Next Article in Journal
Synthesis of Metal Matrix Composites Based on CrxNiy-TiN for Additive Technology
Previous Article in Journal
Study on the Compressive Stress Retention in Quenched Cam of 100Cr6 Steel Based on Coupled Thermomechanical and Metallurgical Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement in NO2 Gas Sensing Properties of Semiconductor-Type Sensors by Loading Pt into BiVO4 Nanocomposites at Room Temperature

1
Department of Center for General Education, St. Mary’s Junior College of Medicine, Nursing and Management, Yilan City 26647, Taiwan
2
Department of Applied Chemistry, Providence University, Taichung City 43301, Taiwan
3
Office of the Dean (Research) & Division of Chemistry, Department of Sciences, Faculty of Sciences & Technology, Alliance University, Karnataka, Bengaluru 562106, India
4
NTRC-MCETRC and 109 Composite Technologies Pvt. Ltd., Andhra Pradesh, Guntur District, Guntur 522201, India
*
Author to whom correspondence should be addressed.
Materials 2021, 14(20), 5913; https://doi.org/10.3390/ma14205913
Submission received: 21 August 2021 / Revised: 2 October 2021 / Accepted: 4 October 2021 / Published: 9 October 2021

Abstract

:
In the present study, we report the first attempt to prepare a conducive environment for Pt/BiVO4 nanocomposite material reusability for the promotion of sustainable development. Here, the Pt/BiVO4 nanocomposite was prepared using a hydrothermal method with various weight percentages of platinum for use in NO2 gas sensors. The surface morphologies and structure of the Pt/BiVO4 nanocomposite were characterized by scanning electron microscope (SEM), transmission electron microscopy (TEM), energy-dispersive X-ray spectroscopy (EDX), and X-ray diffraction (XRD). The results showed that Pt added to BiVO4 with 3 wt.% Pt/BiVO4 was best at a concentration of 100 ppm NO2, with a response at 167.7, and a response/recovery time of 12/35 s, respectively. The Pt/BiVO4 nanocomposite-based gas sensor exhibits promising nitrogen dioxide gas-sensing characteristics, such as fast response, highly selective detection, and extremely short response/recovery time. Additionally, the mechanisms of gas sensing in Pt/BiVO4 nanocomposites were explored in this paper.

Graphical Abstract

1. Introduction

Nitrogen dioxide (NO2) emissions are largely the result of fossil fuel combustion in automobiles and electrical power plants [1,2]. NO2 exposure at concentrations as low as a few ppm can be dangerous to human health and higher concentrations have been linked to smog and acid rain [3,4]. In the current study, we developed a highly sensitive NO2 sensor capable of detection at extremely low concentrations at room temperature [5,6,7]. One common approach to NO2 detection involves nanostructures of bismuth vanadate (BiVO4). Beyond NO2 detection [8,9,10], BiVO4 has been used in a wide variety of applications, such as formaldehyde degradation [11,12], H2S production [13,14], hydrogen production [15,16], CO2 capture [17,18], and other fields [19,20,21]. Table 1 lists a variety of NO2 sensors based on Pt or BiVO4. BiVO4 is inexpensive, highly responsive to visible light, stable, non-toxic, and environmentally benign with a narrow bandgap of Eg = 2.4 eV [22]. However, the low charge transfer rate of BiVO4 impedes efforts to enhance photocatalytic activity [23].
To improve charge transfer, various methods have been planned, including doping with various metals [24,25], the surface deposition of noble metals [26,27,28], and the formation of compound semiconductors [29,30]. The use of dopants has been shown to introduce electronic barriers, which capture photogenerated electrons and transfer them to other materials to prevent electron-hole recombination.
In this study, we developed a novel Pt/BiVO4 composite designed specifically for the sensing of NO2 at room temperature.

2. Materials and Methods

2.1. Materials

(Bi(NO3)3·5H2O), (NH4VO3), ethanol (C2H5OH), nitric acid (HNO3) ammonium hydroxide (NH4OH), sodium borohydride (NaBH4), ethylenediaminetetraacetic acid (EDTA), silver(I) nitrate (AgNO3) and dihydrogen hexachloroplatinate (VI) hexahydrate (H2PtCl6·6H2O) were purchased from Sigma-Aldrich Co., Inc. (St. Louis, MO, USA). Distilled H2O obtained using the distillation water system provided by the Millipore Corporation (Millipore Corp. Molsheim, France) was used.

