Next Article in Journal
Sustainable Preparation of Nanoporous Carbons via Dry Ball Milling: Electrochemical Studies Using Nanocarbon Composite Electrodes and a Deep Eutectic Solvent as Electrolyte
Next Article in Special Issue
Three-Layer PdO/CuWO4/CuO System for Hydrogen Gas Sensing with Reduced Humidity Interference
Previous Article in Journal
Epitaxially Integrated Hierarchical ZnO/Au/SrTiO3 and ZnO/Ag/Al2O3 Heterostructures: Three-Dimensional Plasmo-Photonic Nanoarchitecturing
Previous Article in Special Issue
Sensing Behavior of Metal-Free Porphyrin and Zinc Phthalocyanine Thin Film towards Xylene-Styrene and HCl Vapors in Planar Optical Waveguide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Light-Excited Ag-Doped TiO2−CoFe2O4 Heterojunction Applied to Toluene Gas Detection

1
Key Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, China
2
Harbin Institute of Technology, School of Science, Shenzhen 518055, China
3
Bingtuan Industrial Technology Research Institute, Shihezi University, Shihezi 832003, China
4
Carbon Neutralization and Environmental Catalytic Technology Laboratory, Shihezi University, Shihezi 832003, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(12), 3261; https://doi.org/10.3390/nano11123261
Submission received: 14 October 2021 / Revised: 24 November 2021 / Accepted: 26 November 2021 / Published: 30 November 2021
(This article belongs to the Special Issue Nanostructured Materials for Gas Sensor Applications)

Abstract

:
(1) Background: Toluene gas is widely used in indoor decoration and industrial production, and it not only pollutes the environment but also poses serious health risks. (2) Methods: In this work, TiO2−CoFe2O4−Ag quaternary composite gas-sensing material was prepared using a hydrothermal method to detect toluene. (3) Results: The recombination of electron–hole pairs was suppressed, and the light absorption range was expanded after constructing a heterojunction and doping with Ag, according to ultraviolet–visible (UV–vis) diffuse reflectance spectra and photoluminescence spectroscopy. Moreover, in the detection range of toluene gas (3 ppm–50 ppm), the response value of TiO2−CoFe2O4−Ag increased from 2 to 15, which was much higher than that of TiO2−Ag (1.7) and CoFe2O4−Ag (1.7). In addition, the working temperature was reduced from 360 °C to 263 °C. Furthermore, its response/recovery time was 40 s/51 s, its limit of detection was as low as 10 ppb, and its response value to toluene gas was 3–7 times greater than that of other interfering gases under the same test conditions. In addition, the response value to 5 ppm toluene was increased from 3 to 5.5 with the UV wavelength of 395 nm–405 nm. (4) Conclusions: This is primarily due to charge flow caused by heterojunction construction, as well as metal sensitization and chemical sensitization of novel metal doping. This work is a good starting point for improving gas-sensing capabilities for the detection of toluene gas.

1. Introduction

With the development of the economy and the substantial improvement at industrial level, there are a growing number of toxic and harmful gases in the air, which pollute the environment and also seriously endanger human health [1,2,3]. Toluene gas, for example, produced during indoor decoration, has been linked to vision impairment; asthma; nasopharyngeal cancer; and other diseases and health events, including pregnancy loss [4,5,6,7]. In addition, when people are exposed to high concentrations of toluene vapor (200 ppm–500 ppm), various symptoms, such as headache, nausea, muscle cramps, and dizziness, may occur. When the human body is exposed to a too high concentration of toluene gas, the brain is permanently poisoned and may even die [8]. As a result, an efficient and sensitive gas sensor for the detection of toluene is required.
The gas sensors made of semiconductor metal oxide gas-sensitive materials have low cost, high sensitivity, and simple operation [4], which are easy to integrate into mobile phones and other miniaturized devices [5]. Common gas-sensitive materials include SnO2 [6,7], ZnO [8,9], TiO2 [10,11], WO3 [12,13], Fe2O3 [14], In2O3 [15,16], and Co3O4 [17,18]. TiO2 is one of them, and it is widely used in gas detection because of its advantages, such as good stability, non-toxicity, and recyclability [19,20]. However, pure TiO2-based gas sensors have high resistance, poor selectivity, and low response strength, limiting their wide application. As we all know, it can be modified by some methods, such as doping [21], constructing heterojunctions, defect regulation, and controlling morphology [22,23,24,25,26]. The use of heterostructures can improve the material’s catalytic activity and adsorption capacity, thereby enhancing sensor signal reception and transduction. Li et al. [27] prepared TiO2@SnO2 hollow nanospheres using a hydrothermal method with the aid of templates. When compared to pure hollow TiO2 nanospheres, hollow TiO2@SnO2 heterojunction nanospheres performed better in terms of formaldehyde sensing under ultraviolet (UV) irradiation at room temperature. The response/recovery time was shortened from 52 s/164 s to 20 s/56 s. This phenomenon indicated that the heterostructure could improve the electron transfer efficiency and enhance the gas sensitivity.
Furthermore, element doping can affect the surface defects and electrical properties of the sensing material, which promotes electron transfer. Simultaneously, element doping can alter the material bandgap, improving gas sensitivity. In particular, doping platinum, gold, palladium, silver, and other precious metals with high catalytic activity and low Fermi levels will trigger electronic sensitization and chemical sensitization effects, thus effectively improving the material’s gas sensitivity [28]. Pan et al. [29] formed Pd-doped TiO2 nanofiber membranes with varying doping amounts using rotating surface flame stabilization technology (FSRS). It was discovered that noble metal doping improved the response intensity of the TiO2-based sensor to CO, which also greatly reduced its response/recovery time to NH3.
In this work, TiO2−CoFe2O4−Ag quaternary composite gas-sensitive material is synthesized with a hydrothermal method to detect toluene gas. The method for constructing heterojunctions and doping noble metals greatly improves the detection ability of TiO2-based gas-sensitive materials for toluene gas. It is discovered that the recombination of photogenerated electron–hole pairs is inhibited, and absorbance increases significantly, particularly in the UV region. Moreover, it can be seen that the working temperature is greatly reduced; the response value is increased compared with TiO2−Ag and CoFe2O4−Ag. At the same time, its gas sensitivity response value is further increased with UV irradiation.

