Next Article in Journal
Modification of Chabazite Using Hexadecyltrime-Thylammonium Bromide (HDTMA-Br) for Chromium(VI) Removal from Water Solutions
Next Article in Special Issue
Chlorella vulgaris-Derived Biochars for Metribuzin Removal: Influence of Thermal Processing Pathways on Sorption Properties
Previous Article in Journal
Identification of Laser Parameters Acting on an Axisymmetric Domain Using an Artificial Immune System
Previous Article in Special Issue
Evaluation of the Porosity and Morphology of Microstructured Charcoal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gas Sensor Properties of (CuO/WO3)-CuWO4 Heterostructured Nanocomposite Materials

by
Michael Castaneda Mendoza
1,
Carlos A. Parra Vargas
1,
Miryam Rincón Joya
2,*,
Adenilson J. Chiquito
3 and
Angela M. Raba-Páez
4
1
Grupo Física de Materiales—GFM, Universidad Pedagógica y Tecnológica de Colombia, Avenida Central Del Norte 39-115, Tunja 150003, Boyacá, Colombia
2
Grupo de Física Mesoscópica, Universidad Nacional de Colombia, Carrera 30 Calle 45-03, Bogotá 111321, Cundinamarca, Colombia
3
NanoLab, Departamento de Física, Universidade Federal de São Carlos, Rodovia Washington Luiz km 235, São Carlos 13565-905, São Paulo, Brazil
4
Grupo de Investigación en Materiales Poliméricos—GIMAPOL, Universidad Francisco de Paula Santander, Avenida Gran Colombia 12E-96, Cúcuta 540003, Norte de Santander, Colombia
*
Author to whom correspondence should be addressed.
Materials 2025, 18(12), 2896; https://doi.org/10.3390/ma18122896
Submission received: 12 May 2025 / Revised: 3 June 2025 / Accepted: 6 June 2025 / Published: 18 June 2025

Abstract

In this work, we report the evaluation of a (CuO/WO3)-CuWO4 heterostructured system as a methanol and acetone gas sensor in different configurations, contrasted with the pure oxides CuO and WO3. The samples were synthesized using a modified precipitation route followed by a single thermal treatment step to induce multiphase simultaneous crystallization. The structural characterization by XRD showed that all the materials presented the formation of monoclinic CuO and WO3 and triclinic CuWO4. No additional phases were detected in the samples, and a reduction in the crystallite size of the CuO phase after the crystallization in the heterostructured system was observed. FE-SEM analysis made it possible to directly observe the morphology and the structures of the samples at the nanometer scale, showing a heterogeneous grain formation and supporting the formation of a heterostructure. UV-Vis DRS was used to study the optical properties of the materials, and the presence of two optical band gaps was successfully determined, which provides further evidence of heterostructure formation via this modified synthesis route. The variation in the resistance of the materials was observed in the presence of methanol and acetone vapors, where the heterostructure exhibited a substantial change in performance in the configuration with 40% copper precursor (Cu40:W60), the sample that presented the highest response as a sensor against these VOCs. To our knowledge, this is the first time that this system has been reported as a gas sensor, using the multiple configurations of the (CuO/WO3)-CuWO4 heterostructured system.

1. Introduction

Currently, there is a significant scientific interest in the quantifying pollutant compounds present in different essential natural resources such as water and air. The accelerated development of human activity and industrial development has led to an increase in the emission of highly toxic pollutants to humans. Although these compounds are important in the early stages of industrial processes, when they are released or escape from the processes, they result in potentially harmful and toxic compounds for any living organism. Examples of these gases include ethanol, methanol, acetone, and benzene, among others, generally classified as volatile organic compounds (VOCs). Exposure to concentrations above permissible limits can lead to health problems and even death [1].
Tungsten oxide (WO3) is among the most cost-effective n-type semiconductors, showing promising performance in many applications, such as solar cells, sensors, photoelectrochemical water splitting, photocatalysis, and superconductivity, among others [2,3,4]. Due to its narrow indirect band gap, between 2.4 and 3.0 eV [5,6], it can absorb visible light more efficiently compared to other commonly studied photocatalytic semiconductors [7,8,9]. However, the electron–hole recombination rate tends to be higher compared to other semiconductors. On the other hand, copper oxide (CuO) is one of the most widely studied p-type oxides. It has an indirect band gap between 1.2 and 2.1 eV [10]. In order to overcome some of the transport disadvantages of many semiconductors and increase their photocatalytic efficiency, heterostructured oxide systems have been developed to reduce charge carrier recombination. Specifically, CuO has been used as a counterpart (p-type) for WO3 (n-type) in various applications [11,12,13,14,15,16], such as a catalyst in dye degradation processes, organic gas sensors, and solid oxide batteries [14,17,18]. CuO-WO3 nanofibers (NFs) fabricated via electrospinning method and tested for sarin gas sensor detection demonstrated that the detection performance of WO3-CuO NFs is better compared to WO3 and pure CuO and showed excellent selectivity toward trace amounts of sarin gas at room temperature from among several volatile organic compounds [12]. A hydrothermal synthesis of CuO-WO3 nanocubes for H2S gas sensing showed substantially greater and faster sensitivity and recovery than those of individual pure samples [19]. The increased surface area of the flower-like morphology, compared to the basic WO3–CuO composites, led to a higher density of active sites, which are essential for sensor performance. This morphological modification significantly enhanced H2S detection [16].
Copper tungstate (CuWO4) could be a key component in enhancing the efficiency of heterojunctions, particularly in CuO/WO3 systems. It promotes efficient electron transport and reduces recombination at the interface, while minimizing interfacial defects and enhancing charge carrier separation. CuWO4 can form high-quality interfaces when synthesized via in situ methods, reducing recombination and boosting charge transfer. Studies have shown that CuWO4-based systems exhibit superior performance in photoelectrochemical cells and gas sensors. CuWO4 also enhances light absorption and increases the active catalytic surface area, making it a valuable addition to multicomponent systems. Further exploration could lead to significant improvements in material performance [20]. The (CuO/WO3)-CuWO4 system has not been previously reported or systematically investigated for gas sensing applications.
In this work, pristine CuO, WO3, and (CuO/WO3)-CuWO4 were synthesized via a modified precipitation step. The materials were characterized using X-ray diffraction (XRD), Fourier transform infrared (FTIR) spectroscopy, scanning electron microscopy (SEM), and UV-Vis diffuse reflectance spectroscopy (UV-Vis DRS). XRD analysis, with the Rietveld refinement step, allowed the phase identification, quantification, and determination of lattice parameters. FTIR spectra provided confirmation of vibrational modes associated with Cu–O and W–O bonds. FE-SEM was used to directly observe the morphology at the nanometer scale, and DRS enabled the estimation of optical band gaps using the Kubelka–Munk model. Finally, the samples were tested as gas sensors by exposing the materials to methanol and acetone vapors in a controlled nitrogen gas flow. This study introduces, for the first time, the evaluation of various configurations of the (CuO/WO3)-CuWO4 heterostructure as a gas sensor.

