Next Article in Journal
An Electrically Rechargeable Zinc/Air Cell with an Aqueous Choline Acetate Electrolyte
Next Article in Special Issue
(Ti,Sn) Solid Solution Based Gas Sensors for New Monitoring of Hydraulic Oil Degradation
Previous Article in Journal
Compositional Dependence of Pore Structure, Strengthand Freezing-Thawing Resistance of Metakaolin-Based Geopolymers
Previous Article in Special Issue
Accelerated Life Testing of a Palladium-Doped Tin Oxide Electrode for Zn Electrowinning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

1D Titanium Dioxide: Achievements in Chemical Sensing

SENSOR Laboratory, Department of Information Engineering, University of Brescia, Via Valotti 9, 25133 Brescia, Italy
*
Author to whom correspondence should be addressed.
Materials 2020, 13(13), 2974; https://doi.org/10.3390/ma13132974
Submission received: 28 May 2020 / Revised: 29 June 2020 / Accepted: 30 June 2020 / Published: 3 July 2020
(This article belongs to the Special Issue TiO2-Based Nanostructures, Composites and Hybrid Photocatalysts)

Abstract

:
For the last two decades, titanium dioxide (TiO2) has received wide attention in several areas such as in medicine, sensor technology and solar cell industries. TiO2-based gas sensors have attracted significant attention in past decades due to their excellent physical/chemical properties, low cost and high abundance on Earth. In recent years, more and more efforts have been invested for the further improvement in sensing properties of TiO2 by implementing new strategies such as growth of TiO2 in different morphologies. Indeed, in the last five to seven years, 1D nanostructures and heterostructures of TiO2 have been synthesized using different growth techniques and integrated in chemical/gas sensing. Thus, in this review article, we briefly summarize the most important contributions by different researchers within the last five to seven years in fabrication of 1D nanostructures of TiO2-based chemical/gas sensors and the different strategies applied for the improvements of their performances. Moreover, the crystal structure of TiO2, different fabrication techniques used for the growth of TiO2-based 1D nanostructures, their chemical sensing mechanism and sensing performances towards reducing and oxidizing gases have been discussed in detail.

1. Introduction

With the worldwide industrial and technological growth, there is a continuous demand for portable, low-cost and efficient sensing devices that can be used for various purposes such as environmental monitoring and health care. Due to this, researchers have made continuous efforts to find new materials or to grow existing materials in different forms such as thin films and nanostructures. In the field of chemical sensors, nanostructured metal oxides (MOX) are the most versatile and widely studied active sensing materials due to their unique physical/chemical properties [1]. Their wide band gap, along with the exceptional electrical properties, makes MOXs among the leading candidate for transparent electronic devices [2]. Indeed, tin oxide (SnO2) in different nanostructures forms such as nanowires [3,4], nanobelt [5,6] and nanorods [7,8] is the most widely investigated material in the field of chemical/gas sensing.
Among the different nanostructures’ forms, one-dimensional (1D) nanostructures such as nanowires exhibit unique properties such as high surface to volume ratio, high crystallinity and controlled electrical properties which make them interesting for chemical sensing applications [3,9,10,11,12]. Because there is no requirement of preliminary gas diffusion to the surface, 1D MOX gas sensors showed faster response dynamics. Moreover, the self-heating phenomenon of nanowires can be used to reduce the power consumption of temperature-driven devices such as conductometric gas sensors, as nanowires attain high temperature when they undergo electrical resistance measurements, even when low electrical power has been used during the measurements [1,13].
Recently, MOXs such as zinc oxide (ZnO) [14], nickel oxide (NiO) [15], titanium dioxide (TiO2) [12] and tungsten oxide (WO3) [11] have been grown in the form of 1D nanostructures and were used for the detection of different gas analytes. Among these MOXs, TiO2 is one of the most versatile ones and is widely used for many different applications, including photocatalysis [16], sensors [17,18] and photovoltaic applications [19]. In particular, due to its exceptional physical/chemical properties, TiO2 was extensively used for different sensing applications such as chemical/gas sensors [20,21], biosensors [17,22] and chemical oxygen demand (COD) [23,24] sensors [25]. It typically exhibits n-type conductivity and, in nanostructured form, is chemically stable, biocompatible, biodegradable and inexpensive [25]. It exhibits three different crystal structures—namely, anatase, rutile and brookite, each of them possessing unique structural, electrical and optical properties [26]. Rutile TiO2 is widely used as a pigment, opacifier, isolator, and in switches, etc. due to its light-scattering ability, high refractive index and ultraviolet (UV) absorptivity, while anatase TiO2 is preferred for photovoltaic and photocatalytic applications [27]. On the other hand, brookite is the rarest occurring phase of TiO2, which is formed under a particular set of conditions. Due to this, brookite TiO2 has been very rarely investigated materials and its applications in fields such as photovoltaic applications [28] and lithium ion batteries [29] were recently discovered.
In this review article, recent achievements of 1D TiO2 chemical/gas sensors will be presented. Specifically, crystal structure of TiO2, MOX gas sensing mechanism and techniques used to grow 1D TiO2 nanostructures will be discussed in detail. In order to describe the recent achievements of 1D TiO2 chemical sensors, literature reported over the span of the last five to seven years will be presented.

2. Crystal Structure of TiO2

The structural properties of the active sensing material in a gas sensor (either in bulk, thin films or nanostructures form) are among the most important characteristics, as they influence the electrical, optical and mechanical behavior of the material. In particular, the surface structure and defectivity/stoichiometry of metal oxides has a great influence on local surface chemistry [26] and, consequently, they affect their performances as active sensing materials decisively. It is therefore important to understand the crystal structure of a chemo-resistive MOX to be used in a sensing device and to study how its structural properties may change as a function of the growth parameters (such as temperature). Hence, in this section, we will discuss the crystal structure of TiO2.
TiO2 exists in nature as well-known minerals rutile (tetragonal), anatase (tetragonal) and brookite (orthorhombic) [30]. In addition to these forms, two high-pressure forms of TiO2 also exist, i.e., a monoclinic baddeleyite-like form and an orthorhombic α-PbO2-like form, which were found at Ries Crater in Bavaria [31]. Indeed, rutile is the most thermodynamically stable polymorph of TiO2.
Figure 1 shows the crystal structures of the TiO2 anatase, rutile and brookite phases [32], and their corresponding crystal data are shown in Table 1 [26,33,34]. It should be noted that, due to the rare existence of brookite phase, the anatase and rutile forms of TiO2 are the most exploited in real applications.
All the crystal structures of TiO2 are made up of distorted octahedra, each one representing a TiO6 unit, where Ti4+ is at the center of the unit and coordinates six O2− ions. The way in which octahedra assemble to form the TiO6 chain is different for each crystal structure. Anatase and rutile crystal structures are assembled by connecting the octahedra by their vertices and edges, respectively, while in brookite, both vertices and edges are connected [34].
At room temperature, the anatase phase is kinetically stable and does not undergo any phase transformation. Nevertheless, it can be converted into rutile via heating at high temperature [26]. The anatase-to-rutile thermally-induced transition in bulk TiO2 starts occurring at temperature T > 600 °C. M.N. Asiah et al. [12] have studied the phase transformation of TiO2 nanowires grown by means of a hydrothermal method. They found that, up to 500 °C, the nanowires (anatase phase) retained their morphology, and they started to change into smaller particles at around 600 °C. At 900 °C, the complete transformation from anatase to rutile phase occurred and the nanowires’ shapes changed to rod-like structures. Figure 2 shows the SEM images on TiO2 nanowires corresponding to the different temperatures [8].
Moreover, it was also found that the surface defects and lattice concentration play a crucial role during the phase transformation [34]. Specifically, the surface defects acted as nucleation sites and the rate of rutile transformation increased with an increase in surface defects. The removal of oxygen ions from TiO2 lattice (i.e., the formation of oxygen vacancies) also accelerated the rutile phase formation.

3. Synthesis of TiO2-Based 1D Nanostructures

This thematic study acquaints the fabrication techniques commonly used for the synthesis of TiO2 nanostructures. In the past decade, diverse fabrication methods, including chemical vapor deposition (CVD) [35], atomic layer deposition (ALD) [36], electrochemical deposition [37], hydrothermal methods [38], sol-gel [39] and electrospinning [40] have been successfully used for the preparation of high-quality TiO2 based nanostructures. The goal of this section is to elucidate the fabrication methods most commonly employed in the last five to seven years.

3.1. Hydrothermal Synthesis

Hydrothermal synthesis is one of the most widely used techniques for the growth of TiO2 nanostructures. This synthesis process is normally performed in steel autoclaves with or without Teflon liners under controlled pressure and temperature in an aqueous solution. All the experimental parameters involved in a hydrothermal process (e.g., annealing treatment, specific titanium precursor and process duration) influence the nanostructure growth. However, the nature of aqueous solution (acid/alkali concentration) has a major role in the structure morphology and crystallinity. B. Zaho et al. [38] reported that five different Titania/Titanate phases could be generated by employing a proper composition of acid (HCl) and alkali. High HCl concentration (6 M) lead to the formation of rutile nanorods, while monoclinic trititanate Na2Ti3O7 nanoribbons were formed at high NaOH concentration (10 M), as shown in Figure 3. These transitions have been explained by the Ostwald’s ripening mechanism [41].
On the other hand, recently, many reports have been published on using hydrothermal synthesis for the growth of TiO2 nanowires (NWs) and nanorods (NRs) of rutile and anatase phase [42,43,44]. Generally speaking, typical processes to produce TiO2 nanorod arrays inside a steel autoclave involve a Ti precursor, HCl solution and deionized water, heated up at 150 °C for at least 9 h, followed by thermal annealing [42,43]. Another interesting work by G. Zaho et al. [44] reports about the growth of bridge-like structures of TiO2 NRs using isopropyl titanate as a precursor (Figure 4). Additionally, using Ti foil or TiO2 nanoparticles in appropriate concentrations of NaOH/ HCl and deionized water leads to successful growth of anatase TiO2 nanowires and nanotubes [45,46].

