Next Article in Journal
Adaptive Expectation–Maximization-Based Kalman Filter/Finite Impulse Response Filter for MEMS-INS-Based Posture Capture of Human Upper Limbs
Next Article in Special Issue
Design and Fabrication of Biosensor for a Specific Microbe by Silicon-Based Interference Color System
Previous Article in Journal
Application of Python-Based Abaqus Secondary Development in Laser Shock Forming of Aluminum Alloy 6082-T6
Previous Article in Special Issue
Development of Fluorescent Sensors for Biorelevant Anions in Aqueous Media Using Positively Charged Quantum Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication of Tungsten Oxide Nanowalls through HFCVD for Improved Electrochemical Detection of Methylamine

1
Advanced Materials and Devices Laboratory, Department of Bio-Convergence Science, Jeonbuk National University, Jeongeup Campus, Jeongeup 56212, Republic of Korea
2
Environmental Engineering Laboratory, Department of Bioactive Material Sciences, Jeonbuk National University, Jeonju 54896, Republic of Korea
3
Department of Bio-Convergence Science, Jeonbuk National University, Jeongeup Campus, Jeongeup 56212, Republic of Korea
4
Graduate School of Integrated Energy-AI, Jeonbuk National University, Jeonju 54896, Republic of Korea
5
New & Renewable Energy Material Development Center (NewREC), Jeonbuk National University, Jeonbuk 56332, Republic of Korea
6
Department of JBNU-KIST Industry-Academia Convergence Research, Jeonbuk National University, Jeonju 54896, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Micromachines 2024, 15(4), 441; https://doi.org/10.3390/mi15040441
Submission received: 31 January 2024 / Revised: 10 March 2024 / Accepted: 21 March 2024 / Published: 26 March 2024
(This article belongs to the Special Issue Nanoparticle (Bio)sensing Platform)

Abstract

:
In this study, well-defined tungsten oxide (WO3) nanowall (NW) thin films were synthesized via a controlled hot filament chemical vapor deposition (HFCVD) technique and applied for electrochemical detection of methylamine toxic substances. Herein, for the thin-film growth by HFCVD, the temperature of tungsten (W) wire was held constant at ~1450 °C and gasification was performed by heating of W wire using varied substrate temperatures ranging from 350 °C to 450 °C. At an optimized growth temperature of 400 °C, well-defined and extremely dense WO3 nanowall-like structures were developed on a Si substrate. Structural, crystallographic, and compositional characterizations confirmed that the deposited WO3 thin films possessed monoclinic crystal structures of high crystal quality. For electrochemical sensing applications, WO3 NW thin film was used as an electrode, and cyclic voltammetry (CV) and linear sweep voltammetry (LSV) were measured with a wide concentration range of 20 μM~1 mM of methylamine. The fabricated electrochemical sensor achieved a sensitivity of ~183.65 μA mM−1 cm−2, a limit of detection (LOD) of ~20 μM and a quick response time of 10 s. Thus, the fabricated electrochemical sensor exhibited promising detection of methylamine with considerable stability and reproducibility.

1. Introduction

In today’s landscape, metal oxides have emerged as valuable assets across various industries, owing to their unique chemical, physical, and electronic characteristics. Metal oxides offer versatility and find applications in a wide range of sectors, including environmental remediation [1], medical technology [2], energy solutions [3], water purification [4], and personal care product development [5]. The use of metal oxide is anticipated to grow even further, underscoring their pivotal role in advancing innovation and addressing contemporary challenges [6].
Tungsten oxide (WO3) is a wide-band-gap metal oxide [7] which has good properties for numerous applications, e.g., information science, electronics (nano and micro), computer science, energy (renewable and non-renewable), transportation, safety engineering, military technologies, optoelectronic [8], electrochromic devices [9] and sensing [10]. WO3 is an n-type semiconductor which exhibits high stability, small diffusion length (150–500 nm) and good carrier mobility (~33.9 cm2 V−1 s−1) [11]. Tungsten oxide (WO3) has become increasingly popular due to its superior sensitivity, heightened responsiveness to sensitization, excellent stability, and suitability for lower-temperature operations [12]. In thin-film configuration, WO3 offers optimal electrical resistivity, selectivity, and repeatability, making it a promising candidate in sensor technology. Furthermore, it exhibits the ability to undergo changes in characteristics such as stoichiometry, composition, structure, thickness, and morphology, depending on the synthesis techniques and conditions employed [13]. WO3 predominantly in thin films shows commendable performances due to its cost-effective properties, simple fabrication, and easy deposition process.
Various deposition techniques (vacuum and non-vacuum techniques) have been used to deposit nanostructured WO3 thin films on substrate surfaces, including pulsed laser deposition (PLD) [14]; the electrophoresis deposition process (EDP) [15]; sputtering [16]; chemical spray pyrolysis [17]; solvothermal [18], hydrothermal [19,20], and sol–gel methods [21]; the chemical vapor deposition method [22]; and the physical vapor deposition method [23,24]. Hot filament chemical vapor deposition (HFCVD) is a cost-effective process compared to other deposition techniques. Its economic benefits arise from the use of a hot filament that decomposes preceding gases, which enables effective film development at lower temperatures [25]. It significantly minimizes the consumption of energy and instrument costs. For thin-film deposition, HFCVD is capable of depositing thin films with a high surface-to-volume ratio and intricate architectures. Accurate control enables the production of high-resolution patterns for device electrode fabrication [26]. HFCVD is beneficial for the fabrication of sensor electrodes because it can control the electrical resistivity, selectivity, and durability of thin films [27,28].
Methylamine is a hazardous organic chemical found in liquid and gas forms [29]. It is toxic, colorless, and flammable at room temperature with a typical pungent smell [30]. Methylamine is used as industrial raw material for the production of pesticides as an agriculture product, making rubbers for transportation industries, dyes for textile industries, and in pharmaceutical industries [31]. It is a hazardous chemical, and its accidental release or exposure can pose risks to public health and safety. Thus, it is crucial to monitor the presence of methylamine to ensure safety and prevent potential accidents or exposures [32]. In this regard, electrochemical detection methods can provide real-time detection and monitoring [33] of methylamine, allowing for prompt action to mitigate any risks. In particular, the chemical sensing methods enable the detection and quantification of methylamine in environmental samples [34], helping to identify potential pollution sources and assess the impact on the environment [35,36].
With these motivations, in this work, we have deposited WO3 nanostructured thin film through HFCVD, using tungsten (W) filament in a constant O2 pressure. The unique morphology displays WO3 thin-film nanowalls (WO3 NWs) with a grain size of 20–25 nm. To the best of our knowledge, this is the first report on single-step HFCVD-deposited WO3 thin film with distinct morphology and has been used as an electrode material for the detection of hazardous methylamine.