2. Materials and Methods

2.1. Synthesis

The general synthesis process is illustrated in Figure 1. Initially, a stoichiometric quantity of high-purity copper (II) nitrate 3-hydrate (Cu(NO3)2⋅3H2O, 98–103%, PanReac AppliChem ITW reagents, Germany) was dissolved in double-distilled water (DDW). Subsequently, the pH was adjusted to 12 using a 2M sodium hydroxide (NaOH, 98%, PanReac AppliChem ITW reagents) solution, added dropwise. The resulting solution was aged for 0.5 h at 90 °C and 400 rpm. The resulting precipitate was centrifuged and washed with DDW until it reached a pH of 7. The resultant black material was dried for 1 h at 120 °C (A in Figure 1), then was ground into a black powder using an Agate mortar.
For the CuO sample, a portion of the material was heated in a Terrigeno D8-2471 oven (Colombia), with the temperature gradually increasing at a rate of 3 °C/min until reaching 500 °C, and maintained for 2 h. In the case of the heterostructure samples, a portion of the initial black powder was combined with a stoichiometric amount of tungstic acid (B Figure 1) (H2WO4, 99%, Sigma Aldrich, USA) to obtain mixtures containing 80, 60, 50, 40, 20, and 0 w/w % of copper precursor. The resulting samples were subjected to the same calcination ramp: 500 °C for 2 h, based on thermal stability and optimized as reported previously [21]. In the case of the WO3 sample, the same procedure was followed, omitting the copper precursor addition step.
Following the procedure outlined in Figure 1, a set of 7 different samples were obtained, which were labeled as follows in Table 1.

2.2. Characterization

The structural characterization of the samples was conducted through XRD analysis to identify the crystal structure and quantify the phase composition. An X’pert Pro Panalytical diffractometer (Amsterdam, The Netherlands) with a real-time multiple strip (RTMS) detector in Bragg–Brentano mode (Cu Kα radiation of λ = 1.542 Å) was used; the 2θ angle range was set between 20 and 80° with a step size of 0.026°. FTIR spectroscopy was used to confirm the presence of characteristic vibrational bands on the surface of the samples. A Shimadzu Prestige-21 IR spectrometer (Kyoto, Japan) with a resolution of 8 cm−1 and 8 runs was used for these measurements. Morphological details were observed through FE-SEM using a Tecnai F20 Super-Twin TMP microscope (United States) operating at an accelerating voltage of 200 kV, and the samples were dispersed in absolute ethanol and deposited on copper grids with 200 Mesh Lacey/Carbon membranes. Particle size distribution was estimated from FE-SEM images using ImageJ (1.53j) software by measuring at least 150 particles per sample. Finally, the determination of the band-gap energy of the samples was performed from the reflectance spectra obtained by UV-Vis DRS with a UV-Vis Cary 5000 spectrophotometer (Malaysia) in diffuse reflectance mode, using the Kubelka–Munk based on the absorbance data obtained.

2.3. Methanol and Acetone Sensing Test

The sensor responses were evaluated by depositing a mixture of the powders dispersed in isopropanol on substrates with interdigitated electrodes. The sensors were introduced in a stainless-steel test chamber for the sensing tests. The experimental bench for the electrical characterization of the sensors allows measurements to be carried out in a controlled atmosphere. Gases from certified cylinders were diluted in nitrogen (60 sccm) to the desired concentrations using mass flow controllers. Sensing measurements were carried out in the temperature range of 150–400 °C, with steps of 50 °C, under a gas (methanol or acetone) total stream of 30 sccm, collecting sensor resistance data in two-point mode under the applied bias of +10 V. A digital multimeter was used for data acquisition. The gas response is defined as the ratio Ra/Rg, where Ra is the electrical resistance of the sensor in dry air and Rg is the resistance at different methanol and acetone concentrations [22].

3. Results and Discussion

3.1. Structural Characterization

The structural properties, crystalline structure, and the phase composition of the samples were analyzed through XRD analysis. Rietveld refinement was conducted using GSAS-II software [23,24]. Figure 2 shows the plots generated from the refinement of the samples. The Rietveld adjustment showed a good agreement between the theoretical model and the experimental data, as confirmed by the low chi-squared ( χ 2 ) value (Table 2). This confirms the quality of the adjustment and the accuracy of phase identification. The CuO and WO3 contrast samples did not show the presence of secondary phases. All the samples with a copper precursor showed characteristic peaks around 35.5° and 38.8°, attributed to the diffraction planes, (−111), (111), respectively, and indexed to the monoclinic crystalline phase of CuO with PDF No. 01-072-0629 and the C2/c (15) space group. The presence of these characteristic peaks was reported previously [25,26,27].
Additionally, in the samples with a tungsten precursor, characteristic signals were observed around 23.1°, 23.6°, and 24.4° corresponding to the diffraction planes, (002), (020), and (200), respectively, and these signals are consistent with the monoclinic WO3 phase, PDF No. 01-072-1465 with the P21/n (14) space group system. These results are comparable to those reported by Kang et al. [28]. Finally, in all the heterostructure samples, the CuWO4 triclinic phase was observed, and characteristic signals around 19.2, 28.3, and 30.5° indexed to the diffraction planes, (100), (−1−11), and (111), respectively, confirm the presence of this phase, which matches PDF No. 01-070-1732 and the P-1 (2) space group, previously reported at this crystallization temperature as α-CuWO4 [29]. This additional phase is commonly formed due to the coexistence of copper and tungsten precursors. Its formation results from a secondary reaction between these compounds, occurring concurrently with the formation of CuO and WO3 phases [30]. The presence of these distinctive peaks confirms the coexistence of all the phases in the synthesized samples. Overall, the existence of distinct phases in the samples indicates the formation of a multicomponent system. This coexistence could indicate an interaction between the phases, potentially influencing the material properties [15,20].
The weight percentage of the phases in the synthesized samples, presented in Figure 3, reveals a clear variation in the phase composition among the samples. In Cu80:W20, monoclinic CuO was the majority phase with 88.81%, then a small proportion of CuWO4 (9.47%) and trace amounts of WO3 (1.72%). A significant rise in the percentage of the WO3 phase is seen when the copper precursor concentration decreases, reaching 43.17% in the Cu20:W80 and 100% in the pure WO3 sample. Conversely, the CuWO4 phase rises gradually to reach 40.2% in the Cu20:W80 sample, suggesting that the CuWO4 content stabilizes in the samples with lower CuO content. These differences in phase composition can directly influence the physicochemical properties of materials, particularly their sensitivity as gas sensors, given that the WO3 phase is known for its good response in this type of application [31,32]. The observed phase evolution suggests that controlling the initial composition is a key factor in tuning the final properties of the materials. In addition, the obtained lattice parameters (Table 2) suggest slight variations in the crystal structure of the phases, which could be related to the generation of structural defects or structural distortions inherent to the synthesis methods used.
Additionally, the average crystallite size (L) was determined using the Debye–Scherrer equation. In Equation (1), the Scherrer constant K = 0.94 ,   λ = 0.1540598   n m , β is the full width at half maximum (FWHM), and θ is the Bragg angle. In all the samples, the maximum intensity peak of each phase in the diffractogram was selected for crystallite size estimation. In addition, Rietveld L was also estimated in both perpendicular (L⊥) and parallel (L∥) directions with Equations (2) and (3), based on the Lorentzian component parameters (LX and ptec) from GSAS refinement. K and λ correspond to Scherrer parameters [33,34].
L = K λ β c o s θ
L = 1800   K λ π L X
L = 1800   K λ π ( L X + p t e c )
Figure 4A shows the variation in Scherrer and Rietveld L for CuO, WO3, and CuWO4 phases between the synthesized samples. The parallel and perpendicular Rietveld L mean values show similar trends to the results obtained by Scherrer. The results show that CuO phase crystallite size decreases with lower copper precursor content, with minimum values observed in Cu20:W80. WO3 phase crystallite size remains relatively constant between Cu80:W20 and Cu20:W80, with the lowest values found in Cu80:W20 and pure WO3 samples.
For the CuWO4 phase, crystallite size increases until reaching a maximum in the Cu40:W60 sample, then decreases in samples with lower copper content, such as Cu20:W80. The Rietveld refinement values provide additional information on particle anisotropy, with a noticeable increase at intermediate compositions, especially at Cu40:W60, where both CuWO4 and WO3 reach maximum crystallite sizes. This suggests that crystallite growth is more pronounced in this specific composition, possibly due to phase stabilization or interactions among the constituent oxides. The analysis of crystallite sizes demonstrates the influence of material composition on structural properties, particularly for the WO3 phase, known for its good gas detection performance [35].
The dislocation density ( δ ) and microstrain ( ε ) shown in Figure 4B were also determined with Equations (4) and (5), respectively, as follows in [36]. ε represents lattice resulting from imperfections or irregularities in the atomic structure [37]. The CuO and WO3 phases showed the highest microstrain, possibly due to their nanocrystalline nature and intrinsic properties. This strain could be favorable for applications like gas sensing and photocatalysis, where surface defects and lattice distortions enhance the reactivity and sensitivity [35]. δ is inversely related to crystallite size, affecting both mechanical and electronic properties. CuO shows a higher dislocation density compared to CuWO4 and WO3, indicating a higher density of internal defects associated with its smaller crystallite size. These dislocations can enhance mechanical strength and act as charge carrier traps, affecting the electrical properties of CuO and metal oxides in general, especially in semiconductor devices [38,39].
δ = 1 L 2
ε = B cos θ 4
The vibrational modes v of the synthesized materials were determined through FTIR spectroscopy. The experimental spectra, presented in Figure 5, reveal intense peaks in the region of 500–1100 cm−1. The red shifts (lower wavenumber displacements) and blue shifts (higher wavenumber displacements) presenting the different v n , are attributed to dislocations and structural defects in the materials [40]. This variation is consistent with the structural distortions observed in the XRD analysis. These results provide insight into the bonding environment within the heterostructures. The presence of multiple metal–oxygen stretching modes indicates the coexistence of Cu-O and W-O bonds; the observed vibrational bands are reported and assigned in Table 3. The v 6 band, corresponding to W-O stretching in tetrahedral coordination, confirms the successful formation of the ternary structure. This analysis supports the phase identification obtained from the XRD results.

