Next Article in Journal
Microbial Biosensor for Characterization of a Microorganism: A Review focusing on the Biochemical Activity of Microbial Cells
Previous Article in Journal
Wireless High Temperature Sensing Chipless Tag Based on a Diamond Ring Resonator
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of Zn: WO3 Thin Films as a Sensing Layer for Detection of NH3 Gas

1
Department of Physics, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, India
2
Department of Chemistry, School of Basic Sciences, Manipal University Jaipur, Jaipur 303007, India
3
Department of Physics, School of Basic Sciences, Manipal University Jaipur, Jaipur 303007, India
4
Department of Atomic and Molecular Physics, Manipal Academy of Higher Education, Manipal 576104, India
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(4), 732; https://doi.org/10.3390/mi14040732
Submission received: 14 September 2022 / Revised: 13 March 2023 / Accepted: 24 March 2023 / Published: 25 March 2023

Abstract

:
Pristine WO3 and Zn-doped WO3 were synthesized using the spray pyrolysis technique to detect ammonia gas. The prominent orientation of the crystallites along the (200) plane was evident from X-ray diffraction (XRD) studies. Scanning Electron Microscope (SEM) morphology indicated well-defined grains upon Zn doping with a smaller grain size of 62 nm for Zn-doped WO3 (Zn: WO3) film. The photoluminescence (PL) emission at different wavelengths was assigned to defects such as oxygen vacancies, interstitial oxygens, localized defects, etc. X-ray Photoelectron spectroscopy (XPS) studies confirmed the formation of oxygen vacancies in the deposited films. The ammonia (NH3) sensing analysis of the deposited films was carried out at an optimum working temperature of 250 °C. The sensor performance of Zn: WO3 was enhanced compared to pristine WO3 at 1 ppm NH3 concentration, elucidating the possibility of the films in sensing applications.

1. Introduction

Over the past few decades, chemiresistive-based gas sensors have drawn much attention to the development of the internet of things (IoT) due to the features such as low cost, miniaturization, and low power consumption [1,2,3,4,5]. Different semiconducting oxide materials such as ZnO, WO3, SnO2, TiO2, etc., were utilized to detect toxic gases such as NO2, CO, NH3, H2, etc. [6,7,8,9,10,11,12,13]. These gases in the human environment affect physical health by causing headaches, nausea, damage to the central nervous system and severe illness [14]. Among all metal oxides, Tungsten Oxide (WO3), a transition group metal oxide, was widely used to detect gases such as NH3, NO2, and alcohol vapors [15,16,17]. Ammonia (NH3), a reducing gas, is one of the harmful gases produced by chemical plants, automobiles, and agricultural wastes. According to the OSHA report, exposure to ammonia more than the permissible limit (27 mg/m3 or 35 ppm for 15 min duration) will severely damage human organs and lead to burns, chemical pneumonitis, and sometimes death [18]. Therefore, it is necessary to have fast and real-time ammonia sensors in order to reduce accidental threats. Furthermore, in the medical field, the lowest detection limit (50 ppb–2 ppm) is required for breath analysis. So, the detection of ammonia is indeed essential in monitoring health and the environment. Recently, WO3/Fe2O3 composites synthesized by the hydrothermal method showed an excellent response to ammonia at a working temperature of 300 °C [19]. Furthermore, a response of 13.6 was reported for Pt-loaded WO3 for 200 ppm ammonia [20]. Undoped films hinder sensing performance due to poor sensitivity and selectivity issues [21,22,23]. Hence, metal doping or composite preparation is a crucial step in the synthesis process and in ordering sensing properties.
In this work, we doped WO3 with metal dopant Zn due to its comparable cationic radii (0.6 Å for W6+; 0.74 Å for Zn2+), anticipating the substitution of Zn in the host lattice modulates optical, structural, and electrical properties [24,25]. Additionally, modification of these properties would enhance the gas sensing characteristics of Zn-doped WO3 (Zn: WO3). B. Nam and co-workers demonstrated the ZnO-WO3 composite for CO and NO2 sensing at 300 °C and 200 ppm gas concentration [26]. Based on their research, they concluded that the sensor worked better for reducing gas (CO) rather than oxidizing gas (NO2). To date, reports on Zn: WO3 for ammonia sensing has not been found in the literature. Due to the high operating temperature and high detection limit of the gases, as seen in the above literature, a practical approach to various applications is difficult. Hence, in our present work, we explored the sensing studies of spray-deposited Zn: WO3 films towards ammonia up to 1 ppm in concentration and achieved enhancement in the response value at an optimum working temperature of 250 °C.

