Next Article in Journal
Highly Integrated Elastic Island-Structured Printed Circuit Board with Controlled Young’s Modulus for Stretchable Electronics
Next Article in Special Issue
Direct Measurement of Temperature Diffusivity of Nanocellulose-Doped Biodegradable Composite Films
Previous Article in Journal
The Mechanical and Physical Properties of 3D-Printed Materials Composed of ABS-ZnO Nanocomposites and ABS-ZnO Microcomposites
Previous Article in Special Issue
Randomly Multiplexed Diffractive Lens and Axicon for Spatial and Spectral Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Wetting Properties of Transparent Anatase/Rutile Mixed Phase Glancing Angle Magnetron Sputtered Nano-TiO2 Films

Department of Chemical Engineering, University of Patras, GR-26504 Patras, Greece
*
Author to whom correspondence should be addressed.
Micromachines 2020, 11(6), 616; https://doi.org/10.3390/mi11060616
Submission received: 8 June 2020 / Revised: 21 June 2020 / Accepted: 22 June 2020 / Published: 25 June 2020
(This article belongs to the Special Issue Micro/Nano-surfaces: Fabrication and Applications)

Abstract

:
Transparent polycrystalline TiO2 thin films have been deposited on unheated glass substrates using RF reactive magnetron sputtering. Depositions were carried out at different glancing angles and with different total gas mixture pressures. The variation of these parameters affected the crystal phase composition and the surface morphology. Depending on the glancing angle and the pressure, rutile, mixed anatase/ rutile and pure anatase were deposited at low substrate temperature. Both hydrophilic and hydrophobic TiO2 were obtained, exhibiting fast photoconversion to superhydrophilic upon UV irradiation. The effect of the materials physicochemical properties on the wettability and rate of the UV induced superhydrophilicity is evaluated.

1. Introduction

Wetting properties of solid surfaces are of great importance both from a theoretical point of view and for industrial applications. During the last decades, there has been a great research interest focused on the control of the wetting state of real surfaces, towards the fabrication of thin films resembling biosurfaces which exhibit extreme or peculiar wetting characteristics. The well-established studies of Young, Wenzel, and Cassie and Baxter revealed that the wettability of a solid surface (hydrophobic or hydrophilic) is greatly influenced by the intrinsic surface energy and the surface morphology [1,2,3]. Thin films possessing special wetting characteristics may act as self-cleaning surfaces, anti-biofouling surfaces for biomedical applications, on microfluidic devices for micro-droplet manipulation and many more [4,5,6,7,8].
TiO2 thin films are considered perfect candidates for all the above mentioned applications, due to their chemical inertness, biocompatibility, photocatalytic properties, and high transparency. Moreover, TiO2 films can convert from relatively hydrophilic (50–70°) to superhydrophilic (CA < 10°) by exposure to UV radiation, while the surface gradually regains its original wettability in dark [9]. The prevailing mechanism of the UV-induced hydrophilicity of TiO2 suggests that the photogenerated electron-holes migrate to the TiO2 surface and induce surface oxygen vacancies which act as energetically favored sites for the dissociative adsorption of water [9,10] The intrinsic wettability in dark, and also the rate of the photoinduced superhydrophilic conversion (PSH) determine the extent of the applicability of the TiO2 surfaces. Rough nano/mesoporous surfaces exhibit enhanced hydrophilicity or hydrophobicity and provide large surface area required for the acceleration of the PSH. For this reason, the wetting properties of nanostructured TiO2 surfaces have been thoroughly investigated [11,12]. Superhydrophobic TiO2 coatings have been prepared by different methods which however usually involve wet chemistry, multistep complex processes, annealing at high temperature, or further physical/chemical surface modification [13,14,15,16,17,18,19]. On the other hand, the deposition of long term hydrophilic TiO2 surfaces in the absence of UV exposure, is a more challenging task. In this direction, TiO2/SiO2 composite films are fabricated, or the hydrophilicity of the TiO2 surface is extended by oxygen plasma post treatment [20,21,22,23].
RF magnetron sputtering (MS) is a versatile technique for the deposition of polycrystalline TiO2 thin films at low temperature and is also suitable for large scale depositions. At the same time, the glancing angle physical vapor deposition (GLAD) constitutes a common practice towards the fabrication of rough and porous thin films [24]. The combined GLAD magnetron sputtering deposition of TiO2 films has been widely studied in terms of the effect of the deposition parameters on the films microstructure. However, there are only a few studies focusing on the TiO2 films wettability and PSH conversion, related to the surface morphology and/or structural characteristics induced by the GLAD MS deposition [25,26,27]. In this work, we present the fabrication of TiO2 thin films by GLAD RF magnetron sputtering on unheated glass substrates. The effect of the glancing angle and the working pressure on the TiO2 films structure and phase composition, the surface morphological characteristics, the optical properties, and the wetting characteristics have been evaluated. Subsequently, the relation between the phase composition and roughness with the wettability of the TiO2 films in dark and the photoinduced superhydrophilicity are investigated.

