Next Article in Journal
ZIF-67 Derived MnO2 Doped Electrocatalyst for Oxygen Reduction Reaction
Next Article in Special Issue
Synthesis of N-Doped TiO2 for Efficient Photocatalytic Degradation of Atmospheric NOx
Previous Article in Journal
1D Zn(II) Coordination Polymers as Effective Heterogeneous Catalysts in Microwave-Assisted Single-Pot Deacetalization-Knoevenagel Tandem Reactions in Solvent-Free Conditions
Previous Article in Special Issue
Promoting Photoelectrochemical Water Oxidation on Ti-Doped Fe2O3 Nanowires Photoanode by O2 Plasma Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low Temperature Deposition of TiO2 Thin Films through Atmospheric Pressure Plasma Jet Processing

1
Center for Advance Studies in Material Science and Solid State Physics, Department of Physics, Savitribai PhulePune University, (Formerly University of Pune), Pune 411007, India
2
Photocatalysis International Research Center, Research Institute for Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278-8510, Japan
3
Faculty of Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278-8510, Japan
4
Raje Ramrao Mahavidyalaya, Jath 416404, India
5
Research Center for Space Colony, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278-8510, Japan
6
Research Initiative for Supra-Materials, Shinshu University, Nagano 380-8553, Japan
*
Author to whom correspondence should be addressed.
Submission received: 13 December 2020 / Revised: 5 January 2021 / Accepted: 8 January 2021 / Published: 11 January 2021
(This article belongs to the Special Issue Commemorative Issue in Honor of Professor Akira Fujishima)

Abstract

:
Titanium dioxide (TiO2) has been widely used as a catalyst material in different applications such as photocatalysis, solar cells, supercapacitor, and hydrogen production, due to its better chemical stability, high redox potential, wide band gap, and eco-friendly nature. In this work TiO2 thin films have been deposited onto both glass and silicon substrates by the atmospheric pressure plasma jet (APPJ) technique. The structure and morphological properties of TiO2 thin films are studied using different characterization techniques like X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, and field emission scanning electron microscopy. XRD study reveals the bronze-phase of TiO2. The XPS study shows the presence of Ti, O, C, and N elements. The FE-SEM study shows the substrate surface is well covered with a nearly round shaped grain of different size. The optical study shows that all the deposited TiO2 thin films exhibit strong absorption in the ultraviolet region. The oleic acid photocatalytic decomposition study demonstrates that the water contact angle decreased from 80.22 to 27.20° under ultraviolet illumination using a TiO2 photocatalyst.