3.2. Morphological Characterization

FE-SEM images (Figure 6) show the morphology of the CuO, WO3, Cu60:W40, and Cu50:W50 samples. The morphology of the samples varies significantly, and two predominant particle size distributions were observed for the copper- and tungsten oxide-based samples except for Cu80:W20. The CuO sample (Figure 6A) shows a porous, network-like structure composed of densely agglomerated nanoparticles. Sample Cu60:W40 (Figure 6B) shows more compact and well-defined polyhedral grains, attributed to the presence of the CuWO4 phase, which appears exclusively in samples containing the CuWO4 phase; agglomerated particles were also observed in the sample. The Cu50:W50 sample (Figure 6B) exhibits granular structures similar to those of Cu60:W40., and this morphology was constant across all the heterostructures samples, except Cu80:W20. These findings confirm that the phase composition directly influences the morphology of the phases as the composition of the system changes, and they also indicate that co-crystallization affects structural features. The sample WO3 (Figure 6D) shows irregular nanoparticle agglomeration with smooth surfaces. These morphological changes reflect differences in phase distribution, synthesis conditions, and particle growth mechanisms between samples, which influence both texture and particle size distribution. The particle size distributions for the samples are reported in Table 4.
The small particle sizes range from 80.9 nm to 130.8 nm, with the CuO sample exhibiting the largest average size (130.8 ± 30.6 nm) and the Cu80:W20 sample showing the smallest average (80.9 ± 28.3 nm). Large particles, on the other hand, exhibit more uniform sizes throughout the samples, ranging in size from 407.1 nm to 426.5 nm. For example, samples Cu60:W40, Cu50:W50, and Cu40:W60 contain large particles with comparable diameters around 407.1 nm, suggesting that particle growth in these compositions may be stabilized. This bimodal particle size distribution behavior might be attributed to the synthesis conditions and the heterogeneity of the samples, with various components and phases leading to the formation of particle sizes of varying sizes. Larger particles may result from the agglomeration or coalescence of smaller crystallites during growth, whereas the smaller particles correspond to isolated crystalline domains [46].

3.3. Optical Properties

The optical properties of the obtained samples were analyzed via UV-Vis DRS. Figure 7 presents the absorbance behavior of the samples, where the CuO sample exhibits the broadest absorption range, which extends approximately from 200 to 800 nm with a strong peak at 300 nm, and undergoes a red shift as the CuO content decreases in the heterostructured samples. In these, the absorption range broadens with increasing CuO content, with the Cu80:W20 and CuO samples showing the widest absorption range. This behavior suggests that CuO has a greater ability to absorb light along the visible spectrum due to its narrower band gap compared to the WO3 and CuWO4 phases. The WO3 sample exhibited a limited absorption range, which could be attributed to the absence of oxygen vacancies or defect states [47], which is consistent with previous reports [48]. Previous studies have shown that the formation of the CuWO4 phase also induces a red shift in the optical absorption spectrum. This shift is attributed to the narrowing of the band gap, which enhances light absorption at longer wavelengths, particularly in the visible region [20,49]. In general terms, these findings demonstrate that the heterostructure generated by different oxides outperforms the individual oxides in terms of optoelectronic application due to the optimized optical response within the visible spectrum.
The optical band gap energy ( E g ) of the samples was determined via UV-Vis DRS, using the Kubelka–Munk transform method, which assumes absorption and scattering as first-order phenomena [50,51]. The line segment x-intercept of the α h v 1 / n versus the energy of the incoming photon ( h v ) (Figure 8), derived from Equation (6), was used. In this, C is a proportionality constant, and n is a variable coefficient that changes based on the kind of electronic transition; direct permitted transitions and indirect permitted transitions have n = 1/2 and n = 2, respectively. h is the Planck constant and ν is the light frequency of the incident light [52,53,54].
α h v = C h v E g n
The approximate direct and indirect band-gap values of the samples are displayed in Table 5. The identification of different indirect band gaps suggests the coexistence of multiple semiconductor phases. This result, together with structural and morphological evidence, supports the formation of a heterostructure. From the CuO to Cu40:W60 samples, two different band gaps were observed, although three semiconductor phases coexist in the heterostructured samples (CuO, WO3, and CuWO4), and only two distinct optical transitions were observed in the UV–Vis DRS analysis. This is attributed to the close proximity of the indirect band gaps of WO3 (2.44 eV) and CuWO4 (~2.35 eV [20,55]), which likely leads to overlapping absorption features that cannot be individually resolved in Kubelka–Munk plots [56,57]. As a result, a single indirect transition is observed in that region. In contrast, the CuO phase exhibits a lower band gap (1.28 eV), which is clearly distinguishable. Therefore, the two observed band gaps correspond to CuO ( E g 1 ) and the combined contribution of WO3/CuWO4  ( E g 2 ) . The band gaps of some heterostructures were lower than those of the individual oxides after the simultaneous crystallization step, suggesting a potential extension of the system’s light absorption range. This suggests that the heterostructured samples are capable of absorbing visible light [15]. The optical response in the heterostructure is often enhanced by a reduced optical band gap. This narrower band gap makes it easier for electron–hole pairs to form and move between CuO, CuWO4, and WO3, which may improve the synergistic optical behavior of the combined oxides [56]. Accordingly, the CuO and Cu40:W60 samples exhibit promising optical performance.