2. Materials and Methods

2.1. Glancing Angle RF Magnetron Sputtering Depositions

TiO2 thin film depositions on unheated glass substrates were carried out using a custom RF magnetron sputtering system in the reactive mode. The chamber was evacuated to a base pressure of 5 × 10−7 Torr by a turbomolecular (Oerlikon SL300, Leybold GmbH, Cologne, Germany) and a dry scroll pump (Edwards XDS10, Edwards Vacuum Inc., NY, USA) in series. A Ti metal target (2 in diameter) attached on balanced magnetrons RF source was sputtered with high purity argon while high purity oxygen was used as the reactive gas. The oxygen content of the mixture was tuned to 50% or 32% in terms of partial pressure. Under these conditions, sputtering is carried out in the oxide mode of the target which favors TiO2 stoichiometry. The glass substrates were placed 5 cm above the Ti target at different glancing angles with respect to the target surface (0, 60, 75, and 87°) as illustrated in Figure 1. RF power of 300 W was delivered to the cathode by a RF generator (Cesar, Advanced Energy, Advanced Energy, Fort Collins, CO, USA). The sputtering parameters, as well as the samples nomenclature, are presented in Table 1.

2.2. TiO2 Thin Films Characterization

Contact profilometry (Dektak XT, Bruker, MA, USA) was used for the measurements of the film thickness. The TiO2 thin films structure was determined by micro-Raman spectroscopy (Horiba Jobin-Yvon LabRam HR800, Horiba, Kyoto, Japan) in the visible range. The surface morphology and surface roughness were obtained by atomic force microscopy (Dimension Fast Scan, Bruker, MA, USA). A Lambda 900 (Bruker) UV–vis–nIR spectrophotometer was used for recording the UV-vis transmittance spectra (300 nm–800 nm) of the deposited TiO2 thin films. TiO2 thin films wettability was characterized right after the deposition and during prolonged storage of the films in dark ambient conditions. The static contact angle of 2 μL deionized water drops was measured under room temperature and RH 55–60%, using a commercial contact angle goniometer (Kruss, DSA100, Kruss Scientific, Hamburg, Germany). The photoinduced conversion to superhydrophilicity was evaluated by discrete measurements of the sessile drop in time intervals (2–3 min) under exposure in UV radiation [23].

3. Results and Discussion

3.1. TiO2 Films Crystal Structure

Figure 2 depicts the acquired Raman spectra of the TiO2 samples deposited at different glancing angles (0, 60, 75 and 87°) at 2.4 mTorr, and at 75° glancing angle varying the total working pressure (1 mTorr and 4.2 mTorr). Rutile or anatase structure is identified by the Raman active phonon modes in the visible range at 447 cm−1g) and 610 cm−1 (A1g) and 399 cm−1 (B1g), 519 cm−1 (A1g/B1g) and 639 cm−1 (Eg) respectively [28,29]. Polycrystalline rutile, mixed anatase/rutile and anatase ΤiO2 films were obtained. Phase formation is significantly affected by the glancing angle of the substrate and the working pressure. The increase of glancing angle at constant pressure and the increase of the total working pressure at large glancing angle (75°) resulted in phase shift from rutile to anatase. The effect of glancing angle and pressure on the RF magnetron sputtered TiO2 films phase composition has been reported and thoroughly discussed in our recent study [26]. The content of different phases can be estimated by the integrated intensities of the rutile and anatase peaks using the Equations (1) and (2) [30]:
% R = I 447 + I 610 I 399 + I 447 + I 519 + I 610 + I 639
and
% A = 1 % R
The acquired Raman spectra were deconvoluted and fitted with Lorenzian functions and the calculated values are summarized in Table 2.

3.2. Optical Properties

Figure 3 shows the optical transmittance spectra in the UV-Visible range of the as deposited TiO2 thin films on glass substrates. The films were highly transparent in the visible range. The differences in the transparency are attributed to the thickness and the morphology of the films. Among the thinner G0, G60, and G87 (120–170 nm) films, the G87 exhibits the highest transmittance value in the visible range reaching 90%.
The optical band gap of the indirect transition was evaluated using the well-established Tauc’s relation [31]:
α h v = α ο ( h v E g ) n
where α is the absorption coefficient, α ο is a constant related to the probability of transition, h is Plank’s constant and v is the photon frequency. Depending on the exponent value, the optical band gap, Ε g , for the allowed direct ( n = 1 / 2 ) or indirect transitions ( n = 2 ) , can be estimated by plotting the ( α h v ) 1 n vs. h v , the so called Tauc plot. The absorption coefficient is calculated from the transmittance T and the thickness d as:
α = ln ( T ) d 1
The Tauc plots showed linear regions of both direct (Figure 4a) and indirect (Figure 4b) allowed transitions of the TiO2 films. The corresponding optical Eg values are estimated by the extrapolation of the linear region of the Tauc plots to the x axis.
The calculated band gap values are in the order of 3.1 to 3.3 eV (indirect) and 3.44 to 3.66 (Table 2). The band gap increases with the increase of glancing angle from 0° to 87° or the pressure from 1 mΤorr to 4.2 mTorr, as the phase composition also changes towards anatase. The optical indirect band gap of bulk rutile and anatase is 3.05 eV and 3.2 eV respectively, while larger band gaps are usually reported for the amorphous ΤiO2. The nanorutile G0 shows a first indirect transition at 2.95 eV and a second one at 3.25 eV. The first transition is assigned to the rutile crystals while the second one can be attributed to the amorphous phase developed at the early stages of the sample growth. The pure anatase G75B film also exhibits a band gap value very close to that of the bulk phase (3.18 eV). For mixed phase TiO2, it has been reported that the proper band alignment of the heterojunctions leads to a red shift of the effective band gap compared the pure bulk phases [32]. For example, band gap of 2.89 eV has been reported for 40% anatase TiO2 thin film, which is even smaller than that of pure bulk rutile [33]. However, the mixed phase TiO2 films of this study (G60, G75, G87) exhibit wider band gaps compared to the corresponding pure rutile (G0) or anatase (G75B). The enhancement of the band gap, observed in this work, can be attributed to the porosity induced at large glancing angle deposition [34,35].