3.2. Electrochemical Anodization

Electrochemical anodization is an efficient method for the growth of well-aligned and highly ordered nanotubes (NTs). In a typical process, the NTs are produced at room temperature using a current generator in a cell consisting of an electrolyte and two electrodes—titanium (anode) and platinum (cathode). From the reported literature, it has been seen that the water-based ethylene glycol and ammonium fluoride aqueous electrolytes are the most commonly used for anodization of titanium to fabricate TiO2 NTs [47,48,49,50].
Generally, for the production of highly ordered NTs, an annealing treatment is required after the anodization. Different parameters such as voltage applied, anodization time and type of electrolyte solution are the major factors that influence the growth process [31,51,52]. It has been reported that highly uniform and longer NTs can be achieved when using high voltage, while the anodization time can majorly affect the length of NTs, but not their morphology [53,54]. Reports by B. Munirathinam et al. [53] and H. Sopha et al. [51] stated that the nature and the age of the electrolyte have major effects on the morphology and aspect ratio of NTs, as shown in Figure 5. This technique is usually carried out at room temperature, and it opens up great possibilities in the use of different substrates for an ultimate objective of fabricating new generation devices.

3.3. Electrospinning

In comparison to the above-mentioned growth techniques, electrospinning is one of the most simple and flexible methods to synthesize the large-scale 1D TiO2 nanostructures, especially long nanofibers of TiO2. Figure 6 shows a schematic diagram of the electrospinning equipment [55]. In this method, a liquid precursor solution is injected through a spinneret under an applied electric field and spin force to create fibrous structures. A post-annealing treatment is commonly needed to remove the solvent and to solidify the nanofibers. In recent years, electrospinning technique has been used not only to fabricate pristine nanostructures of TiO2, but also the composite [56,57,58,59,60].
In a recent report by M. Zhou et al. [56], the authors synthesized TiO2 nanowires of different phases using the typical electrospinning method. Authors studied the effect of calcination temperature after the electrospinning of the nanowires at 500 °C, 600 °C, 700 °C and 800 °C for 2 h in air (shown in Figure 7). It was discovered that by increasing the annealing temperature, the grain size and pore size increase as well, and a phase transition from rutile to anatase TiO2 occurs, while the band gap decreases.
Another interesting report, authored by S. Wang et al. [57], reported the fabrication of TiO2 nanowires with two different methods, namely, an ethylene glycol-mediated method and a high-voltage electrospinning method. The nanowires synthesized via the solution method were found to be relatively uniform, having an average diameter around 550 ± 90 nm and a length of 26 ± 5 μm. On the other hand, nanowires grown with the electrospinning method were found to be much smaller, exhibiting an average diameter of 70 ± 5 nm and a length of 12 ± 5.5 μm. In 2017, F. Li et al. [59] presented the coaxial electrospinning method using two precursors to fabricate the core-shell nanostructures of TiO2/SnO2. The structural characterizations showed two mixed phases of tetragonal SnO2 and rutile TiO2 with no other impurity diffraction peaks evidenced by X-ray diffraction analysis.
Moreover, besides the above-mentioned growth process for TiO2 nanostructures, the other least used techniques within the last five year were thermal oxidation [61], acid vapor deposition (AVO) [62] and the vapor phase growth process [63].
Finally, the various reports published in the last five years on the growth of TiO2 nanostructures and its composites have been collected in Table 2, while in Table 3, some the major advantages and disadvantages related to these techniques have been presented.

4. Working Principles of TiO2-Based Chemical Sensors

Most semiconductor-based gas sensors rely on the chemo-resistive effect—that is, the change of the electrical conductance or resistance upon the exposure to chemical compounds. In particular, the semiconductor sensing element resistance can increase or decrease when reacting with oxidizing (O3, CO2 and NO2) or reducing (H2, H2S, NH3 and VOCs) gases [66,67,71,75,76]. The sensing mechanism of TiO2-based chemical sensors is frequently explained by two processes: receptor and transduction function, as shown in Figure 8.
The receptor process takes place on the surface of TiO2, comprising physisorption and chemisorption processes [77,78]. As TiO2 surface is exposed to air, the oxygen molecules start to physically adsorb on its surface. This process is determined by dipole and van der Waals interactions. Furthermore, these adsorbed molecules initiate capturing electrons from the conduction band of TiO2, while oxygen species chemisorb ( O , O 2 ) on the surface. This chemisorption of oxygen ions creates a depletion region near the TiO2 surface.
The receptor process efficiency is majorly dependent on physisorption and chemisorption processes, which in turn depend upon the sensor operating temperature. Specifically, at elevated temperature (100–500 °C), the chemisorption of oxygen ions occurs on the oxide surface and can be explained by the following equations [79]:
O 2 ( gas ) O 2 ( ads )
O 2 ( ads ) + e O 2 ( ads ) ( < 100   ° C )
O 2 ( ads ) + e 2 O ( ads ) ( 100 300   ° C )  
O ( ads ) + e O ( ads ) 2 ( > 300   ° C )
Hence, the type of chemisorbed oxygen ions depends on the sensors’ operating temperature and determines the reaction mechanism with the gas analyte. Due to this, metal oxides gas sensors are operated at different temperatures to find the optimal working conditions for each gas analyte.
Secondly, the transducer process is determined by the transportation of electrons in TiO2 and their transformation into a readable electronic signal. This process is mainly divided into three different modes (shown in Figure 8), i.e., surface-controlled mode, grain-controlled mode and neck-controlled mode [77,78,80,81]. (i) The surface-controlled model is mainly related to compact thin film structures, in which the gases react to the surface of TiO2 rather than the bulk. (ii) On the other hand, in the grain-neck-controlled model, the gases reactivity is much higher. This is due to the fact that in the granular structures with higher porosity, each grain and neck boundary will create a depletion region that leads to an increase in the sensor resistance, thus improving the sensitivity of the TiO2 nanostructures.
Furthermore, the surface to volume ratio, grain size and porosity of the fabricated TiO2 nanostructures can majorly affect the above mention process. In this direction, in the last decade, different 1D nanostructures of TiO2 and their heterostructures, as well as element doping of TiO2, have been highly studied in chemical sensing applications [58,61,67,71,74].

