Next Article in Journal
The Impact of 3D Prism Cavity for Enhanced Oil Recovery Using Different Nanomaterials
Next Article in Special Issue
Hierarchical Surface Structures and Large-Area Nanoscale Gratings in As2S3 and As2Se3 Films Irradiated with Femtosecond Laser Pulses
Previous Article in Journal
Experimental Research on Surface Quality of Titanium Rod Turned by Wire Electrical Discharge Turning Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

UV-Activated Au Modified TiO2/In2O3 Hollow Nanospheres for Formaldehyde Detection at Room Temperature

1
School of Microelectronics, Dalian University of Technology, Dalian 116024, China
2
School of Information and Control Engineering, Qingdao University of Technology, Qingdao 266520, China
3
School of Artificial Intelligence, Dalian University of Technology, Dalian 116024, China
4
Key Laboratory of Integrated Circuit and Biomedical Electronic System, Dalian University of Technology, Dalian 116023, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 4010; https://doi.org/10.3390/ma16114010
Submission received: 25 April 2023 / Revised: 19 May 2023 / Accepted: 24 May 2023 / Published: 26 May 2023
(This article belongs to the Special Issue Synthesis and Characterization of Semiconductor Nanomaterials)

Abstract

:
Au modified TiO2/In2O3 hollow nanospheres were synthesized by the hydrolysis method using the carbon nanospheres as a sacrificial template. Compared to pure In2O3, pure TiO2, and TiO2/In2O3 based sensors, the Au/TiO2/In2O3 nanosphere-based chemiresistive-type sensor exhibited excellent sensing performances to formaldehyde at room temperature under ultraviolet light (UV-LED) activation. The response of the Au/TiO2/In2O3 nanocomposite-based sensor to 1 ppm formaldehyde was about 5.6, which is higher than that of In2O3 (1.6), TiO2 (2.1), and TiO2/In2O3 (3.8). The response time and recovery time of the Au/TiO2/In2O3 nanocomposite sensor were 18 s and 42 s, respectively. The detectable formaldehyde concentration could go down as low as 60 ppb. In situ diffuse reflectance Fourier transform infrared spectroscopy (DRIFTS) was used to analyze the chemical reactions on the surface of the sensor activated by UV light. The improvement in the sensing properties of the Au/TiO2/In2O3 nanocomposites could be attributed to the nanoheterojunctions and electronic/chemical sensitization of the Au nanoparticles.

1. Introduction

Formaldehyde (HCHO) is one of the main air pollutants, and when the concentration of formaldehyde exceeds the standard, it can cause a variety of diseases and has a serious impact on the body organs, mainly manifested in strong stimulation, sensitization, and mutagenesis [1,2,3,4]. In serious cases, it can induce cancer. In view of this, the World Health Organization recommends that long-term exposure to formaldehyde should not exceed 0.8 ppb. It is necessary to develop a fast response sensor for formaldehyde [5,6,7,8]. Compared with traditional detection methods, the use of metal oxide semiconductor (MOS)-based sensors to detect formaldehyde has the advantages of high sensitivity, good stability, and low price, and is a research hotspot in the field of gas sensors.
However, most MOS-based sensors usually need to operate at high temperatures (200–450 °C), which is not conducive to the integration of the device, and the sensing material will be destroyed at high temperature, which will affect the service life of the sensor [9,10,11,12,13]. Due to the limitations of thermal excitation, the activation of these sensors by ultraviolet (UV) irradiation is a promising strategy, especially for the metal oxides with good photocatalytic properties such as TiO2, ZnO, WO3, In2O3, and SnO2, etc. Under UV activation, the sensing oxides absorb photons, and electrons in the oxides transition from low to high energy levels to generate electron–hole pairs [14,15,16]. The photogenerated electrons and holes increase the concentration of the charge carriers, thus reducing the operating temperature of the sensor [17,18,19]. Previous studies have found that TiO2 with a hollow microstructure showed good sensing performance to formaldehyde gas under UV activation. Titanium dioxide (TiO2) is a widely studied gas sensing material with n-type response, and the bandgap of TiO2 is about 3.2 eV [20,21]. TiO2 is widely used in optoelectronics, photocatalysis, gas detection, and other fields [22]. However, the sensitivity and response/recovery performance of pure TiO2 still need to be improved, and the baseline resistance of a TiO2-based chemiresistive-type sensor at room temperature is about tens of million of ohms, which is too high to be conveniently manufactured. A second phase with very small resistance needs to be introduced to reduce its baseline resistance. Introducing the second phase into a single phase sensing material is an important way to improve the gas sensing performance of sensors at present [23,24,25]. The formation of nanoheterojunctions play an important role in regulating the concentration of the carrier and the height of the barrier in the sensing materials, which is conducive to improving the conductivity of the sensing material and improving the gas sensing performance of the sensors [26,27,28,29]. Indium oxide (In2O3) is a novel n-type semiconductor, and the indirect bandgap of In2O3 is about 2.8 eV [30]. Its conductivity is almost the highest among representative semiconductor oxide sensing materials, and the resistance is only a few to tens of thousands of ohms [31,32]. However, there are few reports on In2O3-based sensors under UV illumination. For example, Wang et al. [33] integrated ultra-thin In2O3 films into GaInN/GaN UV-LED structures to create a UV-LED sensor. The results showed that the response of the sensor to 726 ppb O3 was about 10.2 at room temperature.
As the surface reaction mechanism in the gas sensing response is similar to the catalytic reaction at the gas–solid interface, noble metal modified metal oxides are also often used in gas sensing [34,35]. The introduction of noble metal nanoparticles into the sensing material as a catalyst can accelerate the response speed, improve the sensitivity, enhance the stability, and improve the selectivity. Loading noble metal nanoparticles uniformly on the surface of the sensing materials will promote the gas sensing properties [36,37,38]. The difference between the work function of noble metal and the support metal oxide semiconductor can promote the adsorption and desorption of the sensing material to the gas and accelerate the electron transfer. Kamble et al. [39] improved the performance of the sensor by modifying a WO3 thin film with Ag nanoparticles, and the response speed was improved by six times compared with the WO3 thin film based sensor. Gu et al. [40] reported a highly selective and sensitive HCHO sensor based on Au/In2O3 nanocomposites. The sensor had a high response (85.67) to 50 ppm HCHO at low operating temperature (100 °C), and the detection limit could reach 1.42 ppb. Au nanoparticles dispersed on the In2O3 surface provide more active sites for the selective adsorption reaction and reduce the activation energy, so Au/In2O3-based sensors have high sensitivity and good selectivity against HCHO at low temperature. The gas sensor based on noble metal modified sensing materials has advantages of good response and low power consumption. Au nanoparticles as catalysts can control the surface chemical reactions through selective adsorption, reduce the activation energy of formaldehyde surface chemical reactions, and make it easier for formaldehyde to react with the chemisorbed oxygen ions on the surface of sensing materials [41]. The decoration of noble metal nanoparticles on the MOSs can improve the reaction between chemisorbed oxygen ions and target gas molecules, finally, enhancing the sensor’s gas sensing properties.
In this work, Au modified TiO2/In2O3 hollow nanospheres were prepared by a simple hydrolysis method at low temperature. First, hollow In2O3 nanospheres were synthesized by the immersion method using carbon as sacrificial templates. Then, TiO2 nanoparticles were deposited on the surface of the obtained In2O3. Finally, Au nanoparticles were loaded on the TiO2/In2O3 by the chemical reduction method. The microstructure and gas sensing performance of the obtained Au modified TiO2/In2O3 nanocomposites were analyzed, and the results showed that the gas-sensing properties to formaldehyde of the Au/TiO2/In2O3 nanocomposite-based sensor were significantly improved under UV activation at room temperature.

