Next Article in Journal
Synthesis of Ceramic Reinforcements in Metallic Matrices during Spark Plasma Sintering: Consideration of Reactant/Matrix Mutual Chemistry
Next Article in Special Issue
Smart Self-Sensing Piezoresistive Composite Materials for Structural Health Monitoring
Previous Article in Journal
Thermally Activated Al(OH)3 Part II—Effect of Different Thermal Treatments
Previous Article in Special Issue
Recent Progress on Electroactive Polymers: Synthesis, Properties and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Zn2+-Doped TiO2:WO3 Films Prepared by Electrospinning and Sintering: Microstructural Characterization and Electrical Signature to Moisture Sensing

by
Georgenes M. G. Silva
1,2,
Victor N. S. Leão
1,
Michel F. G. Pereira
1,
Pedro M. Faia
3 and
Evando S. Araújo
1,*
1
Research Group on Electrospinning and Nanotechnology Applications, Department of Materials Science, Federal University of San Francisco Valley, Juazeiro 48902-300, Brazil
2
Federal Institute of Education, Science and Technology of the Sertão Pernambucano, Petrolina 56314-520, Brazil
3
CEMMPRE—Electrical and Computer Engineering Department, FCTUC, University of Coimbra, Polo 2, Pinhal de Marrocos, 3030-290 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Ceramics 2021, 4(4), 576-591; https://doi.org/10.3390/ceramics4040041
Submission received: 31 July 2021 / Revised: 8 October 2021 / Accepted: 14 October 2021 / Published: 21 October 2021
(This article belongs to the Special Issue Innovative Processing Routes for Electroactive Materials)

Abstract

:
In this work, Zn2+-doped TiO2:WO3 nanostructured films, with different doping levels, were produced by electrospinning followed by sintering, and tested as potential materials for relative humidity (RH) detection. The materials microstructure was investigated by scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), Fourier-transform infrared spectroscopy (FTIR), Raman spectroscopy, and X-ray diffraction (XRD). The electrical characterization was performed by electrical impedance spectroscopy in the range of 400 HZ–40 MHZ, at 20 °C. The sensors’ sensitivity to moisture was evaluated from the impedance variations in response to changes in RH (10–100%). The analyses confirmed the interaction of water molecules with the oxides surface, and showed that zinc atoms were incorporated into the titanium vacancies in the crystal lattice. All the studied sensors showed a p- to n-type conduction transition taking place at around 40% RH. The nanocomposite with 2 wt% of dopant presented the best sensitivity to moisture, with an impedance variation of about 1 order of magnitude. The results are discussed in relation to the microstructure and fabrication route.

1. Introduction

Relative humidity (RH) is one of the most important environmental parameters for prevention and monitoring of a wide variety of processes in medical, food and industrial sectors [1].
Moisture sensors are the most commonly used instruments in monitoring this parameter. Several materials have been tested and used for this application, such as conductive polymers [2], carbon-based materials [3], metal oxides [4], and their composites [5], based on the notable variation in their electrical response as a function of RH.
Metal oxide-based moisture sensors have received attention due their excellent combined properties such as flexibility, chemical and structural stability, electroactivity, reproducibility and possibility of applications over a wide range of relative humidity [4,6]. More recently, titanium (IV) oxide ( TiO 2 )-based metal oxide heterogeneous nanostructures ( M O H N ) have been receiving increasing interest, and have been studied as potential materials for the development of high sensitivity and low cost moisture sensing devices [7].
In this configuration, the interaction at the atomic and surface level between TiO 2 and the other oxide constituents enables the improvement of the electrical and structural properties of the resulting material (when compared to the use of pure TiO 2 ) and, consequently, the moisture sensing [7]. The choice of selective oxides, dopant ions and fabrication route are other factors that directly influence the obtention of sensors with high surface area/volume ratio, and consequently higher sensitivity to RH [8]. In this sense, Zanetti et al. [9] showed that the addition of tungsten (VI) oxide ( WO 3 ) to TiO 2 by sintering improves the stability of the TiO 2 anatase phase and the sensitivity to relative humidity, due to the occurrence of surface heterojunction between the oxides. The results still revealed decreasing pore size and smaller variance in particle size distribution, which improves the chemisorption of water molecules on the semiconductors surface.
In the fabrication procedure of these sensing materials, in the form of volumetric pellets, the mechanical alloying method was introduced before the usual molding and sintering steps [10]. The results showed that the change in the preparation route was the main reason for obtaining heterogeneous nanostructures with smaller particle size, larger surface area, greater stability to phase transitions, changes in electrical behavior at low RH and improved sensitivity.
More recently, Araújo and Leão [11] indicated that it is possible to produce heterogeneous TiO 2 : WO 3 nanostructures with high moisture sensitivity, using an alternative preparation route that combines electrospinning and sintering. In these systems, the surface area/volume ratio was higher than that observed for the TiO 2 : WO 3 pelletized systems [10,12], while the amount of material by about three orders of magnitude [11] could lead to the miniaturization of sensor devices.
Based on previous studies on moisture sensors produced from TiO 2 and WO 3 heterogeneous nanostructures doped with diverse selective ions ( V 5 + , Nb 4 + , Zn 2 + , In 3 + , among others) [9,10,12,13], using various fabrication routes [10,13], this work investigates the electrical signature to moisture of Zn 2 + -doped TiO 2 : WO 3   M O H N films produced by electrospinning followed by sintering.
The selection of zinc as a dopant was based on previous reports [12,14,15,16,17], which evidenced that the use of zinc could increase the sensitivity of the composite sensors.
All the obtained sensors showed a p- to n-type conduction transition taking place at around 40 % RH, different from what was observed with pelletized samples of the same compositions previously tested, which exhibited n-type conduction along the full RH range [13].
The conduction mechanisms and electrical response of the samples to RH variations were directly related to the alternative fabrication process and the resulting microstructure of the final materials.

2. Materials and Methods

2.1. Materials

Methacrylic acid-methyl methacrylate copolymer (Evonik), ethanol PA ( 99.8 % , Neon Comercial, Suzano, Brazil), titanium (IV) oxide (anatase TiO 2 , < 25 nm particle size, 99.7 % trace metals basis, Sigma Aldrich, Burlington, VT, USA), tungsten (VI) oxide ( WO 3 , < 100 nm particle size, Sigma Aldrich) and zinc oxide ( ZnO , < 50 nm particle size, > 97 % , Sigma Aldrich) chemicals were used as raw materials.