2. Materials and Methods

2.1. Preparation of Materials

First, 5.95 g of cobalt chloride hexahydrate and 13.51 g of ferric chloride hexahydrate were placed into 100 mL of ethylene glycol solution with ammonium acetate. The solution was then stirred at 300 rpm at room temperature for one night before being aged for another. Following that, the suspension was hydrothermally treated for 24 h at 180 °C. After it cooled down, it was centrifuged at a speed of 8000 r/min and cleaned with 1 L of deionized water and anhydrous ethanol alternately. Finally, the product was dried at 60 °C for 12 h before being prepared with CoFe2O4. TiO2−CoFe2O4 was prepared in the same way as CoFe2O4, except that a suspension containing 2.00 g of titanium dioxide was added before the hydrothermal treatment. TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag were all prepared by dipping with silver nitrate solution as a silver source; the doping amount of silver in the three materials was 1% wt.
The concentration of ethylene glycol is 98%. All other materials and reagents are analytical reagents.

2.2. Material Characterization

The crystal structures of the samples were recorded using a powder X-ray diffraction method (XRD, BRUCKERD8 ADVANCE, Karlsruhe, Germany) with Co Kα radiation at a scan rate of 8°/min. Scanning electron microscope (SEM, SU8020, Hitachi Corporation, Tokyo, Japan), transmission electron microscope (TEM), and high-resolution TEM (HRTEM) with FEI Tecnai G2 F30 were used to characterize the morphology and microstructure (FEI Corporation, Hillsboro, OR, USA). The specific area and pore diameter distributions were estimated using the Brunauer–Emmett–Teller equation and the Barrett–Joyner–Halenda method based on N2 adsorption isotherms (ASAP 2460, Micromeritics Instruments Corporation, NJ, USA). Thermo scalable 250 Xi equipment was used for X-ray photoelectron spectroscopy (XPS) measurements (FEI Corporation, Hillsboro, OR, USA). The ultraviolet–visible diffuse reflectance spectra (UV–vis DRS) of the samples were measured using the UV-3600 instrument (SHIMADZU Corporation, Kyoto, Japan). An FLS1000/FS5 fluorescence spectrometer (excitation wavelength = 350 nm) was employed to obtain the photoluminescence (PL) spectra (Edinburgh Company, Edinburgh, UK), which were used to analyze the recombination behavior of photoinduced carriers.

2.3. Sensing Performance Evaluation

The CGS-8 intelligent gas sensor system was used to evaluate sensing performance (Beijing Elite Co., Ltd., Beijing, China). Before the performance test, the gas-sensitive material was mixed evenly with ethanol and applied to the surface of the ceramic tube, which was welded to the hexapod base. Then, it was aged at 150 °C for 48 h after standing for a while. The light excitation performance test was carried out by irradiating about 3 cm above the gas sensor. During the performance test, the relative humidity (RH) was around 12%. The sensors’ responses were calculated using the following equation:
Response = Ra/Rg
where Ra is the resistance of the sensor in air, and Rg is the electrical resistance for the sensor in the tested gas.