2. Materials and Methods

2.1. Materials

Methylamine (CH3NH2, Sigma Aldrich, ≥99.5%, St. Louis, MO, USA), silicon (Si, p-type, 10 mm × 10 mm, Siltron Inc., Seoul, Republic of Korea), tungsten (W) wire (thickness ~0.5 mm, The Nilaco Corporation, Tokyo, Japan), sodium dihydrogen phosphate (NaH2PO4, Sigma Aldrich, ≥99%, St. Louis, MO, USA), and disodium hydrogen phosphate (Na2HPO4, Sigma Aldrich, ≥99%, St. Louis, MO, USA) were used.

2.2. Synthesis of WO3 NW Thin Film

The WO3 NW thin film on the Si substrate was deposited through the HFCVD technique. As shown in Figure 1, the vacuum chamber is equipped with a thermocouple for measuring and monitoring the temperature, tungsten (W) wire filament was used, gas inlet pipes connected to mass flow controllers were used to regulate the flow of the gasses, and a high-current power supply, a rotary vacuum pump, and a cooling facility were attached to the vacuum chamber. For thin-film deposition, the silicon wafers (Si-P100) were cleaned with ultrasonic vibration using deionized (DI) water and acetone. Thereafter, Si substrates were placed on the substrate tray at the distance of ~10 mm from the W wire filament, and the chamber pressure was set to a constant 0.2 Torr. Subsequently, the W filament was heated at 1400 °C and the substrate temperature was raised from 300 to 450 °C for ~30 min under the supply of oxygen at a flow rate of 10 sccm. Herein, hydrogen gas was used as a precursor gas at 5.0 sccm. Finally, the W filament was oxidized at the above temperatures, resulting in the growth of WO3 nanostructured thin film on the Si substrate.

2.3. Characterization Techniques

Morphological analysis was carried out by a field-emission scanning electron microscope (FESEM, Hitachi S-4700, Tokyo, Japan). The elemental composition, mapping and line scan mapping analysis were determined by the SEM-coupled energy-dispersive X-rays spectroscopy (EDS). The absorption properties were obtained through a UV–visible spectrophotometer (JASCO, V-670, Tokyo, Japan). The structural investigations of WO3 nanostructured thin film were performed by X-ray diffraction (XRD, Rigaku, Tokyo, Japan) in the Bragg angle ranging between 20° and 60° to explain the crystal phases and lattice properties using CuKα radiation (λ = 1.5406 Å). Fourier-transform infrared (FTIR, Nicolet, IR300, Wisconsin, USA) was used to study bond vibrations in the range of 400–4000 cm−1. X-ray photoelectron spectroscopy (XPS, KRATOS AXIS-Nova, Manchester, UK) was conducted to study the surface compositional and element states.

2.4. Sensing Performance

To detect the presence of the hazardous methylamine, a three-electrode system of 10 mL electrochemical cells was employed for the measurement of cyclic voltammetry and linear sweep voltammetry using an electrometer (Keithley, 6517A, Aurora Rd, USA), Herein, HFCVD-grown WO3 NW thin film served as the working electrode, AgCl/Ag was employed as a reference electrode and gold wire was utilized as the counter electrode. A targeted analyte (methylamine) was prepared at a broad concentration range of 20 μM–1 mM in a 10 mL solution of 0.1 M phosphate buffer solution (PBS) of pH 7. The use of PBS in electrochemical sensing offers advantages such as stable pH, biocompatibility, consistent ionic strength, enhanced solubility, and compatibility. Its buffering capacity and versatility make PBS a reliable choice for maintaining optimal conditions during electrochemical measurements [37]. Cyclic voltammetry was performed to study the oxidation and reduction peaks and the linear sweep voltammetry technique was used to study the current and voltage responses. The active area of the fabricated electrode was 1 cm2 (WO3 NWs) and the sensitivity was calculated by dividing the slope of the calibrated current–concentration plot by the active area of the sensor, as expressed in Equation (1):
Sensitivity = Slope of calibrated curve Active area
Herein, CV was performed within the range of −0.8~0.8 V, with a scan rate of 50 mV/s. The current responses were analyzed within a voltage range from 0 to 2.0 V, and the response time was determined to be 10 s.

3. Results

3.1. Morphological Properties of WO3 NW Thin Film

The morphology of HFCVD-grown WO3 nanostructured thin films deposited on the Si substrate was investigated by FESEM and the elemental characterization was performed by energy-dispersive X-ray (EDAX). The HFCVD-grown WO3 thin film, as shown in Figure 2a–c, depicts the formation of highly dense and uniformly distributed nanowall (NW) structures. At an optimized substrate temperature (Ts) of ~400 °C, the surface of the thin film appears notably uniform, with a grain size in the range of 20–25 nm. However, upon raising the substrate temperature, the grain size increase might be due to the increased reaction rate at the substrate surface [38]. This improved grain size might boost the overall surface area of the WO3 thin film, indicating the availability of more active sites for chemical interactions. In the context of chemical sensing, the large surface area of the thin film brings about enhanced sensitivity, a faster response time, a low limit of detection and increases the reproducibility of the targeted electrode [39,40]. Figure 2d shows an elemental analysis of the HFCVD-grown WO3 NW thin film, exhibiting the presence of two primary elements: W (tungsten) and O (oxygen). The EDX analysis quantifies the elemental composition, with tungsten accounting for ~24.51% and oxygen making up the remaining ~75.49%.

3.2. Optical Characterizations of WO3 NW Thin Film

The optical properties of HFCVD-grown WO3 NW thin film are studied by UV–vis absorption in the range of 200~800 nm. The UV–vis spectrum, as shown in Figure 3a, exhibits a sharp absorption peak at ~339 nm, which confirms the deposition of the WO3 thin film [41,42]. The band gap value of WO3 NW thin film is calculated by the Tauc relation:
α = A ( h υ E g ) 2 h υ
where A, hυ and Eg are a constant of proportionality, photon energy and optical bandgap energy [43], respectively. In our work, the WO3 NW thin film has a band gap of ~3.321 eV, as shown in Figure 3b. The existence of an optical band gap energy of ~3.321 eV represents the minimum energy required to move an electron from the valence band (the highest energy band filled with electrons at absolute zero) to the conduction band (the lowest energy band with available electron states) within the thin film [44].