5. Chemical Sensing Properties

5.1. Sensing of Reducing Gases

Among many metal oxides, 1D TiO2 nanostructures have been extensively used as sensing materials to build a high-performance chemical sensor for the detection of reducing gases such as H2, VOCs (ethanol, acetone etc.) H2S, NH3 and so on [40,71,72,78].
Among chemical compounds crucial to detect and monitor, H2 is one of the most important. H2 is used in applications such as fuel cells [82]; it is extremely flammable and explosive, and, therefore, its detection is crucial to guarantee a safe exploitation of H2 based energy-related systems. M. Enachi et al. [67] reported the sensing performances of three different phases of TiO2 nanotubes including amorphous, anatase and mixed anatase/rutile. The phase transition was achieved using thermal treatment by annealing the sample after the deposition procedure. The authors separated a single nanotube and transferred it to the chip containing Au metal contacts to study the suitability of an individual TiO2 nanotube as a hydrogen sensor and to investigate the sensing performance for the three obtained phases. The sensing device based on single nanotubes with mixed anatase/rutile phase exhibited a faster response and recovery time of ~0.7 s and ~0.9 s, respectively, and the response (defined as I H 2 I air ) was about 3.5 at room temperature.
In order to improve the sensing performances of 1D TiO2, many strategies have been proposed in the literature, such as increasing the response and reducing the power consumption using photoactivation, improving the selectivity, enhancing the response and reducing the working temperature by functionalization TiO2 with transition metals, and maximizing the sensitivity by bulk doping and by building heterojunctions.
A. Nikfarjam et al. proposed photoactivation as a method to increase the response of the thermoactivated TiO2 nanofibers network [40]. The best response was found 18 and 96 towards 25 and 200 ppm of H2, respectively, under UV illumination at 190 °C, significantly higher compared to 1.8 and 10.1 in dark at 290 °C. The sensor exhibited a fast response, and a long-term stability of 9 months was reported. Moreover, the optimal working temperature decreased from 290 to 190 °C. This increase in performance was attributed to the reduction of the activation energy between H2 molecules and TiO2 surface under UV irradiation, and to the increase in oxygen chemisorption (O-) density for its interaction with H2, even at low temperatures. Such results are interesting for the possibility to achieve battery-operated devices thanks to the reduction in power consumption and miniaturization of the gas sensor device.
Another way to increase sensing performances is the functionalization of the TiO2 surface. For example, TiO2 nanorods produced using the hydrothermal method were modified with Pd nanoparticles [75]. Figure 9 shows Pd/TiO2 NRs exhibiting a great enhancement toward 1000 ppm of H2 at 200 °C with a 35-fold higher response compared with TiO2 NTs due to the spillover effect. Specifically, Pd nanoparticles on the TiO2 NRs catalytically stimulate the dissociation of molecular oxygen and atomic products obtained during this process diffuse to the metal oxide. This process greatly enhances the oxygen vacancies on the TiO2 surface and directly affects the sensing performance. Furthermore, together with the good selectivity towards H2 over other VOCs, the working temperature was reduced from 200 to 30 °C and the response increased to 250, thanks to the hydrogen collector activity of Pd.
Furthermore, the combination of two different catalysts may increase performance even more. Shoutian Ren et al. [45] achieved an enhancement in response and response time (350% and 42 s) for AuPd-TiO2 as compared with Pd-TiO2 (2% and 21 s) at room temperature. They also investigated the sensing mechanism by formation of a local heterostructure between Au and Pd, which enhanced electrocatalytic activity of alloyed macrostructures and increased the hydrogen dissociation process further; a crucial role was played by the size and the content of Pd/Au ratios. Moreover, heterostructures based on titanium oxide such as Co3O4/TiO2 and CuO/TiO2 also revealed high potentialities for hydrogen detection [48,49].
Other chemical species that deserve attention are volatile organic compounds (VOCs). VOCs, which commonly arise from painting, vehicle exhaust emissions and oil refining, are hazardous compounds. They have serious effects on air quality and human health. TiO2 nanostructures have been investigated as VOC sensors. In particular, TiO2 nanorods showed good sensing performances in triethylamine (TEA) detection [43]. The fast response time (2s), low detection limit (0.1 ppm) and good selectivity over acetone, ethanol, benzene, chlorobenzene, methanol and N-propanol make it a good choice for controlling fish and seafood freshness [43]. In this study, the sensing mechanism of TEA has been proved using XPS, confirming that TEA adsorbs on the surface of TiO2 NRs before reacting with pre-adsorbed oxygen. Ethanol and acetone are the most studied among VOCs targets for 1D TiO2 [83,84,85]. In this context, it has been shown that TiO2 nanorods with a diameter of 500 nm show a good response of 9, 13 and 20 towards 10, 300 and 1000 ppm of acetone at 500 °C with response and recovery time ranging from 11 to 14 s and 4 to 8 s, respectively, showing a fast acetone detection [86]. TiO2 nanotubes with 0.85 µm lengths and wall thickness of 13 nm show efficient low-temperature (150 °C) ethanol sensing with an 80% response exposed to 1000 ppm, and response and recovery times of 11 and 117 s, respectively [85].
However, in the last five years, greater interest was focused on heterostructured and composite titania material for the purpose of increasing their sensing performances. Siowwoon Ng et al. [47] investigated TiO2/ZnO core-shell nanostructures for the development of ethanol sensors. In their work, a fascinating strategy has been employed to control the ZnO shell thickness (0.19, 1.9, 7.6 and 19 nm). The response of TiO2/ZnO core shell nanostructures towards 1930 ppm of ethanol is improved by increasing the thickness of ZnO shell; with a 19 nm thickness, it is 11 times higher than the one of pristine TiO2 NTs at 200 °C (Figure 10). The sensor response dynamics are enhanced, and good repeatability and stability have been achieved. Similar results have been established on 1D TiO2 heterostructures with Nb2O5, Al2O3, V2O5 and TiO2−x materials [58,61,74,87].
Enhancing sensing performances of heterojunction are due to band bending at the interface attributed to Fermi level alignment via charge transfer; furthermore, active sites are provided for gas adsorption and synergistic reaction effects [47]. The improvement of sensing characteristics toward acetone have been investigated using different strategies, either by using heterojunction structure as reported by Feng Li et al., [59] who studied TiO2/SnO2 core shell nanofibers and observed a response improvement from 3 to 14 for TiO2 and TiO2/SnO2, respectively, at 280 °C; by surface sensitization [71]; or sometimes by bulk doping [73].
1D TiO2 also showed good potentialities as NH3 and H2S chemical sensors. Capability of detection of low concentrations of NH3 at room temperature in different humidity environments has been proven [44]. The sensor exhibited good long-term stability (29 days), with a response of 102% toward 100 ppm of NH3. Moreover, a decrease in response was observed by increasing humidity levels from 28% to 75%. The study also showed a good selectivity to NH3 over CH4, H2S, NO2, H2, ethanol, methanol, acetone, diethylaniline (DEA) and TEA.
Another important chemical compound in terms of health, security and safety is H2S. H2S is a colorless gas that is very poisonous and extremely flammable. Exposure to H2S concentration ranging from 20 to 100 ppm may cause eye burns and respiratory problems, and sometimes paralysis and even death [88,89]. Therefore, finding equipment for early detection of H2S is highly recommended. In this context, P. M. Perillo et al. fabricated flexible H2S sensors based on anatase TiO2 nanotubes [72]. The tube length was 12 µm, while the pore and wall diameters were approximately 100 nm and 30 nm, respectively. The flexible sensor showed response of 75% towards 25 ppm of H2S at room temperature with excellent long-term stability (6 months).
Table 4 reports a literature survey for the TiO2 nanostructure and heterostructure growth procedures and their chemical/gas sensing performances towards reducing gases.

5.2. Sensing of Oxidizing Gases

Environmentally hazardous oxidizing compounds such as NO2, O3 and CO2, etc. are among the major concerns for the human life and health [11,90,91,92,93]. For example, NO2 is a typical air pollutant byproduct of combustion facilities, aircrafts and automobiles. According to the American standard air quality monitoring protocols for NO2, the concentration should be detectable in the range of 3−25 ppm. Moreover, the decomposition of NO2 by solar irradiation is a common source of ozone (O3) production. The safe concentration of ozone is 50 ppb for continuous exposure and 100 ppb for short-term exposure has been set for the developed countries. Thus, a low-cost, reliable sensor with high selectivity and sensitivity is of huge demand for environmental safety and industrial control purposes. In this context, TiO2 nanostructure-based sensors have been widely investigated for the detection of these hazardous pollutants [48,50,60,61,63,64,65,66,94].
A recent report by Z. Zhu et al. [66] presented highly sensitive NO2 sensors based on TiO2 nanowires. These hydrothermally grown nanowires-based TiO2 sensors showed very fast response and recovery time of 10 to 19 s respectively, with a response of 3.1 towards 100 ppm of NO2. Furthermore, TiO2 nanorods-based sensors have been widely used for the investigation of highly sensitive NO2 sensors. Z. Tshabalala et al. [64] presented a hydrothermal method followed by an annealing process to fabricated pristine TiO2 NRs. Different devices were prepared using different washing solutions such as distilled water and hydrochloric acid of different concentrations 0.25, 0.5 and 1 M, and were annealed at different temperatures afterward. Gas tests, performed at RT in dry air, show that the sensing properties depend on washing solution and annealing temperature. The best response (1300 at 40 ppm of NO2) was achieved by the samples cleaned with hydrochloric acid 1.0 M and annealed at 700 °C. The best optimized samples exhibit high porosity, which leads to the larger surface area, in turn increasing the response.
As discussed in pervious sections, in the last five years, much work has been done to achieve high sensing performance by doping of metal particles or by creating the oxide-to-oxide junction. D. Ponnuvelu et al. [94] reported on Au-decorated mixed phase TiO2 NRs fabricated by hydrothermal method for highly sensitive NO2 sensors. After Au particle decoration, not only did the response value increase, but also the working temperature decreased from 400 °C to 200 °C in comparison to bare TiO2 NRs. The sensor devices were tested toward very low concentrations of NO2 from 500 ppb to 5 ppm with a response of 135.5 to 15, respectively. The enhancement in NRs sensing performances is due to the catalytic effect of Au doping to the NRs surface that induces surface defects and the coexistence of mixed phases (metastable anatase and thermodynamically stable rutile) of TiO2.
Constructing a junction between two similar or dissimilar materials is another successful method to improve the sensing performance of metal oxide-based devices [60,65]. L. Wang et al. [60] presented an interesting report on TiO2-In2O3 composite nanofibers (NFs) with In2O3 porous beads built using the electrospinning method for the detection of NO2. In particular, they developed a single-step process, followed by calcination, to achieve a nanoparticle–nanofiber structure. The porous single crystal NRs loaded on the composite NFs play a crucial role in the formation of Schottky junctions between the semiconductor materials and the gold electrodes. These junctions reduce the interfacial resistance when charge carriers flow through the contacts, thus increasing the overall conductance (shown in Figure 11). Moreover, the surface defects along with the NFs porosity give birth to a high surface-to-volume ratio which allows the adsorption of gases (O2 or NO2) and the electrons capture more effectively, even in the bulk. The resulting sensing devices exhibit superior performances at RT with a response of 95 toward 97 ppm of NO2.
Even though oxygen is not a toxic gas, the results reported by Wang et al. [95] from fabricating TiO2 NRs array using an acid vapor oxidation (AVO) technique are worth mentioning. The sensing tests were carried out in dry air at RT for O2 concentrations ranging from 1% to 16% (volume ratio). The responses were monotonically increasing with the concentration approaching 2.1 at 16%. At the same concentration, the best tradeoff between response and recovery time (55 s and 51 s) were obtained. These sensors have also demonstrated a good repeatability and selectivity against hydrogen and methane.
Table 5 comprises the literature survey for TiO2 nanostructures and heterostructures together with their growth techniques and sensing performance towards oxidizing gases.

5.3. Effect of Humidity

Generally, the response of the metal oxides gas sensor in the presence of humidity is reduced. The reduction in their gas sensing response in a wet environment occurs due to the competition of adsorption between the analyte and the water molecules. The moisture may absorb on the surface metal sites, reducing oxygen adsorption. In addition, hydrogen produced by the homolytic dissociation of water molecules subsequently reacts with the lattice oxygen of the metal oxides (MOX), forming surface hydroxyl species. Hence, in the presence of humidity, water is in competition with the absorption of hydrogen at the MOX surface. The reduction in the gas sensing response in a wet environment is commonly known as the “poisoning effect of water on metal oxide sensors” [96,97]. Moreover, at elevated temperatures, moisture may also be responsible for a restructuring of the surfaces, modifying the density of oxygen absorption sites.

6. Conclusions

Titanium oxide (TiO2) is a nontoxic and environmentally friendly material in nature with excellent biocompatibility and stability, which allows its nanomaterials to be adopted in a wide range of applications, especially in chemical/gas sensing. The 1D TiO2 nanostructures such as nanowires, nanotubes and nanofibers possess larger specific surface areas that contribute to the unique physical, chemical, optical and electronic properties of this material. TiO2-based 1D nanostructures have been grown using large-scale and high-yield techniques such as hydrothermal, electrochemical anodization and the electrospinning growth method showed exceptional sensing performances for the detection of different gas analytes such as VOCs, toxic and explosive gases. However, in the last five to seven years, great attention has been paid to improving the sensing performance of these nanostructures by doping of metal nanoparticle or by creating a junction of different MOX. These heterostructures show an enhancement in sensitivity, selectivity and lower operating temperature with faster responses and recovery times. The change in sensing performances of composite structures compared to pristine TiO2 is attributed to the heterojunction formation within the interface of two semiconducting materials, as well as catalytical and synergetic effects. Furthermore, the formation of nanostructured material also influences gas adsorption and diffusion, which leads to a change in sensing performances.
Concluding, in the field of chemical/gas sensing, there has been much progress achieved in growth of various 1D TiO2 nanostructures and heterostructures using different fabrication techniques. The effect of different phases of TiO2 on its sensing properties has been widely investigated. In this review, we have presented the recent achievements in the field of TiO2-based 1D nanostructures fabrication and understanding of their sensing mechanism and their sensing performance related to different gas analytes.