3.4. Gas Sensing Test

The I–V (current vs. voltage) curves presented in Figure 9 exhibit typical ohmic behavior in the CuO, Cu20:W80, Cu40:W60, Cu50:W50, Cu60:W40, and Cu80:W20 samples, indicating low contact resistance and charge transport within the applied voltage range [12,58,59]. A device with this behavior will ensure functionality at low power and with minimum energy consumption. In contrast, the nonlinear behavior of the WO3 sample at higher voltages is mainly due to a space charge-limited current mechanism. In this case, for low values of voltage, the current exhibits near-ohmic behavior; however, upon increasing the voltage, charge traps within the material limit current flow, resulting in this phenomenon [60,61].
On the other hand, the analysis of resistance as a function of temperature (Figure 10) clearly shows that the resistance decreases with increasing temperature due to the granular nature of the material, which creates multiple potential barriers between grains, as well as depletion layers within the heterostructure channels. The optimum operating temperature of a sensor directly affects its resistivity by balancing the adsorption and desorption of gas molecules on the sensor surface, with higher resistance at moderate temperatures due to oxygen adsorption, whereas at higher temperatures, desorption reduces the sensor response. These findings agree well with metal oxide-based materials, which constitute one of the most studied families of materials in gas sensors due to their high sensitivity and thermal stability [59,62,63].
In the gas sensor analysis, the sample sensors showed a good response to methanol/acetone vapors. Figure 11 shows the resistance profile of all samples exposed to a fixed vapor concentration in N2 over time. Exposure to methanol (Figure 11A) and acetone (Figure 11B) induces a decrease in resistance, characteristic of the semiconductor’s response to reducing gases. The novel heterostructure system increases the number active sites, which are essential for gas sensing performance, observing a higher response or resistance change in the Cu40:W60 and Cu50:W50 samples. This enhanced behavior is attributed to the multicomponent nature of the sensor, made from metal oxides, which works on the principle of heterostructures between p-type and n-type materials, such as CuO and WO3-CuWO4, respectively. In this type of sensor, upon being brought into contact with a p-type material with an n-type material, an electron depletion layer forms in the n-type material, while a hole depletion layer forms in the p-type material, resulting in the formation of a potential barrier at the interface [62].
In contact with oxygen, the sensor surface promotes its adsorption, forming adsorbed oxygen species ( O 2 , O , O 2 ), which subsequently capture electrons from the n-type materials (WO3 and CuWO4), leading to an expansion of the depletion layer and an increase in sensor resistance. When exposed to reducing gases such as acetone and methanol, the oxygen ions combine with such gases, releasing electrons, which reduces both the potential barrier and the depletion layer width, thereby decreasing the resistance. The gases react with the oxygen species on the material surface, producing byproducts such as CO2 and H2O, among others, depending on the gas’s nature [47,63,64,65]. A schematic representation of this mechanism is shown in Figure 12.
Figure 13 shows how a gas sensor measures the change in resistance from its initial state (Ra) to its state after gas exposure (Rg), enabling the calculation of the sensor response as Ra/Rg for the detected gas. The improved behavior observed in the p–n heterostructure exhibits a faster and more sensitive response and recovery when compared to materials without this type of configuration. For the most responsive material configuration (Cu40:W60), the response to methanol was higher than acetone, 2.2 and 1.7, respectively. This difference is usually reported for semiconductor-based materials [66]. For the CuO nanowires, the ethanol response was reported as 1.5 [67] and 1.4 for methanol [66], and for the WO3 thin films, the response was 1.5 for ethanol [68]. The enhanced response of the heterostructured system over the pristine WO3 sample and high WO3 phase percentage (Cu80:W20) sample shows a superior performance in samples with higher CuO content (p-type component), as previously observed in CuO/WO3 heterostructures [69].

4. Conclusions

Gas sensing measurements performed on (CuO/WO3)-CuWO4 heterostructures show promising results in detecting methanol and acetone vapors. The ohmic behavior observed in most samples through the I-V curves indicates low contact resistance and good electrical conductivity, which is essential for the operation of low-power sensors. The heterostructure samples, especially the Cu40:W60 and Cu50:W50 samples, exhibited an improved gas sensitivity, which was tuned by n-type and p-type sensing mechanisms that promote electron transfer between the p-type CuO and n-type WO3-CuWO4 components. This configuration increases the number of active sensing sites, leading to significant changes in resistance upon exposure to reducing gases such as methanol and acetone. The Cu40:W60 sample exhibited excellent sensitivity, especially toward methanol, with a response ratio of 2.2 compared to 1.7 for acetone. These results highlight the potential application of (CuO/WO3)-CuWO4 heterostructures in gas sensing due to their high sensitivity, good stability, and rapid response times.