3.3. Crystalline and Structural Studies of WO3 NW Thin Film

XRD is performed to study the crystal structures, phases, crystallite size, and purity of the synthesized WO3 NW thin film. Figure 4a exhibits the diffraction patterns at 23.40°, 24.54°, 26.84°, 28.91°, 33.59°, 33.89°, 34.46°, 41.78°, 45.83°, 48.61°, 50.27°, 53.78°, 54.45°, 56.09°, and 60.53° relating to miller planes (020), (200), (120), (112), (022), (202), (122), (222), (004), (020), (114), (024), (204), (142), and (320), respectively [45,46]. The obtained XRD diffraction peaks are well matched with JCPDS card no. 83-0950. The crystal size of WO3 NW thin film is calculated as ~83.3472 nm by the Scherrer formula [47], using the most intense peak at 23.40°.
D = k λ β c o s θ
where D is the crystallite size of the particle in nm, θ is the diffraction angle, β is the full width at half maximum observed in radians (FWHM), k is the Scherrer constant (k = 0.94) and λ is the X-ray wavelength (λ = 1.54178 Å) [24,25]. Herein, the WO3 NW thin film prepared at Ts of ~400 °C shows phase purity at the strongest diffraction peak of 020 lattice planes, indicating the preferential growth orientation [48].
Figure 4b displays the results of Raman spectroscopy, performed in the range of 200–1000 cm−1, to measure the structural and molecular properties of the WO3 NW thin film. Raman spectra clearly manifest the signature peaks associated with stoichiometric WO3 with a monoclinic phase [49]. The W-O-W stretching and bending vibrations are located between 700 and 800 cm−1 [48] and O-W-O stretching and bending vibrations are noticed between 250 and 400 cm−1 [49]. The intense band at ~278.08 cm−1 and the weak band at ~330.17 cm−1 are attributed to the bending vibration of δ(O-W-O) [50]. The Raman peak observed at ~801 cm−1 is the typical polycrystalline WO3 in the monoclinic or triclinic crystalline phase [48,49].
Chemical configurations of HFCVD-grown WO3 NW thin films at Ts of ~400 °C is shown in Figure 4c. The spectrum exhibits transmittance in the range of 400–4000 cm−1. The FTIR spectra exhibit a broad peak at ~746 cm−1 and 815 cm−1 [51], which are attributed to the stretching vibration of υ(O-W-O) [51]. The υ(O-W-O) stretching vibration mode is the monoclinic crystal phase, confirming that the WO3 NW thin film is grown well on the Si substrate [52]. The transmittance peaks located at ~746 cm−1 refer to the O-W-O bending mode of the vibration and transmittance peak located at ~846 cm−1 is due to the W-O stretching mode of hexagonal WO, confirming the hexagonal structure [53].

3.4. XPS Studies of WO3 NW Thin Films

The XPS properties provide an in-depth understanding of the oxidation states of W and O in WO3 thin films using high-resolution spectra of W 4f and O 1s binding energies. Figure 5a showcases the fitted W4f XPS spectrum, revealing a distinct doublet pattern, with binding energies of approximately ~35.5 and ~37.7 eV [54]. These values correspond to the W 4f5/2 and W 4f7/2 electronic states, respectively. Peaks at 4f5/2 correspond to the binding energy of ~35.58 eV, aligning precisely with the characteristic energy level of the W6+ oxidation state within WO3 thin films and indicating the predominant oxidation state of tungsten [55] in the HFCVD-grown WO3 NW thin film. This peak serves as compelling evidence affirming that the WO3 thin film predominantly comprise hexavalent tungsten [56]. The second distinct XPS peak emerges at approximately ~37.9 eV, corresponding to the 4f7/2 state of tungsten ions in the HFCVD-grown WO3 NW thin film. This peak reaffirms the dominance of the W6+ oxidation state, which is crucial for understanding the electronic structure and chemical environment of the material [29,57]. Mostly, WO3 thin films are characterized by the presence of two distinct binding energies within the range of ~36–38 eV. These binding energies are indicative of the prevalence of W6+ ions within stoichiometric WO3 [25]. In our work, we observe a similar doublet pattern within the range of 36–38 eV, affirming the stoichiometric nature of the thin film [58,59]. The O 1s XPS plot, as depicted in Figure 5b, reveals two distinct binding energies: one with a higher intensity at ~530.4 eV and another with a lower intensity at ~531.4 and ~532.3 eV. Herein, the higher-intensity peak at ~530.2 eV typically arises from the O component of oxide, providing further evidence for the formation of W-O bonds in HFCVD-grown WO3 NW thin films [60]. Notably, the lower-energy peaks at ~531.4 and ~532.3 eV are associated with -OH or H2O species on the thin film’s surface, due to contamination that might occur with atmospheric moisture or crystal water [61].

