Next Article in Journal
Development of Highly Efficient Lamb Wave Transducers toward Dual-Surface Simultaneous Atomization
Previous Article in Journal
Autonomic Responses Associated with Olfactory Preferences of Fragrance Consumers: Skin Conductance, Respiration, and Heart Rate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving Triethylamine-Sensing Performance of WO3 Nanoplates through In Situ Heterojunction Construction

1
Department of Material and Chemical Engineering, Zhengzhou University of Light Industry, Zhengzhou 450001, China
2
State Key Lab of Transducer Technology, Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Sciences, Shanghai 200050, China
*
Authors to whom correspondence should be addressed.
Sensors 2024, 24(17), 5606; https://doi.org/10.3390/s24175606 (registering DOI)
Submission received: 10 August 2024 / Revised: 21 August 2024 / Accepted: 27 August 2024 / Published: 29 August 2024
(This article belongs to the Special Issue Nanomaterial-Based Biochemical Sensors and Their Applications)

Abstract

:
Surface engineering techniques can be used to develop high-performance gas sensing materials and advance the development of sensors. In this study, we improved the gas sensing performance of two-dimensional (2D) WO3 nanoplates by combining surface Zn modification and the in situ formation of ZnWO4/WO3 heterojunctions. Introducing Zn atoms by surface modification can reconstruct the atomic surface of 2D WO3 nanoplates, creating additional active sites. This allowed for the preparation of various types of ZnWO4/WO3 heterojunctions on the surface of the WO3 nanoplates, which improved the selectivity and sensitivity to the target gas triethylamine. The sensor exhibited good gas sensing performance for triethylamine even at low operating temperatures and strongly resisted humidity changes. The ZnWO4/WO3 material we prepared demonstrated a nearly threefold improvement in the triethylamine (TEA) response, with a gas sensing responsivity of 40.75 for 10 ppm of TEA at 250 °C. The sensor based on ZnWO4/WO3 has a limit of detection (LOD) for TEA of 200 ppb in practical measurements (its theoretical LOD is even as low as 31 ppb). The method of growing ZnWO4 on the surface of WO3 nanoplates using surface modification techniques to form surface heterojunctions differs from ordinary composites. The results suggest that the in situ construction of surface heterojunctions using surface engineering strategies, such as in situ modifying, is a practical approach to enhance the gas sensing properties and resistance to the humidity changes of metal oxide materials.

1. Introduction

Metal oxide semiconductor-based gas sensors play an important role in our lives and are an essential manifestation of technological progress and improved living standards. Currently, these sensors have been widely used in various fields, including safety measures, quality control, and personal health monitoring, where accurate gas detection using sensors is crucial. There are numerous materials with gas-sensitive properties, and the extensive literature suggests that WO3 is an effective gas-sensitive material [1,2]. However, to better meet the application of gas sensors based on WO3 in real life, it is necessary to further enhance the gas sensing performance of WO3, such as low sensitivity, poor selectivity, and poor humidity immunity [3]. Various strategies have been implemented in previous studies to enhance the gas sensing properties of WO3 [2,4,5,6,7]. Surface structure plays a crucial role in improving gas sensing performance, as the conduction behavior of metal oxides is influenced by surface chemical adsorption, reactions, and catalysis during gas molecule interaction [3]. It is conceivable that surface modification of metal oxides can substantially improve their gas sensing capabilities [8].
In this study, we introduced Zn2+ ions through rational modifying into the surface of WO3 nanoplates, forming a unique internal surface heterojunction composed of ZnWO4/WO3 (as illustrated in Figure 1). This heterojunction structure is akin to heteroatom-doped metal oxides but confined explicitly to the surface, generating shallow energy levels within the base material and inducing rearrangements in the atomic structure. The internal heterojunction formed within the substrate oxide effectively segregates electron-hole pairs. Consequently, the ZnWO4/WO3 internal heterojunction structure has been extensively employed in photocatalysis and optoelectronic semiconductor applications [9,10]. Leveraging the efficient charge separation capability of this internal heterojunction, the ZnWO4/WO3 heterojunction improves the adsorption and reaction of target gases on the surface of WO3. Moreover, the incorporation of ZnWO4/WO3 heterojunctions into the surface of WO3 nanoplates enhances the utilization of composites for H2O in the air, promoting the production of more active radicals on the surface while reducing the impact of humidity on gas sensing performance and enhancing the response of the composites [11].
To explore the preparation process of this distinctive heterojunction and its influence on the gas sensing performance of WO3, we synthesized various ZnWO4/WO3 nanoplates with internal heterojunctions using an ion permeability method. The gas sensing performance of the sensors based on ZnWO4/WO3 nanoplates exhibited significant improvement. Notably, the gas sensing response to triethylamine (TEA) at low operating temperatures exhibited a remarkable increase, approximately four times higher than the response of WO3 to 10 ppm TEA at 250 °C (Ra/Rg = 40.75). TEA is a colorless and transparent toxic gas with a strong ammonia odor, widely used in industrial and commercial applications. According to the Occupational Safety and Health Administration (OSHA) suggestion, the threshold limit value (TLV) for TEA is 10 ppm [12], and the American Conference of Governmental Industrial Hygienists (ACGIH) recommends a TLV of 1 ppm [13]. Fast, sensitive, and selective TEA detection is crucial for the chemical industry and our everyday lives.
Structural characterization using XRD, TEM, and EDS revealed that the ZnWO4 phase is embedded within the WO3 nanoplates, forming “nano-dots”, “nano-islands”, or “continuous domain” structures on the surface. This island-like distribution of the ZnWO4/WO3 heterojunctions, coupled with efficient electron-hole separation, results in electron-rich regions on the surface of the WO3 nanoplates. This alteration affects the adsorption and reaction of target gases on the surface of WO3, consequently strongly influencing the electrical properties of WO3. The findings of this study demonstrate that the internal heterojunction of ZnWO4/WO3 constructed on the surface of nanoplates plays a crucial role in enhancing the gas sensing performance of WO3. In comparison to uniformly distributed heterojunctions and additive heterojunctions, the surface internal heterojunction structure presented in this study, achieved through a combination of surface engineering and band structure design technology, offers innovative ideas and methodologies for fabricating high-performance gas sensors based on metal oxides and advancing the development of novel gas sensing materials.

