Next Article in Journal
Datasets of Groundwater Level and Surface Water Budget in a Central Mediterranean Site (21 June 2017–1 October 2022)
Previous Article in Journal
A Global Multiscale SPEI Dataset under an Ensemble Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Data Descriptor

Experimental Spectroscopic Data of SnO2 Films and Powder

1
Department of Physics and Astronomy, Howard University, Washington, DC 20059, USA
2
Physical Measurement Laboratory, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA
3
Theiss Research, La Jolla, CA 92037, USA
*
Author to whom correspondence should be addressed.
Submission received: 10 November 2022 / Revised: 21 December 2022 / Accepted: 28 January 2023 / Published: 9 February 2023
(This article belongs to the Section Chemoinformatics)

Abstract

:
Powders and films composed of tin dioxide (SnO2) are promising candidates for a variety of high-impact applications, and despite the material’s prevalence in such studies, it remains of high importance that commercially available materials meet the quality demands of the industries that these materials would most benefit. Imaging techniques, such as scanning electron microscopy (SEM), atomic force microscopy (AFM), were used in conjunction with Raman spectroscopy and X-ray photoelectron spectroscopy (XPS) to assess the quality of a variety of samples, such as powder and thin film on quartz with thicknesses of 41 nm, 78 nm, 97 nm, 373 nm, and 908 nm. In this study, the dependencies of the corresponding Raman, XPS, and SEM analysis results on properties of the samples, like the thickness and form (powder versus film) are determined. The outcomes achieved can be regarded as a guide for performing quality checks of such products, and as reference to evaluate commercially available samples.

1. Summary

Tin (IV) oxide (SnO2) has a large range of applications in today’s world, and there are fields in which the use of SnO2 has become incredibly important. Some of these efforts seek to develop and optimize environmental detectors [1,2,3,4,5,6,7,8], solar cell technologies [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23], and solid-state chemical sensors [24,25,26,27,28,29,30,31,32,33,34,35,36]. This n-type semiconductor has been utilized in ancient civilizations in a variety of applications from ceramic dye to polishing and glass coatings [37]. SnO2 has also been useful in the development of optical filters [38,39,40], photovoltaic devices [41,42,43,44], and materials dedicated to catalytic processes [45,46,47,48,49,50]. In addition to these applications, this material can also be used for lithium–ion batteries and photocatalysis [51,52].
These applications have led to the wide commercial availability of SnO2. With this ease of sample acquisition, it also remains important for materials to be tested for quality and compared with other counterparts in the literature. Some of the ways in which one may assess the various properties of SnO2 are through the measurement of its surface, optical, and electrical properties.
In this work, imaging techniques, such as scanning electron microscopy (SEM) and atomic force microscopy (AFM), were used in conjunction with Raman spectroscopy and X-ray photoelectron spectroscopy (XPS) to assess the quality of a variety of samples, many of which differ in thickness and were either prepared as films or powders. As such, the dependencies of the corresponding Raman spectra, XPS surveys, and SEM images on properties such as the thickness have been determined and reported here as a guide for those performing similar quality checks as well as to add to the statistics of reproducible results for commercially available samples.

2. Data Description

2.1. Scanning Electron and Atomic Force Microscopy

Scanning electron microscopy was used to gain an initial understanding on how the powder and film SnO2 samples appeared in terms of grain sizes and uniformity. As such, these resulting images, shown in Figure 1 and whose parameter data are provided in the Methods section, show that the superficial layers within thicker samples (notably the 908 nm and 373 nm) are not as uniform. Such an appearance may result by way of thermal cracking or by spontaneous seeding (the samples were not annealed during the SEM process, only during manufacturing). This interpretation can be supported by the behavior where, at 373 nm, smaller clusters appear to form than in the case of the 908 nm film (Figure 2). The SEM images also reveal that, for the films thinner than 373 nm, the superficial surface layers appear more uniform. Given that the same measurement parameters were used for all samples, it appears that the 41 nm has the least disordered appearance from all of the films. As far as the powder samples are concerned, their clusters appear more unbounded, with almost a near-spherical smoothness, at least within the 100 nm measurement scale used for all images. Energy Dispersive Spectroscopy (EDS) analysis was carried out on the 908 nm thick sample to confirm the chemical elements of SnO2 (Figure 3). The composition came out to be 31.3% Sn (by weight) and 68.7% O (by weight). The EDS data confirmed that the film is indeed SnO2.
AFM images were acquired for the samples shown in Figure 4a,b (908 nm and 373 nm, respectively) to gain a better understanding of the expected surface roughness that results from the clustering seen superficially. From the two-dimensional profiles measured, the texture for the 908 nm film appeared more granular than its 373 nm counterpart, whose microscope grains appear slightly smoother. A three-dimensional roughness display was plotted and is displayed below the images in Figure 4 as a visualization of the difference in surface roughness between the two films. Though the difference in smoothness between the two films is more obvious in this three-dimensional picture, it is still beneficial to quantify that difference via the root-mean-square (RMS) roughness. This quantity may be calculated by software such as Gwyddion. It was found that the RMS roughness for the 908 nm and 373 nm films was about 8.8 nm and 5.4 nm, respectively. It should be noted that these two thicknesses were the only ones measurable without error (that is, other images have sections indicative of AFM tip damage and error).

