Next Article in Journal
Green, Sustainable Synthesis of γ-Fe2O3/MWCNT/Ag Nano-Composites Using the Viscum album Leaf Extract and Waste Car Tire for Removal of Sulfamethazine and Bacteria from Wastewater Streams
Next Article in Special Issue
Strengthening and Toughening CNTs/Mg Composites by OpTimizing the Grinding Time of Magnesium Powder
Previous Article in Journal
Investigation on Crystal-Structure, Thermal and Electrical Properties of PVDF Nanocomposites with Cobalt Oxide and Functionalized Multi-Wall-Carbon-Nanotubes
Previous Article in Special Issue
Nonlinear Vibration of a Pre-Stressed Water-Filled Single-Walled Carbon Nanotube Using Shell Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Room Temperature Ammonia Gas Sensing Properties of Fe-Doped MoO3 Thin Films Fabricated Using Nebulizer Spray Pyrolysis

by
Fatemah H. Alkallas
1,
Amira Ben Gouider Trabelsi
1,
Mohd Shkir
2,3,* and
Salem AlFaify
2
1
Department of Physics, College of Science, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Advanced Functional Materials & Optoelectronic Laboratory (AFMOL), Department of Physics, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
3
Department of Chemistry and University Centre for Research & Development, Chandigarh, University, Mohali 140413, Punjab, India
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(16), 2797; https://doi.org/10.3390/nano12162797
Submission received: 21 July 2022 / Revised: 8 August 2022 / Accepted: 10 August 2022 / Published: 15 August 2022

Abstract

:
MoO3 thin films are fabricated using nebulizer spray pyrolysis technique, which is doped with Fe at various concentrations of 1, 2, 3, and 4% for ammonia gas sensors application at room temperature. X-ray diffraction (XRD) study confirms the growth of the crystal by Fe doping up to 3%, nano rods shape morphology of the thin film samples observed by field emission scanning electron microscope (FESEM), reduction in bandgap is evidenced via UV-VIS spectrophotometer. Gas sensing study is performed using gas analyzing chamber attached with Keithley source meter. Since 3% Fe doped MoO3 sample displayed nano rods over the film surface which exhibits highest sensitivity of 38,500%, in a short period of raise and decay time 54 and 6 s. Our findings confirms that the 3% Fe doped MoO3 films suitability for ammonia gas sensing application.

Graphical Abstract

1. Introduction

In today’s world, continuous detection of various gases is required for industrial process monitoring, automotive gas emission, and outdoor air quality monitoring [1]. Conventionally metal oxide semiconductors like SnO2, TiO2, WO3, MoO3, CeO2, Nb2O5, ZnO, etc. have been most frequently used as sensing materials [2,3,4,5,6,7,8,9,10]. The detection ability, operating temperature, and fabrication efficiency are the key factors that limit the real-time use of these gas sensors. Global technological efforts have been made to improve sensor selectivity, sensitivity, stability, repeatability, responsiveness, and recovery times [11,12,13,14,15]. Several materials have been developed so far, including molybdenum-based gas sensors such as molybdenum oxide (MoO3) and molybdenum sulphide (MoS2), allowing more gaseous chemicals adsorbed on its surface, resulting in high sensitivity [16,17,18,19,20,21,22,23]. MoO3 is an n-type semiconductor with a band gap between 2.39-2.9 eV which exhibits orthorhombic (α-MoO3), monoclinic (β-MoO3), hexagonal (h-MoO3), and ε-MoO3 phases [24,25,26]. Moreover, MoO3 suffers from disadvantages while satisfying the practical requirements like lower response and interference of various gases [27,28]. Currently, different strategies related to the surface modification, doping, nanoscale size and shape control are employed to ameliorate MoO3 gas sensing capability. Doping is one of these effective techniques optimizing the optical and electrical properties of MoO3 gas sensing. The ionic radius of Fe element is different than Mo element it generates some lattice deformation and oxygen vacancy formation which improve the gas response [29,30,31]. Fabrication of thin films using suitable technique is also important for producing suitable sensor material. Among different thin film coating techniques, nebulizer spray pyrolysis (NSP) offers various advantages in order to prepare homogeneous thin films [32]. Comparatively, NSP is a low-cost, eco-friendly and easy technique that can construct thin films with high growth rate in a non-vacuum atmosphere as well as it can deposit nanostructured material on the large glass substrate. Few previous studies have reported metal oxide thin film preparation using NSP and related techniques [33,34]. In this work MoO3:Fe thin films were fabricated using NSP technique, and its sensitivity towards the ammonia (NH3) gas is tested.

2. Materials and Methodology

Fe-doped MoO3 thin films were prepared by nebulizer spray pyrolysis method. In a typical thin film coating process, 0.005 M of ammonium heptamolybdate tetrahydrate ((NH4)6Mo7O24·4H2O) and a certain amount of Fe(NO3)3·9H2O (0, 1, 2, 3 and 4 wt.%) were dissolved in deionized water and stirred for 10 min. The as-prepared precursor was coated on glass substrate at temperature of 450 °C. The distance between substrate and spray gun maintained at 7 mm and the rate of flow of precursor solution is 1 mL/min. The coated films were kept at the coating temperature for 30 min. The obtained Fe-doped MoO3 thin films were subjected to know structural, optical, electrical and gas sensing properties. The micro structural spectacles of the fabricated samples were studied using a Pan-analytical X’Pert PRO X-ray diffraction (XRD) system. The surface morphology and surface coarseness were determined by field emission scanning electron microscopy (FESEM). A UV–vis–NIR spectrophotometer (Perkin Elmer Lambda, Waltham, MA, USA) was used to determine the transmittance, absorption spectra. A PerkinElmer LS55 fluorescence spectrophotometer with an excitation wavelength of 325 nm was utilized to detect the photoluminescence spectra at room temperature. A Keithley Source Meter (model 2450) was used to sense ammonia gas at ambient temperature (27 °C). The electrical current of the as-prepared thin films was obtained in both air and ammonia atmosphere.