2. Experimental Section

2.1. Material Synthesis

All reagents were of analytical-reagent grade and used as purchased from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China). without any purification. The preparation procedures for the WO3 nanoplates and the ZnWO4/WO3 nanoplates are demonstrated below.

2.2. Preparation of WO3 Nanoplates

The WO3 nanoplates were prepared using a low-temperature method. The typical process is that 3.3 g Na2WO4·2H2O was dissolved into 50 mL deionized water with stirring. Next, 10 mL hydrochloric acid (36–38%) was added. Then, 0.9 g oxalic acid was added under stirring and aging at 60 °C for 24 h. Finally, the precursor was washed and centrifuged. After drying at 60 °C for 12 h, the precursor was calcined at 450 °C with a rate of 1 °C/min for 2 h, and the WO3 nanoplates were obtained.

2.3. Preparation of Nanoplates with ZnWO4/WO3 Surface Heterojunction

The WO3 nanoplates prepared above were used as a substrate for ZnWO4/WO3. First, 0.46 g WO3 nanoplates (0.002 mol) were impregnated in a 10 mL solution with different concentrations of zinc acetate (0.01 M, 0.1 M, and 0.5 M) with ultrasonic dispersion for 10 min. After that, the suspension solutions were subsequently stirred at room temperature until dried. The obtained samples were annealed at 500 °C with a rate of 1 °C/min for 2 h. The as-prepared samples were labeled as ZnWO4/WO3-0.01, ZnWO4/WO3-0.1, and ZnWO4/WO3-0.5.

2.4. Microscopic Characterization

X-ray diffraction (XRD) patterns of these samples were recorded by a Rigaku D/Max-2500 X-ray diffractometer (Rigaku Corporation, Tokyo, Japan) using Cu-Kα1 as radiation. High-resolution transmission electron microscopy (HR-TEM) analyses and energy dispersive X-ray (EDX) mapping profiles were conducted on an FEI Talos F200X G2 microscope (Thermo Scientific, Waltham, MA, USA) with energy dispersive spectroscope (super-x, Energy resolution < 136 eV (Mn-Ka)) at an accelerating voltage of 200 kV. X-ray photoelectron spectroscopy (XPS) was performed using an ESCA LAB250Xi spectrometer (Thermo Scientific, USA) with an unmonochromated Al Kα X-ray source. The energy resolution of XPS is less than or equal to 0.45 eV. The UV–vis spectra were recorded on a Thermo Scientific Evolution 201 spectrophotometer (Thermo Scientific, USA). The Raman spectra were recorded with a 532 nm wavelength laser using a HORIBA XploRA PLUS.3 (HORIBA, Longjumeau, France).

