. [Google Scholar] [CrossRef]
  76. Mei, H.; Zhou, S.; Lu, M.; Zhao, Y.; Cheng, L. Construction of Pine-Branch-like α-Fe2O3/TiO2 Hierarchical Heterostructure for Gas Sensing. Ceram. Int. 2020, 46, 18675–18682. [Google Scholar] [CrossRef]
  77. Hsu, K.C.; Fang, T.H.; Hsiao, Y.J.; Wu, P.C. Response and Characteristics of TiO2/Perovskite Heterojunctions for CO Gas Sensors. J. Alloys Compd. 2019, 794, 576–584. [Google Scholar] [CrossRef]
  78. Avansi, W.; Catto, A.C.; da Silva, L.F.; Fiorido, T.; Bernardini, S.; Mastelaro, V.R.; Aguir, K.; Arenal, R. One-Dimensional V2O5/TiO2 Heterostructures for Chemiresistive Ozone Sensors. ACS Appl. Nano Mater. 2019, 2, 4756–4764. [Google Scholar] [CrossRef]
  79. Chen, K.; Chen, S.; Pi, M.; Zhang, D. SnO2 Nanoparticles/TiO2 Nanofibers Heterostructures: In Situ Fabrication and Enhanced Gas Sensing Performance. Solid-State Electron. 2019, 157, 42–47. [Google Scholar] [CrossRef]
  80. Yu, Q.; Zhu, J.; Xu, Z.; Huang, X. Facile Synthesis of α-Fe2O3@SnO2 Core–Shell Heterostructure Nanotubes for High Performance Gas Sensors. Sens. Actuators B Chem. 2015, 213, 27–34. [Google Scholar] [CrossRef]
  81. Seekaew, Y.; Wisitsoraat, A.; Phokharatkul, D.; Wongchoosuk, C. Room Temperature Toluene Gas Sensor Based on TiO2 Nanoparticles Decorated 3D Graphene-Carbon Nanotube Nanostructures. Sens. Actuators B Chem. 2019, 279, 69–78. [Google Scholar] [CrossRef]
  82. Lee, E.; Lee, D.; Yoon, J.; Yin, Y.; Lee, Y.N.; Uprety, S.; Yoon, Y.S.; Kim, D.-J. Enhanced Gas-Sensing Performance of GO/TiO2 Composite by Photocatalysis. Sensors 2018, 18, 3334. [Google Scholar] [CrossRef] [Green Version]
  83. Stratakis, E.; Savva, K.; Konios, D.; Petridis, C.; Kymakis, E. Improving the Efficiency of Organic Photovoltaics by Tuning the Work Function of Graphene Oxide Hole Transporting Layers. Nanoscale 2014, 6, 6925–6931. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, C.; Cai, W.; Long, M.; Zhou, B.; Wu, Y.; Wu, D.; Feng, Y. Synthesis of Visible-Light Responsive Graphene Oxide/TiO2 Composites with p/n Heterojunction. ACS Nano 2010, 4, 6425–6432. [Google Scholar] [CrossRef] [PubMed]
  85. Lightcap, I.V.; Kosel, T.H.; Kamat, P. Anchoring Semiconductor and Metal Nanoparticles on a Two-Dimensional Catalyst Mat. Storing and Shuttling Electrons with Reduced Graphene Oxide. Nano Lett. 2010, 10, 577–583. [Google Scholar] [CrossRef] [PubMed]
  86. Ammu, S.; Dua, V.; Agnihotra, S.R.; Surwade, S.P.; Phulgirkar, A.; Patel, S.; Manohar, S.K. Flexible, All-Organic Chemiresistor for Detecting Chemically Aggressive Vapors. J. Am. Chem. Soc. 2012, 134, 4553–4556. [Google Scholar] [CrossRef] [PubMed]
  87. Lam, K.C.; Huang, B.; Shi, S.-Q. Room-Temperature Methane Gas Sensing Properties Based on in Situ Reduced Graphene Oxide Incorporated with Tin Dioxide. J. Mater. Chem. A 2017, 5, 11131–11142. [Google Scholar] [CrossRef]
  88. Ye, Z.; Tai, H.; Xie, T.; Yuan, Z.; Liu, C.; Jiang, Y. Room Temperature Formaldehyde Sensor with Enhanced Performance Based on Reduced Graphene Oxide/Titanium Dioxide. Sens. Actuators B Chem. 2016, 223, 149–156. [Google Scholar] [CrossRef]