2.2. Raman Spectroscopy

Raman spectroscopy was utilized to gain a further understanding of these samples’ optical properties. In the first set of tests, powder samples were compressed into pellets and then loaded into a cryogenic chamber with temperature control. The intent was to conduct thermal cycling starting from the highest temperature available on the measurement system (400 °C), and then repeat every set of data by starting from the lowest temperature available on the measurement system (−193 °C). In all cases, these data provided insight into which vibrational modes were present and whether or not these observations were consistent with what has been reported in earlier work. This optical check is one way by which industries can verify the quality of semiconductor materials. All experimental parameters are listed in the Methods section.
Figure 5 shows data from the powder samples. These data are temperature dependent and have endpoints of −193 °C and 400 °C resulting from the limits of the measurement system.# The corresponding data in Figure 5b for measurements performed with a reverse order of temperature steps, starting with the highest temperature, do not show any obvious hysteresis when compared with Figure 5a. This similarity and lack of hysteresis is best seen in Figure 6. For the datasets in Figure 6b–d, three vibrational modes are tracked, namely the Eg, A1g, and B2g modes, located at (for −193 °C) 476.6 cm−1, 636.5 cm−1, and 779.5 cm−1, respectively. Figure 5a,b shows a subset of peaks (for visual clarity) whose centers align with the three dashed vertical lines. These demarcations have been aligned with the coldest temperature data to provide a visualization for the amount by which a given peak shifts with temperature. These shifts, along with the formal assignment of the modes, are provided in Figure 6.
At the Gamma point of the Brillouin zone for SnO2, one expects various phonon responses, including: 1A1g + 1B1g + 1B2g + 1Eg + 1A2g + 2B1u + 2A2u + 4Eu [53]. The first four are Raman active, followed by two that are optically inactive, and concluded by two that are infrared (IR) active. Several additional modes appear in the data, especially with high temperatures (Figure 6a). One interpretation of these additional peaks involves forbidden modes that appear at high temperatures (greater than 300 °C) due to an increase in defect density or disorder in the material. It should be noted that IR modes can become weakly active when structural changes emerge due to a disorder.
Raman spectra from powder samples were analyzed and interpreted to determine the presence of various vibrational modes. For Figure 6b–d, the full widths at half-maxima (FWHM) are shown as a function of temperature for the Eg, A1g, and B2g modes, respectively. Datasets, where the temperature is incrementally increased and decreased, are both included to show a minimal contribution from hysteresis. The Eg mode exhibits a FWHM that only increases by about 7 cm−1 over the span of the temperature range, whereas the other two modes exhibit an increase of just over 10 cm−1.
In Figure 6e–g, the peak centers are shown as a function of temperature for the Eg, A1g, and B2g modes, respectively. The peak centers were determined by Lorentzian fits to the Raman modes. It is evident from these plots that some of the modes experience stronger shifts than others, but all exhibit redshifting with increased temperature. The Eg, A1g, and B2g modes exhibit a redshift of about 3 cm−1, 10 cm−1, and 14 cm−1, respectively. For the latter two modes, the trends appear to reflect a subtle nonlinear relationship with temperature, but the subtlety is great enough that fitting this behavior to a polynomial or other nonlinear function would have to be carried out arbitrarily until a greater range of temperatures could be measured.
Though the SnO2 powder has provided us with ample information regarding the expected vibrational modes, it is still useful to obtain spectra of the various films to determine if there are any obvious signs of thickness dependence for those modes. As such, Figure 7 shows these data including the background spectrum for the bare quartz upon which the films reside. Raman spectra were collected for thicknesses of 908 nm, 373 nm, 97 nm, 78 nm, and 41 nm. Using the 908 nm film as an exemplary sample for the non-powdered samples, several of the modes were assigned and are indicated in Figure 7b. It shows a peak pattern change between bare quartz and 908 nm SnO2 film thickness. The bare quartz indicates a pronounced peak around 438 cm−1 due to Si-O-Si symmetric stretching [54,55]. Two peaks appeared at the 908 nm film at 238 cm−1 and 352 cm−1 that could be assigned to E U ( 1 ) ( TO ) and A2g (optically inactive mode), respectively [56]. The peaks that are positioned at 487 cm−1 and 602 cm−1 in bare quartz and at 489 cm−1 and 604 cm−1 in the SnO2 film are associated with D1 and D2 defect modes of 4-fold and 3-fold rings in the normal 6-fold rings of SnO2 respectively 55, 57. The shoulder peak that appears on the SnO2 film at 712 cm−1 can be assigned to E U ( 3 ) ( LO ) [56]. The 800 cm−1 peak that appears on bare quartz is associated with peak represented as ω3 band that coincides with Si motion against the tetrahedral O cage as symmetric bond stretching (O-Si-O) [55,56,57,58]. The following peaks at 1055 cm−1 in bare quartz and 1064 cm−1 in SnO2 film are associated with the symmetric stretching of Si-O [55].
Since some of these modes also shift as a function of temperature, it was important to perform measurements with more granular steps. It should be noted that each of the three modes of interest are slightly shifted with respect to other work reported in the literature. For instance, in Bhadrapriya et al. [54], the Eg, A1g, and B2g modes are assigned at 476.6 cm−1, 637.9 cm−1, and 793.4 cm−1, respectively. This aligns nicely with the powder samples, whose corresponding values (all at room temperature) are about 475 cm−1, 634 cm−1, and 775 cm−1, respectively. However, that does not appear to be as true for the films. Using 908 nm, the assigned peaks take on values of about 490 cm−1, 602 cm−1, and 707 cm−1, respectively. The reason for this shift may be attributed to factors such as contamination or oxygen vacancies in the SnO2 samples. Additional data (to be presented in the next section) for both reasons must corroborate that attribution. Raman spectra generally do not show the band at around 582 cm−1 and A2u band at around 694 cm−1, indicating the lack of vacant sites and local lattice disorder as described in Gouvêa et al. and Kamali et al. [59,60].
Table 1 shows the different peak values associated with 908 nm film, bare quartz, and SnO2 powder pellet and its effect with temperature. The majority of the SnO2 film peaks are associated with SiO2 (i.e., quartz glass) and can be summarized as follows:
  • It appeared that the 438 cm−1 Raman mode observed on bare quartz is close to the Zur et al. band of SiO2-SnO2 (transparent glass-ceramic) and could be assigned to vibrations of symmetric stretching of Si-O-Si bonds in six-membered rings which assigned to SiO4 tetrahedra deformation vibration [55,57]. The narrowing of the 438 cm−1 peak is due to the existence of SnO2 that could provoke a tighter Si-O-Si angle of SiO4 tetrahedra or obstruct the construction of higher-membered rings [55,58]. Another study by McMillan associates the narrowing of the peak along with the wavenumber shift to the tighter Si-O-Si angle that is for a more compact SiO2 structure. Therefore, adding SnO2 to SiO2 could have prevented the formation of higher-membered rings [61].
  • The two peaks at 488 cm−1 and 602 cm−1 in SnO2 film and 487 cm−1 and 602 cm−1 in bare quartz are related to D1 and D2 defect bands that are assigned to the defects of 4-fold and 3-fold rings in the normal 6-fold rings of SnO2, respectively [55,57]. It also appears that the addition of SnO2 into the SiO2 matrix has modified the 438 cm−1 peak with a D1 defect band.
  • A shoulder peak of 602 cm−1 in the SnO2 908 nm film at room temperature and at −193 °C around 623 cm−1 could be related to the addition of SnO2 to SiO2. This band is weak and overlaps with the D1 defect band due to the disorder of the SnO2 nanoparticle [55,62]. The broadening of the A1g peak is due to a smaller grain size in the film.
  • The peak at 739 cm−1 is a shoulder peak of that at 783 cm−1, which only appeared at −193 °C and may be due to E U ( 3 ) ( LO ) . This value is closer to the specified value in Lan et al. [56].
  • Next, the 783 cm−1 peak is represented as a ω3 band that coincides with the Si motion against the tetrahedral O cage as symmetric bond stretching (O-Si-O) [55,56,57,58].
  • Finally, the peak that appears on SnO2 film at 1063 cm−1 and on bare quartz at 1055 cm−1 is associated with the symmetric stretching of Si-O [55].
Again, the 908 nm film was used as an example in determining the temperature dependence of various Raman modes. All of these spectra are shown in Figure 8, with the same plotting scheme adopted from Figure 7, where both increasing and decreasing temperature data are shown. All experimental parameters are listed in the Methods section. In Figure 8a (left panel), several of the modes are tracked with dashed vertical lines, with the former panel conducted in a similar fashion to the powder samples. However, for Figure 8a (right panel), dashed lines are drawn to highlight the shifts observed in the three modes with temperature as a guide for the eye. All of the peaks consistently redshift with increasing temperature, as seen in the powder samples. The extent of shifting is even of similar value, but it is more difficult to conclude a similarity in trends with the FWHM because the peaks have a characteristically larger error when being fit with a Lorentzian profile.

