3.1 XRD analysis
The calcined powder's crystallinity and phases were investigated using X-ray diffraction (XRD) spectroscopy using a Bruker D8 Advance apparatus. Cu-Kα radiation (λ = 1.542 Å) was used over a Bragg angle range of 20° to 80°. The phase of the nanocrystalline particles was confirmed by XRD analysis. The Scherrer formula was used to calculate the size of each crystallite.
The X-ray diffraction (XRD) pattern of calcined In2O3 nanocrystalline particles is displayed in Fig. 1. The diffraction pattern closely matches cubic In2O3 (JCPDS card number 06-0416), indicating that high-quality In2O3 nanocrystalline particles were successfully synthesized without any X-ray-detectable impurities. A preferentially oriented crystal plane of (222) is present in the particles generated by the sol-gel method. There is a significant degree of crystallinity in the sample, as seen by the strong diffraction peaks. Using least squares refinement, the powder's lattice constants were found from X-ray diffraction (XRD) peaks by indexing it in a cubic space group (Ia-3). This revealed a lattice parameter of a = b = c = 10.117 Å. The creation of crystalline In2O3 particles with a body-centered cubic (BCC) structure is confirmed by the XRD pattern. Approximately 15 nm is the typical crystallite size, as determined by the Scherrer formula.
Figure 2 depicts the X-ray diffraction (XRD) patterns of undoped and Zn-doped In2O3 samples at various Zn concentrations (1%, 3%, 5%, 7%, and 9% mole). The lattice constant decreased with increasing Zn concentration. The samples have a significant preferred orientation along the (222) direction, as well as lesser intensity reflections corresponding to the (211), (400), (440), and (622) planes, which is compatible with the JCPDS No. 06-0416 standard for In2O3 compounds. The strength of the (222) peak rises with Zn doping, indicating increased crystallinity. The high strength of the diffraction peaks in the XRD pattern indicates that the samples are highly crystalline. The lattice constants of undoped and Zn-doped In2O3 were determined to determine the effect of doped Zn ions.
In X-ray diffraction (XRD), the spacing between crystal planes (denoted as d) in a cubic crystal structure is related to the lattice constant (\(\:a\)) and Miller indices (hkl) by the following Eq. (1).
\(\:a=\frac{\sqrt{3}}{\text{s}\text{i}\text{n}{\theta\:}}\:\) λ (1)
Table 1
Lattice constant and crystallite size calculated for undoped and Zn doped In2O3
Sr.
No
|
Material
|
Lattice constant \(\:\varvec{a}\:\)(Å)
|
Crystallite Size (nm)
|
1
|
Undoped In2O3
|
10.117
|
15
|
2
|
1% Zn doped
|
10.111
|
16.62
|
3
|
3% Zn doped
|
10.108
|
16.22
|
4
|
5% Zn doped
|
10.107
|
14.88
|
5
|
7% Zn doped
|
10.107
|
13.27
|
6
|
9% Zn doped
|
10.103
|
18.36
|
Table 1 shows the lattice constants and crystallite sizes determined for undoped and Zn-doped In2O3 samples. The X-ray diffraction (XRD) spectra show significant alterations in peak positions with increasing Zn doping. We attribute these movements to two possible causes. For starters, because doping occurred at the percentage level rather than the atomic level, complete substitution of lattice sites may not have occurred, resulting in material stress. This behavior is explained by the substituting Zn2 + ion's lesser ionic radius (0.074 nm) compared to the host ion, In3+ (0.081 nm).
3.2 TEM analysis
The TEM picture of calcined In2O3 nanocrystalline powder and the matching selected area electron diffraction (SAED) pattern are displayed together in Fig. 3. The particles have diameters that range from 17 to 23 nm, and their form appears to be primarily spherical. The crystalline cubic structure of the In2O3 sample is indicated by the patchy ring pattern seen in its SAED pattern in the (222) plane, which lacks any extra diffraction spots or rings from second phases. The X-ray diffraction pattern (JCPDS card number 06-0416) and the measured interplanar spacings (d, hkl)) from the SAED pattern in Fig. 4 match and are consistent with the cubic phase of In2O3 nanocrystalline material. Reflections like (211), (222), (400), (411), (332), (431), and (440) correspond to these measurements.
A high-resolution transmission electron microscopy (HRTEM) image of In2O3 nanoparticles is shown in Fig. 3. The image shows well-crystallized particles in the form of single crystals, as indicated by distinct lattice planes that sharply define the borders of the particles. The majority of the particles are spherical in shape and have a tendency to group together.