2. Materials and Methods

2.1. Synthesis of the Au Modified TiO2/In2O3 Nanocomposites

All chemicals were analytical grade reagents and used without further purification. First, the carbon nanospheres used as hard templates were synthesized via the hydrothermal method. A total of 6.44 g of glucose (Aladdin, Shanghai, China) was dissolved into 65 mL of deionized water to form a solution at room temperature. Then, the solution was transferred into a 100 mL Teflon-lined stainless steel autoclave (Hongguan, Shanghai, China). The reaction temperature was 180 °C for 10 h. The as-obtained precipitates were collected by centrifugation [42]. Then, 5.864 g InCl3∙4H2O (Aladdin, Shanghai, China) was dissolved in 20 mL of deionized water under stirring at room temperature. Next, 200 mg as-obtained carbon templates were dispersed in the above solution, and then the suspension was stirred at room temperature for 12 h. The obtained precipitates were centrifugally washed and dried at 60 °C. The dried precipitates were calcined at 500 °C in a muffle furnace (Kejing, Fufei, Anhui, China) for 3 h to remove the carbon template.
The 50 mg as-obtained In2O3 sample was dispersed into a 25 mL ethanol/water mixture (Aladdin, Shanghai, China; the volume ratio was ethanol:water = 4:1) to form a suspension. Then, 0.3 mL tetrabutyl titanate (Aladdin, Shanghai, China) and 0.1 mL ammonia (Aladdin, Shanghai, China) were added into 25 mL of absolute ethanol (Aladdin, Shanghai, China) to form a solution, and the solution was slowly added to the In2O3 suspension. The mixture was stirred for 6 h at 80 °C in the water bath. Then, the obtained precipitates were washed by centrifugation and dried. Finally, the precipitates were calcined at 500 °C for 2 h.
The Au modified TiO2/In2O3 nanocomposites were obtained by the chemical reduction method [43]. First, 50 mg of the as-obtained TiO2/In2O3 sample was dispersed into 15 mL of deionized water to form a suspension, then 0.45 mL of 0.01 M chloroauric acid (Aladdin, Shanghai, China) and 1 mL of 0.01 M L-lysine solution (Aladdin, Shanghai, China) were added into the above suspension. After stirring for 30 min, 0.1 mL of 1 M sodium citrate solution (Aladdin, Shanghai, China) was added into the above mixture under continual stirring. The precipitate was washed with deionized water and absolute ethanol and dried at 60 °C for 12 h. Finally, the precipitates were calcined at 300 °C for 30 min in air to remove the lysine.

2.2. Characterization

The crystalline structures of the hollow In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites were examined by an X-ray diffractometer with Cu Kα radiation (XRD, Bruker D8 Advance, λ = 0.15406 nm, Germany) operated at 40 KV. The morphology was investigated by a field-emission scanning electronic microscope (FESEM, FEI Company, Hillsboro, OR, USA, QUANTA FEG 250) and transmission electron microscope (HRTEM, Hitachi H-800, Japan). The valence state was obtained by an X-ray photoelectron spectrometer (XPS, Physical electronics, PHI 5300, Germany) with an Al Kα = 280 eV excitation source. The pore structure was obtained by Mercury Injection Apparatus (BET, Micromeritics, Autopore 9620, Norcross, GA, USA). In situ diffuse reflectance Fourier transform manufacture spectra of the samples were measured on a Fourier transform infrared spectrometer equipped (in situ DRIFTS, BRUKER OPTICS, Tensor, Germany) at a resolution of 4 cm−1 by accumulating 40 scans.

2.3. Sensor Fabrication and Sensing Measurements

An appropriate amount of deionized water or absolute ethanol was added to the powdered samples and ground into a slurry in a mortar. The slurries were evenly coated on an alumina ceramic substrate printed with gold interdigital electrodes (Beirun, Changchun, China, 15 × 8 × 0.6 mm, finger width of 0.4 mm, finger spacing of 0.2 mm, 8 pairs of electrodes). The as-fabricated sensors were aged at 60 °C for at least 24 h before the gas sensing test to ensure stability. The gas sensing performances were researched by a static testing method, as shown in Figure 1. The test system consists of a signal acquisition system, digital multimeter (Agilent 34465A, Santa Clara, CA, USA), and test chamber. The as-fabricated sensors were placed in a test chamber with a capacity of 50 L. The Agilent 34465A digital multimeter was connected to the computer to record the change curve of the resistance value of the sensors. The gas sensing test process was performed at room temperature and excited by ultraviolet light (UV-LED, λ = 365 nm, power density = 2.5 mW/cm2) at a distance of about 0.5 cm. In the traditional method, the front window of the test chamber is closed to ensure that the chamber is sealed, afterward, the baseline resistance of the sensor was stabilized, and the target gases or liquids were injected into the chamber through a microinjector (Collect, Zhengzhou, Henan, China). The effects of accompanying water in the formaldehyde solutions are discussed in the Supplementary Materials (Table S1 and Figure S1). When the sensor response signal is stable, the front window of the test chamber is opened to exhaust the gas and test the recovery properties of the sensors.
The sensor response (S) was defined as:
S = Rair/Rgas
where Rair represents the resistance of the sensors in the air and Rgas represents the resistance of the sensor in targeted gas vapors.