1. Introduction

TiO2 is one of the most studied semiconductor oxide materials due to its vast applications, which include photocatalytic dye degradation, water splitting, self-cleaning agent, dye sensitized solar cells, photocatalytic hydrogen production, etc. Low cost synthesis, biocompatibility and inertness under ambient conditions are additional factors that make it one of the most studied materials [1,2]. Apart from these, it absorbs ultra violet (UV) light significantly so it is usable in cosmetics, such as sunscreen lotion. The UV light absorptions relate to its electronic transition from valence band (VB) to conduction band (CB). VB of TiO2 is quite oxidative and hence its holes act as strong oxidant under UV light irradiation. Such band-to-band transition of TiO2 by absorbing UV light and strong oxidant nature of the VB holes are the reasons behind the above-mentioned applications [3,4].
The cost of good quality coating depends on the processing parameters. The low pressure and temperature plasma technique is costly as well as inept due to the expensive vacuum system requirement and, in addition, utilization of different kinds and sizes of substrates during the deposition process are limited [5]. Researchers have adopted different methods such as sol-gel, chemical vapor deposition, sputtering, pulsed laser deposition, electrodeposition and the spray pyrolysis method for deposition of the TiO2 thin films. However, low temperature and atmospheric pressure plasma (APP) processing is a cost effective and versatile coating technique, which is tremendously attracting the interest of many research groups. However, it needs post heat treatment to achieve crystalline thin films. Recently, the atmospheric pressure plasma jet (APPJ) technique became more popular to obtain crystalline thin films in a single step without post-annealing. A plasma is generated by applying energy to a gas in order to rearrange the electronic structure of the atoms and molecules and to generate the excited species and ions. The generated energy can be thermal or carried by either an electric current or electromagnetic radiations. The APPJ is a small RF plasma torch that works at low power. It consists of two concentric electrodes through which a mixture of helium, oxygen, and other gases flow. By applying RF power between 1 to 100 MHz to the inner electrode at a voltage between 100–250 V, the gas discharge is ignited. The ionized gas from the plasma jet exits through a nozzle, where it is directed onto a substrate a few millimeters downstream. The low injected power enables the torch to produce a stable discharge and avoids the arc transition. The APP and APPJ techniques are thoroughly discussed in the recent review article [5,6,7].
In recent times, Dong and co-researchers achieved adherent, optically transparent and fine electrically conductive TiNx/TiO2 hybrid films on polycarbonate and silicon substrates using the APP technique [8]. The plasma processing conditions, i.e., power and gas compositions, were precisely varied to achieve the hybrid thin films with 83% transmittance in the visible region, almost 8.5 J/m2 of film adhesion energy towards the polycarbonate and electrical resistivity of ~8.5 × 101 ohm cm before annealing. Post annealing and submission of H2 gas into the primary plasma have successfully reduced the electrical resistivity of the hybrid film to ~6.1 × 10−1 ohm cm. Chou et al. have used the APP method to deposit anatase TiO2 nano-dendrites (TNDs) films on FTO substrates for dye sensitized solar cell (DSSC) applications [9]. Due to high specific surface area of the ordered TNDs films, a photo-conversion efficiency of 12.08 % was achieved under simulated AM 1.5 solar light irradiation. Gazal et al. have studied the growth of TiO2 films on silicon substrates using the APP-ECVD-microwave torch method [10]. A TiO2 film got deposited in circular shape as a consequence of plasma jet geometry, which affects the film thickness as well as crystallinity from the center to periphery. The crystalline anatase phase TiO2 with precise columnar growth was observed at the center, whereas crystallinity eventually vanished at the periphery with cauliflower-like amorphous structure. The improved photocatalytic properties of TiO2 can be observed by synthesizing the TiO2 material with both anatase and rutile phases, which assist in diminishing the possibility of electron-hole recombination [11]. Recently, Wang et al. synthesized the mixed phase anatase-rutile TiO2 nanopowders using the APPJ technique [12]. Chang et al. synthesized dye-sensitized solar cells (DSSCs) with TiO2 photoanodes sintered by N2 and air atmospheric pressure plasma jets (APPJs) with/without air-quenching [13].
In the present research work, the APPJ technique was used to deposit TiO2 thin films onto the glass and silicon substrates. This technique enables thin film deposition on large or any type of substrate without the need for vacuum or inert deposition environments. Also, this technique is useful for obtaining crystalline thin films in a single step without post-annealing or post-deposition treatments. Helium gas and argon (Ar) gas are used as the carrier gas and H2O vapor is also introduced. The fixed flow rate was maintained for Helium gas and H2O vapor 5.0 and 20 mL/min, respectively. By varying the Argon gas flow rate 50 and 300 mL/min for the depositions TiO2 thin films. The deposited TiO2 thin films are characterized using the different characterization tools like, XRD, XPS, Raman Spectroscopy, FE-SEM, contact angle, and UV-Vis spectra in diffused reflection mode.