2.3. X-ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy helps one determine if any impurities get introduced during the synthesis of the films and powders. Furthermore, the technique helps one verify the stoichiometric ratio and general composition of all of the materials. Figure 9 and Figure 10 show an array of data obtained for each film thickness and the powder form. More specifically, Figure 9 shows the survey scans, the oxygen 1s electrons, and the tin 3d electrons, whereas Figure 10 shows the nitrogen 1s electrons, carbon 1s electrons, and silicon 2p electrons.
The survey scans show mostly standard behavior with the one exception that there may be indium contamination, as exemplified by the subtle peaks at about 460 eV (one of the sputter targets had a minuscule amount of indium as well). This observation corroborates the aforementioned suspicion on the possibility of dopants modifying observed Raman modes. As far as the oxygen scan goes, one interesting observation to point out is the presence and trending behavior of the non-oxide peak near 532 eV. Not only does powdered SnO2 show the most subdued non-oxide peak, but the contribution itself increases for the film samples (though it does not have a strong thickness dependence). Despite not linearly increasing in contribution, the non-oxide peak does, in fact, shift to slightly higher binding energies with increasing thickness. For the third column of data, focused on the tin 3d electrons, no reliable trends are observable, with the exception of a generally increased energy for the relevant electrons observed in the 41 nm film and the powder samples.
Additional observations can be noted from the data shown in Figure 8, starting with the nitrogen 1s electrons. The difference seen is mainly between film and powder, where the response from the powder is 4 to 5 times lower than its film counterpart. Lastly, the silicon 2p electrons exhibit the smallest of all the responses, with none recorded from the powder, and very weak signals recorded from the films. No strong measurable trend can be extracted from the data collected on the samples. All experimental parameters are described in the Methods section.
The XPS peak positions were extracted by fitting them with Lorentzian profiles, with the relevant data presented in Table 2 and Table 3. One observation that remains interesting is that the non-oxide oxygen 1s electrons appear to demonstrate a response that shifts with the film thickness, as also mentioned earlier by reviewing the plots in Figure 9. Though the shifting is very gradual, it does fall within the range of uncertainties from the fitting procedure. A closer look at the elemental percentages from Figure 9 reveals that, although Sn percentages are uncorrelated with thickness, the non-oxide oxygen composition is weighted slightly more heavily for thinner samples, within the errors of the fitting procedures. A thickness of about 100 nm appears to divide the elemental composition by a few percent.

3. Methods

Methods are generally summarized in Figure 11. The SnO2 powder was acquired from SIGMA-ALDRICH. It has 99.9% purity with 325 mesh, with a target being six-inch diameter high purity 99.99% ceramic SnO2 with a small percentage (0.5%) of indium. It was annealed at 1000 °C for 1 hour. SnO2 thin samples on 0.1-inch quartz thick were provided by General Monitors and manufactured by radio frequency (RF) sputter deposition in an oxygen/argon mixed atmosphere. The sputter times varied on the single-digit minute scale. The deposited SnO2 film thickness was monitored by sputter deposition onto a polished silicon substrate [63]. Double-sided polished silicon was used to enable infrared transmission measurement studies and visible-ultraviolet reflectance measurements that were used to calculate the SnO2 film thickness. A Filmetrics F20 instrument was used for curve fitting to calculate film thickness. Film thickness was varied by changing the sputter deposition time. UV–quartz substrates were placed in the sputtering machine alongside the polished silicon thickness monitor substrate [64].
SEM data were acquired with a FEI Helios Focused Ion Beam, with a magnification that ranged from 200× to 200,000×. Other parameters are as follows: 2 kV, 0.1 nA, 746 nm horizontal field width, 10 μs dwell time, use of through lens detector, 0° tilt, and a 3.8 mm working distance. The samples were placed on a holder and then loaded into a chamber where plasma cleaning was applied to clean carbon from the sample surfaces. The images were obtained with the Everhart–Thornley detector (TLD) mode for better resolution, especially for the film samples. The SnO2 powder was sprinkled lightly on carbon tape, whereas the SnO2 films on quartz were placed on carbon tape and atop copper tape for better conductivity and resolution. The cross-section images were collected by using Phenom Pure Desktop SEM (Nanoscience Instrument) operated at a magnification that ranges between 20× to 65,000×. The EDS analysis was obtained using JEOL JSM-7100F SEM with magnification ranges from 10× to 500,000×.
The Atomic Force Microscope (AFM) images were recorded by using The Bruker Dimension FastScan atomic force microscope (AFM) combines a high-speed scanning AFM and programmable stage control. It provides rapid nanoscale imaging capabilities on substrates ranging from 200 mm diameter wafers down to small pieces. The system supports scan areas of up to 35 µm × 35 µm and can automatically collect data from multiple locations on a substrate.
Raman spectroscopy data were recorded on a Renishaw via Raman spectrometer using a 514 nm wavelength excitation laser source, laser beam quality of 0.65 mm, and laser maximum power of 50 mW. The spectra were collected using a laser exposure time of 10 s, 50× objective, and 3 accumulations to reduce the signal-to-noise ratio. The Linkam LTS420 heating and freezing stage and T96 LinkPad system controller with a maximum heating rate of 40 °C/min were used with Raman spectroscopy. The maximum stage temperature was 420 °C, and the minimum temperature was −195 °C with a LNP96-S liquid nitrogen cooling pump. To prevent condensation on the stage window surface during cooling, two steps were carried out: (1) a small tube was placed on the stage window that blows warm recycled nitrogen, and (2) air was purged from the stage chamber with dry nitrogen.
XPS measurements were performed using the Kratos Axis Ultra X-ray photoelectron spectrometer operating at a base pressure of about 2.66 × 107 Pa (about 2 × 109 Torr). All samples were analyzed using monochromatic Al K (1486.7 eV) with spot size 300 µm × 700 µm. Pass energies of 160 eV and 20 eV were used to collect the survey and high-resolution core level XPS spectra, respectively. All experimental uncertainties (1σ) arise from the uncertainty of the instrumentation used for this work.