Author Contributions

Conceptualization, investigation, methodology, formal analysis, resources, A.J.C.; Conceptualization, investigation, material preparation, methodology, formal analysis, writing—original draft, data curation, review and editing, funding acquisition, M.C.M.; Conceptualization, investigation, supervision, funding acquisition, C.A.P.V.; Conceptualization, investigation, characterization, resources, M.R.J.; conceptualization, investigation, supervision, material characterization, methodology development, and formal analysis, resources, funding acquisition, review and editing, A.M.R.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Research Directorate of the Pedagogical and Technological University of Colombia SGI 3804 and the Francisco de Paula Santander University grant FINU 010-2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge the institutional support provided by the Pedagogical and Technological University of Colombia, Francisco de Paula Santander University, and the National University of Colombia.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Hao, R.; Sun, J.; Liu, R.; Zhao, H.; Yao, Z.; Wang, H.; Hao, Z. Emission characteristics, environmental impact, and health risk assessment of volatile organic compounds (VOCs) during manicure processes. Sci. Total Environ. 2024, 906, 167464. [Google Scholar] [CrossRef] [PubMed]
  2. Dong, P.; Hou, G.; Xi, X.; Shao, R.; Dong, F. WO3-based photocatalysts: Morphology control, activity enhancement and multifunctional applications. Environ. Sci. Nano 2017, 4, 539–557. [Google Scholar] [CrossRef]
  3. Quan, H.; Gao, Y.; Wang, W. Tungsten oxide-based visible light-driven photocatalysts: Crystal and electronic structures and strategies for photocatalytic efficiency enhancement. Inorg. Chem. Front. 2020, 7, 817–838. [Google Scholar] [CrossRef]
  4. Adhikari, S.; Kim, D.-H.; Madras, G.; Sarkar, D. Understanding the morphological effects of WO3 photocatalysts for the degradation of organic pollutants. Adv. Powder Technol. 2018, 29, 1591–1600. [Google Scholar] [CrossRef]
  5. Gomis-Berenguer, A.; Celorrio, V.; Iniesta, J.; Fermin, D.J.; Ania, C.O. Nanoporous carbon/WO3 anodes for an enhanced water photooxidation. Carbon 2016, 108, 471–479. [Google Scholar] [CrossRef]
  6. Wang, L.; Hu, H.; Xu, J.; Zhu, S.; Ding, A.; Deng, C. WO3 nanocubes: Hydrothermal synthesis, growth mechanism, and photocatalytic performance. J. Mater. Res. 2019, 34, 2955–2963. [Google Scholar] [CrossRef]
  7. Abe, R.; Takami, H.; Murakami, N.; Ohtani, B. Pristine simple oxides as visible light driven photocatalysts: Highly efficient decomposition of organic compounds over platinum-loaded tungsten oxide. J. Am. Chem. Soc. 2008, 130, 7780–7781. [Google Scholar] [CrossRef]
  8. Dursun, S.; Kaya, İ.C.; Kocabaş, M.; Akyildiz, H.; Kalem, V. Visible light active heterostructured photocatalyst system based on CuO plate-like particles and SnO 2 nanofibers. Int. J. Appl. Ceram. Technol. 2020, 17, 1479–1489. [Google Scholar] [CrossRef]
  9. Çinar, B.; Keri̇moğlu, I.; Tönbül, B.; Demi̇rbüken, A.; Dursun, S.; Cihan Kaya, I.; Kalem, V.; Akyildiz, H. Hydrothermal/electrospinning synthesis of CuO plate-like particles/TiO2 fibers heterostructures for high-efficiency photocatalytic degradation of organic dyes and phenolic pollutants. Mater. Sci. Semicond. Process. 2020, 109, 104919. [Google Scholar] [CrossRef]
  10. Li, X.; Yu, J.; Jaroniec, M. Hierarchical photocatalysts. Chem. Soc. Rev. 2016, 45, 2603–2636. [Google Scholar] [CrossRef] [PubMed]
  11. Alizadeh, N.; Salimi, A.; Hallaj, R.; Fathi, F.; Soleimani, F. CuO/WO3 nanoparticles decorated graphene oxide nanosheets with enhanced peroxidase-like activity for electrochemical cancer cell detection and targeted therapeutics. Mater. Sci. Eng. C 2019, 99, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  12. Alali, K.T.; Liu, J.; Aljebawi, K.; Liu, P.; Chen, R.; Li, R.; Zhang, H.; Zhou, L.; Wang, J. Electrospun n-p WO3/CuO heterostructure nanofibers as an efficient sarin nerve agent sensing material at room temperature. J. Alloys Compd. 2019, 793, 31–41. [Google Scholar] [CrossRef]
  13. Wang, C.; Tang, J.; Zhang, X.; Qian, L.; Yang, H. WO3 nanoflakes decorated with CuO clusters for enhanced photoelectrochemical water splitting. Prog. Nat. Sci. Mater. Int. 2018, 28, 200–204. [Google Scholar] [CrossRef]
  14. Dursun, S.; Koyuncu, S.N.; Kaya, İ.C.; Kaya, G.G.; Kalem, V.; Akyildiz, H. Production of CuO–WO3 hybrids and their dye removal capacity/performance from wastewater by adsorption/photocatalysis. J. Water Process Eng. 2020, 36, 101390. [Google Scholar] [CrossRef]
  15. Wang, H.; Xiao, M.; Wang, Z.; Chen, X.; Dai, W.; Fu, X. Visible photocatalytic hydrogen production from CH3OH over CuO/WO3: The effect of electron transfer behavior of the adsorbed CH3OH. Chem. Eng. J. 2023, 459, 141616. [Google Scholar] [CrossRef]
  16. He, M.; Xie, L.; Zhao, X.; Hu, X.; Li, S.; Zhu, Z.-G. Highly sensitive and selective H2S gas sensors based on flower-like WO3/CuO composites operating at low/room temperature. J. Alloys Compd. 2019, 788, 36–43. [Google Scholar] [CrossRef]
  17. Zappa, D.; Galstyan, V.; Kaur, N.; Munasinghe Arachchige, H.M.M.; Sisman, O.; Comini, E. “Metal oxide—Based heterostructures for gas sensors”—A review. Anal. Chim. Acta 2018, 1039, 1–23. [Google Scholar] [CrossRef]
  18. Danish, M.S.S.; Estrella-Pajulas, L.; Alemaida, I.; Lisin, A.; Moiseev, N.; Ahmadi, M.; Nazari, M.; Wali, M.; Zaheb, H.; Senjyu, T. Photocatalytic Applications of Metal Oxides for Sustainable Environmental Remediation. Met.—Open Access Metall. J. 2021, 11, 80. [Google Scholar] [CrossRef]
  19. Yu, W.; Sun, Y.; Zhang, T.; Zhang, K.; Wang, S.; Chen, X.; Dai, N. CuO/WO3 Hybrid Nanocubes for High-Responsivity and Fast-Recovery H2S Sensors Operated at Low Temperature. Part. Part. Syst. Charact. 2016, 33, 15–20. [Google Scholar] [CrossRef]
  20. Li, J.; Hu, S.; Liu, S.; Hou, S.; Li, L.; Huang, J. In situ fabrication of WO3/CuWO4/CuO heterojunction photoanode for boosted interfacial charge transfer and enhanced photoelectrochemical water splitting. Int. J. Hydrog. Energy 2024, 61, 967–974. [Google Scholar] [CrossRef]
  21. Raba-Páez, A.M.; Malafatti, J.O.D.; Parra-Vargas, C.A.; Paris, E.C.; Rincón-Joya, M. Effect of tungsten doping on the structural, morphological and bactericidal properties of nanostructured CuO. PLoS ONE 2020, 15, e0239868. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, J.; Wan, Q. Gas Sensors Based on Semiconducting Metal Oxide One-Dimensional Nanostructures. Sensors 2009, 9, 9903. [Google Scholar] [CrossRef] [PubMed]
  23. Gaona, I.M.S.; Mendoza, M.C.; Vargas, C.A.P. Structural and Magnetic Properties of Nd3Ba5Cu8O18+ẟ Superconductor. J. Low Temp. Phys. 2023, 211, 156–165. [Google Scholar] [CrossRef]
  24. Saavedra Gaona, I.M.; Supelano, G.I.; Suarez Vera, S.G.; Fonseca, L.C.I.; Castaneda Mendoza, M.; Sánchez Saenz, C.L.; Izquierdo, J.L.; Gómez, A.; Morán, O.; Parra Vargas, C.A. Magnetic and electrical behaviour of Yb substitution on Bi1-xYbxFeO3 (0.00 < x < 0.06) ceramic system. J. Magn. Magn. Mater. 2024, 593, 171827. [Google Scholar]
  25. Pallavolu, M.-R.; Banerjee, A.-N.; Joo, S.-W. Battery-Type Behavior of Al-Doped CuO Nanoflakes to Fabricate a High-Performance Hybrid Supercapacitor Device for Superior Energy Storage Applications. Coatings 2023, 13, 1337. [Google Scholar] [CrossRef]
  26. Jansanthea, P.; Inyai, N.; Chomkitichai, W.; Ketwaraporn, J.; Ubolsook, P.; Wansao, C.; Wanaek, A.; Wannawek, A.; Kuimalee, S.; Pookmanee, P. Green synthesis of CuO/Fe2O3/ZnO ternary composite photocatalyst using grape extract for enhanced photodegradation of environmental organic pollutant. Chemosphere 2024, 351, 141212. [Google Scholar] [CrossRef]
  27. Abdelkarem, K.; Saad, R.; Ahmed, A.M.; Fathy, M.I.; Shaban, M.; Hamdy, H. Efficient room temperature carbon dioxide gas sensor based on barium doped CuO thin films. J. Mater. Sci. 2023, 58, 11568–11584. [Google Scholar] [CrossRef]