2. Results and Discussion

An X-ray diffraction study was used to confirm the crystal structure of the prepared TiO2 thin films. Figure 1 shows the XRD patterns of TiO2 thin films deposited at fixed flow rate of He = 5 L/min as a carrier gas, and Ar flow rate of 50 and 300 mL/min and recorded by varying the diffraction angle (2θ) from 20° to 80°.
All XRD patterns shows the peak (111) arises from the Si substrate. For TiO2 samples deposited at He = 5 L/min and 50 mL/min no diffraction peaks TiO2 were detected, which suggests an amorphous nature or that the nanocrystals are smaller than the detection limit of the XRD set-up [14,15]. For TiO2 sample deposited at Ar = 300 mL/min shows the diffraction peak at 2θ = 36.14°, corresponding to the (202) plane, which is in well agreement with JCPDS card no. 46-1237 and confirms the bronze-phase of TiO2 [16].
Symmetry and modes of vibration present in the TiO2 structure is studied with the help of Raman spectroscopy. Figure 2A,B shows the Raman spectra of TiO2 thin films with and without H2O vapor flow.
All TiO2 spectra show the strong peak at 520 cm−1, which is associated with the Si substrate [17]. The Raman peak found at 418 cm−1 corresponds to the Ag vibrational mode found in the sample prepared using H2O vapor and Ar flow rate of 50 and 300 mL/min (Figure 2A) [18]. While, in the case of He gas, the Raman peak at 418 cm−1 is not observed, indicating the amorphous nature of the TiO2. The degree of crystallinity is higher at a low gas flow rate of Ar (50 mL/min). Figure 2B shows without H2O vapor all samples exhibit the amorphous nature of TiO2, while the sample prepared at gas flow rate of Ar (50 mL/min) shows a peak at 418 cm−1.
For crystallized films (i.e., synthesized with H2O vaper), an XPS study was used to analyze chemical composition and surface state of the elements present in the TiO2 sample. Figure 3A displays the survey scan spectrum of the TiO2 sample prepared at a gas flow rate of He = 5L/min + Ar = 300 mL/min. It shows the signals of C1s, O1s, Ti2p, N1s elements and Si substrate peak present in TiO2. No other elements or impurity peaks were found in the prepared sample. Figure 3B shows the Ti 2p spectrum. Due to the spin orbit coupling the Ti 2p spectrum split into two stronger and two weaker peaks at binding energy 464.4, 458.5 eV and 462 and 456.5 eV, respectively.
The peaks at binding energy 464.4 and 458.5 eV associated with the Ti2p1/2 and Ti2p3/2 components, respectively [19]. The binding energy difference between the two peaks is 5.9 eV, which is in good agreement with reported values attributed to Ti4+ confirming the formation of the TiO2 compound [20]. The peaks were at binding energy 458.5 and 462.0 eV associated with Ti-N bonding. The observed higher O1s concentration and N1s species is considered to be due to the atmospheric processes. Elemental analysis using XPS for a synthesized TiO2 thin film with its atomic percentage is given in Table 1. Since the synthesis is carried out in an atmosphere, as expected, the Nitrogen functionality with small atomic percentage is observed. The atomic percentage for Ti2p varies from 37.18% to 43.91% and for O1s it is 38.10%, to 42.41%, respectively. Table 2 shows the ratio of Ti-N and Ti-O at 5 mm distance between substrate and plasma plume.
The film thickness was measured using a cross section of the films for different flow rates of Ar, He gases and water vapor. Figure 4A–C shows the cross section view of TiO2 thin films using SEM. For the He and H2O flow rate (He = 5L/min + H2O = 20 mL/min) TiO2 film thickness was observed to be 4 µm. While in the case of Ar, the gas flow rate increased from 50 to 300 mL/min, and it is observed that film thickness increased from 2 to 3 µm. This increase in film thickness attributed to the Ar gas flow can enhance the energetic ion bombardment, which promotes the mobility of atoms, and also, when there is a further increase in Ar gas flow, the ion density becomes large enough for crystal growth; hence, the grain size [21].
Figure 5A–F shows the SEM images of TiO2 thin films prepared at different deposition conditions. For the TiO2 sample (Figure 5A) prepared at He = 5L/min without the water vapor condition, clubbed shape grains are observed, then fairly spherical shaped after water vapor was introduced (Figure 5B) and particle size reduced. Reduction of particle size was observed after the addition of water vapor in all cases (Figure 5D,F).
In the case (Figure 5 C) of Ar gas of 30 mL/min and without water vapor composition, TiO2 nanoparticles are agglomerated and with water vapor particles, the size reduced. While for the Ar gas of flow rate 300 mL/min and without water vapor, nearly round shaped TiO2 grains are observed (Figure 5E) and after introducing water vapor, grain size is reduced (Figure 5F). For different gases (He and Ar), morphology is different. Also, for the same gas (Ar) but different flow rate (50 and 300 mL/min), different morphology is observed, which suggests that gas flow rate and mixture of different gases affects the morphology [22,23].
Figure 6 shows the UV-vis spectra in diffused reflection mode for TiO2 thin films deposited at fixed flow rate of He = 5 L/min as a carrier gas, and Ar flow rate of 50 and 300 mL/min. All the deposited TiO2 thin films exhibit the strong absorbance in the UV region whereas high transmittance in the visible range. The band gap of the deposited TiO2 thin films varies from 3.87 to 4.07 eV. The band edge is shifted towards the higher energy side indicating a red shift. In the case of Ar, the gas flow rate increased and the reflectance decreased, which suggests that the flow rate of the gas affects the optical properties of TiO2 thin films [24].
Figure 7 presents the photocatalytic decomposition study of Oleic acid using bare glass and TiO2 thin films. The photocatalytic activity of the bare glass and TiO2 thin films (He/Ar: 3L/min / 300 mL/min and water vapor) prepared by the APPJ technique was tested for the photocatalytic decomposition of oleic acid under UV light illumination. For glass substrate coated with oleic acid, a nearly constant (about 43°) water contact angle was observed up to 25 h under UV light illumination. No major changes are observed in the water contact angle after 25 h illumination, which proves that no photocatalytic decomposition of oleic acid takes place on the surface of glass substrate under ultraviolet light illumination in the absence of TiO2 coating.
Similar experiments were performed using TiO2 thin films. Before UV irradiation, the water contact angle was 80.22°, which shows that oleic acid was successfully coated on the TiO2 surface. The oleic acid treated TiO2 photocatalyst was exposed to UV light and, as a result, the water contact angle decreased slowly from 80.22° to 27.2° during 25 h of irradiation time. The decrease in water contact angle after UV irradiation confirms the photocatalytic decomposition of oleic acid using the TiO2 photocatalyst. These results suggest that both light and catalyst is required for the effective photocatalytic decomposition of organic compounds like oleic acid [25,26].