3.3. Wettability and Surface Morphology

Figure 5 shows the water contact angles of the TiO2 films as a function of the glancing angle (Figure 5a) and the working pressure (Figure 5b). The increase of the glancing angle and the working pressure apparently affected the wettability of the TiO2 films. All the films were hydrophilic after 10 days of storage with apparent contact angles (APCA) ranging from 15° to 70°, depending on the deposition conditions. Depositions at high glancing angle or pressure resulted in the most hydrophilic surfaces. The films were then left under dark ambient conditions until the contact angle reached a constant value indicating the saturated state of the samples. After prolonged storage (>30 days) the films deposited at small glancing angle or pressure (G0, G60 and G75A) turned to hydrophobic with APCAs > 100°, while those deposited at large glancing angle or high pressure (G87, G75A and G75B) maintained high hydrophilicity with APCAs 40–45°.
During long term storage the surface energy is inevitably decreased, i.e., the Young contact angle of the TiO2 films is increased, either due to the accumulation of non-polar hydrocarbon adsorbates or due to the depletion of polar surface species. However, it is noteworthy that although the conditions and the period of storage were the same for all the films, the augmentation of the APCA, for some of the films was very large, with ΔCA up to 74°, while for others, ΔCA was only 20–25°. The wettability of rough surfaces depends on the Young contact angle, i.e., the contact angle of a corresponding flat surface, the surface roughness and also on the specific topography features of the surface. The implemented deposition parameters caused simultaneous variation of the crystal structure, roughness, and surface morphology of the GLAD TiO2 films. Therefore, in order to interpretate the different final wetting states, all the above mentioned factors which affect the final wettability are considered.
In Figure 6 the APCAs after 10 days of storage and the stabilized APCAs of the TiO2 films are plotted against the anatase content. The respective rms roughness values, as they were obtained by 5 μm × 5 μm AFM images, are also shown. The APCAs are well correlated with the content of the anatase phase of the films, regardless of the deposition conditions. It can be seen, that only TiO2 films consisted of >45% anatase were highly hydrophilic after 10 days and retained their hydrophilicity. The films that eventually turned from hydrophilic to hydrophobic are the rutile rich films (anatase content < 40%). The effect of the different crystal structure of the surface is more obvious if we compare TiO2 films with the same rms roughness (G87 vs. G75A). The contact angle relaxation on polycrystalline mixed or anatase TiO2 sputtered films of same roughness in atmospheric conditions was examined by Lee and Park [36] using quantitative XPS studies. It was found that, compared to the films containing rutile, the anatase films retained high hydrophilicity due to the formation of strong donor-acceptor complexes by the interaction of Ti-OH basic groups with H2O [36]. Τhis implies that the anatase rich surface can be persistent hydrophilic and inherently more hydrophilic (smaller Young contact angle) compared to the rutile films. However, the large difference of the final APCAs between the rutile rich and anatase rich films (ΔCA > 60°) indicates that the difference of the final wetting states should be further investigated by considering the role of roughness and surface morphology.
The increase of the glancing angle and pressure resulted not only in the preferential growth of anatase over rutile, but also in a remarkable increase of the TiO2 films roughness. Figure 6 shows that as the crystal phase develops towards anatase, the rms roughness is increased from 2 nm to 18.5 nm. However, neither the roughness magnitude alone is an adequate measure for the prediction of the final wetting state of all the TiO2 films. For example, the G87 anatase and G75A rutile films were respectively hydrophilic and hydrophobic, despite the fact that the rms roughness value of both was ~8.5 nm, indicating that indeed the crystal phase may have significant impact on the final wetting state.
The effect of roughness on wettability is usually described by the Wenzel and Cassie-Baxter models [2,3]. In the Wenzel model,   R , which is defined as the ratio of the real surface area to the projected (smooth) area, induces a deviation of the apparent contact angle from the Young contact angle as:
  cos θ A = R   cos θ Υ .
The Wenzel equation implies that surface wettability, either hydrophilicity or hydrophobicity, is amplified by surface roughness. The Cassie–Baxter model describes the effect of roughness on the apparent contact angle for superhydrophobic surfaces exhibiting low solid-liquid adhesion and low contact angle hysteresis. The latter will not be considered since the water droplet showed strong adhesion on the TiO2 films surface and none of the films reached the superhydrophobic state. In Figure 7, the cosine of the apparent contact angle is plotted as a function of roughness for both the rutile and anatase rich films. The rutile rich films follow the Wenzel model. After 10 days of storage, the θY of rutile is probably <90° and the increase of roughness induces higher hydrophilicity (cosθA is increased). A smaller negative slope of cosθA is also observed for the hydrophobic stabilized state of the rutile films which indicates that after long term storage, the Young contact angle just exceeds 90°. Hence, the transition of rutile TiO2 surface from hydrophilic to hydrophobic ΔCA is greater as the surface roughness increases.
On the other hand, the highly hydrophilic character of the anatase rich films 10 days after the deposition (APCAs 20°) and after its stabilization (APCAs 40°) cannot be expressed by the Wenzel model. In Figure 7b, cosθA is shown to be independent of the rms roughness. It is possible that the low apparent contact angles originate from the minimum possible free energy equilibrium of the droplet on the surface among the different wetting states, as described by the Cassie impregnating model (CI) [37,38]:
cos θ A = 1 f s + f s   cos θ Υ
In this case, the surface equilibrium is reached by the lateral propagation of the liquid front beyond the droplet, through the surface microstructure. Water penetrates the surface pores and fs is the fraction of the solid islands of the mixed liquid-solid surface that remain dry. Equation (6) shows that as long as the surface wettability is expressed by the CI model, the APCA is independent of the roughness. The criterion for this wetting mechanism is [37,38]:
cos θ Y > 1 f s R f s
or according to the thermodynamic analysis of Huang et al. [39]:
cos θ Y > d 4 h + d
where d and h are the mean diameter and height of surface pores. Therefore, the thermodynamic equilibrium can be shifted from the Wenzel to the CI state, due to certain texturization and surface porosity.
In Figure 8, we present the AFM 3d images and the cross-section line profiles of two hydrophobic rutile films (A,C) and two hydrophilic anatase films (B,D). Both the hydrophilic anatase films consist of nanorods forming quasi-periodical deep surface pore channels. By the line profiles, the approximate mean height and diameter of the pores for the G87 and G75B anatase films are h = 22 ± 6 nm, d = 150 ± 30 nm, and h = 55 ± 15 nm, d = 140 ± 30 nm respectively. Hence, according to Equation (8), the films are in the CI wetting state if the anatase θΥ < 65–50°, which is very probable according to ref. [36] as it was previously discussed. The nanorutile films are characterized by irregular needle-like asperities, interrupted by larger structures forming shallow cavities (G60), or mountain-like structures with on-top nanocones (G75A), but no deep pore channels. It is worth noticing that, under the same sputtering conditions, the increase of glancing angle from 60° to 87° (Figure 7A,B), or the increase of pressure from 1 mTorr to 4.2 mTorr (Figure 7C,D), induces prominent changes on the surface texturization, which are related to the phase change from rutile to anatase [26].
Subsequently, the wetting state of the TiO2 films derives from the synergistic effect of the crystal phase and surface topography. The anatase films probably retain low Young CA and the fibrous structure with deep surface pores facilitates the establishment of the CI wetting state, permitting persistent high hydrophilicity of the surface. In the work of Chaterjee et al. [25], RF sputtered nanostructured TiO2 with amorphous structure and shallow pores (<10 nm) exhibited hydrophobicity, despite being deposited at high glancing angle. Therefore, both the anatase structure, as well as the mean pore height, seems to be determining factors for the persistent hydrophilicity of the GLAD TiO2 films. On the other hand, the Wenzel state applies for the rutile surfaces with simpler topography. Due to the greater tendency of Young CA to increase during storage, the rutile surfaces exhibit a transition from hydrophilic to hydrophobic, and the ΔCA of this transition depends on the surface roughness.