2.5. Gas Sensing Measurement

The measurement of gas-sensing properties was similar to that of reported work [14,15]. The gas sensing performance was evaluated using the WS-30A test instrument (Hanwei Electronics Co. Ltd., Zhengzhou, China). Target gases of various concentrations were obtained by heating a specific amount of analytically pure liquid chemicals to generate vapor, which was then mixed with clean air. The test chamber we used in this work has a volume of 18 L. According to the ideal gas law, the vapor concentration can be calculated based on the quantity of the liquid sample injected. For example, to achieve the desired concentration of triethylamine or other vapors, precise volumes of solution are injected into the hot plate of the test chamber to ensure complete evaporation. Simultaneously, the fan in the test chamber was activated to ensure an even distribution of the target gas throughout the chamber. During testing, we maintained the relative humidity of the sample chamber at approximately the same level as the laboratory environment, within the range of 35 ± 5%.
After exposing the gas sensor to the sample gas environment for 180 s, we opened the test chamber and used a fan to blow away the sample gas. This restored the test environment to the initial clean air. We then prepared at least 4 sensors of each type and tested each sensor more than 5 times to obtain the corresponding error bars. The response of these sensors is defined as S = Rg/Ra, where Ra is the average resistance of the sensor in the air, and Rg is the average resistance of the sensor in the target gas. The time it takes for the sensor to reach 90% of the complete response and recovery is called the response time (tres) or recovery time (trec). The aging and testing processes were conducted in a separate room. The humidity of the gas mixture was controlled with a saturated inorganic salt solution. Relative humidity of 33%, 43%, 52%, 67%, 75%, and 85% were prepared by using MgCl2, K2CO3, Mg(NO3)2, CuCl2, NaCl, and KCl saturated solutions, respectively.
As investigated in Figure 2, the XRD patterns of the WO3 nanoplates and these different ZnWO4/WO3 samples are clearly displayed. The strong diffraction peaks in the XRD pattern of WO3 nanoplates can be well-indexed to WO3 (JCPDS NO.72-0677). It can be observed from spectra in Figure 2B–D that the intensity of the ZnWO4 (100), (1–11), and (021) crystal planes increase gradually with the increasing concentration of Zn2+. And these peaks indicate precisely the type of monoclinic structure (JCPDS NO.73-0554). For the content of ZnWO4, which is far below the limit of detection of XRD, no prominent peaks of the ZnWO4 phase are observed in the spectrum of ZnWO4/WO3-0.01 (as shown in Figure 2B). It is considered that the doped Zn2+ is too small to form a continuous ZnWO4 phase on the surface of WO3 nanoplates, and Zn2+ modifying into the WO3 lattice will replace the W6+ to create more oxygen vacancies. The results of XPS support this hypothesis. This result demonstrates the possibility of the formation of a phase.
The Raman spectra of the WO3 nanoplates and ZnWO4/WO3 samples are shown in Figure 3 to illustrate the phase information further. It can be found that two intense peaks at 713 and 804 cm−1 can be clearly observed, which are mainly attributed to the monoclinic WO3, which is in good accordance with the XRD [16]. As depicted from the patterns of different ZnWO4/WO3 samples, a distinct peak at 904 cm−1 presents in the patterns of ZnWO4/WO3-0.1 and ZnWO4/WO3-0.5. This peak is ascribed to the stretching vibration of the W-O bonds of ZnWO4 [17]. The Raman results demonstrate that we have successfully prepared the ZnWO4/WO3 composites.
TEM was used to study the surface morphology and structural characteristics of ZnWO4/WO3, as shown in Figure 4. It can be observed that WO3 forms a nanoplate structure with an average size of about 150 × 80 nm, and the size distribution of the WO3 nanoplates is uniform. The substrate WO3 nanoplates’ structure is well preserved after the in situ construction of the ZnWO4/WO3 composites. It precisely shows that the ZnWO4/WO3 composites retain the uniform nanoplates’ morphology of the substrate, and the surface of the ZnWO4/WO3 composites is as smooth as the WO3 nanoplates. This result indicates that these ZnWO4/WO3 heterojunction composites are not composed of uniform separated ZnWO4 and WO3 phases, but a new phase ZnWO4 region has been formed on the surface of the WO3 nanoplates through the Zn2+ ions that have penetrated the WO3 crystal lattice.
Furthermore, no new structures are found in these composites, which indicates that the ZnWO4 phase may be embedded into the surface of WO3 nanoplates. Hence, the improvement in the gas sensing performance of these composites is mainly brought by the change in the surface structure of the WO3 nanoplates. The lattice fringes can be observed in Figure 4B,E,H. The spacing of 0.365 nm can be indexed to the (200) plane of monoclinic WO3, while the other spacing of 0.292 nm matches the (1–11) plane of ZnWO4. It should be noted that no apparent lattice can be found in ZnWO4/WO3-0.01. This is probably attributed to the size of the ZnWO4 phase being tiny. The high-angle annular dark field (HAADF) elemental mapping images (Figure 4C,F,I) can be applied to verify the existence of ZnWO4 phase on the surface of the WO3 nanoplates. It is clearly investigated that the Zn elements are evenly dispersed on the surface of the WO3 nanoplates, and the dispersed concentration of Zn elements will increase with the increase in the impregnating solution concentration of Zn2+. As shown in the EDS mapping results of ZnWO4/WO3-0.01 (Figure 4C), it can be observed that a small amount of the Zn element is scattered on the surface of the WO3 nanoplates. This result indicates that some Zn2+ penetrates the lattice of WO3 to form ZnWO4, and the amount of ZnWO4 distributed as dots on the surface of WO3 nanoplates is too low to be detected by XRD and Raman spectroscopy. As shown in Figure 4F,I, the ZnWO4 has a continuous phase on the surface of WO3 nanoplates, which looks like an “island”.
In order to investigate the definite elemental constituent and chemical states of the ZnWO4/WO3 internal heterojunction, XPS analyses were adopted in this work. The XPS data were calibrated using the 284.4 eV peak of the C element as the standard, employing the Tougaard method to conduct background correction on all spectra, and then utilizing the Lorentz–Gaussian function for peak fitting to determine the corresponding fitted peaks. Figure 5 investigates the high-resolution XPS spectra of Zn 2p, W 4f, and O 1s. The high-resolution Zn 2p spectra of all the ZnWO4/WO3 nanoplates exhibited two distinct spin-orbit doublet peaks located at about 1045 and 1022 eV, which verified that the Zn element holds an oxidation state of +2 in the composites [9,18]. The W 4f spectra can be observed in Figure 5B,E,H,K. Three peaks were clearly observed after fitting with the Gaussian–Lorentzian functions in all samples. The two characteristic peaks of W 4f5/2 and W 4f7/2 at binding energies of 37.7 eV and 35.5 eV are typical for W6+ sites coordinated by O2− [19,20,21]. The O 1s spectra of all samples are shown in Figure 5C,F,I,L. The O 1s region of WO3 and ZnWO4/WO3 nanoplates can be fitted by two peaks corresponding to the lattice oxygen and oxygen vacancy [9,20]. These results indicate that the chemical states of the W and O elements are the same in all samples, and the chemical states of the Zn element are not significantly different in the three different ZnWO4/WO3 nanoplates. All these features indicate that the samples are ZnWO4 and WO3, as reported in previous work [22,23].
The results of the XRD, Raman spectrum, EDS mapping, and XPS characterization indicate that Zn2+ has been successfully incorporated into the WO3 surface lattice, forming a new ZnWO4 phase in the microregion of the WO3 surface. Unlike traditional methods of metal oxide doping, our approach involves immersing WO3 nanosheets in a solution containing a specific concentration of Zn2+, thoroughly stirring to adsorb Zn2+ onto the surface of the WO3 nanosheets, and then treating them at high temperature to allow Zn2+ to enter the WO3 lattice. As a result, the doped Zn2+ gradually penetrates the lattice from the surface of the WO3 nanosheets under high-temperature conditions, becoming concentrated in the surface lattice area. Depending on the concentration of doped Zn2+, this process can lead to a continuous or discontinuous new ZnWO4 phase on the surface of the WO3 nanoflakes. This novel impregnation method results in the preparation of a ZnWO4/WO3 composite material with a unique surface heterojunction structure, which was added prior to Section 4 on gas sensing properties.