3. Experimental

3.1. Materials

Titanium ethoxide (Ti4(OCH2CH3)16) was purchased from Sigma-Aldrich. Oleic acid (C18H34O2), and n-heptane (C7H16) were purchased from the Fujifilm Wako Pure Chemical Corporation (Osaka, Japan). These chemicals were used as obtained.

3.2. Synthesis of TiO2 Thin Film

The APPJ experimental set up was developed and presented through the schematic diagram in Figure 8. The jet was made up of a Pyrex glass tube with a 4 mm diameter, which is capacitively coupled to the pulse generator by using copper electrode (research grade (99.99%)). Helium gas and argon (Ar) gas were used as carrier gas, whereas Titanium ethoxide was used as a precursor source, which was kept at a temperature of 125 °C. H2O vapor was also introduced. The plasma was sustained using an AC pulsed power supply at 15 KHz frequency having 1 µs pulsed rate by keeping discharge voltage constant at 4 kV corresponding to 160 Watts. Optimization of atmospheric pressure plasma deposition parameters for TiO2 thin film were carried out with fixed flow rate of helium gas = 5.0 L/min, and plasma was sustained in the jet, which shows violet plasma. By keeping the precursor (titanium ethoxide MW 228.11 gm/mol) flow rate constant = 500 mL/min and varying the argon gas flow rate from 50 to 300 mL/min, depositions of TiO2 were carried out with and without H2O vapor. The flow rate for H2O vapor was maintained at 20 mL/min. The deposition time was kept constant at 5 min for all the cases. The substrate to plasma jet distance was basically kept constant at 5 mm, throughout the experiments where the plume was uniformly distributed with high plasma density.

3.3. Photocatalytic Activity Test

The photocatalytic activity test of APPJ synthesized TiO2 thin films and bare glass were studied by performing photocatalytic decomposition of oleic acid. In order to remove the residual organic components present on the surface of TiO2 thin film, it was pre-treated before the photocatalytic experiment using UV light with a wavelength of 365 nm at 2.0 mW cm−2 for a single day. After one day of UV illumination, the TiO2 thin film was immersed in 0.5 vol% oleic acid solution diluted with n-heptane, and then withdrawn using a dip-coater at 10 mm s−1. To dry the oleic acid layer, samples were heated at 70 °C for 15 min with a hot plate. The samples were illuminated with 365 nm UV light at 1.0 mW cm−2 to induce photocatalytic decomposition of oleic acid. During the test, samples were taken at different time intervals and the water-contact angle measurement was carried out using a contact angle meter (DM-501, Kyowa Interface Science Co., Ltd.) [26].

3.4. Characterization Technique Details

Structural, morphological, and optical characterization, of the deposited TiO2 thin film was carried out by using XRD, XPS, Raman, FE-SEM and UV-Vis techniques. X-ray diffraction (XRD) analyses were performed on a Bruker D8 Discover diffractometer with a CuKα radiation source in detector mode at 1.0 degree grazing incidence. XPS analysis of the film was carried out using an AXIS Nova X-ray photoelectron spectrometer. The system was equipped with AlKα radiation source (1486.6 eV), which was operated at 150 W, having a spectral resolution of 0.1 eV. The system was operated at the base pressure of 1.0 × 10−9 Pa. The binding energy (BE) values are charge-corrected to C 1s = 284.6 eV as an internal standard, and the pass energy of 40 eV was used. Raman spectra were collected on a NRS-510, JASCO Raman microscope, using an excitation wavelength of 532 nm. Morphologies of as-deposited thin film samples were observed using field emission scanning electron microscopy (FE-SEM, JSM 7600F), operated at 20 kV, with 1.0 nm resolution. The film thickness was measured by using an SEM cross section, using a lapping technique. UV-Vis spectra in diffused reflection mode were recorded using a JASCO V-670 spectrophotometer at 2 nm resolution within the range 200–800 nm.