Author Contributions

H.A.: conceptualization, methodology, software, validation, formal analysis, writing—original draft, visualization. O.Z.F.: conceptualization, validation, formal analysis, writing—original draft. M.L.K.: validation, writing—original draft. A.J.B.: validation, writing—original draft. D.S.: validation, writing—original draft. T.A.: validation, writing—original draft. K.S.: validation, writing—original draft. A.R.H.W.: validation, writing—original draft, supervision. C.A.H.: validation, writing—original draft, supervision. A.F.R.: validation, writing—original draft, supervision. P.M.: validation, writing—original draft, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

NSF Award No. PHY 1950379, NSF Award No. DMR 2101121, DMR190126, and PHY210066.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available from corresponding author upon reasonable request.

Acknowledgments

P.M. would like to acknowledge the financial support from the National Science Foundation REU Site in Physics at Howard University (NSF Award No. PHY 1950379) and Excellence in Research (NSF Award No. DMR 2101121), and allocation support from Advanced Cyberinfrastructure Coordination Ecosystem: Services and Support (ACCESS) (Grant Nos. DMR190126 and PHY210066). He also wishes to acknowledge the thin film samples provided by General Monitors through the auspices of Shankar Baliga, Laser Components DG, Inc., Chandler, AZ 85225. M.K. acknowledges support from the National Institute of Standards and Technology (NIST) Financial Assistance Award with Federal Award ID 70NANB19H152.

Conflicts of Interest

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

Commercial equipment

Instruments and materials are identified in this paper in order to specify the experimental procedure adequately. Such identification is not intended to imply recommendation or endorsement by the National Institute of Standards and Technology or the United States government, nor is it intended to imply that the materials or equipment identified are necessarily the best available for the purpose.

# Though not an SI Unit

This unit is permitted for expressing data by the International Committee for Weights and Measures (CIPM) and the National Institute of Standards and Technology.

