Next Article in Journal
TriNymAuth: Triple Pseudonym Authentication Scheme for VANETs Based on Cuckoo Filter and Paillier Homomorphic Encryption
Previous Article in Journal
Two-Step Approach for Occupancy Estimation in Intensive Care Units Based on Bayesian Optimization Techniques
Previous Article in Special Issue
Functional Microfiber Nonwoven Fabric with Copper Ion-Immobilized Polymer Brush for Detection and Adsorption of Acetone Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Imbedding Pd Nanoparticles into Porous In2O3 Structure for Enhanced Low-Concentration Methane Sensing

1
Key Laboratory for Liquid-Solid Structural Evolution and Processing of Materials Ministry of Education, Shandong University, Jinan 250061, China
2
China Aerospace Components Engineering Center, China Academy of Space Technology, Beijing 100094, China
*
Authors to whom correspondence should be addressed.
Sensors 2023, 23(3), 1163; https://doi.org/10.3390/s23031163
Submission received: 1 January 2023 / Revised: 17 January 2023 / Accepted: 18 January 2023 / Published: 19 January 2023
(This article belongs to the Special Issue Advanced Micro and Nano Technologies for Gas Sensing)

Abstract

:
Methane (CH4), as the main component of natural gas and coal mine gas, is widely used in daily life and industrial processes and its leakage always causes undesirable misadventures. Thus, the rapid detection of low concentration methane is quite necessary. However, due to its robust chemical stability resulting from the strong tetrahedral-symmetry structure, the methane molecules are usually chemically inert to the sensing layers in detectors, making the rapid and efficient alert a big challenge. In this work, palladium nanoparticles (Pd NPs) embedded indium oxide porous hollow tubes (In2O3 PHTs) were successfully synthesized using Pd@MIL-68 (In) MOFs as precursors. All In2O3-based samples derived from Pd@MIL-68 (In) MOFs inherited the morphology of the precursors and exhibited the feature of hexagonal hollow tubes with porous architecture. The gas-sensing performances to 5000 ppm CH4 were evaluated and it was found that Pd@In2O3-2 gave the best response (Ra/Rg = 23.2) at 370 °C, which was 15.5 times higher than that of pristine-In2O3 sensors. In addition, the sensing materials also showed superior selectivity against interfering gases and a rather short response/recovery time of 7 s/5 s. The enhancement in sensing performances of Pd@In2O3-2 could be attributed to the large surface area, rich porosity, abundant oxygen vacancies and the catalytic function of Pd NPs.

1. Introduction

As the main component of natural gas, liquefied gas, and coal mine gas, methane has gained large-scale application of in daily life and industrial processes [1,2]. It is noted that CH4 is inflammable and explosive at ambient temperature and pressure [3]. Once the concentration of CH4 reaches the explosive limit (4.9–15.4%) in a relatively airtight space, a severe explosion is likely to occur, resulting in serious casualties and property losses [4]. Furthermore, methane is also considered as a powerful greenhouse gas, of which the greenhouse effect reported by the Environmental Protection Agency (EPA) is about 20~25 times higher than that of CO2 [5,6]. What is worse, CH4 is a considerable threat to environment because it can remain for over 12 years in the atmosphere until complete decomposition [7]. Therefore, it is of great significance to develop cost-effective technique that can track methane in time so as to ensure the safety of the environment.
Methane is highly thermodynamic stable owing to its firm tetrahedral-symmetry structure with the bond energy of 431 kJ/mol [8]. Thus, it is distinctly an enormous challenge to detect methane efficiently. Gas sensors based on metal oxide semiconductors (MOSs), including SnO2 [9,10], ZnO [11,12], WO3 [2], VOx [13], NiO [14,15], and In2O3 [16], have been widely studied to detect methane because of their high response, wide linear range, low detection limit, excellent anti-interference capacity, facile operation, and low cost [17]. As an n-type semiconductor with a wide band gap of 3.5~3.7 eV, In2O3, with various nanostructures such as nanosheets, nanoparticles, and nanospheres, has attracted considerable attention due to its superior electrical properties, showing potential to fabricate methane sensors for application [18,19,20,21]. However, pristine In2O3 sensors are still suffering from some drawbacks, such as low sensitivity, long-time response and recovery, as well as poor selectivity to methane. Therefore, it is crucial to enhance their methane-sensing performance by changing structure of the materials.
MOSs derived from metal-organic frameworks (MOFs) have been extensively applied in catalysis [22,23], battery [24] and gas sensors [25]. Moreover, MOFs offer great opportunities for accommodating myriad guest materials, including noble metal nanoparticles, and still remain the internal MOF structure for further application [26]. Their large specific surface area, regular porous morphologies, and tunable microstructures are convenient for gas diffusion [27,28]. Li et al. fabricated a productive H2S gas sensor based on the bamboo-like CuO/In2O3 derived from Cu2+-impregnated CPP-3 (In) with a low detection limit (200 ppb) at a low operating temperature (70 °C) [29]. Ma et al. successfully synthesized a chlorine gas sensor using MIL-68 (In) as template and the improved sensitivity and selectivity were obtained at 160 °C, which can be attributed to abundant oxygen vacancies and active sites [30]. Furthermore, modification with noble metals, especially palladium (Pd), has been proved to be an extremely effective approach to enhance the sensing properties of MOSs towards methane. Wang et al. prepared Pd-modified ZnO nanosheets, exhibiting not only appreciable response but also a declined optimum working temperature when the loading amount of Pd was 1 at% [31]. In addition, a reliable CH4 sensor based on Pd-loaded SnO2 hollow spheres was manufactured by Yang et al. and demonstrated a response value of 4.88 to 250 ppm CH4 at a decreased temperature of 300 °C [32]. Based on these pioneer works, it is possibly a marvelous idea to combine the large surface area of In2O3 MOSs using metal-organic frameworks as template, with the catalytic effect of Pd NPs in order to improve the methane sensing performance. Nonetheless, there is still no relevant literature reported constructing methane gas sensors based on Pd@In2O3 derived from Pd@MIL-68 (In) precursors with core-shell structure.
In this work, Pd@In2O3 porous hollow tubes (PHTs) were successfully synthesized through a facile solvothermal method combined with subsequent calcination and reduction. Along with the construction of MIL-68 (In) prisms, different amounts of Pd NPs were injected to form the core-shell structure, ensuring intimate interaction between Pd NPs and MOFs. The sensing performance towards 5000 ppm CH4 of all prepared sensors were investigated and compared. We found that Pd@In2O3-2 sensor could be a promising candidate for methane detection. The mechanism for the enhanced methane-sensing performance of the materials were thoroughly discussed.