4. Conclusions

The TiO2 thin films can successfully be deposited onto glass and Si substrates using the APPJ technique. The XRD, Raman and XPS study confirms the formation of TiO2 by the APPJ technique. This technique is useful for thin film deposition of any type of substrate without the need for vacuum or inert deposition environments. Also, it is useful to obtain crystalline thin films in a single step without post-annealing or post-deposition treatments. Also, the effect of gases like He, Ar, water vapor and gas flow rate on the structural, morphological, optical properties of TiO2 thin films was studied. Different gas composition and flow rate affect the structural, morphology and optical properties. From the SEM study, it is observed that particle size of TiO2 decreased after introducing water vapor during synthesis. Also, the SEM cross sectional study shows that an increase in Ar gas flow rate increased film thickness. The oleic acid photocatalytic decomposition experiment shows that the water contact angle decreased, which confirms the photocatalytic decomposition of oleic acid under ultraviolet light illumination.

Author Contributions

S.G.: Data curation, Formal analysis, Investigation, Writing-original draft. R.T.: Investigation, Experimental, Analysis, Data curation. N.R.: Formal analysis, Investigation. S.S.L.: Formal analysis, Investigation. Y.M.H.: Writing-original draft. N.S.: Investigation, Methodology. T.K.: Formal analysis, Investigation. M.Y.: Data curation, Formal analysis, Investigation. K.T.: Methodology. A.F.: Conceptualization, Project administration, Supervision. C.T.: Project administration, Supervision, Conceptualization, Methodology. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly supported by MEXT Promotion of Distinctive Joint Research Center Program Grant Number JPMXP0618217662.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  2. Fan, Q.; McQuillin, B.; Ray, A.K.; Turner, M.L.; Seddon, A.B. High density, non-porous anatase titania thin films for device applications. J. Phys. D Appl. Phys. 2000, 3, 2683–2686. [Google Scholar] [CrossRef]
  3. Hashimoto, K.; Irie, H.; Fujishima, A. TiO2 photocatalysis: A historical overview and future prospects. Jpn. J. Appl. Phys. 2005, 44, 8269. [Google Scholar] [CrossRef]
  4. Hunge, Y.M.; Yadav, A.A.; Khan, S.; Takagi, K.; Suzuki, N.; Teshima, K.; Terashima, C.; Fujishima, A. Photocatalytic degradation of bisphenol A using titanium dioxide@nanodiamond composites under UV light illumination. J. Colloid Interface Sci. 2021, 582, 1058–1066. [Google Scholar] [CrossRef] [PubMed]
  5. Schutze, A.; Jeong, J.Y.; Babayan, S.E.; Park, J.; Selwyn, G.S.; Hicks, R.F. The atmospheric-pressure plasma jet: A review and comparison to other plasma sources. IEEE Trans. Plasma Sci. 1998, 26, 1685–1693. [Google Scholar] [CrossRef] [Green Version]
  6. Mariotti, D.; Belmonte, T.; Benedikt, J.; Velusamy, T.; Jain, G.; Švrček, V. Low-Temperature Atmospheric Pressure Plasma Processes for “Green” Third Generation Photovoltaics. Plasma Process. Polym. 2016, 13, 70–90. [Google Scholar] [CrossRef] [Green Version]
  7. Tendero, C.; Tixier, C.; Tristant, P.; Desmaison, J.; Leprince, P. Atmospheric pressure plasmas: A review. Spectrochim. Acta B 2006, 61, 2–30. [Google Scholar] [CrossRef]
  8. Dong, S.; Watanabe, M.; Dauskardt, R.H. Conductive Transparent TiNx/TiO2 Hybrid Films Deposited on Plastics in Air Using Atmospheric Plasma Processing. Adv. Funct. Mater. 2014, 24, 3075–3081. [Google Scholar] [CrossRef]