Author Contributions

Conceptualization, N.K., M.S., A.M., E.C. and G.D.; writing of the original draft preparation, N.K. and E.C. writing of review and editing, E.C.; validation, E.C.; visualization, E.C.; supervision, E.C.; project administration, E.C.; funding acquisition, N.K., M.S., A.M., E.C. and G.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially funded by MIUR “Smart Cities and Communities and social innovation” project titled “SWaRM Net/Smart Water Resource Management—Neworks and by the NATO Science for Peace and Security Programmer (SPS) under grant G5634 AMOXES—“Advanced Electro-Optical Chemical Sensors”.

Conflicts of Interest

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

References

  1. Comini, E. Metal oxide nanowire chemical sensors: Innovation and quality of life. Mater. Today 2016, 19, 559–567. [Google Scholar] [CrossRef]
  2. Zhou, Z.; Lan, C.; Wei, R.; Ho, J.C. Transparent metal-oxide nanowires and their applications in harsh electronics. J. Mater. Chem. C 2019, 7, 202–217. [Google Scholar] [CrossRef]
  3. Comini, E.; Bianchi, S.; Faglia, G.; Ferroni, M.; Vomiero, A.; Sberveglieri, G. Functional nanowires of tin oxide. Appl. Phys. A 2007, 89, 73–76. [Google Scholar] [CrossRef]
  4. Shen, Y.; Yamazaki, T.; Liu, Z.; Meng, D.; Kikuta, T.; Nakatani, N.; Saito, M.; Mori, M. Microstructure and H2 gas sensing properties of undoped and Pd-doped SnO2 nanowires. Sens. Actuators B Chem. 2009, 135, 524–529. [Google Scholar] [CrossRef]
  5. Suman, P.H.; Felix, A.A.; Tuller, H.L.; Varela, J.A.; Orlandi, M.O. Comparative gas sensor response of SnO2, SnO and Sn3O4 nanobelts to NO2 and potential interferents. Sens. Actuators B Chem. 2015, 208, 122–127. [Google Scholar] [CrossRef]
  6. Sharma, A.P.; Dhakal, P.; Pradhan, D.K.; Behera, M.K.; Xiao, B.; Baha, M. Fabrication and characterization of SnO2 nanorods for room temperature gas sensors. AIP Adv. 2018, 8, 095219. [Google Scholar] [CrossRef] [Green Version]
  7. Forleo, A.; Francioso, L.; Capone, S.; Casino, F.; Siciliano, P.; Tan, O.K.; Hui, H. Wafer-Level Fabrication and Gas Sensing Properties of miniaturized gas sensors based on Inductively Coupled Plasma Deposited Tin Oxide Nanorods. Procedia Chem. 2009, 1, 196–199. [Google Scholar] [CrossRef] [Green Version]
  8. Qian, L.H.; Wang, K.; Li, Y.; Fang, H.T.; Lu, Q.H.; Ma, X.L. CO sensor based on Au-decorated SnO2 nanobelt. Mater. Chem. Phys. 2006, 100, 82–84. [Google Scholar] [CrossRef]
  9. Kaur, N.; Singh, M.; Comini, E. One-Dimensional Nanostructured Oxide Chemoresistive Sensors. Langmuir 2020, 36, 6326–6344. [Google Scholar] [CrossRef]
  10. Kaur, N.; Comini, E.; Zappa, D.; Poli, N.; Sberveglieri, G. Nickel oxide nanowires: Vapor liquid solid synthesis and integration into a gas sensing device. Nanotechnology 2016, 27, 205701. [Google Scholar] [CrossRef]
  11. Kaur, N.; Zappa, D.; Poli, N.; Comini, E. Integration of VLS-Grown WO3 Nanowires into Sensing Devices for the Detection of H2S and O3. ACS Omega 2019, 4, 16336–16343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Asiah, M.N.; Mamat, M.H.; Khusaimi, Z.; Achoi, M.F.; Abdullah, S.; Rusop, M. Thermal stability and phase transformation of TiO2 nanowires at various temperatures. Microelectron. Eng. 2013, 108, 134–137. [Google Scholar] [CrossRef]
  13. Fàbrega, C.; Casals, O.; Hernández-Ramírez, F.; Prades, J.D. A review on efficient self-heating in nanowire sensors: Prospects for very-low power devices. Sens. Actuators B Chem. 2018, 256, 797–811. [Google Scholar] [CrossRef] [Green Version]
  14. Baratto, C.; Comini, E.; Ferroni, M.; Faglia, G.; Sberveglieri, G. Plasma-induced enhancement of UV photoluminescence in ZnO nanowires. CrystEngComm 2013, 15, 7981. [Google Scholar] [CrossRef]
  15. Kaur, N.; Comini, E.; Poli, N.; Zappa, D.; Sberveglieri, G. Nickel Oxide Nanowires Growth by VLS Technique for Gas Sensing Application. Procedia Eng. 2015, 120, 760–763. [Google Scholar] [CrossRef]
  16. Topcu, S.; Jodhani, G.; Gouma, P. Optimized Nanostructured TiO2 Photocatalysts. Front. Mater. 2016, 3, 35. [Google Scholar] [CrossRef] [Green Version]
  17. Mavrič, T.; Benčina, M.; Imani, R.; Junkar, I.; Valant, M.; Kralj-Iglič, V.; Iglič, A. Electrochemical Biosensor Based on TiO2 Nanomaterials for Cancer Diagnostics. In Advances in Biomembranes and Lipid Self-Assembly; Elsevier B.V.: Cambridge, MA, USA, 2018; Volume 27, pp. 63–105. [Google Scholar]
  18. Maziarz, W.; Kusior, A.; Trenczek-Zajac, A. Nanostructured TiO2-based gas sensors with enhancedsensitivity to reducing gases. Beilstein J. Nanotechnol. 2016, 7, 1718–1726. [Google Scholar] [CrossRef] [Green Version]
  19. Bai, Y.; Mora-Seró, I.; De Angelis, F.; Bisquert, J.; Wang, P. Titanium Dioxide Nanomaterials for Photovoltaic Applications. Chem. Rev. 2014, 114, 10095–10130. [Google Scholar] [CrossRef]
  20. Wang, Y.; Zu, M.; Zhou, X.; Lin, H.; Peng, F.; Zhang, S. Designing efficient TiO2-based photoelectrocatalysis systems for chemical engineering and sensing. Chem. Eng. J. 2020, 381, 122605. [Google Scholar] [CrossRef]
  21. Mardare, D.; Cornei, N.; Mita, C.; Florea, D.; Stancu, A.; Tiron, V.; Manole, A.; Adomnitei, C. Low temperature TiO2 based gas sensors for CO2. Ceram. Int. 2016, 42, 7353–7359. [Google Scholar] [CrossRef]
  22. Cui, H.-F.; Wu, W.-W.; Li, M.-M.; Song, X.; Lv, Y.; Zhang, T.-T. A highly stable acetylcholinesterase biosensor based on chitosan-TiO2-graphene nanocomposites for detection of organophosphate pesticides. Biosens. Bioelectron. 2018, 99, 223–229. [Google Scholar] [CrossRef] [PubMed]
  23. Kim, Y.C.; Lee, K.H.; Sasaki, S.; Hashimoto, K.; Ikebukuro, K.; Karube, I. Photocatalytic sensor for chemical oxygen demand determination based on oxygen electrode. Anal. Chem. 2000, 72, 3379–3382. [Google Scholar] [CrossRef] [PubMed]
  24. Li, X.; Yin, W.; Li, J.; Bai, J.; Huang, K.; Li, J.; Zhou, B. TiO2 Nanotube Sensor for Online Chemical Oxygen Demand Determination in Conjunction with Flow Injection Technique. Water Environ. Res. 2014, 86, 532–539. [Google Scholar] [CrossRef] [PubMed]
  25. Bai, J.; Zhou, B. Titanium Dioxide Nanomaterials for Sensor Applications. Chem. Rev. 2014, 114, 10131–10176. [Google Scholar] [CrossRef] [PubMed]
  26. Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53–229. [Google Scholar] [CrossRef]
  27. Kumar, S.G.; Rao, K.S.R.K. Polymorphic phase transition among the titania crystal structures using a solution-based approach: From precursor chemistry to nucleation process. Nanoscale 2014, 6, 11574–11632. [Google Scholar] [CrossRef]
  28. Xu, J.; Wu, S.; Jin, J.; Peng, T. Preparation of brookite TiO2 nanoparticles with small sizes and the improved photovoltaic performance of brookite-based dye-sensitized solar cells. Nanoscale 2016, 8, 18771–18781. [Google Scholar] [CrossRef]
  29. Zhang, W.; Shen, D.; Liu, Z.; Wu, N.-L.; Wei, M. Brookite TiO2 mesocrystals with enhanced lithium-ion intercalation properties. Chem. Commun. 2018, 54, 11491–11494. [Google Scholar] [CrossRef]
  30. Gupta, S.M.; Tripathi, M. A review of TiO2 nanoparticles. Chinese Sci. Bull. 2011, 56, 1639. [Google Scholar] [CrossRef] [Green Version]
  31. El Goresy, A.; Chen, M.; Gillet, P.; Dubrovinsky, L.; Graup, G.; Ahuja, R. A natural shock-induced dense polymorph of rutile with α-PbO2 structure in the suevite from the Ries crater in Germany. Earth Planet. Sci. Lett. 2001, 192, 485–495. [Google Scholar] [CrossRef]