3. Results

3.1. Structural and Morphological Characteristics

The phase and composition of the as-prepared samples were characterized by X-ray diffraction. Figure 2a shows the XRD patterns of the hollow In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites. From the results, all the XRD peaks of the TiO2/In2O3 nanocomposites could be well-assigned to the anatase phase TiO2 (JCPDS card No. 21-1272) and cubic phase In2O3 (JCPDS card No. 06-0416), respectively. After Au nanoparticles were modified on the surface of the TiO2/In2O3 nanocomposites, the peaks of Au appeared at 38.18°, 44.39°, 64.58°, and 77.55° assigned to the (111), (200), (220), and (311) planes of face-centered cubic Au (JCPDS card No. 04-0784), respectively. In addition, the diffraction peaks of the Au/TiO2/In2O3 nanocomposites did not shift after decorating the Au nanoparticles, which indicates that the Au atoms were not doped into the crystal structure. In the nanocrystals, inherent strain exists due to size limitations, and this important elastic property can affect the optical and electrical properties of materials [44]. XRD analysis of the nanocrystals can confirm the crystallinity of the sample, which exhibits different peaks associated with different reflection planes. The Williamson–Hall (W–H) method is a suitable method to study various elastic properties including strain as well as calculate the average size. In the XRD data, the broadening ( β T ) of the peaks was due to the combined effect of the crystallite size ( β D ) and micro strain ( β ε ).
β T = β D + β ε
From the Scherer equation,
β D = K λ D cos θ
where β D is the full width at half maxima (FWHM) in radians, K = 0.9 is the shape factor, λ = 0.15406 nm is the wavelength of X-ray source, D is the crystallite size, and θ is the peak position in radians. The XRD peak broadening due to micro strain is given as:
β ε = 4 ε tan θ
where β ε is broadening due to strain, ε is the strain, and θ is the peak position in radians. Putting Equations (3) and (4) in Equation (2):
β T =   K λ D cos θ   + 4 ε tan θ
Therefore, Equation (5) can be written as:
β T cos θ =   ε 4 sin θ + K λ D
Equation (6) is an equation of a straight line as the uniform deformation model (UDM) equation, which considers the isotropic nature of the crystals. Figure 2b shows the UDM plots for the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites. The slope of the line provides the value of the strain, while the intercept provides the average particle size of the crystal. According to the UDM, the average particle sizes of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites are about 23 nm, 53 nm, 29 nm, and 29 nm, respectively. The slope of the fitted lines in the UDM diagram is positive, indicating that the lattice expands to generate an intrinsic strain in the nanocrystals.
Figure 2. (a) X-ray diffraction (XRD) spectra of the hollow In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites; (b) Uniform deformation model (UDM) plots for the In2O3, TiO2, TiO2/In2O3 and Au/TiO2/In2O3 nanocomposites.
Figure 2. (a) X-ray diffraction (XRD) spectra of the hollow In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites; (b) Uniform deformation model (UDM) plots for the In2O3, TiO2, TiO2/In2O3 and Au/TiO2/In2O3 nanocomposites.
Materials 16 04010 g002
Figure 3 shows the SEM images of the carbon nanospheres, hollow In2O3 nanospheres, and TiO2/In2O3 and Au/TiO2/In2O3 nanocomposites. As shown in Figure 3a, the particle size of the carbon nanospheres was relatively uniform, and the size was about 100–200 nm. Figure 3b shows the SEM image of the prepared In2O3 nanospheres. As can be seen, the In2O3 sample was composed of a large number of spherical nanoparticles, and the diameter of the In2O3 nanospheres was about 150–250 nm. The hollow structure could be observed from the broken area of some nanoparticles. As can be seen from Figure 3c, after forming the nanocomposite material with TiO2, the surface of the nanospheres became obviously rough, and was observed that the TiO2 nanoparticles were uniformly covered on the surface of the In2O3 nanospheres. The particle size was about 150–250 nm. Figure 3d shows the SEM image of the Au/TiO2/In2O3 nanocomposites, and as can be seen, the hollow structure of the TiO2/In2O3 nanocomposites was still relatively intact, and the size of the TiO2/In2O3 nanospheres was not significantly changed. Due to the limitation of magnification, the existence of Au nanoparticles could not be observed. Therefore, the information of the Au nanoparticles and hollow structure of the Au/TiO2/In2O3 nanocomposites can be further observed by high resolution TEM.
The hollow structure of the Au modified TiO2/In2O3 nanocomposites could be clearly observed in the TEM images, as shown in Figure 4. As it can be seen in Figure 4a,b, the particle size of the Au/TiO2/In2O3 nanocomposites was 150–250 nm, and it could be observed that the Au nanoparticles were evenly dispersed on the surface of the nanospheres, and the size was about 10–15 nm. From Figure 4c, the wall thicknesses of the In2O3 shell and TiO2 shell were about 10 nm and 15 nm, respectively. The lattice fringes could be observed with an interplanar distance of 0.294 nm and 0.353 nm, which corresponded to the In2O3 (111) and TiO2 (101) crystal planes, respectively. Figure 4d shows the typical HRTEM images at the interface of the Au and TiO2 nanoparticles. The lattice fringes of Au and TiO2 interlace at the interface, which may be the effect of the overlapping of different lattice fringes of Au nanoparticles and TiO2 nanoparticles at the interface. The interplanar distance of Au was calculated to be 0.234 nm, corresponding to the Au (111) crystal planes. Figure 4j shows the energy dispersive spectrometer (EDS) mapping of the Au modified TiO2/In2O3 nanocomposites. It can be seen from the energy spectrum that the sample was composed of O, In, Ti, and Au elements, which indicates that the sample is a ternary composite material. The Au element was uniformly distributed across the whole sample. The weight percentages of the Ti, In, O, and Au elements were tested as 24.92 wt%, 48.10 wt%, 25.21 wt%, and 1.77 wt%, respectively, which was close to the theoretical additions (Ti: 24.75 wt%, In: 48.54 wt%, O: 25.00 wt%, and Au: 1.71 wt%).
XPS analysis was used to characterize the element composition and valence state information of the Au/TiO2/In2O3 nanocomposites, as shown in Figure 5. The characteristic peak of carbon can be used as a reference (C 1s 284.6 eV) for calibration [45]. As can be seen in Figure 5a, the characteristic peaks of elements In, Ti, Au, and O were observed in the full survey scan spectrum of the sample. The weight percentages of Ti 2p, In 3d, O 1s, and Au 4f elements were calculated as 23.41 wt%, 47.57 wt%, 27.20 wt%, and 1.82 wt%, respectively, which is basically consistent with the EDS results. Figure 5b presents the high-resolution spectrum of Ti 2p. The characteristic peaks at 458.5 eV and 464.2 eV were attributed to Ti 2p3/2 and Ti 2p1/2, respectively, indicating that the valence state of the Ti element is Ti4+ [46]. Figure 5c shows the XPS spectrum of In 3d. It can be seen that there are two characteristic peaks at the binding energy of 444.0 eV and 451.6 eV, which corresponded to the In 3d5/2 and In 3d3/2 electron orbitals, respectively [47]. The results showed that the In element exists in the form of In3+. Figure 5d shows the XPS spectrum of Au 4f. There are two characteristic peaks at the binding energy of about 83.1 eV and 86.7 eV, which corresponded to the electron orbitals of Au 4f7/2 and Au 4f5/2, respectively [48]. In addition, the O 1s spectrum of the TiO2/In2O3 and Au/TiO2/In2O3 samples were analyzed as shown in Figure 5e–f. The O1s spectrum was fitted and analyzed by the Gaussian fitting method, and three characteristic peaks were obtained, which corresponded to the lattice oxygen (OL), oxygen vacancies (Ov), and chemisorbed oxygen (OC) [49,50,51], respectively. As shown in Figure 5f, the peaks of OL, OV, and OC were located at about 529.32 eV, 530.10 eV, and 531.63 eV, respectively. The content of different oxygen ions before and after Au modification was analyzed, and the results are shown in Table 1. Compared with the TiO2/In2O3 samples, the oxygen vacancy content ratio of the Au/TiO2/In2O3 increased from 20.5% to 36.0%, and the chemisorbed oxygen content ratio increased from 11.5% to 23.7%. Chemisorbed oxygen will participate in the subsequent gas sensing response, therefore, the increase in chemisorbed oxygen content on the sample surface is conducive to the improvement in sensing performance. When TiO2 is irradiated by UV light, the electrons in the valence band are excited to the conduction band, where electrons and holes migrate to the surface of TiO2, forming electron–hole pairs on the surface; electrons react with Ti4+, and holes react with surface bridging oxygen ions to form Ti3+ and oxygen vacancy, respectively. However, the loading of Au nanoparticles on the surface of TiO2 can further promote the generation of electron–hole pairs, thus promoting the formation of oxygen vacancy. Therefore, the relative content of oxygen vacancy increases in the O 1 s spectra of the Au/TiO2/In2O3 nanocomposites.
In order to further characterize the porous structures of the Au/TiO2/In2O3 nanocomposites, nitrogen adsorption–desorption tests were conducted, and the results are shown in Figure 6. The curves show typical IV-type isotherm characteristics with H3-type hysteresis loops, which indicates that the sample had a mesoporous structure [52]. According to the pore size distribution curves, the pore size was mainly distributed around 7.5 nm. The specific surface area of the sample was 49.73 m2/g. Mesoporous structures can promote the diffusion, adsorption and desorption of target gas molecules, and it also provides more chemical active sites for the reaction of gas molecules with adsorbed oxygen ions on the surface, which plays an important role in improving the gas sensing properties.