2.2. Sample Preparation

In brief, the samples for the humidity sensitivity tests were prepared by electrospinning of the metal oxide-containing copolymer solution, followed by the sintering process of the electrospun composite fibers [9].
The polymer solutions for the electrospinning procedure were prepared from 1.6 g of copolymer diluted in 6.0 mL of ethanol.
The metal oxide heterogeneous nanosctructures ( M O H N samples) were defined as: T W ( Ti O 2 : W O 3 base mixture, with 48.9 : 51.1   wt % proportion of these oxides, respectively); T W Z n 2 and T W Z n 4 ( 98 : 2 and 96 : 4 weight ratio of the T W mixture and ZnO , respectively).
Then, 4   mL of polymer solution containing 0.50   g of each M O H N   sample were introduced into a syringe ( 10   mL capacity and 0.7 mm metallic capillary diameter) and injected under constant pressure, with a flow rate of 100   µ L   min 1 at the capillary tip (needle). An electric potential difference of 15   kV was applied between the needle and a grounded plane metal collector ( 10 × 10   cm 2 ), separated by a distance of   10   cm , to allow the electrospun polymer/ M O H N fibers production on the collector surface. The collected fibers were then deposited over a pair of gold ( A u ) interdigitated electrodes ( 1.1   cm   × 1.1   cm ), previously printed on an alumina substrate. All experiments were carried out at 20 °C.
Each sample on the substrates was then sintered in an oven at 500 °C for 2 h, in order to obtain the final M O H N films for electrical characterization. The choice of the just mentioned sintering conditions was based on previous studies [13,15], which showed favourable moisture behaviour under similar sintering conditions. Figure 1 illustrates the procedure for the preparation of M O H N and the characterization of the humidity sensing properties.

2.3. Microstructural Characterization

The materials microstructure was investigated by scanning electron microscopy (SEM) and energy dispersive spectrometry (EDS), by which elemental composition mappings were generated, using a Vega 3XM Tescan equipment (Tescan, Brno, Czech Republic), 10 to 20   kV accelerating voltage. Fourier Transform Infrared spectroscopy (FTIR) was carried out in transmission mode in the wavenumber range from 400 to 4000   cm 1 , using K B r pellets (Schimadzu Prestige 21 equipment, Shimadzu Corporation, Kyoto, Japan). Raman spectra were obtained using 532 nm laser wavelength, 20 25   mW (HORIBA Scientific, Osaka, Japan). X-ray diffraction (XRD) patterns in the range 15 70 ° with a scan rate of 0.02 ° s 1 , were generated using a Miniflex Rigaku equipment (Rigaku Corporation, Tokyo, Japan), Cu Kα radiation ( λ C u = 1.5406   Å ), 15   mA current and 40   kV voltage. The crystalline phases were identified using X-Pert HighScore software, while the crystallographic data were obtained from the International Center for Diffraction Data (ICDD).

2.4. Electrical Response

The obtained M O H N samples’ electrical response dependence with relative humidity, RH, was investigated by electrical impedance spectroscopy (impedance analyzer HP4194A, Hewlett-Packard, Santa Clara, CA, USA), in the frequency range from 400 Hz to 40 MHz. The samples were placed inside a test chamber with a volume of approximately 6.5 L, at a temperature of 20 °C, in which moisture concentration and temperature were automatically controlled, with an error smaller than ±1%. RH values ranging from 10 to 100%, were obtained by mixing dry air and fully humidified air in the desired proportions. The sensitivity of the samples to RH changes was evaluated using calculated impedance module values, |Z| (|Z|= (   Z ) 2 + (   Z ) 2 ), where Z and   Z are the resistance and the reactance of the measured impedance by impedance spectroscopy. All depicted electrical results, are the medium values calculated for each frequency of two sets of sweep measures taken in two different moments in time.

2.5. Statistical Analysis

The analyses of fiber diameter and crystallite and particle size of the sintered structures were given as the average of three independent measurements, for each M O H N configuration. Normality and equality of means tests were performed with significance level, α, of 0.05 and compared to the p value returned by statistical tests. Sample sizes (n) were obtained from SEM images using ImageJ software.