2. Experimental Section

2.1. Materials

1,4-benzenedicarboxylic acid (H2BDC), N,N-dimethylformamide (DMF), polyvinylpyrrolidone K30 (PVP K30), potassium iodide (KI), and formamide (FA) were acquired from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. Palladium chloride (PdCl2) and indium nitrate hydrate (In(NO3)3·xH2O) were purchased from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China. All the chemicals and reagents were of chemical purity and used as received.

2.2. Preparation of Pd Nanoparticles Suspension, Porous Hollow In2O3 Tubes and Pd@In2O3 Porous Hollow Tubes

The synthesis of Pd nanoparticles was carried out by modifying the previously reported method [33]. First, 50 mg PVP K30 and 170 mg KI were dissolved into 5 mL FA under vigorously stirring and the mixture was heated in oil bath until its temperature reached 120 °C. Afterwards, 36 mg PdCl2 was rapidly added into the solution and the temperature maintained at 120 °C for 1 h. Ultimately, the Pd nanoparticles suspension was obtained and used without any other processing.
The MIL-68 (In) precursors were synthesized by a modified solvothermal method [34]. In a typical process, 0.27 g In(NO3)3·xH2O and 0.33 g H2BDC were dissolved in 120 mL DMF. After being stirred for 15 min, the transparent solution was subject to be heated at 120 °C for 45 min in an oil bath. Subsequently, the obtained white precipitate was filtrated and washed with DMF and ethanol three times, respectively, and dried at 60 °C overnight to obtain MIL-68 (In) precursors. Ultimately, the porous In2O3 nanorods were acquired through calcination of the precursors at 500 °C for 50 min with a ramping rate of 5 °C/min in a muffle furnace.
The fabrication process of Pd@In2O3 is schematically illustrated in Figure 1. Typically, a certain volume of Pd NPs suspension (0.63 mL, 1.26 mL, and 1.88 mL) was rapidly injected into the solution of In(NO3)3·xH2O and H2BDC at 120 °C and maintained for 45 min to obtain the Pd@MIL-68 (In) rods. Afterwards, the Pd@MIL-68 (In) precursors were calcined in a muffle furnace at 500 °C for 50 min in air to gain the PdO@In2O3 hollow tubes. Eventually, the Pd@In2O3 PHTs were acquired by reducing PdO@In2O3 using H2 (5 vol.% in N2) at 500 °C for 2 h with a temperature ramp of 2 °C/min in the tube furnace. The as-prepared Pd@In2O3 PHTs were denoted as Pd@In2O3-1, Pd@In2O3-2, Pd@In2O3-3, respectively corresponding to the contents of Pd NPs.

2.3. Characterizations

The crystal structure was characterized by the X-ray diffraction (XRD, DMAX-2500PC) instrument with Cu-Kα radiation (λ = 1.5418 Å) in the range of 10°–90°. In order to observe the morphology and microstructure of the fabricated samples, a SU-70 field emission scanning electron microscope (FE-SEM) with an accelerating voltage of 15 kV, a JEM-1011 transmission electron microscopy (TEM) with an accelerating voltage of 100 kV and a JEM-2100 high-resolution transmission electron microscopy (HR-TEM) with an accelerating voltage of 200 kV were employed. X-ray photoelectron spectroscopy (XPS, Thermo Scientific ESCALAB Xi+, Waltham, MA, USA) using a monochromatic Al-Kα radiation resource (hʋ = 1486.6 eV) was implemented. The Electron Paramagnetic Resonance (EPR) spectra were recorded using a Bruker A300 at 9.85 GHz at 77 K. Room temperature UV–vis diffuse reflectance spectra of In2O3 and Pd@In2O3 were recorded on a Shimadzu UV-3600Plus ultraviolet spectrophotometer.

2.4. Sensing Performances Test

The gas-sensitivity sensors based on prepared sensing material were fabricated as follows. Generally, the synthesized materials were mixed with a mixture of terpineol and ethyl cellulose (M70) with a fixed weight ratio of 1:99, and then the mixture was ground until a uniform slurry was obtained. Afterwards, the paste was uniformly coated on the Al2O3 ceramic substrate with a size of 2 mm in length and 1 mm in width, equipped with a pair of Au electrodes and four Pt wires, as graphically illustrated in Figure 2a. After that, the substrates smeared with sensing materials were calcined at 400 °C for 2 h in air. Ultimately, the substrates were welded on pedestals to gain gas sensors. The gas sensors were aged with a voltage of 5 V added across for about 7 days to stabilize the electrical properties before tested. The gas-sensing properties were evaluated on a WS-30A gas sensitivity instrument (Zhengzhou Winsen Electronics Co., Ltd., Zhengzhou, China) at 25 °C without air humidity exceeding 30%. The concentrations of methane were regulated by adjusting the volume of methane and dry air injected into the glass chamber. The testing system is illustrated in Figure 2b, in which the Rh, Rx, and RL are heating resistance, material resistance, and load resistance, respectively. Generally, adjustment of operating temperatures is realized by tuning the voltage (Vh) applied across the Rh. Meanwhile, a circle voltage (Vc = 5 V) is supplied across the sensing material and load resistance. The output electric signal is the voltage (Vout) across the load resistor. Furthermore, the resistance of the sensing material could be expressed by the formula Rx = (Vc − Vout) RL/Vc. The sensitivity (S) of the reducing gas can be defined as S = Ra/Rg, where Ra and Rg are resistances of sensing material in air and methane, respectively. In addition to the sensitivity, the optimal operating temperature, response and recovery time, long-term stability, as well as selectivity are also evaluated as they are significant indicators to evaluate the sensing performance of a methane gas sensor [35]. The response or recovery time is well-defined as the time for a sensor required to achieve 90% of the whole resistance change when methane is input or output.