3.2. Electrical Characterization and Gas-Sensing Property

Figure 7a shows the I–V polarization curves of sensors based on the TiO2/In2O3 nanocomposites and the Au/TiO2/In2O3 nanocomposites under the UV activation. At room temperature, the I–V characteristic curves of these two sensors were linear without UV excitation, which indicates the contact type between the materials and the gold interdigital electrodes. Under UV activation, the I–V polarization curves of the two sensors showed a slight nonlinear relationship between the applied voltage and the measured current. Due to the presence of nanoheterojunctions in the samples, photogenerated electrons were generated and transferred under UV excitation [53]. This process increases the energy barrier at the nanoheterojunctions, which illustrates that UV activation enhances the nanoheterojunction properties. Figure 7b shows the photo–response curves of hollow In2O3 and TiO2 nanospheres at room temperature. When the UV-LED was turned on, the resistance of the TiO2 nanosphere-based sensor decreased rapidly, and after the UV-LED was turned off, the resistance quickly returned to the initial resistance, and the resistance change rate was about 52%. In contrast, the respond and recovery speeds of the In2O3 nanosphere-based sensor were slow for the UV-LED. The photo–response curves of the sensor based on TiO2/In2O3 and Au/TiO2/In2O3 nanocomposites to UV activation at room temperature is shown in Figure 7c. The baseline resistance of the TiO2/In2O3 and Au/TiO2/In2O3 nanocomposite-based sensors was about 7 MΩ and 33 MΩ, respectively, which was much lower compared to the baseline resistance of the TiO2 nanospheres (~630 MΩ). The TiO2/In2O3 and Au/TiO2/In2O3 nanocomposite-based sensor also had a fast response to UV-LED, and the response signal was very stable. When UV light irradiates to the surface of the sensing material, a large number of photogenerated electron–hole pairs will be generated, and for the Fermi level to be in equilibrium, more electrons tend to be transferred from the TiO2 to the Au nanoparticle side, so the Au nanoparticles on the surface act as electron acceptors, which promotes the effective separation of electrons and holes at the interface, thus promoting the generation of more photogenerated charge carriers. In this process, the resistance of the sensor changes greatly. Therefore, Au modification improves the UV activation of the Au/TiO2/In2O3 nanocomposites.
Figure 8a shows the response curve of the sensor based on pure In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposites to 1 ppm formaldehyde under UV activation at room temperature. When formaldehyde gas is introduced into the test environment, the response signals of the In2O3 and TiO2-based sensors were significantly less than those of the nanocomposites, while the response curves of the TiO2/In2O3 and Au/TiO2/In2O3-based sensors increased rapidly and reached a stable state. The response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3 nanocomposite-based sensors were about 1.6, 2.1, 3.8, and 5.6, respectively. It can be observed that the response/recovery time of a pure In2O3-based sensor was the longest (136/233 s). In contrast, the response/recovery times of both the TiO2/In2O3 and Au/TiO2/In2O3-based sensors were significantly improved under the same test conditions, which were about 28/50 s and 18/42 s, respectively. This indicates that the addition of Au nanoparticles can effectively improve the adsorption and desorption rate of gas molecules on the surface of the sensing materials. Figure 8b shows the repeatability of response of the In2O3, TiO2, TiO2/In2O3 and Au/TiO2/In2O3-based sensors to 1 ppm formaldehyde for five consecutive cycles. It can be seen that in the recovery process, the resistance of the sensor can be completely restored to the initial state. The response value deviation of the Au/TiO2/In2O3-based sensor was less than 2%, which indicates that the Au/TiO2/In2O3-based sensor had good signal repeatability and stability. In can be seen from Figure 8c that the response curves of these four sensors to formaldehyde had concentrations ranging from 30 ppb to 10 ppm at room temperature. As can be seen, with the increase in the formaldehyde concentration, the response of these four sensors also increased correspondingly. The detection limit of the Au/TiO2/In2O3-based sensor was 60 ppb. Figure 8d shows the relationship between the response of the sensors and the formaldehyde concentrations. As can be seen, the response of the Au/TiO2/In2O3-based sensor was much higher than the other three sensors. These four sensors showed a good linear relationship between the response and formaldehyde concentrations. The slope of the calibration plot of the Au/TiO2/In2O3-based sensor was 2.64, which was significantly higher than the other sensors.
Figure 9a shows the response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm of various gases under UV activation at room temperature. We compared the response value of the Au/TiO2/In2O3-based sensor to 1 ppm formaldehyde gas with the response of 1 ppm of acetone (S = 3.1), ammonia (S = 2.6), methanol (S = 3.7), ethanol (S = 3.1), toluene (S = 2.7), and benzene (S = 2.4); the anti-interference ability (Selectivity coefficient: S = SHCHO/Sinterference) of the Au/TiO2/In2O3-based sensor was improved through a comparison with these common interference gases. This indicates that the selectivity of Au/TiO2/In2O3-based sensors is improved by Au modification. In order to research the effect of ambient humidity on the gas sensing performance of the sensors, a different humidity is introduced as a comparison condition during the test, and the results are shown in Figure 9b. As can be seen, the response values of these four sensors showed a gradual decline trend with the increase in humidity. When the humidity exceeded 75%, the pure In2O3 and TiO2 sensors had no response signals. When the humidity was 85%, the response value of the Au/TiO2/In2O3-based sensor maintained 55% of the original value. This indicates that the moisture resistance of the sensor is improved after the modified Au nanoparticles. Long-term stability is considered as one of the important indices to measure the performance of sensors. Figure 9c shows the stability of the Au/TiO2/In2O3-based sensor to 1 ppm formaldehyde at room temperature for 90 days. It can be observed that the response curve of the sensor had not changed significantly. It was calculated that the variation range of the response value was less than 3.5%. This demonstrates that the Au/TiO2/In2O3-based sensor has great long-term stability for formaldehyde detection at room temperature. Table 2 summarizes the comparison of the gas sensing properties of the Au/TiO2/In2O3 nanocomposite-based sensor and other sensing materials reported in the literature to formaldehyde [40,43,54,55,56,57]. It can be seen that the Au modified TiO2/In2O3-based sensor had a low detection limit, and its response value, response/recovery time, selectivity, and stability were excellent. These results indicate that the Au/TiO2/In2O3 sample has good application prospects as sensing materials for the detection of formaldehyde.
Figure 10 shows the sensing mechanism of the Au modified TiO2/In2O3 nanocomposites to formaldehyde under UV activation at room temperature. When the sensor is in the air, the Au/TiO2/In2O3 is activated to produce photogenerated electron–hole pairs under the excitation of UV light. O2, as a strong oxidizing gas, can be quickly adsorbed on the surface of the material under UV activation and react with photogenerated electrons to form chemisorbed oxygen ions (O2, O2−, O). In this process, the resistance of the sensor increases. When the operating temperature is less than 100 °C, the chemisorbed oxygen ion type is O2. The photogenerated holes are then combined with OH to form the neutral OH∙ with higher oxidation properties. The equation for the reaction is as follows:
hυ → h+ + e
O2 + e → O2 (hυ)
h+ + OH → OH∙
When the sensor is in contact with formaldehyde gas, the formaldehyde gas molecules will react with the pre-absorbed chemisorbed oxygen ions O2, and the released electrons in the reaction will return to the conduction band of the oxide; during this process, the resistance of the sensor decreases [58]. The chemical reactions on the surface of the sensing materials during this process were analyzed by in situ DRIFTS spectra detection.
In situ DRIFTS spectra detection technology plays a very important role in the real-time detection of the gas adsorption and reaction process on the surface of sensors, which contributes to speculating on the reaction mechanism. Figure 11a–c shows the in situ DRIFTS spectra of the TiO2, In2O3, and Au/TiO2/In2O3 nanocomposites exposed to air and formaldehyde at room temperature under UV activation. As can be seen, the vibration absorption peaks of formaldehyde molecules can be observed at about 1060 cm−1 and 1150 cm−1. It indicates that some formaldehyde is adsorbed on the surface sensing materials in a molecular state at room temperature [59]. The dissociation of water from the air on a TiO2 surface forms two distinctive hydroxyl groups: one OH- group bridges two Ti4+ (3741 cm−1) and the other forms a terminal Ti4+–OH- group (3670 cm−1), which correspond to the observed bands at 3650 and 3742 cm−1, respectively [60,61]. From Figure 11c, the band at 2936 cm−1 corresponded to the characteristic absorption peak of dioxymethylene (H2COO), and the intensity increased significantly with the reaction, which indicates that the formaldehyde molecules adsorbed are first oxidized to dioxymethylene, so one possible involved redox reaction can be proposed as:
HCHO (ads) + O2 (ads) → H2COO (ads) + e
Compared with the absorption peaks of TiO2 and In2O3, the relative strength of the absorption peaks of the nanocomposites was higher, which indicates that the Au/TiO2/In2O3 nanoheterojunctions greatly promoted the chemical reaction of the oxidation of formaldehyde to dioxymethylene. The bands at 2835 cm−1, 1591 cm−1, 1460 cm−1, and 1320 cm−1 were the absorption peaks of formate (HCOO), and their intensities increased with the extension of reaction time. Among them, 2835 cm−1 is the C–H stretching vibration absorption peak of formate; 1591 cm−1 is the asymmetric stretching vibration absorption peak of COO; 1320 cm−1 is the symmetric stretching vibration absorption peak of COO; 1460 cm−1 is the C–H bending vibration absorption peak of formate [62]. The presence of these peaks indicates that dioxymethylene is further oxidized by chemisorbed oxygen ions to form formate, as shown in the equation as follows:
H2COO (ads) + O2 → HCOO (ads) + OH (ads)
The bands at 2328 cm−1 and 2358 cm−1 are the characteristic absorption peaks of CO2, and the intensity of the peaks is relatively strong, indicating that the generated formate was further oxidized, as shown in in equation as follow:
HCOO (ads) + OH (ads) →CO2 + H2O + 2e
In situ DRIFTS spectra were used to analyze the changes of adsorbates on the surface of sensing materials during the sensing process. It was concluded that formaldehyde was first oxidized to H2COO, then oxidized to HCOO, and finally to CO2. The reaction process was more intense on the surface of the Au/TiO2/In2O3 nanocomposites, and UV light promoted the oxidation of formaldehyde to CO2. Figure 11d shows the in situ DRIFTS spectra of the Au/TiO2/In2O3 nanocomposites without UV activation. The bands at 1062 cm−1 and 1142 cm−1 were due to the vibration absorption peaks of the formaldehyde molecules. This indicates that formaldehyde molecules can be adsorbed on the surface of sensing materials without UV light. The presence of bands at 2836 cm−1, 1540 cm−1, and 1440 cm−1 indicates that a small amount of formaldehyde was oxidized to H2COO and HCOO, which also illustrates that a large number of formaldehyde molecules adsorbed on the surface of the Au/TiO2/In2O3 nanocomposites without significant dissociation. No characteristic absorption peak of CO2 was observed in the spectra, which indicates that the HCOO on the surface cannot be further oxidized to CO2 without UV activation. The results show that UV activation can significantly enhance the catalytic conversion of formaldehyde gas.
There are two reasons for the obvious improvement in the gas sensing of Au modified TiO2/In2O3 nanocomposites. On one hand, due to the different positions of the Fermi levels of In2O3 and TiO2, these two oxides contact and form a nanoheterojunction at the interface, so electrons will transfer from the conduction band of In2O3 to TiO2 through the interface and accumulate [63]. This can effectively separate the photogenerated electron–hole pair at the interface and increase the lifetime of the photogenerated electron–hole pair. With the increase in the electron content in the TiO2 nanoparticles, more formaldehyde molecules can participate in the REDOX reaction, which further increases the range of conductivity of the sensing materials, which improves the gas sensing properties of the Au/TiO2/In2O3 nanocomposites [64]. In addition, the relatively thin shell thickness also facilitates UV light to penetrate into the interior of the hollow materials. On the other hand, the modification of noble metals also plays an important role in improving the gas sensing performance of Au/TiO2/In2O3. First, due to the “spillover effect” of Au, Au nanoparticles modified on TiO2/In2O3 nanospheres can be used as a chemical sensitizer [65]. The catalytic capacity of noble metal nanoparticles for oxygen decomposition is much higher than that of In2O3 and TiO2, and the reactive oxygen species obtained after catalytic decomposition will overflow to the surface of the sensing material, which increases the width of the electron depletion layer of the sensing material and further increases the resistance of the sensing material in the air [66]. In addition, noble metal modification can significantly increase the proportion of active oxygen in Au/TiO2/In2O3, which promotes the reaction between formaldehyde and O2 and improves the sensing of the sensor, which is consistent with the results of XPS.