2.2. Preparation of Bismuth Vanadate

Briefly, a few mmol of Bismuth nitrate (4 g) was added to 6.8 mmol of EDTA (2 g) in acidic medium (0.3 M), whereupon gradual heating to 85 °C was performed for the mixture to procure a colourless solution (named Solution A). A total of 8.5 mmol (1 g) of NH4VO3 was then dissolved in 50 mL of H2O at 60 °C under vigorous stirring for the generation of a yellow solution (named Solution B). The above two solutions mentioned (i.e., Solution A + Solution B) were then further processed with mixing at 50 °C for 1 hour with controlled stirring, after which, the resulting mixture was controlled with pH adjustment to 3.0 by the addition of salt (1 M). The prepared mixed product was poured into the autoclave (Dogger, New Taipei City, Taiwan) made of stainless steel with Teflon lining and further covered with sealant, then kept for heating, which was maintained at 180 °C for 6 h. After cooling of the autoclave, the resultant product was withdrawn via centrifugation, before undergoing multiple washings using ultrapure water and ethanol, followed by drying in an oven at 80 °C overnight and calcination at 450 °C for 4 h to complete the synthesis of BiVO4 [11].

2.3. Synthesis Method for Pt/BiVO4 Nanocomposite

Distilled water containing a pre-calculated quantity of H2PtCl6⋅6H2O was added to the as-prepared product of BiVO4, which was dispersed in 100 mL and further maintained for 1 h at a continuous stirring rate. The precursor suspension was produced as the product was extracted for centrifugation, washed three times using DI water, and then three times again using alcohol. The precipitate obtained was prepared by drying at 80 °C for more than 6 h. Pt/BiVO4 preparation was done with Pt at various weight concentrations labelled as follows: 0, 0.5, 1, 3, 5, or 10 wt.% Pt/BiVO4.

2.4. Characterization

The proposed Pt/BiVO4 nanocomposite was determined using a transmission electron microscope for morphology and structure with an energy-dispersive X-ray spectroscope (TEM/EDS; JEM-2100F), and a field emission scanning electron microscope (FESEM; JEOL JSM-7500F) operated at 30 kV. Pt/BiVO4 was characterized to understand its crystal structure using a Shimadzu X-ray diffractometer at 1.5405 Å at 40 kV and 30 mA, supported by a vertical goniometer in the range of 10° to 80° (2theta) at a scan speed of 2 °/min.

2.5. Sensor Fabrication and Measurements

The designed sensors were prepared by a dipping and coating procedure (Binder: PVA) on an alumina-based solid substrate (10 × 5 mm2; and fab of chip-based sensor; rotational speed, 1000 rpm) in the prepared material, to fabricate an electrode with a comb-like structure. The chips were subsequently pretreated to 80 °C over a period of 0.5 h and then calcined at a temperature of 400 °C for 2 h.
Figure 1 presents a schematic diagram showing the experimental setup used to measure the electrical response of the sensors [3]. A homemade arrangement for the sensor in the glass chamber was designed with a dynamic flow rate system to evaluate gas sensing performance. The target gas was injected into the chamber at the desired concentrations with a design including a mass controller arrangement for controlling the flow rate (1, 10, 30, 50, 70, or 100 ppm NO2) via a hole in the cover of the chamber. A simple circuit was utilized to understand signals coming from the head of the sensor, which is further showcased by resistance values; later on, collection of data for evaluation and processing was done with a PC. All resistance measurements were averaged from multiple measurements obtained using a Jiehan5000 data acquisition system with the input voltage from a power supply (Vs) set at 4.0 V. S = Rg/Ra—which is the ratio that is used for the calculation of the response from the sensor [31], where Ra indicates the resistance in the presence of air and Rg indicates the resistance in the presence of NO2 gas in the provided system. The times required for a 90% variation in resistance upon exposure to NO2 or air are described as the response and recovery times. Selectivity toward NO2 was assessed by exposing the sensor individually, including carbon monoxide, nitric oxide, methane (concentrations equal to 100 ppm), and further recording the characteristics of the response with corresponding values. Long-term stability was assessed by repeating the sensing measurements on multiple consecutive days.