3. Results and Discussion

3.1. Characterizations

The phase structure of the fabricated sensing materials was investigated via XRD. Figure 3a shows the XRD pattern of prepared precursors, which agrees well with the previously reported data [34], confirming the successful synthesis of MIL-68 (In) precursors. Meanwhile, XRD patterns of pristine In2O3 and Pd@In2O3 with various Pd loadings are shown in Figure 3b. For the pristine In2O3, it can be observed that all diffraction peaks can be well indexed to cubic structure of In2O3 (JCPDS card No. 06-0416). As for Pd@In2O3 samples, there is no obvious Pd signal observed, and this can be chiefly attributed to the high dispersion of Pd with low loadings.
The morphologies of prepared samples are displayed in Figure 4. As illustrated in Figure 4a, the TEM image demonstrates that the size of prepared Pd NPs ranges from 4 to 8 nm. As can be observed in Figure 4b,c, MIL-68 (In) precursors are solid hexagonal prisms with smooth surface, and the length of them varies from approximately 5 to 18 μm and the diameter is about 2 μm. After the calcination at 500 °C, the solid hexagonal rods transformed into the hollow microstructure with rough surface and porous shell, almost maintaining the geometric characteristics of MIL-68 (In) rods, which is illustrated in Figure 4d and the inset image. Due to the Kirkendall effect, the transformation can be attributed to the inter-diffusion between In ions and O ions [36]. To further investigate the morphology and microstructure of Pd@In2O3-2, TEM and HRTEM images are taken as shown in Figure 4e,f. It is obvious that Pd@In2O3-2 PHTs are composed of abundant small nanoparticles, which is accordant with the observation from SEM images. Figure 4f reveals the lattice fringes, among which the lattice spacings of 0.223 nm and 0.413 nm correspond with the (111) plane of Pd and (211) plane of In2O3, respectively, implying the successful hybridization of Pd nanoparticles into In2O3.
The surface properties of pristine In2O3 and Pd@In2O3-2 were characterized by XPS. In the In 3d spectra shown in Figure 5a, the peaks at 444.1 eV and 451.69 eV in pristine In2O3 and the peaks at 444.12 eV and 451.64 eV in Pd@In2O3-2 correspond to In 3d5/2 and In 3d3/2 orbitals of In3+, respectively [37,38]. The XPS spectra of Pd 3d orbitals illustrated in Figure 5b signify that the two characteristic peaks situated at ca. 335.2 eV (Pd 3d5/2) and ca. 340.45 eV (Pd 3d3/2), with an energy interval of 5.25 eV are related to Pd0 [39]. Compared with the previous report [40], a shift of 0.3 eV towards the lower binding energy is observed, which is primarily triggered by the charge transfer between In2O3 and Pd NPs due to their different work functions. The peaks located at 337.23 eV and 342.6 eV are associated with Pd2+, indicating a small amount of PdO [41], mainly owing to the inevitable oxidation during storage.
To study the presence and content of oxygen species in pristine In2O3 and Pd@In2O3-2 PHTs, XPS spectra of O 1s orbital were analyzed. In Figure 5c,d, three deconvoluted peaks can be observed at ca. 529.8 eV, ca. 531.5, and ca. 532.7 eV, corresponding to the lattice oxygen (Olatt), the surface oxygen vacancies (Ov), and chemically absorbed oxygen species (Ochem)(O, O2− and O 2 ), respectively [38,42]. The contents of the surface oxygen vacancies in pristine In2O3 and Pd@In2O3-2 are estimated to be 25.78% and 31.01%, respectively, presenting an increase in surface oxygen vacancies after the addition of Pd NPs. The increase of the defect was further confirmed by EPR. The g factor which is corresponding to ionized Ov is calculated to be 2.003, as shown in Figure 5e. The signal intensity of g value is correlated with the amount of oxygen defect, thus more oxygen vacancies can be observed in the Pd@In2O3-2 sample, which is consistent with the results of XPS. The augment of oxygen vacancies could make a positive effect on methane sensing performance. On one hand, the Ov species can provide additional active sites for reactants [43,44]. On the other hand, as electron donors, the increased oxygen vacancies enlarge the number of free electrons in the conduction band, thereby elevating the electronic mobility. In order to investigate the effect of the improvement of the defect on the energy band gap, UV-vis spectra were recorded in Figure 5f. The band gaps were calculated based on the Tauc model by the formula (αhν)n = A (hν − Eg), where α, h, ν, A, and Eg are the adsorption coefficient, Planck constant, light frequency, a material-dependent constant, and the band gap, respectively. For In2O3-based materials, n was chosen to be 2 [45]. As shown in the inset of Figure 5f, the band gaps are 2.69 eV and 2.50 eV for pristine In2O3 and Pd@In2O3-2, respectively. As shallow donor states under the conduction band, oxygen vacancies in metal oxide semiconductor could narrow the forbidden band gap. Hence, the improvement results in decrease of the forbidden gap width, so as to lower the energy barrier of methane adsorption and adjust the electronic structure of In2O3-based material, which is conducive to enhancing the sensitivity [46,47].

3.2. Gas-Sensing Performances towards Methane

Generally speaking, it is laborious to detect methane at low temperatures due to their high chemical thermal stability [15,48,49]. Thus, the working temperature should be elevated to increase the atomic activity of the surface between the material and methane molecules. For a gas sensor, there is an optimum operating temperature (O. T.) when the sensor exhibits the most outstanding sensitivity. A lower optimal operating temperature is expected to consume less energy, prolong the service life of sensor devices, and ensure the operation safety. The operating temperature has great impact on the reaction kinetics of the gas-sensing process, thus significantly influence the sensing performance [50].
In order to confirm the optimal operating temperature of the sensors, responses of all the fabricated sensors towards 5000 ppm methane at different temperatures were tested and recorded. As shown in Figure 6 and Table 1, a typical volcanic shape is observed in all samples. Originally, the sensitivity of the sensors towards methane is improved with the working temperature rising mainly because of the increased surface reaction kinetics. However, further increasing the temperature could not bring in the growth of responsibility. It is observed that the optimal operating temperature of Pd@In2O3-2 is 370 °C, which is 50 °C lower than that of the pristine In2O3 sensor (420 °C). Moreover, as shown in Figure 6 and Table 1, the Pd@In2O3-2 sensor exhibits the highest selectivity (S = 23.2) among all the samples at the optimum operating temperature of 370 °C, approximately 15.5 times higher than that of In2O3 sensor. When the response reaches the maximum, gas desorption assumes the predominance of the sensing reaction and restrains the sensing performance because of the higher temperature. The methane sensing performances of the Pd@In2O3-2 were further studied at the optimal operating temperature.
High sensitivity is not the only criterion for evaluation of the gas sensors. The real-time kinetic curve of Pd@In2O3-2 sensors towards CH4 at 370 °C was further investigated and displayed in Figure 7a. The response gradually increases with the concentration of CH4 rising from 100 to 5000 ppm. Even when the concentration of methane is as low as 100 ppm, an obvious response value of 2.5 is observed. Figure 7b reveals that an evident linear correlation can be fitted as y = −1.82 + 0.57x between log S and log C, where S is the response and C is the concentration of methane. The satisfying linear fitting makes the quantitative detection of methane possible, signifying the potential application in device manufacture. According to the Martins-Naes theory for multivariate calibration, the detection limit (DL) can be extrapolated by the empirical formula as DL = 3 rms/slope, where rms is the root-mean-square deviation of the sensor noise calculated from the photocurrent signals in ambient air and the slope is obtained from the linear function [51]. Additionally, the real-time response curve of the Pd@In2O3-2 sensor to 5000 ppm CH4 is illustrated in Figure 7c, demonstrating a high-speed response and recovery. Its response and recovery time exposed to 5000 ppm methane is 7 s and 5 s, respectively. The fast response/recovery times are basically assigned to the efficient catalytic effect of Pd nanoparticles, which accelerates the reaction between adsorbed methane molecules and the oxygen species on the surface of sensors.
Selectivity is another crucial aspect to assess gas sensing properties. The response of pure In2O3 and Pd@In2O3-2 towards various interfering gases, such as nitrogen dioxide (NO2, 1 ppm), formaldehyde (HCHO, 1 ppm), sulfur dioxide (SO2, 5 ppm), carbon monoxide (CO, 24 ppm), sulfureted hydrogen (H2S, 40 ppm), and ammonia (NH3, 40 ppm), were investigated, which is depicted in Figure 7d. The selection of the concentrations of interfering gases is confirmed by the latest edition of coal mine security regulations for reference [15,52]. Compared with the extremely low response to CH4, pristine In2O3 performed higher responses to some interfering gases, such as H2S (57.9) and NH3 (20.2). However, it is observed that the responses of Pd@In2O3-2 to interferential gases are pretty weak (approximately 1) after adding moderate Pd NPs, which are much lower than that to methane (23.2). It is generally acknowledged that the value of response (Ra/Rg > 2) is usually chosen as the standard for a valid response and thus the influence of interfering gases can be neglected in this work [29]. The modest introduction of Pd NPs promote not only the performance towards methane, but also the ability to resist interference of both H2S and NH3.
Repeatability is another essential factor to evaluate the performances of gas sensors. As depicted in Figure 8a, the almost identical responses of Pd@In2O3-2 exposed to 5000 ppm methane based on three-times measurements demonstrate the satisfactory repeatability of Pd@In2O3-2. Furthermore, as shown in Figure 8b, the response maintains relatively stable during the one-month test, demonstrating an excellent long-term stability. In this work, the Pd@In2O3-2 based sensor exhibits prominent sensing properties and is highly competitive in methane detection. The improved sensing performance can be attributed to the unique porous morphology of hexagonal hollow tubes, as well as abundant oxygen vacancies originating from the thermal decomposition of MIL-68 (In) precursors and the existence of Pd NPs. Table 2 compares the methane sensing performances of Pd@In2O3-2 in this work with those in other literatures.