3. Results and Discussion

3.1. Microstructural Analysis

The SEM micrographs and EDS mappings of the T W Z n 2 and T W Z n 4   M O H N -loaded fibers are shown in Figure 2. The surface morphology analysis returned regular diameter composite fibers with no apparent structural defects, indicating that the electrospinning process parameters were adequate for preparation of these materials (first step of sensors production).
The mean diameters of T W , T W Z n 2 and T W Z n 4 fibers were estimated as 6.4 ± 1.7   μ m ( n   = 59 ) [11], 6.4 ± 1.5   μ m ( n = 69 ) and 6.7 ± 2.3   μ m ( n = 62 ), respectively. All M O H N -loaded fiber configurations presented diameters with normal distribution. In addition, the hypothesis tests on the population mean difference, with unknown standard deviation, revealed that there is no statistically significant difference in the mean diameter of the composite fibers ( p value > α ).
The EDS elemental composition and maps of the M O H N models-loaded fibers confirmed that the   Ti ,   W , Zn and O element populations are evenly distributed in the polymer matrix, as exemplified for the T W Z n 2 sample (Figure 2c). In addition, these combined analyses indicate that the moisture sensitivity of the nanocomposites is mainly influenced by the microstructure of the oxides and their interactions, since there is no significant variation in the volume fraction of nanoparticles in the produced polymeric matrices.
With electrospinning, particle aggregation is minimized and dispersion of the M O H N combinations over the interdigitated electrodes is promoted. As a consequence, higher surface contact areas, and in turn, with higher sensitivity materials to RH are expected [11]. In other words, electrospinning proves to be a simple and low-cost technique (when compared with other methods such as coating and sputtering [18]) that can be viably used for the deposition of metal oxides on substrates.
Thus, it is possible to produce sensing materials with a high surface area to volume ratio, which directly leads to better RH sensitivity. The application of M O H N as RH sensors involves the processes of chemisorption and physisorption of water molecules on the oxides surface. Additionally, the improved efficiency of these processes is also directly related with the larger porosity of sensing materials, as expected for the case of M O H N   produced by electrospinning [11].
The surface morphology of the materials used for the electrical response tests to moisture, resulting from the sintering process at 500 °C, is shown in Figure 3. After sintering, the obtained initial fibrous morphology is no longer present, however, based on the SEM images, the obtained nanostructures seem to present significant overall contact surface together with open porosity.
The EDS analyses of the sintered M O H N confirmed the continuity of the uniform distribution of   Ti : W metal matrix and of the zinc dopant, as shown for the T W Z n 2 sample (Figure 3c). The micrographs of T W Z n 2 and T W Z n 4 samples revealed a porous nanostructure, with spherical particles in the order of 100–250 nm, and free of polymeric phase. These features are known to favor the application of these structures for RH sensing, as discussed earlier. The mean sample diameters were calculated to be 145.7   ±   47.0   nm   ( n   = 135 ) [11], 165.1   ±   64.1   nm   ( n   = 115 ) and   179.7   ±   73.5   nm   ( n   = 122 ) , for   T W , T W Z n 2 and T W Z n 4 systems, respectively.
The hypothesis tests’ analyses for the difference in mean particle diameter showed that there were no significant differences in particle size among the all produced sintered M O H N ( p value > α , for all mean comparisons). This proves that the best results for the sensitivity of these systems to RH are predominantly related to the atomic and surface interaction levels between the constituent oxides of the nanocomposites. It is important to, once more, point out that all morphological characteristics discussed here (such as nanometric size, spherical shape, porosity and regular distribution) are fundamental for RH sensing, since they favor the adsorption processes of water vapor from the environment, on the oxides surface and the porous structure.
The FTIR spectra of the polymer matrix and sintered samples are shown in Figure 4a. For all three M O H N samples, it is possible to observe a broad transmission band in the 470 1000   cm 1 wavenumber range, due to the superposition of bands belonging to the different metal oxides present in the structures. In addition, the characteristic bands of the polymer matrix in the range of 1000 to 3600   cm 1 (including the broad band around 3380 cm 1 due to the adsorbed water molecules by the polymer) [19], are not observed in the M O H N spectra, proving that the sensor materials are free from the organic phase, as suggested by the results obtained by SEM.
The X-ray diffractograms of the   T W , T W Z n 2 , and   T W Z n 4   M O H N are presented in Figure 4b. The characteristic crystallographic planes of the anatase TiO 2   with tetragonal structure ( a   = b = 3.7821   Å and c = 9.5022   Å lattice constants [20]) and monoclinic WO 3   ( a   = 7.306   Å , b = 7.540   Å and c = 7.692   Å [21]) were confirmed in the experimental data of the three composite systems, from 01-071-1166 and 01-083-0951 ICDD cards, respectively. The results show that the electrospinning and sintering processes at 500 °C causes no phase change or solid solution formation from these oxides in the tested M O H N [1,22].
In all three samples, the experimental values of WO 3   unit cell parameters ( a = 7.31   Å , b = 7.54 Å,   c = 7.69   Å , α = γ = 90 ° and β = 90.88 ° ) showed no significant changes, related to pristine oxide. The lattice parameters for TiO 2   in the T W sample were calculated to be a = b = 3.78   Å and   c = 9.51   Å , α = β = γ = 90 ° [11], in accordance with the data of the starting material. In addition, it is possible to verify in Table 1 the dependence of   TiO 2   parameters on Zn concentrations in the T W Z n 2   and T W Z n 4 samples. A decrease in the values of a and b parameters, and an increase in the value of   c , from the undoped material to the doped material with 2   wt % of zinc, can be noticed, due to a variation in the volume of the TiO 2 unit cell. The T W Z n 4 sample, on the other hand, does not present significant changes in the values of these parameters (no change in the cell volume), compared to the T W Z n 2 sample, indicating that part of the zinc concentration remains on the surface material after the interaction saturation between zinc and titanium ions at the atomic level.
In fact, there is the occurrence of zinc ion doping in the atomic structure of titanium: the ionic radius of Zn 2 + ions (0.74 Å [23]) is close to that of Ti 4 + ions, promoting the occupation of vacancies of the Ti atom sites by Zn elements, after the sintering process.
The crystallite average size ( D ) of the anatase TiO 2   (Table 1) and monoclinic WO 3   in the investigated   T W , T W Z n 2   and   T W Z n 4   nanocomposites was determined by the modified Scherrer equation (MSE),   D = λ / 3 β cos ( θ ) [24], where 1/3 is the correction value of the shape factor K in Scherrer Equation [25]; λ is the X-ray radiation wavelength; β and θ are the full width half maxima and the half of the diffraction angle (2θ) at the crystal planes detected in the XRD spectrum, respectively.
The average size ( D ) of the WO 3 crystallites was estimated to be 15.01 ( ± 0.75 ) nm, 15.32 ( ± 1.22 ) nm and 15.95 ( ± 1.49 ) nm for T W , T W Z n 2 , and T W Z n 4 samples, respectively. For   TiO 2 , D values were calculated to be 19.12 ( ± 1.12 ) nm, 24.27 ( ± 2.03 ) nm, and 23.75 ( ± 2.33 ) nm, with increasing zinc concentration.
The equality of means test (with sample size n = 3 and level of significance α = 0.05) showed that the WO 3 crystallite average sizes present no significant differences in the three tested systems. On the other hand, TiO 2 average size values calculated for the T W Z n 2 and T W Z n 4 compositions are significantly higher than that observed for the T W model. The D values in T W Z n 2 and T W Z n 4 samples did not show significant difference. These results corroborate the previously reported effects of dopant and indicate that ZnO at 4   wt % promotes saturation of the Zn / Ti atomic interactions, which increases the presence of surface interactions between the oxides. This behavior is confirmed by the emergence of ZnO peaks with a hexagonal crystal structure (from the ( 100 )   plane at   2 θ = 32.2 ° , and according to the supplier) in the diffractograms of doped samples.
The Raman spectra of the sintered M O H N are presented in Figure 5. The spectrum of pristine anatase TiO 2   sintered at 500   ° C was also analyzed for comparison with the M O H N spectra. The characteristic molecular vibration bands of the anatase phase were observed at   141 ,   195 ,   396 ,   515   and   637   cm 1   wavenumbers [26].
These bands were also found in the T W and T W Z n samples with 2 and 4   wt % of Zinc, but shifted to lower wavenumber regions (about 8 11   cm 1 shift) of the spectrum (red shift). In addition, the spectral analysis of undoped and Zn-doped samples proved the continuity of the WO 3   monoclinic structure after sintering at 500   ° C , with characteristic vibration bands at 262, 316, 704, and 798 cm 1   wavenumbers, as described in the literature [27,28].
The red shift phenomenon of   TiO 2   vibrational bands is attributed to distortions and confinement of phonons in its crystalline lattice, promoted by the introduction of impurities in the atomic structure of the material [29]. In fact, the combined analysis of the XRD and Raman results confirmed that Zn 2 + ions occupied the available sites of Ti atoms in the anatase crystal lattice.