  9. Chou, W.C.; Liu, W.J. Study of dye sensitized solar cell application of TiO2 films by atmospheric pressure plasma deposition method. In Proceedings of the International Conference on Electronics Packaging (ICEP), Sapporo, Japan, 20–22 April 2016. [Google Scholar]
  10. Gazal, Y.; Dublanche-Tixier, C.; Chazelas, C.; Colas, M.; Carles, P.; Tristant, P. Multi-structural TiO2 film synthesized by an atmospheric pressure plasma-enhanced chemical vapor deposition microwave torch. Thin Solid Films 2016, 600, 43–52. [Google Scholar] [CrossRef]
  11. Liu, G.; Yan, X.; Chen, Z.; Wang, X.; Wang, L.; Lu, G.Q.; Cheng, H.M. Synthesis of rutile-anatase core-shell structured TiO2 for photocatalysis. J. Mater. Chem. 2009, 19, 6590–6596. [Google Scholar] [CrossRef]
  12. Wang, Y.; Yuan, Q.; Yin, G.; Zhang, Y.; Zhang, Y.; Li, Y.; Li, J.; Wang, T.; Ma, S. Synthesis of Mixed-Phase TiO2 Nanopowders Using Atmospheric Pressure Plasma Jet Driven by Dual-Frequency Power Sources. Plasma Chem. Plasma Process. 2016, 36, 1471–1484. [Google Scholar] [CrossRef]
  13. Chang, H.; Hsu, C.M.; Kao, P.K.; Yang, Y.J.; Hsu, C.C.; Cheng, I.C.; Chen, J.Z. Dye-sensitized solar cells with nanoporous TiO2 photoanodes sintered by N2 and air atmospheric pressure plasma jets with/without air-quenching. J. Power Source 2014, 251, 215–221. [Google Scholar] [CrossRef]
  14. Suzuki, N.; Terashima, C.; Nakata, K.; Katsumata, K.; Fujishima, A. Synthesis of a mesoporous titania thin film with a pseudo-single-crystal framework by liquid-phase epitaxial growth, and enhancement of photocatalytic activity. RSC Adv. 2020, 10, 40658–40662. [Google Scholar] [CrossRef]
  15. Yasuda, H.; Hirotsu, T. Critical Evaluation of Conditions of Plasma Polymerization. J. Polym. Sci. Polym. Chem. Ed. 1978, 16, 743–759. [Google Scholar] [CrossRef]
  16. Fakhouri, H.; Salem, D.B.; Carton, O.; Pulpytel, J.; Khonsari, F.A. Highly efficient photocatalytic TiO2 coatings deposited by open air atmospheric pressure plasma jet with aerosolized TTIP precursor. J. Phys. D Appl. Phys. 2014, 47, 265301. [Google Scholar] [CrossRef]
  17. Zhou, Y.; Zhu, Q.; Tian, J.; Jiang, F. TiO2 Nanobelt@Co9S8 Composites as Promising Anode Materials for Lithium and Sodium Ion Batteries. Nanomaterials 2017, 7, 252. [Google Scholar] [CrossRef]
  18. Borowicz, P.; Latek, M.; Rzodkiewicz, W.; Łaszcz, A.; Czerwinski, A.; Ratajczak, J. Deep-ultraviolet Raman investigation of silicon oxide: Thin film on silicon substrate versus bulk material. Adv. Nat. Sci. Nanosci. Nanotechnol. 2012, 3, 045003. [Google Scholar] [CrossRef]
  19. Jokisaari, J.R.; Bayerl, D.; Zhang, K.; Xie, L.; Nie, Y.; Schlom, D.G.; Kioupakis, E.; Graham, G.W.; Pan, X. Polarization-Dependent Raman Spectroscopy of Epitaxial TiO2(B) Thin Films. Chem. Mater. 2015, 27, 7896–7902. [Google Scholar] [CrossRef]
  20. Hunge, Y.M.; Yadav, A.A.; Dhodamani, A.G.; Suzuki, N.; Terashima, C.; Fujishima, A.; Mathe, V.L. Enhanced photocatalytic performance of ultrasound treated GO/TiO2 composite for photocatalytic degradation of salicylic acid under sunlight illumination. Ultrason. Sonochem. 2020, 61, 104849. [Google Scholar] [CrossRef]
  21. Hernandez-Rodriguez, E.; Marquez-Herrera, A.; Zaleta-Alejandre, E.; Melendez-Lira, M.; Cruz, W.D.; Zapata-Torres, M. Effect of electrode type in the resistive switching behavior of TiO2 thin films. J. Phys. D Appl. Phys. 2013, 46, 045103. [Google Scholar] [CrossRef]