3. Results and Discussion

3.1. Structure Property

Figure 2 shows XRD measurements, indicating the phases and structure of pure BivO4 and various Pt/BiVO4 nanocomposites. The samples exhibited characteristic peaks corresponding to a structure of BiVO4 (JCPDS 14-0688) with a monoclinic arrangement, as follows: 18.6° (110), 19.0° (011), 29.3° (121), 30.5° (040), 34.5° (200), 35.2° (002), 39.5° (211), 43.2° (051), 46.0° (042), 47.6° (240), 50.3° (202), 53.6° (161), 57.9° (321), and 59.3° (132) [19]. Peaks at 39.8° (111) and 46.0° (200) correspond to face-centred cubic Pt (JCPDS card: No. 04-0802) [32]. No other impurity peak was detected, indicating that the prepared samples were of high purity. Moreover, with increases in the amounts of deposited Pt, the diffraction peaks of Pt (111) were gradually intensified. Using Scherrer’s equation, the mean Pt (111), BiVO4 (051) and (161) crystalline sizes were estimated and listed in Table 1. Table 1 reveals that the crystalline sizes of Pt wt.% loading at 0.5%, 1%, 3%, 5% and 10% were 10.1 nm, 10.6 nm, 12.1 nm, 12.9 nm and 16.3 nm, respectively. The BiVO4 (051) and (161) crystalline sizes were 24.2–25.8 nm and 25.8–28.2 nm, respectively. This shows that the Pt (111) crystalline size was increased by enlarging the amount of Pt loading. However, the BiVO4 crystal size did not change greatly with the addition of platinum. These results indicate that the sensing material was a simple mixture.
The morphology and microstructure of the BiVO4 and Pt/BiVO4 were characterized by FESEM and TEM evaluation of the nanocomposite structures. As shown in Figure 3a, SEM images revealed an arrangement comprising a large number of irregularly stacked BiVO4 structures with an average diameter of 0.7–2.5 μm. Figure 3b shows the Pt/BiVO4 nanocomposite containing Pt particles, with uneven diameters on the BiVO4 sample. As shown in Figure 3c, TEM micrographs revealed d-spacing in the prescribed range of 0.227 and 0.196 nm, respectively corresponding to the original Pt-based (111) and (200) lattice plane [32]. As shown in Figure 3d, the EDS spectra confirmed the presence of Pt, V, Bi, Cu and O. The presence of strong signals from Cu can be ascribed to the Cu grid. Taken together, these confirm the formation of fully developed nanocomposites with a pure phase structure.

3.2. Gas-Sensing Performance of Pt/BiVO4

Figure 4a illustrates fluctuations in the response of sensors comprising (0, 0.5, 1, 3, 5 and 10) wt.% Pt/BiVO4 nanocomposites following exposure to NO2 gas at concentrations of 1–100 ppm at 25 °C. The increased response at all gas concentrations is indicative of typical n-type semiconductor behavior. The unloaded BiVO4 materials exhibited a strong response (91.3) when exposed to high concentrations of NO2 gas (100 ppm). Loading samples with a high concentration of Pt (10 wt.% Pt) increased the response strength substantially to 181.4. Table 1 lists the response times, recovery times, sensor responses, and base linearity results. The response and recovery times of sensors based on Pt/BiVO4 nanocomposites were shorter than those of Pt/BiVO4 nanocomposites. The 3 wt.% Pt/BiVO4 nanocomposites achieved a T90 of only 12 s and a Tr90 of 35 s. As shown in Figure 4b, this sample also presented the highest linearity (R2 = 0.992) and sensor response (S = 167.7).
Note that the response and recovery times of samples with high Pt concentrations (5 wt.% and 10 wt.%) were slower than those of the 3 wt.% Pt/BiVO4 nanocomposite. These results are similar to those reported in [31,33], indicating that at higher Pt concentrations, the formation of Pt aggregates tends to hinder the response and recovery of Pt/BiVO4 nanocomposites. Overall, the 3 wt.% Pt/BiVO4 sensor presented the best NO2 sensing performance and was therefore selected as the gas sensor material for all subsequent experiments.
Response and recovery times are important parameters in sensor characterization and should be as short as possible. Figure 5a presents typical dynamic response curves as a function of the weight ratios of Pt/BiVO4 when exposed to NO2 at a concentration of 100 ppm at room temperature. Figure 5b presents the dynamic sensing curve of the Pt/BiVO4 sensor when exposed to NO2 gas at various concentrations (1–100 ppm). Table 1 lists sensing properties, sensor response times, recovery times, and linearity (R2). The dramatic response of the sensors upon exposure to NO2 is typical of n-type semiconductors.
We can see in Figure 5a that the strength of the responses increased with low NO2 concentrations (1–100 ppm), as shown in Figure 5b. As shown in Figure 5c, the results from the proposed sensor maintained high reproducibility throughout the test cycle, thereby demonstrating that the Pt/BiVO4 sensor could reliably be at this particular concentration level to monitor the concentration of NO2.
Gas sensors are utilized and mostly evaluated with a range of selectivity, which is an important parameter. Figure 6 illustrates the responses of 3 wt.% Pt/BiVO4 nanocomposites to NO2, CO, NO, and CH4 (100 ppm) at room temperature. Under these conditions, the response to NO2 was 167.7, whereas the responses provided by the other selected gases were all lower, at a value of 30—thereby confirming the extraordinary selectivity demonstrated in 3 wt.% Pt/BiVO4 nanocomposites to NO2.
As shown in Figure 7, we did not observe a significant decrease in the response of the 3 wt.% Pt/BiVO4 nanocomposite despite continuous exposure to NO2 (10 ppm) for a period of 10 days, thereby demonstrating the stability of the sensors.