3.2. Electrical Response

The electrical response of T W Z n   M O H N to RH variations was investigated by electrical impedance spectroscopy. The M O H N electrical response to moisture depends on the moisture concentration and is the consequence of chemisorption and physisorption of water molecules along the surface, and on the porous structure of these nanostructures.
For low moisture concentrations, only chemisorption takes place; this occurs when the first layer of water molecules that will cover the oxides surface and porous structure gets formed. This layer, which is strongly bonded to the oxides, covers entirely the contact surface (including the open porous surface) of the materials at around 30 to 40% RH concentration. At this point, oxygen from the humidified air at room temperature is typically adsorbed in the form of   O ,   O 2 , O 2 and O 2 2 ions. As RH concentration keeps increasing, a first physiosorbed layer forms on top of the chemisorbed one: with continuous increment of moisture concentration, subsequent physiosorbed layers form one after the other. However, all physiosorbed layers are unstable and can be easily removed by increasing the temperature or by decreasing RH [7,30]. At this stage, H 3 O + ions with a high quantity of H + protons, get formed and start participating as charge carriers, a process known as Grotthuss mechanism [31]. Once the hydronium chemical bonds are instable, H + protons hopping to the nearest water molecule takes place; consequently, H 2 O molecules get ionized and form H 3 O + ions, which perpetuate the charge transportation, unless, the material is heated or the RH concentration is reduced.
As the amount of water molecules begins to get saturated over the materials surface, they start to condense in the existing pore structure, within the bulk of the sensor material. When the pores in the mesoporous region (pores around 10–100 nm) get totally filled with water, new carrier paths are established for ions migration.
The measured impedance spectra highlight the different phenomena of electrical conduction and polarization that occur in these materials in the presence of water molecules, making the use of electrical impedance spectroscopy a very powerful technique in the evaluation of the materials future use in sensing devices.
In Figure 6 are plotted the Nyquist plots of T W , T W Z n 2 , and T W Z n 4 , in the full range of relative humidity (10–100% RH).
With the increase in RH concentration, a maximum value in impedance is observed for all tested samples. From Figure 6 this characteristic is perceived in the Nyquist plots (imaginary versus real parts of the impedance, Im   Z   ×   Re   Z ) for all samples. This type of behavior is a typical of a p-to n-type conduction transition in materials [32,33,34,35].
Although this phenomenon has already been reported in gas sensing metal oxide based devices, the mechanism that governs this transition has not been widely described in the literature. For RH sensing with high surface area M O H N (produced by the suggested preparation route), at low RH regime, the oxygen ions from ambient air and moisture molecules act as different agents when adsorbed on the material surface, with direct influence on the conduction properties. Herein, at low RH concentrations, oxygen from the humidified air at room temperature (in the form of O ,   O 2 , O 2 and O 2 2 ions) are preferentially adsorbed on the surface of M O H N films, by capturing electrons from the conduction band [35].
This process results in the formation of an electron depletion layer [32,35], which causes the decrease in the density of free electrons and their entrapment by holes in the valence band. Consequently, the impedance of the material increases due to the removal of charge carriers from the conduction band. With the beginning of exposure to moisture, the surface of the material also begins to adsorb oxygen from the water molecules, but at a much lower rate than that of ambient oxygen, which does not favor their interaction. At this point, it is suggested that there are no significant surface active sites available for H2O gas adsorption. Thus, the increase in system resistance is attributed to the occupying of negatively charged surface of oxygen from the air instead of moisture ions. With increasing RH, moisture adsorption tends to occur primarily by replacing the initially adsorbed oxygen ions.
When the   M O H N is exposed to increasing RH concentrations in a low range (up to 40% of RH), the water vapor ions will progressively substitute the oxygen from the air and promote the distribution of surface adsorbed active sites. At this stage, the distance between ions from both oxygen sources will become smaller, with a consequent increase in the chemical interaction between them on the materials surface.
At this stage, water adsorption can be favoured if the sensing materials possess high surface area to volume ratio. It is know that adsorption of water vapour on the materials surface is promoted by a higher contact surface area, due to its porous nature, which affects the electrical conduction; this change encompasses the resistance, capacitance or electrolytic conduction depending on the sensor type [36].
Adsorption of ions from moisture becomes dominant above 30–40% RH in the tested M O H N , allowing the electrons to return to the conduction band and thus defining the transition from p- to n-type conduction in these materials.
In fact, when M O H N is known to adsorb a large amount of oxygen from the environment at a low RH range, a significant water vapor concentration is required so that the expected usual n-type conduction process can be established (as noted for pristine anatase titanium dioxide). In sequence, when RH concentration keeps raising above   40 % , the Grotthuss chain reaction dominates and ion hopping becomes the main conduction process.
As a consequence, electrons continue to be released to the conduction band, and consequently the sensors conductance increases (n-type behavior). In general, the equilibrium between these reactions defines the dominant behavior of sensing mechanism at the level of electronic bands.
The impedance of these systems continues to decrease progressively with the continuous increase of the moisture concentration, with the formation of the first and following physiosorbed layers, mainly due to transport and electrical polarization mechanisms, especially the hopping conduction of protons described earlier.
As already seen in a previous work by the authors [11], the production of M O H N sensing films by electrospinning may allow the obtention of nanostructured systems with high surface area to volume ratio. On close inspection of the micrographs, that seems to be the case, by a regular distribution of small spherical grains with low agglomeration, contributing to a higher contact surface area. Nevertheless, the larger presence of agglomerates in samples with higher zinc content is most likely the key factor that determines their lower sensitivity, when compared to samples with 2   wt %   Zn content. This further confirms the validity of electrospinning as an efficient and low-cost fabrication process compared to other approaches, such as vapor deposition, or sputtering techniques, among others.