  28. Kang, M.; Liang, J.; Wang, F.; Chen, X.; Lu, Y.; Zhang, J. Structural design of hexagonal/monoclinic WO3 phase junction for photocatalytic degradation. Mater. Res. Bull. 2020, 121, 110614. [Google Scholar] [CrossRef]
  29. Reis, L.R.M.; Costa, M.J.S.; Oliveira, Y.L.; Santos, R.S.; Sczancoski, J.C.; Cavalcante, L.S. Structure, optical, colorimetric, and supercapacitor properties of anode α-CuWO4 crystals. Mater. Lett. 2024, 354, 135340. [Google Scholar] [CrossRef]
  30. Ohyama, J.; Iwai, H.; Takahashi, D.; Tsushida, M.; Machida, M.; Nishimura, S.; Takahashi, K. Improved Catalytic Partial Oxidation of Methane via Lattice Oxygen Modification on Supported Copper Oxide Catalyst System. ChemCatChem 2024, 16, e202401045. [Google Scholar] [CrossRef]
  31. Zou, Z.; Zhao, Z.; Zhang, Z.; Tian, W.; Yang, C.; Jin, X.; Zhang, K. Room-Temperature Optoelectronic Gas Sensor Based on Core–Shell g-C3N4@WO3 Heterocomposites for Efficient Ammonia Detection. Anal. Chem. 2023, 95, 2110–2118. [Google Scholar] [CrossRef] [PubMed]
  32. Jeong, Y.; Hong, S.; Jung, G.; Shin, W.; Lee, C.; Park, J.; Kim, D.; Lee, J.-H. Effects of oxygen gas in the sputtering process of the WO3 sensing layer on NO2 sensing characteristics of the FET-type gas sensor. Solid-State Electron. 2023, 200, 108563. [Google Scholar] [CrossRef]
  33. Murugesan, S.; Thirumurugesan, R.; Mohandas, E.; Parameswaran, P. X-ray diffraction Rietveld analysis and Bond Valence analysis of nano titania containing oxygen vacancies synthesized via sol-gel route. Mater. Chem. Phys. 2019, 225, 320–330. [Google Scholar] [CrossRef]
  34. Cuervo Farfán, J.A. Producción y Propiedades Físicas de Nuevas Perovskitas Complejas del Tipo RAMOX (R = La, Nd, Sm, Eu; A = Sr, Bi; M = Ti, Mn, Fe). Ph.D. Thesis, Universidad Nacional de Colombia, Bogotá, Colombia, 2021. Available online: https://repositorio.unal.edu.co/handle/unal/79915 (accessed on 15 August 2024).
  35. Li, X.; Fu, L.; Karimi-Maleh, H.; Chen, F.; Zhao, S. Innovations in WO3 gas sensors: Nanostructure engineering, functionalization, and future perspectives. Heliyon 2024, 10, e27740. [Google Scholar] [CrossRef]
  36. Sen, S.K.; Dutta, S.; Paik, L.; Paul, T.C.; Manir, M.S.; Hossain, M.; Hossain, M.N. Dy-doped MoO3 nanobelts synthesized via hydrothermal route: Influence of Dy contents on the structural, morphological and optical properties. J. Alloys Compd. 2021, 876, 160070. [Google Scholar] [CrossRef]
  37. Sutapa, I.W.; Wahid Wahab, A.; Taba, P.; Nafie, N.L. Dislocation, crystallite size distribution and lattice strain of magnesium oxide nanoparticles. J. Phys. Conf. Ser. 2018, 979, 012021. [Google Scholar] [CrossRef]
  38. Jia, Y.; Zhou, K.; Sun, W.; Ding, M.; Wang, Y.; Kong, X.; Jia, D.; Wu, M.; Fu, Y. Enhancement mechanisms of mechanical, electrical and thermal properties of carbon nanotube-copper composites: A review. J. Mater. Res. Technol. 2024, 32, 1395–1415. [Google Scholar] [CrossRef]
  39. Hegde, V.N.; V, M.V.; M, P.T.; C, H.B. Study on structural, morphological, elastic and electrical properties of ZnO nanoparticles for electronic device applications. J. Sci. Adv. Mater. Devices 2024, 9, 100733. [Google Scholar] [CrossRef]
  40. Mariammal, R.N.; Ramachandran, K.; Kalaiselvan, G.; Arumugam, S.; Renganathan, B.; Sastikumar, D. Effect of magnetism on the ethanol sensitivity of undoped and Mn-doped CuO nanoflakes. Appl. Surf. Sci. 2013, 270, 545–552. [Google Scholar] [CrossRef]
  41. Vanasundari, K.; Ponnarasi, P.; Mahalakshmi, G. A eco-friendly, green synthesis of Ag loaded WO3/rGO nanocomposites for effective UV light photocatalytic degradation of 4-nitrophenol and antimicrobial activity. Diam. Relat. Mater. 2024, 142, 110859. [Google Scholar] [CrossRef]
  42. Fatima, R.; Warsi, M.F.; Sarwar, M.I.; Shakir, I.; Agboola, P.O.; Aly Aboud, M.F.; Zulfiqar, S. Synthesis and Characterization of Hetero-metallic Oxides-Reduced Graphene Oxide Nanocomposites for Photocatalytic Applications. Ceram. Int. 2021, 47, 7642–7652. [Google Scholar] [CrossRef]
  43. Capeli, R.A.; Belmonte, T.; Caierão, J.; Dalmaschio, C.J.; Teixeira, S.R.; Mastelaro, V.R.; Chiquito, A.J.; Teodoro, M.D.; Domenegueti, J.F.M.; Longo, E.; et al. Effect of hydrothermal temperature on the antibacterial and photocatalytic activity of WO3 decorated with silver nanoparticles. J. Sol-Gel Sci. Technol. 2021, 97, 228–244. [Google Scholar] [CrossRef]
  44. Pourmortazavi, S.M.; Rahimi-Nasrabadi, M.; Khalilian-Shalamzari, M.; Ghaeni, H.R.; Hajimirsadeghi, S.S. Facile Chemical Synthesis and Characterization of Copper Tungstate Nanoparticles. J. Inorg. Organomet. Polym. 2014, 24, 333–339. [Google Scholar] [CrossRef]
  45. Sreekanth, T.V.M.; Prasad, K.; Yoo, J.; Kim, J.; Yoo, K. CuWO4 as a cost-effective electrocatalyst for urea oxidation reaction. Inorg. Chem. Commun. 2023, 154, 110933. [Google Scholar] [CrossRef]
  46. Thanh, N.T.K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. 2014, 114, 7610–7630. [Google Scholar] [CrossRef]
  47. Xue, S.; Cao, S.; Huang, Z.; Yang, D.; Zhang, G. Improving Gas-Sensing Performance Based on MOS Nanomaterials: A Review. Materials 2021, 14, 4263. [Google Scholar] [CrossRef]
  48. Mohammed Harshulkhan, S.; Janaki, K.; Velraj, G.; Sakthi Ganapthy, R.; Nagarajan, M. Effect of Ag doping on structural, optical and photocatalytic activity of tungsten oxide (WO3) nanoparticles. J. Mater. Sci: Mater. Electron. 2016, 27, 4744–4751. [Google Scholar] [CrossRef]
  49. Wang, Z.; Wang, X.; Wang, H.; Chen, X.; Dai, W.; Fu, X. The role of electron transfer behavior induced by CO chemisorption on visible-light-driven CO conversion over WO3 and CuWO4/WO3. Appl. Catal. B Environ. 2020, 265, 118588. [Google Scholar] [CrossRef]
  50. Landi, S.; Segundo, I.R.; Freitas, E.; Vasilevskiy, M.; Carneiro, J.; Tavares, C.J. Use and misuse of the Kubelka-Munk function to obtain the band gap energy from diffuse reflectance measurements. Solid State Commun. 2022, 341, 114573. [Google Scholar] [CrossRef]
  51. Asiri, A.M.; Nawaz, T.; Tahir, M.B.; Fatima, N.; Khan, S.B.; Alamry, K.A.; Alfifi, S.Y.; Marwani, H.M.; Al-Otaibi, M.M.; Chakraborty, S. Fabrication of WO3 based nanocomposites for the excellent photocatalytic energy production under visible light irradiation. Int. J. Hydrog. Energy 2021, 46, 39058–39066. [Google Scholar] [CrossRef]
  52. Escobedo-Morales, A.; Ruiz-López, I.I.; Ruiz-Peralta, M.D.; Tepech-Carrillo, L.; Sánchez-Cantú, M.; Moreno-Orea, J.E. Automated method for the determination of the band gap energy of pure and mixed powder samples using diffuse reflectance spectroscopy. Heliyon 2019, 5, e01505. [Google Scholar] [CrossRef] [PubMed]
  53. Jubu, P.R.; Obaseki, O.S.; Yam, F.K.; Stephen, S.M.; Avaa, A.A.; McAsule, A.A.; Yusof, Y.; Otor, D.A. Influence of the secondary absorption and the vertical axis scale of the Tauc’s plot on optical bandgap energy. J. Opt. 2023, 52, 1426–1435. [Google Scholar] [CrossRef]
  54. Jubu, P.R.; Obaseki, O.S.; Ajayi, D.I.; Danladi, E.; Chahrour, K.M.; Muhammad, A.; Landi, S.; Igbawua, T.; Chahul, H.F.; Yam, F.K. Considerations about the determination of optical bandgap from diffuse reflectance spectroscopy using the tauc plot. J. Opt. 2024, 53, 5054–5064. [Google Scholar] [CrossRef]
  55. Zhang, H.; Yilmaz, P.; Ansari, J.O.; Khan, F.F.; Binions, R.; Krause, S.; Dunn, S. Incorporation of Ag nanowires in CuWO4 for improved visible light-induced photoanode performance. J. Mater. Chem. A 2015, 3, 9638–9644. [Google Scholar] [CrossRef]
  56. Raba-Páez, A.M.; Malafatti, J.O.D.; Parra-Vargas, C.A.; Paris, E.C.; Rincón-Joya, M. Structural evolution, optical properties, and photocatalytic performance of copper and tungsten heterostructure materials. Mater. Today Commun. 2021, 26, 101886. [Google Scholar] [CrossRef]
  57. Morales-Morales, G.; Manzanares-Martinez, J. Enlargement of band gaps on thermal wave crystals by using heterostructures. Results Phys. 2022, 42, 106019. [Google Scholar] [CrossRef]
  58. Yao, Y.; Sang, D.; Zou, L.; Wang, Q.; Liu, C. A Review on the Properties and Applications of WO3 Nanostructure-Based Optical and Electronic Devices. Nanomaterials 2021, 11, 2136. [Google Scholar] [CrossRef]