3.3. Underlying Sensing Mechanism

This study demonstrated that the addition of Pb to BiVO4 notably enhanced the response characteristics of the sensors. We assume that these effects are described by the provided reaction mechanism (1)–(4) [9,26,31,34,35]:
O 2 g a s + e O 2 a d s
N O 2 g a s N O 2 a d s
N O 2 a d s + O a d s N O g a s + O 2 a d s
N O 2 g a s + O 2 a d s + 2 e N O 2 g a s + 2 O a d s
The enhancement mechanism was projected and illustrated in Figure 8. The Pt-based BiVO4 nanocomposite exhibited typical n-type semiconductor behaviour [9,36]. When exposed to air, the active sites on the BiVO4 started adsorbing the oxygen molecules, which in turn gave free electrons from the BiVO4 material for the process of chemisorbed behaviour of oxygen ions ( O 2 ) at RT (Equation (1)). The subsequent adsorption of NO2 molecules at available adsorption sites on the BVO4 surface enables the direct extraction of the surface electrons of the prepared sensor, producing a breaking of bond in the form of NO—the result of which leaves free oxygen ions ( O 2 ) (Equations (2) and (3)) [37]. Catalysis of platinum can help in the enhancement of responsiveness of materials used for sensors to the various provided gases, due to a phenomenon referred to as the spillover effect [38,39]. The activity of Pt in this catalysis can help in accelerating the adsorption behaviour of molecular level oxygen; the adding of Pt dopants has a significant effect on surface area, which can lead to further chemisorbed oxygen species spillover, which increases the number of active sites on the BiVO4 [11,26,33,40]. This increases the number of electrons that are captured, and in so doing enhances the response of the sensor. As shown in Equation (4), the availability of additional oxygen molecules for reactions with incoming NO2 gas molecules increases the number of interactions between gas molecules in nitrous oxide and the sensing layer. The surface modification of BiVO4 with platinum also increases the number of dissociated NO2 molecules migrating to the surface of BiVO4, thereby increasing responsivity to NO2 gas. For comparison, the various NO2 sensors based on Pt or BiVO4 are listed in Table 2.

4. Conclusions

In this study, a Pt/BiVO4 nanocomposite was prepared using a hydrothermal method with various weight percentages of platinum for use in NO2 gas sensors. The structure and morphology of samples were characterized by XRD, TEM, FESEM, and EDX. The experiment results demonstrated that the addition of Pt to BiVO4 at a concentration of 3 wt.% can greatly enhance the responsivity of Pt/BiVO4 nanocomposite sensors to NO2 at relatively low concentrations (100 ppm) as follows: sensor response (167.7), response time (12 s), and recovery time (35 s). The proposed Pt/BiVO4 nanocomposite-based gas sensor exhibited promising nitrogen dioxide gas-sensing characteristics, including high sensitivity, high selectivity, and extremely short response/recovery times.

Author Contributions

Conceptualization, W.-D.L. and S.-Y.L.; methodology, W.-D.L.; software, W.-D.L.; validation, W.-D.L.; formal analysis, W.-D.L.; investigation, W.-D.L.; resources, W.-D.L.; data curation, W.-D.L. and S.-Y.L.; writing—original draft preparation, W.-D.L.; writing—review and editing, W.-D.L. and M.C.; visualization, W.-D.L. and S.-Y.L.; supervision, W.-D.L. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to a complicated structure that requires additional explanations.