Gas Sensing Set Up

The Fe-doped MoO3 films were prepared, and the concentration of ammonia can be obtained using formula given in Equation (1) [25].
C ppm = δ VRT MP b V b × 10 6
Here, δ represents the testing gas density (g mL−1), T is the temperature of the sensor (K), R is the universal gas constant (8.415 J mol−1 K−1), Pb is the pressure in the chamber, Vb is the chamber volume in liters, V is the volume of the infused gas (mL), M specifies the molecular weight of the test gas (g mol−1). Here the sensor which is to be tested by placing the films inside the chamber.
Then, an adequate amount of ammonia was taken in to the chamber through a metallic beaker which is connected to a heating coil and a thermocouple was used to detect the temperature in the chamber. The ammonia vapor was produced by heating the ammonia solution up to the necessary temperature. As a result, the films began to sense the ammonia at room temperature, and the response current was recorded using a Keithley Source Meter. Once the response current reached to saturation condition, the chamber holding the films was opened and exposed to the atmospheric air. The procedure was continued for a range of NH3 concentrations. The metal contacts to the films are prepared using the silver (Ag) paste over the films surface for measuring the current response for all the variations in NH3 concentration. Finally, the overall current response towards the analyte gas was measured from the ratio of current in air (Ia) condition and the current in the test gas (Ig) condition.

3. Result and Discussion

3.1. XRD Analysis

Figure 1, shows the XRD pattern of MoO3:Fe (0, 1, 2, 3 and 4%) thin films deposited using the nebulizer spray pyrolysis technique. It is noticeable from the figure, that all films display sharp peaks matched to the JCPDS card data 47–1320, of monoclinic structure [35]. It was clearly observed that the intensity of all the peaks of MoO3 varied due to the inclusion of Fe dopant, moreover the intensity of (002) plane increases with the increment in concentration of Fe dopants. This indicates that the crystallinity increases with respect to the incorporation of Fe dopants in MoO3. However, the further rise in the Fe dopants beyond the optimal level i.e., at 4% associated with a slight decline in the intensity of the (002) plane which might be originated by the domination of Fe ions in the target material MoO3 and the increase of lattice distortion and strain in MoO3.
The average crystallite sizes are calculated via the Debye Scherrer formula and the strain ε was calculated using Equations (2) and (3), respectively.
D   =   0.9   λ β C o s θ  
ε = β   C o t   θ 4
where β depicts the full width half maximum (FWHM), λ denotes the wavelength of X-ray and θ is Bragg’s angle.
In Table 1, the crystallite size of MoO3:Fe (0, 1, 2, 3 and 4%) are shown, in which the crystallite size augments from 70 to 82 nm and greater concentration of Fe ion on the MoO3 matrix decreases to the size of 75 nm. This might be due to substitution of higher concentration of Fe atoms integrated in to the MoO3 lattice which distresses the crystallite size with the generation of large number of nucleation centers and increase in the lattice strain of the fabricated thin films [36]. Lattice constant values changed due to Fe doping and cell volume of the MoO3 sample increases up to 3% of Fe doping owning to the higher ionic radius of Fe2+ compared with host element of Mo6+.

3.2. Surface Morphologyanalysis

The MoO3 nanostructures possess larger surface area, so they can exposed to absorb more gases, and the porosity of the material surface is favorable to support the gas diffusion, thereby enhancing the exploitation of sensing mechanism, and eventually improving its sensitivity [37]. Figure 2a–e, shows the nanobelts morphology with some nano rods over the MoO3:Fe (0, 1, 2, 3 and 4%) thin films. In 3% Fe doped MoO3 film has nano rods with porosity which is clearly visible in Figure 2d. Further increase in the dopant concentration i.e., Fe (4%), the high porosity gets decreased, which would reduce the gas sensing response. Figure 2f, shows EDX pattern of as prepared thin film, which confirms the presence of Mo, Fe and O in the final deposited film.

3.3. Optical Property Analysis

Figure 3 represents the transmission spectra of MoO3:Fe (0, 1, 2, 3 and 4%) thin films observed within the wavelength range 300–900 nm. It has been noticed that the transparency of films increases with the increase of dopant concentration and reaches maximum transparency for MoO3:Fe (3%) due to the enlarged voids generated in the thin films, which can be identified through the SEM image. This is ascribed to the structural deformations occurred by the amalgamation of Fe2+ ions which substitutes the Mo6+ ions from the host MoO3 matrix [38]. The diminution in transmittance for Fe dopants (1, 2, and 4%) are due to the amplified scattering of photons originated by the crystal defects, where the voids are less compared to the Fe (3%) dopant concentration [39,40].
Figure 3 illustrates the Tauc’s plot for obtaining the optical band gap (Eg) of the fabricated thin films MoO3:Fe (0, 1, 2, 3, and 4%) calculated using Equation (4) [41].
α h v 2 = A ( h v E g )
where denotes the photon energy, A indicates a constant for a direct transition, and α represents the optical absorption coefficient, the energy gap of the developed thin films was obtained through hv vs. (αhv)2 plot by extending the linear segment of the curve to the energy axis. The optical band gap of MoO3:Fe doped thin film shows decreasing from 3.13 to 2.79 eV for the doped concentration Fe (0%) to Fe (3%) dopant ions on the MoO3 matrix. On further rise of doping concentration from Fe (3%) to Fe (4%) the band gap of MoO3 tends to increase to 2.82 eV which was similar to the earlier reported values [42]. The increase in carrier concentration causes the narrowing of band gap as the concentration of Fe doping increases, which is related to the raise in film defects. According to the Burstein–Moss effect, as the carrier concentration increases from Fe (0 to 4%), the lowest energy state of the conduction band is filled with the electrons, thus a red shift is observed in the Fermi level while doping from Fe (0%) to Fe (3%), and a blue shift is observed for Fe (4%) film, due to the increase in band gap, as a result of the heavily doped Fe2+ ion on the host MoO3 [40,41,42,43].