4. Conclusions

Au modified TiO2/In2O3 hollow nanospheres were prepared via a facile hydrolysis method with the assistance of a template. The results of the TEM, EDS, and XPS showed that Au nanoparticles were successfully modified on the surface of the TiO2/In2O3 nanocomposites, and the Au/TiO2/In2O3 nanocomposites maintained a double-layer hollow nanosphere structure. The size of the Au/TiO2/In2O3 nanospheres was about 150–250 nm. The thickness of the outer TiO2 shell and inner In2O3 wall was about 15 nm and 10 nm, respectively. The size of the Au nanoparticles was about 10–15 nm. Compared with the pure In2O3 and TiO2-based sensor, the Au modified TiO2/In2O3 nanocomposite-based sensor had better gas sensing properties to formaldehyde. The response of the Au/TiO2/In2O3 nanocomposite-based sensor to 1 ppm formaldehyde was about 5.6, and the detection limit was 60 ppb. The response and recovery times of the Au/TiO2/In2O3 nanocomposite-based sensor were 18 s and 42 s, respectively, which was faster than that of the pure In2O3, TiO2, and TiO2/In2O3 nanocomposites. The results of the in situ DRIFTS show that the nanoheterojunctions promoted the catalytic conversion of formaldehyde on the surface of the Au/TiO2/In2O3 nanocomposites under UV activation. The improvement in the gas sensing properties of the Au/TiO2/In2O3 nanocomposites is mainly due to the electronic and chemical sensitization of the Au nanoparticles.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16114010/s1; Table S1. Composition of the aqueous formaldehyde solutions used for the preparation of the formaldehyde vapors; Figure S1. Response curves of the Au/TiO2/In2O3 nanocomposites-based sensor under UV activation to different concentrations of water vapor at room temperature.