3.4. UV-Induced Conversion to Superhydrophilic TiO2

Figure 9a shows the variation of the water contact angle on the TiO2 films as a function of the exposure time to UV radiation. All the films convert to superhydrophilic within some minutes, regardless of the initial hydrophobic or hydrophilic state of the surface, or the crystal phase composition. Interestingly, complete wetting (CA = 0°) occurs within 10–20 min even for the most hydrophobic G75A TiO2 film. However, the apparent contact angle of the mixed phase, anatase rich hydrophilic films exponentially decreases with illumination time, while the initially hydrophobic rutile rich films exhibit quasi zero order dependence on the UV exposure time. Therefore, within a very short UV illumination time, the initially hydrophilic high anatase content films reach a critical contact angle, approaching the superhydrophilic condition (CA ≈ 10°), while the corresponding CA of the enhanced rutile hydrophobic films is still substantially large. The different photo-conversion rate order indicates that the high content anatase surfaces become almost saturated by OH groups significantly faster compared to the hydrophobic rutile rich films. Using photoluminescence spectroscopy, Baiju et al. [40] reported that mixed phase anatase-rich TiO2 exhibits longer photo-induced carrier lifetime compared to rutile rich mixed phase. Better photoactivity of anatase rich (60–80% A) TiO2 materials, compared to pure anatase or rutile rich films has been also reported by several studies and it is attributed to the effective charge separation and transport occurring at the anatase–rutile interface [40,41,42].
Our results are in accordance with those observations. The film deposited at the largest glancing angle G87 (76% anatase) approached the superhydrophilic state within 2 min and required in total 10 min for complete wetting (CA ≈ 0°). The pure anatase G75B and the G75 TiO2 film (<50% anatase), converted to superhydrophilic 5 and 10 min slower than the G87 film (76% anatase), despite that their initial contact angle was also ≈45° and that they both exhibited higher surface roughness. Therefore, faster photoconversion of the G87 film is attributed to the crystal phase composition, obtained by the deposition at the largest glancing angle and at Pt = 2.4 mTorr. Among the rutile rich TiO2 films, the most hydrophobic G75A (13% anatase) and G60 (35% anatase) reach the superhydrophilicity (CA ≈ 10°) in about 8 min, whereas the pure nanorutile TiO2 deposited at 0° requires double UV illumination time. Moreover, the G75A deposited at large glancing angle and low pressure, achieved complete wetting very fast, in 11 min. In this case the photoconversion rate of the hydrophobic rutile rich films seems to be better correlated with the films surface roughness rather than the crystal phase composition. These observations are better illustrated in Figure 9c,d where the UV irradiation time required for the TiO2 films to attain complete wetting along with the rms roughness are plotted versus the anatase content. Due to the different initial wetting state of the rutile or anatase rich films, two separate graphs are presented.
After the UV-induced superhydrophilic conversion, all the films were stored in dark and ambient conditions. The recovery of the surface was recorded by contact angle measurements for 30 days (Figure 9b). The TiO2 surfaces are gradually reconstructed back to their initial state due to the substitution of the metastable hydroxyl groups by ambient oxygen. Full recovery of the surface is only attained after 25 days for most of the films. Nevertheless, the rate of contact angle relaxation in dark could be enhanced by the application of mechanochemical methods [43].