3. Results and Discussion

3.1. Structural and Morphological Characterization

From the XRD test results in Figure 1a, the typical anatase peak pattern of TiO2 can be seen; the peaks appearing at 2θ = 25.347°, 37.861°, 48.073°, 53.922°, 55.164°, and 62.788° correspond with (101), (004), (200), (105), (211), and (213) and other crystal planes of TiO2, respectively (JCPDS No.21-1272) [30,31]. When compared to other peaks, the intensity of the peak at 25.347° is the highest, indicating that the corresponding (101) crystal plane is the most exposed. The peak state compares to that reported in the literature, confirming the presence of titanium dioxide in the crystal phase of anatase [32,33]. The anatase nanocrystals have a higher specific surface area and a greater concentration of oxygen vacancies, so there may be more active centers and higher charge separation efficiency. Furthermore, because the bandgap of anatase is relatively large, its redox ability is slightly higher than that of rutile. As a result, anatase has improved photocatalytic performance and is the ideal TiO2 nanocrystalline phase for gas-sensitive reactions [34,35]. Compared with TiO2−Ag, the peak intensity of TiO2−CoFe2O4−Ag becomes weaker, which indicates that titanium dioxide nanoparticles may be affected after being compounded with a cobalt ferrite. Similar reports have been reported in the previous literature [36].
The peaks of CoFe2O4 at 2θ = 18.231°, 30.016°, 35.376°, 43.001°, 53.347°, 56.878°, and 62.483° correspond with (111), (220), (311), (400), (422), (511), and (440) of CoFe2O4 crystal planes (JCPDS No.22-1086) [33,37]. The peak state is consistent with what has been reported in the literature [35,38]. The intensity of the peak at 35.376° is greater than that of the other peaks, indicating that the corresponding (311) crystal plane is more exposed. Due to too little Ag doping, there is no obvious characteristic peak of Ag in the spectrum. Furthermore, there are no impurity peaks in all three materials, which shows that the synthesized samples are relatively pure.
Figure 1b shows that the pore diameters of the three types of gas-sensitive materials are concentrated in the 10 nm–40 nm range, with mesopores being the most common, followed by a few micropores. It can be seen in Table 1 that the specific surface area, pore volume, and pore size of the TiO2−CoFe2O4−Ag gas-sensing material are all closer to those of TiO2−Ag after recombination but slightly reduced. While these characteristics are related to gas-sensing performance, they are not the only deciding factors. The gas sensitivity of a material is also affected by its internal structure and surface defects [39].
The SEM and TEM images in Figure 2a–f show that TiO2−Ag is made up of small rice-shaped particles that stack together to form a coral-like structure. CoFe2O4−Ag, on the other hand, has a nearly spherical structure with a regular shape. The diameter varies from about 300 nm to 700 nm. The size of TiO2−CoFe2O4−Ag is bigger than that of TiO2−Ag and smaller than that of CoFe2O4−Ag. TiO2−CoFe2O4−Ag is in the shape of a sphere, which stacks together, and some individual boundaries blend. The structure has numerous holes and voids to ensure that the gas-sensitive material fully contacts and reacts with the air and target gas, improving gas-sensitive performance and shortening reaction time. At the same time, it can be seen that there is some debris on the surface of the three materials, which may be the Ag substance.
As the binding energy increases from low to high, the XPS total spectrum (Figure 3a) shows six features of C, Ag, Ti, O, Fe, and Co. Apart from the C element derived from the test process, no other impurity elements are present, indicating that the gas-sensitive material is pure and pollution free.
In Figure 3b, the binding energy of the two peaks of Ti 2p from low to high is attributed to Ti 2p3/2 and Ti 2p1/2, respectively, and the difference between the two peaks is about 5.5 eV, indicating that the material contains TiO2. The Ti 2p1/2 peak can be divided into two peaks, 463.54/463.67 eV and 464.61/464.65 eV, indicating the presence of Ti3+. To maintain the charge balance, there must be oxygen vacancies, which can improve the gas sensitivity response.
Figure 3c shows the O 1s peaks that can be attributed to lattice oxygen (OL), defect oxygen (OV), and adsorbed oxygen (OC). Among them, the binding energy of OL is 529.94/529.57/529.77 eV, which is primarily derived from the oxygen within the TiO2 and CoFe2O4 lattices. The binding energy of OV is located at 529.95/531.11/530.96 eV, indicating that there are some oxygen defects, which mainly exist in the form of oxygen vacancies. This corresponds to the result of the Ti 2p elemental analysis. The binding energy of OC is 531.55/532.25/532.35 eV, which comes from the chemically adsorbed and dissociated oxygen elements, indicating that there is a certain amount of oxygen anions in different states ((O2 (ads), O (ads), and O2− (ads)), which are helpful for the response of metal oxide semiconductor gas-sensitive materials.
The presence of Ag in Figure 3d indicates that the element is doped successfully. The two peaks are Ag 3d5/2 and Ag 3d3/2, and their binding energies are 368.06/367.42/366.9 eV and 374.07/373.44/372.29 eV. The binding energy difference is approximately 6 eV, indicating that the deposited silver nanoparticles are metallic silver [40,41]. At the same time, because the element Ag occupies a relatively small proportion, the peaks are not obvious, and the peak intensity is low.
It can be seen that the TiO2−CoFe2O4−Ag quaternary composite material and TiO2−Ag have the same peaking law in Figure 4a. Peaks between 400 nm and 450 nm are near-band edge peaks that are typically accompanied by the generation of electron–hole pairs. When an electron is excited to transition to the conduction band, the holes in the valence band interact and self-annihilate, releasing the energy difference between the highest occupied molecular orbital energy state in the valence band and the lowest unoccupied molecular orbital energy state in the valence band minus the excitons’ binding energy [42]. The peak between 600 nm and 650 nm corresponds to the red emission band, indicating that there are a large number of defects in the material, which are conducive to the improvement of its gas-sensing performance. Among the three materials, the peak intensity of CoFe2O4−Ag and TiO2−CoFe2O4−Ag is significantly lower than that of TiO2−Ag, indicating that the recombination of photogenerated electron–hole pairs is significantly suppressed and that more carriers can be generated under light conditions. It is helpful to improve the conductivity of the material and reduce the operating temperature.
A large portion of solar radiation is visible light (43%) and infrared (52%), with only 5% falling into the UV range. Only colorless and white substances can absorb energy in the UV range. Pure TiO2 is a white crystalline solid that only absorbs UV energy. This is related to the intrinsic bandgap absorption of typical anatase TiO2, which is caused by the transition of electrons from the valence band to the conduction band [43]. Moreover, by modification through a series of methods, such as doping and constructing heterojunctions, the light absorption range can be expanded to include visible light [44]. Figure 4b shows that its light absorption range is 200 nm–700 nm, which covers the visible, UV, and infrared regions, indicating that the addition of the noble metal Ag improves its utilization ability to different wavelengths of light. This may cause different wavelengths of light to enhance the gas sensitivity of the material.
The absorption in the visible region of the spectrum is low, but the absorption in the UV region is high, indicating that it preferentially absorbs UV light. At the same time, due to the effect of local surface plasmon resonance, the absorption of TiO2−Ag in the visible region (400 nm–700 nm) will be increased accordingly [45,46]. Furthermore, the absorption light intensity of TiO2−CoFe2O4−Ag is significantly greater than that of TiO2−Ag and CoFe2O4−Ag, indicating that it has a higher light utilization rate. When a gas-sensitive reaction occurs, the light enhancement may be more pronounced.
The bandgap energy is calculated using the Kubelka–Munk (KM) method:
αhυ = A ( − Eg)n
where A is an absorbance constant, is the energy of discrete photons, and α is the absorption coefficient, which can be calculated from the diffuse reflectance data using the KM method. The bandgap energy is denoted by the symbol Eg. n is 0.5 because titanium dioxide and other materials have an indirect bandgap. Finally, the bandgap value can be obtained by extrapolating the intersection of (α)n and photon energy () [42,47].
It can be seen in Figure 4c,d that the bandgap of TiO2−Ag is about 2.74 eV, the bandgap of CoFe2O4−Ag is 1.91 eV, and the bandgap of TiO2−CoFe2O4−Ag is about 1.47 eV. The bandgap is reduced by constructing heterojunctions and doping Ag. As a result, electrons transition more easily, and gas sensitivity may be improved.