  22. Vijaya, G.; Singh, M.M.; Krupashankara, M.S.; Kulkarni, R.S. Effect of Argon Gas Flow Rate on the Optical and Mechanical Properties of Sputtered Tungsten Thin Film Coatings. IOP Conf. Ser. Mater. Sci. Eng. 2016, 149, 012075. [Google Scholar] [CrossRef] [Green Version]
  23. Zurek, J.; Michalik, M.; Schmitz, F.; Kern, T.U.; Singheiser, L.; Quadakkers, W.J. The Effect of Water-Vapor Content and Gas Flow Rate on the Oxidation Mechanism of a 10% Cr-Ferritic Steel in Ar-H2O Mixtures. Oxid. Met. 2005, 63, 401–422. [Google Scholar] [CrossRef]
  24. Rebrov, A.K.; Emelyanov, A.A.; Plotnikov, M.Y.; Timoshenko, N.I.; Terekhov, V.V.; Yudin, I.B. Effect of the gas mixture flow rate on the process of diamond synthesis from a high-velocity microwave plasma. J. Appl. Mech. Tech. Phys. 2020, 61, 819–827. [Google Scholar] [CrossRef]
  25. Jena, S.; Tokas, R.B.; Misal, J.S.; Rao, K.D.; Udupa, D.V.; Thakur, S.; Sahoo, N.K. Effect of O2/Ar gas flow ratio on the optical properties and mechanical stress of sputtered HfO2 thin films. Thin Solid Films 2015, 592, 135–142. [Google Scholar] [CrossRef]
  26. Adachi, T.; Latthe, S.S.; Gosavi, S.W.; Roy, N.; Suzuki, N.; Ikari, H.; Kato, K.; Katsumata, K.; Nakata, K.; Furudate, M.; et al. Photocatalytic, superhydrophilic, self-cleaning TiO2 coating on cheap, light-weight, flexible polycarbonate substrates. Appl. Surf. Sci. 2018, 458, 917–923. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of TiO2 thin films deposited using the atmospheric pressure plasma jet technique with H2O vapor at fixed flow rate (500 mL/min) of titanium ethoxide precursor and (a) He = 5 L/min (b) He = 5L/min + Ar = 5 0mL/min and (c) He = 5L/min + Ar = 300 mL/min.
Figure 1. X-ray diffraction patterns of TiO2 thin films deposited using the atmospheric pressure plasma jet technique with H2O vapor at fixed flow rate (500 mL/min) of titanium ethoxide precursor and (a) He = 5 L/min (b) He = 5L/min + Ar = 5 0mL/min and (c) He = 5L/min + Ar = 300 mL/min.
Catalysts 11 00091 g001
Figure 2. (A) Raman spectra of TiO2 thin films with H2O flow rate 20 mL/min, (B) without H2O flow rate.
Figure 2. (A) Raman spectra of TiO2 thin films with H2O flow rate 20 mL/min, (B) without H2O flow rate.
Catalysts 11 00091 g002
Figure 3. (A) Survey scan spectrum and (B) Ti2p spectrum of TiO2 thin films deposited using Atmospheric Pressure Plasma Jet Jet with H2O vapor at fixed flow rate (500 mL/min) of Titanium ethoxide precursor and He = 5L/min + Ar = 300 mL/min.
Figure 3. (A) Survey scan spectrum and (B) Ti2p spectrum of TiO2 thin films deposited using Atmospheric Pressure Plasma Jet Jet with H2O vapor at fixed flow rate (500 mL/min) of Titanium ethoxide precursor and He = 5L/min + Ar = 300 mL/min.
Catalysts 11 00091 g003
Figure 4. Cross sectional view of TiO2 thin films deposited at (A) He = 5L/min + H2O = 20 mL/min (B) He = 5L/min + Ar = 50 mL/min +H2O = 20mL/min and (C) He = 5L/min + Ar = 300 mL/min + H2O = 20mL/min; using atmospheric pressure plasma jet.
Figure 4. Cross sectional view of TiO2 thin films deposited at (A) He = 5L/min + H2O = 20 mL/min (B) He = 5L/min + Ar = 50 mL/min +H2O = 20mL/min and (C) He = 5L/min + Ar = 300 mL/min + H2O = 20mL/min; using atmospheric pressure plasma jet.
Catalysts 11 00091 g004