4. Conclusions

Polycrystalline TiO2 thin films were deposited using RF reactive magnetron sputtering in glancing angle configuration on unheated substrates. Depending on the glancing angle, or the working pressure the films exhibited different phase composition. The increase of glancing angle or pressure resulted in phase shift from rutile to anatase and in roughness enhancement. All films were transparent in the visible range and the indirect band gaps were estimated from 2.95 tο 3.3 eV. By altering the glancing angle or pressure the realization of either hydrophobic or persistent hydrophilic TiO2 films was achieved. After long term storage in dark environment, the TiO2 films deposited at low glancing angle or pressure turned to hydrophobic, while those deposited at high glancing angle or pressure showed permanent hydrophilicity. It was revealed that the final wettability, as well as the surface roughness of the nanostructured TiO2 films is well correlated with the crystal phase composition. Permanent hydrophilicity is achieved for the films which are rich in anatase and exhibit high surface roughness, while the rutile rich films with lower mean roughness became hydrophobic. Regardless of the rms roughness value, the hydrophilic anatase films deposited at large glancing angle exhibited completely different surface topography features, compared to those of the rutile rich film. By considering fundamental wetting models, the mechanism of different wetting was identified. Permanent hydrophilicity of the anatase films was associated to the Cassie impregnating state, the establishment of which is facilitated by the deep surface pores formed between the anatase nanorods, while the simpler topography rutile rich films were found to follow the Wenzel model. Both the intrinsic wettability and the crystal phase formed under different glancing angles or pressure seem to be crucial for the final wetting state. Regardless of the initial wettability, all the glancing angle deposited TiO2 films showed fast photoinduced superhydrophilic conversion with exposure to UV irradiation for a few minutes.

Author Contributions

V.V., Conceptualization, Investigation, Writing—Original Draft, Visualization; E.F., Writing—Review & Editing; D.M., Supervision, Project Administration All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the Operational Program "Human Resources Development, Education and Lifelong Learning" and is co-financed by the European Union (European Social Fund) and Greek national funds.