Acknowledgments

The authors are pleased for the help and support provided by Providence University, Taiwan and its Applied Chemistry section, in fabricating sensor humidity control.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lim, S.H.; Feng, L.; Kemling, J.W.; Musto, C.J.; Suslick, K.S. An optoelectronic nose for the detection of toxic gases. Nat. Chem. 2009, 1, 562–567. [Google Scholar] [CrossRef] [Green Version]
  2. Kampa, M.; Castanas, E. Human health effects of air pollution. Environ. Pollut. 2008, 151, 362–367. [Google Scholar] [CrossRef]
  3. Zhu, Z.; Lin, S.-J.; Wu, C.-H.; Wu, R.-J. Synthesis of TiO2 nanowires for rapid NO2 detection. Sens. Actuator A Phys. 2018, 272, 288–294. [Google Scholar] [CrossRef]
  4. Zhang, W.; Hu, M.; Liu, X.; Wei, Y.; Li, N.; Qin, Y. Synthesis of the cactus-like silicon nanowires/tungsten oxide nanowires composite for room-temperature NO2 gas sensor. J. Alloys Compd. 2016, 679, 391–399. [Google Scholar] [CrossRef]
  5. Liu, Y.; Gao, X.; Li, F.; Lu, G.; Zhang, T.; Barsan, N. Pt-In2O3 mesoporous nanofibers with enhanced gas-sensing performance towards ppb-level NO2 at room temperature. Sens. Actuator B Chem. 2018, 260, 927–936. [Google Scholar] [CrossRef]
  6. Sharma, B.; Sharma, A.; Myung, J.-h. Selective ppb-level NO2 gas sensor based on SnO2-boron nitride nanotubes. Sens. Actuator B Chem. 2021, 331, 129464. [Google Scholar] [CrossRef]
  7. Bai, S.; Zhang, K.; Zhao, Y.; Li, Q.; Luo, R.; Li, D.; Chen, A. rGO decorated NiO-BiVO4 heterojunction for detection of NO2 at low temperature. Sens. Actuator B Chem. 2021, 329, 128912. [Google Scholar] [CrossRef]
  8. Choi, S.W.; Katoch, A.; Sun, G.J.; Kim, S.S. Bimetallic Pd/Pt nanoparticle-functionalized SnO2 nanowires for fast response and recovery to NO2. Sens. Actuator B Chem. 2013, 181, 446–453. [Google Scholar] [CrossRef]
  9. Li, Q.; Han, N.; Zhang, K.; Bai, S.; Guo, J.; Luo, R.; Li, D.; Chen, A. Novel p-n heterojunction of BiVO4/Cu2O decorated with rGO for low concentration of NO2 detection. Sens. Actuator B Chem. 2020, 320, 128284. [Google Scholar] [CrossRef]
  10. Bai, S.; Tian, K.; Fu, H.; Feng, Y.; Luo, R.; Li, D.; Chen, A.; Liu, C.C. Novel α-Fe2O3/BiVO4 heterojunctions for enhancing NO2 sensing properties. Sens. Actuator B Chem. 2018, 268, 136–143. [Google Scholar] [CrossRef]
  11. Zhu, Z.; Lin, Y.-C.; Chung, C.-L.; Wu, R.-J.; Huang, C.-L. A novel composite of triangular silver nanoplates on BiVO4 for gaseous formaldehyde degradation. Appl. Surf. Sci. 2021, 543, 148784. [Google Scholar] [CrossRef]
  12. Yang, J.; Shi, Q.; Zhang, R.; Xie, M.; Jiang, X.; Wang, F.; Cheng, X.; Han, W. BiVO4 quantum tubes loaded on reduced graphene oxide aerogel as an efficient photocatalyst for gaseous formaldehyde degradation. Carbon 2018, 138, 118–124. [Google Scholar] [CrossRef]
  13. Qiao, X.; Xu, Y.; Yang, K.; Ma, J.; Li, C.; Wang, H.; Jia, L. Mo doped BiVO4 gas sensor with high sensitivity and selectivity towards H2S. Chem. Eng. J. 2020, 395, 125144. [Google Scholar] [CrossRef]
  14. Li, C.; Qiao, X.; Jian, J.; Feng, F.; Wang, H.; Jia, L. Ordered porous BiVO4 based gas sensors with high selectivity and fast-response towards H2S. Chem. Eng. J. 2019, 375, 121924. [Google Scholar] [CrossRef]
  15. Li, F.; Leung, D.Y.C. Highly enhanced performance of heterojunction Bi2S3/BiVO4 photoanode for photoelectrocatalytic hydrogen production under solar light irradiation. Chem. Eng. Sci. 2020, 211, 115266. [Google Scholar] [CrossRef]