3.5. Sensing Parameters of WO3 NW Thin-Film-Based Electrodes

A HFCVD-grown WO3 NW thin film-modified electrode is employed for the detection of methylamine. The electrochemical sensing studies were performed in 0.1 M phosphate buffer solution (PBS, pH = 7). Cyclic voltammetry (CV) measurements were used to analyze the electrochemical behavior at a scan rate of 50 mVs−1. The CV graphs for the detection of methylamine are shown in Figure 6a, exhibiting a promising electrochemical reversibility and efficiency with a redox response (oxidation and reduction) of the WO3 NW thin-film electrode of 0.2303 V redox potential [62]. These peaks usually occur due to the transfer of electrons displaying, the oxidation peak as a result of electron-losing behavior and the reduction peak occurs due to the electron-gaining behavior. In our work, the detection of methylamine by the HFCVD-grown WO3 NW thin-film electrode exhibits efficient electro chemical sensing behavior, which might be due to the high conductivity [33] and large surface area [63] of WO3 NW thin films. From linear sweep voltammetry, as shown in Figure 6b, a low current value of ~1.3 μA is observed in pristine PBS. However, it has been observed that upon the addition of the lowest amount of methylamine (20 μM), there is a significant change in current due to the high sensing properties of the HFCVD-grown WO3 NW thin film. Electrolytes with different concentrations of methylamine (20 μM–1 mM) display a lower current response of ~14.15 μA and the highest current value of ~23.23 μA. This gradual increase in the current indicates the rapid sensing response of HFCVD-grown WO3 NW thin-film electrodes in the detection of methylamine, which might result from the better electrocatalytic or electrochemical behavior and the fast electron exchange of HFCVD-grown WO3 NW thin films [64]. The increase in current upon the further addition of the targeted chemical in an electrolyte usually results in an increase in the ionic strength of the electrolyte. Herein, when increasing the methylamine concentration in the 0.1 M PBS, a large number of ions are generated; due to this response, more electrons are exchanged and this increases the opportunity for electron transfer at the electrode surface. As such, there is an increase in the rate of reaction, which enhances the sensitivity of the electrode. This result suggests that HFCVD-grown WO3 NWs show effective sensing capabilities in the detection of methylamine [65].
To examine the sensitivity of HFCVD-grown WO3 NW thin-film-based electrodes, Figure 7a shows the current vs. concentration calibration curve [33]. Herein, the current increases linearly with an increase in analyte concentration. The HFCVD-deposited WO3 NW thin-film-based electrode shows a reproducible, reliable and considerable promising sensitivity of ~183.65 μA mM−1 cm−2 with a linearity of 20 μM−1 mM, a detection limit of ~20 μM and a correlation coefficient (R2) of ~0.97708 in 10 s response time. The existence of a good current response and reliable sensitivity might suggest high electron mobility and electrochemical activity [66] over the surface of HFCVD-grown WO3 NW thin-film electrodes. The stability performance of HFCVD-grown WO3 NW thin-film electrode materials was performed by a linear sweep voltammetry graph (current vs. voltage graph) in the presence of 20 μM methylamine. The electrochemical behavior was studied twice a day for 1 month. As seen in Figure 7b, ~95% of the current response remains the same as compared to the measurements performed on the first day of the analysis and no significant change in current is observed, which shows the good stability of HFCVD-grown WO3 NW thin-film electrodes. The performance of HFCVD-grown WO3 NW thin-film-based sensor is compared to other reported sensors [67,68,69,70,71], as shown in Table 1.
The detection mechanism of methylamine by HFCVD-grown WO3 NW thin-film-based electrodes is related to the changes in electrode conductance during the interaction of fabricated electrodes with an analyte. As shown in Figure 8, oxygen species adhere to the surface of WO3 NW thin films grown through the HFCVD process. The HFCVD-grown WO3 NW thin film shows n-type semiconducting behavior [33]. Due to the nature of WO3 thin films, electrons in the conduction band tend to bond with the surface area of WO3 thin films. By capturing electrons within WO3 thin films, these oxygen species (O2 adsorbed) transform into active sites, converting into anionic species containing oxygen [72]. The presence of these adsorbed oxygen species on the surface stimulates low-energy electrons in the valence band, leading to an increase in the number of holes within the HFCVD-grown WO3 NW thin film. This accumulation of holes eventually results in reduced resistance in the fabricated sensor. When the target methylamine molecule interacts with the adsorbed oxygen ions, it triggers the release of trapped electrons, as described by the following equations:
O 2   ( gas ) O 2   ( ads )
O 2   ( ads ) + ē   ( CB of WO 3 ) O 2   ( ads ) O   ( ads )
8 O + C 2 H 4 ( N H 2 ) 2 4 C O 2 + H 2 O + 2 N 2 + 4 ē
During the oxidation process, electrons emitted from the conduction band of the HFCVD-grown WO3 thin-film NWs enhance the electrical conductivity, resulting in an increased current associated with the methylamine [73]. In our work, the deposited WO3 thin film exhibits uniformly distributed nanowalls of an average grain size of ~20–25 nm. This structure might provide a large surface area for adsorption, making it highly efficient for the detection of methylamine, ensuring exceptional electrochemical performance.

4. Conclusions

This study focuses on the development of a sensitive sensor for detecting the hazardous chemical methylamine. A uniform WO3 thin-film nanowall structure on a Si substrate is obtained at a relatively low temperature of ~400 °C through HFCVD. The resulting WO3 NW-based thin-film-based electrode exhibits highly effective detection of methylamine at very low concentrations. The obtained results are attributed to the unique nanowall-like structure of WO3 thin films, which might offer a large surface area, facilitating efficient electron transfer during the electrochemical detection of methylamine. The WO3 NW-based fabricated chemical sensor demonstrates a promising sensitivity of ~196.33 μA μM−1 cm−2, a low limit of detection (LOD) of ~12 μM, and a strong retention coefficient of ~0.97708. Thus, our fabricated sensor has the potential for application in environmental monitoring, making it adaptable for the detection of other harmful chemicals in the future.