Acknowledgments

The authors would like to gratefully thank Eleftherios Amanatides (Department of Chemical Engineering, University of Patras) for the valuable advice, scientific and technical support throughout the project. Angelos Kalampounias (Department of Chemistry, University of Ioannina) is gratefully acknowledged for performing the laser Raman measurements and Georgios Petrakis and Constantine Galiotis (Department of Chemical Engineering, University of Patras) for the AFM measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Young, T., III. An essay on the cohesion of fluids. Philos. Trans. R. Soc. Lond. 1805, 95, 65–87. [Google Scholar]
  2. Wenzel, R.N. Resistance of solid surfaces to wetting by water. Ind. Eng. Chem. 1936, 28, 988–994. [Google Scholar] [CrossRef]
  3. Cassie, A.B.D.; Baxter, S. Wettability of porous surfaces. Trans. Faraday Soc. 1944, 40, 546. [Google Scholar] [CrossRef]
  4. Liu, K.; Jiang, L. Bio-Inspired Self-Cleaning Surfaces. Available online: https://www.annualreviews.org/doi/abs/10.1146/annurev-matsci-070511-155046 (accessed on 19 January 2020).
  5. Xu, Q.; Zhang, W.; Dong, C.; Sreeprasad, T.S.; Xia, Z. Biomimetic self-cleaning surfaces: Synthesis, mechanism and applications. J. R. Soc. Interface 2016, 13, 20160300. [Google Scholar] [CrossRef]
  6. Su, B.; Wang, S.; Song, Y.; Jiang, L. Utilizing superhydrophilic materials to manipulate oil droplets arbitrarily in water. Soft Matter 2011, 7, 5144–5149. [Google Scholar] [CrossRef]
  7. Lin, X.; Yang, M.; Jeong, H.; Chang, M.; Hong, J. Durable superhydrophilic coatings formed for anti-biofouling and oil–water separation. J. Membr. Sci. 2016, 506, 22–30. [Google Scholar] [CrossRef]
  8. Gogolides, E.; Ellinas, K.; Tserepi, A. Hierarchical micro and nano structured, hydrophilic, superhydrophobic and superoleophobic surfaces incorporated in microfluidics, microarrays and lab on chip microsystems. Microelectron. Eng. 2015, 132, 135–155. [Google Scholar] [CrossRef]
  9. Wang, R.; Hashimoto, K.; Fujishima, A.; Chikuni, M.; Kojima, E.; Kitamura, A.; Shimohigoshi, M.; Watanabe, T. Light-induced amphiphilic surfaces. Nature 1997, 388, 431–432. [Google Scholar] [CrossRef]
  10. Takeuchi, M.; Sakamoto, K.; Martra, G.; Coluccia, S.; Anpo, M. Mechanism of Photoinduced Superhydrophilicity on the TiO2 Photocatalyst Surface. Available online: https://pubs.acs.org/doi/abs/10.1021/jp058075i (accessed on 19 January 2020).
  11. Lai, Y.; Huang, J.; Cui, Z.; Ge, M.; Zhang, K.-Q.; Chen, Z.; Chi, L. Recent Advances in TiO2-Based Nanostructured Surfaces with Controllable Wettability and Adhesion. Small 2016, 12, 2203–2224. [Google Scholar] [CrossRef]
  12. Borras, A.; González-Elipe, A.R. Wetting Properties of Polycrystalline TiO2 Surfaces: A Scaling Approach to the Roughness Factors. Available online: https://pubs.acs.org/doi/full/10.1021/la101975e (accessed on 4 February 2020).
  13. Wang, Y.; Li, B.; Liu, T.; Xu, C.; Ge, Z. Controllable fabrication of superhydrophobic TiO2 coating with improved transparency and thermostability. Colloids Surf. A Physicochem. Eng. Asp. 2014, 441, 298–305. [Google Scholar] [CrossRef]
  14. Sun, W.; Zhou, S.; Chen, P.; Peng, L. Reversible switching on superhydrophobic TiO2 nano-strawberry films fabricated at low temperature. Chem. Commun. 2008, 5, 603–605. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, X.; Jin, M.; Liu, Z.; Tryk, D.A.; Nishimoto, S.; Murakami, T.; Fujishima, A. Superhydrophobic TiO2 Surfaces: Preparation, Photocatalytic Wettability Conversion, and Superhydrophobic−Superhydrophilic Patterning. J. Phys. Chem. C 2007, 111, 14521–14529. [Google Scholar] [CrossRef]
  16. Holtzinger, C.; Niparte, B.; Wächter, S.; Berthomé, G.; Riassetto, D.; Langlet, M. Superhydrophobic TiO2 coatings formed through a non-fluorinated wet chemistry route. Surf. Sci. 2013, 617, 141–148. [Google Scholar] [CrossRef]
  17. Feng, X.; Zhai, J.; Jiang, L. The Fabrication and Switchable Superhydrophobicity of TiO2 Nanorod Films. Angew. Chem. Int. Ed. 2005, 44, 5115–5118. [Google Scholar] [CrossRef]
  18. Kobayashi, T.; Konishi, S. TiO2 patterns with wide photo-induced wettability change by a combination of reactive sputtering process and surface modification in a microfluidic channel. J. Micromech. Microeng. 2015, 25, 115014. [Google Scholar] [CrossRef]
  19. Lai, Y.; Tang, Y.; Gong, J.; Gong, D.; Chi, L.; Lin, C.; Chen, Z. Transparent superhydrophobic/superhydrophilic TiO2-based coatings for self-cleaning and anti-fogging. J. Mater. Chem. 2012, 22, 7420–7426. [Google Scholar] [CrossRef]
  20. Houmard, M.; Berthomé, G.; Joud, J.C.; Langlet, M. Enhanced cleanability of super-hydrophilic TiO2–SiO2 composite surfaces prepared via a sol-gel route. Surf. Sci. 2011, 605, 456–462. [Google Scholar] [CrossRef]
  21. Permpoon, S.; Houmard, M.; Riassetto, D.; Rapenne, L.; Berthomé, G.; Baroux, B.; Joud, J.C.; Langlet, M. Natural and persistent superhydrophilicity of SiO2/TiO2 and TiO2/SiO2 bi-layer films. Thin Solid Film. 2008, 516, 957–966. [Google Scholar] [CrossRef]
  22. Zhang, K.-X.; Wang, W.; Hou, J.-L.; Zhao, J.-H.; Zhang, Y.; Fang, Y.-C. Oxygen plasma induced hydrophilicity of TiO2 thin films. Vacuum 2011, 85, 990–993. [Google Scholar] [CrossRef]