3.3. Gas-Sensing Mechanism

The porous, hollow, rod-like structure with large specific surface area inherited from MIL-68 (In) precursors provides abundant channels for methane molecules to diffuse in the sensing layer and affluent active sites for reaction, which is illustrated in Figure 9a. The surface adsorbed oxygen model is widely recognized as the gas-sensing mechanism, and the detailed interpretation is as follows. During gas adsorption-desorption process, electronic concentration in conduction band will make a huge difference on sensor resistance. First, when the sensing materials are exposed to the air, oxygen is physically adsorbed on the surface of the sensing materials without electron transfer during this process. With temperature gradually elevated, oxygen molecules capture electrons from the conduction band, forming chemisorbed oxygen species such as O 2 , O, and O2− on the surface. Meanwhile, the thickness of electron depletion layer (EDL) increases, resulting in an increase in resistance. This process can be elaborated as follows [59,60]:
O2gas → O2ads,
O 2 ads + e O 2 ads
O 2 ads + e 2   O ads
Oads + e → O2−ads.
Indium oxide (In2O3) is a typically n-type semiconductor metal oxide. Once In2O3 encounters with methane, the reducing gas will react with the active oxygen ions ( O 2 , O, and O2−) and generates CO2 and H2O with electrons releasing back to the conduction band. Consequently, the thickness of EDL and resistance tend to descend. Therefore, measuring the resistance in air and methane atmosphere is an efficient method to monitor the CH4 in real time [61]. The process is shown as Equations (5) and (6) [62]:
CH4gas + 4 Oads → CO2gas + 2 H2Ogas+ 4 e,
CH4gas + 4 O2−ads → CO2gas + 2 H2Ogas + 8 e.
In this work, the improved sensing performance can be mainly attributed to the catalytic effect of Pd nanoparticles. The schematic diagram of methane sensing mechanism for Pd@In2O3-2 PHTs in air and methane was shown in Figure 9b.
On one hand, the Fermi levels of noble metals usually lie below the conduction band of semiconductor metal oxide, and the work function of Pd (Wm = 5.60 eV) is higher than that of In2O3 (Ws = 5.0 eV) [63,64]. Thus, the energy bands of In2O3 and Pd on contact surface will bend until the Fermi energy levels reach the equilibrium. The change of the Fermi level will promote the electrons to transfer from the surface of In2O3 to Pd NPs, motivating the sensing material to perform enhanced performance. On the other hand, the catalytic effect of Pd NPs effectively activates dissociation of oxygen molecules, which will flow over from the complex and diffuse on the surface of In2O3, capturing electrons from the metal oxide [65]. This process increases the concentration of adsorbed oxygen ions and makes the depletion layer wider, providing more active sites for the adsorption reaction of methane [66]. Moreover, when the sensing material is exposed to methane, Pd will facilitate the dissociation of C–H bond to form H and CH3, which spill over to the surface and react with chemically adsorbed oxygen ions on the surface of sensitive film, generating water molecules. During this process, the trapped electrons release back to the conduction band of In2O3, thus the thickness of EDL and the resistance of Pd@In2O3 interface decreases, reducing the barrier height of reaction [67]. The specific reactions on the surface of Pd@In2O3-2 are as follows:
CH4 → CH3ads + Hads,
CH3 + H + 4 O2− → CO2 + 2 H2O + 8 e.

4. Conclusions

In this work, Pd@In2O3 porous hollow tubes derived from Pd@MIL-68 (In) precursors with core-shell structure were successfully synthesized through a hydrothermal method with subsequent calcination and reduction processes. Among all prepared samples, Pd@In2O3-2 exhibited the best performance in methane detection. With a response value of 23.2 at a relatively lower working temperature (370 °C), the sensitivity of Pd@In2O3-2 towards 5000 ppm methane was about 15.5 times higher than that of pristine In2O3, and the response and recovery times were only 7 s and 5 s, respectively. Furthermore, Pd@In2O3-2 PHTs also enable to selectively track methane with the presence of interfering gases, exhibiting an outstanding anti-interference capacity. Thus, this research indicates that the porous Pd@In2O3 are promising candidates for fabrication of methane sensors due to the large specific surface area, porous microstructure, increased oxygen vacancies and the catalytic effect of Pd nanoparticles.

Author Contributions

X.Z.: Conceptualization, Investigation, Formal analysis, Writing—original draft & editing. Z.Y.: Investigation, Formal analysis. J.K.: Supervision, Writing—review & editing. Z.H.: Investigation, Formal analysis, Writing—review & editing. J.Z.: Data curation, Supervision. X.M.: Investigation, Formal analysis. S.H.: Investigation, Formal analysis. L.W.: Data curation, Writing—review & editing. S.W.: Investigation, Formal analysis. J.L.: Funding acquisition, Writing—review & editing. Z.W.: Data curation. F.W.: Conceptualization, Funding acquisition, Supervision, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science and Development Foundation of Shenzhen (JCYJ20190807093205660) And The APC was waived as an invited paper.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All data are included in the paper.