3. Gas Sensing Properties

The sensors prepared with different samples were measured by their gas sensing performance to other volatile organic compounds (VOCs) at various concentrations (0.2–10 ppm) and operating temperatures (100–400 °C). The gas sensing performance of the WO3 and ZnWO4/WO3 nanoplates with varying concentrations of Zn ions are shown in Figure 6 and Figure 7. The gas sensors based on these nanoplates present the typical n-type semiconductor behavior toward TEA under an indoor humidity of about 35 ± 5%. Figure 6A exhibits the temperature-dependent response of all the gas sensors to 10 ppm TEA, with the working temperature from room temperature (RT) to 400 °C. It was found that the response of the sensors based on the different WO3 and ZnWO4/WO3 nanoplates increased with the increasing temperature, from RT to 250 °C. However, the response of all the sensors gradually decreased with further increases in the operating temperature. From then on, 250 °C was selected as the optimum working temperature for the gas sensing performance test. As shown in Figure 6C, the response of WO3 sensors is about 13.41 to 10 ppm TEA at 250 °C, and the response/recovery time is about 17/72 s. The response of the ZnWO4/WO3 sensors has been obviously improved, reaching about 44.86 (ZnWO4/WO3-0.01), 38.72 (ZnWO4/WO3-0.1), and 40.75 (ZnWO4/WO3-0.5), and the speed of response was also enhanced by the construction of surface heterojunction. The response time shortened from 17 s to 10 s as the concentration of Zn2+ content increased. Furthermore, it can also be found that the response of ZnWO4/WO3 nanoplates to TEA is improved at 250 °C. The selectivity of these different kinds of sensors to 10 ppm of various VOC gases at the optimum temperature of 250 °C is presented in Figure 6B. The ZnWO4/WO3 sensors exhibit good selectivity to TEA at optimum temperature. Significantly, the response of ZnWO4/WO3-0.5 sensors reaches about 40.75 towards TEA, while it only reaches 1.4–3.5 towards other VOC gases.
To investigate the repeatability and sensitivity performance of the WO3 and ZnWO4/WO3 nanoplates to TEA, the response–recovery curves of these four kinds of sensors’ exposure to various concentrations of TEA (0.2, 0.5, 1, 2, 5, and 10 ppm) at the optimum temperature are illustrated in Figure 7A. It can be observed that each concentration point has been tested twice, and the responses of these sensors increased with the increase in the TEA concentration. This result indicates that these sensors have good repeatability among individual alternating tests. The response of the ZnWO4/WO3-0.5 sensors gradually increases from 3.62 at 200 ppb to 40.75 at 10 ppm. Even as the concentration of TEA is 200 ppb, the response can reach about 3.62, which is two times higher than that of WO3, and the response–recovery time is 24/47 s. Although the response of ZnWO4/WO3-0.01 and ZnWO4/WO3-0.1 is very good, the recovery time of ZnWO4/WO3-0.01 and ZnWO4/WO3-0.1 increases more than 100 s. For verifying the effect of the surface heterojunction on the humidity immunity, the humidity-dependent testing of the WO3 and ZnWO4/WO3 sensors to 10 ppm TEA was measured under different RH conditions (Figure 7B). It can be found that the response values of the pure WO3 and ZnWO4/WO3-0.01 nanoplates dramatically decrease with the increasing RH. This is mainly due to the lesser amount of doped Zn2+ on the WO3 surface, and the formation of a new phase is not enough to offset the effect of water molecules. With the increase in the content of the ZnWO4 phase disturbed on the surface, the humidity immunity of the composites is significantly improved. The results revealed that at a relative humidity of 85%, the response of the sensor based on ZnWO4/WO3-0.5 was only 16% lower than that of the device at a relative humidity of 25%, indicating nice humidity immunity of the sensors based on ZnWO4/WO3-0.5. The influence of moisture on the gas sensing performance of metal oxides is mainly because the water molecules can further react with the negative oxygen ions formed on the surface of metal oxides to produce hydroxyl and directly change the resistance of the sensors. However, ZnWO4, as previous works have reported [23,24], presented a strong adsorption and decomposition ability to water molecules, which mainly depends on the generated reaction of the holes with the molecular water. As in the previous reports, the WO3 will transfer holes to the valence band of ZnWO4 in the ZnWO4/WO3 heterojunctions [21,25]. The ZnWO4 phase disturbed on the surface of the WO3 nanoplates will preferentially adsorb and consume a lot of water molecules. Meanwhile, the electrons of the WO3 surface were increased to form more negative oxygen ions. The above results offset the influence of humidity on the gas sensing performance of composites. In addition, Figure 7C illustrates the responses R of these sensors exhibiting a Freundlich isotherm relationship with TEA concentration C. After fitting with the Freundlich equation as follows:
R = K C n
the relevant parameters are presented in Table 1. The detection limit of the ZnWO4/WO3-0.5 sensors was estimated to be 1.40 toward 31 ppb of TEA as the S/N = 3. The long-term stability and reproducibility have also been characterized to evaluate the Zn2+ doped WO3 nanoplates. As shown in Figure 7D, the response/recovery curves of these ZnWO4/WO3 composites toward 10 ppm TEA at 250 °C revealed that the long-term stability and reproducibility of these ZnWO4/WO3 sensors are very good. The response value deviation of each sensor based on ZnWO4/WO3 is less than 2.2, indicating a very low drift deviation of the device. The fundamental resistance drift of each sensor at 250 °C remains below 5%, indicating its reliability in applications. This good stability and reproducibility are mainly ascribed to the composites having uniform nanoplates structural without any hierarchical nanostructure, and the WO3 and ZnWO4 have excellent thermal and chemical stability. As shown in Table 2, the obtained results demonstrated that the sensor based on ZnWO4/WO3 heterojunctions is very promising for practical application.

4. Gas Sensing Mechanism

As described above, the ZnWO4 phase disturbed on the surface of the WO3 nanoplates can adsorb a large number of O2 and H2O to activate them and produce active radicals that participate in the gas sensing reaction (as shown in Figure 8A) [23,38]. Furthermore, more oxygen vacancies have been created during the impregnation process. Compared to ZnWO4/WO3-0.1 and ZnWO4/WO3-0.5, the gas sensing performance of ZnWO4/WO3-0.01 is mainly improved by a higher concentration of oxygen vacancy. This is probably because of the small amount of Zn2+, which cannot construct a continuous ZnWO4 phase after infiltration into the WO3 lattice, resulting in the local substitution of W6+ for producing more oxygen vacancies on the surface. Therefore, the humidity immunity of ZnWO4/WO3-0.01 is still poor. Apart from these favorable factors for enhancing the gas sensing performance of these composites, the internal heterojunctions formed on the surface of the WO3 nanoplates can efficiently separate the electron hole in the ZnWO4/WO3, resulting in the increase in electrons in the conduction band of WO3 to generate more negative oxygen ions, and the valence band of ZnWO4 will then attract more holes from WO3, which further forms active radicals (as shown in Figure 8B). This process can further improve the gas sensing performance of WO3 nanoplates toward TEA. Obviously, the gas sensing performance of these internal heterojunction composites did not increase geometrically with the increase in ZnWO4 content on the surface of the WO3 nanoplates. The first reason is the homogeneity of the ZnWO4/WO3 internal heterojunctions prepared by the impregnation method. It is impossible to guarantee that the ZnWO4 phase content and distribution on the surface of each WO3 nanoplate will be the same (as shown in Figure 8C). Secondly, the more the ZnWO4 phase is distributed on the surface, the smaller the area of WO3 which will participate in the gas sensing reaction, which makes the response of ZnWO4/WO3-0.5 to TEA unable to be further improved. Therefore, the rational design of the content and distribution of ZnWO4 on the surface of the WO3 nanoplates is essential for enhancing gas sensing performance. In this work, the ZnWO4/WO3-0.1 nanoplates, which demonstrate good response, selectivity, and long-term stability gas sensing performance towards TEA and fine humidity immunity, are sufficient to be applied as a gas sensing material for TEA testing in practical application. Excessive doping of Zn2+ on the surface of the WO3 nanosheets cannot further improve the gas sensing performance of TEA but will cause the unnecessary waste of the zinc element.