  89. Buchsteiner, A.; Lerf, A.; Pieper, J. Water Dynamics in Graphite Oxide Investigated with Neutron Scattering. J. Phys. Chem. B 2006, 110, 22328–22338. [Google Scholar] [CrossRef]
  90. Phan, D.T.; Chung, G.S. Effects of Rapid Thermal Annealing on Humidity Sensor Based on Graphene Oxide Thin Films. Sens. Actuators B Chem. 2015, 220, 1050–1055. [Google Scholar] [CrossRef]
  91. Wang, P.; Zhai, Y.; Wang, D.; Dong, S. Synthesis of Reduced Graphene Oxide-Anatase TiO2 Nanocomposite and Its Improved Photo-Induced Charge Transfer Properties. Nanoscale 2011, 3, 1640–1645. [Google Scholar] [CrossRef]
  92. Cui, S.; Wen, Z.; Huang, X.; Chang, J.; Chen, J. Stabilizing MoS2 Nanosheets through SnO2 Nanocrystal Decoration for High-Performance Gas Sensing in Air. Small 2015, 11, 2305–2313. [Google Scholar] [CrossRef]
  93. Mirzaei, A.; Janghorban, K.; Hashemi, B.; Neri, G. Metal-Core@metal Oxide-Shell Nanomaterials for Gas-Sensing Applications: A Review. J. Nanoparticle Res. 2015, 17, 371. [Google Scholar] [CrossRef]
  94. Rieu, M.; Camara, M.; Tournier, G.; Viricelle, J.P.; Pijolat, C.; de Rooij, N.F.; Briand, D. Fully Inkjet Printed SnO2 Gas Sensor on Plastic Substrate. Sens. Actuators B Chem. 2016, 236, 1091–1097. [Google Scholar] [CrossRef] [Green Version]
  95. Chung, F.C.; Wu, R.J.; Cheng, F.C. Fabrication of a Au@SnO2 Core–Shell Structure for Gaseous Formaldehyde Sensing at Room Temperature. Sens. Actuators B Chem. 2014, 190, 1–7. [Google Scholar] [CrossRef]
  96. Chen, G.; Ji, S.; Li, H.; Kang, X.; Chang, S.; Wang, Y.; Yu, G.; Lu, J.; Claverie, J.; Sang, Y.; et al. High-Energy Faceted SnO2-Coated TiO2 Nanobelt Heterostructure for Near-Ambient Temperature-Responsive Ethanol Sensor. ACS Appl. Mater. Interfaces 2015, 7, 24950–24956. [Google Scholar] [CrossRef]
  97. Li, F.; Gao, X.; Wang, R.; Zhang, T.; Lu, G. Study on TiO2-SnO2 Core-Shell Heterostructure Nanofibers with Different Work Function and Its Application in Gas Sensor. Sens. Actuators B Chem. 2017, 248, 812–819. [Google Scholar] [CrossRef]
  98. Zeng, W.; Liu, T.; Wang, Z. UV Light Activation of TiO2 Doped SnO2 Thick Film for Sensing Ethanol at Room Temperature. Mater. Trans. 2010, 51, 243–245. [Google Scholar] [CrossRef] [Green Version]
  99. Lee, H.C.; Hwang, W.S. Substrate Effects on the Oxygen Gas Sensing Properties of SnO2/TiO2 Thin Films. Appl. Surf. Sci. 2006, 253, 1889–1897. [Google Scholar] [CrossRef]
  100. Lee, J.H.; Mirzaei, A.; Kim, J.H.; Kim, J.Y.; Nasriddinov, A.F.; Rumyantseva, M.N.; Kim, H.W.; Kim, S.S. Gas-Sensing Behaviors of TiO2-Layer-Modified SnO2 Quantum Dots in Self-Heating Mode and Effects of the TiO2 Layer. Sens. Actuators B Chem. 2020, 310, 127870. [Google Scholar] [CrossRef]
  101. Ng, S.; Prášek, J.; Zazpe, R.; Pytlíček, Z.; Spotz, Z.; Pereira, J.R.; Michalička, J.; Přikryl, J.; Krbal, M.; Sopha, H.; et al. Atomic Layer Deposition of SnO2-Coated Anodic One-Dimensional TiO2 Nanotube Layers for Low Concentration NO2 Sensing. ACS Appl. Mater. Interfaces 2020, 12, 33386–33396. [Google Scholar] [CrossRef]
  102. Song, Z.; Wei, Z.; Wang, B.; Luo, Z.; Xu, S.; Zhang, W.; Yu, H.; Li, M.; Huang, Z.; Zang, J.; et al. Sensitive Room-Temperature H2S Gas Sensors Employing SnO2 Quantum Wire/Reduced Graphene Oxide Nanocomposites. Chem. Mater. 2016, 28, 1205–1212. [Google Scholar] [CrossRef]