References

  1. Lee, I.; Choi, S.-J.; Park, K.-M.; Lee, S.S.; Choi, S.; Kim, I.-D.; Park, C.O. The stability, sensitivity and response transients of ZnO, SnO2 and WO3 sensors under acetone, toluene and H2S environments. Sens. Actuators B Chem. 2014, 197, 300–307. [Google Scholar] [CrossRef]
  2. Yuliarto, B.; Gumilar, G.; Septiani, N.L.W. SnO2 nanostructure as pollutant gas sensors: Synthesis, sensing performances, and mechanism. Adv. Mater. Sci. Eng. 2015, 2015, 694823. [Google Scholar] [CrossRef]
  3. Sharma, A.; Tomar, M.; Gupta, V. SnO2 thin film sensor with enhanced response for NO2 gas at lower temperatures. Sens. Actuators B Chem. 2011, 156, 743–752. [Google Scholar] [CrossRef]
  4. Gong, J.; Chen, Q.; Lian, M.-R.; Liu, N.-C.; Stevenson, R.G.; Adami, F. Micromachined nanocrystalline silver doped SnO2 H2S sensor. Sens. Actuators B Chem. 2006, 114, 32–39. [Google Scholar] [CrossRef]
  5. Korotcenkov, G.; Brinzari, V.; Cho, B.K. In2O3-and SnO2-based thin film ozone sensors: Fundamentals. J. Sens. 2016, 2016, 1–32. [Google Scholar] [CrossRef]
  6. Suman, P.H.; Felix, A.A.; Tuller, H.L.; Varela, J.A.; Orlandi, M.O. Comparative gas sensor response of SnO2, SnO and Sn3O4 nanobelts to NO2 and potential interferents. Sens. Actuators B Chem. 2015, 208, 122–127. [Google Scholar] [CrossRef]
  7. Mills, S.; Lim, M.; Lee, B.; Misra, V. Atomic layer deposition of SnO2 for selective room temperature low ppb level O3 sensing. ECS J. Solid State Sci. Technol. 2015, 4, S3059. [Google Scholar] [CrossRef]
  8. Kim, B.; Lu, Y.; Hannon, A.; Meyyappan, M.; Li, J. Low temperature Pd/SnO2 sensor for carbon monoxide detection. Sens. Actuators B Chem. 2013, 177, 770–775. [Google Scholar] [CrossRef]
  9. Vishwakarma, S.R.; Prasad, H.C. Low cost SnO2: P/SiO2/n-Si (textured) heterojunction solar cells. J. Phys. D Appl. Phys. 1993, 26, 959. [Google Scholar] [CrossRef]
  10. Vasu, V.; Subrahmanyam, A. Photovoltaic properties of indium tin oxide (ITO)/silicon junctions prepared by spray pyrolysis-dependence on oxidation time. Semiconductor Sci. Technol. 1992, 7, 320. [Google Scholar] [CrossRef]
  11. Ferlauto, A.S.; Ferreira, G.M.; Pearce, J.M.; Wronski, C.R.; Collins, R.W.; Deng, X.; Ganguly, G. Analytical model for the optical functions of amorphous semiconductors from the near-infrared to ultraviolet: Applications in thin film photovoltaics. J. Appl. Phys. 2002, 92, 2424–2436. [Google Scholar] [CrossRef]
  12. Tennakone, K.; Kumara, G.R.R.A.; Kottegoda, I.R.M.; Perera, V.P.S. An efficient dye-sensitized photoelectrochemical solar cell made from oxides of tin and zinc. Chem. Commun. 1999, 15–16. [Google Scholar] [CrossRef]
  13. Zheng, J.P.; Kwok, H.S. Preparation of indium tin oxide films at room temperature by pulsed laser deposition. Thin Solid Films 1993, 232, 99–104. [Google Scholar] [CrossRef]
  14. Ito, S.; Makari, Y.; Kitamura, T.; Wada, Y.; Yanagida, S. Fabrication and characterization of mesoporous SnO2/ZnO-composite electrodes for efficient dye solar cells. J. Mater. Chem. 2004, 14, 385–390. [Google Scholar] [CrossRef]
  15. Bagheri-Mohagheghi, M.M.; Shokooh-Saremi, M. The influence of Al doping on the electrical, optical and structural properties of SnO2 transparent conducting films deposited by the spray pyrolysis technique. J. Phys. D Appl. Phys. 2004, 37, 1248. [Google Scholar] [CrossRef]
  16. Liu, C.; Matsutani, T.; Yamamoto, N.; Kiuchi, M. High-quality indium tin oxide films prepared at room temperature by oxygen ion beam assisted deposition. EPL Europhys. Lett. 2002, 59, 606–611. [Google Scholar] [CrossRef]
  17. Reddy, V.S.; Das, K.; Dhar, A.; Ray, S.K. The effect of substrate temperature on the properties of ITO thin films for OLED applications. Semicond. Sci. Technol. 2006, 21, 1747–1752. [Google Scholar] [CrossRef]
  18. LavrenčičáŠtangar, U. Preparation and characterization of Mo-and Sb: Mo-doped SnO2 sol-gel-derived films for counter-electrode applications in electrochromic devices. J. Mater. Chem. 1995, 5, 617–624. [Google Scholar]
  19. Kityk, I.V.; Eboth, J.; Miedzinski, R.; Addou, M.; Sieder, H.; Karafiat, A. Optically-induced non-linear optical effects in indium–tin oxide crystalline films. Semicond. Sci. Technol. 2003, 18, 549–553. [Google Scholar] [CrossRef]
  20. Morgado, J.; Barbagallo, N.; Charas, A.; Matos, M.; Alcácer, L.; Cacialli, F. Self-assembly surface modified indium-tin oxide anodes for single-layer light-emitting diodes. J. Phys. D Appl. Phys. 2003, 36, 434–438. [Google Scholar] [CrossRef]
  21. Suzuki, T.; Yamazaki, T.; Oda, H. Effects of composition and thickness on the electrical properties of indium oxide/tin oxide multilayered films. J. Mater. Sci. 1989, 24, 1383. [Google Scholar] [CrossRef]
  22. Guillén, C.; Herrero, J. Influence of the film thickness on the structure, optical and electrical properties of ITO coatings deposited by sputtering at room temperature on glass and plastic substrates. Semicond. Sci. Technol. 2008, 23, 075002. [Google Scholar] [CrossRef]
  23. Munir, M.M.; Iskandar, F.; Yun, K.M.; Okuyama, K.; Abdullah, M. Optical and electrical properties of indium tin oxide nanofibers prepared by electrospinning. Nanotechnology 2008, 19, 145603. [Google Scholar] [CrossRef] [PubMed]
  24. Ferrere, S.; Zaban, A.; Gregg, B.A. Dye sensitization of nanocrystalline tin oxide by perylene derivatives. J. Phys. Chem. B 1997, 101, 4490–4493. [Google Scholar] [CrossRef]
  25. Huang, H.; Tan, O.K.; Lee, Y.C.; Tran, T.D.; Tse, M.S.; Yao, X. Semiconductor gas sensor based on tin oxide nanorods prepared by plasma-enhanced chemical vapor deposition with post plasma treatment. Appl. Phys. Lett. 2005, 87, 163123. [Google Scholar] [CrossRef]
  26. Varghese, O.K.; Malhotra, L.K. Electrode-sample capacitance effect on ethanol sensitivity of nano-grained SnO2 thin films. Sens. Actuators B Chem. 1998, 53, 19–23. [Google Scholar] [CrossRef]
  27. Sakanoue, T.; Nakatani, S.; Ueda, Y.; Izumi, H.; Ishihara, T.; Motoyama, M. Fabrication of well-ordered indium-tin-oxide film and characterization of organic films vacuum-deposited on it. Mol. Cryst. Liq. Cryst. 2003, 405, 59–66. [Google Scholar] [CrossRef]
  28. Khanna, A.; Kumar, R.; Bhatti, S.S. CuO doped SnO2 thin films as hydrogen sulfide gas sensor. Appl. Phys. Lett. 2003, 82, 4388. [Google Scholar] [CrossRef]