5. Conclusions

In this work, we prepared various surface-intra-heterojunction nanostructured ZnWO4/WO3 composites using a controlled impregnation method. We found that using a double impregnation method effectively created heterojunction structures on the surface of WO3 nanoplates. These surface heterojunction composites significantly improved gas sensing performance, particularly for TEA, through surface engineering. The ZnWO4/WO3-0.5 composite showed a nearly three times increased response and improved selectivity. It exhibited a gas sensing response of 40.75 at 250 °C to 10 ppm TEA, with a response recovery time of 10/89 s, faster than WO3 nanoplates. The sensor based on ZnWO4/WO3-0.5 achieved a minimum detection concentration of 200 ppb, with a response of about 3.62, which was higher than the response to other VOC gases (about 1.4~3.5). The theoretical LOD was approximately 31 ppb. The sensor based on ZnWO4/WO3-0.5 exhibited ample long-term stability. Even after 12 months of placement, the response of the sensor decreased by only 2.2. Additionally, this sensor demonstrated good humidity resistance, with the sensor response value decreasing by less than 16%, even at a relative humidity of up to 85%. These results demonstrated that sensors based on ZnWO4/WO3-0.5 present good stability and a low failure rate. Our results demonstrate the effectiveness of surface engineering in designing and optimizing the surface structures of metal oxides, especially for enhancing gas sensing performance. Our research provides a promising strategy for improving the gas sensing performance of oxide materials.

Author Contributions

Conceptualization, K.T. and Y.S.; methodology, K.Y. and X.R.; formal analysis, X.R. and K.Y.; investigation, K.T., X.R., P.X. and Y.M.; data curation, M.L., M.S. and J.W.; writing—original draft preparation, K.T.; writing—review and editing, P.X., K.T., X.R. and K.Y.; supervision, P.X. and Y.S.; funding acquisition, K.T.; The manuscript was written through contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the [National Natural Science Foundation of China–Henan United Fund] [Grant No. U20041102], [Henan Province key R&D and promotion special application] [Grant No. 222102210082], [National Natural Science Foundation of China Fund and State Key Laboratory of Materials Processing and Die and Mold Technology, Huazhong University of Science and Technology].