  32. Karthick, S.N.; Hemalatha, K.V.; Balasingam, S.K.; Manik Clinton, F.; Akshaya, S.; Kim, H. Dye-Sensitized Solar Cells: History, Components, Configuration, and Working Principle. In Interfacial Engineering in Functional Materials for Dye-Sensitized Solar Cells; Wiley: Hoboken, NJ, USA, 2019; pp. 1–16. [Google Scholar]
  33. Van de Krol, R. Erratum: “Mott-Schottky Analysis of Nanometer-Scale Thin-Film Anatase TiO2” [J. Electrochem. Soc., 144, 1723 (1997)]. J. Electrochem. Soc. 1998, 145, 3697. [Google Scholar] [CrossRef]
  34. Carp, O.; Huisman, C.L.; Reller, A. Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 2004, 32, 33–177. [Google Scholar] [CrossRef]
  35. Shi, J.; Wang, X. Growth of rutile titanium dioxide nanowires by pulsed chemical vapor deposition. Cryst. Growth Des. 2011, 11, 949–954. [Google Scholar] [CrossRef]
  36. Kemell, M.; Pore, V.; Tupala, J.; Ritala, M.; Leskelä, M. Atomic layer deposition of nanostructured TiO2 photocatalysts via template approach. Chem. Mater. 2007, 19, 1816–1820. [Google Scholar] [CrossRef]
  37. Maijenburg, A.W.; Veerbeek, J.; De Putter, R.; Veldhuis, S.A.; Zoontjes, M.G.C.; Mul, G.; Montero-Moreno, J.M.; Nielsch, K.; Schäfer, H.; Steinhart, M.; et al. Electrochemical synthesis of coaxial TiO2-Ag nanowires and their application in photocatalytic water splitting. J. Mater. Chem. A 2014, 2, 2648–2656. [Google Scholar] [CrossRef] [Green Version]
  38. Zhao, B.; Lin, L.; He, D. Phase and morphological transitions of titania/titanate nanostructures from an acid to an alkali hydrothermal environment. J. Mater. Chem. A 2013, 1, 1659–1668. [Google Scholar] [CrossRef]
  39. Rodríguez-Reyes, M.; Dorantes-Rosales, H.J. A simple route to obtain TiO2 nanowires by the sol-gel method. J. Sol Gel Sci. Technol. 2011, 59, 658–661. [Google Scholar] [CrossRef]
  40. Nikfarjam, A.; Salehifar, N. Improvement in gas-sensing properties of TiO2 nanofiber sensor by UV irradiation. Sens. Actuators B Chem. 2015, 211, 146–156. [Google Scholar] [CrossRef]
  41. Cushing, B.L.; Kolesnichenko, V.L.; O’Connor, C.J. Recent advances in the liquid-phase syntheses of inorganic nanoparticles. Chem. Rev. 2004, 104, 3893–3946. [Google Scholar] [CrossRef]
  42. Alev, O.; Şennik, E.; Kilinç, N.; Öztürk, Z.Z. Gas sensor application of hydrothermally growth TiO2 nanorods. Procedia Eng. 2015, 120, 1162–1165. [Google Scholar] [CrossRef] [Green Version]
  43. Yang, H.Y.; Cheng, X.L.; Zhang, X.F.; Zheng, Z.K.; Tang, X.F.; Xu, Y.M.; Gao, S.; Zhao, H.; Huo, L.H. A novel sensor for fast detection of triethylamine based on rutile TiO2 nanorod arrays. Sens. Actuators B Chem. 2014, 205, 322–328. [Google Scholar] [CrossRef]
  44. Zhao, G.; Xuan, J.; Gong, Q.; Wang, L.; Ren, J.; Sun, M.; Jia, F.; Yin, G.; Liu, B. In Situ Growing Double-Layer TiO2 Nanorod Arrays on New-Type FTO Electrodes for Low-Concentration NH3 Detection at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 8573–8582. [Google Scholar] [CrossRef] [PubMed]
  45. Ren, S.; Liu, W. One-step photochemical deposition of PdAu alloyed nanoparticles on TiO2 nanowires for ultra-sensitive H2 detection. J. Mater. Chem. A 2016, 4, 2236–2245. [Google Scholar] [CrossRef]
  46. Alam, U.; Fleisch, M.; Kretschmer, I.; Bahnemann, D.; Muneer, M. One-step hydrothermal synthesis of Bi-TiO2 nanotube/graphene composites: An efficient photocatalyst for spectacular degradation of organic pollutants under visible light irradiation. Appl. Catal. B Environ. 2017, 218, 758–769. [Google Scholar] [CrossRef]
  47. Ng, S.; Kuberský, P.; Krbal, M.; Prikryl, J.; Gärtnerová, V.; Moravcová, D.; Sopha, H.; Zazpe, R.; Yam, F.K.; Jäger, A.; et al. ZnO Coated Anodic 1D TiO2 Nanotube Layers: Efficient Photo-Electrochemical and Gas Sensing Heterojunction. Adv. Eng. Mater. 2018, 20, 1700589. [Google Scholar] [CrossRef] [Green Version]
  48. Alev, O.; Kılıç, A.; Çakırlar, Ç.; Büyükköse, S.; Öztürk, Z. Gas Sensing Properties of p-Co3O4/n-TiO2 Nanotube Heterostructures. Sensors 2018, 18, 956. [Google Scholar] [CrossRef] [Green Version]
  49. Alev, O.; Şennik, E.; Öztürk, Z.Z. Improved gas sensing performance of p-copper oxide thin film/n-TiO2 nanotubes heterostructure. J. Alloys Compd. 2018, 749, 221–228. [Google Scholar] [CrossRef]
  50. Zhao, P.X.; Tang, Y.; Mao, J.; Chen, Y.X.; Song, H.; Wang, J.W.; Song, Y.; Liang, Y.Q.; Zhang, X.M. One-Dimensional MoS2-Decorated TiO2 nanotube gas sensors for efficient alcohol sensing. J. Alloys Compd. 2016, 674, 252–258. [Google Scholar] [CrossRef]
  51. Sopha, H.; Hromadko, L.; Nechvilova, K.; Macak, J.M. Effect of electrolyte age and potential changes on the morphology of TiO2 nanotubes. J. Electroanal. Chem. 2015, 759, 122–128. [Google Scholar] [CrossRef]
  52. Varghese, O.K.; Paulose, M.; Grimes, C.A. Long vertically aligned titania nanotubes on transparent conducting oxide for highly efficient solar cells. Nat. Nanotechnol. 2009, 4, 592–597. [Google Scholar] [CrossRef]
  53. Munirathinam, B.; Pydimukkala, H.; Ramaswamy, N.; Neelakantan, L. Influence of crystallite size and surface morphology on electrochemical properties of annealed TiO2 nanotubes. Appl. Surf. Sci. 2015, 355, 1245–1253. [Google Scholar] [CrossRef]
  54. Mazierski, P.; Nischk, M.; Gołkowska, M.; Lisowski, W.; Gazda, M.; Winiarski, M.J.; Klimczuk, T.; Zaleska-Medynska, A. Photocatalytic activity of nitrogen doped TiO2 nanotubes prepared by anodic oxidation: The effect of applied voltage, anodization time and amount of nitrogen dopant. Appl. Catal. B Environ. 2016, 196, 77–88. [Google Scholar] [CrossRef]
  55. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef]
  56. Zhou, M.; Liu, Y.; Wu, B.; Zhang, X. Different crystalline phases of aligned TiO2 nanowires and their ethanol gas sensing properties. Phys. E Low Dimens. Syst. Nanostruct. 2019, 114, 113601. [Google Scholar] [CrossRef]
  57. Wang, S.; Lin, Z.X.; Wang, W.H.; Kuo, C.L.; Hwang, K.C.; Hong, C.C. Self-regenerating photocatalytic sensor based on dielectrophoretically assembled TiO2 nanowires for chemical vapor sensing. Sens. Actuators B Chem. 2014, 194, 1–9. [Google Scholar] [CrossRef]
  58. Li, G.; Zhang, X.; Lu, H.; Yan, C.; Chen, K.; Lu, H.; Gao, J.; Yang, Z.; Zhu, G.; Wang, C.; et al. Ethanol sensing properties and reduced sensor resistance using porous Nb2O5-TiO2 n-n junction nanofibers. Sens. Actuators B Chem. 2019, 283, 602–612. [Google Scholar] [CrossRef]
  59. 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]
  60. Wang, L.; Gao, J.; Wu, B.; Kan, K.; Xu, S.; Xie, Y.; Li, L.; Shi, K. Designed Synthesis of In2O3 Beads@TiO2-In2O3 Composite Nanofibers for High Performance NO2 Sensor at Room Temperature. ACS Appl. Mater. Interfaces 2015, 7, 27152–27159. [Google Scholar] [CrossRef]
  61. Arafat, M.M.; Haseeb, A.S.M.A.; Akbar, S.A.; Quadir, M.Z. In-situ fabricated gas sensors based on one dimensional core-shell TiO2-Al2O3 nanostructures. Sens. Actuators B Chem. 2017, 238, 972–984. [Google Scholar] [CrossRef]
  62. Wang, H.; Yao, Y.; Wu, G.; Sun, Q.; Wang, M.; Chen, X.; Wang, J. A room temperature oxygen gas sensor based on hierarchical TiO2. In Proceedings of the IEEE 3M-NANO 2016–6th IEEE International Conference on Manipulation, Manufacturing and Measurement on the Nanoscale, Chongqing, China, 18–22 July 2016; Institute of Electrical and Electronics Engineers Inc.: Piscataway, NJ, USA; pp. 199–202. [Google Scholar]
  63. Lee, J.S.; Katoch, A.; Kim, J.H.; Kim, S.S. Growth of networked TiO2 nanowires for gas-sensing applications. J. Nanosci. Nanotechnol. 2016, 16, 11580–11585. [Google Scholar] [CrossRef]