Author Contributions

Conceptualization, B.H., X.L., J.Z., Q.F. and S.Z.; Methodology, X.L., B.H., S.Z. and Z.J.; Data collection and analysis, S.Z., Z.J., J.Q. and J.C.; Validation, X.L., B.H., J.Z., Q.F. and S.Z.; Investigation, J.Q., J.C., B.H., X.L., J.Z. and S.Z.; Resources, X.L., B.H. and S.Z.; Writing—original draft preparation, S.Z., B.H. and X.L.; Writing—review and editing, X.L. and S.Z.; Supervision, X.L., J.Z. and B.H.; Funding acquisition, X.L. and B.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors highly appreciate the financial support from the National Key R&D Program of China (No. 2021YFB3201302), the National Natural Science Foundation of China (Nos. 61971085, 62201117, 62111530055), the Fundamental Research Funds for the Central Universities (No. DUT19RC(3)054) for the financial support, and the Opening Project of Key Laboratory of Microelectronic Devices & Integrated Technology, Institute of Microelectronics, Chinese Academy of Sciences, Natural Science Foundation of Shandong Province (No. ZR2021MF081).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tie, Y.; Ma, S.Y.; Yang, G.J.; Chen, Q.; Pei, S.T.; Ma, L.; Wang, W.Q.; Zhu, K.M.; Zhang, Q.X.; Almamoun, O. Improved formaldehyde sensor of Zn2SnO4/SnO2 microcubes by compositional evolution and Y2O3 decoration. Ceram. Int. 2019, 45, 5384–5391. [Google Scholar] [CrossRef]
  2. Xu, C.B.; Yang, W.S.; Guo, Q.; Dai, D.X.; Minton, T.K.; Yang, X.M. Photoinduced decomposition of formaldehyde on a TiO2(110) surface, assisted by bridge-bonded oxygen atoms. J. Phys. Chem. Lett. 2013, 4, 2668–2673. [Google Scholar] [CrossRef]
  3. Kim, H.R.; Haensch, A.; Kim, I.D.; Barsan, N.; Weimar, U.; Lee, J.H. The role of NiO doping in reducing the impact of humidity on the performance of SnO2-based gas sensors: Synthesis strategies, and phenomenological and spectroscopic studies. Adv. Funct. Mater. 2011, 21, 4456–4463. [Google Scholar] [CrossRef]
  4. Ma, Z.Z.; Yang, K.; Xiao, C.L.; Jia, L.C. C-doped LaFeO3 porous nanostructures for highly selective detection of formaldehyde. Sens. Actuators B Chem. 2021, 347, 130550. [Google Scholar] [CrossRef]
  5. Lou, C.M.; Huang, Q.X.; Li, Z.S.; Lei, G.L.; Liu, X.H.; Zhang, J. Fe2O3-sensitized SnO2 nanosheets via atomic layer deposition for sensitive formaldehyde detection. Sens. Actuators B Chem. 2021, 345, 130429. [Google Scholar] [CrossRef]
  6. Yu, H.M.; Li, J.Z.; Tian, Y.W.; Li, Z.Y. Environmentally friendly recycling of SnO2/Sn3O4 from tin anode slime for application in formaldehyde sensing material by Ag/Ag2O modification. J. Alloys Compd. 2018, 765, 624–634. [Google Scholar] [CrossRef]
  7. Zhou, P.; Zhu, X.; Yu, J.; Xiao, W. Effects of adsorbed F, OH, and Cl ions on formaldehyde adsorption performance and mechanism of anatase TiO2 nanosheets with exposed {001} facets. ACS Appl. Mater. Inter. 2013, 5, 8165–8172. [Google Scholar] [CrossRef]
  8. Wang, T.S.; Jiang, B.; Yu, Q.; Kou, X.Y.; Sun, P.; Liu, F.M.; Lu, H.Y.; Yan, X.; Lu, G.Y. Realizing the control of electronic energy level structure and gas-sensing selectivity over heteroatom-doped In2O3 spheres with an inverse opal microstructure. ACS Appl. Mater. Inter. 2019, 11, 9600–9611. [Google Scholar] [CrossRef]
  9. Gu, D.; Wang, X.Y.; Liu, W.; Li, X.G.; Lin, S.W.; Wang, J.; Rumyantseva, M.N.; Gaskov, A.M.; Akbar, S.A. Visible-light activated room temperature NO2 sensing of SnS2 nanosheets based chemiresistive sensors. Sens. Actuators B Chem. 2020, 305, 127455. [Google Scholar] [CrossRef]
  10. Rong, X.R.; Chen, D.L.; Qu, G.P.; Li, T.; Zhang, R.; Sun, J. Effects of graphene on the microstructures of SnO2@rGO nanocomposites and their formaldehyde-sensing performance. Sens. Actuators B Chem. 2018, 269, 223–237. [Google Scholar] [CrossRef]
  11. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B Chem. 2014, 204, 250–272. [Google Scholar] [CrossRef]
  12. Walker, J.; Karnati, P.; Akbar, S.A.; Morris, P.A. Selectivity mechanisms in resistive-type metal oxide heterostructural gas sensors. Sens. Actuators B Chem. 2022, 355, 131242. [Google Scholar] [CrossRef]
  13. Yu, W.L.; Zeng, W.; Li, Y.Q. A nest-like TiO2 nanostructures for excellent performance ethanol sensor. Mater. Lett. 2019, 248, 82–85. [Google Scholar] [CrossRef]
  14. Lei, T.; Zhang, S.P.; Li, D.; Zhang, W.; Huang, S.; Xie, C.S. The influence of Au and Pt electrodes on the stability of TiO2 under UV light activation for sensing formaldehyde in moisture circumstances. Sens. Actuators B Chem. 2014, 199, 15–21. [Google Scholar] [CrossRef]
  15. Zhang, S.P.; Lei, T.; Li, D.; Zhang, G.Z.; Xie, C.S. UV light activation of TiO2 for sensing formaldehyde: How to be sensitive, recovering fast, and humidity less sensitive. Sens. Actuators B Chem. 2014, 202, 964–970. [Google Scholar] [CrossRef]
  16. Li, X.G.; Li, X.X.; Wang, J.; Lin, S.W. Highly sensitive and selective room-temperature formaldehyde sensors using hollow TiO2 microspheres. Sens. Actuators B Chem. 2015, 219, 158–163. [Google Scholar] [CrossRef]
  17. Mintcheva, N.; Srinivasan, P.; Rayappan, J.B.B.; Kuchmizhak, A.A.; Gurbatov, S.; Kulinich, S.A. Room-temperature gas sensing of Laser-modified anatase TiO2 secorated with Au nanoparticles. Appl. Surf. Sci. 2020, 507, 145169. [Google Scholar] [CrossRef]
  18. Kumar, G.; Li, X.J.; Du, Y.; Geng, Y.F.; Hong, X.M. UV-light enhanced high sensitive hydrogen (H2) sensor based on spherical Au nanoparticles on ZnO nanostructured thin films. J. Alloys Compd. 2019, 798, 467–477. [Google Scholar] [CrossRef]
  19. Wang, Y.H.; Lai, X.H.; Liu, B.Y.; Chen, Y.B.; Lu, Y.Z.; Wang, F.P.; Zhang, L.W. UV-induced desorption of oxygen at the TiO2 surface for highly sensitive room temperature O2 sensing. J. Alloys Compd. 2019, 793, 583–589. [Google Scholar] [CrossRef]
  20. Wang, Z.; Wang, K.; Peng, X.; Geng, Q.; Chen, X.; Dai, W.; Fu, X.; Wang, X. Comparative study of ultraviolet light and visible light on the photo-assisted conductivity and gas sensing property of TiO2. Sens. Actuators B Chem. 2017, 248, 724–732. [Google Scholar] [CrossRef]
  21. Seekaew, Y.; Wisitsoraat, A.; Phokharatkul, D.; Wongchoosuk, C. Room temperature toluene gas sensor based on TiO2 nanoparticles decorated 3D graphene-carbon nanotube nanostructures. Sens. Actuators B Chem. 2019, 279, 69–78. [Google Scholar] [CrossRef]
  22. Zhang, J.N.; Chen, C.J.; Lu, H.B.; Leng, D.Y.; Li, G.; Liu, Y.M.; Liang, Q.F.; Gao, J.Z.; Wang, C.L.; Zhu, B.P. Construction of ana-tase@rutile core@shell TiO2 nanosheets with controllable shell layer thicknesses for enhanced ethanol sensing. Sens. Actuators B Chem. 2020, 325, 128815. [Google Scholar] [CrossRef]
  23. Zhang, S.; Zhao, L.J.; Huang, B.Y.; Li, X.G. UV-activated formaldehyde sensing properties of hollow TiO2@SnO2 heterojunctions at room temperature. Sens. Actuators B Chem. 2020, 319, 128264. [Google Scholar] [CrossRef]
  24. Zhao, S.K.; Shen, Y.B.; Zhou, P.F.; Hao, F.L.; Xu, X.Y.; Gao, S.L.; Wei, D.Z.; Ao, Y.X.; Shen, Y.S. Enhanced NO2 sensing performance of ZnO nanowires functionalized with ultra-fine In2O3 nanoparticles. Sens. Actuators B Chem. 2020, 308, 127729. [Google Scholar] [CrossRef]
  25. Zhang, S.; Sun, S.P.; Huang, B.Y.; Wang, N.; Li, X.G. UV-enhanced formaldehyde sensor using hollow In2O3@TiO2 double-layer nanospheres at room temperature. ACS Appl. Mater. Inter. 2023, 15, 4329–4342. [Google Scholar] [CrossRef]
  26. Chen, K.Q.; Chen, S.J.; Pi, M.Y.; Zhang, D.K. SnO2 nanoparticles/TiO2 nanofibers heterostructures: In situ fabrication and enhanced gas sensing performance. Solid State Electronics 2019, 157, 42–47. [Google Scholar] [CrossRef]
  27. Li, W.W.; Guo, J.H.; Cai, L.; Qi, W.Z.; Sun, Y.L.; Xu, J.L.; Sun, M.X.; Zhu, H.W.; Xiang, L.; Xie, D.; et al. UV light irradiation enhanced gas sensor selectivity of NO2 and SO2 using rGO functionalized with hollow SnO2 nanofibers. Sens. Actuators B Chem. 2019, 290, 443–452. [Google Scholar] [CrossRef]
  28. Nagabandi, J.; Madhukar, P.; Julakanti, S.; Ramana, R.M.V. Semi shield driven p-n heterostructures and their role in enhancing the room temperature ethanol gas sensing performance of NiO/SnO2 nanocomposites. Ceram. Int. 2019, 45, 15134–15142. [Google Scholar]