Author Contributions

Conceptualization, M.I. and E.-B.K.; methodology, M.I.; software, M.I., D.-H.K., T.-G.K. and E.-B.K.; validation, M.I., M.S.A. and E.-B.K.; formal analysis, M.I. and E.-B.K.; investigation, M.I.; resources, M.I and T.-G.K.; data curation, M.I., E.-B.K. and S.A.; writing—original draft preparation, M.I. and E.-B.K.; writing—review and editing, M.S.A. and S.A.; supervision, S.A.; project administration, M.S.A. and D.-H.K.; funding acquisition, M.S.A., D.-H.K. and E.-B.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by National University Development Project at Jeonbuk National University in 2022.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Theerthagiri, J.; Chandrasekaran, S.; Salla, S.; Elakkiya, V.; Senthil, R.A.; Nithyadharseni, P.; Maiyalagan, T.; Micheal, K.; Ayeshamariam, A.; Arasu, M.V.; et al. Recent Developments of Metal Oxide Based Heterostructures for Photocatalytic Applications towards Environmental Remediation. J. Solid State Chem. 2018, 267, 35–52. [Google Scholar] [CrossRef]
  2. Kanazawa, E.; Sakai, G.; Shimanoe, K.; Kanmura, Y.; Teraoka, Y.; Miura, N.; Yamazoe, N. Metal Oxide Semiconductor N2O Sensor for Medical Use. Sens. Actuators B Chem. 2001, 77, 72–77. [Google Scholar] [CrossRef]
  3. Li, F.; Ran, J.; Jaroniec, M.; Qiao, S.Z. Solution Combustion Synthesis of Metal Oxide Nanomaterials for Energy Storage and Conversion. Nanoscale 2015, 7, 17590–17610. [Google Scholar] [CrossRef] [PubMed]
  4. Bashambu, L.; Singh, R.; Verma, J. Metal/Metal Oxide Nanocomposite Membranes for Water Purification. Mater. Today Proc. 2021, 44, 538–545. [Google Scholar] [CrossRef]
  5. Jolly, A.; Kim, H.; Moon, J.-Y.; Mohan, A.; Lee, Y.-C. Exploring the Imminent Trends of Saponins in Personal Care Product Development: A Review. Ind. Crops Prod. 2023, 205, 117489. [Google Scholar] [CrossRef]
  6. Sun, Y.-F.; Liu, S.-B.; Meng, F.-L.; Liu, J.-Y.; Jin, Z.; Kong, L.-T.; Liu, J.-H. Metal Oxide Nanostructures and Their Gas Sensing Properties: A Review. Sensors 2012, 12, 2610–2631. [Google Scholar] [CrossRef] [PubMed]
  7. Subramani, T.; Thimmarayan, G.; Balraj, B.; Chandrasekar, N.; Palanisamy, M.; Nagarajan, S.K.; Amirtharajan, S.; Kumar, M.; Sivakumar, C. Surfactants Assisted Synthesis of WO3 Nanoparticles with Improved Photocatalytic and Antibacterial Activity: A Strong Impact of Morphology. Inorg. Chem. Commun. 2022, 142, 109709. [Google Scholar] [CrossRef]
  8. Diez-Cabanes, V.; Segalina, A.; Pastore, M. Effects of Size and Morphology on the Excited-State Properties of Nanoscale WO3 Materials from First-Principles Calculations: Implications for Optoelectronic Devices. ACS Appl. Nano Mater. 2022, 5, 16289–16297. [Google Scholar] [CrossRef]
  9. Lima, L.V.C.; Rodriguez, M.; Freitas, V.A.A.; Souza, T.E.; Machado, A.E.H.; Patrocínio, A.O.T.; Fabris, J.D.; Oliveira, L.C.A.; Pereira, M.C. Synergism between n-type WO3 and p-type δ-FeOOH semiconductors: High interfacial contacts and enhanced photocatalysis. Appl. Catal. B Environ. 2015, 165, 579–588. [Google Scholar] [CrossRef]
  10. Azam, A.; Kim, J.; Park, J.; Novak, T.G.; Tiwari, A.P.; Song, S.H.; Kim, B.; Jeon, S. Two-Dimensional WO3 Nanosheets Chemically Converted from Layered WS2 for High-Performance Electrochromic Devices. Nano Lett. 2018, 18, 5646–5651. [Google Scholar] [CrossRef]
  11. Jadkar, V.; Pawbake, A.; Waykar, R.; Jadhavar, A.; Date, A.; Late, D.; Pathan, H.; Gosavi, S.; Jadkar, S. Synthesis of g-WO3 thin films by hot wire-CVD and investigation of its humidity sensing properties. Phys. Status Solidi A 2017, 214, 1600717. [Google Scholar] [CrossRef]
  12. 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]
  13. Zhu, L.; Zheng, W. Room-temperature gas sensing of ZnO-based gas sensor: A review. Sens. Actuators A Phys. 2017, 267, 242–261. [Google Scholar] [CrossRef]
  14. Zou, Y.S.; Zhang, Y.C.; Lou, D.; Wang, H.P.; Gu, L.; Dong, Y.H.; Dou, K.; Song, X.F.; Zeng, H.B. Structural and optical properties of WO3 films deposited by pulsed laser deposition. J. Alloys Compd. 2014, 583, 465–470. [Google Scholar] [CrossRef]
  15. Heidari, E.K.; Zamani, C.; Marzbanrad, E.; Raissi, B.; Nazarpour, S. WO3-based NO2 sensors fabricated through low frequency AC electrophoretic deposition. Sens. Actuators B Chem. 2010, 146, 165–170. [Google Scholar] [CrossRef]
  16. Kim, T.S.; Kim, Y.B.; Yoo, K.S.; Sung, G.S.; Jung, H.J. Sensing characteristics of dc reactive sputtered WO3 thin films as an NOx gas sensor. Sens. Actuators B Chem. 2000, 62, 102–108. [Google Scholar] [CrossRef]
  17. Arutanti, O.; Ogi, T.; Nandiyanto, A.B.D.; Iskandar, F.; Okuyama, K. Controllable crystallite and particle sizes of WO3 particles prepared by a spray-pyrolysis method and their photocatalytic activity. Inorg. Mater. Synth. Process. 2014, 60, 41–49. [Google Scholar] [CrossRef]
  18. Li, Y.; McMaster, W.A.; Wei, H.; Chen, D.; Caruso, R.A. Enhanced Electrochromic Properties of WO3 Nanotree-like Structures Synthesized via a Two-Step Solvothermal Process Showing Promise for Electrochromic Window Application. ACS Appl. Nano Mater. 2018, 1, 2552–2558. [Google Scholar] [CrossRef]
  19. Su, X.; Xiao, F.; Li, Y.; Jian, J.; Sun, Q.; Wang, J. Synthesis of uniform WO3 square nanoplates via an organic acid-assisted hydrothermal process. Mater. Lett. 2010, 64, 1232–1234. [Google Scholar] [CrossRef]
  20. Salmaoui, S.; Sediri, F.; Gharbi, N. Characterization of h-WO3 nanorods synthesized by hydrothermal process. Polyhedron 2010, 29, 1771–1775. [Google Scholar] [CrossRef]
  21. Djaoued, Y.; Priya, S.; Balaji, S. Low temperature synthesis of nanocrystalline WO3 films by sol–gel process. J. Non-Cryst. Solids 2008, 354, 673–679. [Google Scholar]