3.2. Gas-Sensing Characteristics

The working temperature has a direct effect on the surface activity state of the gas-sensitive material and the diffusion rate of gas molecules on the surface of the gas-sensitive material, which in turn affects the sensor device’s sensitivity. Therefore, the influence of the working temperature on the gas-sensing performance is tested first. It can be seen from Figure 5a, b that the response value of TiO2−CoFe2O4−Ag to 50 ppm toluene presents the trend of first increasing and then decreasing with an increase in working temperature. It is hypothesized that this is because the energy required for the reaction of toluene with chemically adsorbed oxygen on the surface of gas-sensitive material is insufficient in a low-temperature environment, resulting in a relatively low response value. As the temperature rises, more gas molecules can obtain enough energy to increase the diffusion and reaction speed, and the response value increases. However, as the test temperature rises, the gas molecules diffuse too quickly and desorb too fast, preventing them from reaching further, and the response value falls accordingly. Furthermore, adsorption and desorption are in a state of dynamic equilibrium. When the temperature is too high, the gas desorption rate may be greater than the adsorption rate, and the gas sensitivity response value will decrease. At the same time, the sensor devices based on TiO2−Ag and CoFe2O4−Ag are produced as a comparison. It is discovered that their resistance is excessively high; only when the temperature exceeds 360 °C does the resistance fall within the instrument’s test range. Furthermore, their response value is extremely low, less than 1.7 for 50 ppm toluene, and they do not exhibit a similar response value to TiO2−CoFe2O4−Ag with a significant temperature-related change trend. At the same time, their response value error is larger than that of TiO2−CoFe2O4−Ag. Finally, it is found that the best working temperature of TiO2−CoFe2O4−Ag is 263 °C, and the response value is about 15 for 50 ppm toluene gas. As a result, all subsequent performance tests are conducted using TiO2−CoFe2O4−Ag for toluene at 263 °C.
As shown in Figure 5c, it can be found that the response/recovery time generally shows a gradually increasing trend as the concentration of toluene gas increases. When the gas concentration is 3 ppm, the response/recovery time is 40 s/51 s.
The dynamic response test results in Figure 5d show that when the toluene concentration is between 3 ppm and 50 ppm, the response value increases linearly from 2 to 15, indicating that the synthesized TiO2−CoFe2O4−Ag gas-sensitive material can provide a reliable quantitative analysis of toluene. In addition, the limit of detection (LOD) can be estimated as 3δ/s based on the linear fitting, in which δ is the standard deviation and s is the slope of the linear fit. Finally, the LOD of TiO2−CoFe2O4−Ag to toluene is calculated to be about 10 ppb.
Then, the selectivity of the TiO2−CoFe2O4−Ag sensor to toluene gas is tested, using acetone, methanol, ethanol, and acetaldehyde as reference interference gases. Figure 6a depicts the gas sensitivity response of TiO2−CoFe2O4−Ag to different gases with the same concentration (50 ppm) at 263 °C, with the response values for each type of gas being approximately 3.5, 2.5, 4.5, 3, and 15, respectively. It can be observed that the TiO2−CoFe2O4−Ag-based sensor shows a much higher response to toluene than to other gases, indicating that it can accurately identify toluene from the above various interference gases.
The TiO2-CoFe2O4-Ag sensor is then tested for moisture resistance by placing it in different RH (10%–60% RH) conditions. Figure 6b shows that as the RH increases, the sensor response value to 20 ppm toluene decreases from about 6 to about 1.7. It is speculated that water vapor molecules in the environment occupy the surface of the gas-sensitive material, which affects the adsorption and reaction of oxygen molecules and target gases on the surface of the gas-sensitive material.
Following that, the sensor is tested for repeatability and long-term stability (Figure 6c, d), and it is discovered that the response value is stable at around 6, and the baseline does not drift significantly during the seven cycles of testing in the 20 ppm toluene atmosphere. In the long-term stability test, the response of TiO2−CoFe2O4−Ag to 10 ppm toluene gas was tested every 5 days; the result shows that about 90% of its response value is still maintained after 30 days.
The light-excited gas sensitivity response test is then performed. As shown in Figure 7a, light irradiation with the same light intensity but different wavelengths have different enhancement effects on gas sensitivity. However, the gas sensitivity response value for 20 ppm toluene with light is higher than that in darkness. The response value of the device is the largest (about 6) with a wavelength of 395 nm–405 nm, which is much greater than that in darkness (about 3).
Finally, the dynamic response curve of TiO2−CoFe2O4−Ag with the light of 395 nm–405 nm wavelength and in darkness is displayed in Figure 7b. It is discovered that as the concentration of toluene gas increases within the measured concentration range (3 ppm–30 ppm), whether with or without light, the response value increases. The response value in the presence of light is much higher than that in darkness. This is mainly due to the effect of photogenerated electron–holes generated under light conditions on the gas-sensing performance.

3.3. Gas-Sensing Mechanism

When a metal oxide resistive semiconductor gas sensor comes into contact with gas molecules, the surface adsorption or reaction causes carrier movement, which causes changes in electrical conductivity, volt–ampere characteristics, or surface potential. Furthermore, various gases can be detected based on changes before and after exposure to the target gas [48,49,50]. At the same time, according to the literature, the gas sensitivity can be improved by surface modification, applying UV light illumination, and other methods [51,52].
As illustrated in Figure 8a,b, when n-type TiO2 and n-type CoFe2O4 combine to form a heterojunction, electrons will flow from the side with the smaller work function to the side with the larger work function until the Fermi level reaches equilibrium [53,54,55]. An electron depletion layer will form on the TiO2 side, and the resistance in the air increases. After contacting the target gas, the change in resistance increases, and the sensitivity is improved. Furthermore, due to the current electronic effects, the formation of heterostructures between different metal oxide semiconductors can significantly improve O2 adsorption [55]. These effects work synergistically, which further improves the sensitivity of TiO2−CoFe2O4−Ag.
Moreover, the doped noble metal will form a Schottky junction with the gas-sensitive material [56]. More chemisorbed oxygen will be adsorbed as a result of the catalytic effect of the noble metals. Noble metal spillover allows chemisorbed oxygen species to be easily transported and distributed on the surface of gas-sensitive material, increasing the reaction rate [57]. In addition, chemical sensitization benefits from the catalytic properties of noble metals can reduce the surface reaction barrier, thereby lowering the working temperature and shortening the response time [58,59,60,61]. Furthermore, silver has good electrical conductivity, which promotes electron transfer and accelerates the ionization of oxygen and the surface redox reaction, ultimately improving response speed [62].
Furthermore, adding light irradiation will produce photogenerated carriers to further improve its gas sensitivity [63,64]. Light excitation can also provide energy for the transition of carriers between energy bands and can activate the surface to adsorb oxygen, resulting in increased sensitivity [42]. These are the reasons why adding light irradiation leads to the enhancement of the response intensity in this work.
At the same time, in addition to the chemical sensitization caused by noble metals, which can lower the working temperature, the composite cobalt ferrite has good catalytic activity and conductivity. Because cobalt ferrite is a multivalent metal element substance, the electron can jump in the cation valence of the octahedral position in its spinel structure [38,65]. The bandgap is significantly reduced after combining titanium dioxide, cobalt ferrite, and silver as shown in Figure 4d. As a result, electrons transition more easily, and the material’s conductivity increases. Eventually, this leads to a decrease in operating temperature from 360 °C to 263 °C.
Furthermore, during the sensing process, there may be some physical charge transfer between the toluene gas and the sensing layer, resulting in a decrease in resistance.

4. Conclusions

In summary, the TiO2−CoFe2O4−Ag quaternary composite gas-sensing material was prepared using the hydrothermal method. The detection capability of the heterojunction to toluene was significantly improved by doping noble metal, which could induce the electronic sensitization and chemical sensitization effects. The working temperature was reduced from 360 °C to 263 °C, and the response value to 50 ppm toluene was increased from 1.7 to 15. At the same time, TiO2−CoFe2O4−Ag has good selectivity and long-term stability. Furthermore, the response value of TiO2−CoFe2O4−Ag to UV irradiation was increased for toluene gas. This article introduces a novel approach to the development of an excellent gas-sensitive reaction system.