  29. Zhu, L.Y.; Yuan, K.P.; Yang, J.G.; Ma, H.P.; Wang, T.; Ji, X.M.; Feng, J.J.; Devi, A.; Lu, H.L. Fabrication of heterostructured p-CuO/n-SnO2 core-shell nanowires for enhanced sensitive and selective formaldehyde detection. Sens. Actuators B Chem. 2019, 290, 233–241. [Google Scholar] [CrossRef]
  30. Yang, X.H.; Fu, H.T.; Tian, Y.; Xie, Q.; Xiong, S.X.; Han, D.Z.; Zhang, H.; An, X.Z. Au decorated In2O3 hollow nanospheres: A novel sensing material toward amine. Sens. Actuators B Chem. 2019, 296, 126696. [Google Scholar] [CrossRef]
  31. Park, S.; Kim, S.; Sun, G.J.; Lee, C. Synthesis, structure and ethanol gas sensing properties of In2O3 nanorods decorated with Bi2O3 nanoparticles. ACS Appl. Mater. Inter. 2015, 7, 8138–8146. [Google Scholar] [CrossRef]
  32. Li, Z.J.; Yan, S.N.; Zhang, S.C.; Wang, J.Q.; Shen, W.Z.; Wang, Z.G.; Fu, Y.Q. Ultra-sensitive UV and H2S dual functional sensors based on porous In2O3 nanoparticles operated at room temperature. J. Alloys Compd. 2019, 770, 721–731. [Google Scholar] [CrossRef]
  33. Wang, C.Y.; Cimalla, V.; Kups, T. Integration of In2O3 nanoparticle based ozone sensors with GaInN/GaN light emitting diodes. Appl. Phys. Lett. 2007, 91, 103509. [Google Scholar] [CrossRef]
  34. Zhao, S.K.; Shen, Y.B.; Zhou, P.F.; Zhong, X.X.; Han, C.; Zhao, Q.; Wei, D.Z. Design of Au@WO3 core-shell structured nanospheres for ppb-level NO2 sensing. Sens. Actuators B Chem. 2019, 282, 917–926. [Google Scholar] [CrossRef]
  35. Ma, N.; Suematsu, K.; Yuasa, M.; Kida, T.; Shimanoe, K. Effect of water vapor on Pd-loaded SnO2 nanoparticles gas sensor. ACS Appl. Mater. Inter. 2015, 7, 5863–5869. [Google Scholar] [CrossRef]
  36. Saito, N.; Haneda, H.; Watanabe, K.; Shimanoe, K.; Sakaguchi, I. Highly sensitive isoprene gas sensor using Au-loaded pyramid-shaped ZnO particles. Sens. Actuators B Chem. 2020, 326, 128999. [Google Scholar] [CrossRef]
  37. Barbosa, M.S.; Suman, P.H.; Kim, J.J.; Tuller, H.L.; Orlandi, M.O. Investigation of electronic and chemical sensitization effects promoted by Pt and Pd nanoparticles on single-crystalline SnO nanobelt-based gas sensors. Sens. Actuators B Chem. 2019, 301, 127055. [Google Scholar] [CrossRef]
  38. Lupan, O.; Postica, V.; Pauportéc, T.; Viana, B.; Terasa, M.I.; Adelung, R. Room temperature gas nanosensors based on individual and multiple networked Au-modified ZnO nanowires. Sens. Actuators B Chem. 2019, 299, 126977. [Google Scholar] [CrossRef]
  39. Kamble, C.; Panse, M.; Nimbalkar, A. Ag decorated WO3 sensor for the detection of sub-ppm level NO2 concentration in air. Mat. Sci. Semicon. Proc. 2019, 103, 104613. [Google Scholar] [CrossRef]
  40. Gu, F.B.; Di, M.Y.; Han, D.M.; Hong, S.; Wang, Z.H. Atomically dispersed Au on In2O3 nanosheets for highly sensitive and selective detection of formaldehyde. ACS Sens. 2020, 5, 2611–2619. [Google Scholar] [CrossRef]
  41. Chung, F.C.; Zhu, Z.; Luo, P.Y. Au@ZnO core-shell structure for gaseous formaldehyde sensing at room temperature. Sens. Actuators B Chem. 2014, 199, 314–319. [Google Scholar] [CrossRef]
  42. Yu, L.; Yu, X.Y.; Lou, X.W. The design and synthesis of hollow micro-/nanostructures: Present and future trends. Adv. Mater. 2018, 30, 1800. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.W.; Liu, J.Y.; Guo, H.; Zhou, X.; Wang, C.; Sun, P.; Hu, X.L.; Lu, G.Y. Au@In2O3 core-shell composites: Metal-semiconductor heterostructure for gas sensing applications. RSC Adv. 2015, 5, 545–551. [Google Scholar] [CrossRef]
  44. Nath, D.; Singh, F.; Das, R. X-ray diffraction analysis by Williamson-Hall, Halder-Wagner and size-strain plot methods of CdSe nanoparticles- a comparative study. Mater. Chem. Phys. 2020, 239, 122021. [Google Scholar] [CrossRef]
  45. Li, C.H.; Ming, T.; Wang, J.X.; Wang, J.F.; Yu, J.C.; Yu, S.H. Ultra-sonic aerosol spray-assisted preparation of TiO2/In2O3 Composite for visible-light-driven photocatalysis. J Catal. 2014, 310, 84–90. [Google Scholar] [CrossRef]
  46. Wang, H.M.; Yan, Y.; Chen, G. The effects of confinement on TiO2@SnO2@TiO2 hollow spheres for high reversible Lithium storage capacity. J. Alloys Compd. 2019, 778, 375–381. [Google Scholar] [CrossRef]
  47. Ramirez-Ortega, D.; Acevedo-Peña, P.; Tzompantzi, F.; Ar-royo, R.; González, F.; Gonzaález, I. Energetic states in SnO2-TiO2 structures and their impact on interfacial charge transfer process. J. Mater. Sci. 2017, 52, 260–275. [Google Scholar] [CrossRef]
  48. Zhang, J.; Liu, X.; Wang, L.; Yang, T.; Guo, X.; Wu, S.; Wang, S.; Zhang, S. Au-functionalized hematite hybrid nanospindles: General synthesis, gas sensing and catalytic properties. J. Phys. Chem. C 2011, 115, 5352–5357. [Google Scholar] [CrossRef]
  49. Zhang, S.; Zheng, Y.K.; Song, P.; Sun, J.; Wang, Q. Enhanced trimethylamine gas-sensing performance of CeO2 nanoparticles-decorated MoO3 nanorods. J. Mater. Sci: Mater. Electron 2022, 33, 3453–3464. [Google Scholar] [CrossRef]
  50. Zhang, D.Z.; Yang, Z.M.; Li, P.; Zhou, X.Y. Ozone gas sensing properties of metal-organic frameworks-derived In2O3 hollow microtubes decorated with ZnO nanoparticles. Sens. Actuators B Chem. 2019, 301, 127081. [Google Scholar] [CrossRef]
  51. Wang, X.J.; Wang, W.; Liu, Y.L. Enhanced acetone sensing performance of Au nanoparticles functionalized flower-like ZnO. Sens. Actuators B Chem. 2012, 168, 39–45. [Google Scholar] [CrossRef]
  52. Guo, W.W.; Zhao, B.Y.; Huang, L.L.; He, Y.Z. One-Step Synthesis of ZnWO4/ZnSnO3 Composite and the Enhanced Gas Sensing Performance to Formaldehyde. Mater. Lett. 2020, 277, 128327. [Google Scholar] [CrossRef]
  53. Li, Z.; Haidry, A.A.; Plecenik, T.; Vidis, M.; Grancic, B.; Roch, T.; Gregor, M.; Durina, P.; Yao, Z.J.; Plecenik, A. Influence of nanoscale TiO2 film thickness on gas sensing properties of capacitor-like Pt/TiO2/Pt sensing structure. Appl. Surf. Sci. 2020, 499, 143909. [Google Scholar] [CrossRef]
  54. Zhang, S.; Zhao, L.J.; Huang, B.Y.; Li, X.G. Enhanced sensing performance of Au-decorated TiO2 nanospheres with hollow structure for formaldehyde detection at room temperature. Sens. Actuators B Chem. 2022, 358, 131465. [Google Scholar] [CrossRef]
  55. Chung, F.C.; Wu, R.J.; Cheng, F.C. Fabrication of a Au@SnO2 core-shell structure for gaseous formaldehyde sensing at room temperature. Chem. Phys. Lett. 2014, 190, 1–7. [Google Scholar] [CrossRef]
  56. Huang, J.Y.; Liang, H.; Ye, J.X. Ultrasensitive formaldehyde gas sensor based on Au-loaded ZnO nanorod arrays at low temperature. Sens. Actuators B Chem. 2021, 346, 130568. [Google Scholar] [CrossRef]
  57. Liu, D.; Wan, J.W.; Wang, H. Mesoporous Au@ZnO flower-like nanostructure for enhanced formaldehyde sensing performance. Inorg. Chem. Commun. 2019, 102, 203–209. [Google Scholar] [CrossRef]
  58. Zhang, L.; Zhang, B.J.; Xue, P.; Li, J.W.; Zhang, Z.; Yang, Y.L.; Wang, S.; Lin, J.D.; Liao, H.G.; Wang, Y.; et al. The effect of pretreatment on the reactivity of Pd/Al2O3 in room temperature formaldehyde oxidation. ChemCatChem 2021, 13, 4133–4141. [Google Scholar] [CrossRef]
  59. Sun, S.; Ding, J.J.; Bao, J.; Gao, C.; Qi, Z.M.; Li, C.X. Photo-catalytic oxidation of gaseous formaldehyde on TiO2: An in situ DRIFTS study. Catal. Lett. 2010, 137, 239–246. [Google Scholar] [CrossRef]
  60. Zhu, L.; Wang, J.N.; Liu, J.W.; Xu, Z.C.; Nasir, M.S.; Chen, X.; Wang, Z.; Sun, S.Y.; Ma, Q.Y.; Liu, J.B.; et al. In situ enrichment amplification strategy enabling highly sensitive formaldehyde gas sensor. Sens. Actuators B Chem. 2022, 354, 131206. [Google Scholar] [CrossRef]
  61. Fan, G.J.; Nie, L.F.; Wang, H.; Zhang, L.; Chai, S.H.; Wang, A.Q.; Guan, J.; Han, N.; Chen, Y.F. Ce doped SnO/SnO2 heterojunc-tions for highly formaldehyde gas sensing at low temperature. Sens. Actuators B Chem. 2022, 373, 132640. [Google Scholar] [CrossRef]
  62. He, M.; Cao, Y.Q.; Ji, J.; Li, K.; Huang, H.B. Superior catalytic performance of Pd-loaded oxygen-vacancy-rich TiO2 for for-maldehyde oxidation at room temperature. J. Catal. 2021, 396, 122–135. [Google Scholar] [CrossRef]
  63. Liu, Y.N.; Xiao, S.; Du, K. Chemiresistive gas sensors based on hollow heterojunction: A review. Adv. Mater. Interfaces 2021, 8, 2002122. [Google Scholar] [CrossRef]