  64. Tshabalala, Z.P.; Shingange, K.; Dhonge, B.P.; Ntwaeaborwa, O.M.; Mhlongo, G.H.; Motaung, D.E. Fabrication of ultra-high sensitive and selective CH4 room temperature gas sensing of TiO2 nanorods: Detailed study on the annealing temperature. Sens. Actuators B Chem. 2017, 238, 402–419. [Google Scholar] [CrossRef]
  65. 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]
  66. Zhu, Z.; Lin, S.J.; Wu, C.H.; Wu, R.J. Synthesis of TiO2 nanowires for rapid NO2 detection. Sens. Actuators A Phys. 2018, 272, 288–294. [Google Scholar] [CrossRef]
  67. Enachi, M.; Lupan, O.; Braniste, T.; Sarua, A.; Chow, L.; Mishra, Y.K.; Gedamu, D.; Adelung, R.; Tiginyanu, I. Integration of individual TiO 2 nanotube on the chip: Nanodevice for hydrogen sensing. Phys. Status Solidi—Rapid Res. Lett. 2015, 9, 171–174. [Google Scholar] [CrossRef]
  68. Kumar, K.G.; Avinash, B.S.; Rahimi-Gorji, M.; Majdoubi, J. Photocatalytic activity and smartness of TiO2 nanotube arrays for room temperature acetone sensing. J. Mol. Liq. 2020, 300, 112418. [Google Scholar] [CrossRef]
  69. Moon, J.; Hedman, H.P.; Kemell, M.; Tuominen, A.; Punkkinen, R. Hydrogen sensor of Pd-decorated tubular TiO2 layer prepared by anodization with patterned electrodes on SiO2/Si substrate. Sens. Actuators B Chem. 2016, 222, 190–197. [Google Scholar] [CrossRef]
  70. Liu, Y.; Wang, L.; Wang, H.; Xiong, M.; Yang, T.; Zakharova, G.S. Highly sensitive and selective ammonia gas sensors based on PbS quantum dots/TiO2 nanotube arrays at room temperature. Sens. Actuators B Chem. 2016, 236, 529–536. [Google Scholar] [CrossRef]
  71. Bhattacharyya, P.; Bhowmik, B.; Fecht, H.J. Operating Temperature, Repeatability, and Selectivity of TiO2 Nanotube-Based Acetone Sensor: Influence of Pd and Ni Nanoparticle Modifications. IEEE Trans. Device Mater. Reliab. 2015, 15, 376–383. [Google Scholar] [CrossRef]
  72. Perillo, P.M.; Rodríguez, D.F. TiO2 nanotubes membrane flexible sensor for low-temperature H2S detection. Chemosensors 2016, 4, 15. [Google Scholar] [CrossRef] [Green Version]
  73. Tong, X.; Shen, W.; Chen, X. Enhanced H2S sensing performance of cobalt doped free-standing TiO2 nanotube array film and theoretical simulation based on density functional theory. Appl. Surf. Sci. 2019, 469, 414–422. [Google Scholar] [CrossRef]
  74. Wang, Y.; Zhou, Y.; Meng, C.; Gao, Z.; Cao, X.; Li, X.; Xu, L.; Zhu, W.; Peng, X.; Zhang, B.; et al. A high-response ethanol gas sensor based on one-dimensional TiO2/V2O5 branched nanoheterostructures. Nanotechnology 2016, 27, 425503. [Google Scholar] [CrossRef] [PubMed]
  75. Şennik, E.; Alev, O.; Öztürk, Z.Z. The effect of Pd on the H2 and VOC sensing properties of TiO2 nanorods. Sens. Actuators B Chem. 2016, 229, 692–700. [Google Scholar] [CrossRef]
  76. Kusior, A.; Radecka, M.; Zych, L.; Zakrzewska, K.; Reszka, A.; Kowalski, B.J. Sensitization of TiO2/SnO2 nanocomposites for gas detection. Sens. Actuators B Chem. 2013, 189, 251–259. [Google Scholar] [CrossRef]
  77. Zakrzewska, K. Gas sensing mechanism of TiO2-based thin films. Pergamon 2004, 74, 335–338. [Google Scholar] [CrossRef]
  78. 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]
  79. Li, Z.; Li, H.; Wu, Z.; Wang, M.; Luo, J.; Torun, H.; Hu, P.; Yang, C.; Grundmann, M.; Liu, X.; et al. Advances in designs and mechanisms of semiconducting metal oxide nanostructures for high-precision gas sensors operated at room temperature. Mater. Horizons 2019, 6, 470–506. [Google Scholar] [CrossRef] [Green Version]
  80. Göpel, W.; Schierbaum, K.D. SnO2 sensors: Current status and future prospects. Sens. Actuators B Chem. 1995, 26, 1–12. [Google Scholar] [CrossRef]
  81. Williams, D.E. Semiconducting oxides as gas-sensitive resistors. Sens. Actuators B Chem. 1999, 57, 1–16. [Google Scholar] [CrossRef]
  82. Staffell, I.; Scamman, D.; Velazquez Abad, A.; Balcombe, P.; Dodds, P.E.; Ekins, P.; Shah, N.; Ward, K.R. The role of hydrogen and fuel cells in the global energy system. Energy Environ. Sci. 2019, 12, 463–491. [Google Scholar] [CrossRef] [Green Version]
  83. Sabri, Y.M.; Kandjani, A.E.; Rashid, S.S.A.A.H.; Harrison, C.J.; Ippolito, S.J.; Bhargava, S.K. Soot template TiO2 fractals as a photoactive gas sensor for acetone detection. Sens. Actuators B Chem. 2018, 275, 215–222. [Google Scholar] [CrossRef]
  84. Avinash, B.S.; Chaturmukha, V.S.; Harish, B.M.; Jayanna, H.S.; Rajeeva, M.P.; Naveen, C.S.; Lamani, A.R. Synthesis, Characterization and Room Temperature Acetone Sensing of TiO2 Nanotubes. Sens. Lett. 2018, 16, 105–109. [Google Scholar] [CrossRef]
  85. Hazra, A.; Bhowmik, B.; Dutta, K.; Chattopadhyay, P.P.; Bhattacharyya, P. Stoichiometry, length, and wall thickness optimization of TiO2 nanotube array for efficient alcohol sensing. ACS Appl. Mater. Interfaces 2015, 7, 9336–9348. [Google Scholar] [CrossRef] [PubMed]
  86. Bian, H.; Ma, S.; Sun, A.; Xu, X.; Yang, G.; Gao, J.; Zhang, Z.; Zhu, H. Characterization and acetone gas sensing properties of electrospun TiO2 nanorods. Superlattices Microstruct. 2015, 81, 107–113. [Google Scholar] [CrossRef]
  87. 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] [Green Version]
  88. Pandey, S.K.; Kim, K.H.; Tang, K.T. A review of sensor-based methods for monitoring hydrogen sulfide. TrAC Trends Anal. Chem. 2012, 32, 87–99. [Google Scholar] [CrossRef]
  89. Moon, C.H.; Zhang, M.; Myung, N.V.; Haberer, E.D. Highly sensitive hydrogen sulfide (H2S) gas sensors from viral-templated nanocrystalline gold nanowires. Nanotechnology 2014, 25, 135205. [Google Scholar] [CrossRef]
  90. Carslaw, D.C.; Farren, N.J.; Vaughan, A.R.; Drysdale, W.S.; Young, S.; Lee, J.D. The diminishing importance of nitrogen dioxide emissions from road vehicle exhaust. Atmos. Environ. X 2019, 1, 100002. [Google Scholar] [CrossRef]
  91. Predic, B.; Yan, Z.; Eberle, J.; Stojanovic, D.; Aberer, K. ExposureSense: Integrating daily activities with air quality using mobile participatory sensing. In Proceedings of the 2013 IEEE International Conference on Pervasive Computing and Communications Workshops, PerCom Workshops, San Diego, CA, USA, 18–22 March 2013; pp. 303–305. [Google Scholar]
  92. Kaur, N.; Zappa, D.; Comini, E. Shelf Life Study of NiO Nanowire Sensors for NO2 Detection. Electron. Mater. Lett. 2019, 15, 743–749. [Google Scholar] [CrossRef]
  93. Engineer, C.B.; Hashimoto-Sugimoto, M.; Negi, J.; Israelsson-Nordström, M.; Azoulay-Shemer, T.; Rappel, W.J.; Iba, K.; Schroeder, J.I. CO2 Sensing and CO2 Regulation of Stomatal Conductance: Advances and Open Questions. Trends Plant Sci. 2016, 21, 16–30. [Google Scholar] [CrossRef] [Green Version]
  94. Ponnuvelu, D.V.; Pullithadathil, B.; Prasad, A.K.; Dhara, S.; Mohamed, K.; Tyagi, A.K.; Raj, B. Highly sensitive, atmospheric pressure operatable sensor based on Au nanoclusters decorated TiO2@Au heterojunction nanorods for trace level NO2 gas detection. J. Mater. Sci. Mater. Electron. 2017, 28, 9738–9748. [Google Scholar] [CrossRef]
  95. Wang, H.; Sun, Q.; Yao, Y.; Li, Y.; Wang, J.; Chen, L. A micro sensor based on TiO2 nanorod arrays for the detection of oxygen at room temperature. Ceram. Int. 2016, 42, 8565–8571. [Google Scholar] [CrossRef]
  96. Raza, M.H.; Kaur, N.; Comini, E.; Pinna, N. Toward Optimized Radial Modulation of the Space-Charge Region in One-Dimensional SnO2-NiO Core-Shell Nanowires for Hydrogen Sensing. ACS Appl. Mater. Interfaces 2020, 12, 4594–4606. [Google Scholar] [CrossRef] [PubMed]
  97. Fomekong, R.L.; Saruhan, B. Influence of humidity on NO2-sensing and selectivity of spray-CVD grown ZnO thin film above 400 °C. Chemosensors 2019, 7, 42. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Different crystal structures of titanium dioxide (anatase, rutile and brookite). Reprinted with permission from ref. [32] WILEY Copyright 2019.