By looking again to the Nyquist plots for all M O H N nanostructures, the presence of compressed semicircles indicates that conduction in these materials is dominated by relaxation processes at the surface of grains and through grain boundaries [37]. Diffusion processes are present, at least at the interface between the measuring electrodes and the material and inside the pores that get filled with water vapour. However, their diffusion contributions are clearly small with respect to the overall measured impedance of the sensors, and in comparison to the remaining conduction contributions present [12,38]. Despite the SEM images showing that the samples are highly porous, there is low evidence for the existence of a significant amount of pores in the mesoporous region (pore size ranging from units to tens of nm). This behavior is directly associated with the arrangement of nanoparticles as films on the electrodes instead of the use of volumetric pellets, in which the impedance variation has a significant influence of micro/nanopores presence [12,38]. As an example, TiO2/WO3 pellets (with 100/36 Titanium/tungsten atoms proportion and sintered at 600 °C) showed a typical n-type conduction behavior in the range of 0 to 100% RH, with strong influence of diffusion processes and, consequently, an impedance variation around 2 order of magnitude [38].
The TiO2:WO3 volumetric systems (1:1 in mol) have been also doped with 7 wt% of ZnO [12]. The results showed that the doped samples electrically respond to RH (with a n-type conduction) but with a smaller impedance variation (close to 1 order of magnitude) [12]. The transition from p- to n-type conduction has also been reported in the TiO2:WO3 pelletized system (1:1 in mol), induced by vanadium ions doping [27]. The dopant introduction improved the stability of the anatase phase and considerably increased the sensitivity of the final material (relative to the initial binary mixture).
Electrospun TiO2:WO3 models [11] were also tested as moisture sensors, with the same binary mixture proportion as that used in other previous works [12,27] and sintered at 500 °C. The p- to n-type transition was observed and related to the change in the preparation route of the sensors by electrospinning. These systems showed excellent sensitivity to RH variations (around 1 order of magnitude) compared to pelletized samples, but using a much smaller amount of sensor material.
This configuration promotes the use of a small amount of sensor material (2.24 ± 0.21 mg (n = 3), 2.18 ± 0.33 mg (n = 3) and 2.20 ± 0.29 mg (n = 3), from 0 to 4 wt% ZnO, respectively). The M O H N mass is invariant from one sample to another, with uniform distribution of the material over the electrode area. This is a result of the proper control of the electrospinning process parameters, the use of electrospinning as part of the material preparation and the sintering process. Thus, the errors related to the impedance measurements of the samples allocated in the humidity chamber (with ambient RH stabilized before each measurement), given a M O H N composition, are shown negligible and the best sensitivity results can be discussed based on the material properties.
On the other hand, in this work, the sensitivity of the produced films was increased in relation to that detected in the volumetric and electrospun TiO2:WO3 models [11,38]. The obtained impedance variations greater than 1 order of magnitude are an indication that the transport and electrical polarization mechanisms, especially the hopping conduction of protons, were considerably improved by: (i) the Zn ions doping and; (ii) high surface area to volume ratio of M O H N promoted by the preparing alternative method of the sensor materials. Figure 7 shows the variations in the impedance modulus obtained for the T W and T W Z n systems as a function of RH, in the 10–100% RH range.
Frequencies above 1 kHz were also analyzed but the overall observed impedance variations were smaller than the ones here reported and plotted. The T W ,     T W Z n 2 ,   and Z n 4 samples display an overall impedance variation of 0.95, 1.52 and 1.29 order of magnitude, respectively, at the frequency of 750 Hz, where all tested sensors display the highest variations to RH changes.
These results allow to state that the use of zinc as dopant is an effective approach to improve the sensitivity of electrospun TiO2/WO3   M O H N . In fact, the M O H N systems containing zinc exhibit higher overall impedance to RH variations (higher overall impedance modulus variation) than those obtained with the T W systems (produced in the form of pellets or prepared by electrospinning, without addition of dopants) [11]. Besides, the depicted results also allow to clearly observe the mentioned p- to n-type transition, and to notice that the M O H N system with 2 wt% zinc content presents the highest sensitivity to moisture. Additionally, and for all sensors, it is possible to see that the frequency at which the higher overall impedance modulus variation is observable, is of 750 Hz.
In Figure 8, a close comparison of the impedance modulus variation between the three   M O H N , done for the two test frequencies, allows to: (i) confirm that the best testing frequency is 750 Hz (the overall impedance modulus variation found at 1060 Hz is clearly smaller than the observed at 750 Hz); (ii) confirm that the system containing 2 wt% of zinc does not only exhibit always a higher overall impedance variation to moisture changes, but it also exhibits an almost linear variation with it.
The electrical response stability to moisture of T W and T W Z n samples was tested at three RH working values (at 50, 70 and 90% RH—after the p-n transition) at 750 Hz, in terms of the relative variation of impedance, over the seven-day measurement interval under constant RH regime. Table 2 reports the stability of the electrical response based on the working values Z i   and   Z f of | Z | measured after the p-n transition, and at 100 % of relative humidity, respectively. The results showed that all doped samples have small variation in the measured impedance modulus values ( 2 4 % ) with time, considering a fixed RH (similar to the variation observed for the undoped one), with high potential usage in moisture sensing devices. The actual obtained materials, and in particular the composition doped with 2   wt % of zinc exhibits, specifically in the high humidity range, does not only show better sensitivity, but also a linear variation of the impedance modulus, characteristics not found in other materials reported in literature [39,40,41,42].
In resume, the electrical moisture sensitivity results obtained with the use of a dopant (see Figure 8 and Table 2), confirm that electrospinning is an effective and convenient processing route to prepare RH sensing devices.