  64. Liu, W.; Gu, D.; Li, X.G. Ultrasensitive NO2 detection utilizing mesoporous ZnSe/ZnO heterojunction-based chemiresistive-type sensors. ACS Appl. Mater. Inter. 2019, 11, 29029–29040. [Google Scholar] [CrossRef]
  65. Kim, J.H.; Kim, J.Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Synergistic effects of SnO2 and Au nanoparticles decorated on WS2 nanosheets for flexible, room-temperature CO gas sensing. Sens. Actuators B Chem. 2021, 332, 129493. [Google Scholar] [CrossRef]
  66. Zhang, Y.Q.; Li, D.; Qin, L.G.; Liu, D.Y.; Liu, Y.Y.; Liu, F.M.; Song, H.W.; Wang, Y.; Lu, G.Y. Preparation of Au-loaded TiO2 pecan-kernel-like and its enhanced toluene sensing performance. Sens. Actuators B Chem. 2018, 255, 2240–2247. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram of the sensing measurement apparatus.
Figure 1. The schematic diagram of the sensing measurement apparatus.
Materials 16 04010 g001
Figure 3. Scanning electron microscope (SEM) images of (a) carbon nanospheres; (b) hollow In2O3 nanospheres; (c) TiO2/In2O3 nanocomposites, and (d) Au/TiO2/In2O3 nanocomposites.
Figure 3. Scanning electron microscope (SEM) images of (a) carbon nanospheres; (b) hollow In2O3 nanospheres; (c) TiO2/In2O3 nanocomposites, and (d) Au/TiO2/In2O3 nanocomposites.
Materials 16 04010 g003
Figure 4. (ae) Transmission electron microscope (TEM) and high resolution-TEM images of the Au/TiO2/In2O3 nanocomposites; (fj) Energy dispersive spectrometer (EDS) mapping of the Au/TiO2/In2O3 nanocomposites for the overlay, O element, Ti element, In element, and Au element.
Figure 4. (ae) Transmission electron microscope (TEM) and high resolution-TEM images of the Au/TiO2/In2O3 nanocomposites; (fj) Energy dispersive spectrometer (EDS) mapping of the Au/TiO2/In2O3 nanocomposites for the overlay, O element, Ti element, In element, and Au element.
Materials 16 04010 g004
Figure 5. X-ray photoelectron spectroscopy (XPS) spectra of: (a) full survey scan spectrum; (b) Ti 2p spectrum of Au/TiO2/In2O3; (c) In 3d spectrum of Au/TiO2/In2O3; (d) Au 4f spectrum of Au/TiO2/In2O3; (e) O 1s spectrum of TiO2/In2O3; (f) O 1s spectrum of Au/TiO2/In2O3.
Figure 5. X-ray photoelectron spectroscopy (XPS) spectra of: (a) full survey scan spectrum; (b) Ti 2p spectrum of Au/TiO2/In2O3; (c) In 3d spectrum of Au/TiO2/In2O3; (d) Au 4f spectrum of Au/TiO2/In2O3; (e) O 1s spectrum of TiO2/In2O3; (f) O 1s spectrum of Au/TiO2/In2O3.
Materials 16 04010 g005
Figure 6. N2 adsorption–desorption isotherms and the pore size distribution plots (inset) of the Au/TiO2/In2O3 nanocomposites.
Figure 6. N2 adsorption–desorption isotherms and the pore size distribution plots (inset) of the Au/TiO2/In2O3 nanocomposites.
Materials 16 04010 g006
Figure 7. (a) I–V polarization curves of the TiO2/In2O3 and Au/TiO2/In2O3 nanocomposite-based sensors at room temperature; (b) photo–response curves of In2O3 and TiO2 nanosphere-based sensors to UV-LED at room temperature; (c) photo–response curves of TiO2/In2O3 and Au/TiO2/In2O3-based sensors to UV-LED at room temperature.
Figure 7. (a) I–V polarization curves of the TiO2/In2O3 and Au/TiO2/In2O3 nanocomposite-based sensors at room temperature; (b) photo–response curves of In2O3 and TiO2 nanosphere-based sensors to UV-LED at room temperature; (c) photo–response curves of TiO2/In2O3 and Au/TiO2/In2O3-based sensors to UV-LED at room temperature.
Materials 16 04010 g007
Figure 8. (a) Response curves of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm HCHO under UV at room temperature; (b) repeatability of response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm HCHO; (c) response curves of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 0.03–10 ppm HCHO; (d) correlation of the response and HCHO concentrations of the sensors (where Ra represents the resistance of the sensors in the air and Rg represents the resistance of the sensor in targeted gas vapors).
Figure 8. (a) Response curves of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm HCHO under UV at room temperature; (b) repeatability of response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm HCHO; (c) response curves of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 0.03–10 ppm HCHO; (d) correlation of the response and HCHO concentrations of the sensors (where Ra represents the resistance of the sensors in the air and Rg represents the resistance of the sensor in targeted gas vapors).
Materials 16 04010 g008
Figure 9. (a) Response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm of various gases under UV activation at room temperature; (b) response of sensors to 1 ppm HCHO under various humidity; (c) the long-term stability of the Au/TiO2/In2O3 nanocomposite-based sensor for 90 days under UV activation at room temperature.
Figure 9. (a) Response of the In2O3, TiO2, TiO2/In2O3, and Au/TiO2/In2O3-based sensors to 1 ppm of various gases under UV activation at room temperature; (b) response of sensors to 1 ppm HCHO under various humidity; (c) the long-term stability of the Au/TiO2/In2O3 nanocomposite-based sensor for 90 days under UV activation at room temperature.
Materials 16 04010 g009
Figure 10. Schematic illustration of the sensing process of the Au/TiO2/In2O3 nanocomposite-based sensor under UV activation at room temperature.
Figure 10. Schematic illustration of the sensing process of the Au/TiO2/In2O3 nanocomposite-based sensor under UV activation at room temperature.
Materials 16 04010 g010
Figure 11. In situ DRIFTS spectra of the (a) TiO2, (b) In2O3, and (c) Au/TiO2/In2O3 nanocomposites exposed to air and HCHO at room temperature with UV activation; (d) in situ DRIFTS spectra of the Au/TiO2/In2O3 nanocomposites exposed to air and HCHO at room temperature without UV activation.
Figure 11. In situ DRIFTS spectra of the (a) TiO2, (b) In2O3, and (c) Au/TiO2/In2O3 nanocomposites exposed to air and HCHO at room temperature with UV activation; (d) in situ DRIFTS spectra of the Au/TiO2/In2O3 nanocomposites exposed to air and HCHO at room temperature without UV activation.
Materials 16 04010 g011
Table 1. Curve-fitting results of the high-resolution X-ray photoelectron spectroscopy (XPS) spectra for the O 1s region.
Table 1. Curve-fitting results of the high-resolution X-ray photoelectron spectroscopy (XPS) spectra for the O 1s region.
SampleOLOVOC
Eb (eV) 1ri (%) 2Eb (eV)ri (%)Eb (eV)ri (%)
TiO2/In2O3529.5268.0530.6320.5531.6811.5
Au/TiO2/In2O3529.3240.3530.1036.0531.6323.7
1 Eb (eV) represents the binding energy of peaks; 2 ri (%) represents the ratio Ai/∑Ai (Ai is the area of each peak).
Table 2. Comparison of the gas sensing properties of the sensors based on different materials to formaldehyde reported in the literature.
Table 2. Comparison of the gas sensing properties of the sensors based on different materials to formaldehyde reported in the literature.
Sensing MaterialsT (°C)ResponseConc. (ppm)Range (ppm)Res./Recov. (s)Ref.
Au/In2O310085.67500.00142–10028/198[40]
Au/In2O3200171005–1007/135[43]
Au/TiO2RT (UV) *8.550.1–1036/110[54]
Au/SnO2RT (UV)2.95020–5080/62[55]
Au/ZnO7068.81001–100216/106[56]
Au/ZnO22011.741010–150038/18[57]
Au/TiO2/In2O3RT (UV)5.610.03–1018/42This work
* RT (UV): Room temperature at ultraviolet activation.
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

Zhang, S.; Huang, B.; Jiang, Z.; Qian, J.; Cao, J.; Feng, Q.; Zhang, J.; Li, X. UV-Activated Au Modified TiO2/In2O3 Hollow Nanospheres for Formaldehyde Detection at Room Temperature. Materials 2023, 16, 4010. https://doi.org/10.3390/ma16114010

AMA Style

Zhang S, Huang B, Jiang Z, Qian J, Cao J, Feng Q, Zhang J, Li X. UV-Activated Au Modified TiO2/In2O3 Hollow Nanospheres for Formaldehyde Detection at Room Temperature. Materials. 2023; 16(11):4010. https://doi.org/10.3390/ma16114010

Chicago/Turabian Style

Zhang, Su, Baoyu Huang, Zenghao Jiang, Junfan Qian, Jiawei Cao, Qiuxia Feng, Jianwei Zhang, and Xiaogan Li. 2023. "UV-Activated Au Modified TiO2/In2O3 Hollow Nanospheres for Formaldehyde Detection at Room Temperature" Materials 16, no. 11: 4010. https://doi.org/10.3390/ma16114010

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