3.3 UV-vis spectrum analysis
UV-Visible spectroscopy was used to examine the optical characteristics of In2O3. A clear absorption peak at 293 nm, or 4.23 eV, can be seen in the UV-visible absorption spectra of In2O3.
Fitting the absorption data to the direct transition equation yielded the sample's direct band gap energy (Eg) as given by Eq. (2).
αhν = ED (hν-Eg )1/2 (2)
Where, α is absorption coefficient, hν is photon energy, Eg is direct band gap and ED is constant. Plotting (αhν) 2 as function of photon energy and extrapolating linear portion of curve to absorption equal to zero. The direct band gap (Eg) deduced from this plot was found to be 3.6 eV. This value of optical band gap exactly matches with reported value. Plotting the changes in the band gap of produced In2O3 doped with different mole percentages of Zn (1%, 3%, 5%, 7%, 9%) against incident photon energy (hv) using the sol-gel process is shown in Fig. 6. Eq. (2) was used to compute the optical band gap. The link between photon energy (hν) and (αhν)2 is seen in Fig. 5.
3.4 FTIR analysis
In order to determine whether molecular species were present in undoped Indium Oxide (In2O3) and to see how Zn doping affected the vibrational modes of In2O3 nanoparticles, the study used FTIR spectroscopy. The Zn-doped In2O3 (Fig. 7) and undoped In2O3 (Fig. 8) FTIR transmission mode spectra were recorded. Peaks were found at 406, 428.96, 501.28, and 590.56 cm− 1, which correspond to the cubic In2O3 In-O phonon vibration mode. While CO2 absorption was responsible for the peak at 2347.07 cm− 1, OH stretching was indicated by a large absorption band about 3447.50 cm− 1. Absorption bands at 1340.52, 1507.72, and 1636.30 cm-1 were associated with nitrate groups and water bend deformation. Minor variations in the frequency or strength of these bands indicate susceptibility to varied In3+ distributions in interstitial locations and defects, such as oxygen vacancies. The findings verified complete phase development and the absence of organic intermediates in the samples. Zn-doped In2O3 (at 1, 3, 5, 7, and 9 mol %) showed a decrease in the intensity of In2O3 bands.
3.5 Thermal gravimetric analysis
The thermal behavior of the precursor material was investigated using thermal gravimetric analysis (TGA). The precursor employed was Indium Nitrate Hydrate (In(NO3)3 x H2O), and 9.446 mg of the sample was placed in a quartz pan to reduce bed effects during decomposition. The sample was heated in air from 30 to 800°C at a rate of 15°C per minute. The purge gas was nitrogen, flowing at a rate of 20 ml per minute. Figure 9 illustrates the sample's thermal behavior. The thermogram shows that the weight loss process continues up to 400°C, with the complete disintegration of Indium Nitrate Hydrate and the creation of In2O3 nanophase powder. The weight loss is due to the removal of physisorbed and chemisorbed water from In (NO3)3 x H2O. A weight loss of 10.04% was reported during the first step of thermal degradation (at about 150°C). The considerable weight loss observed between 150 and 250°C is most likely caused by the elimination of nitrate from the sample. The TGA curve shows essentially no weight loss between 400 and 800°C.
3.6 Application in gas sensing
The study was investigated the response of pure, 1, 3, 5, 7, and 9 mol% Zn-doped In2O3 thick film resistors (TFRs) to 100 ppm H2S at various operating temperatures as shown in Fig. 10. Sensitivity peaked at 150°C for both undoped and 7 mol% Zn-doped In2O3, with the latter being the most sensitive. Selectivity experiments at 150°C against C2H5OH and NH3 revealed that the 7 mol% Zn-doped sensor had superior H2S selectivity as shown in Fig. 11. The calibration curve for this sensor showed a linear response to H2S values of up to 200 ppm, followed by saturation. The 7 mol % Zn-doped sensor had a faster recovery time (~ 240 sec) compared to the undoped sensor, but the response time remained identical at ~ 7 sec. Table 2 shows the comparison between the sensing characteristics different materials and sensor performances.
Table 2
Comparison between the sensing characteristics different materials
Sr.
No
|
Materials
|
H2S
(ppm)
|
Optimal Temp.
(0C)
|
Sensitivity / Response
(%)
|
References
|
1
|
In2O3
|
50
|
150
|
68
|
[16]
|
2
|
CO: In2O3
|
50
|
125
|
82
|
[17]
|
3
|
La: In2O3
|
50
|
125
|
86
|
[18]
|
4
|
In2O3
|
10
|
195
|
50.2
|
[19]
|
5
|
Gd: In2O3
|
10
|
195
|
52.3
|
[20]
|
6
|
Zn: In2O3
|
100
|
150
|
98.74
|
Present work
|