Conflicts of Interest

There are no conflict of interest to declare.

References

  1. Schoonbaert, S.; Tyner, D.; Johnson, M. Remote ambient methane monitoring using fiber-optically coupled optical sensors. Appl. Phys. B Lasers Opt. 2015, 119, 133–142. [Google Scholar] [CrossRef] [Green Version]
  2. Xue, D.; Wang, Y.; Cao, J.; Sun, G.; Zhang, Z. Improving methane gas sensing performance of flower-like SnO2 decorated by WO3 nanoplates. Talanta 2019, 199, 603–611. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, D.; Chang, H.; Sun, Y.; Jiang, C.; Yao, Y.; Zhang, Y. Fabrication of platinum-loaded cobalt oxide/molybdenum disulfide nanocomposite toward methane gas sensing at low temperature. Sens. Actuators B Chem. 2017, 252, 624–632. [Google Scholar] [CrossRef]
  4. Smedt, G.D.; Corte, F.; Notele, R.; Berghmans, J. Comparison of two standard test methods for determining explosion limits of gases at atmospheric conditions. J. Hazard. Mater. 1999, 70, 105–113. [Google Scholar] [CrossRef] [PubMed]
  5. Eugster, W.; Kling, G. Performance of a low-cost methane sensor for ambient concentration measurements in preliminary studies. Atmos. Meas. Tech. 2012, 5, 1925–1934. [Google Scholar] [CrossRef] [Green Version]
  6. Hao, S.; Wen, J.; Yu, X.; Chu, W. Effect of the surface oxygen groups on methane adsorption on coals. Appl. Surf. Sci. 2013, 264, 433–442. [Google Scholar] [CrossRef]
  7. Lawrence, N. Analytical detection methodologies for methane and related hydrocarbons. Talanta 2006, 69, 385–392. [Google Scholar] [CrossRef]
  8. Zhang, B.; Wang, Y.; Meng, X.; Zhang, Z.; Mu, S. High response methane sensor based on Au-modified hierarchical porous nanosheets-assembled ZnO microspheres. Mater. Chem. Phys. 2020, 250, 123027. [Google Scholar] [CrossRef]
  9. Kooti, M.; Keshtkar, S.; Askarieh, M.; Rashidi, A. Progress toward a novel methane gas sensor based on SnO2 nanorods-nanoporous graphene hybrid. Sens. Actuators B Chem. 2019, 281, 96–106. [Google Scholar] [CrossRef]
  10. Vuong, N.; Hieu, N.; Hieu, H.; Yi, H.; Kim, D.; Han, Y.-S.; Kim, M. Ni2O3-decorated SnO2 particulate films for methane gas sensors. Sens. Actuators B Chem. 2014, 192, 327–333. [Google Scholar] [CrossRef]
  11. Wang, J.; Hu, C.; Xia, Y.; Zhang, B. Mesoporous ZnO nanosheets with rich surface oxygen vacancies for UV-activated methane gas sensing at room temperature. Sens. Actuators B Chem. 2021, 333, 129547. [Google Scholar] [CrossRef]
  12. Barreca, D.; Bekermann, D.; Comini, E.; Devi, A.; Fischer, R.; Gasparotto, A.; Maccato, C.; Sberveglieri, G.; Tondello, E. 1D ZnO nano-assemblies by Plasma-CVD as chemical sensors for flammable and toxic gases. Sens. Actuators B Chem. 2010, 149, 1–7. [Google Scholar] [CrossRef]
  13. Prasad, A.; Amirthapandian, S.; Dhara, S.; Dash, S.; Murali, N.; Tyagi, A. Novel single phase vanadium dioxide nanostructured films for methane sensing near room temperature. Sens. Actuators B Chem. 2014, 191, 252–256. [Google Scholar] [CrossRef] [Green Version]
  14. Zhang, S.; Zhang, B.; Zhang, B.; Wang, Y.; Bala, H.; Zhang, Z. Structural evolution of NiO from porous nanorods to coral-like nanochains with enhanced methane sensing performance. Sens. Actuators B Chem. 2021, 334, 129645. [Google Scholar] [CrossRef]
  15. Wang, Y.; Yao, M.; Guan, R.; Zhang, Z.; Cao, J. Enhanced methane sensing performance of NiO decorated In2O3 nanospheres composites at low temperature. J. Alloys Compd. 2021, 854, 157169. [Google Scholar] [CrossRef]
  16. Shaalan, N.; Rashad, M.; Abdel-Rahim, M. Repeatability of indium oxide gas sensors for detecting methane at low temperature. Mater. Sci. Semicond. Process. 2016, 56, 260–264. [Google Scholar] [CrossRef]
  17. Shen, Y.; Bi, H.; Li, T.; Zhong, X.; Chen, X.; Fan, A.; Wei, D. Low-temperature and highly enhanced NO2 sensing performance of Au-functionalized WO3 microspheres with a hierarchical nanostructure. Appl. Surf. Sci. 2018, 434, 922–931. [Google Scholar] [CrossRef]
  18. Han, S.; Herman, G.; Chang, C. Low-temperature, high-performance, solution-processed indium oxide thin-film transistors. J. Am. Chem. Soc. 2011, 133, 5166–5169. [Google Scholar] [CrossRef]
  19. Xue, D.; Zhang, S.; Zhang, Z. Hydrothermal synthesis of methane sensitive porous In2O3 nanosheets. Mater. Lett. 2019, 252, 169–172. [Google Scholar] [CrossRef]
  20. Vuong, N.; Hieu, N.; Kim, D.; Choi, B.; Kim, M. Ni2O3 decoration of In2O3 nanostructures for catalytically enhanced methane sensing. Appl. Surf. Sci. 2014, 317, 765–770. [Google Scholar] [CrossRef]
  21. Xue, D.; Wang, Y.; Zhang, Z.; Cao, J. Porous In2O3 nanospheres with high methane sensitivity: A combined experimental and first-principle study. Sens. Actuator A Phys. 2020, 305, 111944. [Google Scholar] [CrossRef]
  22. Nguyen, C.; Vu, N.; Do, T.-O. Recent advances in the development of sunlight-driven hollow structure photocatalysts and their applications. J. Mater. Chem. A 2015, 3, 18345–18359. [Google Scholar] [CrossRef]
  23. Corma, A.; Garcia, H.; Xamena, F. Engineering Metal Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606–4655. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, W.; Sun, W.; Wang, Y. Multilayer CuO@NiO Hollow Spheres: Microwave-Assisted Metal-Organic-Framework Derivation and Highly Reversible Structure-Matched Stepwise Lithium Storage. ACS Appl. Nano Mater. 2015, 9, 11462–11471. [Google Scholar] [CrossRef]