  16. Li, F.; Zhao, W.; Leung, D.Y.C. Enhanced photoelectrocatalytic hydrogen production via Bi/BiVO4 photoanode under visible light irradiation. Appl. Catal. B Environ. 2019, 258, 117954. [Google Scholar] [CrossRef]
  17. Chen, L.; Zhang, M.; Yang, J.; Li, Y.; Sivalingam, Y.; Shi, Q.; Xie, M.; Han, W. Synthesis of BiVO4 quantum dots/reduced graphene oxide composites for CO2 reduction. Mater. Sci. Semicond. Process. 2019, 102, 104578. [Google Scholar] [CrossRef]
  18. Li, X.; Wei, D.; Ye, L.; Li, Z. Fabrication of Cu2O-RGO/BiVO4 nanocomposite for simultaneous photocatalytic CO2 reduction and benzyl alcohol oxidation under visible light. Inorg. Chem. Commun. 2019, 104, 171–177. [Google Scholar] [CrossRef]
  19. Talasila, G.; Sachdev, S.; Srivastava, U.; Saxena, D.; Ramakumar, S.S.V. Modified synthesis of BiVO4 and effect of doping (Mo or W) on its photoelectrochemical performance for water splitting. Energy Rep. 2020, 6, 1963–1972. [Google Scholar] [CrossRef]
  20. Luo, J.; Fu, P.; Qu, Y.; Lin, Z.; Zeng, W. The n-butanol gas-sensing properties of monoclinic scheelite BiVO4 nanoplates. Phys. E 2018, 103, 71–75. [Google Scholar] [CrossRef]
  21. Lin, Y.; Cai, H.; Chen, H.; Luo, H. One–pot synthesis of Bi4V2O11/BiVO4 heterostructure with enhanced photocatalytic activity for dye degradation. Appl. Surf. Sci. 2021, 544, 148921. [Google Scholar] [CrossRef]
  22. Malathi, A.; Madhavan, J.; Ashokkumar, M.; Prabhakarn, A. A review on BiVO4 photocatalyst: Activity enhancement methods for solar photocatalytic applications. Appl. Catal. 2018, 555, 47–74. [Google Scholar]
  23. Abdi, F.F.; Savenije, T.J.; May, M.M.; Dam, B.; van de Krol, R. The origin of slow carrier transport in BiVO4 thin film photoanodes: A time-resolved microwave conductivity study. J. Phys. Chem. Lett. 2013, 4, 2752–2757. [Google Scholar] [CrossRef]
  24. Zhou, B.; Zhao, X.; Liu, H.; Qu, J.; Huang, C.P. Visible-light sensitive cobalt-doped BiVO4 (Co-BiVO4) photocatalytic composites for the degradation of methylene blue dye in dilute aqueous solutions. Appl. Catal. B Environ. 2010, 99, 214–221. [Google Scholar] [CrossRef]
  25. Wang, M.; Guo, P.; Chai, T.; Xie, Y.; Han, J.; You, M.; Wang, Y.; Zhu, T. Effects of Cu dopants on the structures and photocatalytic performance of cocoon-like Cu-BiVO4 prepared via ethylene glycol solvothermal method. J. Alloys Compd. 2017, 691, 8–14. [Google Scholar] [CrossRef]
  26. Mohamed, R.M.; Mkhalid, I.A.; Shawky, A. Facile synthesis of Pt–In2O3/BiVO4 nanospheres with improved visible-light photocatalytic activity. J. Alloys Compd. 2019, 775, 542–548. [Google Scholar] [CrossRef]
  27. Cao, S.-W.; Yin, Z.; Barber, J.; Boey, F.Y.C.; Loo, S.C.J.; Xue, C. Preparation of Au-BiVO4 Heterogeneous Nanostructures as Highly Efficient Visible-Light Photocatalysts. ACS Appl. Mater. Interfaces 2012, 4, 418–423. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, M.; Han, J.; Lv, C.; Zhang, Y.; You, M.; Liu, T.; Li, S.; Zhu, T. Ag, B, and Eu tri-modified BiVO4 photocatalysts with enhanced photocatalytic performance under visible-light irradiation. J. Alloys Compd. 2018, 753, 465–474. [Google Scholar] [CrossRef]
  29. Wang, W.; Wang, J.; Wang, Z.; Wei, X.; Liu, L.; Ren, Q.; Gao, W.; Liang, Y.; Shi, H. p–n junction CuO/BiVO4 heterogeneous nanostructures: Synthesis and highly efficient visible-light photocatalytic performance. Dalton Trans. 2014, 43, 6735–6743. [Google Scholar] [CrossRef] [PubMed]
  30. Fujimoto, I.; Wang, N.; Saito, R.; Miseki, Y.; Gunji, T.; Sayama, K. WO3/BiVO4 composite photoelectrode prepared by the improved auto-combustion method for highly efficient water splitting. Int. J. Hydrogen Energy 2014, 39, 2454–2461. [Google Scholar] [CrossRef]