  22. Kafizas, A.; Francàs, L.; Sotelo-Vazquez, C.; Ling, M.; Li, Y.; Glover, E.; McCafferty, L.; Blackman, C.; Darr, J.; Parkin, I. Optimizing the Activity of Nanoneedle Structured WO3 Photoanodes for Solar Water Splitting: Direct Synthesis via Chemical Vapor Deposition. J. Phys. Chem. C 2017, 121, 5983–5993. [Google Scholar] [CrossRef]
  23. Keshri, S.; Kumar, A.; Kabiraj, D. Effect of Annealing on Structural, Optical and Electrical Behaviors of WO3 Thin Films Prepared by Physical Vapor Deposition Method. J. Nano-Electron. Phys. 2011, 3, 260–267. [Google Scholar]
  24. Buch, V.R.; Chawla, A.K.; Rawal, S.K. Review on electrochromic property for WO3 thin films using different deposition techniques. Mater. Today Proc. 2016, 3, 1429–1437. [Google Scholar] [CrossRef]
  25. Godbole, R.; Ameen, S.; Nakate, U.T.; Akhtar, M.S.; Shin, H.-S. Low temperature HFCVD synthesis of tungsten oxide thin film for high response hydrogen gas sensor application. Mater. Lett. 2019, 19, 398–401. [Google Scholar] [CrossRef]
  26. Zhai, Z.; Shen, H.; Chen, J.; Li, X.; Jiang, Y. Evolution of Structural and Electrical Properties of Carbon Films from Amorphous Carbon to Nanocrystalline Graphene on Quartz Glass by HFCVD. ACS Appl. Mater. Interfaces 2018, 10, 17427–17436. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, T.; Zou, Y. The Effect of Deposition Parameters on the Growth Rate of Microcrystalline Diamond Powders Synthesized by HFCVD Method. Coating 2017, 7, 95. [Google Scholar] [CrossRef]
  28. Tan, G.L.; Tang, D.; Dastan, D.; Jafari, A.; Silva, J.P.B.; Yin, X.T. Effect of heat treatment on electrical and surface properties of tungsten oxide thin films grown by HFCVD technique. Mater. Sci. Semicon. Process. 2021, 122, 105506. [Google Scholar] [CrossRef]
  29. Jacob, S.P.C. The influence of substrate temperature variation on tungsten oxide thin film growth in an HFCVD system. Appl. Surf. Sci. 2007, 253, 3317–3325. [Google Scholar]
  30. Jafari, A.; Ghoranneviss, M.; Elahi, A.S. HFCVD application for growth of monoclinic tungsten trioxide crystal nano-walls. J. Inorg. Organomet. Polym. Mater. 2016, 26, 254–258. [Google Scholar] [CrossRef]
  31. White, C.M.; Gillaspie, D.T.; Whitney, E.; Lee, S.-H.; Dillon, A.C. Flexible electrochromic devices based on crystalline WO3 nanostructures produced with hot-wire chemical vapor deposition. Thin Solid Films 2009, 517, 3596–3599. [Google Scholar] [CrossRef]
  32. Shankar, N.; Yu, M.-F.; Vanka, S.P.; Glumac, N.G. Synthesis of tungsten oxide (WO3) nanorods using carbon nanotubes as templates by hot filament chemical vapor deposition. Mater. Lett. 2006, 60, 771–774. [Google Scholar] [CrossRef]
  33. Imran, M.; Kim, E.-B.; Kwak, D.-H.; Akhtar, M.S.; Ameen, S. Controlled Growth of WO3 Pyramidal Thin Film via Hot-Filament Chemical Vapor Deposition: Electrochemical Detection of Ethylenediamine. Chemosensors 2021, 9, 257. [Google Scholar] [CrossRef]
  34. Lu, D.; Ogino, A.; Liang, B.; Liu, J.; Nagatsu, M. Field-Emission Properties of Nanostructured WO3 Arrays Fabricated Using Tungsten Hot-Filament Chemical Vapor Deposition. Jpn. J. Appl. Phys. 2009, 48, 090206. [Google Scholar] [CrossRef]
  35. Sahu, C.; Sircar, A.; Sangwai, J.S.; Kumar, R. Effect of Methylamine, Amylamine, and Decylamine on the Formation and Dissociation Kinetics of CO2 Hydrate Relevant for Carbon Dioxide Sequestration. Ind. Eng. Chem. Res. 2022, 61, 2672–2684. [Google Scholar] [CrossRef]
  36. Vinogradoff, V.; Duvernay, F.; Danger, G.; Theulé, P.; Borget, F.; Chiavassa, T. Formaldehyde and methylamine reactivity in interstellar ice analogues as a source of molecular complexity at low temperature. Astron. Astrophys. 2013, 549, A40. [Google Scholar] [CrossRef]
  37. Noorhashimah, M.N.; Razak, K.A.; Lockman, Z. Effect of Concentration and pH of PBS to the Electrocatalytic Performance of Enzymatic Glucose Biosensor. Solid State Phenom. 2019, 290, 195–198. [Google Scholar] [CrossRef]
  38. Murugappan, K.; Kang, C.; Silvester, D.S. Electrochemical Oxidation and Sensing of Methylamine Gas in Room Temperature Ionic Liquids. J. Phys. Chem. C 2014, 118, 19232–19237. [Google Scholar] [CrossRef]
  39. Dai, X.; Rabeah, J.; Yuan, H.; Brückner, A.; Cui, X.; Shi, F. Glycerol as a Building Block for Prochiral Aminoketone, N-Formamide, and N-Methyl Amine Synthesis. ChemSusChem 2016, 9, 3133–3138. [Google Scholar] [CrossRef]
  40. Liu, Y.; Xu, X.; Lu, H.; Yan, B. Dual-emission ratiometric fluorescent probe-based lanthanide-functionalized hydrogen-bonded organic framework for the visual detection of methylamine. J. Mater. Chem. C 2022, 10, 1212–1219. [Google Scholar] [CrossRef]
  41. Huang, G.; Zhou, Y.; Li, F.; Tan, X.; Cai, Z.; Luo, D.; Chen, T.; Zhang, M. An effective and reliable fluorescent sensor for selective detection of methylamine gas based on in-situ formation of MAPbBr3 perovskite nanocrystals in electrospun fibers. Sens. Actuators B Chem. 2021, 347, 130618. [Google Scholar] [CrossRef]
  42. Kim, T.-H.; Yoon, J.-W.; Kang, Y.C.; Abdel-Hady, F.; Wazzan, A.A.; Lee, J.-H. A strategy for ultrasensitive and selective detection of methylamine using p-type Cr2O3: Morphological design of sensing materials, control of charge carrier concentrations, and configurational tuning of Au catalysts. Sens. Actuators B Chem. 2017, 240, 1049–1057. [Google Scholar] [CrossRef]
  43. Rajagopal, S.; Nataraj, D.; Mangalaraj, D.; Djaoued, Y.; Robichaud, J.; Khyzhun, O.Y. Controlled Growth of WO3 Nanostructures with Three Different Morphologies and Their Structural, Optical, and Photodecomposition Studies. Nanoscale Res. Lett. 2009, 4, 1335–1342. [Google Scholar] [CrossRef] [PubMed]