  103. Nasriddinov, A.; Rumyantseva, M.; Marikutsa, A.; Gaskov, A.; Lee, J.-H.; Kim, J.-H.; Kim, J.-Y.; Kim, S.S.; Kim, H.W. Sub-Ppm Formaldehyde Detection by n-n TiO2@SnO2 Nanocomposites. Sensors 2019, 19, 3182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Li, X.; Li, X.; Wang, J.; Lin, S. Highly Sensitive and Selective Room-Temperature Formaldehyde Sensors Using Hollow TiO2 Microspheres. Sens. Actuators B Chem. 2015, 219, 158–163. [Google Scholar] [CrossRef]
  105. Righettoni, M.; Tricoli, A.; Pratsinis, S.E. Minimal Cross-Sensitivity to Humidity during Ethanol Detection by SnO2-TiO2 Solid Solutions Related Content Toward Portable Breath Acetone Analysis for Diabetes Detection Minimal Cross-Sensitivity to Humidity during Ethanol Detection by SnO2-TiO2 Solid Solutions. Nanotechnology 2009, 20, 10. [Google Scholar] [CrossRef]
  106. Li, Z.; Yao, Z.J.; Haidry, A.A.; Plecenik, T.; Xie, L.J.; Sun, L.C.; Fatima, Q. Resistive-Type Hydrogen Gas Sensor Based on TiO2: A Review. Int. J. Hydrogen Energy 2018, 43, 21114–21132. [Google Scholar] [CrossRef]
  107. Shaposhnik, D.; Pavelko, R.; Llobet, E.; Gispert-Guirado, F.; Vilanova, X. Hydrogen Sensors on the Basis of SnO2–TiO2 Systems. Sens. Actuators B: Chem. 2012, 174, 527–534. [Google Scholar] [CrossRef]
  108. Plecenik, T.; Moško, M.; Haidry, A.A.; Ďurina, P.; Truchlý, M.; Grančič, B.; Gregor, M.; Roch, T.; Satrapinskyy, L.; Mošková, A.; et al. Fast Highly-Sensitive Room-Temperature Semiconductor Gas Sensor Based on the Nanoscale Pt–TiO2–Pt Sandwich. Sens. Actuators B Chem. 2015, 207, 351–361. [Google Scholar] [CrossRef] [Green Version]
  109. Lin, J.; Heo, Y.-U.; Nattestad, A.; Sun, Z.; Wang, L.; Kim, J.H.; Dou, S.X. 3D Hierarchical Rutile TiO2 and Metal-Free Organic Sensitizer Producing Dye-Sensitized Solar Cells 8.6% Conversion Efficiency. Sci. Rep. 2014, 4, 5769. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal Oxide Gas Sensors: Sensitivity and Influencing Factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [Green Version]
  111. 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]
  112. Wang, C.; Yin, L.; Zhang, L.; Qi, Y.; Lun, N.; Liu, N. Large Scale Synthesis and Gas-Sensing Properties of Anatase TiO2 Three-Dimensional Hierarchical Nanostructures. Langmuir 2010, 26, 12841–12848. [Google Scholar] [CrossRef]
  113. Barreca, D.; Comini, E.; Ferrucci, A.P.; Gasparotto, A.; Maccato, C.; Maragno, C.; Sberveglieri, G.; Tondello, E. First Example of ZnO−TiO2 Nanocomposites by Chemical Vapor Deposition: Structure, Morphology, Composition, and Gas Sensing Performances. Chem. Mater. 2007, 19, 5642–5649. [Google Scholar] [CrossRef]
  114. Lü, R.; Zhou, W.; Shi, K.; Yang, Y.; Wang, L.; Pan, K.; Tian, C.; Ren, Z.; Fu, H. Alumina Decorated TiO2 Nanotubes with Ordered Mesoporous Walls as High Sensitivity NOx Gas Sensors at Room Temperature. Nanoscale 2013, 5, 8569–8576. [Google Scholar] [CrossRef] [PubMed]