Figure 1. Different crystal structures of titanium dioxide (anatase, rutile and brookite). Reprinted with permission from ref. [32] WILEY Copyright 2019.
Materials 13 02974 g001
Figure 2. FESEM images of the as-prepared nanowires (a) and nanowires annealed at (b) 400 °C, (c) 500 °C, (d) 600 °C, (e) 700 °C and (f) 900 °C. Reprinted with permission from Ref. [12] Elsevier Copyright 2013.
Figure 2. FESEM images of the as-prepared nanowires (a) and nanowires annealed at (b) 400 °C, (c) 500 °C, (d) 600 °C, (e) 700 °C and (f) 900 °C. Reprinted with permission from Ref. [12] Elsevier Copyright 2013.
Materials 13 02974 g002
Figure 3. The effect of acid/alkali concentration. Rutile TiO2, anatase TiO2, brookite TiO2, dititanate Na2Ti2O5 and trititanate Na2Ti3O7 are represented by “R, A, B, T2 and T3”, respectively. Reprinted with permission from Ref. [38] RSC Copyright 2012.
Figure 3. The effect of acid/alkali concentration. Rutile TiO2, anatase TiO2, brookite TiO2, dititanate Na2Ti2O5 and trititanate Na2Ti3O7 are represented by “R, A, B, T2 and T3”, respectively. Reprinted with permission from Ref. [38] RSC Copyright 2012.
Materials 13 02974 g003
Figure 4. SEM images of TiO2 nanorods prepared using the hydrothermal method. (a) Top and (b) cross-sectional views of the nanorods. Reprinted with permission from Ref. [44] ACS Copyright 2020.
Figure 4. SEM images of TiO2 nanorods prepared using the hydrothermal method. (a) Top and (b) cross-sectional views of the nanorods. Reprinted with permission from Ref. [44] ACS Copyright 2020.
Materials 13 02974 g004
Figure 5. Cross-sectional images of TiO2 nanotubes synthesized by applying 60 V for 6 h using different electrolyte ages: (a) fresh electrolyte, (b) 6 h, (c) 25 h and (d) 50 h. Reprinted with permission from Ref. [51] Copyright 2015 Elsevier.
Figure 5. Cross-sectional images of TiO2 nanotubes synthesized by applying 60 V for 6 h using different electrolyte ages: (a) fresh electrolyte, (b) 6 h, (c) 25 h and (d) 50 h. Reprinted with permission from Ref. [51] Copyright 2015 Elsevier.
Materials 13 02974 g005
Figure 6. Schematic diagram of the electrospinning equipment (a) vertical and (b) horizontal setups. Reprinted with permission from Ref. [55], Copyright 2013 Elsevier.
Figure 6. Schematic diagram of the electrospinning equipment (a) vertical and (b) horizontal setups. Reprinted with permission from Ref. [55], Copyright 2013 Elsevier.
Materials 13 02974 g006
Figure 7. SEM images of aligned TiO2 NWs calcinated at different temperatures: (a) 500 °C, (b) 600 °C, (c) 700 °C and (d) 800 °C. Reprinted with permission from Ref. [56], Copyright 2019 Elsevier.
Figure 7. SEM images of aligned TiO2 NWs calcinated at different temperatures: (a) 500 °C, (b) 600 °C, (c) 700 °C and (d) 800 °C. Reprinted with permission from Ref. [56], Copyright 2019 Elsevier.
Materials 13 02974 g007
Figure 8. Schematic representation of gas sensing at different modes.
Figure 8. Schematic representation of gas sensing at different modes.
Materials 13 02974 g008
Figure 9. (a) Dynamic response of TiO2 nanorods at 200 °C. (b) Dynamic response of Pd-TiO2 at 200 °C toward different H2 concentrations. (c) Dynamic response of Pd-TiO2 at 30 °C toward 100 ppm of H2 (d) Sensor responses of TiO2 and Pd TiO2 nanorod device for H2 and VOCs at 200 °C. Reprinted with permission from Ref. [75] Copyright 2016 Elsevier.
Figure 9. (a) Dynamic response of TiO2 nanorods at 200 °C. (b) Dynamic response of Pd-TiO2 at 200 °C toward different H2 concentrations. (c) Dynamic response of Pd-TiO2 at 30 °C toward 100 ppm of H2 (d) Sensor responses of TiO2 and Pd TiO2 nanorod device for H2 and VOCs at 200 °C. Reprinted with permission from Ref. [75] Copyright 2016 Elsevier.
Materials 13 02974 g009
Figure 10. Ethanol sensing responses of TiO2 NTs and TiO2/ZnO (0.19, 1.9, 7.6 and 19 nm) core/shell (a) at 100 °C and (b) at 200 °C. Reprinted with permission from Ref. [47] Copyright 2018 WILEY.
Figure 10. Ethanol sensing responses of TiO2 NTs and TiO2/ZnO (0.19, 1.9, 7.6 and 19 nm) core/shell (a) at 100 °C and (b) at 200 °C. Reprinted with permission from Ref. [47] Copyright 2018 WILEY.
Materials 13 02974 g010
Figure 11. (a) Energy band diagram of In2O3 and TiO2, EC: conduction band and EV: valence band; (b) I-V curves measured for INFs and TINF2 thin film sensors in air at RT, which were treated at 110°C (the gate voltage Vg = 0.1); (c) The gas sensing reactions based on Schottky junction between Au electrode and In2O3 beads. Reprinted with permission from Ref. [46] Copyright 2015 American Chemical Society.
Figure 11. (a) Energy band diagram of In2O3 and TiO2, EC: conduction band and EV: valence band; (b) I-V curves measured for INFs and TINF2 thin film sensors in air at RT, which were treated at 110°C (the gate voltage Vg = 0.1); (c) The gas sensing reactions based on Schottky junction between Au electrode and In2O3 beads. Reprinted with permission from Ref. [46] Copyright 2015 American Chemical Society.
Materials 13 02974 g011
Table 1. Crystal structure data of the three main polymorphs of TiO2 [26,33,34].
Table 1. Crystal structure data of the three main polymorphs of TiO2 [26,33,34].
Crystal StructureSystemSpace GroupLattice Constants (nm)
abcc/a
RutileTetragonal D 4 h 14 —P42/mmm0.4584-0.29530.644
AnataseTetragonal D 4 h 14 —I41/amd0.3733-0.9372.51
BrookiteRhombohedral D 2 h 15 —Pbca0.54360.9166-0.944
Table 2. Summarization of different materials, parameters, crystal phase involved in the growth process of TiO2 nanostructures and its composites for different fabrication techniques.
Table 2. Summarization of different materials, parameters, crystal phase involved in the growth process of TiO2 nanostructures and its composites for different fabrication techniques.
Hydrothermal
Metal OxideMaterials/ParametersNanostructure TypeSizedCrystalline PhasesRef./Year
TiO2Titanium butoxide, hydrochloric acid and deionized water/
Temp. = 150 °C
Time = 18 h
Nanorodslength = 5 μm diameter = 100 nmrutile [42]/2015
TiO2Isopropyl titanate, hydrochloric acid
Temp. = 150 °C
Time = 3, 6, and 9 h
Nanorodslength = 4200 nm diameter = 120 nm rutile[44]/2020
TiO2TiCl4, HCl and deionized water Temp. = 180 °C
Time = 3 h
Nanorodsdiameter = 21.20 nm
length = NA
anatase/rutile mixed phases[64]/2017
TiO2–Ag2OTiCl4, deionized water
Temp. = 180 °C
Time = 3 h
Composite Nanorodsdiameter = 100 nmrutile[65]/2019
Bi-TiO2[Bi(NO3)3·5H2O] and TiO2 nanoparticles with NaOH and water.
Temp. = 140 °C
Time = 24 h
Nanotubesdiameter = NA
length = NA
anatase[46]/2017
TiO2Titanium (IV) isopropoxide, NaOH and ethanol.
Temp. = 150 °C
Time = 20 h
Nanowireslength = 1825 nm
diameter = 30–50 nm
anatase[66]/(2018)
Electrochemical Anodization
TiO2Ti foil in ethylene glycol, acide phosphorique (H3PO4) and hydrofluoric acid HF.
Anodization = 120 V
Temp. = RT
Time = 2 h
Nanotubeslength = 20 µm diameter = 80–120 nmanatase/rutile mixed phases[67]/2015
7TiO2Ti foil in ethylene glycol, deionized water and ammonium fluoride (NH4F)
Anodization = 1
From 0 V to 60 V
Time = 4 h
Temp. = RT
Nanotubeslength = 50 μm diameter = 120 nmanatase[51]/2015
TiO2Ti foil in NH4F and (NH4)2SO4 and deionized water
Anodization = 20 V
Time = 2 h
Temp. = RT
Thermal annealing at 750 °C for 4 h in air
Nanotubeslength = NA
diameter = 50–80 nm
anatase[68]/2020
p-Co3O4/n-TiO2Ti foil, NH4F ethylene glycol, cobalt (II) nitrate hexahydrate and water
Anodization = 50 V
Time = 1 h
Temp. = 500 °C
Nanotubeslength = NA