3.4. Photoluminescence Analysis

Figure 4 represents the photoluminescence spectra, obtained at room temperature within the wavelength range of 300 to 600 nm. It is well known that all the emission peaks like 396, 416, 441, 453, 480, 523 nm corresponds to transitions among the diverse intrinsic defect states within the band gap of MoO3:Fe thin films [44]. Nevertheless, the change in morphology and crystallite size caused by Fe doping has a noteworthy effect on the emissivity of as-prepared thin films. It can be clarified that the broad PL spectra shows the event of UV emission and visible light emission from MnO3:Fe film. Here the PL spectra exhibit two emission bands such as UV (389 nm) and visible [Violet (416 nm), blue (441 nm) and bluish green (453 nm and 480 nm) green (523 nm)]. The UV emissions are due to the well-built near-band edge emission (NBE) owing to the free-exciton recombination in the fabricated thin films and, similarly visible light emissions are owed to the changeover of excited optical centers into the deep levels. These deep level emissions are due to the existence of impurities as dopants in the host MoO3, along with the surface defects, and its potency is positively allied with the defect density. Thus, in the as-prepared thin films shows the enhanced crystallinity leads to an increase in the strength of band-edge emission [42]. The emission peaks such as 389, 416, 441, 453 and 480 nm bands of Fe (3%) shows red shift towards a longer wavelength, similarly a blue shift is observed for the higher dopant concentrations towards the shorter wavelength compared to Fe (0%) is due to the quantum confinement, followed by the widening of band gap with the decrease in particle size with respect to the dopant concentration. The green emission was observed at 523 nm, which is ascribed to the transitions sandwiched between the electrons close to the conduction band and the deep spellbound holes which are due to the ionized oxygen vacancies. This green emission is also credited to doubly charged oxygen vacancy centers situated at the surface of the thin film [45]. The intensity of Fe (3%) of NBE is higher than the visible deep level states, due to the fact that surface defects, lattice defects and stoichiometric ratio of as-prepared thin films, which generates more oxygen vacancies in the host material. The improvement of PL intensity in case of Fe (3%) dopant concentration is due the rise of porosity and roughness due ti increase in the surface areathereby producing an energetic response towards the incoming ultraviolet radiation. On other hand the smooth surfaced thin films shows specular reflection, but a coarse nature of thin film will permit the excitation light to be reflected onto a different segment of the film surface, hence it results highly well-organized utilization of the excitation energy and consequently possess a higher PL intensity [42].