  115. Li, Z.; Ding, D.; Liu, Q.; Ning, C.; Wang, X. Ni-Doped TiO2 Nanotubes for Wide-Range Hydrogen Sensing. Nanoscale Res. Lett. 2014, 9, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Galstyan, V.; Comini, E.; Faglia, G.; Sberveglieri, G. TiO2 Nanotubes: Recent Advances in Synthesis and Gas Sensing Properties. Sensors 2013, 13, 14813–14838. [Google Scholar] [CrossRef] [PubMed]
  117. Ratautaite, V.; Bagdziunas, G.; Ramanavicius, A.; Ramanaviciene, A. An Application of Conducting Polymer Polypyrrole for the Design of Electrochromic PH and CO2 Sensors. J. Electrochem. Soc. 2019, 166, B297–B303. [Google Scholar] [CrossRef]
  118. Celiesiute, R.; Ramanaviciene, A.; Gicevicius, M.; Ramanavicius, A. Electrochromic Sensors Based on Conducting Polymers, Metal Oxides, and Coordination Complexes. Crit. Rev. Anal. Chem. 2019, 49, 195–208. [Google Scholar] [CrossRef]
  119. Popov, A.; Brasiunas, B.; Mikoliunaite, L.; Bagdziunas, G.; Ramanavicius, A.; Ramanaviciene, A. Comparative Study of Polyaniline (PANI), Poly(3,4-Ethylenedioxythiophene) (PEDOT) and PANI-PEDOT Films Electrochemically Deposited on Transparent Indium Thin Oxide Based Electrodes. Polym. (Guildf) 2019, 172, 133–141. [Google Scholar] [CrossRef]
  120. Turemis, M.; Zappi, D.; Giardi, M.T.; Basile, G.; Ramanaviciene, A.; Kapralovs, A.; Ramanavicius, A.; Viter, R. ZnO/Polyaniline Composite Based Photoluminescence Sensor for the Determination of Acetic Acid Vapor. Talanta 2020, 211, 120658. [Google Scholar] [CrossRef] [Green Version]
  121. 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]
  122. Wu, Y.; Xing, S.; Fu, J. Examining the Use of TiO2 to Enhance the NH3 Sensitivity of Polypyrrole Films. J. Appl. Polym. Sci. 2010, 118, 3351–3356. [Google Scholar] [CrossRef]
  123. Tai, H.; Jiang, Y.; Xie, G.; Yu, J.; Zhao, M. International Journal of Environmental Analytical Chemistry Self-Assembly of TiO2/Polypyrrole Nanocomposite Ultrathin Films and Application for an NH3 Gas Sensor. Int. J. Environ. Anal. Chem. 2007, 87, 539–551. [Google Scholar] [CrossRef]
  124. Bulakhe, R.N.; Patil, S.V.; Deshmukh, P.R.; Shinde, N.M.; Lokhande, C.D. Fabrication and Performance of Polypyrrole (Ppy)/TiO2 Heterojunction for Room Temperature Operated LPG Sensor. Sens. Actuators B Chem. 2013, 181, 417–423. [Google Scholar] [CrossRef]
  125. Gong, J.; Li, Y.; Hu, Z.; Zhou, Z.; Deng, Y. Ultrasensitive NH3 Gas Sensor from Polyaniline Nanograin Enchased TiO2 Fibers. J. Phys. Chem. C 2010, 114, 9970–9974. [Google Scholar] [CrossRef]
  126. Pawar, S.G.; Chougule, M.A.; Sen, S.; Patil, V.B. Development of Nanostructured Polyaniline–Titanium Dioxide Gas Sensors for Ammonia Recognition. J. Appl. Polym. Sci. 2012, 125, 1418–1424. [Google Scholar] [CrossRef]
  127. Wang, Q.; Dong, X.; Pang, Z.; Du, Y.; Xia, X.; Wei, Q.; Huang, F. Ammonia Sensing Behaviors of TiO2-PANI/PA6 Composite Nanofibers. Sensors 2012, 12, 17046–17057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Du, P.; Song, L.; Xiong, J.; Li, N.; Xi, Z.; Wang, L.; Jin, D.; Guo, S.; Yuan, Y. Coaxial Electrospun TiO2/ZnO Core–Sheath Nanofibers Film: Novel Structure for Photoanode of Dye-Sensitized Solar Cells. Electrochim. Acta 2012, 78, 392–397. [Google Scholar] [CrossRef]
  129. Ding, Y.; Wang, Y.; Zhang, L.; Zhang, H.; Li, C.M.; Lei, Y. Preparation of TiO2–Pt Hybrid Nanofibers and Their Application for Sensitive Hydrazine Detection. Nanoscale 2011, 3, 1149–1157. [Google Scholar] [CrossRef]