diameter = 80 nm
anatase[48]/2018
Pd decorated TiO2 Ti foil in NH4F and ethylene glycol. Anodization = 60 V
Time = 45 min
Temp. = RT
Thermal annealing at 6 h at 500 °C in ambient air
Nanotubeslength = NA
diameter = 40 nm
anatase[69]/2016
PbS quantum dots/TiO2Ti foil in ethylene glycol solution containing NH4F and H2O
Pb(NO3)2 dissolved in methanol and Na2S in methanol.
Anodization = 50 V
Time = 5 h
Temp. = RT
Nanotubeslength = 18 μm
diameter = 80 nm
anatase[70]/2016
Ni and Pd-modified TiO2Ti foil, NH4F, ethylene glycol and water.
NiCl2 and PdCl2
Anodization = 20 V
Temp. = 300 C
Time = 3 h
Nanotubesdiameter = 50–60 nm
length = 470 nm
anatase[71]/2015
TiO2 Ti foil and NH4F with H2O in ethylene glycol
Anodization = 50 V
Time = 3 h
Temp. = RT
Nanotubesdiameter = 100 nm
length = NA
anatase[72]/2016
Co-doped TiO2 Ti foil in glycol solution with ammonium fluoride and deionized water.
Anodization = 30 V
Time 2 h
Temp. = RT
Nanotubesdiameter = 110 nm
length = NA
anatase[73]/2019
Electrospinning
TiO2Titanium tetraisopropoxid (C12H28O4Ti), ethanol, acetic acid Stirring time = 15 min Voltage = 18 kV Rate = 2 mL/min
Calcination temp. = 300
°C, 500 °C, 700 °C, 900 °C
Nanofiberslengths = NA
diameters = 50 nm, 80 nm, 130 nm, 200 nm
anatase/rutile mixed phase[40]/2015
In2O3 beads @ TiO2-In2O3 compositeTBT (tetrabutyl titanate), indium nitrate hydrate, ethanol, DMF (dimethylformamide), PVP (polyvinylpyrrolidone)
Stirring time = 6 h
Voltage = 16.0 kV
Rate = 0.25 mL·h−1
Nanofiberslengths = tens of µm diameter = 150–250 nmpolycrystalline TiO2-In2O3 composite[60]/(2015)
Nb2O5-TiO2Titanium isopropoxide, polyvinylpyrrolidone, acetic acid, ethanol stirring time = 12 h
Voltage = 18.0 kV, −4.0 kV
Rate = 1.5 mL/h,
Calcination temp. = 500 °C
Nanofiberslengths = NA
diameter = 121.3 nm
anatase/ rutile
mixed phase
[58]/2019
TiO2/V2O5Tetrabutyl titanate, poly-vinylpyrrolidone, ethanol, acetic acid stirred time = 20 min Ti/V molar ratio of 4:1
Annealing Temp = 450 °C
Nanofiberslength = µm range
diameter = 60 nm
anatase/ rutile
mixed phase
[74]/2016
TiO2-SnO2Dimethylformamide, ethanol
stirring time = 10 min
SnCl2·2H2O stirring
time = 8 h
voltage = 18 kV
NanofibersN/Arutile[59]/2017
Table 3. Advantages and disadvantages of hydrothermal, electrochemical anodization and electrospinning techniques.
Table 3. Advantages and disadvantages of hydrothermal, electrochemical anodization and electrospinning techniques.
HydrothermalElectrochemical AnodizationElectrospinning
AdvantagesDisadvantagesAdvantagesDisadvantagesAdvantagesDisadvantages
Simple, easy and low-cost synthesis methodThe reaction takes long timeProduction of High quality of 1D nanostructures specially nanotubes Mainly used for nanotubes growthEase to fabricated compositesLimited control of structure porosity
Production of High quality of 1D nanostructures specially nanorodsUtilization of highly concentrated NaOH solutionOrdered and aligned structure The mass-produced is limitedHigh efficiency Use of toxic solvents
The morphology is controlled by synthesis parametersDifficult in achieving uniform sizehigh aspect ratio (length/diameter ratio)Utilization of toxic electrolyte: Hydrofluoric acid HFProcess simplicity-
Easy addition of additives for doping length/diameter ratio is smaller than the ratio produced by electrochemical anodizationGrowth at room temperaturehigh production costMass production-
--Aspect ratio controlled by synthesis parametersDifficulties in separation of film from substrate--
Table 4. Summary of the chemical/gas sensing properties of TiO2-based nanostructures and heterostructures towards reducing gases.
Table 4. Summary of the chemical/gas sensing properties of TiO2-based nanostructures and heterostructures towards reducing gases.
MaterialSynthesis MethodTarget Gas/
Concentration
Response (S)Temperature/
Humidity
Response/
Recovery Time
Ref./Year
TiO2 NTsElectrochemical anodizationH2/100 ppmS = (Ig/Ia)
3.5
RT/Dry air0.7 s/0.9 s[67]/2015
TiO2 NFsElectrospinningH2/50 ppmS = (Ra/Rg)
30
190 °C with UV irradiation/Dry air2s/6.9 s[40]/2015
TiO2 NRsHydrothermalH2/2000 ppmS = (∆I/I × 100)
215%
200 °C/Dry airNA[42]/2015
p-Co3O4/n-TiO2 NTsElectrochemical anodizationH2/1000 ppmS = (∆I/I)
6
200 °C/50%10 min/5 min[48]/2018
Pd decorated TiO2 NTsElectrochemical anodizationH2/10ppmS = (∆R/R)
1.25
180 °C/in dry synthetic air20 s/40 s[69]/2016
PdAu decorated TiO2 NWsHydrothermalH2/5 ppmS = (∆I/I×100)
350%
RT/-42 s/NA[45]/2016
TiO2/ZnO core-shell NTsElectrochemical anodizationEthanol/1930 ppmS = (Ra/Rg)
0.8
100 °C/-NA[47]/2018
Al2O2/TiO2Thermal oxidationEthanol/1000 ppmS = (Ra/Rg)
1108.9
650 °C/-4 min/20 min[61]/2017
Nb2O5-TiO2 NFsElectrospinningEthanol/500 ppmS = (Ra/Rg)
21.64
250 °C0/45%NA[58]/2019
TiO2/V2O5 NFsElectrospinningEthanol/100 ppmS = (Ra/Rg)
24.6
350 °C/30%6 s/7s[74]/2016
Ni-TiO2 NTsElectrochemical anodizationAcetone/1000 ppmS = (∆R/R×100)
82%
100 °C/-NA[71]/2015
TiO2-SnO2 core-shell NFsElectrospinningAcetone/100 ppmS = (Ra/Rg)
13.5
280 °C/-2 s/60 s[59]/2017
TiO2 NTsElectrochemical anodizationAcetone/50 vol%S = (∆R/R×100)
115%
Light irradiation at RT/-NA[68]/2020
TiO2 NWsVapor-Phase growthCO/1 ppmS = (NA)
11%
400 °C/-NA[63]/2016
TiO2 NRsHydrothermalAmmonia/20 ppmS = (∆R/R × 100)
14.1%
RT/50%61 s/9 s[44]/ 2020
Pbs QDs/TiO2 NTsElectrochemical anodizationAmmonia/100 ppmS = (Ra/Rg)
17.49
RT/-NA[70]/2016
TiO2 NTsElectrochemical anodizationH2S/6 to38 ppmS = (∆R/R × 100)
12 to 144%
70 °C/10%NA[72]/2016
Co-doped TiO2 NTsElectrochemical anodizationH2S/50 ppmS = (Ra/Rg)
199.16%
300 °C/50%15 s/4 s[73]/2019
Table 5. Summary of the gas sensing properties of TiO2-based nanostructure and heterostructure sensors towards oxidizing gases.
Table 5. Summary of the gas sensing properties of TiO2-based nanostructure and heterostructure sensors towards oxidizing gases.
MaterialSynthesis
Method
Target Gas/
Concentration
Response (S)Temperature/
Humidity
Response/
Recovery Time
Ref./Year
TiO2 NWsHydrothermalNO2/100 ppmS = (Ra/Rg)
3.1
RT/50%RH10 s/19 s[66]/(2018)
TiO2 NRsHydrothermal and annealingNO2/40 ppmS = (Rg/Ra)
1300
RT/dry airNA[64]/(2017)
TiO2-Al2O3 Core-shell NWsThermal oxidationNO2/1000 ppmS = (Rg/Ra)
1.9
650 °C/dry air180 s/180 s [61]/(2017)
MoS2-Decorated TiO2 NTsAnodization and HydrothermalNO2/100 ppmS = (Rg/Ra)
1.1
150 °C/45%RHNA[50]/(2016)
TiO2-In2O3 Composite NFsElectrospinningNO2/97 ppmS = (∆R/Ra)
95
RT/26%RH6.7 s/NA[60]/(2015)
TiO2 NTsAnodizationNO2/50 ppmS = (∆I/Ig)
17
200 °C/50%RHNA[48]/(2018)
TiO2 Networked NWsVapor-phase growthNO2/50 ppmS = (∆R/Ra×100)
8%
400 °C/dry airNA[63]/(2016)
TiO2-Ag2O Composite NRsHydrothermal and sputteringNO2/0.5 ppmS = (Rg/Ra)
3.1
250 °C/dry air87 s/112 s[65]/(2019)
TiO2@Au Heterojunction NRsHydrothermal and chemical approachNO2/5 ppmS = (∆R/Ra)
135.5
250 °C/dry air40 s/43 s[94]/(2017)
TiO2 NRs ArrayAcid vapor oxidationO2/16% vol.S = (Rg/Ra)
2.1
RT/dry air55 s/51 s[95]/(2016)

Share and Cite

MDPI and ACS Style

Kaur, N.; Singh, M.; Moumen, A.; Duina, G.; Comini, E. 1D Titanium Dioxide: Achievements in Chemical Sensing. Materials 2020, 13, 2974. https://doi.org/10.3390/ma13132974

AMA Style

Kaur N, Singh M, Moumen A, Duina G, Comini E. 1D Titanium Dioxide: Achievements in Chemical Sensing. Materials. 2020; 13(13):2974. https://doi.org/10.3390/ma13132974

Chicago/Turabian Style

Kaur, Navpreet, Mandeep Singh, Abderrahim Moumen, Giorgio Duina, and Elisabetta Comini. 2020. "1D Titanium Dioxide: Achievements in Chemical Sensing" Materials 13, no. 13: 2974. https://doi.org/10.3390/ma13132974

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