3.5. Gas Sensing

In wide range of semiconducting materials, the electrical spectacles of thin films may change when definite amount of gas molecules are adsorbed on its surface. Here the sensing properties of MoO3:Fe (0, 1, 2, 3 and 4%) thin films can be identified as the function of electrical current to the NH3 with different exposed on the film surface as mentioned in schematic Figure 5. Since the extent of sensing depends on the interaction between MoO3:Fe thin films and surrounding gases. Typically the sensing of NH3, gas molecules related to the surface coarseness, crystallite size, the absorptive nature of the surface, dopants concentration and surface defects of the as-prepared thin films affect the sensing response [37,46]. In the prior reports, the sensing efficiency of Fe doped MoO3 materials towards ammonia gases is higher due to the catalytic consequence of large surface area [37]. MoO3:Fe sensing behavior depends on the composition, crystallite size, surface morphology, which is enhanced by the impact of Fe dopants in MoO3 nanostructures [47,48]. In case of critical surface state density, the total carrier concentration within the grains, decreases with the crystallite size. Because of bigger grains observed in Fe (3%) on MoO3 thin films, increases the amount of charge carriers which is responsible for the conductivity mechanism, and the responsivity reaches 18,900% which is higher than the earlier reported values [49,50], along with the raise or ON state occurs at 51 s and decay or OFF state at 8 s.
From the Figure 6, it was observed higher response for Fe (3%) film than with other doped films of Fe (1%), Fe (2%), Fe (4%) and the previous reported values [51]. On other hand, it was noticed that for the certain number of surface state density, the net carrier concentration decreased in the case of 1, 2 and 4% Fe doped MoO3 film compared to the 3% Fe doped film. Figure 6 also shows that the electrical resistance of MoO3:Fe thin films varies as the function of ammonia (NH3) gas concentrations of 50–250 ppm at room temperature. The NH3 gas sensitivities are drastically increased, with the rise in concentration. The optimum NH3 concentration is found to be 250 ppm for the as- prepared thin films of MoO3:Fe (0, 1, 2, 3, and 4%). PL spectra revealed that MoO3:Fe (3%) film has more oxygen vacancies than the other films. Consequently, the doping process appreciably improves the gas sensing properties of MoO3:Fe nanostructures.
The film sensing response is calculated from the ratio of the electrical current of the film in air to that of the film exposed to the NH3 gas [52]:
S % = I g I a I a
where S, Ia and Ig stands for the sensitivity, film current in air (darkness) and film current in the presence of NH3 gas. Figure 7, shows the disparity of sensing reaction with different NH3 gas concentrations for the as-prepared thin films. It can be seen from Table 2 that the response (%) for 250 ppm of maximum concentration of NH3 gas was 2620% for MoO3:Fe (0%), with response time of 62 s for raise or on state, and 8s decay or off state of gas sensor. This response (%) probably increased while adding dopants of about Fe (3%), resulting in maximum value of 38,500% with short period of raise and fall time of about 54 s and 6 s. In this case, for Fe (3%) on the MoO3 host reaches proper neutral charge vacancy, thereby creating high level of oxygen vacancies in the system, but for the other dopant concentrations like Fe (1, 2 and 4%), due to the mismatch of stochiometric ratio, proper oxygen vacancies are not allowed in to the host system, hence the Fe (3%) shows good sensing response than the other films with high response current [53]. This might be because of the better crystallinity, surface area, adsorption and defect sites sensors [54,55,56]. Herein, we have compared the current ammonia gas sensing results of MoO3 sample with other published work and the comparison table is displayed in Table 3. Andreas et al. [57] fabricated α-MoO3 sample by Si-doping via spray pyrolysis technique exhibited notable response and recovery time of 3.6 min/7 min for 400 ppm ammonia concentration. The MoO3 sample synthesized using sol–gel technique by Prasad et al. [58] displayed reasonable raise and decay speed of 15 s/20 s, but the sample operating temperature was 500 °C. Sunu et al. [59] readied Pulsed laser deposition of MoO3 thin films as ammonia sensor at 280 °C working temperature showed response and recovery speed of 15 s/300 s. Recently Neha et al. [60] and Chua et al. [61] manufactured 2D hexagonal MoO3 and h-MoO3 samples responded towards 5 ppm and 500 ppm ammonia gas limit at room temperature with the improved response and recovery speed of 69 s/129 s and 210 s/241 s respectively. The sample prepared in the present work i.e., Fe (3%) doped MoO3 sample displayed better ammonia gas response (250 ppm) at room temperature with smaller response and recovery time of 54 s/6 s.
The stability and the reproducibility preside over the device excellence. The film MoO3:Fe (3%), shows good sensing performance, therefore, it is investigated for its performance after 2 months. Importantly, the film was stored under atmospheric condition. Figure 8, shows the reproducibility of the sensing performance of MnO3:Fe (3%)film was used after a couple of months which is stored in laboratory conditions are checked for its performance stability, by taking 5 cycles each time for the maximum 250 ppm NH3. Therefore, it is clearly understood that the electrical current of the newly fabricated film and the old film shows identical behavior response as the freshly prepared thin film. It predicts that the NH3 gas molecules interaction with MoO3:Fe (3%) is not affected by any of the atmospheric condition. From this inspection, it is understood that the stability and lifetime of the MoO3:Fe (3%) prepared film shows better enhancement in the gas sensing application even after aging its stability remains constant.
Cross-sensitivity of the as-prepared sensors was checked by other gases and the gas response values of the fabricated films are 216, 364, 412 and 2089 for isopropanol, ethanol, methanol and ammonia. It is observed that the film shows maximum response to ammonia compared to isopropanol, ethanol, and methanol. Such a high response of ammonia could be assigned to its electron donating ability which is elevated compared to other tested gases due to the presence of lone pair of electrons.
Typically, a number of mechanisms can be well thought-out for the sensing response of MoO3 thin films with regards to ammonia gas. It is noteworthy to mention that the presence and enhancement of oxygen vacancies seems to be one of the major reasons which modifies the electrical properties of the as-prepared sensor. Since the oxygen could be simply adsorbed on the surface of the film, when exposed to gas or even heating at room temperature. The adsorbed oxygen on the surface of the film, when exposed to air leads to the abduction electrons tends to form O2 and O, which relies on the temperature of the film. O2 is typically produced below 100 °C, whereas O is produced between 100 and 300 °C, and O2 is produced beyond 300 °C [53]. These different species are produced when ambient oxygen chemically adsorbs to the film surface, eliminating the electrons that were previously present in the conduction band [54]. From the earlier, at room temperature, MoO3:Fe films with 0, 1, 2, 3, and 4% Fe content may have formed O2 species. The chemisorbed oxygen increases the films current when it reacts to being exposed to incoming NH3 gas. MoO3 is an n-type semiconducting material with a wide bandgap, consequently a large number of oxygen can be adsorbed onto surface of thin films in atmospheric conditions. Furthermore, some of the precise type gases can interrelate greatly with the oxygen of MoO3 and thus alter the electrical response of the film towards the incoming gas. Since the defects created by the donor like oxygen vacancies and interstitials of Fe on MoO3 are the most constructive path for the oxygen species, and thus enhancing the probability of interface effect to the NH3 molecules. Besides, when the donor content increases, which lead to the formation of defects, more oxygen molecules get adsorbed the film surface and can respond to the NH3 gas, thus enriching the gas sensing response of the prepared thin film [55,56]. The interaction of gas with MoO3 is expressed by the following expression using Equations (6) and (7):
NH 3 ads + O ads   NH 3 ( ads )
2 NH 3 ads + 3 O N 2 ads + 3 H 2 O + 3 e
Since the straight forward absorption and desorption of NH3 gas and the oxygen molecule are given by Equations (8) and (9).
NH 3 gas NH 3 ads
and
O 2 ads O 2 ads
hence, the negatively charged O and O2 are formed on the film surface, using Equation (10):
O 2 ads + e 2 O ads 2
From the above metioned relations, it is noted that the discharge of electrons spellbound in the film surface, which causes the increase in the response current when NH3 molecules responds to the O species. The maximum gas response of the samples depends on two factors: the quantity of electrons trapped from the conduction band through the process of adsorption of oxygen species, and the quantity of electrons released by the adsorbed oxygen species by reaction with ammonia gas. These pathways allow for the complete release of all trapped electrons from the adsorbed oxygen species, demonstrating the critical function that ammonia gas can play in the exchange of electrons. When ammonia interacts with the MoO3:Fe film surface, the electron depletion region, grain boundaries, and Fermi energy (EF) are also reduced. The schematic ammonia gas sensing diagram of MoO3 film is shown in Figure 9. The speedy response to ammonia gas can also be attributed to the distribution of porous nanobelts with greater surface coarseness with the increased surface area while going towards the higher dopant concentration. On the other hand, the switch over of electrons in the Fe2+ ions with MoO3 are responsible for the decrease in the oxidation state, and thereby increasing the depletion region, resulting in the enhanced of the sensor response for MoO3:Fe (3%).

4. Conclusions

This work aims to prepare reproducible chemo resistive gas sensors made ofMoO3:Fe (0, 1, 2, 3 and 4%) thin film sensor. Here four different concentrations of Fe are used as a dopant on MoO3 matrix, which was successfully synthesized and confirmed the monoclinic phase of MoO3 thin films through XRD. FESEM analysis revealed that the 3 % Fe doped MoO3 film possess nano belt morphology with some porosity which enhances the gas sensing behaviour. The MoO3:Fe (3%) thin film shows better transparency and least bandgap value of 2.79 eV. PL shows visible, blue and green emission doubly charged oxygen vacancy centers situated at the surface of the as-prepared thin film, which makes it superior to enrich the gas sensing response. The gas sensing response of the MoO3:Fe (3%) thin film showed a drastically improved responsivity with short period of on/off response to the 50 ppm to 250 ppm of NH3 gas. Thus, MoO3:Fe (3%) thin film could be better suited as gas sensing material for room temperature ammonia sensing.

Author Contributions

F.H.A., A.B.G.T. and M.S.: Investigation, Methodology, Conceptualization, Data curation, Formal analysis, Writing—original draft, F.H.A., A.B.G.T. and S.A.: Funding acquisition, Project administration, Resources, Software, Validation, Visualization, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R223), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia. The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through a Research Groups Program under Grant No. R.G.P. 2/249/43.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also forms part of an ongoing study.