Data Availability Statement

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

Acknowledgments

National Natural Science Foundation of China Fund and State Key Laboratory of Materials Processing and Die and Mold Technology, Huazhong University of Science and Technology. The Analytical and Testing Center of Zhengzhou University of Light Industry is acknowledged for the XRD, SEM, and XPS. The authors thank the Shiyanjia Lab for supporting the HRTEM and EDX mapping test.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Guo, M.; Luo, N.; Chen, Y.; Fan, Y.; Wang, X.; Xu, J. Fast-response MEMS xylene gas sensor based on CuO/WO3 hierarchical structure. J. Hazard. Mater. 2022, 429, 127471. [Google Scholar] [CrossRef] [PubMed]
  2. Staerz, A.; Liu, Y.; Geyik, U.; Brinkmann, H.; Weimar, U.; Zhang, T.; Barsan, N. The effect of platinum loading on WO3 based sensors. Sens. Actuators B Chem. 2019, 291, 378–384. [Google Scholar] [CrossRef]
  3. Korotcenkov, G. Gas response control through structural and chemical modification of metal oxide films: State of the art and approaches. Sens. Actuators B Chem. 2005, 107, 209–232. [Google Scholar] [CrossRef]
  4. Liang, Y.; Yang, Y.; Zou, C.; Xu, K.; Luo, X.; Luo, T.; Li, J.; Yang, Q.; Shi, P.; Yuan, C. 2D ultra-thin WO3 nanoplates with dominant {002} crystal facets for high-performance xylene sensing and methyl orange photocatalytic degradation. J. Alloys Compd. 2019, 783, 848–854. [Google Scholar] [CrossRef]
  5. Wang, X.; Han, W.; Yang, J.; Jiang, B.; Cheng, P.; Wang, Y.; Sun, P.; Zhang, H.; Sun, Y.; Lu, G. Facet effect on diverse WO3 morphologies and ideal work function for ppb-level triethylamine detection. Sens. Actuators B Chem. 2022, 363, 131849. [Google Scholar] [CrossRef]
  6. Gui, Y.; Yang, L.; Tian, K.; Zhang, H.; Fang, S. P-type Co3O4 nanoarrays decorated on the surface of n-type flower-like WO3 nanoplates for high-performance gas sensing. Sens. Actuators B Chem. 2019, 288, 104–112. [Google Scholar] [CrossRef]
  7. Dong, C.; Zhao, R.; Yao, L.; Ran, Y.; Zhang, X.; Wang, Y. A review on WO3 based gas sensors: Morphology control and enhanced sensing properties. J. Alloys Compd. 2020, 820, 153194. [Google Scholar] [CrossRef]
  8. Degler, D.; Pereira de Carvalho, H.W.; Kvashnina, K.; Grunwaldt, J.-D.; Weimar, U.; Barsan, N. Structure and chemistry of surface-doped Pt:SnO2 gas sensing materials. RSC Adv. 2016, 6, 28149–28155. [Google Scholar] [CrossRef]
  9. Cui, Y.; Pan, L.; Chen, Y.; Afzal, N.; Ullah, S.; Liu, D.; Wang, L.; Zhang, X.; Zou, J.-J. Defected ZnWO4-decorated WO3 nanorod arrays for efficient photoelectrochemical water splitting. RSC Adv. 2019, 9, 5492–5500. [Google Scholar] [CrossRef]
  10. Li, W.; Cao, L.; Kong, X.; Huang, J.; Yao, C.; Fei, J.; Li, J. In situ synthesis and photocatalytic performance of WO3/ZnWO4 composite powders. RSC Adv. 2016, 6, 23783–23789. [Google Scholar] [CrossRef]
  11. Sheng, Z.; Ma, D.; He, Q.; Wu, K.; Yang, L. Mechanism of photocatalytic toluene oxidation with ZnWO4: A combined experimental and theoretical investigation. Catal. Sci. Technol. 2019, 9, 5692–5697. [Google Scholar] [CrossRef]
  12. Li, W.; Xu, H.; Zhai, T.; Yu, H.; Chen, Z.; Qiu, Z.; Song, X.; Wang, J.; Cao, B. Enhanced triethylamine sensing properties by designing Au@SnO2/MoS2 nanostructure directly on alumina tubes. Sens. Actuators B Chem. 2017, 253, 97–107. [Google Scholar] [CrossRef]
  13. Pan, H.; Wang, C.; Zhang, Z.; Li, Y.; Hou, X.; Zheng, W.; Liu, X.; Wan, Y.; Zhang, J. Oxygen vacancy-enriched ALD NiO sub-50 nm thin films for enhanced triethylamine detection. Appl. Phys. Lett. 2022, 121, 111603. [Google Scholar] [CrossRef]
  14. Tian, K.; Yang, X.-Y.; Ren, X.-N.; Huang, Y.-X.; Li, H.-Y.; Gui, Y. Metal–Organic Framework-Derived Hierarchical Nanostructured Ni0.7Zn0.3O Hollow Porous Microspheres for Selective and Sensitive aniline Sensing. ACS Appl. Nano Mater. 2023, 6, 5508–5516. [Google Scholar] [CrossRef]
  15. Tian, K.; Zhang, W.; Sun, S.-N.; Xing, L.; Li, Z.-Y.; Zhang, T.-T.; Yang, N.-N.; Kuang, B.-B.; Li, H.-Y. Design and fabrication of spinel nanocomposites derived from perovskite hydroxides as gas sensing layer for volatile organic compounds detection. Sens. Actuators B Chem. 2021, 329, 129076. [Google Scholar] [CrossRef]
  16. Guillén, C.; Herrero, J. Amorphous WO3-x thin films with color characteristics tuned by the oxygen vacancies created during reactive DC sputtering. J. Mater. Sci. Technol. 2021, 78, 223–228. [Google Scholar] [CrossRef]
  17. Bai, X.; Wang, L.; Zhu, Y. Visible Photocatalytic Activity Enhancement of ZnWO4 by Graphene Hybridization. ACS Catal. 2012, 2, 2769–2778. [Google Scholar] [CrossRef]
  18. Jung, J.; Jeong, J.R.; Dang Van, C.; Yoo, H.Y.; Lee, M.H. Morphology-Controlled ZnO@ZnWO4 Hetero-Nanostructures for Efficient Photooxidation of Water in Near-Neutral pH. ACS Appl. Mater. Interfaces 2024, 16, 4700–4707. [Google Scholar] [CrossRef]
  19. Sun, D.; Wang, Q.; Wang, W.; Chen, Y.; Ruan, S. Hollow ZnWO4/In2O3 Nanotubes for Ultrasensitive and Rapid Trace Detection of Triethylamine. ACS Appl. Nano Mater. 2023, 6, 10581–10589. [Google Scholar] [CrossRef]
  20. Bai, S.; Yin, P.; Zhao, Y.; Tang, P.; Luo, R.; Li, D.; Chen, A. rGO-Modified ZnWO4/WO3 Nanocomposite for Detection of NH3. J. Phys. Chem. C 2023, 127, 9315–9326. [Google Scholar] [CrossRef]
  21. Kumar, G.M.; Lee, D.J.; Jeon, H.C.; Ilanchezhiyan, P.; Young, K.D.; Won, K.T. One dimensional ZnWO4 nanorods coupled with WO3 nanoplates heterojunction composite for efficient photocatalytic and photoelectrochemical activity. Ceram. Int. 2022, 48, 4332–4340. [Google Scholar] [CrossRef]
  22. Wei, L.; Zhang, H.; Cao, J. Electrospinning of Ag/ZnWO4/WO3 composite nanofibers with high visible light photocatalytic activity. Mater. Lett. 2019, 236, 171–174. [Google Scholar] [CrossRef]