  29. Comini, E.; Faglia, G.; Sberveglieri, G.; Pan, Z.; Wang, Z.L. Stable and highly sensitive gas sensors based on semiconducting oxide nanobelts. Appl. Phys. Lett. 2002, 81, 1869–1871. [Google Scholar] [CrossRef]
  30. Song, J.E.; Lee, D.K.; Kim, Y.H.; Kang, Y.S. Preparation of water dispersed indium tin oxide sol solution. Mol. Cryst. Liq. Cryst. 2006, 444, 247–255. [Google Scholar] [CrossRef]
  31. Coles, G.S.V.; Williams, G.; Smith, B.J. The effect of oxygen partial pressure on the response of tin (IV) oxide based gas sensors. J. Phys. D Appl. Phys. 1991, 24, 633–641. [Google Scholar] [CrossRef]
  32. Panchapakesan, B.; DeVoe, D.L.; Widmaier, M.R.; Cavicchi, R.; Semancik, S. Nanoparticle engineering and control of tin oxide microstructures for chemical microsensor applications. Nanotechnology 2001, 12, 336–349. [Google Scholar] [CrossRef]
  33. Choi, Y.J.; Hwang, I.-S.; Park, J.-G.; Choi, K.J.; Park, J.H.; Lee, J.-H. Novel fabrication of an SnO2 nanowire gas sensor with high sensitivity. Nanotechnology 2008, 19, 095508. [Google Scholar] [CrossRef]
  34. Eastwood, P.G.; Claypole, T.C.; Watson, J.; Coles, G.S.V. The behaviour of tin dioxide sensors in exhaust environments at low and intermediate temperatures. Meas. Sci. Technol. 1993, 4, 524–533. [Google Scholar] [CrossRef]
  35. Llobet, E.; Vilanova, X.; Brezmes, J.; Alcubilla, R.; Calderer, J.; Sueiras, J.E.; Corrie, X. Conductance-transient analysis of thick-film tin oxide gas sensors under successive gas-injection steps. Meas. Sci. Technol. 1997, 8, 1133–1138. [Google Scholar] [CrossRef]
  36. Partridge, J.G.; Field, M.R.; Peng, J.L.; Sadek, A.Z.; Kalantar-zadeh, K.; Du Plessis, J.; McCulloch, D.G. Nanostructured SnO2 films prepared from evaporated Sn and their application as gas sensors. Nanotechnology 2008, 19, 125504. [Google Scholar] [CrossRef]
  37. Colomban, P.; Gironda, M.; Franci, G.S.; d’Abrigeon, P. Distinguishing Genuine Imperial Qing Dynasty Porcelain from Ancient Replicas by On-Site Non-Invasive XRF and Raman Spectroscopy. Materials 2022, 15, 5747. [Google Scholar] [CrossRef]
  38. Alamrani, F.; Alsharaeh, E. Controlling the Bandgaps of One-Dimensional TiO2/SiO2, TiO2/SnO2, and SiO2/SnO2 Photonic Crystals Using the Transfer Matrix Method. Optics Photon. J. 2022, 12, 171–189. [Google Scholar] [CrossRef]
  39. Gotschy, W.; Vonmetz, K.; Leitner, A.; Aussenegg, F.R. Thin films by regular patterns of metal nanoparticles: Tailoring the optical properties by nanodesign. Appl. Phys. B 1996, 63, 381–384. [Google Scholar] [CrossRef]
  40. Dhatarwal, P.; Sengwa, R.J.; Choudhary, S. Multifunctional (PVP/PEO)/SnO2 nanocomposites of tunable optical and dielectric properties. Optik 2020, 221, 165368. [Google Scholar] [CrossRef]
  41. Kavan, L. Conduction band engineering in semiconducting oxides (TiO2, SnO2): Applications in perovskite photovoltaics and beyond. Catal. Today 2019, 328, 50–56. [Google Scholar] [CrossRef]
  42. Subbiah, A.S.; Mathews, N.; Mhaisalkar, S.G.; Sarkar, S.K. Novel Plasma-assisted low-temperature-processed SnO2 thin films for efficient flexible perovskite photovoltaics. ACS Energy Lett. 2018, 3, 1482–1491. [Google Scholar] [CrossRef]
  43. Dalapati, G.K.; Sharma, H.; Guchhait, A.; Chakrabarty, N.; Bamola, P.; Liu, Q.; Saianand, G.; Krishna, A.M.S.; Mukhopadhyay, S.; Dey, A.; et al. Tin oxide for optoelectronic, photovoltaic and energy storage devices: A review. J. Mater. Chem. A 2021, 9, 16621–16684. [Google Scholar] [CrossRef]
  44. Yu, H.; Yeom, H.-I.; Lee, J.W.; Lee, K.; Hwang, D.; Yun, J.; Ryu, J.; Lee, J.; Bae, S.; Kim, S.K.; et al. Superfast room-temperature activation of SnO2 thin films via atmospheric plasma oxidation and their application in planar perovskite photovoltaics. Adv. Mater. 2018, 30, 1704825. [Google Scholar] [CrossRef]
  45. Pyke, D.R.; Reid, R.; Tilley, R.J.D. Structures of tin oxide—antimony oxide catalysts. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1980, 76, 1174–1182. [Google Scholar] [CrossRef]
  46. Dodd, A.; McKinley, A.; Saunders, M.; Tsuzuki, T. Mechanochemical synthesis of nanocrystalline SnO2–ZnO photocatalysts. Nanotechnology 2006, 17, 692. [Google Scholar] [CrossRef]
  47. Cai, X.; Ogorevc, B.; Tavčar, G.; Wang, J. Indium—Tin oxide film electrode as catalytic amperometric sensor for hydrogen peroxide. Analyst 1995, 120, 2579–2583. [Google Scholar] [CrossRef]
  48. Jyothi, T.M.; Sugunan, S.; Rao, B.S. Selective methylation of anisole to 2,6-xylenol over rare earth promoted SnO2 catalysts. Green Chem. 2000, 2, 269–271. [Google Scholar] [CrossRef]
  49. Bond, G.C.; Molloy, L.R.; Fuller, M.J. Oxidation of carbon monoxide over palladium—tin(IV) oxide catalysts: An example of spillover catalysis. J. Chem. Soc. Chem. Commun. 1975, 796–797. [Google Scholar] [CrossRef]
  50. Hines, R.I.; Allan, N.L.; Flavell, W.R. Oxidation catalysts: A comparative simulation study of the lattice, defect and surface structure of the stannates ASnO3(A = Ca, Sr and Ba) and SnO2. J. Chem. Soc. Faraday Trans. 1996, 92, 2057–2063. [Google Scholar] [CrossRef]
  51. He, Z.-K.; Sun, Q.; Xie, K.; Lu, P.; Shi, Z.; Kamali, A.R. Reactive molten salt synthesis of natural graphite flakes decorated with SnO2 nanorods as high performance, low cost anode material for lithium ion batteries. J. Alloy. Compd. 2019, 792, 1213–1222. [Google Scholar] [CrossRef]
  52. Kohl, D.; Thoren, W.; Schnakenberg, U.; Schüll, G.; Heiland, G. Decomposition of gaseous acetic acid on SnO2. J. Chem. Soc. Faraday Trans. 1991, 87, 2647–2653. [Google Scholar] [CrossRef]
  53. Bonu, V.; Das, A.; Sivadasan, A.K.; Tyagi, A.K.; Dhara, S. Invoking forbidden modes in SnO2 nanoparticles using tip enhanced Raman spectroscopy. J. Raman Spectrosc. 2015, 46, 1037–1040. [Google Scholar] [CrossRef]
  54. Bhadrapriya, B.C.; Varghese, A.R.; Amarendra, G.; Hussain, S. Sb:SnO2 thin films—Synthesis and characterization. AIP Conf. Proc. 2018, 1951, 030006. [Google Scholar]
  55. Zur, L.; Tran, L.T.N.; Massella, D.; Vaccari, A.; Chiappini, A.; Chiasera, A.; Varas, S.; Armellini, C.; Carpentiero, A.; Boulard, B.; et al. SiO2-SnO2 transparent glass-ceramics activated by rare earth ions. Opt. Compon. Mater. 2019, 10914, 180–188. [Google Scholar]