2. Experimental

2.1. Chemicals

Ammonium metatungstate hydrate [(NH4)6H2W12O40·xH2O] from Sigma-Aldrich, St. Louis, Mo, USA, Zinc acetate dihydrate [Zn(CH3COO)2·2H2O] from Merck Life Science Pvt. Ltd., Mumbai, India, Polyethylene Glycol 400 (PEG 400) from Molychem, Mumbai, India was used without further purification.

2.2. Synthesis and Material Characterization

Pristine WO3 and Zn: WO3 film at 5 wt.% were synthesized using the spray pyrolysis deposition technique. The desired amount of tungsten and zinc precursors were dissolved in double distilled water to obtain a homogeneous solution of 0.01 M concentration. Solvent PEG 400, which acts as a surfactant, was added to the final solution and stirred using a magnetic stirrer for about 30 min. During deposition, spray parameters such as the flow rate (2 mL/min), the distance from the nozzle to the substrate (19 cm), and substrate temperature (400 °C) were kept constant throughout the experiment.
Structural information was obtained from an X-ray diffractometer that used Cu K-α radiation (1.54 Å). Morphological properties were analyzed via Scanning Electron Microscope. Various defects and sources of origin were determined using a Photoluminescence spectrofluorometer. Structural conformation and vibrational mode information were obtained from the Raman spectrometer. Composition and oxidation state analysis was carried out using X-ray Photoelectron Spectroscope.

2.3. Gas sensing Measurements

Synthesized films were kept in a closed chamber and were exposed to the air (79% N2 + 21% O2) and ammonia gas at different ppm levels. The ammonia concentration was controlled via a flow-through method wherein the carrier gas (air) was mixed with the target gas (NH3) with the help of mass flow controllers. I-V measurements were performed using silver paste as electrical contacts and variation in the resistance was noted from the Keithley source meter. The obtained values were used to calculate additional sensing parameters such as sensor response, response, and recovery times.

3. Results and Discussion

3.1. XRD Studies

XRD patterns of pristine WO3 and Zn: WO3 films by the Rigaku SmartLab X-ray diffractometer are illustrated in Figure 1. The diffraction peaks obtained were consistent with the monoclinic structure of WO3 and agreed with the JCPDS data card (no. 43-1035). Observed diffraction patterns at 22.97°, 23.44°, 24.18°, 33.99°, 48.15°, 49.73° and 55.74° corresponded to the peaks (002), (020), (200), (202), (040), (140) and (240). It indicates that the films have well crystalline and exhibit a strong preferential orientation (200), which is confirmed by an intense high peak located around 24.18° for both samples. It also confirms an increase in FWHM, which suggests the possibility of lowering the crystallite size, generating a non-uniform strain, or loss in crystalline nature with doping. Again, there was no evidence of any secondary phases corresponding to ZnO or ZnWO4, etc., during doping which validated the pure nature of the deposited samples. A decrease in crystallinity with Zn incorporation was reconfirmed by the reduction in peak intensities with the doping amount. The ionic radius of Zn2+ was 0.74 Å, and W6+ was 0.60 Å [24]. It is expected that during doping, some Zn2+ displace W6+ without changing its monoclinic crystal structure. Still, due to size and charge imbalance, a distortion in the crystal structure is reflected in its interplanar spacing values. This may be attributed to microstrain generation in the film during doping. The crystallite size for the dominant peak was estimated via Sherrer’s equation [27] and was found to decrease for the Zn: WO3 film. A similar trend was noticed by M.Arshad et al. [28] for Zn-doped WO3 nanoparticles. Structural parameters extracted from XRD patterns are given in Table 1.

3.2. Raman Studies

Raman spectra give an insight into crystallinity and bonding within the nanomaterials under investigation. The spectra (Figure 2) were obtained for pristine WO3 and Zn: WO3 films in 100–1000 cm−1 using a Horiba JOBINYVON LabRAM HR spectrometer. Intense wavenumber bands at 716 cm−1 and 806 cm−1 provided information about O-W-O asymmetric and symmetric stretches. In comparison, bands at 273 cm−1 and 326 cm−1 corresponded to O-W-O bending vibrations, and these two sets of bands validate the monoclinic phase of crystalline WO3 [29,30]. A slight shift (~2 cm−1) for bending vibrations might be due to the lattice distortion caused by Zn doping, as Zn2+ and W6+ have different ionic radii. Bands below 200 cm−1 may be attributed due to lattice vibrations. The evolution of a fragile band at 958 cm1 for Zn: WO3 film corresponds to the symmetric stretching of W=O terminal bonds at the grain surface [31].