  44. Phuruangrat, A.; Ham, D.J.; Hong, S.J.; Thongtem, S.; Lee, J.S. Synthesis of hexagonal WO3 nanowires by microwave-assisted hydrothermal method and their electrocatalytic activities for hydrogen evolution reaction. J. Mater. Chem. 2010, 20, 1780–1786. [Google Scholar] [CrossRef]
  45. Hunge, Y.M.; Yadav, A.A.; Mathe, V.L. Ultrasound assisted synthesis of WO3-ZnO nanocomposites for brilliant blue dye degradation. Ultrason. Sonochem. 2018, 45, 116–122. [Google Scholar] [CrossRef] [PubMed]
  46. Hersh, H.N.; Kramer, W.E.; McGee, J.H. Mechanism of electrochromism in WO3. Appl. Phys. Lett. 1975, 27, 646–648. [Google Scholar] [CrossRef]
  47. González-Borrero, P.P.; Sato, F.; Medina, A.N.; Baesso, M.L.; Bento, A.C.; Baldissera, G.; Persson, C.; Niklasson, G.A.; Granqvist, C.G.; Ferreira da Silva, A. Optical band-gap determination of nanostructured WO3 film. Appl. Phys. Lett. 2010, 96, 061909. [Google Scholar] [CrossRef]
  48. Vemuri, R.S.; Engelhard, M.H.; Ramana, C.V. Correlation between Surface Chemistry, Density, and Band Gap in Nanocrystalline WO3 Thin Films. ACS Appl. Mater. Interfaces 2012, 4, 1371–1377. [Google Scholar] [CrossRef]
  49. Colton, R.J.; Guzman, A.M.; Rabalais, J.W. Electrochromism in some thin-film transition-metal oxides characterized by X-ray electron spectroscopy. J. Appl. Phys. 1978, 49, 409–416. [Google Scholar] [CrossRef]
  50. Daniel, M.F.; Desbat, B.; Lassegues, J.C.; Gerand, B.; Figlarz, M. Infrared and Raman study of WO3 tungsten trioxides and WO3, xH2O tungsten trioxide hydrates. J. Solid State Chem. 1987, 67, 235–247. [Google Scholar] [CrossRef]
  51. Lee, S.H.; Cheong, H.M.; Liu, P.; Smith, D.; Tracy, C.E.; Mascarenhas, A.; Pitts, J.R.; Deb, S.K. Raman spectroscopic studies of gasochromic a-WO3 thin films. Electrochim. Acta 2011, 46, 1995–1999. [Google Scholar] [CrossRef]
  52. Boulova, M.; Lucazeau, G. Crystallite Nanosize Effect on the Structural Transitions of WO3 Studied by Raman Spectroscopy. J. Solid State Chem. 2002, 167, 425–434. [Google Scholar] [CrossRef]
  53. Cazzanelli, E.; Vinegoni, C.; Mariotto, G.; Kuzmin, A.; Purans, J. Raman study of the phase transitions sequence in pure WO3 at high temperature and in HxWO3 with variable hydrogen content. Solid State Ionics 1999, 123, 67–74. [Google Scholar] [CrossRef]
  54. Kanan, S.M.; Tripp, C.P. Synthesis, FTIR studies and sensor properties of WO3 powders. Curr. Opin. Solid State Mater. Sci. 2007, 11, 19–27. [Google Scholar] [CrossRef]
  55. Deepa, M.; Sharma, N.; Varshney, P.; Varma, S.P.; Agnihotry, S.A. FTIR investigations of solid precursor materials for sol-gel deposition of WO3 based electrochromic films. J. Mater. Sci. 2000, 35, 5313–5318. [Google Scholar] [CrossRef]
  56. Charton, P.; Gengembre, L.; Armand, P. TeO2-WO3 Glasses: Infrared, XPS and XANES Structural Characterizations. J. Solid State Chem. 2002, 168, 175–183. [Google Scholar] [CrossRef]
  57. Shpak, A.P.; Korduban, A.M.; Medvedskij, M.M.; Kandyba, V.O. XPS studies of active elements surface of gas sensors based on WO3−x nanoparticles. J. Electron Spectrosc. Relat. Phenom. 2007, 156–158, 172–175. [Google Scholar]
  58. Lozzi, L.; Ottaviano, L.; Passacantando, M.; Santucci, S.; Cantalini, C. The influence of air and vacuum thermal treatments on the NO2 gas sensitivity of WO3 thin films prepared by thermal evaporation. Thin Solid Films 2001, 391, 224–228. [Google Scholar] [CrossRef]
  59. Leftheriotis, G.; Papaefthimiou, S.; Yianoulis, P.; Siokou, A. Effect of the tungsten oxidation states in the thermal coloration and bleaching of amorphous WO3 films. Thin Solid Films 2001, 384, 298–306. [Google Scholar] [CrossRef]
  60. Soto, G.; De La Cruz, W.; Díaz, J.A.; Machorro, R.; Castillón, F.F.; Farías, M.H. Characterization of tungsten oxide films produced by reactive pulsed laser deposition. Appl. Surf. Sci. 2003, 218, 281–289. [Google Scholar] [CrossRef]
  61. Wang, Q.; Wu, H.; Wang, Y.; Li, J.; Yang, Y.; Cheng, X.; Luo, Y.; An, B.; Pan, X.; Xie, E. Ex-situ XPS analysis of yolk-shell Sb2O3/WO3 for ultra-fast acetone resistive sensor. J. Hazard Mater. 2021, 412, 125175. [Google Scholar] [CrossRef]
  62. Joshi, N.; da Silva, L.F.; Jadhav, H.S.; Shimizu, F.M.; Suman, P.H.; M’Peko, J.C.; Orlandie, M.O.; Seo, J.G.; Mastelaro, V.R.; Oliveira, O.N., Jr. Yolk-shelled ZnCo2O4 microspheres: Surface properties and gas sensing application. Sens. Actuators B Chem. 2018, 257, 906–915. [Google Scholar] [CrossRef]
  63. Kim, E.-B.; Imran, M.; Shaheer Akhtar, M.; Shin, H.-S.; Ameen, S. Enticing 3D peony-like ZnGa2O4 microstructures for electrochemical detection of N, N-dimethylmethanamide chemical. J. Hazard. Mater. 2021, 404, 124069. [Google Scholar] [CrossRef]
  64. Imran, M.; Kim, E.-B.; Kwak, D.-H.; Ameen, S. Porous MgNiO2 Chrysanthemum Flower Nanostructure Electrode for Toxic Hg2+ Ion Monitoring in Aquatic Media. Sensors 2023, 23, 7910. [Google Scholar] [CrossRef] [PubMed]
  65. Li, N.; Fan, Y.; Shi, Y.; Xiang, Q.; Wang, X.; Xu, J. A low temperature formaldehyde gas sensor based on hierarchical SnO/SnO2 nano-flowers assembled from ultrathin nanosheets: Synthesis, sensing performance and mechanism. Sens. Actuators B Chem. 2019, 294, 106–115. [Google Scholar] [CrossRef]
  66. Ibrahim, A.A.; Dar, G.N.; Zaidi, S.A.; Umar, A.; Abaker, M.; Bouzid, H.; Baskoutas, S. Growth and properties of Ag-doped ZnO nanoflowers for highly sensitive phenyl hydrazine chemical sensor application. Talanta 2012, 93, 257–263. [Google Scholar] [CrossRef]