  31. Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Wisitsoraat, A.; Phanichphant, S. NO2 gas sensing of flame-made Pt-loaded WO3 thick films. J. Solid State Chem. 2014, 214, 47–52. [Google Scholar] [CrossRef]
  32. Mirzaei, A.; Bang, J.H.; Choi, M.S.; Han, S.; Lee, H.Y.; Kim, S.S.; Kim, H.W. Changes in characteristics of Pt-functionalized RGO nanocomposites by electron beam irradiation for room temperature NO2 sensing. Ceram. Int. 2020, 46, 21638–21646. [Google Scholar] [CrossRef]
  33. Wang, C.-Y.; Hong, Z.-S.; Wu, R.-J. Promotion effect of Pt on a SnO2–WO3 material for NOx sensing. Phys. E 2015, 69, 191–197. [Google Scholar] [CrossRef]
  34. Bai, S.; Li, Q.; Han, N.; Zhang, K.; Tang, P.; Feng, Y.; Luo, R.; Li, D.; Chen, A. Synthesis of novel BiVO4/Cu2O heterojunctions for improving BiVO4 towards NO2 sensing properties. J. Colloid Interface Sci. 2020, 567, 37–44. [Google Scholar] [CrossRef]
  35. Liu, H.; Xu, Y.; Zhang, X.; Zhao, W.; Ming, A.; Wei, F. Enhanced NO2 sensing properties of Pt/WO3 films grown by glancing angle deposition. Ceram. Int. 2020, 46, 21388–21394. [Google Scholar] [CrossRef]
  36. He, H.; Zhou, Y.; Ke, G.; Zhong, X.; Yang, M.; Bian, L.; Lv, K.; Dong, F. Improved Surface Charge Transfer in MoO3/BiVO4 Heterojunction Film for Photoelectrochemical Water Oxidation. Electrochim. Acta 2017, 257, 181–191. [Google Scholar] [CrossRef]
  37. Zhang, C.; Luo, Y.; Xu, J.; Debliquy, M. Room temperature conductive type metal oxide semiconductor gas sensors for NO2 detection. Sens. Actuators A Phys. 2019, 289, 118–133. [Google Scholar] [CrossRef]
  38. Cabot, A.; Arbiol, J.; Morante, J.R.; Weimar, U.; Bârsan, N.; Göpel, W. Analysis of the noble metal catalytic additives introduced by impregnation of as obtained SnO2 sol-gel nanocrystals for gas sensors. Sens. Actuator B Chem. 2000, 70, 87–100. [Google Scholar] [CrossRef]
  39. Bang, J.H.; Mirzaei, A.; Han, S.; Lee, H.Y.; Shin, K.Y.; Kim, S.S.; Kim, H.W. Realization of low-temperature and selective NO2 sensing of SnO2 nanowires via synergistic effects of Pt decoration and Bi2O3 branching. Ceram. Int. 2021, 47, 5099–5111. [Google Scholar] [CrossRef]
  40. Lu, Y.; Jing, H.; Yu, H.; Zhao, Y.; Li, Y.; Huo, M. Enhanced catalytic performance of BiVO4/Pt under the combination of visible-light illumination and ultrasound waves. J. Taiwan Inst. Chem. E 2019, 102, 133–142. [Google Scholar] [CrossRef]
  41. Zhang, Q.; Wang, T.; Sun, Z.; Xi, L.; Xu, L. Performance improvement of photoelectrochemical NO2 gas sensing at room temperature by BiVO4-polyoxometalate nanocomposite photoanode. Sens. Actuator B Chem. 2018, 272, 289–295. [Google Scholar] [CrossRef]
Figure 1. Diagrammatic representation of the experimental setup.
Figure 1. Diagrammatic representation of the experimental setup.
Materials 14 05913 g001
Figure 2. XRD patterns of different contents for Pt/BiVO4 nanocomposites.
Figure 2. XRD patterns of different contents for Pt/BiVO4 nanocomposites.
Materials 14 05913 g002
Figure 3. (a) FESEM images of BiVO4 (b) Pt/BiVO4; (c) TEM images of Pt/BiVO4 (d) EDS spectrum of 3wt.% Pt/BiVO4 nanocomposites.
Figure 3. (a) FESEM images of BiVO4 (b) Pt/BiVO4; (c) TEM images of Pt/BiVO4 (d) EDS spectrum of 3wt.% Pt/BiVO4 nanocomposites.
Materials 14 05913 g003
Figure 4. (a) Discrepancies in response relative to variations in 1–100 ppm NO2 for varying Pt/BiVO4 nanocomposites contents. (b) Response of the 3wt.% Pt/BiVO4 sensor to 1–100 ppm NO2 at room temperature.