  23. Vrakatseli, V.E.; Pagonis, E.; Amanatides, E.; Mataras, D. Photoinduced superhydrophilicity of amorphous TiOx-like thin films by a simple room temperature sol-gel deposition and atmospheric plasma jet treatment. J. Phys. Conf. Ser. 2014, 550, 012034. [Google Scholar] [CrossRef]
  24. Barranco, A.; Borras, A.; Gonzalez-Elipe, A.R.; Palmero, A. Perspectives on oblique angle deposition of thin films: From fundamentals to devices. Prog. Mater. Sci. 2016, 76, 59–153. [Google Scholar] [CrossRef] [Green Version]
  25. Chatterjee, S.; Kumar, M.; Gohil, S.; Som, T. An oblique angle radio frequency sputtering method to fabricate nanoporous hydrophobic TiO2 film. Thin Solid Film. 2014, 568, 81–86. [Google Scholar] [CrossRef]
  26. Vrakatseli, V.; Kalarakis, A.; Kalampounias, A.; Amanatides, E.; Mataras, D. Glancing Angle Deposition Effect on Structure and Light-Induced Wettability of RF-Sputtered TiO2 Thin Films. Micromachines 2018, 9, 389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Horprathum, M.; Eiamchai, P.; Kaewkhao, J.; Chananonnawathorn, C.; Patthanasettakul, V.; Limwichean, S.; Nuntawong, N.; Chindaudom, P. Fabrication of nanostructure by physical vapor deposition with glancing angle deposition technique and its applications. AIP Conf. Proc. 2014, 7, 1617. [Google Scholar]
  28. Porto, S.P.S.; Fleury, P.A.; Damen, T.C. Raman Spectra of TiO2, MgF2, ZnF2, FeF2 and MnF2. Phys. Rev. 1967, 154, 522. [Google Scholar] [CrossRef]
  29. Ohsaka, T.; Izumi, F.; Fujiki, Y. Raman spectrum of anatase, TiO2. J. Raman Spectrosc. 1978, 7, 321–324. [Google Scholar] [CrossRef]
  30. Zanatta, A.R. A fast-reliable methodology to estimate the concentration of rutile or anatase phases of TiO2. AIP Adv. 2017, 7, 075201. [Google Scholar] [CrossRef] [Green Version]
  31. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  32. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.J.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef]
  33. Paul, S.; Choudhury, A. Investigation of the optical property and photocatalytic activity of mixed phase nanocrystalline titania. Appl. Nanosci. 2014, 4, 839–847. [Google Scholar] [CrossRef] [Green Version]
  34. Vorobiev, Y.V.; Horley, P.P.; Hernández-Borja, J.; Esparza-Ponce, H.E.; Ramírez-Bon, R.; Vorobiev, P.; Pérez, C.; González-Hernández, J. The effects of porosity on optical properties of semiconductor chalcogenide films obtained by the chemical bath deposition. Nanoscale Res. Lett. 2012, 7, 483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lee, Y.-C.; Chang, Y.S.; Teoh, L.G.; Huang, Y.L.; Shen, Y.C. The effects of the nanostructure of mesoporous TiO2 on optical band gap energy. J. Sol. Gel. Sci. Technol. 2010, 56, 33–38. [Google Scholar] [CrossRef]
  36. Lee, M.-K.; Park, Y.-C. Contact Angle Relaxation and Long-Lasting Hydrophilicity of Sputtered Anatase TiO2 Thin Films by Novel Quantitative XPS Analysis. Langmuir 2019, 35, 2066–2077. [Google Scholar] [CrossRef] [PubMed]
  37. Bormashenko, E.; Pogreb, R.; Stein, T.; Whyman, G.; Erlich, M.; Musin, A.; Machavariani, V.; Aurbach, D. Characterization of rough surfaces with vibrated drops. Phys. Chem. Chem. Phys. 2008, 10, 4056. [Google Scholar] [CrossRef] [PubMed]
  38. Bico, J.; Thiele, U.; Quéré, D. Wetting of textured surfaces. Colloids Surf. A Physicochem. Eng. Asp. 2002, 206, 41–46. [Google Scholar] [CrossRef]
  39. Huang, W.; Lei, M.; Huang, H.; Chen, J.; Chen, H. Effect of polyethylene glycol on hydrophilic TiO2 films: Porosity-driven superhydrophilicity. Surf. Coat. Technol. 2010, 204, 3954–3961. [Google Scholar] [CrossRef]
  40. Baiju, K.V.; Zachariah, A.; Shukla, S.; Biju, S.; Reddy, M.L.P.; Warrier, K.G.K. Correlating Photoluminescence and Photocatalytic Activity of Mixed-phase Nanocrystalline Titania. Catal. Lett. 2009, 130, 130–136. [Google Scholar] [CrossRef]
  41. Su, R.; Bechstein, R.; Sø, L.; Vang, R.T.; Sillassen, M.; Esbjörnsson, B.; Palmqvist, A.; Besenbacher, F. How the Anatase-to-Rutile Ratio Influences the Photoreactivity of TiO2. J. Phys. Chem. C 2011, 115, 24287–24292. [Google Scholar] [CrossRef]
  42. Zhang, X.; Lin, Y.; He, D.; Zhang, J.; Fan, Z.; Xie, T. Interface junction at anatase/rutile in mixed-phase TiO2: Formation and photo-generated charge carriers properties. Chem. Phys. Lett. 2011, 504, 71–75. [Google Scholar] [CrossRef]
  43. Kamei, M.; Mitsuhashi, T. Hydrophobic drawings on hydrophilic surfaces of single crystalline titanium dioxide: Surface wettability control by mechanochemical treatment. Surf. Sci. 2000, 463, L609–L612. [Google Scholar] [CrossRef]
Figure 1. Substrate geometry with respect to target surface for glancing angle depositions.
Figure 1. Substrate geometry with respect to target surface for glancing angle depositions.
Micromachines 11 00616 g001
Figure 2. Raman spectra of the TiO2 films deposited at different glancing angles and sputtering conditions.
Figure 2. Raman spectra of the TiO2 films deposited at different glancing angles and sputtering conditions.
Micromachines 11 00616 g002
Figure 3. UV-visible optical transmission spectra of the TiO2 films on glass substrate.
Figure 3. UV-visible optical transmission spectra of the TiO2 films on glass substrate.
Micromachines 11 00616 g003