  56. Lan, T.; Li, C.W.; Fultz, B. Phonon anharmonicity of rutile SnO2 studied by Raman spectrometry and first principles calculations of the kinematics of phonon-phonon interactions. Phys. Rev. B 2012, 86, 134302. [Google Scholar] [CrossRef]
  57. Manghnani, M.H.; Hushur, A.; Sekine, T.; Wu, J.; Stebbins, J.F.; Williams, Q. Raman, Brillouin, and nuclear magnetic resonance spectroscopic studies on shocked borosilicate glass. J. Appl. Phys. 2011, 109, 113509. [Google Scholar] [CrossRef]
  58. Van Tran, T.; Turrell, S.; Eddafi, M.; Capoen, B.; Bouazaoui, M.; Roussel, P.; Berneschi, S.; Righini, G.C.; Ferrari, M.; Bhaktha, S.N.B.; et al. Investigations of the effects of the growth of SnO2 nanoparticles on the structural properties of glass–ceramic planar waveguides using Raman and FTIR spectroscopies. J. Mol. Struct. 2010, 976, 314–319. [Google Scholar] [CrossRef]
  59. Gouvêa, D.; Pereira, G.J.; Gengembre, L.; Steil, M.C.; Roussel, P.; Rubbens, A.; Hidalgo, P.; Castro, R.H. Quantification of MgO surface excess on the SnO2 nanoparticles and relationship with nanostability and growth. Appl. Surf. Sci. 2011, 257, 4219–4226. [Google Scholar] [CrossRef]
  60. Kamali, A.R.; Fray, D.J. Solid phase growth of tin oxide nanostructures. Mater. Sci. Eng. B 2012, 177, 819–825. [Google Scholar] [CrossRef]
  61. McMillan, P.F.; Remmele, R.L. Hydroxyl sites in SiO2 glass: A note on infrared and Raman spectra. Am. Mineral. 1986, 71, 772–778. [Google Scholar]
  62. Van Tran, T.T.; Bui, T.S.; Turrell, S.; Capoen, B.; Roussel, P.; Bouazaoui, M.; Ferrari, M.; Cristini, O.; Kinowski, C. Controlled SnO2 nanocrystal growth in SiO2-SnO2 glass-ceramic monoliths. J. Raman Spectrosc. 2012, 43, 869–875. [Google Scholar] [CrossRef]
  63. Saji, K.J.; Mary, A.P.R. Tin oxide based p and n-type thin film transistors developed by RF sputtering. ECS J. Solid State Sci. Technol. 2015, 4, Q101–Q104. [Google Scholar] [CrossRef]
  64. Alghamdi, H.; Concepcion, B.; Baliga, S.; Misra, P. Synthesis, spectroscopic characterization and applications of tin dioxide. In Contemporary Nanomaterials in Material Engineering Applications; Springer: Cham, Switzerland, 2021; pp. 285–318. [Google Scholar]
Figure 1. SEM images show cross-section images of SnO2 films of six different thicknesses (i.e., 41 nm, 78 nm, 97 nm, 373 nm, and 908 nm.
Figure 1. SEM images show cross-section images of SnO2 films of six different thicknesses (i.e., 41 nm, 78 nm, 97 nm, 373 nm, and 908 nm.
Data 08 00037 g001
Figure 2. These SEM images show SnO2 films of six different thicknesses or forms at similar magnifications: (a) powder, (b) 908 nm thick, (c) 373 nm thick, (d) 97 nm thick, (e) 78 nm thick, and (f) 41 nm thick. Relevant measurement parameters can be found in the Methods section.
Figure 2. These SEM images show SnO2 films of six different thicknesses or forms at similar magnifications: (a) powder, (b) 908 nm thick, (c) 373 nm thick, (d) 97 nm thick, (e) 78 nm thick, and (f) 41 nm thick. Relevant measurement parameters can be found in the Methods section.
Data 08 00037 g002
Figure 3. EDS analysis of 908 nm film thickness.
Figure 3. EDS analysis of 908 nm film thickness.
Data 08 00037 g003
Figure 4. Comparative data are shown for two atomic force microscope (AFM) images obtained on a (a) 908 nm film and a (b) 373 nm film. A three-dimensional roughness display is shown below the images as a visualization of the difference in surface roughness between the two films. The scale on the roughness ranges from 0 nm to 39 nm, where the average thickness value is the center of the scale (18.5 nm). The variation generally falls within the ranges shown.
Figure 4. Comparative data are shown for two atomic force microscope (AFM) images obtained on a (a) 908 nm film and a (b) 373 nm film. A three-dimensional roughness display is shown below the images as a visualization of the difference in surface roughness between the two films. The scale on the roughness ranges from 0 nm to 39 nm, where the average thickness value is the center of the scale (18.5 nm). The variation generally falls within the ranges shown.
Data 08 00037 g004
Figure 5. Powder samples were measured at experimental parameters listed in the Methods section. Powder was compressed into standard pellets prior to measurement and subsequently loaded into a cryogenic chamber with temperature control. (a) A series of Raman spectra are shown as a function of temperature, starting with −193 °C and ending at 400 °C with five spectra were plotted with greater vertical scaling to distinguish the changes apparent with some of the vibrational modes. (b) Measurements were generally repeated with temperatures decreasing from 400 °C to assess any hysteresis or other unaccounted effects.
Figure 5. Powder samples were measured at experimental parameters listed in the Methods section. Powder was compressed into standard pellets prior to measurement and subsequently loaded into a cryogenic chamber with temperature control. (a) A series of Raman spectra are shown as a function of temperature, starting with −193 °C and ending at 400 °C with five spectra were plotted with greater vertical scaling to distinguish the changes apparent with some of the vibrational modes. (b) Measurements were generally repeated with temperatures decreasing from 400 °C to assess any hysteresis or other unaccounted effects.
Data 08 00037 g005
Figure 6. Spectra from powder samples were analyzed and interpreted to determine the presence of various vibrational modes. (a) The comparison of the two temperature extrema allows one to determine the extent of vibrational mode shifting. (bd) The temperature dependence of the full width at half-maxima (FWHM) is shown for the Eg, A1g, and B2g modes, respectively. Datasets where temperature is being incrementally increased and decreased are both included to show a minimal contribution from hysteresis. (eg) The temperature dependence of the peak center is shown for the Eg, A1g, and B2g modes.
Figure 6. Spectra from powder samples were analyzed and interpreted to determine the presence of various vibrational modes. (a) The comparison of the two temperature extrema allows one to determine the extent of vibrational mode shifting. (bd) The temperature dependence of the full width at half-maxima (FWHM) is shown for the Eg, A1g, and B2g modes, respectively. Datasets where temperature is being incrementally increased and decreased are both included to show a minimal contribution from hysteresis. (eg) The temperature dependence of the peak center is shown for the Eg, A1g, and B2g modes.