Acknowledgments

This research was funded by Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R223), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia. The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through a Research Groups Program under Grant No. R.G.P. 2/249/43.

Conflicts of Interest

The authors declared that there is no conflict of interest.

References

  1. Meixner, H.; Lampe, U. Metal oxide sensors. Sens. Actuators B Chem. 1996, 33, 198–202. [Google Scholar] [CrossRef]
  2. Neri, G. First fifty years of chemoresistive gas sensors. Chemosensors 2015, 3, 1–20. [Google Scholar] [CrossRef]
  3. Kukkola, J.; Mohl, M.; Leino, A.R.; Tóth, G.; Wu, M.C.; Shchukarev, A.; Popov, A.; Mikkola, J.P.; Lauri, J.; Riihimäki, M.; et al. Inkjet-printed gas sensors: Metal decorated WO 3 nanoparticles and their gas sensing properties. J. Mater. Chem. 2012, 22, 17878–17886. [Google Scholar] [CrossRef]
  4. Kukkola, J.; Mohl, M.; Leino, A.R.; Mäklin, J.; Halonen, N.; Shchukarev, A.; Konya, Z.; Jantunen, H.; Kordas, K. Room temperature hydrogen sensors based on metal decorated WO3 nanowires. Sens. Actuators B Chem. 2013, 186, 90–95. [Google Scholar] [CrossRef]
  5. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2018, 229, 206–217. [Google Scholar] [CrossRef]
  6. Wang, Y.; Zhou, Y.; Wang, Y.; Zhang, R.; Li, J.; Li, X.; Zang, Z. Conductometric room temperature ammonia sensors based on titanium dioxide nanoparticles decorated thin black phosphorus nanosheets. Sens. Actuators B Chem. 2021, 349, 130770. [Google Scholar] [CrossRef]
  7. Aponte-Luis, J.; Gómez-Galán, J.A.; Gómez-Bravo, F.; Sánchez-Raya, M.; Alcina-Espigado, J.; Teixido-Rovira, P.M. An Efficient Wireless Sensor Network for Industrial Monitoring and Control. Sensors 2018, 18, 182. [Google Scholar] [CrossRef]
  8. Somov, A.; Baranov, A.; Savkin, A.; Spirjakin, D.; Spirjakin, A.; Passerone, R. Development of wireless sensor network for combustible gas monitoring. Sens. Actuators A Phys. 2011, 171, 398–405. [Google Scholar] [CrossRef]
  9. Järvinen, T.; Lorite, G.S.; Rautio, A.-R.; Juhász, K.L.; Kukovecz, Á.; Kónya, Z.; Kordas, K.; Toth, G. Portable cyber-physical system for indoor and outdoor gas sensing. Sens. Actuators B Chem. 2017, 252, 983–990. [Google Scholar] [CrossRef]
  10. Akhtar, F.; Rehmani, M.H. Energy replenishment using renewable and traditional energy resources for sustainable wireless sensor networks: A review. Renew. Sustain. Energy Rev. 2015, 45, 769–784. [Google Scholar] [CrossRef]
  11. Kuc, A. Low-dimensional transition-metal dichalcogenides. In Chemical Modelling; Royal Society of Chemistry: London, UK, 2015; Volume 11, pp. 1–29. [Google Scholar]
  12. Li, B.L.; Wang, J.; Zou, H.L.; Garaj, S.; Lim, C.T.; Xie, J.; Li, N.B.; Leong, D.T. Low-Dimensional Transition Metal Dichalcogenide Nanostructures Based Sensors. Adv. Funct. Mater. 2016, 26, 7034–7056. [Google Scholar] [CrossRef]
  13. Kim, T.H.; Kim, Y.H.; Park, S.Y.; Kim, S.Y.; Jang, H.W. Two-dimensional transition metal disulfides for chemoresistive gas sensing: Perspective and challenges. Chemosensors 2017, 5, 15. [Google Scholar] [CrossRef]
  14. Eftekhari, A. Tungsten dichalcogenides (WS2, WSe2, and WTe2): Materials chemistry and applications. J. Mater. Chem. A 2017, 5, 18299–18325. [Google Scholar] [CrossRef]
  15. Hasani, A.; Tekalgne, M.; Van Le, Q.; Jang, H.W.; Kim, S.Y. Two-dimensional materials as catalysts for solar fuels: Hydrogen evolution reaction and CO2 reduction. J. Mater. Chem. A 2019, 7, 430–454. [Google Scholar] [CrossRef]
  16. Shokri, A.; Salami, N. Gas sensor based on MoS2 monolayer. Sens. Actuators B Chem. 2016, 236, 378–385. [Google Scholar] [CrossRef]
  17. Hung, C.M.; Vuong, V.A.; Van Duy, N.; Van An, D.; Van Hieu, N.; Kashif, M.; Hoa, N.D. Controlled Growth of Vertically Oriented Trilayer MoS2 Nanoflakes for Room-Temperature NO2 Gas Sensor Applications. Phys. Status Solidi Appl. Mater. Sci. 2020, 217, 2000004. [Google Scholar] [CrossRef]
  18. Yu, M.; Shao, H.; Wang, G.; Yang, F.; Liang, C.; Rozier, P.; Wang, C.Z.; Lu, X.; Simon, P.; Feng, X. Interlayer gap widened α-phase molybdenum trioxide as high-rate anodes for dual-ion-intercalation energy storage devices. Nat. Commun. 2020, 11, 1348. [Google Scholar] [CrossRef]
  19. Liu, T.; Liu, Z. 2D MoS2 Nanostructures for Biomedical Applications. Adv. Healthc. Mater. 2018, 7, 1701158. [Google Scholar] [CrossRef]
  20. Krishnamoorthy, K.; Veerapandian, M.; Yun, K.; Kim, S.J. New function of molybdenum trioxide nanoplates: Toxicity towards pathogenic bacteria through membrane stress. Colloids Surf. B Biointerfaces 2013, 112, 521–524. [Google Scholar] [CrossRef]