3.3. SEM Studies

Morphological features of pristine WO3 and Zn: WO3 films were investigated via ZEISS ULTRA55 Scanning Electron Microscope. Figure 3a–d illustrates the closely packed grains with uniform distribution all over the surface. It is evident from the obtained morphographs that the grain size of Zn: WO3 films decreased compared to the pristine WO3 film. The estimated grain size for pristine WO3 and Zn: WO3 films were around 98 nm and 62 nm, respectively. A smaller grain size facilitated the chemisorption phenomena and the sensitivity of sensors by diffusing the gas throughout the grains and reducing the barrier for charge transfer between the adjacent grains [32].

3.4. PL Studies

PL measurements of pristine WO3 and Zn: WO3 films were performed at room temperature via JASCO FP 8300 spectrofluorometer excited at 256 nm using a Xe lamp source, as shown in Figure 4. Luminescent centers at 4.25 eV and 4.02 eV correspond to NUV emissions predominantly, which arise owing to the deep level of oxygen vacancies present in the conduction band of the films [33]. A PL emission at 3.47 eV was due to band-to-band transitions. In contrast, the emission center at 2.75 eV is probably due to the interstitial oxygen or other impurities present in the films [34]. The lower energy emission center at 2.12 eV can be ascribed to the localized defects lying within the bandgap of the films [35]. The enhanced emission intensity for Zn: WO3 films signified an increased radiative defect density upon Zn doping. The surface states generally increased as the grain size decreased due to a large surface-to-volume ratio. In addition, PL intensity increased as the density of the surface states increased. Thus, films with smaller grain sizes show higher luminescence than larger ones. The results observed are consistent with SEM analysis.

3.5. XPS Studies

Chemical state and composition studies were performed using a Kratos AXIS ULTRA X-ray Photoelectron spectroscope. Figure 5a–e represents the deconvoluted XPS spectra composed of W 4f, O 1s and Zn 2p core levels in WO3 and Zn: WO3 films. The high resolution W 4f peaks (Figure 5a) signifies the W4f7/2 and W4f5/2 spin-orbit doublet at the binding energies (Eb) 36.15 eV and 38.28 eV, respectively, corresponding to the +6 oxidation state [11]. Upon Zn dopng, W 4f energy levels shifted (~0.1 eV) toward lower Eb, suggesting the emission of electrons from lower oxidation states of W (substoichiometric WO3−x) [36]. A satellite peak at 42.24 eV was attributed to W 5p3/2. Oxygen 1s spectra (Figure 5c) contained two peaks, namely OI and OII at 530.68 eV and 531.61 eV, respectively, the former of which is associated with the lattice oxygen ( O 2 ) bonded to tungsten ions, and latter corresponded to oxygen vacancy (O) formed in the lattice [37]. The slight shift (~0.09 eV) of these peaks towards lower Eb was observed in the case of Zn: WO3, which could be attributed to the formation of substoichiometric WO3−x. Oxygen vacancies in the films served as better adsorption centres for the gas molecules and contributed to enhancement in the sensing performance [37]. Figure 5e depicts the core level spectra of Zn 2p at the binding energies 1021.84 eV (Zn 2p3/2) and 1045.06 eV (Zn 2p1/2) with respect to the +2 oxidation state, which confirmed the successful doping of Zn into WO3 [6,7].