  130. Li, Z.; Zhang, H.; Zheng, W.; Wang, W.; Huang, H.; Wang, C.; MacDiarmid, A.G.; Wei, Y. Highly Sensitive and Stable Humidity Nanosensors Based on LiCl Doped TiO2 Electrospun Nanofibers. J. Am. Chem. Soc. 2008, 130, 5036–5037. [Google Scholar] [CrossRef]
  131. Zeng, W.; Liu, T.; Wang, Z. Enhanced Gas Sensing Properties by SnO2 Nanosphere Functionalized TiO2 Nanobelts. J. Mater. Chem. 2012, 22, 3544–3548. [Google Scholar] [CrossRef]
  132. Zakrzewska, K. Gas Sensing Mechanism of TiO2-Based Thin Films. Vacuum 2004, 74, 335–338. [Google Scholar] [CrossRef]
  133. Liang, Y.-C.; Liu, Y.-C. Design of Nanoscaled Surface Morphology of TiO2–Ag2O Composite Nanorods through Sputtering Decoration Process and Their Low-Concentration NO2 Gas-Sensing Behaviors. Nanomaterials 2019, 9, 1150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Wang, H.; Chen, L.; Wang, J.; Sun, Q.; Zhao, Y. A Micro Oxygen Sensor Based on a Nano Sol-Gel TiO2 Thin Film. Sensors 2014, 14, 16423–16433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. SA, M.C.; Hamidon, M.N.; Mamat, M.S.; Ertugrul, M.; Abdullah, N.H. A Hydrogen Gas Sensor Based on TiO2 Nanoparticles on Alumina Substrate. Sensors 2018, 18, 2483. [Google Scholar] [CrossRef] [Green Version]
  136. Park, S.; Kim, S.; Park, S.; Lee, W.I.; Lee, C. Effects of Functionalization of TiO2 Nanotube Array Sensors with Pd Nanoparticles on Their Selectivity. Sensors 2014, 14, 15849–15860. [Google Scholar] [CrossRef] [PubMed]
  137. Yavuz, A.G.; Gök, A. Preparation of TiO2/PANI Composites in the Presence of Surfactants and Investigation of Electrical Properties. Synth. Met. 2007, 157, 235–242. [Google Scholar] [CrossRef]
  138. Oh, M.; Park, S.J.; Jung, Y.; Kim, S. Electrochemical Properties of Polyaniline Composite Electrodes Prepared by In-Situ Polymerization in Titanium Dioxide Dispersed Aqueous Solution. Synth. Met. 2012, 162, 695–701. [Google Scholar] [CrossRef]
  139. Guo, N.; Liang, Y.; Lan, S.; Liu, L.; Zhang, J.; Ji, G.; Gan, S. Microscale Hierarchical Three-Dimensional Flowerlike TiO2/PANI Composite: Synthesis, Characterization, and Its Remarkable Photocatalytic Activity on Organic Dyes under UV-Light and Sunlight Irradiation. J. Phys. Chem. C 2014, 118, 18343–18355. [Google Scholar] [CrossRef]
  140. Gottam, R.; Srinivasan, P. One-Step Oxidation of Aniline by Peroxotitanium Acid to Polyaniline–Titanium Dioxide: A Highly Stable Electrode for a Supercapacitor. J. Appl. Polym. Sci. 2015, 132. [Google Scholar] [CrossRef]
  141. Radoičić, M.; Šaponjić, Z.; Nedeljković, J.; Ćirić-Marjanović, G.; Stejskal, J. Self-Assembled Polyaniline Nanotubes and Nanoribbons/Titanium Dioxide Nanocomposites. Synth. Met. 2010, 160, 1325–1334. [Google Scholar] [CrossRef]
  142. Su, S.J.; Kuramoto, N. Processable Polyaniline–Titanium Dioxide Nanocomposites: Effect of Titanium Dioxide on the Conductivity. Synth. Met. 2000, 114, 147–153. [Google Scholar] [CrossRef]
  143. Tai, H.; Jiang, Y.; Xie, G.; Yu, J.; Chen, X. Fabrication and Gas Sensitivity of Polyaniline–Titanium Dioxide Nanocomposite Thin Film. Sens. Actuators B Chem. 2007, 125, 644–650. [Google Scholar] [CrossRef]