  21. Jiao, Y.; Hafez, A.M.; Cao, D.; Mukhopadhyay, A.; Ma, Y.; Zhu, H. Metallic MoS2 for High Performance Energy Storage and Energy Conversion. Small 2018, 14, 1800640. [Google Scholar] [CrossRef]
  22. Cheng, J.; Wang, C.; Zou, X.; Liao, L. Recent Advances in Optoelectronic Devices Based on 2D Materials and Their Heterostructures. Adv. Opt. Mater. 2019, 7, 1800441. [Google Scholar] [CrossRef]
  23. Singh, E.; Singh, P.; Kim, K.S.; Yeom, G.Y.; Nalwa, H.S. Flexible Molybdenum Disulfide (MoS2) Atomic Layers for Wearable Electronics and Optoelectronics. ACS Appl. Mater. Interfaces 2019, 11, 11061–11105. [Google Scholar] [CrossRef] [PubMed]
  24. Yan, H.; Song, P.; Zhang, S.; Zhang, J.; Yang, Z.; Wang, Q. Au nanoparticles modified MoO3 nanosheets with their enhanced properties for gas sensing. Sens. Actuators B Chem. 2016, 236, 201–207. [Google Scholar] [CrossRef]
  25. Zhang, S.; Song, P.; Zhang, J.; Li, Z.; Yang, Z.; Wang, Q. In2O3-functionalized MoO3 heterostructure nanobelts with improved gas-sensing performance. RSC Adv. 2016, 6, 50423–50430. [Google Scholar] [CrossRef]
  26. Yan, H.; Song, P.; Zhang, S.; Yang, Z.; Wang, Q. Facile fabrication and enhanced gas sensing properties of hierarchical MoO3 nanostructures. RSC Adv. 2015, 5, 72728–72735. [Google Scholar] [CrossRef]
  27. de Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalantar-Zadeh, K. Molybdenum Oxides—From Fundamentals to Functionality. Adv. Mater. 2017, 29, 1701619. [Google Scholar] [CrossRef]
  28. Luo, X.; You, K.; Hu, Y.; Yang, S.; Pan, X.; Wang, Z.; Chen, W.; Gu, H. Rapid hydrogen sensing response and aging of A-MoO3 nanowires paper sensor. Int. J. Hydrogen Energy 2017, 42, 8399–8405. [Google Scholar] [CrossRef]
  29. Wang, S.; Xie, J.; Hu, J.; Qin, H.; Cao, Y. Fe-doped α-MoO3 nanoarrays: Facile solid-state synthesis and excellent xylene-sensing performance. Appl. Surf. Sci. 2020, 512, 145722. [Google Scholar] [CrossRef]
  30. Yang, S.; Lei, G.; Tan, L.; Xu, H.; Xiong, J.; Wang, Z.; Gu, H. Fe-doped MoO3 nanoribbons for high-performance hydrogen sensor at room temperature. J. Alloys Compd. 2021, 877, 160200. [Google Scholar] [CrossRef]
  31. Ouyang, Q.Y.; Li, L.; Wang, Q.S.; Zhang, Y.; Wang, T.S.; Meng, F.N.; Chen, Y.J.; Gao, P. Facile synthesis and enhanced H2S sensing performances of Fe-doped α-MoO3 micro-structures. Sens. Actuators B Chem. 2012, 169, 17–25. [Google Scholar] [CrossRef]
  32. Arun Kumar, K.D.; Ganesh, V.; Valanarasu, S.; Shkir, M.; Kulandaisamy, I.; Kathalingam, A.; AlFaify, S. Effect of solvent on the key properties of Al doped ZnO films prepared by nebulized spray pyrolysis technique. Mater. Chem. Phys. 2018, 212, 167–174. [Google Scholar] [CrossRef]
  33. Rao, C.N.R.; Biswas, K. Essentials of Inorganic Materials Synthesis; Wiley: Hoboken, NJ, USA, 2015. [Google Scholar]
  34. Arun Kumar, K.D.; Valanarasu, S.; Ponraj, J.S.; Fernandes, B.J.; Shkir, M.; AlFaify, S.; Murahari, P.; Ramesh, K. Effect of Er doping on the ammonia sensing properties of ZnO thin films prepared by a nebulizer spray technique. J. Phys. Chem. Solids 2020, 144, 109513. [Google Scholar] [CrossRef]
  35. Bhojanaa, K.B.; Kannadhasan, S.; Santhosh, N.; Vijayakumar, P.; Senthil Pandian, M.; Ramasamy, P.; Pandikumar, A. Enhanced electrochemical and photovoltaic performance for MoO3 nanorods at different calcination temperature based counter electrode in Pt-free dye-sensitized solar cells applications. SN Appl. Sci. 2020, 2, 1750. [Google Scholar] [CrossRef]
  36. Garcés, F.A.; Budini, N.; Koropecki, R.R.; Arce, R.D. Structural Analysis of ZnO(:Al,Mg) Thin Films by X-ray Diffraction. Procedia Mater. Sci. 2015, 8, 551–560. [Google Scholar] [CrossRef]
  37. Guan, W.; Tang, N.; He, K.; Hu, X.; Li, M.; Li, K. Gas-Sensing Performances of Metal Oxide Nanostructures for Detecting Dissolved Gases: A Mini Review. Front. Chem. 2020, 8, 76. [Google Scholar] [CrossRef] [PubMed]
  38. Sahoo, B.; Behera, D.; Pradhan, S.K.; Mishra, D.K.; Sahoo, S.K.; Nayak, R.R.; Sekhar, K.P.C. Analysis of structural, optical and electrical properties of nano-particulate indium doped zinc oxide thin films. Mater. Res. Express 2019, 6, 1150a6. [Google Scholar] [CrossRef]
  39. Shinde, S.S.; Shinde, P.S.; Bhosale, C.H.; Rajpure, K.Y. Optoelectronic properties of sprayed transparent and conducting indium doped zinc oxide thin films. J. Phys. D Appl. Phys. 2008, 41, 105109. [Google Scholar] [CrossRef]