  23. Huang, Y.; Gao, Y.; Zhang, Q.; Cao, J.-J.; Huang, R.-J.; Ho, W.; Lee, S.C. Hierarchical porous ZnWO4 microspheres synthesized by ultrasonic spray pyrolysis: Characterization, mechanistic and photocatalytic NOx removal studies. Appl. Catal. A: Gen. 2016, 515, 170–178. [Google Scholar] [CrossRef]
  24. Leonard, K.C.; Nam, K.M.; Lee, H.C.; Kang, S.H.; Park, H.S.; Bard, A.J. ZnWO4/WO3 Composite for Improving Photoelectrochemical Water Oxidation. J. Phys. Chem. C 2013, 117, 15901–15910. [Google Scholar] [CrossRef]
  25. Li, H.; Chu, S.; Ma, Q.; Wang, J.; Che, Q.; Wang, G.; Yang, P. Hierarchical WO3/ZnWO4 1D fibrous heterostructures with tunable in-situ growth of WO3 nanoparticles on surface for efficient low concentration HCHO detection. Sens. Actuators B Chem. 2019, 286, 564–574. [Google Scholar]
  26. Zhang, M.; Zhao, Z.; Hui, B.; Sun, J.; Sun, J.; Tian, W.; Zhang, Z.; Zhang, K.; Xia, Y. Carbonized polymer dots activated hierarchical tungsten oxide for efficient and stable triethylamine sensor. J. Hazard. Mater. 2021, 416, 126161. [Google Scholar] [CrossRef]
  27. Bai, S.; Zuo, Y.; Zhang, K.; Zhao, Y.; Luo, R.; Li, D.; Chen, A. WO3-ZnFe2O4 heterojunction and rGO decoration synergistically improve the sensing performance of triethylamine. Sens. Actuators B Chem. 2021, 347, 130619. [Google Scholar] [CrossRef]
  28. Han, Y.; Liu, Y.; Su, C.; Chen, X.; Zeng, M.; Hu, N.; Su, Y.; Zhou, Z.; Wei, H.; Yang, Z. Sonochemical synthesis of hierarchical WO3 flower-like spheres for highly efficient triethylamine detection. Sens. Actuators B Chem. 2020, 306, 127536. [Google Scholar] [CrossRef]
  29. Wang, X.; Han, W.; Yang, J.; Cheng, P.; Wang, Y.; Feng, C.; Wang, C.; Zhang, H.; Sun, Y.; Lu, G. Conductometric ppb-level triethylamine sensor based on macroporous WO3-W18O49 heterostructures functionalized with carbon layers and PdO nanoparticles. Sens. Actuators B Chem. 2022, 361, 131707. [Google Scholar] [CrossRef]
  30. Tian, W.; Wang, Y.; Zhang, Y.; Cao, J.; Guan, R.-F. WO3 Nanoflakes Coupled with Hexagonal Boron Nitride Nanoplates for Triethylamine Sensing. ACS Appl. Nano Mater. 2021, 4, 6316–6327. [Google Scholar] [CrossRef]
  31. Hu, Q.; He, J.; Chang, J.; Gao, J.; Huang, J.; Feng, L. Needle-Shaped WO3 Nanorods for Triethylamine Gas Sensing. ACS Appl. Nano Mater. 2020, 3, 9046–9054. [Google Scholar] [CrossRef]
  32. Zhai, C.; Zhu, M.; Jiang, L.; Yang, T.; Zhao, Q.; Luo, Y.; Zhang, M. Fast triethylamine gas sensing response properties of nanoplates assembled WO3 hollow microspheres. Appl. Sur. Sci. 2019, 463, 1078–1084. [Google Scholar] [CrossRef]
  33. Zhang, M.; Liu, K.; Zhang, X.; Wang, B.; Xu, X.; Du, X.; Yang, C.; Zhang, K. Interfacial energy barrier tuning of hierarchical Bi2O3/WO3 heterojunctions for advanced triethylamine sensor. J. Adv. Ceram. 2022, 11, 1860–1872. [Google Scholar] [CrossRef]
  34. Zeb, S.; Yang, Z.; Hu, R.; Umair, M.; Naz, S.; Cui, Y.; Qin, C.; Liu, T.; Jiang, X. Electronic structure and oxygen vacancy tuning of Co&Ni co-doped W18O49 nanourchins for efficient TEA gas sensing. Chem. Eng. J. 2023, 465, 142815–142851. [Google Scholar]
  35. Dong, J.; Liu, H.; Sun, C.; Shao, J.; Wang, M.; He, P.; Qi, Y.; Pan, G.; Yang, X. Low-temperature and high-efficient detection of triethylamine based on Pt/PtO2 loaded WO3 gas sensors. J. Alloy. Compd. 2023, 966, 171642. [Google Scholar] [CrossRef]
  36. Zhao, W.; Zhao, Q.; Hao, H.; Lin, H.; Gao, W.; Yan, S. Hydrophobic modification of WO3/WS2 heterostructure for construction of humidity immune TEA sensors. Mater. Chem. Phys. 2023, 294, 126993. [Google Scholar] [CrossRef]
  37. Zhang, H.; Liu, B.; Lv, L.; Shao, J.; Du, Y.; Li, Y.; Chang, W. Enhanced triethylamine sensing characteristics of In-doped WO3 cubic nanoblocks at low operating temperature. Vacuum 2023, 218, 112640. [Google Scholar] [CrossRef]
  38. Li, J.; Guo, C.; Li, L.; Gu, Y.; BoK-Hee, K.; Huang, J. In Situ Fabrication of 1D WO3 Nanorod/2D ZnWO4 Nanosheet Heterojunction for Enhanced Photoelectrochemical Performance. Catal. Lett. 2022, 152, 1611–1620. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram illustrating the in-situ construction of a heterojunction on WO3 nanoplates and the preparation of gas sensors.
Figure 1. Schematic diagram illustrating the in-situ construction of a heterojunction on WO3 nanoplates and the preparation of gas sensors.
Sensors 24 05606 g001
Figure 2. XRD patterns of (A) WO3, (B) ZnWO4/WO3-0.01, (C) ZnWO4/3-0.1, and (D) ZnWO4/WO3-0.5.
Figure 2. XRD patterns of (A) WO3, (B) ZnWO4/WO3-0.01, (C) ZnWO4/3-0.1, and (D) ZnWO4/WO3-0.5.
Sensors 24 05606 g002
Figure 3. Raman patterns of the prepared four samples.
Figure 3. Raman patterns of the prepared four samples.
Sensors 24 05606 g003
Figure 4. TEM images and EDS mapping patterns of the prepared ZnWO4/WO3 nanoplates. (A) low-magnification image of the ZnWO4/WO3-0.01 sample. The scale bar is 20 nm; (B) the high-magnification image of the ZnWO4/WO3-0.01 sample; (C) EDS mapping patterns of the ZnWO4/WO3-0.01 sample; (D) low-magnification image of the ZnWO4/WO3-0.1 sample. The scale bar is 20 nm; (E) high-magnification image of the ZnWO4/WO3-0.1 sample; (F) EDS elements mappings of the ZnWO4/WO3-0.1 sample, (G) low-magnification image of the ZnWO4/WO3-0.5 sample. The scale bar is 20 nm; (H) high-magnification image of the ZnWO4/3-0.5 sample; (I) EDS mapping results of the ZnWO4/WO3-0.5 sample.