  25. Zou, X.; Zhu, G.; Hewitt, I.; Sun, F.; Qiu, S. Synthesis of a metal-organic framework film by direct conversion technique for VOCs sensing. Dalton Trans. 2009, 3009–3013. [Google Scholar] [CrossRef]
  26. Zheng, G.; Pastoriza-Santos, I.; Pérez-Juste, J.; Liz-Marzán, L. Plasmonic metal-organic frameworks. SmartMat 2021, 2, 446–465. [Google Scholar] [CrossRef]
  27. Sun, J.-K.; Xu, Q. Functional materials derived from open framework templates/precursors: Synthesis and applications. Energy Environ. Sci. 2014, 7, 2071–2100. [Google Scholar] [CrossRef]
  28. Li, R.; Sun, L.; Zhan, W.; Li, Y.-A.; Wang, X.; Han, X. Engineering an effective noble-metal-free photocatalyst for hydrogen evolution: Hollow hexagonal porous micro-rods assembled from In2O3@carbon core-shell nanoparticles. J. Mater. Chem. A 2018, 6, 15747–15754. [Google Scholar] [CrossRef]
  29. Li, S.; Xie, L.; He, M.; Hu, X.; Luo, G.; Chen, C.; Zhu, Z. Metal-Organic frameworks-derived bamboo-like CuO/In2O3 Heterostructure for high-performance H2S gas sensor with Low operating temperature. Sens. Actuators B Chem. 2020, 310, 127828. [Google Scholar] [CrossRef]
  30. Ma, J.; Fan, H.; Zheng, X.; Wang, H.; Zhao, N.; Zhang, M.; Yadava, A.; Wang, W.; Dong, W.; Wang, S. Facile metal-organic frameworks-templated fabrication of hollow indium oxide microstructures for chlorine detection at low temperature. J. Hazard. Mater. 2020, 387, 122017. [Google Scholar] [CrossRef]
  31. Wang, Y.; Meng, X.; Yao, M.; Sun, G.; Zhang, Z. Enhanced CH4 sensing properties of Pd modified ZnO nanosheets. Ceram. Int. 2019, 45, 13150–13157. [Google Scholar] [CrossRef]
  32. Yang, L.; Wang, Z.; Zhou, X.; Wu, X.; Han, N.; Chen, Y. Synthesis of Pd-loaded mesoporous SnO2 hollow spheres for highly sensitive and stable methane gas sensors. RSC Adv. 2018, 8, 24268–24275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Xu, B.; Zhang, Z.; Wang, X. Formamide: An efficient solvent to synthesize water-soluble and sub-ten-nanometer nanocrystals. Nanoscale 2013, 5, 4495–4505. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, S.; Guan, B.; Lu, Y.; Lou, X. Formation of Hierarchical In2S3-CdIn2S4 Heterostructured Nanotubes for Efficient and Stable Visible Light CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 17305–17308. [Google Scholar] [CrossRef] [PubMed]
  35. Tshabalala, Z.; Shingange, K.; Dhonge, B.; Ntwaeaborwa, O.; Mhlongo, G.; Motaung, D. Fabrication of ultra-high sensitive and selective CH4 room temperature gas sensing of TiO2 nanorods: Detailed study on the annealing temperature. Sens. Actuators B Chem. 2017, 238, 402–419. [Google Scholar] [CrossRef]
  36. Wang, H.; Zhuo, S.; Liang, Y.; Han, X.; Zhang, B. General Self-Template Synthesis of Transition-Metal Oxide and Chalcogenide Mesoporous Nanotubes with Enhanced Electrochemical Performances. Angew. Chem. Int. Ed. 2016, 55, 9055–9059. [Google Scholar] [CrossRef] [PubMed]
  37. Lei, F.; Sun, Y.; Liu, K.; Gao, S.; Liang, L.; Pan, B.; Xie, Y. Oxygen vacancies confined in ultrathin indium oxide porous sheets for promoted visible-light water splitting. J. Am. Chem. Soc. 2014, 136, 6826–6829. [Google Scholar] [CrossRef]
  38. Wang, L.; Gao, J.; Wu, B.; Kan, K.; Xu, S.; Xie, Y.; Li, L.; Shi, K. Designed Synthesis of In2O3 Beads@TiO2-In2O3 Composite Nanofibers for High Performance NO2 Sensor at Room Temperature. ACS Appl. Mater. Interfaces 2015, 7, 27152–27159. [Google Scholar] [CrossRef]
  39. Lupan, O.; Postica, V.; Hoppe, M.; Wolff, N.; Polonskyi, O.; Pauporte, T.; Viana, B.; Majerus, O.; Kienle, L.; Faupel, F.; et al. PdO/PdO2 functionalized ZnO: Pd films for lower operating temperature H2 gas sensing. Nanoscale 2018, 10, 14107–14127. [Google Scholar] [CrossRef]
  40. Rai, P.; Yoon, J.-W.; Kwak, C.-H.; Lee, J.-H. Role of Pd nanoparticles in gas sensing behaviour of Pd@In2O3 yolk-shell nanoreactors. J. Mater. Chem. A 2016, 4, 264–269. [Google Scholar] [CrossRef]
  41. Mahara, Y.; Tojo, T.; Murata, K.; Ohyama, J.; Satsuma, A. Methane combustion over Pd/CoAl2O4/Al2O3 catalysts prepared by galvanic deposition. RSC Adv. 2017, 7, 34530–34537. [Google Scholar] [CrossRef] [Green Version]
  42. Li, X.; He, X.; Liu, X.; He, L.-N. Ruthenium-promoted reductive transformation of CO2. Sci. China Chem. 2017, 60, 841–852. [Google Scholar] [CrossRef]
  43. Dai, Z.; Lee, C.-S.; Kim, B.-Y.; Kwak, C.-H.; Yoon, J.-W.; Jeong, H.-M.; Lee, J.-H. Honeycomb-like periodic porous LaFeO3 thin film chemiresistors with enhanced gas-sensing performances. ACS Appl. Mater. Interfaces 2014, 6, 16217–16226. [Google Scholar] [CrossRef] [PubMed]
  44. Dai, Z.; Lee, C.-S.; Tian, Y.; Kim, I.-D.; Lee, J.-H. Highly reversible switching from P- to N-type NO2 sensing in a monolayer Fe2O3 inverse opal film and the associated P–N transition phase diagram. J. Mater. Chem. A 2015, 3, 3372–3381. [Google Scholar] [CrossRef]
  45. Shanmugasundaram, A.; Ramireddy, B.; Basak, P.; Manorama, S.; Srinath, S. Hierarchical In(OH)3 as a Precursor to Mesoporous In2O3 Nanocubes: A Facile Synthesis Route, Mechanism of Self-Assembly, and Enhanced Sensing Response toward Hydrogen. J. Phys. Chem. C 2014, 118, 6909–6921. [Google Scholar] [CrossRef]