  144. Lee, I.S.; Lee, J.Y.; Sung, J.H.; Choi, H.J. Synthesis and Electrorheological Characteristics of Polyaniline-Titanium Dioxide Hybrid Suspension. Synth. Met. 2005, 152, 173–176. [Google Scholar] [CrossRef]
  145. Kang, K.S. Effect of Excess Amount of Aniline for TiO2 and Polyaniline Composite. Synth. Met. 2016, 217, 197–201. [Google Scholar] [CrossRef]
  146. Maldonado-Larios, L.; Mayen-Mondragón, R.; Martínez-Orozco, R.D.; Páramo-García, U.; Gallardo-Rivas, N.V.; García-Alamilla, R. Electrochemically-Assisted Fabrication of Titanium-Dioxide/Polyaniline Nanocomposite Films for the Electroremediation of Congo Red in Aqueous Effluents. Synth. Met. 2020, 268, 116464. [Google Scholar] [CrossRef]
  147. Ilieva, M.; Ivanov, S.; Tsakova, V. Electrochemical Synthesis and Characterization of TiO2-Polyaniline Composite Layers. J. Appl. Electrochem. 2008, 38, 63–69. [Google Scholar] [CrossRef]
  148. Saeb, E.; Asadpour-Zeynali, K. Facile Synthesis of TiO2@PANI@Au Nanocomposite as an Electrochemical Sensor for Determination of Hydrazine. Microchem. J. 2021, 160, 105603. [Google Scholar] [CrossRef]
  149. Cai, G.; Tu, J.; Zhou, D.; Zhang, J.; Xiong, Q.; Zhao, X.; Wang, X.; Gu, C. Multicolor Electrochromic Film Based on TiO2@Polyaniline Core/Shell Nanorod Array. J. Phys. Chem. C 2013, 117, 15967–15975. [Google Scholar] [CrossRef]
  150. Gobal, F.; Faraji, M. Electrodeposited Polyaniline on Pd-Loaded TiO2 Nanotubes as Active Material for Electrochemical Supercapacitor. J. Electroanal. Chem. 2013, 691, 51–56. [Google Scholar] [CrossRef]
  151. Jagminas, A.; Balčiūnaitė, A.; Niaura, G.; Tamašauskaitė-Tamašiūnaitė, L. Electrochemical Synthesis and Characterisation of Polyaniline in TiO2 Nanotubes. Trans. IMF 2012, 90, 311–315. [Google Scholar] [CrossRef]
  152. Patil, B.H.; Jang, K.; Lee, S.; Kim, J.H.; Yoon, C.S.; Kim, J.; Kim, D.H.; Ahn, H. Periodically Ordered Inverse Opal TiO2/Polyaniline Core/Shell Design for Electrochemical Energy Storage Applications. J. Alloys Compd. 2017, 694, 111–118. [Google Scholar] [CrossRef]
  153. Gao, L.; Yin, C.; Luo, Y.; Duan, G. Facile Synthesis of the Composites of Polyaniline and TiO2 Nanoparticles Using Self-Assembly Method and Their Application in Gas Sensing. Nanomaterials 2019, 9, 493. [Google Scholar] [CrossRef] [Green Version]
  154. Ma, X.; Wang, M.; Li, G.; Chen, H.; Bai, R. Preparation of Polyaniline–TiO2 Composite Film with in Situ Polymerization Approach and Its Gas-Sensitivity at Room Temperature. Mater. Chem. Phys. 2006, 98, 241–247. [Google Scholar] [CrossRef]
  155. Huyen, D.N.; Tung, N.T.; Thien, N.D.; Thanh, L.H. Effect of TiO2 on the Gas Sensing Features of TiO2/PANi Nanocomposites. Sensors 2011, 11, 1924–1931. [Google Scholar] [CrossRef] [Green Version]
  156. Tai, H.; Jiang, Y.; Xie, G.; Yu, J.; Chen, X.; Ying, Z. Influence of Polymerization Temperature on NH3 Response of PANI/TiO2 Thin Film Gas Sensor. Sens. Actuators B Chem. 2008, 129, 319–326. [Google Scholar] [CrossRef]
  157. Bairi, V.G.; Bourdo, S.E.; Sacre, N.; Nair, D.; Berry, B.C.; Biris, A.S.; Viswanathan, T. Ammonia Gas Sensing Behavior of Tanninsulfonic Acid Doped Polyaniline-TiO2 Composite. Sensors 2015, 15, 26415–26429. [Google Scholar] [CrossRef]