Figure 4. TEM images and EDS mapping patterns of the prepared ZnWO4/WO3 nanoplates. (A) low-magnification image of the ZnWO4/WO3-0.01 sample. The scale bar is 20 nm; (B) the high-magnification image of the ZnWO4/WO3-0.01 sample; (C) EDS mapping patterns of the ZnWO4/WO3-0.01 sample; (D) low-magnification image of the ZnWO4/WO3-0.1 sample. The scale bar is 20 nm; (E) high-magnification image of the ZnWO4/WO3-0.1 sample; (F) EDS elements mappings of the ZnWO4/WO3-0.1 sample, (G) low-magnification image of the ZnWO4/WO3-0.5 sample. The scale bar is 20 nm; (H) high-magnification image of the ZnWO4/3-0.5 sample; (I) EDS mapping results of the ZnWO4/WO3-0.5 sample.
Sensors 24 05606 g004
Figure 5. XPS spectra of the samples: (AC) Zn 2p, W 4f and O 1s level of WO3 nanoplates, (DF) Zn 2p, W 4f and O 1s level of ZnWO4/WO3-0.01 nanoplates, (GI) Zn 2p, W 4f and O 1s level of ZnWO4/WO3-0.1 nanoplates, (JL) Zn 2p, W 4f, and O 1s level of ZnWO4/WO3-0.5 nanoplates.
Figure 5. XPS spectra of the samples: (AC) Zn 2p, W 4f and O 1s level of WO3 nanoplates, (DF) Zn 2p, W 4f and O 1s level of ZnWO4/WO3-0.01 nanoplates, (GI) Zn 2p, W 4f and O 1s level of ZnWO4/WO3-0.1 nanoplates, (JL) Zn 2p, W 4f, and O 1s level of ZnWO4/WO3-0.5 nanoplates.
Sensors 24 05606 g005
Figure 6. (A) Response of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm TEA at different operating temperatures, (B) Selectivity of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm of various gases at 250 °C, (C) Response–recovery curves of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm TEA at 250 °C, (D) the relationship between the resistance value (Ra) of the four sensors in air and the operating temperature.
Figure 6. (A) Response of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm TEA at different operating temperatures, (B) Selectivity of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm of various gases at 250 °C, (C) Response–recovery curves of WO3 nanoplates, ZnWO4/WO3-0.01 nanoplates, ZnWO4/WO3-0.1 nanoplates and ZnWO4/WO3-0.5 nanoplates sensors to 10 ppm TEA at 250 °C, (D) the relationship between the resistance value (Ra) of the four sensors in air and the operating temperature.
Sensors 24 05606 g006
Figure 7. (A) Response–recovery curves of WO3 and ZnWO4/WO3 sensors to different concentrations (0.2, 0.5, 1, 2, 5 and 10 ppm) of TEA at 250 °C and the RH = 35% ± 5% (this humidity here is the ambient humidity at room temperature), (B) Humidity stability of different ZnWO4/WO3 sensors at 250 °C to 10 ppm TEA (the relative humidity were prepared by different saturated solution at 25 °C), (C) The Freundlich fitting of the sensing response to TEA concentration, (D) Long-term stability of ZnWO4/WO3 sensors 10 ppm TEA at 250 °C and the RH = 35% ± 5% (this humidity here is the ambient humidity at room temperature).
Figure 7. (A) Response–recovery curves of WO3 and ZnWO4/WO3 sensors to different concentrations (0.2, 0.5, 1, 2, 5 and 10 ppm) of TEA at 250 °C and the RH = 35% ± 5% (this humidity here is the ambient humidity at room temperature), (B) Humidity stability of different ZnWO4/WO3 sensors at 250 °C to 10 ppm TEA (the relative humidity were prepared by different saturated solution at 25 °C), (C) The Freundlich fitting of the sensing response to TEA concentration, (D) Long-term stability of ZnWO4/WO3 sensors 10 ppm TEA at 250 °C and the RH = 35% ± 5% (this humidity here is the ambient humidity at room temperature).
Sensors 24 05606 g007
Figure 8. (A,B) Schematic diagram showing the sensing mechanism; (C) Schematic of ZnWO4/WO3 internal heterojunction nanoplates.
Figure 8. (A,B) Schematic diagram showing the sensing mechanism; (C) Schematic of ZnWO4/WO3 internal heterojunction nanoplates.
Sensors 24 05606 g008
Table 1. Parameters were obtained by fitting using the Freundlich equation.
Table 1. Parameters were obtained by fitting using the Freundlich equation.
SensorR2Kn
WO30.99983.8300.543
ZnWO4/WO3-0.010.99978.4910.761
ZnWO4/WO3-0.10.99997.9570.693
ZnWO4/WO3-0.50.999910.6510.586
Table 2. Comparison of the gas sensing performance of various WO3-based sensors toward TEA.
Table 2. Comparison of the gas sensing performance of various WO3-based sensors toward TEA.
MaterialOperation ConditionConcentration (ppm)Responsetres/trec (s)LOD (ppb)
Hierarchical CPDs/WO3 [26]140 °C4014.6267/34046
0.8% rGO-WO3-ZnFe2O4 [27]130 °C1026.9251/14420
Hierarchical WO3 Flower-like Spheres [28]205 °C1011.63/5583
Micro-flower WO3 [5]325 °C12.23/550
Carbon Modified WO3-W18O49 with PdO [29]325 °C10035.71/250
WO3/h-BN-5wt% [30]260 °C500390.68/6041
Needle-Shaped WO3 [31]250 °C16.445/781000
WO3 Hollow Microspheres [32]220 °C60161.5/2250,000
Hierarchical Bi2O3/WO3 [33]140 °C509.289/16232
Co&Ni Co-Doped W18O49 Nanourchins [34]250 °C5015416/131000
Layered Pt/PtO2-WO3 [35]137.5 °C503323.516/262100
WO3/WS2 [36]240 °C5021.81−/−3000
In-doped WO3 cubic nanoblocks [37]115 °C5011.211/401000
ZnWO4/WO3-0.01 (This work)250 °C1044.8616/12293
0.22.4132/188
ZnWO4/WO3-0.1 (This work)250 °C1038.7213/11381
0.22.8926/178
ZnWO4/WO3-0.5 (This work)250 °C1040.7510/8931
0.23.6224/47
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

Tian, K.; Yang, K.; Ren, X.; Miao, Y.; Liu, M.; Su, M.; Wu, J.; Sun, Y.; Xu, P. Improving Triethylamine-Sensing Performance of WO3 Nanoplates through In Situ Heterojunction Construction. Sensors 2024, 24, 5606. https://doi.org/10.3390/s24175606

AMA Style

Tian K, Yang K, Ren X, Miao Y, Liu M, Su M, Wu J, Sun Y, Xu P. Improving Triethylamine-Sensing Performance of WO3 Nanoplates through In Situ Heterojunction Construction. Sensors. 2024; 24(17):5606. https://doi.org/10.3390/s24175606

Chicago/Turabian Style

Tian, Kuan, Kai Yang, Xuening Ren, Yuxin Miao, Mengyao Liu, Mingxing Su, Jiawen Wu, Yu’an Sun, and Pengcheng Xu. 2024. "Improving Triethylamine-Sensing Performance of WO3 Nanoplates through In Situ Heterojunction Construction" Sensors 24, no. 17: 5606. https://doi.org/10.3390/s24175606

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