Author Contributions

Conceptualization, methodology, validation, writing—original draft preparation, and writing—review and editing, W.W.; data curation, Y.K. and L.Z.; resources, supervision, and project administration, F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Innovation Talents Program of Bingtuan, No.2019CB025.

Acknowledgments

The authors thank Yingchun Li (Harbin Institute of Technology) for all kinds of help in the experiment and submission process.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dai, J.; Ogbeide, O.; Macadam, N.; Sun, Q.; Yu, W.; Li, Y.; Su, B.L.; Hasan, T.; Huang, X.; Huang, W. Printed gas sensors. Chem. Soc. Rev. 2020, 49, 1756–1789. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, C.; Sun, J.; Sun, Y.; Tan, Z.; Xu, X.; Fu, Y.; Feng, Z.; Zhu, J. Fabrication of cubic Co3O4-hexagonal ZnO disk/rGO as a two-phase benzaldehyde sensor via a sequential nucleation strategy. Sens. Actuators B 2021, 330, 129384–129396. [Google Scholar] [CrossRef]
  3. Malik, R.; Tomer, V.K.; Chaudhary, V.; Dahiya, M.S.; Rana, P.S.; Nehra, S.P.; Duhan, S. Facile synthesis of hybridized mesoporous Au@TiO2/SnO2 as efficient photocatalyst and selective VOC sensor. ChemistrySelect 2016, 1, 3247–3258. [Google Scholar] [CrossRef]
  4. Fiedot-Tobola, M.; Suchorska-Wozniak, P.; Startek, K.; Rac-Rumijowska, O.; Szukiewicz, R.; Kwoka, M.; Teterycz, H. Correlation between microstructure and chemical composition of zinc oxide gas sensor layers and their gas-sensitive properties in chlorine atmosphere. Sensors 2020, 20, 6951. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, T.H.; Jeong, S.Y.; Moon, Y.K.; Lee, J.H. Dual-mode gas sensor for ultrasensitive and highly selective detection of xylene and toluene using Nb-doped NiO hollow spheres. Sens. Actuators B 2019, 301, 127140–127178. [Google Scholar] [CrossRef]
  6. Huang, J.; Xu, X.; Gu, C.; Wang, W.; Geng, B.; Sun, Y.; Liu, J. Effective VOCs gas sensor based on porous SnO2 microcubes prepared via spontaneous phase segregation. Sens. Actuators B 2012, 173, 599–606. [Google Scholar] [CrossRef]
  7. Wang, X.; Ren, P.; Tian, H.; Fan, H.; Cai, C.; Liu, W. Enhanced gas sensing properties of SnO2: The role of the oxygen defects induced by quenching. J. Alloys Compd. 2016, 669, 29–37. [Google Scholar] [CrossRef]
  8. Wang, L.; Wang, S.; Xu, M.; Hu, X.; Zhang, H.; Wang, Y.; Huang, W. A Au-functionalized ZnO nanowire gas sensor for detection of benzene and toluene. Phys. Chem. Chem. Phys. 2013, 15, 17179–17186. [Google Scholar] [CrossRef]
  9. Peng, C.; Guo, J.; Yang, W.; Shi, C.; Liu, M.; Zheng, Y.; Xu, J.; Chen, P.; Huang, T.; Yang, Y. Synthesis of three-dimensional flower-like hierarchical ZnO nanostructure and its enhanced acetone gas sensing properties. J. Alloys Compd. 2016, 654, 371–378. [Google Scholar] [CrossRef]
  10. Li, X.; Zhao, Y.; Wang, X.; Wang, J.; Gaskov, A.M.; Akbar, S.A. Reduced graphene oxide (rGO) decorated TiO2 microspheres for selective room-temperature gas sensors. Sens. Actuators B 2016, 230, 330–336. [Google Scholar] [CrossRef]
  11. Fomekong, R.L.; Kelm, K.; Saruhan, B. High-Temperature Hydrogen Sensing Performance of Ni-Doped TiO2 Prepared by Co-Precipitation Method. Sensors 2020, 20, 5992. [Google Scholar] [CrossRef]
  12. Ma, J.; Ren, Y.; Zhou, X.; Liu, L.; Zhu, Y.; Cheng, X.; Xu, P.; Li, X.; Deng, Y.; Zhao, D. Pt nanoparticles sensitized ordered mesoporous WO3 semiconductor: Gas sensing performance and mechanism study. Adv. Funct. Mater. 2018, 1705268–1705280. [Google Scholar] [CrossRef]
  13. Li, Z.; Li, J.; Song, L.; Gong, H.; Niu, Q. Ionic liquid-assisted synthesis of WO3 particles with enhanced gas sensing properties. J. Mater. Chem. A 2013, 1, 15377–15382. [Google Scholar] [CrossRef]
  14. Wang, S.; Wang, Y.; Zhang, H.; Gao, X.; Wang, Y. Fabrication of porous α-Fe2O3 nanoshuttles and their application for toluene sensors. RSC Adv. 2014, 4, 30840–30850. [Google Scholar] [CrossRef]
  15. Li, Z.; Li, Y.; Luan, Y.; Li, J.; Song, A. In(OH)3 particles from an ionic liquid precursor and their conversion to porous In2O3 particles for enhanced gas sensing properties. CrystEngComm 2013, 15, 1659–1822. [Google Scholar]
  16. Naberezhnyi, D.; Rumyantseva, M.; Filatova, D.; Batuk, M.; Hadermann, J.; Baranchikov, A.; Khmelevsky, N.; Aksenenko, A.; Konstantinova, E.; Gaskov, A. Effects of Ag Additive in Low Temperature CO Detection with In2O3 Based Gas Sensors. Nanomaterials 2018, 8, 801. [Google Scholar] [CrossRef] [Green Version]
  17. Zhao, C.; Huang, B.; Zhou, J.; Xie, E. Synthesis of porous Co3O4 nanonetworks to detect toluene at low concentration. Phys. Chem. Chem. Phys. 2014, 16, 19327–19332. [Google Scholar] [CrossRef]
  18. Yang, Y.; Wang, X.; Yi, G.; Li, H.; Shi, C.; Sun, G.; Zhang, Z. Hydrothermal Synthesis of Co3O4/ZnO Hybrid Nanoparticles for Triethylamine Detection. Nanomaterials 2019, 9, 1599. [Google Scholar] [CrossRef] [Green Version]
  19. Song, G.; Xin, F.; Yin, X. Photocatalytic reduction of carbon dioxide over ZnFe2O4/TiO2 nanobelts heterostructure in cyclohexanol. J. Colloid Interface Sci. 2015, 442, 60–66. [Google Scholar] [CrossRef]
  20. Lai, Y.; Huang, J.; Cui, Z.; Ge, M.; Zhang, K.Q.; Chen, Z.; Chi, L. Recent advances in TiO2-based nanostructured surfaces with controllable wettability and adhesion. Small 2016, 12, 2203–2224. [Google Scholar] [CrossRef]
  21. Luan, V.H.; Tien, H.N.; Hur, S.H.; Han, J.H.; Lee, W. Three-Dimensional Porous Nitrogen-Doped NiO Nanostructures as Highly Sensitive NO2 Sensors. Nanomaterials 2017, 7, 313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Sun, P.; Wang, C.; Liu, J.; Zhou, X.; Li, X.; Hu, X.; Lu, G. Hierarchical assembly of alpha-Fe2O3 nanosheets on SnO2 hollow nanospheres with enhanced ethanol sensing properties. ACS Appl. Mater. Inter. 2015, 7, 19119–19125. [Google Scholar] [CrossRef]
  23. 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] [PubMed]
  24. Zhang, D.; Sun, Y.E.; Li, P.; Zhang, Y. Facile fabrication of MoS2-modified SnO2 hybrid nanocomposite for ultrasensitive humidity sensing. ACS Appl. Mater. Inter. 2016, 8, 14142–14182. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, S.; Qiao, Y.; Huang, J.; Yao, H.; Zhang, Y.; Li, Y.; Du, J.; Fan, W. One-pot synthesis of mesoporous spherical SnO2@graphene for high-sensitivity formaldehyde gas sensors. RSC Adv. 2016, 6, 25198–25202. [Google Scholar] [CrossRef]
  26. 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]
  27. Zhang, S.; Zhao, L.; Huang, B.; Li, X. UV-activated formaldehyde sensing properties of hollow TiO2@SnO2 heterojunctions at room temperature. Sens. Actuators B 2020, 319, 128264–128274. [Google Scholar] [CrossRef]
  28. Platonov, V.; Rumyantseva, M.; Khmelevsky, N.; Gaskov, A. Electrospun ZnO/Pd nanofibers: CO sensing and humidity effect. Sensors 2020, 20, 7333. [Google Scholar] [CrossRef]
  29. Pan, F.; Lin, H.; Zhai, H.; Miao, Z.; Zhang, Y.; Xu, K.; Guan, B.; Huang, H.; Zhang, H. Pd-doped TiO2 film sensors prepared by premixed stagnation flames for CO and NH3 gas sensing. Sens. Actuators B 2018, 261, 451–459. [Google Scholar] [CrossRef]
  30. Tamgadge, R.M.; Shukla, A. Fluorine-doped anatase for improved supercapacitor electrode. Electrochim. Acta 2018, 289, 342–353. [Google Scholar] [CrossRef]
  31. Qiu, C.; Shen, J.; Lin, J.; Liu, D.; Li, D.; Zhang, J.; Zhang, Z.; Lin, H.; Wang, X.; Fu, X. Construction of the rutile/anatase micro-heterophase junction photocatalyst from anatase by liquid nitrogen quenching method. ACS Appl. Energy Mater. 2021, 4, 10172–10186. [Google Scholar] [CrossRef]
  32. Malik, R.; Tomer, V.K.; Joshi, N.; Dankwort, T.; Lin, L.; Kienle, L. Au–TiO2-loaded cubic g-C3N4 nanohybrids for photocatalytic and volatile organic amine sensing applications. ACS Appl. Mater. Inter. 2018, 10, 34087–34097. [Google Scholar] [CrossRef]
  33. Li, Y.; Han, Y.; Song, X.; Li, T.; Liu, G.; Chen, Z.; Dong, Z.; Liu, Y. Sized dependence and microstructural defects on highly photocatalytic activity based on multisized CdTe quantum dots sensitized TiO2. Surf. Interface Anal. 2019, 51, 968–981. [Google Scholar] [CrossRef]
  34. Monamary, A.; Vijayalakshmi, K.; Jereil, S.D. Fe overlayered hybrid TiO2/ITO nanocomposite sensor for enhanced hydrogen sensing at room temperature by novel two step process. Sens. Actuators B 2019, 287, 278–289. [Google Scholar] [CrossRef]
  35. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997–1902023. [Google Scholar] [CrossRef]
  36. Ullah, K.; Meng, Z.-D.; Ye, S.; Zhu, L.; Oh, W.-C. Synthesis and characterization of novel PbS–graphene/TiO2 composite with enhanced photocatalytic activity. J. Ind. Eng. Chem. 2014, 20, 1035–1042. [Google Scholar] [CrossRef]
  37. Sun, Q.; Wu, S.; Li, K.; Han, B.; Chen, Y.; Pang, B.; Yu, L.; Dong, L. The favourable synergistic operation of photocatalysis and catalytic oxygen reduction reaction by a novel heterogeneous CoFe2O4-TiO2 nanocomposite. Appl. Surf. Sci. 2020, 516, 146142–146155. [Google Scholar] [CrossRef]
  38. Sagu, J.S.; Wijayantha, K.G.U.; Tahir, A.A. The pseudocapacitive nature of CoFe2O4 thin films. Electrochim. Acta 2017, 246, 870–878. [Google Scholar] [CrossRef] [Green Version]
  39. Thu Do, T.A.; Giang, H.T.; Van Huong, D.; Ngan, P.Q.; Thai, G.H.; Thu, D.T.; Lam, T.D. Correlation between photoluminescence spectra with gas sensing and photocatalytic activities in hierarchical ZnO nanostructures. RSC Adv. 2017, 7, 9826–9832. [Google Scholar] [CrossRef]
  40. Cheng, X.; Xu, Y.; Gao, S.; Zhao, H.; Huo, L. Ag nanoparticles modified TiO2 spherical heterostructures with enhanced gas-sensing performance. Sens. Actuators B 2011, 155, 716–721. [Google Scholar] [CrossRef]
  41. Zhang, X.; Song, D.; Liu, Q.; Chen, R.; Hou, J.; Liu, J.; Zhang, H.; Yu, J.; Liu, P.; Wang, J. Designed synthesis of Ag-functionalized Ni-doped In2O3 nanorods with enhanced formaldehyde gas sensing properties. J. Mater. Chem. C 2019, 7, 7219–7229. [Google Scholar] [CrossRef]