  59. Wei, Z.; Zhou, Q.; Lu, Z.; Xu, L.; Gui, Y.; Tang, C. Morphology controllable synthesis of hierarchical WO3 nanostructures and C2H2 sensing properties. Phys. E Low-Dimens. Syst. Nanostructures 2019, 109, 253–260. [Google Scholar] [CrossRef]
  60. Wang, L.; Cheng, S.; Wu, C.; Pei, K.; Song, Y.; Li, H.; Wang, Q.; Sang, D. Fabrication and high temperature electronic behaviors of n-WO3 nanorods/p-diamond heterojunction. Appl. Phys. Lett. 2017, 110, 052106. [Google Scholar] [CrossRef]
  61. Shen, Z.; Peng, Z.; Zhao, Z.; Fu, X. Nonlinear current-voltage characteristics of WO3-x nano-/micro-rods. Solid State Sci. 2018, 78, 126–132. [Google Scholar] [CrossRef]
  62. Krishna, K.G.; Parne, S.; Pothukanuri, N.; Kathirvelu, V.; Gandi, S.; Joshi, D. Nanostructured metal oxide semiconductor-based gas sensors: A comprehensive review. Sens. Actuators A Phys. 2022, 341, 113578. [Google Scholar] [CrossRef]
  63. Cho, S.-Y.; Jang, D.; Kang, H.; Koh, H.-J.; Choi, J.; Jung, H.-T. Ten Nanometer Scale WO3/CuO Heterojunction Nanochannel for an Ultrasensitive Chemical Sensor. Anal. Chem. 2019, 91, 6850–6858. [Google Scholar] [CrossRef] [PubMed]
  64. Goel, N.; Kunal, K.; Kushwaha, A.; Kumar, M. Metal oxide semiconductors for gas sensing. Eng. Rep. 2023, 5, e12604. [Google Scholar] [CrossRef]
  65. Karnati, P.; Akbar, S.; Morris, P.A. Conduction mechanisms in one dimensional core-shell nanostructures for gas sensing: A review. Sens. Actuators B Chem. 2019, 295, 127–143. [Google Scholar] [CrossRef]
  66. Gou, X.; Wang, G.; Yang, J.; Park, J.; Wexler, D. Chemical synthesis, characterisation and gas sensing performance of copper oxide nanoribbons. J. Mater. Chem. 2008, 18, 965–969. [Google Scholar] [CrossRef]
  67. Raksa, P.; Gardchareon, A.; Chairuangsri, T.; Mangkorntong, P.; Mangkorntong, N.; Choopun, S. Ethanol sensing properties of CuO nanowires prepared by an oxidation reaction. Ceram. Int. 2009, 35, 649–652. [Google Scholar] [CrossRef]
  68. Ahmad, M.Z.; Kang, J.H.; Sadek, A.Z.; Moafi, A.; Sberveglieri, G.; Wlodarski, W. Synthesis of WO3 Nanorod based Thin Films for Ethanol and H2 Sensing. Procedia Eng. 2012, 47, 358–361. [Google Scholar] [CrossRef]
  69. Yang, F.; Wang, F.; Guo, Z. Characteristics of binary WO3@CuO and ternary WO3@PDA@CuO based on impressive sensing acetone odor. J. Colloid Interface Sci. 2018, 524, 32–41. [Google Scholar] [CrossRef]
Figure 1. Schematic general synthesis process, where A = copper precursor, B = tungsten precursor, and x corresponds to the mass fraction of the Cu−based precursor (A) and W−based precursor (B). For example, A0.5B0.5 represents a 50:50 weight ratio.
Figure 1. Schematic general synthesis process, where A = copper precursor, B = tungsten precursor, and x corresponds to the mass fraction of the Cu−based precursor (A) and W−based precursor (B). For example, A0.5B0.5 represents a 50:50 weight ratio.
Materials 18 02896 g001
Figure 2. The Rietveld refinement plots of the WO3 (A), Cu20:W80 (B), Cu40:W60 (C), Cu50:W50 (D), Cu60:W40 (E), Cu80:W20 (F), and CuO (G) samples.
Figure 2. The Rietveld refinement plots of the WO3 (A), Cu20:W80 (B), Cu40:W60 (C), Cu50:W50 (D), Cu60:W40 (E), Cu80:W20 (F), and CuO (G) samples.
Materials 18 02896 g002
Figure 3. Weight percentage of the phases in the synthesized samples.
Figure 3. Weight percentage of the phases in the synthesized samples.
Materials 18 02896 g003
Figure 4. Variation in Scherrer and Rietveld crystallite size (A), dislocation density, and microstrain (B) in synthesized samples.
Figure 4. Variation in Scherrer and Rietveld crystallite size (A), dislocation density, and microstrain (B) in synthesized samples.
Materials 18 02896 g004
Figure 5. FTIR spectra of the synthesized samples in the region of 500 to 1100 cm−1.
Figure 5. FTIR spectra of the synthesized samples in the region of 500 to 1100 cm−1.
Materials 18 02896 g005
Figure 6. FE-SEM images at 20–50 kx and particle size distributions estimated from SEM image analysis using ImageJ of (A) CuO, (B) Cu60:W40 (Orange—Small Particle Size; Gray—Large Particle Size (nm)), (C) Cu50:W50, and (D) WO3 samples.
Figure 6. FE-SEM images at 20–50 kx and particle size distributions estimated from SEM image analysis using ImageJ of (A) CuO, (B) Cu60:W40 (Orange—Small Particle Size; Gray—Large Particle Size (nm)), (C) Cu50:W50, and (D) WO3 samples.
Materials 18 02896 g006
Figure 7. The UV-Vis DRS spectra of the synthesized materials.
Figure 7. The UV-Vis DRS spectra of the synthesized materials.
Materials 18 02896 g007
Figure 8. (A) Direct ( α h v )2 and (B) indirect ( α h v )1/2 transitions vs. photon energy. The dotted line indicates the linear fit used to estimate the optical band gap.
Figure 8. (A) Direct ( α h v )2 and (B) indirect ( α h v )1/2 transitions vs. photon energy. The dotted line indicates the linear fit used to estimate the optical band gap.
Materials 18 02896 g008
Figure 9. Current in function of voltage curves of synthesized samples at 300 K.
Figure 9. Current in function of voltage curves of synthesized samples at 300 K.
Materials 18 02896 g009
Figure 10. Variation in sample resistance with temperature.
Figure 10. Variation in sample resistance with temperature.
Materials 18 02896 g010
Figure 11. Resistance changes in sample sensors for methanol (A) and acetone (B) vapors at 350 K.
Figure 11. Resistance changes in sample sensors for methanol (A) and acetone (B) vapors at 350 K.
Materials 18 02896 g011
Figure 12. Schematic representation of methanol sensing mechanism in (CuO/WO3) − CuWO4 heterostructure.
Figure 12. Schematic representation of methanol sensing mechanism in (CuO/WO3) − CuWO4 heterostructure.
Materials 18 02896 g012
Figure 13. Cu40:W60 sample response in methanol and acetone gas sensor test.
Figure 13. Cu40:W60 sample response in methanol and acetone gas sensor test.
Materials 18 02896 g013
Table 1. Sample labeling.
Table 1. Sample labeling.
LabelWeight Percentage (w-w %)
Copper Precursor (A)Tungsten Precursor (B)
WO30100
Cu20:W802080
Cu40:W604060
Cu50:W505050
Cu60:W406040
Cu80:W208020
CuO1000
Table 2. Rietveld refinement parameters of the synthesized samples.
Table 2. Rietveld refinement parameters of the synthesized samples.
PhaseCuOCu80:W20Cu60:W40Cu50:W50Cu40:W60Cu20:W80WO3
a (Å)CuO4.694.694.684.694.694.68-
WO3-7.507.307.317.317.337.32
CuWO4-4.704.704.714.714.70-
b (Å)CuO3.433.433.433.423.423.40-
WO3-7.607.527.537.537.567.54
CuWO4-5.845.835.845.845.83-
c (Å)CuO5.135.145.135.145.135.12-
WO3-7.657.677.697.697.727.71
CuWO4-4.884.874.884.884.88-
V (Å3)CuO82.5282.6982.3582.4482.2881.47-
WO3-436.05421.05423.29423.29427.80425.54
CuWO4-133.95133.44134.23134.23133.72-
α (°)CuO90-
WO3-90
CuWO4-91.6791.6691.6691.6591.66-
β (°)CuO99.5299.5299.5599.5499.5499.60-
WO3-90.4290.5090.4590.4290.4490.56
CuWO4-92.5192.5092.5092.5092.50-
γ (°)CuO90-
WO3-90
CuWO4-82.7882.7982.7882.7882.79-
Chi ( χ 2 ) 1.281.781.851.481.922.811.70
Table 3. IR band assignment for the vibrational bands.
Table 3. IR band assignment for the vibrational bands.
Wavenumber Range (cm⁻¹)AssignationReference
v 1 555–560Metal–oxygen–metal bond vibrations (W-O-W, Cu-O-Cu)[41,42]
v 2 590–605
v 3 620–650W-O-W stretch vibration modes[14]
v 4 670–720Vibration of O-W-O-W-O flexion[43]
v 5 800–810O-W-O stretch modes[14]
v 6 870–920W-O bond stretch mode in the octahedral WO6 structure (CuWO4)[29,44,45]
Table 4. Particle size distributions of the samples.
Table 4. Particle size distributions of the samples.
SampleSmall Particle Size (nm)Large Particle Size (nm)
CuO130.8 ± 30.6-
Cu80:W2080.9 ± 28.3-
Cu60:W40100.5 ± 63.3407.5 ± 95.2
Cu50:W5096.3 ± 47.9408.8 ± 80.6
Cu40:W6084.9 ± 24.9407.1 ± 81.2
Cu20:W8082.6 ± 45.9426.5 ± 159.3
WO399.7 ± 47.5-
Table 5. Variations in the direct and indirect optical band gaps of the samples.
Table 5. Variations in the direct and indirect optical band gaps of the samples.
Direct Gap (eV)Indirect Gap (eV)
E g 1 E g 1 E g 2
CuO2.441.28-
Cu80:W202.681.39-
Cu60:W402.431.481.62
Cu50:W502.451.481.60
Cu40:W602.461.331.93
Cu20:W802.72-2.04
WO32.96-2.44
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