4. Conclusions

Zinc-doped TiO 2 : WO 3 samples were successfully prepared by electrospinning followed by sintering. The obtained electrical response to RH changes proves that electrospinning is a powerful and inexpensive technique to obtain nanostructures of mixed metal oxides for potential use in moisture sensing devices.
Besides, the present results also showed that doping is a highly valuable path that can be used to improve M O H N moisture electrical response. The microstructural characterization supports the discussion of the observed electrical response to moisture variations, obtained for the zinc-doped samples.
The highest impedance variation was obtained in sample doped with 2 wt% of zinc, which showed an overall impedance variation of more than 1 order of magnitude, with a linear electrical response. The research with these potential materials as gas sensors will be continued with the evaluation of other important parameters such as hysteresis and response and recovery time.

Author Contributions

Sample preparation and characterization, writing, editing and review, G.M.G.S., V.N.S.L., M.F.G.P., P.M.F., and E.S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Project 202451/2015-1) and Fundação de Amparo à Pesquisa do Estado da Bahia (FAPESB, Project 1252/2018) brazilian research agencies. This research was also sponsored by national funds through FCT—Fundação para a Ciência e a Tecnologia, under the Project UIDB/00285/2020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Farahani, H.; Wagiran, R.; Hamidon, M.N. Humidity sensors principle, mechanism, and fabrication technologies: A comprehensive review. Sensors 2014, 14, 7881–7939. [Google Scholar] [CrossRef] [Green Version]
  2. Morais, R.M.; Klem, M.D.S.; Nogueira, G.L.; Gomes, T.C.; Alves, N. Low Cost Humidity Sensor Based on PANI/PEDOT:PSS Printed on Paper. IEEE Sens. J. 2018, 18, 2647–2651. [Google Scholar] [CrossRef] [Green Version]
  3. Tulliani, J.-M.; Inserra, B.; Ziegler, D. Carbon-Based Materials for Humidity Sensing: A Short Review. Micromachines 2019, 10, 232. [Google Scholar] [CrossRef] [Green Version]
  4. McGhee, J.R.; Sagu, J.S.; Southee, D.J.; Evans, P.S.A.; Wijayantha, K.G.U. Printed, Fully Metal Oxide, Capacitive Humidity Sensors Using Conductive Indium Tin Oxide Inks. ACS Appl. Electron. Mater. 2020, 2, 3593–3600. [Google Scholar] [CrossRef]
  5. Zhu, Z.; Lin, W.-D.; Lin, Z.-Y.; Chuang, M.-H.; Wu, R.-J.; Chavali, M. Conductive Polymer (Graphene/PPy)–BiPO4 Composite Applications in Humidity Sensors. Polymers 2021, 13, 2013. [Google Scholar] [CrossRef]
  6. Ashok, C.; Venkateswara, R.K.; Shilpa, C.C. Comparison of Metal Oxide Nanomaterials: Humidity Sensor Applications. In Materials, Energy and Environment Engineering; Mohan, B.R., Srinikethan, G., Meikap, B., Eds.; Springer: Singapore, 2017; pp. 267–275. [Google Scholar] [CrossRef]
  7. Cappelli, I.; Fort, A.; Lo Grasso, A.; Panzardi, E.; Mugnaini, M.; Vignoli, V. RH Sensing by Means of TiO2 Nanoparticles: A Comparison among Different Sensing Techniques Based on Modeling and Chemical/Physical Interpretation. Chemosensors 2020, 8, 89. [Google Scholar] [CrossRef]
  8. Silva, G.M.G.; Leão, V.N.S.; Pereira, M.F.G.; Faia, P.M.; Araújo, E.S. Effect of Nb5+ and In3+ Ions on Moisture Sensitivity of Electrospun Titanium/Tungsten Oxide Nanostructures: Microstructural Characterization and Electrical Response. Processes 2021, 9, 1336. [Google Scholar] [CrossRef]
  9. Zanetti, S.M.; Rocha, K.O.; Rodrigues, J.A.J.; Longo, E. Soft-chemical synthesis, characterization and humidity sensing behavior of WO3/TiO2 nanopowders. Sens. Actuators B Chem. 2014, 190, 40–47. [Google Scholar] [CrossRef]
  10. Faia, P.M.; Libardi, J. Response to humidity of TiO2:WO3 sensors doped with V2O5: Influence of fabrication route. Sens. Actuators B Chem. 2016, 236, 682–700. [Google Scholar] [CrossRef]
  11. Araújo, E.S.; Leão, V.N.S. TiO2:WO3 heterogeneous structures prepared by electrospinning and sintering steps: Characterization and analysis of the impedance variation to humidity. Adv. Ceram. 2019, 8, 238–246. [Google Scholar] [CrossRef] [Green Version]
  12. Faia, P.M.; Jesus, E.L.; Louro, C.S. TiO2:WO3 composite humidity sensors doped with ZnO and CuO investigated by impedance spectroscopy. Sens. Actuators B Chem. 2014, 203, 340–348. [Google Scholar] [CrossRef]
  13. Araújo, E.S.; Libardi, J.; Faia, P.M.; Oliveira, H.P. Characterization and Electrical Response to Humidity of Sintered Polymeric Electrospun Fibers of Vanadium Oxide-(TiO2/WO3). J. Electron. Mater. 2018, 47, 2018. [Google Scholar] [CrossRef]
  14. Faia, P.M.; Furtado, C.S. Effect of composition on electrical response to humidity of TiO2:ZnO sensors investigated by impedance spectroscopy. Sens. Actuators B Chem. 2013, 181, 720–729. [Google Scholar] [CrossRef]
  15. Thiwawong, T.; Onlaor, K.; Chaithanatkun, N.; Tunhoo, B. Preparation of copper doped zinc oxide nanoparticles by precipitation process for humidity sensing device. AIP Conf. Proc. 2010, 020022. [Google Scholar] [CrossRef]
  16. Yawale, S.P.; Yawale, S.S.; Lamdhade, G.T. Tin oxide and zinc oxide based doped humidity sensors. Sens. Actuators A Phys. 2007, 135, 388–393. [Google Scholar] [CrossRef]
  17. Peng, X.; Chu, J.; Yang, B.; Peter, F.X. Mn-doped zinc oxide nanopowders for humidity sensors. Sens. Actuators B Chem. 2012, 174, 258–262. [Google Scholar] [CrossRef]
  18. Usha, K.S.; Sivakumar, R.; Sanjeeviraja, C. Optical constants and dispersion energy parameters of NiO thin films prepared by radio frequency magnetron sputtering technique. J. Appl. Phys. 2013, 114, 123501. [Google Scholar] [CrossRef]