3.6. NH3 Gas Sensing Analysis

Pristine WO3 and Zn: WO3 films were exposed to air and ammonia atmosphere at an optimized operating temperature of 250 °C. The temperature optimization plot is given in Figure 6. Transient response curves were recorded for ammonia concentrations of 1, 3, and 5 ppm, as depicted in Figure 7a,b. The sensor response for reducing gases such as NH3 is defined by S = ( R a R g R g ) [38,39]. Ra and Rg are resistances of the film when subjected to air and ammonia, respectively. Response and recovery times are described with a 90% resistance change in the film when it was kept in ammonia and the air, respectively. The sensor response, response time, and recovery time of the films were calculated and reported in Table 2. The variation in the rate of response/recovery is attributed to the change in surface reactions such as the adsorption, diffusion of gas molecules, etc. A higher sensor response of 1.40 was observed for the Zn: WO3 film with a response time of 167 s and recovery time of 199 s at 5 ppm NH3 concentration. Moreover, Zn: WO3 exhibited good sensor performance at concentrations as low as 1 ppm. A smaller grain size observed from SEM increased the surface states observed in PL, and the oxygen vacancy formation noticed from XPS perhaps induced the increment in the response of the Zn: WO3 film. Table 3 depicts the comparison study of our work with the previously reported ammonia sensors. The present sensor stands out among the reported literature due to its better response, low detection limit, and moderate operating temperature.
Figure 8 demonstrates the sensing mechanism of metal oxides in the presence of air and ammonia gas. When sensing material (i.e., Zn: WO3) is kept in an air atmosphere at a specific temperature, oxygen molecules adsorb on the surface and form ionic species ( O 2 , O a n d   O 2 ) by removing electrons from the conduction band of the film. As a result, the formation of a space-charge layer (w) within the grain and potential barrier (∆φ) and between the grains increases the film’s resistance. When ammonia interacts with the sensor material, the release of electrons to the conduction band occurs through a reduction in the width of ‘w’ and ′∆φ′ and decreases the film’s resistance. The following equations represent the reactions of chemisorbed oxygen species at lower temperatures and their interaction with ammonia [11].
O 2 a d s +   e   O 2
O 2 + e     2 O
4 NH 3 ( g ) + 5 O 2 4 N O + 6 H 2 O + 5 e
2NH3(g) + 3O → N2 + 3H2O + 3e
Selectivity and repeatability are the key aspects that govern the sensing performance of the films. Figure 9 shows the bar graph representing the selectivity of the Zn: WO3 film at 5 ppm ammonia concentration. The Zn: WO3 film showed a higher response to ammonia compared to the other tested gases, such as CO, NO2, and CH4. Figure 10 depicts the repeatable signals of the Zn: WO3 film at 5 ppm NH3 concentration. The repeatability test conducted for 5 cycles of ammonia exhibited a steady response after each cycle, signifying the excellent repeatability characteristics of the Zn: WO3 film. Since the repeatability measurements were carried out after 8 months due to atmospheric conditions, a change in the resistance was noted. However, we did not observe a large variation in the sensor response of the films.

4. Conclusions

Spray-deposited pristine WO3 and Zn: WO3 films were investigated for their structural, morphological, optical, and ammonia-sensing properties. XRD and Raman confirmed the monoclinic structure of the deposited films. Smaller grains were observed from SEM and oxygen vacancy formation, which were confirmed via XPS and might have ameliorated the sensor response in the Zn: WO3 film. Repeatability measurements revealed the good reproducible characteristics of the Zn: WO3 sensor toward ammonia gas. Hence, our study confirmed that WO3 doped with Zn was a potential approach to enhance the ammonia sensing characteristics at a moderate operating temperature of 250 °C.

Author Contributions

Conceptualization, P.P.; Methodology, A. and S.C.; Software, P.K.; Validation, S.D.K.; Formal analysis, A., P.P. and A.R.; Investigation, A. and S.C.; Resources, P.P.; Data curation, P.K., A.R. and S.D.K.; Writing—original draft, A.; Supervision, P.P.; Funding acquisition, P.P. All authors have read and agreed to the published version of the manuscript.

Funding

Poornesh P acknowledges Manipal Academy of Higher Education, Manipal, for funding the research project through Intramural Funding (Grant no.: MAHE/DREG/Ph.D./IMF/2019).