Figure 5. SEM images of TiO2 thin films (A) without and (B) with water vapor for He = 5 L/min, (C) without and (D) with water vapor for He = 5 L/min + Ar = 50 mL/min, and (E) without and (F) with water vapor for He = 5 L/min + Ar = 300 mL/min.
Figure 5. SEM images of TiO2 thin films (A) without and (B) with water vapor for He = 5 L/min, (C) without and (D) with water vapor for He = 5 L/min + Ar = 50 mL/min, and (E) without and (F) with water vapor for He = 5 L/min + Ar = 300 mL/min.
Catalysts 11 00091 g005
Figure 6. UV-vis diffused reflectance spectra for TiO2 thin films deposited using atmospheric pressure plasma jet at fixed flow rate (500 mL/min) of titanium ethoxide precursor and (a) He = 5L/min (b) He = 5L/min + Ar = 50mL/min and (c) He = 5L/min + Ar = 300 mL/min.
Figure 6. UV-vis diffused reflectance spectra for TiO2 thin films deposited using atmospheric pressure plasma jet at fixed flow rate (500 mL/min) of titanium ethoxide precursor and (a) He = 5L/min (b) He = 5L/min + Ar = 50mL/min and (c) He = 5L/min + Ar = 300 mL/min.
Catalysts 11 00091 g006
Figure 7. Photocatalytic decomposition of oleic acid using bare glass and TiO2 thin films.
Figure 7. Photocatalytic decomposition of oleic acid using bare glass and TiO2 thin films.
Catalysts 11 00091 g007
Figure 8. Schematic diagram of experimental setup used for the synthesis of TiO2 thin films using atmospheric pressure plasma jet.
Figure 8. Schematic diagram of experimental setup used for the synthesis of TiO2 thin films using atmospheric pressure plasma jet.
Catalysts 11 00091 g008
Table 1. Difference in atomic percentage of Ti, O, C and N elements at different deposition conditions. Distance between the plasma jet and a substrate are shown in the parenthesis.
Table 1. Difference in atomic percentage of Ti, O, C and N elements at different deposition conditions. Distance between the plasma jet and a substrate are shown in the parenthesis.
Mass (%)
Ti (2p)O (1s)C (1s)N (1s)
Ar 50 mL/min (5 mm)37.1842.4117.632.78
Ar 50 mL/min (1 mm)39.0244.4714.581.93
Ar 300 mL/min (5 mm)36.7243.2917.652.34
Ar 300 mL/min (1 mm)43.9138.1016.581.40
Table 2. Ratio of Ti-N and Ti-O at 5 mm distance between substrate and plasma plume.
Table 2. Ratio of Ti-N and Ti-O at 5 mm distance between substrate and plasma plume.
Mass (%)
Ti-N
(456 eV)
Ti-O
(459 eV)
Ti-N
(462 eV)
Ti-O
(464 eV)
Ar 50 mL/min (5 mm)22.9745.9216.3414.77
Ar 300 mL/min (5 mm)10.2957.305.8826.53
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gosavi, S.; Tabei, R.; Roy, N.; Latthe, S.S.; Hunge, Y.M.; Suzuki, N.; Kondo, T.; Yuasa, M.; Teshima, K.; Fujishima, A.; et al. Low Temperature Deposition of TiO2 Thin Films through Atmospheric Pressure Plasma Jet Processing. Catalysts 2021, 11, 91. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11010091

AMA Style

Gosavi S, Tabei R, Roy N, Latthe SS, Hunge YM, Suzuki N, Kondo T, Yuasa M, Teshima K, Fujishima A, et al. Low Temperature Deposition of TiO2 Thin Films through Atmospheric Pressure Plasma Jet Processing. Catalysts. 2021; 11(1):91. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11010091

Chicago/Turabian Style

Gosavi, Suresh, Rena Tabei, Nitish Roy, Sanjay S. Latthe, Yuvaraj M. Hunge, Norihiro Suzuki, Takeshi Kondo, Makoto Yuasa, Katsuya Teshima, Akira Fujishima, and et al. 2021. "Low Temperature Deposition of TiO2 Thin Films through Atmospheric Pressure Plasma Jet Processing" Catalysts 11, no. 1: 91. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11010091

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop