Next Article in Journal
Influence of the Phosphor Layer Composition on Flexible Electroluminescent Device Performance
Previous Article in Journal / Special Issue
Research and Development of Online Monitoring Protection Sensors for Paper Books Based on TiO2 NT/MoS2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development and Application of a Nano-Gas Sensor for Monitoring and Preservation of Ancient Books in the Library

1
Library, Lingnan Normal University, Zhanjiang 524048, China
2
Key Laboratory of Advanced Coating and Surface Engineering, Lingnan Normal University, Zhanjiang 524048, China
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(5), 553; https://doi.org/10.3390/coatings14050553
Submission received: 21 March 2024 / Revised: 20 April 2024 / Accepted: 23 April 2024 / Published: 30 April 2024
(This article belongs to the Special Issue Current Trends in Coatings and Films for Optical Sensors)

Abstract

:
Monitoring the gas composition in library environments is crucial for the preservation of ancient books. In this study, TiO2 NTs/CNTs composites were synthesized via a hydrothermal method and utilized as nano-gas sensors for NO2 detection. The surface morphology and element composition of the samples were characterized using scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS). Additionally, the gas sensitivity of the prepared TiO2 nanocomposites was evaluated at different temperatures, both with and without ultraviolet light irradiation. The results demonstrate that the synthesized TiO2 NTs/CNTs samples exhibit a large specific surface area due to their titanium dioxide nanotubes (TiO2 NTs) and carbon nanotubes (CNTs) composition. Moreover, these samples display excellent gas sensitivity under ultraviolet light irradiation at temperatures of 120 °C. Compared to uncomposited and non-ultraviolet light irradiated samples, the sensor response rate is significantly improved, enabling effective monitoring of NO2 gas in library environments conducive to preserving ancient books. Overall, our findings highlight that the developed TiO2 NTs/CNTs nano gas sensor holds great potential for monitoring and safeguarding ancient books.

1. Introduction

As a crucial component of cultural heritage, ancient books and literature embody the legacy and evolution of exceptional traditional culture. To ensure their continuous preservation, significant time and effort must be invested [1]. Comprehensive protection of ancient bookstores necessitates monitoring internal corrosive and oxidizing gases. NO2 dissolves in water to form nitric acid, which not only exhibits acidity but also strong oxidation potential. Acidic conditions can degrade paper cellulose, leading to book acidification, while the potent oxidizing properties of NO2 can cause oxidative decomposition of paper cellulose [2]. Detection methods for these gases encompass selected ion flow tube mass spectrometry, gas chromatography–mass spectrometry, proton transfer reaction mass spectrometry, and gas sensor detection technology [3,4]. Among these, gas sensor detection technology is gaining popularity due to its affordability and ease of use [5]. Therefore, developing rapid and accurate NO2 gas sensors with low cost, low power consumption, and excellent sensitivity is critical for gas-monitoring equipment in ancient bookstores.
The types of gas sensors primarily include electrochemical, contact combustion, light reaction, and metal-oxide semiconductor (MOS) gas sensors. Among these, MOS gas sensors have garnered significant attention due to their advantageous features such as portability, high sensitivity, good stability, and low preparation cost [6]. MOS sensors have been used to detect toxic and harmful gases in libraries, but the problems of low sensitivity and high operating temperature have not been completely solved. TiO2 exhibits ultra-high chemical stability, non-toxicity, and excellent thermal stability at elevated temperatures. It finds extensive application in the field of gas-sensitive sensor research; however, TiO2 suffers from a wide band gap (3.2 eV) and low efficiency when utilizing natural light [7]. Carbon nanotubes (CNTs) possess remarkable chemical stability, mechanical durability, exceptional electrical properties, and an extremely high specific surface area. Moreover, the electrical characteristics of CNTs are easily influenced by favorable changes in surface adsorption [8], while their gas sensitivity can be enhanced under ultraviolet light irradiation [9]. Consequently, employing CNTs as sensing materials in gas sensors can yield heightened sensitivity towards electrons or gases with strong electron-withdrawing properties (e.g., NH3 gas and NO2 gas) even at room temperature. Zhao et al. synthesized In2O3 nanowires through solvothermal method, which showed a high response value of 54.6 to 1 ppm NO2 at 150 °C [10]. Liu et al. utilized the atomic layer deposition to prepare sensor fabricated on nanotubes, which exhibited good response and acceptable response time of 65 s to 0.5 ppm NO2 under UV illumination [11]. However, the above reported methods for nanostructured materials have weaknesses of intricate manipulation, the usage of UV illumination and noble metals also increases in sensor costs [12]. Hence, It is necessary to invent the NO2 sensor that can quickly and sensitively detect the trace amount of NO2 in the air at low operating temperature.
In this manuscript, we employed a low-cost hydrothermal method and non-metallic doping technique to synthesize composite products of TiO2 nanotubes/C nanotubes (TiO2 NTs/CNTs). The gas sensing properties of the composite products were systematically investigated under different temperatures and ultraviolet light irradiation. By incorporating heterostructures and employing photoactivation techniques, significant improvements in the gas sensing performance are achieved [13,14]. Moreover, we developed a gas sensor based on multidimensional nanostructures to improve its sensitivity and lower the detection temperature for NO2 gas, thereby effectively reducing the risk of damage to ancient books stored in libraries.

2. Materials and Methods

2.1. Preparation of TiO2NTs and TiO2NTs/CNTs

To start, 2 g TiO2 nanoparticles and 80 mL 10 M NaOH solution were added to the beaker, and they were transferred to the reaction kettle for hydrothermal reaction at 130 °C for 24 h with ultrasonic stirring for 30 min each. The solution was washed to neutral and the precipitation was taken out and stirred in 5% HCl solution for 1 h. The white precipitation was washed to neutral, dried at 70 °C for several hours, and calcined at different temperatures to obtain TiO2 NTs. The CNTs were ultrasonically dissolved in an alcohol solution and magnetically stirred for 30 min to form a uniform CNT solution (0.1 mg/mL). TiO2 nanoparticles, HCl solution, CNTs, sodium hydroxide (NaOH), and absolute ethanol (CH3CH2OH) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) without further purification.
Deionized water was added to the beaker, followed by the addition of titanium dioxide, carbon nanotubes, and sodium hydroxide in sequence. To achieve homogeneity, the solution was stirred using a magnetic stirrer for 30 min. Subsequently, ultrasonic treatment was conducted on the solution for 30 min to obtain a mixed solution which was then transferred into a 100 mL liner and placed inside a reaction kettle. The reaction kettle was positioned in a drying oven set at 150 °C for hydrothermal reaction. After 24 h, the reaction concluded and the bottom layer of the liner was extracted as the first composite product [15]. Subsequently, it was soaked in 30 mL of dilute hydrochloric acid with pH adjusted to 1 for another 24 h. Further centrifugal cleaning ensured neutralization before drying and grinding were carried out. The ground powder underwent calcination in a muffle furnace set at 400 °C for two hours, resulting in obtaining a TiO2 NTs/CNTs composite product. The final product contained about 3% carbon nanotubes, and experimental tests have shown that it exhibits excellent gas sensing properties.

2.2. Characterization of TiO2 NTs/CNTs

The morphology of the sample was observed and analyzed using scanning electron microscopy (SEM/EDS, SU3800, Hitachi-Hightech, Tokyo, Japan). X-ray diffraction (XRD, RINT2000) was employed to analyze the phase of TiO2 NTs/CNTs. In XRD measurements, a Ni-filtered Cu source with a wavelength of 1.54 nm (Kα) was used to generate X-rays and the measurements were taken between 10° and 90° with a step size of 0.01°.

2.3. Fabrication of Gas Monitoring Sensors

The gas-sensing device was custom-built based on the description provided in a previous publication [14]. Firstly, grind the sample and set it aside. Next, place the fork finger electrode into the sample bottle and extract 10 g of the ground sample. Add anhydrous ethanol to the sample bottle and subject it to ultrasonic treatment for 10 min, resulting in a suspension of the sample in anhydrous ethanol. Subsequently, heat the sample in a water bath to evenly precipitate it onto the fork finger electrode. The resistance of the sample is measured using a multi-function electricity meter, with only components exhibiting resistance at MΩ level considered as gas-sensitive components. We employed a self-built gas-sensitive test bench consisting of an air-expelling system, an air-filling system, a test chamber, tungsten steel needles, a heating table, gas-sensitive components, a resistance tester, and a computer (Figure 1). To initiate testing procedures, insert the fork finger electrode into the vacuum chamber while crossing pairs of tungsten steel needles.
The central gas-sensitive element is irradiated with UV light (365 nm), and a temperature controller controls the experimental environment. The test gas from the gas source is controlled for flow rate and total input using a D07–19BM mass flow controller (sevenstar, Beijing, China) and a D08–8CM flow meter (sevenstar, Beijing, China). The change in the resistance value is received and transmitted using the Keithley Model 2450 digital source meter (Tektronix Inc., Beaverton, OR, USA) by the vacuum four-probe resistance testing system. After the set parameters reach a stable value, the gas-sensitive material is tested for gas sensing. The resistance change curve is obtained by detecting the resistance value before and after introducing the gas. The measurement was conducted at room temperature (T = 25.0 ± 0.5 °C) and dry air (relative humidity (RH) = 15.5 ± 1.7%) during the test.

3. Results and Discussion

3.1. Morphology and Phase Analysis of TiO2 NTs/CNTs

Figure 2 illustrates the surface morphology of the sample at different calcination temperatures. As depicted, at a calcination temperature of 300 °C, the TiO2 NTs/CNTs product exhibits a disordered tubular structure with some unformed accumulations still present. At a calcination temperature of 400 °C, compared to the product calcined at 300 °C, a significant number of intertwined TiO2 NTs and CNTs and are observed in the product, forming numerous pore structures that result in a larger specific surface area for TiO2 NT/CNTs. However, when the temperature reaches 500 °C, the nanotube morphology of TiO2 NTs/CNTs completely disappears and the sample transforms into nanoparticles with a diameter above 30 nm. As can be seen from EDS images shown in Figure 3, the weight ratio of carbon nanotubes to TiO2 was found to be 3%.
As depicted in Figure 4, TiO2 NTs exhibit a prominent diffraction peak at 2θ = 27.45° [16], corresponding to the crystal structure of TiO2 (110). Notably, as illustrated in the figure, an increase in calcination temperature leads to a progressively sharper characteristic diffraction peak for anatase, indicating enhanced crystallization with elevated calcination temperatures. Simultaneously, it can be observed from the figure that when calcination temperature is below 500 °C, hydrothermally treated TiO2 NTs lack the rutile diffraction peak. This suggests that rutile-phase formation does not occur in TiO2 NTs under calcination temperatures below 500 °C and implies complete destruction of the rutile–anatase (P90) structure of raw material by strong alkali solution. Energy spectrum analysis of the composite product comprising TiO2 NTs/CNTs reveals presence of titanium (Ti), carbon (C), and oxygen (O), confirming stoichiometric proportionality within the composite product. In Figure 4b, all diffraction peaks are consistent with the peak positions of TiO2 nanotubes. The sharpness of the diffraction peaks indicates that the crystallinity of TiO2 remains high in the presence of carbon nanotubes. However, no distinct characteristic peaks of carbon nanotubes are observed in the figure, indicating that in cases where the content is only 3% and 5%, the disorderly accumulation of carbon nanotubes causes significant damage, resulting in relatively small or even absent diffraction peaks.

3.2. Gas-Sensitive Properties of TiO2 NTs/CNTs

The gas-sensing response of the devices to oxidizing gases can be defined as S = Rg/Ra, where Rg and Ra represent the resistance of the gas sensor in the detected gas and air, respectively. The response time and recovery time were defined by the time for the gas sensor to achieve 90% of the total response change with the exposure and evacuation of the detected gases, respectively. The response recovery curves of TiO2 material with NO2 gas concentrations of 2, 4, 6, 8, and 10 ppm are presented in Figure 5. The resistance of the material shows minimal change after exposure to NO2 gas. The sensor exhibits a gas-sensing response of approximately 1. At 10 ppm, the sensor demonstrates maximum gas-sensing response with a response time of 91 s and a recovery time of 84 s. This study selected this as the research basis to test whether the response time and recovery time of TiO2 NTs/CNTs nanocomposites to NO2 gas at different temperatures were improved, and whether the gas-sensitive response of the sensor was improved.
The response recovery curves of the TiO2 NTs/CNTs nanocomposites following exposure to 2, 4, 6, 8, and 10 ppm NO2 gas at 120 °C exhibited an enhanced gas-sensing response compared to Figure 5. As shown in Figure 6, the nanocomposites displayed the best response to 8 ppm NO2 gas with a gas-sensing response of 2.28, a response time of 97 s, and a recovery time of 225 s. In comparison to pure TiO2 materials without nanocomposites products, both the response time and recovery time are prolonged by about 6 s and 171 s, respectively, compared to Figure 5. However, the gas-sensing response of the sensor is increased by 1.3 times.
Figure 7 shows the gas response recovery curve of TiO2 NTs/CNTs in response to NO2 with concentrations of 2, 4, 6, 8, and 10 ppm at 200 °C. With the increase in NO2 gas concentrations, the gas-sensing response of the sensor reaches the maximum value of 12 at 10 ppm, and the response time is 163 s. The response time is extended by 66 s and the recovery time is 107 s compared with that at 120 °C, and the recovery time is reduced by 85 s. Compared with pure TiO2 materials, although the response time and recovery time are slightly longer, the gas-sensing response of the sensor is greatly improved.
Figure 8 shows the response recovery curve of TiO2 NTs/CNTs materials at 120 °C with NO2 gas concentrations of 2, 4, 6, 8, and 10 ppm under ultraviolet irradiation. As can be seen from the figure, the gas-sensing response of the sensor increases with the increase in NO2 gas concentrations, and reaches the maximum value when the gas concentration is 10 ppm (S = Ra/Rg = 6.57). The response time of the sensor is 234 s, and the recovery time is 133 s. Compared with sensors without UV irradiation at 120 °C, the gas-sensing performance is significantly improved, and the gas-sensing response is increased by about six times, although the response time and recovery time are prolonged. Compared with sensors without ultraviolet irradiation at 200 °C, the gas-sensing response decreased, and the response time and recovery time were relatively long. The response time and recovery time were extended by 71 s and 26 s compared with 200 °C without ultraviolet irradiation, which failed to achieve good gas-sensitive performance under low temperature and ultraviolet irradiation at the same time. This shows that at the same temperature, the gas sensitive performance of the sensor can be improved under ultraviolet irradiation, but the response recovery time cannot be shortened by ultraviolet irradiation [17].
The continuous transient response recovery of the TiO2 NTs array to NO2 at different temperatures is shown in Figure 9. The test demonstrates that the TiO2 NTs array exhibits excellent selectivity for NO2 at a temperature of 320 °C [18]. As depicted in Figure 9a, when the temperature is set at 320 °C and the concentration of NO2 increases from 168 to 1020 ppm, the response value of the TiO2 NTs array to NO2 rises from 100 to 400. Upon removal of the NO2, the resistance of the TiO2 NTs array increases and returns to its original state. This not only suggests that higher concentrations enhance sensitivity for detecting NO2 gas but also highlights good reversibility exhibited by TiO2 NTs. Even after multiple repeated operations, the response value can still return to its initial level. As illustrated in Figure 9b, when exposed to a concentration of 1020 ppm of NO2, it takes approximately 32 s for TiO2 NTs to respond, indicating a favorable response value towards NO2 gas at both a temperature of 320 °C and this particular concentration level. However, it requires around 90 s for complete recovery due to a strong linear correlation between the resistance of the nanocomposites and NO2 gas concentration during this process [19].
As shown in Figure 10, the curve showed a relatively stable trend during a long-term stability measurement of 30 days toward 8 ppm NO2 gas. Moreover, after 30 consecutive tests, no apparent decrease trend in response was observed for detecting 8 ppm NO2. These results demonstrate that the TiO2 NTs/CNTs-based sensor possesses good repeatability and long-term stability. In addition, the effects of environmental factors such as humidity, temperature fluctuations, and light exposure on the sensors are currently in the process for optimizing experimental parameters. The properties of TiO2 NTs/CNTs heterojunction nanostructure are potentially superior to their individual component because of greater active sites for adsorption of the target gas, which means a higher response.

4. Gas-Sensitive Reaction Mechanism

The gas monitoring can be explained by the combination of the adsorption–desorption model [20] and the energy band theory [21]. Firstly, TiO2 NTs array acts as an N-type semiconductor. In this study, a non-metallic material P-type semiconductor material CNT is selected for doping on the basis of a TiO2 NTs array. After doping, a heterojunction PN junction is formed at the interface of these two materials [21]. Moreover, electrons are carriers in N-type semiconductors and participate in conduction while holes are carriers in P-type semiconductors and also participate in conduction. As electrons move from an N-type semiconductor TiO2 NTs to P-type semiconductor CNTs, TiO2 NTs become positively charged whereas holes move from P-type semiconductor CNTs to N-type semiconductor TiO2 NTs resulting in negative charge accumulation on CNTs. Consequently, an electric field is established at the heterojunction which increases the Schottky barrier leading to the excellent gas-sensitive performance of the TiO2 NTs/CNT composite [22]. Owing to the relative position of the CNT and TiO2 CB edge [19], the electrons injected into the oxide can be shuttled into the graphitic network, as proposed for photo-injected carriers in TiO2/CNTs catalysts [23]. Such a process finally brings about a reduction in the h+ concentration in the CNTs, thus causing the electrical resistance of the nanocomposite to increase [24]. Upon contact with oxidizing gases (such as NO2), strong electron absorption occurs on the surface of CNTs causing tilting effect towards oxidizing gases. On a macroscopic scale, this results in a charge transfer between gas molecules and CNT surfaces, increasing hole carrier concentration and conductivity while losing electrons, thus enhancing sensitivity of CNT towards oxidizing gases [25].

5. Conclusions

In this study, TiO2 NTs/CNTs composites were synthesized via a hydrothermal method and utilized as nano gas sensors for NO2 detection. A heterojunction junction is formed at the interface of hydrothermal synthesis of TiO2 NTs and CNT nanocomposites, which increases the Schottky barrier leading to an excellent gas-sensitive performance. This improves the power and sensitivity of the sensor. Increasing calcination temperature enhances the crystal structure of TiO2 NTs, resulting in increased sensitivity towards NO2 gas detection with higher concentrations. At 200 °C, the sensor exhibits a 12-fold increase in power compared to 120 °C. Furthermore, under UV irradiation conditions, significant improvements are observed in the material’s gas-sensing performance compared to non-irradiated conditions at similar temperatures. However, increasing temperature and adding UV irradiation prolong both response and recovery times relative to initial values for this gas-sensitive sensor. This study has the potential to decrease production expenses and this sensor can be potentially applied for rapidly and precisely detecting NO2 in air quality monitoring. Future work will focus on optimizing test parameters for real-time monitoring of ancient book environments to enable timely and efficient protective measures.

Author Contributions

Formal analysis, J.W.; Investigation, J.W., Q.W., S.H., Z.C., W.Q. and Y.Y.; Writing—original draft, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Science and Technology Project of Zhanjiang (2022B01044, 2022A0100) and Special Fund for Science and Technology Innovation Strategy in Guangdong Province (pdih2022b0319).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, H. Research on the Protection Measures of Ancient Books in University Libraries in the New Era. J. Northwest Adult Educ. Coll. 2022, 1, 99. [Google Scholar]
  2. Lu, A. Harmful gases in the air to books and their relationship with the humidity of the library. Nandu Xuetan 1997, 2, 96–97. [Google Scholar]
  3. Borowik, P.; Adamowicz, L.; Tarakowski, R.; Siwek, K.; Grzywacz, T. Odor detection using an E-nose with a reduced sensor array. Sensors 2020, 20, 3542. [Google Scholar] [CrossRef] [PubMed]
  4. Tomic, M.; Setka, M.; Vojkuvka, L.; Vallejos, S. VOCs sensing by metal oxides, conductive polymers, and carbon-based materials. Nanomaterials 2021, 11, 552. [Google Scholar] [CrossRef]
  5. Mu, S.; Shen, W.; Lv, D.; Song, W.; Tan, R. Inkjet-printed MOS-based MEMS sensor array combined with one-dimensional convolutional neural network algorithm for identifying indoor harmful gases. Sens. Actuators A Phys. 2024, 369, 11520. [Google Scholar] [CrossRef]
  6. Shi, X.; He, Y.; Gong, T.; Xie, Y.; Chen, J.; Yang, X. Preparation of three-dimensional ZnO nanoflowers and their gas sensing performance to indoor polluted gases. Sens. Microsyst. 2022, 41, 10–13. [Google Scholar]
  7. Xie, L. Study on Metal-Modified Nano TiO2 and Its Gas Sensing Performance; Nanjing University of Aeronautics and Astronautics: Nanjing, China, 2021. [Google Scholar]
  8. Xu, J.; Chen, Y.; Li, Y.; Shen, J. Application of one-dimensional nanomaterials in gas sensors. Sens. Technol. 2005, 1, 4–6. [Google Scholar]
  9. Chen, G.; Paronyan, T.M.; Pigos, E.M.; Harutyunyan, A.R. Enhanced gas sensing in pristine carbon nanotubes under continuous ultraviolet light illumination. Sci. Rep. 2012, 2, 343. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, S.; Shen, Y.; Zhou, P.; Hao, F.; Xu, X.; Gao, S.; Wei, D.; Ao, Y.; Shen, Y. Enhanced NO2 sensing performance of ZnO nanowires functionalized with ultra-fine In2O3 nanoparticles. Sens. Actuators B Chem. 2020, 308, 127729. [Google Scholar] [CrossRef]
  11. Liu, Y.; Zhang, J.; Li, G.; Liu, J.; Liang, Q.; Wang, H.; Zhu, Y.; Gao, J.; Lu, H. In2O3-ZnO nanotubes for the sensitive and selective detection of ppb-level NO2 under UV irradiation at room temperature. Sens. Actuators B Chem. 2022, 355, 131322. [Google Scholar] [CrossRef]
  12. Dang, T.; Nguyen, T.T.O.; Truong, T.H.; Le, A.T.; Nguyen, T.D. Facile synthesis of different ZnO nanostructures for detecting sub-ppm NO2 gas. Mater. Today Commun. 2020, 22, 100826. [Google Scholar] [CrossRef]
  13. Liang, Y.; Wang, H.; Sanchez Casalongue, H.; Chen, Z.; Dai, H. TiO2 nanocrystals grown on graphene as advanced photocatalytic hybrid materials. Nano Res. 2010, 3, 701–705. [Google Scholar] [CrossRef]
  14. Li, M.; Zou, C.; Liang, F.; Hou, E.; Lin, J. Preparation and gas-sensing performance under photoactivation of TiO2 nanotubes/carbon nanotubes/ZnS quantum dots gas-sensitive materials. Mater. Today Commun. 2023, 36, 106487. [Google Scholar] [CrossRef]
  15. Guo, R.-B. Preparation of g-C3N4 Composite Nanomaterials and Their Photocatalytic Performance; Nanchang Hangkong University: Nanchang, China, 2018. [Google Scholar]
  16. Huang, D.-G.; Mo, Z.-H.; Quan, S.-Q.; Yang, T.-Z.; Liu, Z.-B.; Liu, M. Preparation and photocatalytic reduction performance of graphene/nano-TiO2 composites. Acta Mater. Compos. Sin. 2016, 33, 155–162. [Google Scholar]
  17. Giampiccolo, A.; Tobaldi, M.D.; Leonardi, G.S.; Murdoch, B.J.; Seabra, M.P.; Ansell, M.P.; Ansell, M.P.; Neri, G.; Ball, R.J. Sol gel graphene/TiO2 nanoparticles for the photocatalytic-assisted sensing and abatement of NO2. Appl. Catal. B Environ. 2019, 243, 183–194. [Google Scholar] [CrossRef]
  18. Tong, X. Preparation and Gas Sensing Performance of TiO2 Nanotube Array Membranes for the Main Components of Papermaking Pollution Gases; South China University of Technology: Guangzhou, China, 2017. [Google Scholar]
  19. Li, J.; Lu, Y.; Ye, Q.; Cinke, M.; Han, J.; Meyyappan, M. Carbon Nanotube Sensorsfor Gas and Organic Vapor Detection. Nano Lett. 2023, 3, 929–933. [Google Scholar] [CrossRef]
  20. Chang, J.; Jiang, D.-G.; Zhan, Z.L.; Song, W.H. Overview of semiconductor metal oxide gas sensing materials. Sens. World 2003, 9, 14–18. [Google Scholar]
  21. Tian, K. Construction of Metal Oxide Semiconductor Heterojunction and Its Gas Sensing and Photoelectrochemical Performance and Mechanism; Beijing University of Chemical Technology: Beijing, China, 2023. [Google Scholar]
  22. Santangelo, S.; Faggio, G.; Messina, G.; Fazio, E.; Neri, F.; Neri, G. On the hydrogen sensing mechanism of Pt/TiO2/CNTs based devices. Sens. Actuators B 2013, 178, 473–484. [Google Scholar] [CrossRef]
  23. Xu, Y.J.; Zhuang, Y.; Fu, X. New insight for enhanced photocatalytic activity of TiO2 by doping carbon nanotubes: A case study on degradation of benzene and methyl orange. J. Phys. Chem. C 2010, 114, 2669–2676. [Google Scholar] [CrossRef]
  24. De Luca, L.; Donato, A.; Santangelo, S.; Faggio, G.; Messina, G.; Donato, N.; Neri, G. Hydrogen sensing characteristics of Pt/TiO2/MWCNTs composites. Int. J. Hydrogen Energy 2012, 37, 842–851. [Google Scholar] [CrossRef]
  25. Lee, J.H.; Kim, J.Y.; Nam, M.S.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Synergistic effect between UV light and PANI/Co3O4 content on TiO2 composite nanoparticles for room-temperature acetone sensing. Sens. Actuators B Chem. 2023, 375, 132868–132879. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of vacuum gas sensitivity test system: exhausting system; ① Vacuum pumping system; ② aeration system; ③ test chamber; ④ tungsten steel needle; ⑤ heating platform; ⑥ gas sensitive element; ⑦ resistance tester; ⑧ computer. (b) The picture of the manufactured sensor based on TiO2 NTs/CNTs.
Figure 1. (a) Schematic diagram of vacuum gas sensitivity test system: exhausting system; ① Vacuum pumping system; ② aeration system; ③ test chamber; ④ tungsten steel needle; ⑤ heating platform; ⑥ gas sensitive element; ⑦ resistance tester; ⑧ computer. (b) The picture of the manufactured sensor based on TiO2 NTs/CNTs.
Coatings 14 00553 g001
Figure 2. SEM images of TiO2 NTs/CNTs with different calcination temperatures: (a,b) 300 °C; (c,d) 400 °C; (e,f) 500 °C.
Figure 2. SEM images of TiO2 NTs/CNTs with different calcination temperatures: (a,b) 300 °C; (c,d) 400 °C; (e,f) 500 °C.
Coatings 14 00553 g002
Figure 3. EDS images for TiO2 NTs/CNTs with calcination temperatures of 300 °C.
Figure 3. EDS images for TiO2 NTs/CNTs with calcination temperatures of 300 °C.
Coatings 14 00553 g003
Figure 4. XRD patterns of TiO2 NTs with different calcination temperatures (a) and TiO2 NTs/CNTs with CNTs contents of 3% and 5% (b).
Figure 4. XRD patterns of TiO2 NTs with different calcination temperatures (a) and TiO2 NTs/CNTs with CNTs contents of 3% and 5% (b).
Coatings 14 00553 g004
Figure 5. The response–recovery curves of TiO2 at temperatures of 120 °C subsequent to the introduction of NO2 gases at concentrations of 2, 4, 6, 8, and 10 ppm.
Figure 5. The response–recovery curves of TiO2 at temperatures of 120 °C subsequent to the introduction of NO2 gases at concentrations of 2, 4, 6, 8, and 10 ppm.
Coatings 14 00553 g005
Figure 6. Response recovery curves of TiO2 NTs/CNTs nanocomposites at 120 °C and with concentrations of 2, 4, 6, 8, and 10 ppm.
Figure 6. Response recovery curves of TiO2 NTs/CNTs nanocomposites at 120 °C and with concentrations of 2, 4, 6, 8, and 10 ppm.
Coatings 14 00553 g006
Figure 7. The recovery curve of TiO2 NTs/CNTs in response to successive introduction of NO2 with different concentrations at a temperature of 200 °C.
Figure 7. The recovery curve of TiO2 NTs/CNTs in response to successive introduction of NO2 with different concentrations at a temperature of 200 °C.
Coatings 14 00553 g007
Figure 8. The response recovery curve of TiO2 NTs/CNTs under ultraviolet light irradiation at 120 °C after introduction of NO2 ranging from 2 to 10 ppm.
Figure 8. The response recovery curve of TiO2 NTs/CNTs under ultraviolet light irradiation at 120 °C after introduction of NO2 ranging from 2 to 10 ppm.
Coatings 14 00553 g008
Figure 9. (a) Transient response recovery of TiO2 NTs/CNTs to 168, 686, and 1020 ppm NO2 gas at 320 °C; (b) transient response of TiO2 NTs array to 1020 ppm NO2 gas at 320 °C.
Figure 9. (a) Transient response recovery of TiO2 NTs/CNTs to 168, 686, and 1020 ppm NO2 gas at 320 °C; (b) transient response of TiO2 NTs array to 1020 ppm NO2 gas at 320 °C.
Coatings 14 00553 g009
Figure 10. Stability test during a long-term of 30 days toward 8 ppm NO2 gas.
Figure 10. Stability test during a long-term of 30 days toward 8 ppm NO2 gas.
Coatings 14 00553 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, J.; Wang, Q.; He, S.; Chen, Z.; Qiu, W.; Yu, Y. Development and Application of a Nano-Gas Sensor for Monitoring and Preservation of Ancient Books in the Library. Coatings 2024, 14, 553. https://doi.org/10.3390/coatings14050553

AMA Style

Wang J, Wang Q, He S, Chen Z, Qiu W, Yu Y. Development and Application of a Nano-Gas Sensor for Monitoring and Preservation of Ancient Books in the Library. Coatings. 2024; 14(5):553. https://doi.org/10.3390/coatings14050553

Chicago/Turabian Style

Wang, Jia, Qingyu Wang, Susu He, Zhiyin Chen, Wentong Qiu, and Yunjiang Yu. 2024. "Development and Application of a Nano-Gas Sensor for Monitoring and Preservation of Ancient Books in the Library" Coatings 14, no. 5: 553. https://doi.org/10.3390/coatings14050553

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