Data 08 00037 g006
Figure 7. (a) Raman spectra are collected and shown from SnO2 films of various thicknesses (908 nm, 373 nm, 97 nm, 78 nm, and 41 nm) as well as the bare quartz substrate. (b) Comparison between SnO2 908 nm film thickness with bare quartz substrate at room temperature.
Figure 7. (a) Raman spectra are collected and shown from SnO2 films of various thicknesses (908 nm, 373 nm, 97 nm, 78 nm, and 41 nm) as well as the bare quartz substrate. (b) Comparison between SnO2 908 nm film thickness with bare quartz substrate at room temperature.
Data 08 00037 g007
Figure 8. Samples with a 908 nm thick SnO2 film were measured at experimental parameters listed in the Methods section. (a) A series of Raman spectra are shown as a function of temperature, starting with −193 °C and ending with 400 °C with sample of five spectra were plotted with greater vertical scaling to distinguish the changes apparent with some of the vibrational modes. (b) In the case of the 908 nm-thick film, the two temperature extrema (400 °C and −193 °C in black and purple, respectively) are shown with demarcations of various modes’ peak positions.
Figure 8. Samples with a 908 nm thick SnO2 film were measured at experimental parameters listed in the Methods section. (a) A series of Raman spectra are shown as a function of temperature, starting with −193 °C and ending with 400 °C with sample of five spectra were plotted with greater vertical scaling to distinguish the changes apparent with some of the vibrational modes. (b) In the case of the 908 nm-thick film, the two temperature extrema (400 °C and −193 °C in black and purple, respectively) are shown with demarcations of various modes’ peak positions.
Data 08 00037 g008
Figure 9. Many XPS measurements are shown in descending order of thickness, concluded by the powder samples. Each column is designated as one for either a survey scan, O 1s electrons, or Sn 3d electrons.
Figure 9. Many XPS measurements are shown in descending order of thickness, concluded by the powder samples. Each column is designated as one for either a survey scan, O 1s electrons, or Sn 3d electrons.
Data 08 00037 g009
Figure 10. Many XPS measurements are shown in descending order of thickness, concluded by the powder samples. Each column is designated as one for either N 1s electrons, C 1s electrons, or Si 2p electrons. Minimal Si counts relative to other electrons suggests adequate enough coverage of films on quartz substrates such that substrate is not probed.
Figure 10. Many XPS measurements are shown in descending order of thickness, concluded by the powder samples. Each column is designated as one for either N 1s electrons, C 1s electrons, or Si 2p electrons. Minimal Si counts relative to other electrons suggests adequate enough coverage of films on quartz substrates such that substrate is not probed.
Data 08 00037 g010
Figure 11. Summary of all samples and corresponding tests carried out in this paper.
Figure 11. Summary of all samples and corresponding tests carried out in this paper.
Data 08 00037 g011
Table 1. Comparison of vibrational modes observed for SnO2 908 nm film, bare quartz, and SnO2 powder [55].
Table 1. Comparison of vibrational modes observed for SnO2 908 nm film, bare quartz, and SnO2 powder [55].
Modes (All Units cm−1)908 nm FilmBare QuartzZur et al. [55]ModesSnO2 Powder
Ambient Condition –193 °C400 °C 30 °C –193 °C400 °C
Si-O-Si (symmetric stretching)438440
D1 defect bands488 460487489Eg476  472
A2u(TO)501
B1u542
D2 defect bands602 606602600A1g636  626
A2u(LO)697  686
739  —
O-Si-O (symmetric stretching)783800 800B2g779  764
Si-O (symmetric stretching)1063 10501055
Table 2. Tabulated sample peak positions are determined with Lorentzian profile fitting. The positions are averaged within the relevant set of data and accompanied by a 1σ error from the peak fitting procedures.
Table 2. Tabulated sample peak positions are determined with Lorentzian profile fitting. The positions are averaged within the relevant set of data and accompanied by a 1σ error from the peak fitting procedures.
Sample Peak Positions (Averaged, with 1σ Error from Peak Fitting)
Sample TypeSn 3d 5/2Sn 3d 3/2O 1s (Oxide)O 1s
(Non-Oxide)
908 nm486.65 ± 0.02495.07 ± 0.02530.48 ± 0.03532.05 ± 0.02
373 nm486. 74 ± 0.02495.16 ± 0.02530.57 ± 0.03532.03 ± 0.05
97 nm486.71 ± 0.01495.12 ± 0.01530.52 ± 0.02531.98 ± 0.03
78 nm486.80 ± 0.04495.21 ± 0.04530.48 ± 0.04531.73 ± 0.06
41 nm486.90 ± 0.03495.31 ± 0.03530.53 ± 0.03531.66 ± 0.08
Powder486.81 ± 0.13495.23 ± 0.14530.49 ± 0.36531.40 ± 0.35
Table 3. Tabulated composition percentages from XPS spectra of various forms of SnO2.
Table 3. Tabulated composition percentages from XPS spectra of various forms of SnO2.
Sample Elemental Percentages (Averaged, with 1σ Error from Peak Fitting)
Sample TypeSn %Oxide O %Non-Oxide O %Total O %
908 nm37.75 ± 0.2546.43 ± 0.4115.81 ± 0.6662.24 ± 0.26
373 nm38.31 ± 0.1946.62 ± 0.5015.02 ± 0.3961.69 ± 0.19
97 nm37.79 ± 0.4146.03 ± 1.2916.19 ± 1.6962.21 ± 0.41
78 nm38.27 ± 0.1542.17 ± 0.8719.55 ± 1.0261.73 ± 0.15
41 nm37.83 ± 0.0240.44 ± 1.1621.73 ± 1.1862.17 ± 0.02
Powder40.06 ± 0.2043.44 ± 0.2416.49 ± 0.1259.93 ± 0.20
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

Alghamdi, H.; Farinre, O.Z.; Kelley, M.L.; Biacchi, A.J.; Saha, D.; Adel, T.; Siebein, K.; Walker, A.R.H.; Hacker, C.A.; Rigosi, A.F.; et al. Experimental Spectroscopic Data of SnO2 Films and Powder. Data 2023, 8, 37. https://doi.org/10.3390/data8020037

AMA Style

Alghamdi H, Farinre OZ, Kelley ML, Biacchi AJ, Saha D, Adel T, Siebein K, Walker ARH, Hacker CA, Rigosi AF, et al. Experimental Spectroscopic Data of SnO2 Films and Powder. Data. 2023; 8(2):37. https://doi.org/10.3390/data8020037

Chicago/Turabian Style

Alghamdi, Hawazin, Olasunbo Z. Farinre, Mathew L. Kelley, Adam J. Biacchi, Dipanjan Saha, Tehseen Adel, Kerry Siebein, Angela R. Hight Walker, Christina A. Hacker, Albert F. Rigosi, and et al. 2023. "Experimental Spectroscopic Data of SnO2 Films and Powder" Data 8, no. 2: 37. https://doi.org/10.3390/data8020037

Article Metrics

Back to TopTop