Figure 4. Tauc plots for the estimation of (a) indirect band gap and (b) direct band gap of the TiO2 films.
Figure 4. Tauc plots for the estimation of (a) indirect band gap and (b) direct band gap of the TiO2 films.
Micromachines 11 00616 g004
Figure 5. Measured apparent contact angles 10 days after the deposition and after prolonged storage in dark when the surface is stabilized versus (a) the glancing angle of the substrate (Pt =2.4 mTorr, O2 content:50%) and (b) the working pressure (glancing angle 75°, O2 content: 32%).
Figure 5. Measured apparent contact angles 10 days after the deposition and after prolonged storage in dark when the surface is stabilized versus (a) the glancing angle of the substrate (Pt =2.4 mTorr, O2 content:50%) and (b) the working pressure (glancing angle 75°, O2 content: 32%).
Micromachines 11 00616 g005
Figure 6. Apparent water contact angles of the GLAD TiO2 films, 10 days after the deposition and >30 days after deposition (stabilized), along with the respective RMS surface roughness (in nm), as a function of the TiO2 phase composition.
Figure 6. Apparent water contact angles of the GLAD TiO2 films, 10 days after the deposition and >30 days after deposition (stabilized), along with the respective RMS surface roughness (in nm), as a function of the TiO2 phase composition.
Micromachines 11 00616 g006
Figure 7. The cosine of the apparent contact angle of glancing angle deposited TiO2 films (a) rich in rutile (65–100%) and (b) rich in anatase (45–100%), as a function of the RMS roughness of the surface.
Figure 7. The cosine of the apparent contact angle of glancing angle deposited TiO2 films (a) rich in rutile (65–100%) and (b) rich in anatase (45–100%), as a function of the RMS roughness of the surface.
Micromachines 11 00616 g007
Figure 8. 3D views of the AFM images (5 μm × 5 μm) and line height profiles acquired for two hydrophobic and two hydrophilic TiO2 thins films. (A,C): G60 and G75A (hydrophobic rutile), (B,D): G87 and G75B (hydrophilic anatase)
Figure 8. 3D views of the AFM images (5 μm × 5 μm) and line height profiles acquired for two hydrophobic and two hydrophilic TiO2 thins films. (A,C): G60 and G75A (hydrophobic rutile), (B,D): G87 and G75B (hydrophilic anatase)
Micromachines 11 00616 g008
Figure 9. (a) Contact angle variation as a function of the elapsed time of exposure of the TiO2 films under UV radiation (8 mW/cm2) and (b) contact angle relaxation during storage of the films in dark ambient conditions after the UV exposure. (c,d) Required time of UV illumination in order the TiO2 films to achieve complete wetting and corresponding surface roughness versus the content in anatase. Note: separate graphs of the initially hydrophobic rutile rich TiO2 and the initially hydrophilic anatase rich TiO2 are plotted due to the difference in the extent of the superhydrophilic photoconversion. (c,d) are not comparable.
Figure 9. (a) Contact angle variation as a function of the elapsed time of exposure of the TiO2 films under UV radiation (8 mW/cm2) and (b) contact angle relaxation during storage of the films in dark ambient conditions after the UV exposure. (c,d) Required time of UV illumination in order the TiO2 films to achieve complete wetting and corresponding surface roughness versus the content in anatase. Note: separate graphs of the initially hydrophobic rutile rich TiO2 and the initially hydrophilic anatase rich TiO2 are plotted due to the difference in the extent of the superhydrophilic photoconversion. (c,d) are not comparable.
Micromachines 11 00616 g009
Table 1. Samples nomenclature and deposition parameters.
Table 1. Samples nomenclature and deposition parameters.
Sample CodingGlancing Angle (Degrees)RF Power (Watts)Working Pressure Pt (mTorr)Oxygen Content (%)
G0, G60, G75, G870, 60, 75, 873002.450%
G75A, G75B75300A:1, B:4.232%
Table 2. TiO2 thin films thickness, phase composition, % transmittance in the visible range on glass substrate and indirect/direct band gap.
Table 2. TiO2 thin films thickness, phase composition, % transmittance in the visible range on glass substrate and indirect/direct band gap.
Sample Thickness (nm)Anatase Content (%)Average Transmittance (Vis) %Direct Allowed Transition Eg (eV)Indirect Allowed Transition Eg (eV)
G0140 ± 35074.3 ± 10.43.532.95, 3.25
G60160 ± 2135 ± 573.8 ± 6.93.613.2
G75240 ± 1147 ± 378.5 ± 5.13.53.23
G87165 ± 2576 ± 682.2 ± 6.13.663.3
G75A246 ± 2913 ± 374.6 ± 5.03.443.13
G75Β310 ± 2298 ± 174.3 ± 8.83.533.18

Share and Cite

MDPI and ACS Style

Vrakatseli, V.; Farsari, E.; Mataras, D. Wetting Properties of Transparent Anatase/Rutile Mixed Phase Glancing Angle Magnetron Sputtered Nano-TiO2 Films. Micromachines 2020, 11, 616. https://doi.org/10.3390/mi11060616

AMA Style

Vrakatseli V, Farsari E, Mataras D. Wetting Properties of Transparent Anatase/Rutile Mixed Phase Glancing Angle Magnetron Sputtered Nano-TiO2 Films. Micromachines. 2020; 11(6):616. https://doi.org/10.3390/mi11060616

Chicago/Turabian Style

Vrakatseli, Vasiliki, Ergina Farsari, and Dimitrios Mataras. 2020. "Wetting Properties of Transparent Anatase/Rutile Mixed Phase Glancing Angle Magnetron Sputtered Nano-TiO2 Films" Micromachines 11, no. 6: 616. https://doi.org/10.3390/mi11060616

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