  42. Nakarungsee, P.; Srirattanapibul, S.; Issro, C.; Tang, I.M.; Thongmee, S. High performance Cr doped ZnO by UV for NH3 gas sensor. Sens. Actuators A 2020, 314, 112230–112261. [Google Scholar] [CrossRef]
  43. Raghu, A.V.; Karuppanan, K.K.; Pullithadathil, B. Controlled carbon doping in anatase TiO2(101) facets: Superior trace-level ethanol gas sensor performance and adsorption kinetics. Adv. Mater. Interfaces 2019, 6, 1801714–1801726. [Google Scholar] [CrossRef]
  44. Ullattil, S.G.; Narendranath, S.B.; Pillai, S.C.; Periyat, P. Black TiO2 nanomaterials: A review of recent advances. Chem. Eng. J. 2018, 297–395. [Google Scholar] [CrossRef]
  45. Šutka, A.; Käämbre, T.; Pärna, R.; Döbelin, N.; Vanags, M.; Smits, K.; Kisand, V. Ag sensitized TiO2 and NiFe2O4 three-component nanoheterostructures: Synthesis, electronic structure and strongly enhanced visible light photocatalytic activity. RSC Adv. 2016, 6, 18834–18842. [Google Scholar] [CrossRef]
  46. Zhao, J.; Xu, J.; Jian, X.; Xu, J.; Gao, Z.; Song, Y.-Y. NIR light-driven photocatalysis on amphiphilic TiO2 nanotubes for controllable drug release. ACS Appl. Mater. Inter. 2020, 12, 23606–23616. [Google Scholar] [CrossRef]
  47. Shaikh, S.K.; Ganbavle, V.V.; Inamdar, S.I.; Rajpure, K.Y. Multifunctional zinc oxide thin films for high-performance UV photodetectors and nitrogen dioxide gas sensors. RSC Adv. 2016, 6, 25641–25650. [Google Scholar] [CrossRef]
  48. Lei, T.; Rao, Z.; Zhang, S.; Cai, S.; Xie, C. The irreversible R-T curves of metal oxide gas sensor under programmed temperature cycle. Sens. Actuators B 2016, 235, 481–491. [Google Scholar] [CrossRef]
  49. Tonezzer, M.; Le, D.T.T.; Iannotta, S.; Van Hieu, N. Selective discrimination of hazardous gases using one single metal oxide resistive sensor. Sens. Actuators B 2018, 277, 121–128. [Google Scholar] [CrossRef]
  50. Wang, Z.; Zhu, L.; Sun, S.; Wang, J.; Yan, W. One-dimensional nanomaterials in resistive gas sensor: From material design to application. Chemosensors 2021, 9, 198. [Google Scholar] [CrossRef]
  51. Zhu, L.; Li, Y.; Zeng, W. Hydrothermal synthesis of hierarchical flower-like ZnO nanostructure and its enhanced ethanol gas-sensing properties. Appl. Surf. Sci. 2018, 427, 281–287. [Google Scholar] [CrossRef]
  52. Joshi, N.; da Silva, L.F.; Shimizu, F.M.; Mastelaro, V.R.; M’Peko, J.C.; Lin, L.; Oliveira, O.N., Jr. UV-assisted chemiresistors made with gold-modified ZnO nanorods to detect ozone gas at room temperature. Mikrochim. Acta 2019, 186, 418–427. [Google Scholar] [CrossRef]
  53. Wang, Z.; Gao, S.; Fei, T.; Liu, S.; Zhang, T. Construction of ZnO/SnO2 heterostructure on reduced graphene oxide for enhanced nitrogen dioxide sensitive performances at room temperature. ACS Sens. 2019, 4, 2048–2057. [Google Scholar] [CrossRef] [PubMed]
  54. Bai, S.; Tian, K.; Fu, H.; Feng, Y.; Luo, R.; Li, D.; Chen, A.; Liu, C.C. Novel α-Fe2O3/BiVO4 heterojunctions for enhancing NO2 sensing properties. Sens. Actuators B 2018, 268, 136–143. [Google Scholar] [CrossRef]
  55. Zhang, L.; Khan, K.; Zou, J.; Zhang, H.; Li, Y. Recent advances in emerging 2D material-based gas sensors: Potential in disease diagnosis. Adv. Mater. Interfaces 2019, 6, 1901329–1901356. [Google Scholar] [CrossRef]
  56. Noh, J.; Kwon, S.H.; Park, S.; Kim, K.K.; Yoon, Y.J. TiO2 nanorods and Pt nanoparticles under a UV-LED for an NO2 gas sensor at room temperature. Sensors 2021, 21, 1826. [Google Scholar] [CrossRef] [PubMed]
  57. Zhong, F.; Wu, Z.; Guo, J.; Jia, D. Ni-doped ZnS nanospheres decorated with Au nanoparticles for highly improved gas sensor performance. Sensors 2018, 18, 2882. [Google Scholar] [CrossRef] [Green Version]
  58. Xue, D.; Wang, P.; Zhang, Z.; Wang, Y. Enhanced methane sensing property of flower-like SnO2 doped by Pt nanoparticles: A combined experimental and first-principle study. Sens. Actuators B 2019, 296, 126710–126720. [Google Scholar] [CrossRef]
  59. Mirzaei, A.; Bang, J.H.; Oum, W.; Kwon, Y.J.; Kim, J.H.; Choi, S.W.; Kim, S.S.; Kim, H.W. Selective H2S-sensing performance of Si nanowires through the formation of ZnO shells with Au functionalization. Sens. Actuators B 2019, 289, 1–14. [Google Scholar]
  60. Horprathum, M.; Srichaiyaperk, T.; Samransuksamer, B.; Wisitsoraat, A.; Eiamchai, P.; Limwichean, S.; Chananonnawathorn, C.; Aiempanakit, K.; Nuntawong, N.; Patthanasettakul, V. Ultrasensitive hydrogen sensor based on Pt-decorated WO nanorods prepared by glancing-angle dc magnetron sputtering. ACS Appl. Mater. Inter. 2014, 6, 22051–22060. [Google Scholar] [CrossRef]
  61. Yang, X.; Fu, H.; Tian, Y.; Xie, Q.; Xiong, S.; Han, D.; Zhang, H.; An, X. Au decorated In2O3 hollow nanospheres: A novel sensing material toward amine. Sens. Actuators B 2019, 296, 126696–126706. [Google Scholar] [CrossRef]
  62. Yang, X.; Fu, H.; Zhang, L.; An, X.; Xiong, S.; Jiang, X.; Yu, A. Enhanced gas sensing performance based on the fabrication of polycrystalline Ag@TiO2 core-shell nanowires. Sens. Actuators B 2019, 286, 483–492. [Google Scholar] [CrossRef]