  158. Pawar, S.G.; Chougule, M.A.; Patil, S.L.; Raut, B.T.; Godse, P.R.; Sen, S.; Patil, V.B. Room Temperature Ammonia Gas Sensor Based on Polyaniline-TiO $ _ {2} $ Nanocomposite. IEEE Sens J 2011, 11, 3417–3423. [Google Scholar] [CrossRef]
  159. Seif, A.; Nikfarjam, A.; Hajghassem, H. UV Enhanced Ammonia Gas Sensing Properties of PANI/TiO2 Core-Shell Nanofibers. Sens. Actuators B Chem. 2019, 298, 126906. [Google Scholar] [CrossRef]
  160. Cui, S.; Wang, J.; Wang, X. Fabrication and Design of a Toxic Gas Sensor Based on Polyaniline/Titanium Dioxide Nanocomposite Film by Layer-by-Layer Self-Assembly. RSC Adv. 2015, 5, 58211–58219. [Google Scholar] [CrossRef]
  161. Zhu, C.; Cheng, X.; Dong, X. Enhanced Sub-Ppm NH3 Gas Sensing Performance of PANI/TiO2 Nanocomposites at Room Temperature. Front. Chem. 2018, 6, 493. [Google Scholar] [CrossRef]
  162. Li, Y.; Gong, J.; He, G.; Deng, Y. Fabrication of Polyaniline/Titanium Dioxide Composite Nanofibers for Gas Sensing Application. Mater. Chem. Phys. 2011, 129, 477–482. [Google Scholar] [CrossRef]
  163. Nasirian, S.; Milani Moghaddam, H. Effect of Different Titania Phases on the Hydrogen Gas Sensing Features of Polyaniline/TiO2 Nanocomposite. Polymer (Guildf) 2014, 55, 1866–1874. [Google Scholar] [CrossRef]
  164. Gawri, I.; Ridhi, R.; Singh, K.P.; Tripathi, S.K. Chemically Synthesized TiO2 and PANI/TiO2 Thin Films for Ethanol Sensing Applications. Mater. Res. Express 2018, 5, 025303. [Google Scholar] [CrossRef]
  165. Wang, Z.; Peng, X.; Huang, C.; Chen, X.; Dai, W.; Fu, X. CO Gas Sensitivity and Its Oxidation over TiO2 Modified by PANI under UV Irradiation at Room Temperature. Appl. Catal. B Environ. 2017, 219, 379–390. [Google Scholar] [CrossRef]
  166. Su, P.G.; Huang, L.N. Humidity Sensors Based on TiO2 Nanoparticles/Polypyrrole Composite Thin Films. Sens. Actuators B Chem. 2007, 123, 501–507. [Google Scholar] [CrossRef]
  167. Srivastava, S.; Kumar, S.; Singh, V.N.; Singh, M.; Vijay, Y.K. Synthesis and Characterization of TiO2 Doped Polyaniline Composites for Hydrogen Gas Sensing. Int. J. Hydrogen Energy 2011, 36, 6343–6355. [Google Scholar] [CrossRef]
  168. Dhawale, D.S.; Salunkhe, R.R.; Patil, U.M.; Gurav, K.V.; More, A.M.; Lokhande, C.D. 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]
  169. Parveen, A.; Koppalkar, A.; Roy, A.S. Liquefied Petroleum Gas Sensing of Polyaniline–Titanium Dioxide Nanocomposites. Sens. Lett. 2013, 11, 242–248. [Google Scholar] [CrossRef]
  170. Zheng, J.; Li, G.; Ma, X.; Wang, Y.; Wu, G.; Cheng, Y. Polyaniline–TiO2 Nano-Composite-Based Trimethylamine QCM Sensor and Its Thermal Behavior Studies. Sens. Actuators B Chem. 2008, 133, 374–380. [Google Scholar] [CrossRef]
Figure 1. Temperature dependence of electrical resistance (R(T)) for the TiO2−x/TiO2(400 °C)-based hetero-structure. Temperature was changed in two ways (indicated by black and red arrows): (i) black cycles shows points measured by cooling down, (ii) red squares shows points by increasing temperature. Measurements were performed in vacuum using helium cryostat. Figure adapted from [1].