  40. Jellal, I.; Ahmoum, H.; Khaaissa, Y.; Nouneh, K.; Boughrara, M.; Fahoume, M.; Chopra, S.; Naja, J. Experimental and ab-initio investigation of the microstructure and optoelectronic properties of FCM–CVD-prepared Al-doped ZnO thin films. Appl. Phys. A Mater. Sci. Process. 2019, 125, 650. [Google Scholar] [CrossRef]
  41. Abinaya, M.; Saravanakumar, K.; Jeyabharathi, E.; Muthuraj, V. Synthesis and Characterization of 1D-MoO3 Nanorods Using Abutilon indicum Extract for the Photoreduction of Hexavalent Chromium. J. Inorg. Organomet. Polym. Mater. 2019, 29, 101–110. [Google Scholar] [CrossRef]
  42. Navas, I.; Vinodkumar, R.; Mahadevan Pillai, V.P. Self-assembly and photoluminescence of molybdenum oxide nanoparticles. Appl. Phys. A Mater. Sci. Process. 2011, 103, 373–380. [Google Scholar] [CrossRef]
  43. Saw, K.G.; Aznan, N.M.; Yam, F.K.; Ng, S.S.; Pung, S.Y. New insights on the burstein-moss shift and band gap narrowing in indium-doped zinc oxide thin films. PLoS ONE 2015, 10, e0141180. [Google Scholar] [CrossRef] [PubMed]
  44. Tan, S.T.; Chen, B.J.; Sun, X.W.; Fan, W.J.; Kwok, H.S.; Zhang, X.H.; Chua, S.J. Blueshift of optical band gap in ZnO thin films grown by metal-organic chemical-vapor deposition. J. Appl. Phys. 2005, 98, 013505. [Google Scholar] [CrossRef]
  45. Thu Do, T.A.; Giang, H.T.; Van Huong, D.; Ngan, P.Q.; Thai, G.H.; Thu, D.T.; Lam, T.D. Correlation between photoluminescence spectra with gas sensing and photocatalytic activities in hierarchical ZnO nanostructures. RSC Adv. 2017, 7, 9826–9832. [Google Scholar] [CrossRef]
  46. Bhati, V.S.; Hojamberdiev, M.; Kumar, M. Enhanced sensing performance of ZnO nanostructures-based gas sensors: A review. Energy Rep. 2020, 6, 46–62. [Google Scholar] [CrossRef]
  47. Sarf, F. Metal Oxide Gas Sensors by Nanostructures. In Gas Sensors; IntechOpen: London, UK, 2020. [Google Scholar]
  48. Fan, C.; Sun, F.; Wang, X.; Huang, Z.; Keshvardoostchokami, M.; Kumar, P.; Liu, B. Synthesis of ZnO hierarchical structures and their gas sensing properties. Nanomaterials 2019, 9, 1277. [Google Scholar] [CrossRef]
  49. Kumar, A.; Kumar, M.; Kumar, R.; Singh, R.; Prasad, B.; Kumar, D. Gas sensitivity as a function of nano sized crystallite in metal oxide semiconductor based gas sensors. Mater. Today Proc. 2019, 17, 161–167. [Google Scholar] [CrossRef]
  50. Tu, Y.; Kyle, C.; Luo, H.; Zhang, D.W.; Das, A.; Briscoe, J.; Dunn, S.; Titirici, M.M.; Krause, S. Ammonia Gas Sensor Response of a Vertical Zinc Oxide Nanorod-Gold Junction Diode at Room Temperature. ACS Sens. 2020, 5, 3568–3575. [Google Scholar] [CrossRef]
  51. Nechepurenko, I.A.; Kulikova, D.P.; Kornienko, V.V.; Afanasiev, K.N.; Shekoyan, L.A.; Baryshev, A.V.; Dorofeenko, A.V. Evaluating the response time of an optical gas sensor based on gasochromic nanostructures. Sensors 2021, 21, 8472. [Google Scholar] [CrossRef]
  52. Alenezi, M.R.; Alshammari, A.S.; Jayawardena, K.D.G.I.; Beliatis, M.J.; Henley, S.J.; Silva, S.R.P. Role of the exposed polar facets in the performance of thermally and UV activated ZnO nanostructured gas sensors. J. Phys. Chem. C 2013, 117, 17850–17858. [Google Scholar] [CrossRef]
  53. Vaishnav, V.S.; Patel, P.D.; Patel, N.G. Indium tin oxide thin film gas sensors for detection of ethanol vapours. Thin Solid Film. 2005, 490, 94–100. [Google Scholar] [CrossRef]
  54. Shingange, K.; Tshabalala, Z.P.; Ntwaeaborwa, O.M.; Motaung, D.E.; Mhlongo, G.H. Highly selective NH3 gas sensor based on Au loaded ZnO nanostructures prepared using microwave-assisted method. J. Colloid Interface Sci. 2016, 479, 127–138. [Google Scholar] [CrossRef]
  55. Kulandaisamy, A.J.; Reddy, J.R.; Srinivasan, P.; Babu, K.J.; Mani, G.K.; Shankar, P.; Rayappan, J.B.B. Room temperature ammonia sensing properties of ZnO thin films grown by spray pyrolysis: Effect of Mg doping. J. Alloys Compd. 2016, 688, 422–429. [Google Scholar] [CrossRef]
  56. Bai, Y.; Zhao, C.; Guo, Q.; Zhang, J.; Hu, S.; Liu, J.; Hayat, T.; Alsaedi, A.; Tan, Z. Enhancing the electron blocking ability of n-type MoO3 by doping with p-type NiO for efficient nonfullerene polymer solar cells. Org. Electron. 2019, 68, 168–175. [Google Scholar] [CrossRef]
  57. Güntner, A.T.; Righettoni, M.; Pratsinis, S.E. Selective sensing of NH3 by Si-doped α-MoO3 for breath analysis. Sens. Actuators B Chem. 2016, 223, 266–273. [Google Scholar] [CrossRef]
  58. Prasad, A.; Kubinski, D.; Gouma, P. Comparison of sol–gel and ion beam deposited MoO3 thin film gas sensors for selective ammonia detection. Sens. Actuators B Chem. 2003, 93, 25–30. [Google Scholar] [CrossRef]
  59. Sunu, S.; Prabhu, E.; Jayaraman, V.; Gnanasekar, K.; Gnanasekaran, T. Gas sensing properties of PLD made MoO3 films. Sens. Actuators B Chem. 2003, 94, 189–196. [Google Scholar] [CrossRef]