  67. Murugappan, K.; Silvester, D.S. Sensors for Highly Toxic Gases: Methylamine and Hydrogen Chloride Detection at Low Concentrations in an Ionic Liquid on Pt Screen Printed Electrodes. Sensors 2015, 15, 26866–26876. [Google Scholar] [CrossRef] [PubMed]
  68. Rahman, M.M.; Khan, S.B.; Jamal, A.; Faisal, M.; Asiri, A.M. Fabrication of highly sensitive acetone sensor based on sonochemically prepared as-grown Ag2O nanostructures. Chem. Eng. J. 2012, 192, 122–128. [Google Scholar] [CrossRef]
  69. Hussain, M.M.; Rahman, M.M.; Asiri, A.M. Efficient 2-Nitrophenol Chemical Sensor Development Based on Ce2O3 Nanoparticles Decorated CNT Nanocomposites for Environmental Safety. PLoS ONE 2016, 11, e0166265. [Google Scholar] [CrossRef] [PubMed]
  70. Godbole, R.; Imran, M.; Kim, E.-B.; Park, J.B.; Ameen, S. Novel approach to synthesize morphology variant tungsten oxide thin films for efficient chemical sensing. Ceram. Int. 2022, 48, 12506–12514. [Google Scholar] [CrossRef]
  71. Rahman, M.; Khan, S.B.; Faisal, M.; Asiri, A.M.; Alamry, K.A. Highly sensitive formaldehyde chemical sensor based on hydrothermally prepared spinel ZnFe2O4 nanorods. Sens. Actuators B Chem. 2012, 171, 932–937. [Google Scholar] [CrossRef]
  72. Kim, E.B.; Ameen, S.; Akhtar, M.S.; Shin, H.S. Iron-nickel co-doped ZnO nanoparticles as scaffold for field effect transistor sensor: Application in electrochemical detection of hexahydropyridine chemical. Sens. Actuators B Chem. 2018, 275, 422–431. [Google Scholar] [CrossRef]
  73. Ameen, S.; Akhtar, M.S.; Shin, H.S. Hydrazine chemical sensing by a modified electrode based on in situ electrochemically synthesized polyaniline/graphene composite thin film. Sens. Actuators B Chem. 2012, 173, 177–183. [Google Scholar] [CrossRef]
Figure 1. A schematic representation of the HFCVD process for the synthesis of WO3 NW thin films.
Figure 1. A schematic representation of the HFCVD process for the synthesis of WO3 NW thin films.
Micromachines 15 00441 g001
Figure 2. FESEM images (ac) and EDX spectrum (d) of the HFCVD-grown WO3 NW thin film.
Figure 2. FESEM images (ac) and EDX spectrum (d) of the HFCVD-grown WO3 NW thin film.
Micromachines 15 00441 g002
Figure 3. UV–vis spectroscopy (a) and its corresponding Tauc plot (b) of the HFCVD-grown WO3 NW thin film.
Figure 3. UV–vis spectroscopy (a) and its corresponding Tauc plot (b) of the HFCVD-grown WO3 NW thin film.
Micromachines 15 00441 g003
Figure 4. The XRD patterns (a), Raman spectrum (b) and FTIR spectrum (c) of HFCVD-grown WO3 NW thin films.
Figure 4. The XRD patterns (a), Raman spectrum (b) and FTIR spectrum (c) of HFCVD-grown WO3 NW thin films.
Micromachines 15 00441 g004
Figure 5. W 4f (a) and O 1s (b) XPS plots of HFCVD-grown WO3 NW thin films.
Figure 5. W 4f (a) and O 1s (b) XPS plots of HFCVD-grown WO3 NW thin films.
Micromachines 15 00441 g005
Figure 6. (a) Cyclic voltammetry plot with methylamine (1 mM) in 0.1 M PBS of a HFCVD-grown nanostructured WO3 thin film and (b) I–V curves of the HFCVD-grown nanostructured WO3 thin-film-based chemical.
Figure 6. (a) Cyclic voltammetry plot with methylamine (1 mM) in 0.1 M PBS of a HFCVD-grown nanostructured WO3 thin film and (b) I–V curves of the HFCVD-grown nanostructured WO3 thin-film-based chemical.
Micromachines 15 00441 g006
Figure 7. (a) Calibrated current versus the concentration of methylamine of HFCVD-grown WO3 NW thin-film-based chemical sensors. (b) Stability test through I–V measurements in the presence of 20 μM methylamine in 0.1 M PBS.
Figure 7. (a) Calibrated current versus the concentration of methylamine of HFCVD-grown WO3 NW thin-film-based chemical sensors. (b) Stability test through I–V measurements in the presence of 20 μM methylamine in 0.1 M PBS.
Micromachines 15 00441 g007
Figure 8. Possible sensing mechanism for the detection of methylamine over the surface of HFCVD-grown WO3 NW thin-film-based electrodes.
Figure 8. Possible sensing mechanism for the detection of methylamine over the surface of HFCVD-grown WO3 NW thin-film-based electrodes.
Micromachines 15 00441 g008
Table 1. Sensing parameters of WO3 NWs electrode-based sensors compared with reported chemical sensors.
Table 1. Sensing parameters of WO3 NWs electrode-based sensors compared with reported chemical sensors.
MaterialsPreparation MethodChemicalsSensitivityLODR2Refs.
WO3HFCVDethylenediamine161.33 μA μM−1 cm−29.56 μM0.98[33]
PANI/GrSpin coatinghydrazine32.54 × 10−5 μA cm−2 mM−115.38 mM0.78578[38]
Ag2OSonochemical methodacetone1.689 μA cm−2 mM−10.11 μM0.946[67]
Ce2O3Wet chemical method2-nitrophenol1.689 μA mM−1 cm−2 0.9030[68]
WO3HFCVDdiethylamine3.5 μA μM−1 cm−27 μM [69]
ZnFe2O4Hydrothermal methodformaldehyde4.10 μA cm−2 mM−10.89 μM [70]
MAPbBr3Electrospunmethylamine-0.8 ppm0.9904[71]
WO3HFCVDmethylamine183.65 μA μM−1 cm−220 μM0.97708This 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

Imran, M.; Kim, E.-B.; Kim, T.-G.; Ameen, S.; Akhtar, M.S.; Kwak, D.-H. Fabrication of Tungsten Oxide Nanowalls through HFCVD for Improved Electrochemical Detection of Methylamine. Micromachines 2024, 15, 441. https://doi.org/10.3390/mi15040441

AMA Style

Imran M, Kim E-B, Kim T-G, Ameen S, Akhtar MS, Kwak D-H. Fabrication of Tungsten Oxide Nanowalls through HFCVD for Improved Electrochemical Detection of Methylamine. Micromachines. 2024; 15(4):441. https://doi.org/10.3390/mi15040441

Chicago/Turabian Style

Imran, Mohammad, Eun-Bi Kim, Tae-Geum Kim, Sadia Ameen, Mohammad Shaheer Akhtar, and Dong-Heui Kwak. 2024. "Fabrication of Tungsten Oxide Nanowalls through HFCVD for Improved Electrochemical Detection of Methylamine" Micromachines 15, no. 4: 441. https://doi.org/10.3390/mi15040441

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