  19. Gulshan, M.; Sai, M.L.S.; Hemalatha, T.; Sri, U.J.; Ramarao, N. Formulation and development of microspheres for the treatment of familial adenomatous polyposis. Int. J. Appl. Pharm. 2017, 9, 66–72. [Google Scholar] [CrossRef] [Green Version]
  20. Treacy, J.P.W.; Hussain, H.; Torrelles, X.; Grinter, D.C.; Cabailh, G.; Bikondoa, O.; Nicklin, C.; Selcuk, S.; Selloni, A.; Lindsay, R.; et al. Geometric structure of anatase TiO2(101). Phys. Rev. B 2017, 95, 075416. [Google Scholar] [CrossRef] [Green Version]
  21. Leng, X.; Pereiro, J.; Strle, J.; Bollinger, A.T.; Božović, I. Epitaxial growth of high quality WO3 thin films. APL Mater. 2015, 3, 096102. [Google Scholar] [CrossRef] [Green Version]
  22. Schumm, M.; Koerdel, M.; Müller, S.; Ronning, C.; Dynowska, E.; Gołacki, Z.; Szuszkiewicz, W.; Geurts, J. Secondary phase segregation in heavily transition metal implanted ZnO. J. Appl. Phys. 2009, 105, 083525. [Google Scholar] [CrossRef]
  23. Yu, Q.; Fu, W.; Yu, C.; Yang, H.; Wei, R.; Sui, Y.; Liu, S.; Liu, Z.; Li, M.; Wang, G.; et al. Structural, electrical and optical properties of yttrium-doped ZnO thin films prepared by sol–gel method. J. Phys. D Appl. Phys. 2007, 40, 5592. [Google Scholar] [CrossRef]
  24. Lima, F.M.; Martins, F.M.; Júnior, P.H.F.M.; Almeida, A.F.L.; Freire, F.N.A. Nanostructured titanium dioxide average size from alternative analysis of Scherrer’s Equation. Matéria 2018, 23, 1–9. [Google Scholar] [CrossRef]
  25. Miranda, M.A.R.; Sasaki, J.M. The limit of application of the Scherrer equation. Acta Crystallogr. Sect. A Found. Adv. 2018, 74, 54–65. [Google Scholar] [CrossRef]
  26. Sahoo, S.; Arora, A.K.; Sridharan, V. Raman line shapes of optical phonons of different symmetries in anatase TiO2 nanocrystals. J. Phys. Chem. C 2009, 113, 16927–16933. [Google Scholar] [CrossRef]
  27. Faia, P.M.; Libardi, J.; Louro, C.S. Effect of V2O5 doping on P- to N-conduction type transition of TiO2:WO3 composite humidity sensors. Sens. Actuators B Chem. 2016, 222, 952–964. [Google Scholar] [CrossRef]
  28. Su, C.Y.; Lin, H.C.; Lin, C.K. Fabrication and optical properties of Ti-doped W18O49 nanorods using a modified plasma-arc gas-condensation technique. J. Vac. Sci. Technol. B 2009, 27, 2170–2174. [Google Scholar] [CrossRef]
  29. Yang, C.C.; Li, S. Size-dependent Raman red shifts of semiconductor nanocrystals. J. Phys. Chem. B 2008, 112, 14193–14197. [Google Scholar] [CrossRef]
  30. Raji, P.; Binitha, H.S.; Kumar, K.B. Synthesis and humidity sensing properties of Sn-doped nano-TiO2. J. Nanotechnol. 2011, 2011, 569036. [Google Scholar] [CrossRef] [Green Version]
  31. Agmon, N. The Grotthuss mechanism. Chem. Phys. Lett. 1995, 244, 456–462. [Google Scholar] [CrossRef]
  32. Gurlo, A.; Bârsan, N.; Oprea, A.; Sahm, M.; Sahm, T.; Weimar, U. An n-to p-type conductivity transition induced by oxygen adsorption on α-Fe2O3. Appl. Phys. Lett. 2004, 85, 2280. [Google Scholar] [CrossRef]
  33. Sumets, M.; Dannangoda, G.C.; Kostyuchenko, A.; Ievlev, V.; Dybov, V.; Martirosyan, K.S. Temperature transition of p- to n-type conduction in the LiNbO3/Nb2O5 polycrystalline films. Mater. Chem. Phys. 2017, 191, 35–44. [Google Scholar] [CrossRef]
  34. Bohra, A.K.; Bhatt, R.; Singh, A.; Bhattacharya, S.; Basu, R.; Meshram, K.N.; Sarkarl, S.K.; Bhatt, P.; Patro, P.K.; Aswal, D.K.; et al. Transition from n-to p-type conduction concomitant with enhancement of figure-of-merit in Pb doped bismuth telluride: Material to device development. Mater. Des. 2018, 159, 127–137. [Google Scholar] [CrossRef]
  35. Gurlo, A.; Sahm, M.; Oprea, A.; Bârsan, N.; Weimar, U. A p-to n-transition on α-Fe2O3-based thick film sensors studied by conductance and work function change measurements. Sens. Actuators B Chem. 2004, 102, 291–298. [Google Scholar] [CrossRef]
  36. Blank, T.A.; Eksperiandova, L.P.; Belikov, K.N. Recent trends of ceramic humidity sensors development: A review. Sens. Actuators B Chem. 2016, 228, 416–442. [Google Scholar] [CrossRef]
  37. Bârsan, N.; Grigorovici, R.; Ionescu, R.; Vancu, A. Mechanism of gas detection in polycrystalline thick film SnO2 sensors. Thin Solid Films 1989, 171, 53–63. [Google Scholar] [CrossRef]
  38. Faia, P.M.; Ferreira, A.J.; Furtado, C.S. Establishing and interpreting an electrical circuit representing a TiO2–WO3 series of humidity thick film sensors. Sens. Actuators B 2009, 140, 128–133. [Google Scholar] [CrossRef]
  39. Ghalamboran, M.; Saedi, Y. TiO2-TiO2 composite resistive humidity sensor:ethanol crosssensitivity. IOP Conf. Ser. Mater. Sci. Eng. 2016, 108, 012039. [Google Scholar] [CrossRef] [Green Version]
  40. Shakya, V.; Pandey, N.K.; Misra, S.K.; Roy, A. Electrical and optical properties of ZnO–WO3 nanocomposite and its application as a solid-state humidity sensor. Bull. Mater. Sci. 2017, 40, 253–262. [Google Scholar] [CrossRef]
  41. Li, P.; Zheng, X.; Zhang, Y.; Yuan, M.; Jiang, B.; Deng, S. Humidity sensor based on electrospun(Na0.5Bi0.5)0.94TiO3– Ba0.06TiO3 nanofibers. Ceram. Int. 2015, 41, 14251–14257. [Google Scholar] [CrossRef]
  42. Sentürk, E.; Duman, S.; Bağcı, S.; Soykan, H.S.; Aslanoğlu, Z. Humidity sensing properties of steatite ceramic containing B2O3. Sens. Actuators A Phys. 2016, 240, 80–84. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the sensor materials production and electrical response testing to RH. (a) Metal oxides into polymer solution; (b) electrospinning process; (c) fibers on electrodes; (d) sintered metal oxides; (e) humidity chamber used for electrical characterization; (f) impedance spectroscopy.