  46. Du, W.; Si, W.; Wang, F.; Lv, L.; Wu, L.; Wang, Z.; Liu, J.; Liu, W. Creating oxygen vacancies on porous indium oxide nanospheres via metallic aluminum reduction for enhanced nitrogen dioxide detection at low temperature. Sens. Actuators B Chem. 2020, 303, 127221. [Google Scholar] [CrossRef]
  47. Li, K.; Luo, Y.; Liu, B.; Gao, L.; Duan, G. High-performance NO2-gas sensing of ultrasmall ZnFe2O4 nanoparticles based on surface charge transfer. J. Mater. Chem. A 2019, 7, 5539–5551. [Google Scholar] [CrossRef]
  48. Xue, D.; Wang, J.; Wang, Y.; Sun, G.; Cao, J.; Bala, H.; Zhang, Z. Enhanced Methane Sensing Properties of WO3 Nanosheets with Dominant Exposed (200) Facet via Loading of SnO2 Nanoparticles. Nanomaterials 2019, 9, 351. [Google Scholar] [CrossRef] [Green Version]
  49. Chen, A.; Bai, S.; Shi, B.; Liu, Z.; Li, D.; Liu, C. Methane gas-sensing and catalytic oxidation activity of SnO2–In2O3 nanocomposites incorporating TiO2. Sens. Actuators B Chem. 2008, 135, 7–12. [Google Scholar] [CrossRef]
  50. Wang, D.; Wan, K.; Zhang, M.; Li, H.; Wang, P.; Wang, X.; Yang, J. Constructing hierarchical SnO2 nanofiber/nanosheets for efficient formaldehyde detection. Sens. Actuators B Chem. 2019, 283, 714–723. [Google Scholar] [CrossRef]
  51. Peng, L.; Zhao, Q.; Wang, D.; Zhai, J.; Wang, P.; Pang, S.; Xie, T. Ultraviolet-assisted gas sensing: A potential formaldehyde detection approach at room temperature based on zinc oxide nanorods. Sens. Actuators B Chem. 2009, 136, 80–85. [Google Scholar] [CrossRef]
  52. Wang, Y.; Cui, Y.; Meng, X.; Zhang, Z.; Cao, J. A gas sensor based on Ag-modified ZnO flower-like microspheres: Temperature-modulated dual selectivity to CO and CH4. Surf. Interfaces 2021, 24, 101110. [Google Scholar] [CrossRef]
  53. Liu, S.; Xie, M.; Li, Y.; Guo, X.; Ji, W.; Ding, W. Synthesis and selective gas-sensing properties of hierarchically porous intestine-like SnO2 hollow nanostructures. Mater. Chem. Phys. 2010, 123, 109–113. [Google Scholar] [CrossRef]
  54. Li, G.; Wang, X.; Yan, L.; Wang, Y.; Zhang, Z.; Xu, J. PdPt Bimetal-Functionalized SnO2 Nanosheets: Controllable Synthesis and its Dual Selectivity for Detection of Carbon Monoxide and Methane. ACS Appl. Mater. Interfaces 2019, 11, 26116–26126. [Google Scholar] [CrossRef]
  55. Min, B.-K.; Choi, S.-D. Undoped and 0.1wt.% Ca-doped Pt-catalyzed SnO2 sensors for CH4 detection. Sens. Actuators B Chem. 2005, 108, 119–124. [Google Scholar] [CrossRef]
  56. Wagner, T.; Bauer, M.; Sauerwald, T.; Kohl, C.; Tiemann, M. X-ray absorption near-edge spectroscopy investigation of the oxidation state of Pd species in nanoporous SnO2 gas sensors for methane detection. Thin Solid Film. 2011, 520, 909–912. [Google Scholar] [CrossRef]
  57. Lu, W.; Ding, D.; Xue, Q.; Du, Y.; Xiong, Y.; Zhang, J.; Pan, X.; Xing, W. Great enhancement of CH4 sensitivity of SnO2 based nanofibers by heterogeneous sensitization and catalytic effect. Sens. Actuators B Chem. 2018, 254, 393–401. [Google Scholar] [CrossRef]
  58. Cao, Y.; Zhao, J.; Zou, X.; Jin, P.-P.; Chen, H.; Gao, R.; Zhou, L.-J.; Zou, Y.-C.; Li, G.-D. Synthesis of porous In2O3 microspheres as a sensitive material for early warning of hydrocarbon explosions. RSC Adv. 2015, 5, 5424–5431. [Google Scholar] [CrossRef]
  59. Takata, M.; Tsubone, D.; Yanagida, H. Dependence of Electrical Conductivity of ZnO on Degree of Sintering. J. Am. Ceram. Soc. 1976, 59, 4–8. [Google Scholar] [CrossRef]
  60. Song, F.; Su, H.; Han, J.; Xu, J.; Zhang, D. Controllable synthesis and gas response of biomorphic SnO2 with architecture hierarchy of butterfly wings. Sens. Actuators B Chem. 2010, 145, 39–45. [Google Scholar] [CrossRef]
  61. Shinde, V.; Gujar, T.; Lokhande, C. Enhanced response of porous ZnO nanobeads towards LPG: Effect of Pd sensitization. Sens. Actuators B Chem. 2007, 123, 701–706. [Google Scholar] [CrossRef]
  62. Amutha, A.; Amirthapandian, S.; Prasad, A.; Panigrahi, B.; Thangadurai, P. Methane gas sensing at relatively low operating temperature by hydrothermally prepared SnO2 nanorods. J. Nanopart. Res. 2015, 17, 289. [Google Scholar] [CrossRef]
  63. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  64. Wang, Y.; Li, Y.; Yu, K.; Zhu, Z. Controllable synthesis and field emission enhancement of Al2O3 coated In2O3 core–shell nanostructures. J. Phys. D Appl. Phys. 2011, 44, 105301. [Google Scholar] [CrossRef]
  65. Tait, S.; Dohnálek, Z.; Campbell, C.; Kay, B. Methane adsorption and dissociation and oxygen adsorption and reaction with CO on Pd nanoparticles on MgO(100) and on Pd(111). Surf. Sci. 2005, 591, 90–107. [Google Scholar] [CrossRef]
  66. Jang, B.-H.; Landau, O.; Choi, S.-J.; Shin, J.; Rothschild, A.; Kim, I.-D. Selectivity enhancement of SnO2 nanofiber gas sensors by functionalization with Pt nanocatalysts and manipulation of the operation temperature. Sens. Actuators B Chem. 2013, 188, 156–168. [Google Scholar] [CrossRef]
  67. Basu, S.; Basu, P. Nanocrystalline Metal Oxides for Methane Sensors: Role of Noble Metals. J. Sens. 2009, 2009, 861968. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the preparation process of Pd@In2O3 PHTs.