  60. Sakhuja, N.; Jha, R.; Bhat, N. Facile green synthesis of 2D hexagonal MoO3 for selective detection of ammonia at room temperature. Mater. Sci. Eng. B 2021, 271, 115249. [Google Scholar] [CrossRef]
  61. Chua, W.H.; Yaacob, M.H.; Tan, C.Y.; Ong, B.H. Chemical bath deposition of h-MoO3 on optical fibre as room-temperature ammonia gas sensor. Ceram. Int. 2021, 47, 32828–32836. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Fe doped MoO3 thin films.
Figure 1. XRD patterns of Fe doped MoO3 thin films.
Nanomaterials 12 02797 g001
Figure 2. FESEM images of Fe doped MoO3 thin films (a) 0%, (b) 1%, (c) 2%, (d) 3%, (e) 4% of Fe and (f) EDX image of MoO3:Fe (3%) thin film.
Figure 2. FESEM images of Fe doped MoO3 thin films (a) 0%, (b) 1%, (c) 2%, (d) 3%, (e) 4% of Fe and (f) EDX image of MoO3:Fe (3%) thin film.
Nanomaterials 12 02797 g002
Figure 3. (a) Transmission spectra and Tauc’s plot of Fe doped MoO3 thin films (b) 0%, (c) 1%, (d) 2% (e) 3%, (f) 4% of Fe concentrations.
Figure 3. (a) Transmission spectra and Tauc’s plot of Fe doped MoO3 thin films (b) 0%, (c) 1%, (d) 2% (e) 3%, (f) 4% of Fe concentrations.
Nanomaterials 12 02797 g003
Figure 4. PL spectra of Fe doped MoO3 thin films. (a) 0%, (b) 1%, (c) 2%, (d) 3%, (e) 4% of Fe concentrations.
Figure 4. PL spectra of Fe doped MoO3 thin films. (a) 0%, (b) 1%, (c) 2%, (d) 3%, (e) 4% of Fe concentrations.
Nanomaterials 12 02797 g004
Figure 5. Schematic diagram of gas sensing setup for ammonia gas detection.
Figure 5. Schematic diagram of gas sensing setup for ammonia gas detection.
Nanomaterials 12 02797 g005
Figure 6. Time−Current of Fe doped MoO3 thin films (a) 0%, (b) 1%, (c) 2%,(d) 3% and (e) 4% of Fe concentrations.
Figure 6. Time−Current of Fe doped MoO3 thin films (a) 0%, (b) 1%, (c) 2%,(d) 3% and (e) 4% of Fe concentrations.
Nanomaterials 12 02797 g006
Figure 7. Gas response of Fe doped MoO3 thin films.
Figure 7. Gas response of Fe doped MoO3 thin films.
Nanomaterials 12 02797 g007
Figure 8. Cycle test of MoO3:Fe (3%) thin film.
Figure 8. Cycle test of MoO3:Fe (3%) thin film.
Nanomaterials 12 02797 g008
Figure 9. Schematic diagram of MoO3 gas sensing mechanism.
Figure 9. Schematic diagram of MoO3 gas sensing mechanism.
Nanomaterials 12 02797 g009
Table 1. Structural parameters and lattice constant values of Fe doped MoO3 thin films with different Fe doping concentrations.
Table 1. Structural parameters and lattice constant values of Fe doped MoO3 thin films with different Fe doping concentrations.
Fe Doping
Concentration
(%)
Crystallite
Size
(nm)
Strain
×10−3
Lattice ConstantsCell
Volume
3)
a (Å) b (Å) c (Å)
0704.403.9683.6847.060100.09
1724.243.9713.6867.073100.39
2783.903.9503.6847.165100.94
3823.753.9433.6917.225101.63
4754.113.9443.6917.237101.81
Table 2. Gas sensing parameters of Fe doped MoO3 thin films with different Fe doping concentrations.
Table 2. Gas sensing parameters of Fe doped MoO3 thin films with different Fe doping concentrations.
Fe Doping
Concentration
(%)
Response
(%)
Rise
Time (s)
Decay
Time (s)
02620628
112,700607
217,100607
338,500546
421,500599
Table 3. Comparison of ammonia sensing performances of Fe doped MoO3 thin films with other MoO3 based sensors.
Table 3. Comparison of ammonia sensing performances of Fe doped MoO3 thin films with other MoO3 based sensors.
MaterialOperating
Temperature
Concentration
(ppm)
Response
Time
Recovery
Time
Ref.
Si doped MoO3400 °C400 ppm3.6 min7 min[57]
MoO3500 °C-15 s20 s[58]
MoO3280 °C500 ppm15 s300 s[59]
MoO3RT5 ppm69 s129 s[60]
MoO3RT500 ppm210 s241 s[61]
Fe doped MoO3RT250 ppm54 s6 sPresent 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

Alkallas, F.H.; Ben Gouider Trabelsi, A.; Shkir, M.; AlFaify, S. Enhanced Room Temperature Ammonia Gas Sensing Properties of Fe-Doped MoO3 Thin Films Fabricated Using Nebulizer Spray Pyrolysis. Nanomaterials 2022, 12, 2797. https://doi.org/10.3390/nano12162797

AMA Style

Alkallas FH, Ben Gouider Trabelsi A, Shkir M, AlFaify S. Enhanced Room Temperature Ammonia Gas Sensing Properties of Fe-Doped MoO3 Thin Films Fabricated Using Nebulizer Spray Pyrolysis. Nanomaterials. 2022; 12(16):2797. https://doi.org/10.3390/nano12162797

Chicago/Turabian Style

Alkallas, Fatemah H., Amira Ben Gouider Trabelsi, Mohd Shkir, and Salem AlFaify. 2022. "Enhanced Room Temperature Ammonia Gas Sensing Properties of Fe-Doped MoO3 Thin Films Fabricated Using Nebulizer Spray Pyrolysis" Nanomaterials 12, no. 16: 2797. https://doi.org/10.3390/nano12162797

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