Figure 1. Schematic illustration of the sensor materials production and electrical response testing to RH. (a) Metal oxides into polymer solution; (b) electrospinning process; (c) fibers on electrodes; (d) sintered metal oxides; (e) humidity chamber used for electrical characterization; (f) impedance spectroscopy.
Ceramics 04 00041 g001
Figure 2. Scanning electron microscopy (SEM) images of T W Z n   M O H N -loaded electrospun fibers with (a)   2   wt % and (b)   4   wt % of zinc oxide (   ZnO ); (c) representative SEM micrograph of the fiber surface and (d) energy dispersive spectroscopy (EDS) elemental composition and mapping of     T W Z n 2 samples.
Figure 2. Scanning electron microscopy (SEM) images of T W Z n   M O H N -loaded electrospun fibers with (a)   2   wt % and (b)   4   wt % of zinc oxide (   ZnO ); (c) representative SEM micrograph of the fiber surface and (d) energy dispersive spectroscopy (EDS) elemental composition and mapping of     T W Z n 2 samples.
Ceramics 04 00041 g002
Figure 3. SEM micrographs of (a) T W Z n 2 and (b) T W Z n 4   M O H N after the sintering process; and (c) EDS elemental composition and mapping of sintered T W Z n samples, with indication of titanium ( Ti ), tungsten ( W ), zinc ( Zn ), and oxygen ( O ) elements.
Figure 3. SEM micrographs of (a) T W Z n 2 and (b) T W Z n 4   M O H N after the sintering process; and (c) EDS elemental composition and mapping of sintered T W Z n samples, with indication of titanium ( Ti ), tungsten ( W ), zinc ( Zn ), and oxygen ( O ) elements.
Ceramics 04 00041 g003
Figure 4. (a) FTIR spectra and (b) X-ray diffractograms of the   T W , T W Z n 2 , and   T W Z n 4   M O H N .
Figure 4. (a) FTIR spectra and (b) X-ray diffractograms of the   T W , T W Z n 2 , and   T W Z n 4   M O H N .
Ceramics 04 00041 g004
Figure 5. Raman spectra of anatase TiO 2 , T W   and   T W Z n sintered   M O H N (varying dopant concentrations).
Figure 5. Raman spectra of anatase TiO 2 , T W   and   T W Z n sintered   M O H N (varying dopant concentrations).
Ceramics 04 00041 g005
Figure 6. Nyquist Plot of the M O H N obtained samples as a function of RH, in the frequency range from 400 Hz to 40 MHz, for the (a) TW, (b) T W Z n 2 , and (c) T W Z n 4 .
Figure 6. Nyquist Plot of the M O H N obtained samples as a function of RH, in the frequency range from 400 Hz to 40 MHz, for the (a) TW, (b) T W Z n 2 , and (c) T W Z n 4 .
Ceramics 04 00041 g006
Figure 7. (a) Impedance modulus plots of the M O H N obtained samples as a function of RH, at two different operating frequencies, for the (a) TW, (b) T W Z n 2 , and (c) T W Z n 4 .
Figure 7. (a) Impedance modulus plots of the M O H N obtained samples as a function of RH, at two different operating frequencies, for the (a) TW, (b) T W Z n 2 , and (c) T W Z n 4 .
Ceramics 04 00041 g007aCeramics 04 00041 g007b
Figure 8. (a) Impedance modulus plots comparison of the M O H N obtained samples as a function of RH, at the two selected operating frequencies: (a) 750 Hz; (b) 1060 Hz.
Figure 8. (a) Impedance modulus plots comparison of the M O H N obtained samples as a function of RH, at the two selected operating frequencies: (a) 750 Hz; (b) 1060 Hz.
Ceramics 04 00041 g008
Table 1. Lattice parameters and crystallite size of TiO 2   in   T W , T W Z n 2   and   T W Z n 4   M O H N .
Table 1. Lattice parameters and crystallite size of TiO 2   in   T W , T W Z n 2   and   T W Z n 4   M O H N .
Lattice Parameters ( Å )Average Crystallite Size, D (nm)
M O H N a = b C
T W 3.789.5119.12 ( ± 1.12 )
T W Z n 2 3.699.7324.27 ( ± 2.03 )
T W Z n 4 3.709.7023.75 ( ± 2.33 )
Table 2. Electrical response stability to moisture (at 50, 70 and 90% RH) for the T W and T W Z n   M O H N produced with   2 and 4   wt % of ZnO , respectively. Similarly to the way exactitude is estimated in percentage, the sensing materials electrical response stability is estimated by the formulation indicated in Table 2.
Table 2. Electrical response stability to moisture (at 50, 70 and 90% RH) for the T W and T W Z n   M O H N produced with   2 and 4   wt % of ZnO , respectively. Similarly to the way exactitude is estimated in percentage, the sensing materials electrical response stability is estimated by the formulation indicated in Table 2.
TW ZnO   ( wt % )
24
RH (%)Electrical response stability ( 100 Z i Z f Z i ) (%)
5096.3296.0995.83
7096.4396.1296.29
9097.4097.2296.31
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Silva, G.M.G.; Leão, V.N.S.; Pereira, M.F.G.; Faia, P.M.; Araújo, E.S. Zn2+-Doped TiO2:WO3 Films Prepared by Electrospinning and Sintering: Microstructural Characterization and Electrical Signature to Moisture Sensing. Ceramics 2021, 4, 576-591. https://doi.org/10.3390/ceramics4040041

AMA Style

Silva GMG, Leão VNS, Pereira MFG, Faia PM, Araújo ES. Zn2+-Doped TiO2:WO3 Films Prepared by Electrospinning and Sintering: Microstructural Characterization and Electrical Signature to Moisture Sensing. Ceramics. 2021; 4(4):576-591. https://doi.org/10.3390/ceramics4040041

Chicago/Turabian Style

Silva, Georgenes M. G., Victor N. S. Leão, Michel F. G. Pereira, Pedro M. Faia, and Evando S. Araújo. 2021. "Zn2+-Doped TiO2:WO3 Films Prepared by Electrospinning and Sintering: Microstructural Characterization and Electrical Signature to Moisture Sensing" Ceramics 4, no. 4: 576-591. https://doi.org/10.3390/ceramics4040041

Article Metrics

Back to TopTop