Figure 1. Temperature dependence of electrical resistance (R(T)) for the TiO2−x/TiO2(400 °C)-based hetero-structure. Temperature was changed in two ways (indicated by black and red arrows): (i) black cycles shows points measured by cooling down, (ii) red squares shows points by increasing temperature. Measurements were performed in vacuum using helium cryostat. Figure adapted from [1].
Coatings 12 00699 g001
Figure 2. Band diagram of TiO2/GO hetero-structure (a) before the formation of hetero-structure, (b) after the formation of hetero-structure, (c) when UV irradiation is applied, and (d) when UV irradiation is switched off. (‘e’ is an electron; ‘h’ is a hole).
Figure 2. Band diagram of TiO2/GO hetero-structure (a) before the formation of hetero-structure, (b) after the formation of hetero-structure, (c) when UV irradiation is applied, and (d) when UV irradiation is switched off. (‘e’ is an electron; ‘h’ is a hole).
Coatings 12 00699 g002
Figure 3. Representation of typical analytical signal for amperometric gas sensors. It should be noted that the duration of signal development and the regeneration of sensors highly depends on sensing material and sensing gases [1].
Figure 3. Representation of typical analytical signal for amperometric gas sensors. It should be noted that the duration of signal development and the regeneration of sensors highly depends on sensing material and sensing gases [1].
Coatings 12 00699 g003
Figure 4. Representation of conductivity mechanism of semiconductor particle-based structure during the determination of CO gas. (A) Structure before the interaction with CO, (B) structure during the interaction with CO. Figure adapted from [2].
Figure 4. Representation of conductivity mechanism of semiconductor particle-based structure during the determination of CO gas. (A) Structure before the interaction with CO, (B) structure during the interaction with CO. Figure adapted from [2].
Coatings 12 00699 g004
Figure 5. Different morphology TiO2 structures used in design of a gas sensors. (A) Nanorods [133], (B) thin film [134], (C) nanoparticles [135], (D) nanotubes [136].
Figure 5. Different morphology TiO2 structures used in design of a gas sensors. (A) Nanorods [133], (B) thin film [134], (C) nanoparticles [135], (D) nanotubes [136].
Coatings 12 00699 g005
Table 1. Characteristics of titanium oxide-based sensors.
Table 1. Characteristics of titanium oxide-based sensors.
Sensing MaterialWorking TemperatureGas ConcentrationResponse Value (Ra/Rg) or ((ΔR/Rg) × 100%)Response TimeRecovery TimeReference
TiO2 (rutile), Ti8O15 and Ti9O17 mixture210 °C12.5–100 ppm (NH3)1–7%2 min8 min[68]
TiOx-NiO250–350 °C100 ppm (H2)
100 ppm (NO2)
100 ppm (NH3)
17 for H2 (250 °C)
16 for NO2 (250 °C)
4 for NH3 (250 °C)
2 min2, 3 min[69]
β-Ti3O5150 °C50 ppm (H2)11%--[70]
Ti3O5-TiO2 mixture25–180 °C105 ppm (H2O)
118 ppm (methanol)
53 ppm (ethanol)
18 ppm n-propanol
220 ppm (acetone)
0.5–18%-4–35 s[1]
TiO2-Ti6O150–450 °C2000 pm (H2)
20 ppm (NO2)
500 ppb (O3)
1.6 ppm (acetone)
80 ppm (NOx)
2.9–348 8–21 s20–32 s[34]
Ti3+-TiO2RT100 ppm (CO)39%10 s30 s[71]
TiO2150 °C100 ppm (ethanol)75.4%155 s779 s[72]
TiO2270 °C500 ppm (acetone)9.1910 s9 s[73]
TiO2350 °C400 ppm (ethanol)22.95 s7 s[74]
TiO2RT200 ppm (NH3)6428 s24 s[75]
α-Fe2O3-TiO2325 °C100 ppm (ethanol)446 s16 s[76]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramanavicius, S.; Jagminas, A.; Ramanavicius, A. Gas Sensors Based on Titanium Oxides (Review). Coatings 2022, 12, 699. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings12050699

AMA Style

Ramanavicius S, Jagminas A, Ramanavicius A. Gas Sensors Based on Titanium Oxides (Review). Coatings. 2022; 12(5):699. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings12050699

Chicago/Turabian Style

Ramanavicius, Simonas, Arunas Jagminas, and Arunas Ramanavicius. 2022. "Gas Sensors Based on Titanium Oxides (Review)" Coatings 12, no. 5: 699. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings12050699

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