Figure 4. (a) Discrepancies in response relative to variations in 1–100 ppm NO2 for varying Pt/BiVO4 nanocomposites contents. (b) Response of the 3wt.% Pt/BiVO4 sensor to 1–100 ppm NO2 at room temperature.
Materials 14 05913 g004
Figure 5. (a) The response for varying Pt/BiVO4 nanocomposites contents of 100 ppm nitrogen dioxide at room temperature; (b) Curves for responses of the 3wt.% Pt/BiVO4 in the range of 1–100 ppm nitrogen dioxide at room temperature; (c) Three-cycle repeated response curves of the 3wt.% Pt/BiVO4 to 100 ppm nitrogen dioxide at room temperature.
Figure 5. (a) The response for varying Pt/BiVO4 nanocomposites contents of 100 ppm nitrogen dioxide at room temperature; (b) Curves for responses of the 3wt.% Pt/BiVO4 in the range of 1–100 ppm nitrogen dioxide at room temperature; (c) Three-cycle repeated response curves of the 3wt.% Pt/BiVO4 to 100 ppm nitrogen dioxide at room temperature.
Materials 14 05913 g005
Figure 6. Selectivity test of the 3wt.% Pt/BiVO4 nanocomposites to 100 ppm of various gases at room temperature.
Figure 6. Selectivity test of the 3wt.% Pt/BiVO4 nanocomposites to 100 ppm of various gases at room temperature.
Materials 14 05913 g006
Figure 7. Stability of 3 wt.% Pt/BiVO4 nanocomposites to 10 ppm NO2 operating at room temperature.
Figure 7. Stability of 3 wt.% Pt/BiVO4 nanocomposites to 10 ppm NO2 operating at room temperature.
Materials 14 05913 g007
Figure 8. Schematic diagram of NO2 gas sensing mechanism for Pt/BiVO4 nanocomposite.
Figure 8. Schematic diagram of NO2 gas sensing mechanism for Pt/BiVO4 nanocomposite.
Materials 14 05913 g008
Table 1. Comparison of the NO2-sensing performance of different contents of Pt/BiVO4 nanocomposites.
Table 1. Comparison of the NO2-sensing performance of different contents of Pt/BiVO4 nanocomposites.
Sensing
Material
Response
Time (T90, s)
Recovery
Time (Tr90, s)
Response
(100 ppm)
Pt Crystalline Size (by XRD) (nm), hkl, 111Bi Crystalline
Size BiVO4 (by XRD) (nm), hkl
Linearity
(R2)
(051)(161)
BiVO4
0.5 wt.% Pt/BiVO4
1 wt.% Pt/BiVO4
3 wt.% Pt/BiVO4
5 wt.% Pt/BiVO4
10 wt.% Pt/BiVO4
72
79
64
12
95
43
44
24
51
35
102
67
91.3
99.2
143.5
167.7
173.9
181.4
-
10.1
10.6
12.1
12.9
16.3
24.8
24.2
25.2
25.5
25.8
24.6
26.5
25.8
27.7
27.7
28.2
27.4
0.983
0.950
0.982
0.992
0.980
0.928
Table 2. Comparison of working temperature and various NO2 sensors based on Pt or BiVO4 nanocomposites.
Table 2. Comparison of working temperature and various NO2 sensors based on Pt or BiVO4 nanocomposites.
Sensing MaterialResponse
(Rg/Ra or Ra/Rg)
NO2
(ppm)
Temperature (°C)References
Pt-SnO2
α-Fe2O3/BiVO4
BiVO4/Cu2O
BiVO4/Cu2O/rGO
rGO-NiO-BiVO4
Pt/WO3
1.3
7.8
4.2
8.1
8.1
11.24
30
2
4
1
2
1
50
110
60
60
60
150
[41]
[10]
[34]
[9]
[7]
[35]
Pt/BiVO4167.710025This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lin, W.-D.; Lin, S.-Y.; Chavali, M. Improvement in NO2 Gas Sensing Properties of Semiconductor-Type Sensors by Loading Pt into BiVO4 Nanocomposites at Room Temperature. Materials 2021, 14, 5913. https://doi.org/10.3390/ma14205913

AMA Style

Lin W-D, Lin S-Y, Chavali M. Improvement in NO2 Gas Sensing Properties of Semiconductor-Type Sensors by Loading Pt into BiVO4 Nanocomposites at Room Temperature. Materials. 2021; 14(20):5913. https://doi.org/10.3390/ma14205913

Chicago/Turabian Style

Lin, Wang-De, Shu-Yun Lin, and Murthy Chavali. 2021. "Improvement in NO2 Gas Sensing Properties of Semiconductor-Type Sensors by Loading Pt into BiVO4 Nanocomposites at Room Temperature" Materials 14, no. 20: 5913. https://doi.org/10.3390/ma14205913

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