  63. Reddeppa, M.; Mitta, S.B.; Park, B.-G.; Kim, S.-G.; Park, S.H.; Kim, M.-D. DNA-CTMA functionalized GaN surfaces for NO2 gas sensor at room temperature under UV illumination. Org. Electron. 2019, 65, 334–340. [Google Scholar] [CrossRef]
  64. Zhang, L.; Li, Z.; Liu, J.; Peng, Z.; Zhou, J.; Zhang, H.; Li, Y. Optoelectronic gas sensor based on few-layered InSe nanosheets for NO2 detection with ultrahigh antihumidity ability. Anal. Chem. 2020, 92, 11277–11287. [Google Scholar] [CrossRef]
  65. Mahala, C.; Sharma, M.D.; Basu, M. 2D Nanostructures of CoFe2O4 and NiFe2O4: Efficient Oxygen Evolution Catalyst. Electrochim. Acta 2018, 273, 462–473. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffraction (XRD) patterns and (b) N2 adsorption–desorption isotherms, with pore size distributions (inset) of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
Figure 1. (a) X-ray diffraction (XRD) patterns and (b) N2 adsorption–desorption isotherms, with pore size distributions (inset) of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
Nanomaterials 11 03261 g001
Figure 2. Scanning electron microscope (SEM) and transmission electron microscope (TEM) morphologies of TiO2−Ag (a,d), CoFe2O4−Ag (b,e), and TiO2−CoFe2O4−Ag (c,f).
Figure 2. Scanning electron microscope (SEM) and transmission electron microscope (TEM) morphologies of TiO2−Ag (a,d), CoFe2O4−Ag (b,e), and TiO2−CoFe2O4−Ag (c,f).
Nanomaterials 11 03261 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra: (a) total spectrum, (b) Ti 2p, (c) O 1s, and (d) Ag 3d.
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra: (a) total spectrum, (b) Ti 2p, (c) O 1s, and (d) Ag 3d.
Nanomaterials 11 03261 g003
Figure 4. (a) Photoluminescence (PL) spectra; (b) ultraviolet–visible diffuse reflectance spectra (UV–vis DRS); and (c,d) optical band gap of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
Figure 4. (a) Photoluminescence (PL) spectra; (b) ultraviolet–visible diffuse reflectance spectra (UV–vis DRS); and (c,d) optical band gap of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
Nanomaterials 11 03261 g004
Figure 5. (a) Gas-sensitive response test to toluene at different temperatures (n = 3), (b) gas-sensitive response test to toluene at 360–455 °C (n = 3), (c) response/recovery time (n = 3), and (d) dynamic response of TiO2−CoFe2O4−Ag to toluene gas at 263 °C.
Figure 5. (a) Gas-sensitive response test to toluene at different temperatures (n = 3), (b) gas-sensitive response test to toluene at 360–455 °C (n = 3), (c) response/recovery time (n = 3), and (d) dynamic response of TiO2−CoFe2O4−Ag to toluene gas at 263 °C.
Nanomaterials 11 03261 g005
Figure 6. (a) Selectivity of five gases, (b) gas sensitivity test of toluene gas in different relative humidity (RH) environments, (c) repeatability and (d) long-term stability for toluene gas of TiO2−CoFe2O4−Ag at 263 °C (n = 3).
Figure 6. (a) Selectivity of five gases, (b) gas sensitivity test of toluene gas in different relative humidity (RH) environments, (c) repeatability and (d) long-term stability for toluene gas of TiO2−CoFe2O4−Ag at 263 °C (n = 3).
Nanomaterials 11 03261 g006
Figure 7. (a) Test of gas sensitivity of TiO2−CoFe2O4−Ag to toluene at 263 °C in darkness and under the light of different wavelengths, and (b) dynamic response curve of TiO2−CoFe2O4−Ag to toluene gas in darkness and under UV irradiation.
Figure 7. (a) Test of gas sensitivity of TiO2−CoFe2O4−Ag to toluene at 263 °C in darkness and under the light of different wavelengths, and (b) dynamic response curve of TiO2−CoFe2O4−Ag to toluene gas in darkness and under UV irradiation.
Nanomaterials 11 03261 g007
Figure 8. (a) Band structure diagram of TiO2, CoFe2O4, and Ag. (b) Band structure diagram of TiO2−CoFe2O4−Ag.
Figure 8. (a) Band structure diagram of TiO2, CoFe2O4, and Ag. (b) Band structure diagram of TiO2−CoFe2O4−Ag.
Nanomaterials 11 03261 g008
Table 1. Physical properties of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
Table 1. Physical properties of TiO2−Ag, CoFe2O4−Ag, and TiO2−CoFe2O4−Ag.
SampleSpecific Surface Area
(m2/g)
Pore Volume
(cm3/g)
Pore Diameter
(nm)
TiO2−Ag63.210.32420.21
CoFe2O4−Ag102.750.1558.84
TiO2−CoFe2O4−Ag59.040.30419.67
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, W.; Zhang, L.; Kang, Y.; Yu, F. Light-Excited Ag-Doped TiO2−CoFe2O4 Heterojunction Applied to Toluene Gas Detection. Nanomaterials 2021, 11, 3261. https://doi.org/10.3390/nano11123261

AMA Style

Wang W, Zhang L, Kang Y, Yu F. Light-Excited Ag-Doped TiO2−CoFe2O4 Heterojunction Applied to Toluene Gas Detection. Nanomaterials. 2021; 11(12):3261. https://doi.org/10.3390/nano11123261

Chicago/Turabian Style

Wang, Wenhao, Lu Zhang, Yanli Kang, and Feng Yu. 2021. "Light-Excited Ag-Doped TiO2−CoFe2O4 Heterojunction Applied to Toluene Gas Detection" Nanomaterials 11, no. 12: 3261. https://doi.org/10.3390/nano11123261

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