Data Availability Statement

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

Acknowledgments

Anusha acknowledges the Indian Nano electronics User Program (INUP), supported by the Ministry of Human Resource Development (MHRD), Govt. of India, located at CeNSE, Indian Institute of Science Bengaluru for providing the facilities to carry out part of the work. Priyanka Kumari and Saikat Chattopadhyay thank the Sophisticated Analytical Instrumentation Facility (SAIF) Manipal University Jaipur (MUJ), India for XRD measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ma, J.; Ren, Y.; Zhou, X.; Liu, L.; Zhu, Y.; Cheng, X.; Xu, P.; Li, X.; Deng, Y.; Zhao, D. Pt Nanoparticles Sensitized Ordered Mesoporous WO3 Semiconductor: Gas Sensing Performance and Mechanism Study. Adv. Funct. Mater. 2017, 28, 1705268. [Google Scholar] [CrossRef]
  2. Chow, L.; Lupan, O.; Chai, G.; Khallaf, H.; Ono, L.; Cuenya, B.R.; Tiginyanu, I.; Ursaki, V.; Sontea, V.; Schulte, A. Synthesis and characterization of Cu-doped ZnO one-dimensional structures for miniaturized sensor applications with faster response. Sensors Actuators A Phys. 2013, 189, 399–408. [Google Scholar] [CrossRef]
  3. Guan, Y.; Wang, D.; Zhou, X.; Sun, P.; Wang, H.; Ma, J.; Lu, G. Hydrothermal preparation and gas sensing properties of Zn-doped SnO2 hierarchical architectures. Sensors Actuators B Chem. 2014, 191, 45–52. [Google Scholar] [CrossRef]
  4. Upadhyay, S.; Mishra, R.; Sahay, P. Cr-doped WO3 nanosheets: Structural, optical and formaldehyde sensing properties. Ceram. Int. 2016, 42, 15301–15310. [Google Scholar] [CrossRef]
  5. Kaur, J.; Anand, K.; Kaur, A.; Singh, R.C. Sensitive and selective acetone sensor based on Gd doped WO3/reduced graphene oxide nanocomposite. Sens. Actuators B Chem. 2018, 258, 1022–1035. [Google Scholar] [CrossRef]
  6. Zhou, Q.; Chen, W.; Xu, L.; Kumar, R.; Gui, Y.; Zhao, Z.; Tang, C.; Zhu, S. Highly sensitive carbon monoxide (CO) gas sensors based on Ni and Zn doped SnO2 nanomaterials. Ceram. Int. 2017, 44, 4392–4399. [Google Scholar] [CrossRef]
  7. Shewale, P.S.; Yun, K.-S. Synthesis and characterization of Cu-doped ZnO/RGO nanocomposites for room-temperature H2S gas sensor. J. Alloy. Compd. 2020, 837, 155527. [Google Scholar] [CrossRef]
  8. Tian, X.; Cui, X.; Lai, T.; Ren, J.; Yang, Z.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Gas sensors based on TiO2 nanostructured materials for the detection of hazardous gases: A review. Nano Mater. Sci. 2021, 3, 390–403. [Google Scholar] [CrossRef]
  9. Zeng, Y.; Hua, Z.; Tian, X.; Li, X.; Qiu, Z.; Zhang, C.; Wang, M.; Li, E.-P. Selective detection of methanol by zeolite/Pd-WO3 gas sensors. Sensors Actuators B Chem. 2018, 273, 1291–1299. [Google Scholar] [CrossRef]
  10. Khudiar, A.I.; Oufi, A.M. Influence of the aluminium doping on the physical and gas sensing properties of SnO2 for H2 gas detection. Sensors Actuators B Chem. 2021, 340, 129633. [Google Scholar] [CrossRef]
  11. Chao, J.; Liu, Z.; Xing, S.; Gao, Q.; Zhao, J. Enhanced ammonia detection of gas sensors based on square-like tungsten oxide loaded by Pt nanoparticles. Sensors Actuators B Chem. 2021, 347, 130621. [Google Scholar] [CrossRef]
  12. Yin, X.T.; Zhou, W.D.; Dastan, D.; Li, J.; Tan, X.M.; Liu, Y.; Gao, X.C.; Ma, X.G. Selectivity sensing response of ZnO-xCo3O4 based sensor to CO against CH4. Mater. Sci. Semicond. Process. 2022, 149, 106883. [Google Scholar] [CrossRef]
  13. Li, A.; Zhao, S.; Bai, J.; Gao, S.; Wei, D.; Shen, Y.; Yuan, Z.; Meng, F. Optimal construction and gas sensing properties of SnO2@TiO2 heterostructured nanorods. Sensors Actuators B Chem. 2021, 355, 131261. [Google Scholar] [CrossRef]
  14. Balamurugan, C.; Song, S.-J.; Kim, H.-S. Enhancing Gas Response Characteristics of Mixed Metal Oxide Gas Sensors. J. Korean Ceram. Soc. 2018, 55, 1–20. [Google Scholar] [CrossRef] [Green Version]
  15. Liu, G.; Zhu, L.; Yu, Y.; Qiu, M.; Gao, H.; Chen, D. WO3 nanoplates for sensitive and selective detections of both acetone and NH3 gases at different operating temperatures. J. Alloy. Compd. 2020, 858, 157638. [Google Scholar] [CrossRef]
  16. Adhyapak, P.V.; Bang, A.D.; More, P.; Munirathnam, N.R. Nanostructured WO3/graphene composites for sensing NOx at room temperature. RSC Adv. 2018, 8, 34035–34040. [Google Scholar] [CrossRef] [Green Version]
  17. Hung, C.M.; Dat, D.Q.; Van Duy, N.; Van Quang, V.; Van Toan, N.; Van Hieu, N.; Hoa, N.D. Facile synthesis of ultrafine rGO/WO3 nanowire nanocomposites for highly sensitive toxic NH3 gas sensors. Mater. Res. Bull. 2020, 125, 110810. [Google Scholar] [CrossRef]
  18. NIOSH, CDC, and USDHHS. Occupational Safety and Health Guideline for Ammonia. In Occupational Safety and Health Guidelines; US Department of Health and Human Services: Washington, DC, USA, 1992; pp. 1–7. [Google Scholar]
  19. Dien, N.D.; Phuoc, L.H.; Hien, V.X.; Vuong, D.D.; Chien, N.D. Hydrothermal Synthesis and Ammonia Sensing Properties of WO3/Fe2O3 Nanorod Composites. J. Electron. Mater. 2017, 46, 3309–3316. [Google Scholar] [CrossRef]
  20. Wang, Y.; Liu, J.; Cui, X.; Gao, Y.; Ma, J.; Sun, Y.; Sun, P.; Liu, F.; Liang, X.; Zhang, T.; et al. NH3 gas sensing performance enhanced by Pt-loaded on mesoporous WO3. Sensors Actuators B Chem. 2016, 238, 473–481. [Google Scholar] [CrossRef]
  21. Takács, M.; Dücső, C.; Lábadi, Z.; Pap, A. Effect of Hexagonal WO3 Morphology on NH3 Sensing. Procedia Eng. 2014, 87, 1011–1014. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, X.; Miura, N.; Yamazoe, N. Study of WO3-based sensing materials for NH3 and NO detection. Sensors Actuators B Chem. 2000, 66, 74–76. [Google Scholar] [CrossRef]
  23. Xu, Y.; Tan, Q.; Tang, Z.; Zhang, Z.; Yuan, Z. WO3-Based Gas sensors. Prog. Chem. 2009, 21, 2734–2743. [Google Scholar]
  24. Cheng, X.F.; Leng, W.H.; Liu, D.P.; Zhang, J.Q.; Cao, C.N. Enhanced photoelectrocatalytic performance of Zn-doped WO3 photocatalysts for nitrite ions degradation under visible light. Chemosphere 2007, 68, 1976–1984. [Google Scholar] [CrossRef]
  25. Madhan, D.; Parthibavarman, M.; Rajkumar, P.; Sangeetha, M. Influence of Zn doping on structural, optical and photocatalytic activity of WO3 nanoparticles by a novel microwave irradiation technique. J. Mater. Sci. Mater. Electron. 2015, 26, 6823–6830. [Google Scholar] [CrossRef]
  26. Nam, B.; Ko, T.-K.; Hyun, S.-K.; Lee, C. Sensitivities of a 6:4 (by molar ratio) ZnO/WO3 composite nanoparticle sensor to reducing and oxidizing gases. Appl. Surf. Sci. 2020, 504, 144104. [Google Scholar] [CrossRef]
  27. Scherrer, P. Bestimmung der Grösse und der inneren Struktur von Kolloidteilchen mittels Röntgensrahlen [Determination of the size and internal structure of colloidal particles using X-rays]. Nachr. Ges. Wiss. Göttingen Math. Phys. Kl. 1918, 26, 98–100. [Google Scholar]
  28. Arshad, M.; Ehtisham-ul-Haque, S.; Bilal, M.; Ahmad, N.; Ahmad, A.; Abbas, M.; Nisar, J.; Khan, M.I.; Iqbal, M. Synthesis and characterization of Zn doped WO3 nanoparticles: Photocatalytic, antifungal and antibacterial activities evaluation. Mater. Res. Express 2020, 7, 1. [Google Scholar] [CrossRef]
  29. Thummavichai, K.; Wang, N.; Xu, F.; Rance, G.; Xia, Y.; Zhu, Y. In situ investigations of the phase change behaviour of tungsten oxide nanostructures. R. Soc. Open Sci. 2018, 5, 171932. [Google Scholar] [CrossRef] [Green Version]
  30. Garcia-Sanchez, R.F.; Ahmido, T.; Casimir, D.; Baliga, S.; Misra, P. Thermal Effects Associated with the Raman Spectroscopy of WO3Gas-Sensor Materials. J. Phys. Chem. A 2013, 117, 13825–13831. [Google Scholar] [CrossRef] [PubMed]
  31. Shigesato, Y.; Murayama, A.; Kamimori, T.; Matsuhiro, K. Characterization of evaporated amorphous WO3 films by Raman and FTIR spectroscopies. Appl. Surf. Sci. 1988, 34, 804–811. [Google Scholar] [CrossRef]
  32. Rothschild, A.; Komem, Y. The effect of grain size on the sensitivity of nanocrystalline metal-oxide gas sensors. J. Appl. Phys. 2004, 95, 6374–6380. [Google Scholar] [CrossRef]
  33. Kriti, K.P.; Kaur, S.; Arora, D.; Asokan, K.; Singh, D.P. Influence of defect structure on colour tunability and magneto optical behaviour of WO3 nanoforms. RSC Adv. 2019, 9, 20536–20548. [Google Scholar] [CrossRef] [Green Version]
  34. Ganbavle, V.; Mohite, S.; Agawane, G.; Kim, J.; Rajpure, K. Nitrogen dioxide sensing properties of sprayed tungsten oxide thin film sensor: Effect of film thickness. J. Colloid Interface Sci. 2015, 451, 245–254. [Google Scholar] [CrossRef]
  35. Yang, C.; Zhu, Q.; Zhang, S.; Zou, Z.; Tian, K.; Xie, C. A comparative study of microstructures on the photoelectric properties of tungsten trioxide films with plate-like arrays. Appl. Surf. Sci. 2014, 297, 116–124. [Google Scholar] [CrossRef]
  36. Vuong, N.M.; Kim, D.; Kim, H. Porous Au-embedded WO3 Nanowire Structure for Efficient Detection of CH4 and H2S. Sci. Rep. 2015, 5, srep11040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Yu, W.; Shen, Z.; Peng, F.; Lu, Y.; Ge, M.; Fu, X.; Sun, Y.; Chen, X.; Dai, N. Improving gas sensing performance by oxygen vacancies in sub-stoichiometric WO3−x. RSC Adv. 2019, 9, 7723–7728. [Google Scholar] [CrossRef] [Green Version]
  38. Büyükköse, S. Highly selective and sensitive WO3 nanoflakes based ammonia sensor. Mater. Sci. Semicond. Process. 2020, 110, 104969. [Google Scholar] [CrossRef]
  39. Kwak, D.; Wang, M.; Koski, K.J.; Zhang, L.; Sokol, H.; Maric, R.; Lei, Y. Molybdenum Trioxide (α-MoO3) Nanoribbons for Ultrasensitive Ammonia (NH3) Gas Detection: Integrated Experimental and Density Functional Theory Simulation Studies. ACS Appl. Mater. Interfaces 2019, 11, 10697–10706. [Google Scholar] [CrossRef]
  40. Hieu, N.V.; Quang, V.V.; Hoa, N.D.; Kim, D. Preparing large-scale WO3 nanowire-like structure for high sensitivity NH3 gas sensor through a simple route. Curr. Appl. Phys. 2011, 11, 657–661. [Google Scholar] [CrossRef]
  41. Godbole, R.; Godbole, V.; Bhagwat, S. Palladium enriched tungsten oxide thin films: An efficient gas sensor for hazardous gases. Eur. Phys. J. B 2019, 92, 78. [Google Scholar] [CrossRef]
  42. Qiu, Z.; Tian, X.; Li, Y.; Zeng, Y.; Fan, C.; Wang, M.; Hua, Z. NH3 sensing properties and mechanism of Ru-loaded WO3 nanosheets. J. Mater. Sci. Mater. Electron. 2018, 29, 11336–11344. [Google Scholar] [CrossRef]
  43. Yuan, K.-P.; Zhu, L.-Y.; Yang, J.-H.; Hang, C.-Z.; Tao, J.-J.; Ma, H.-P.; Jiang, A.-Q.; Zhang, D.W.; Lu, H.-L. Precise preparation of WO3@SnO2 core shell nanosheets for efficient NH3 gas sensing. J. Colloid Interface Sci. 2020, 568, 81–88. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of WO3 and Zn: WO3 films.
Figure 1. XRD patterns of WO3 and Zn: WO3 films.
Micromachines 14 00732 g001
Figure 2. Raman spectra of WO3 and Zn: WO3 films.
Figure 2. Raman spectra of WO3 and Zn: WO3 films.
Micromachines 14 00732 g002
Figure 3. SEM micrographs of (a,b) WO3 and (c,d) Zn: WO3 films.
Figure 3. SEM micrographs of (a,b) WO3 and (c,d) Zn: WO3 films.
Micromachines 14 00732 g003
Figure 4. PL spectra of WO3 and Zn: WO3 films.
Figure 4. PL spectra of WO3 and Zn: WO3 films.
Micromachines 14 00732 g004
Figure 5. Deconvoluted core level XPS spectra of (a,b) W 4f (c,d) O 1s (e) Zn 2p in WO3 and Zn: WO3, respectively.
Figure 5. Deconvoluted core level XPS spectra of (a,b) W 4f (c,d) O 1s (e) Zn 2p in WO3 and Zn: WO3, respectively.
Micromachines 14 00732 g005
Figure 6. Temperature optimization plot of Zn: WO3.
Figure 6. Temperature optimization plot of Zn: WO3.
Micromachines 14 00732 g006
Figure 7. (a) Transient response curves of (a) WO3 and (b) Zn: WO3 films.
Figure 7. (a) Transient response curves of (a) WO3 and (b) Zn: WO3 films.
Micromachines 14 00732 g007
Figure 8. Mechanism of sensing in the presence of (a) Air (b) Ammonia for Zn:WO3 film.
Figure 8. Mechanism of sensing in the presence of (a) Air (b) Ammonia for Zn:WO3 film.
Micromachines 14 00732 g008
Figure 9. Selectivity studies of Zn: WO3 toward various gases.
Figure 9. Selectivity studies of Zn: WO3 toward various gases.
Micromachines 14 00732 g009
Figure 10. Repeatability measurements of Zn: WO3 film.
Figure 10. Repeatability measurements of Zn: WO3 film.
Micromachines 14 00732 g010
Table 1. Structural parameters of the films extracted from XRD patterns.
Table 1. Structural parameters of the films extracted from XRD patterns.
Samples2θ, (200)
(deg.)
FWHM (deg.)Interplanar Spacing d (Å)Crystallite Size D (nm) Dislocation Density δ (×1014 Lines/m2)Microstrain
ε (×10−3)
WO324.180.252973.677329.71.1
Zn: WO324.180.270973.6763011.11.2
Table 2. Sensing parameters of WO3 and Zn: WO3 films.
Table 2. Sensing parameters of WO3 and Zn: WO3 films.
Conc.
(ppm)
Sensor ResponseResponse Time (s)Recovery Time (s)
WO310.15156200
30.40198159
50.93164123
Zn: WO310.2489188
30.65118176
51.40167199
Table 3. Comparison study with the previously reported NH3 sensors.
Table 3. Comparison study with the previously reported NH3 sensors.
MaterialNH3 Conc. (ppm)Sensor ResponseOperating Temperature (°C)MethodRef.
WO3 nanowires15009.7250Sputtering and calcination[40]
WO3-Fe2O3 composites3006300Hydrothermal[19]
Pd-WO3 films100.27225Spray Pyrolysis[41]
rGO/WO3nanowire composites10011300Hydrothermal[17]
Ru loaded WO3 nanosheets2017.8300Acidification with impregnation[42]
WO3@SnO2 core shell nanostructures151.5200Hydrothermal[43]
Zn: WO3 nanostructures51.40250Spray PyrolysisThis 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

Anusha; Kumari, P.; Poornesh, P.; Chattopadhyay, S.; Rao, A.; Kulkarni, S.D. Evaluation of Zn: WO3 Thin Films as a Sensing Layer for Detection of NH3 Gas. Micromachines 2023, 14, 732. https://doi.org/10.3390/mi14040732

AMA Style

Anusha, Kumari P, Poornesh P, Chattopadhyay S, Rao A, Kulkarni SD. Evaluation of Zn: WO3 Thin Films as a Sensing Layer for Detection of NH3 Gas. Micromachines. 2023; 14(4):732. https://doi.org/10.3390/mi14040732

Chicago/Turabian Style

Anusha, Priyanka Kumari, P. Poornesh, Saikat Chattopadhyay, Ashok Rao, and Suresh D. Kulkarni. 2023. "Evaluation of Zn: WO3 Thin Films as a Sensing Layer for Detection of NH3 Gas" Micromachines 14, no. 4: 732. https://doi.org/10.3390/mi14040732

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