Castaneda Mendoza, M.; Parra Vargas, C.A.; Rincón Joya, M.; Chiquito, A.J.; Raba-Páez, A.M. Gas Sensor Properties of (CuO/WO3)-CuWO4 Heterostructured Nanocomposite Materials. Materials 2025, 18, 2896. https://doi.org/10.3390/ma18122896

AMA Style

Castaneda Mendoza M, Parra Vargas CA, Rincón Joya M, Chiquito AJ, Raba-Páez AM. Gas Sensor Properties of (CuO/WO3)-CuWO4 Heterostructured Nanocomposite Materials. Materials. 2025; 18(12):2896. https://doi.org/10.3390/ma18122896

Chicago/Turabian Style

Castaneda Mendoza, Michael, Carlos A. Parra Vargas, Miryam Rincón Joya, Adenilson J. Chiquito, and Angela M. Raba-Páez. 2025. "Gas Sensor Properties of (CuO/WO3)-CuWO4 Heterostructured Nanocomposite Materials" Materials 18, no. 12: 2896. https://doi.org/10.3390/ma18122896

APA Style

Castaneda Mendoza, M., Parra Vargas, C. A., Rincón Joya, M., Chiquito, A. J., & Raba-Páez, A. M. (2025). Gas Sensor Properties of (CuO/WO3)-CuWO4 Heterostructured Nanocomposite Materials. Materials, 18(12), 2896. https://doi.org/10.3390/ma18122896

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