Figure 1. Schematic illustration of the preparation process of Pd@In2O3 PHTs.
Sensors 23 01163 g001
Figure 2. (a) Schematic diagram of gas sensor element and (b) the sensor testing principle.
Figure 2. (a) Schematic diagram of gas sensor element and (b) the sensor testing principle.
Sensors 23 01163 g002
Figure 3. XRD patterns of (a) MIL-68 (In) and (b) pristine In2O3, Pd@In2O3-1, Pd@In2O3-2 and Pd@In2O3-3.
Figure 3. XRD patterns of (a) MIL-68 (In) and (b) pristine In2O3, Pd@In2O3-1, Pd@In2O3-2 and Pd@In2O3-3.
Sensors 23 01163 g003
Figure 4. (a) TEM images of Pd nanoparticles (the inset is the size distribution histograms); SEM images of (b,c) MIL-68 (In) precursors, and (d) pristine In2O3 (the inset is the enlargement of the cross section of In2O3); (e,f) HRTEM images of Pd@In2O3-2.
Figure 4. (a) TEM images of Pd nanoparticles (the inset is the size distribution histograms); SEM images of (b,c) MIL-68 (In) precursors, and (d) pristine In2O3 (the inset is the enlargement of the cross section of In2O3); (e,f) HRTEM images of Pd@In2O3-2.
Sensors 23 01163 g004
Figure 5. XPS spectra of the elements: (a) In 3d of pristine In2O3 and Pd@In2O3-2, (b) Pd 3d of Pd@In2O3-2, O 1s of (c) pristine In2O3 and (d) Pd@In2O3-2; characterization of oxygen vacancies: (e) electron paramagnetic resonance spectra, (f) UV-vis absorbance spectra, the inset is curves of (αhν)2 vs photo energy (hν).
Figure 5. XPS spectra of the elements: (a) In 3d of pristine In2O3 and Pd@In2O3-2, (b) Pd 3d of Pd@In2O3-2, O 1s of (c) pristine In2O3 and (d) Pd@In2O3-2; characterization of oxygen vacancies: (e) electron paramagnetic resonance spectra, (f) UV-vis absorbance spectra, the inset is curves of (αhν)2 vs photo energy (hν).
Sensors 23 01163 g005
Figure 6. Gas response of pristine In2O3, Pd@In2O3-1, Pd@In2O3-2, and Pd@In2O3-3 to 5000 ppm CH4 at different temperatures.
Figure 6. Gas response of pristine In2O3, Pd@In2O3-1, Pd@In2O3-2, and Pd@In2O3-3 to 5000 ppm CH4 at different temperatures.
Sensors 23 01163 g006
Figure 7. (a) Responses of Pd@In2O3-2 towards different concentrations of methane; (b) the linear fitting between ln[C] and ln[S]; (c) response and recovery time of Pd@In2O3-2 when exposed in 5000 ppm methane; (d) responses of pure In2O3 and Pd@In2O3-2 to 5000 ppm methane and other interfering gases.
Figure 7. (a) Responses of Pd@In2O3-2 towards different concentrations of methane; (b) the linear fitting between ln[C] and ln[S]; (c) response and recovery time of Pd@In2O3-2 when exposed in 5000 ppm methane; (d) responses of pure In2O3 and Pd@In2O3-2 to 5000 ppm methane and other interfering gases.
Sensors 23 01163 g007
Figure 8. (a) The repeatability of Pd@In2O3-2 to 5000 ppm CH4; (b) long-term stability of Pd@In2O3-2 sensor towards 5000 ppm CH4 for 30 days.
Figure 8. (a) The repeatability of Pd@In2O3-2 to 5000 ppm CH4; (b) long-term stability of Pd@In2O3-2 sensor towards 5000 ppm CH4 for 30 days.
Sensors 23 01163 g008
Figure 9. (a) The illustration of gas diffusion; Schematic diagram of methane sensing mechanism for Pd@In2O3-2 PHTs exposed to (b) air and (c) methane.
Figure 9. (a) The illustration of gas diffusion; Schematic diagram of methane sensing mechanism for Pd@In2O3-2 PHTs exposed to (b) air and (c) methane.
Sensors 23 01163 g009
Table 1. Response of pristine In2O3 and Pd@In2O3-1, Pd@In2O3-2, and Pd@In2O3-3 samples at different temperatures.
Table 1. Response of pristine In2O3 and Pd@In2O3-1, Pd@In2O3-2, and Pd@In2O3-3 samples at different temperatures.
Sample240 °C300 °C340 °C370 °C400 °C420 °C440 °C
In2O31.01.11.11.51.94.44.3
Pd@In2O3-11.52.94.35.45.34.67.0
Pd@In2O3-21.78.722.423.218.915.112.1
Pd@In2O3-31.22.43.55.16.05.14.1
Table 2. Comparison of methane gas sensing properties of different materials from recent publications.
Table 2. Comparison of methane gas sensing properties of different materials from recent publications.
Sensing MaterialsSensitivityO. T. (°C)Concentration (ppm)Tres/Trec (s)Ref.
SnO21.25400100014/37[53]
PdPt/SnO25.232010005/4[54]
Pt/SnO22.133255000-/-[55]
SnO2-Pd204006600-/-[56]
Pt-SnO24.5350100030/150[57]
In2O3 film2.2530010,000200/-[16]
In2O3 spheres6.503505000-/-[58]
Pd@In2O3 PHTs23.237050007/5This work
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

Zuo, X.; Yang, Z.; Kong, J.; Han, Z.; Zhang, J.; Meng, X.; Hao, S.; Wu, L.; Wu, S.; Liu, J.; et al. Imbedding Pd Nanoparticles into Porous In2O3 Structure for Enhanced Low-Concentration Methane Sensing. Sensors 2023, 23, 1163. https://doi.org/10.3390/s23031163

AMA Style

Zuo X, Yang Z, Kong J, Han Z, Zhang J, Meng X, Hao S, Wu L, Wu S, Liu J, et al. Imbedding Pd Nanoparticles into Porous In2O3 Structure for Enhanced Low-Concentration Methane Sensing. Sensors. 2023; 23(3):1163. https://doi.org/10.3390/s23031163

Chicago/Turabian Style

Zuo, Xiaoyang, Zhengyi Yang, Jing Kong, Zejun Han, Jianxin Zhang, Xiangwei Meng, Shuyan Hao, Lili Wu, Simeng Wu, Jiurong Liu, and et al. 2023. "Imbedding Pd Nanoparticles into Porous In2O3 Structure for Enhanced Low-Concentration Methane Sensing" Sensors 23, no. 3: 1163. https://doi.org/10.3390/s23031163

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