Next Article in Journal
Dynamic Compression Properties of Ni-Fe Slag Powder Soil Cement under Impact Load
Next Article in Special Issue
Electrodeposition of Copper Oxides as Cost-Effective Heterojunction Photoelectrode Materials for Solar Water Splitting
Previous Article in Journal
Argon and Other Defects in Amorphous SiO2 Coatings for Gravitational-Wave Detectors
Previous Article in Special Issue
Intrinsic Properties and Future Perspective of HfO2/V2O5/HfO2 Multi-Layer Thin Films via E-Beam Evaporation as a Transparent Heat Mirror
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Porous vs. Nanotubular Anodic TiO2: Does the Morphology Really Matters for the Photodegradation of Caffeine?

1
Department of Inorganic Chemistry, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovicova 6, Mlynska Dolina, 842 15 Bratislava, Slovakia
2
Institute of Environmental Technology, CEET, VSB-Technical University of Ostrava, 17, listopadu 15/2172, 70800 Ostrava-Poruba, Czech Republic
3
Department of Environmental Ecology and Landscape Management, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovicova 6, Mlynska Dolina, 842 15 Bratislava, Slovakia
4
Department of Experimental Physics, Mathematics, and Informatics, Comenius University in Bratislava, 842 48 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Coatings 2022, 12(7), 1002; https://doi.org/10.3390/coatings12071002
Submission received: 29 June 2022 / Revised: 14 July 2022 / Accepted: 15 July 2022 / Published: 16 July 2022
(This article belongs to the Special Issue Advanced Electrochemical Surface Properties)

Abstract

:
Herein, the preparation of nanotubular and porous TiO2 structures (TNS) is presented for photocatalytic applications. Different TNS were prepared in three different types of glycerol- and ethylene glycol-based electrolytes on a large area (approx. 20 cm2) via anodization using different conditions (applied potential, fluoride concentration). Morphology, structure, and optical properties of TNS were characterized by Scanning Electron Microscopy (SEM), X-ray Diffractometry (XRD), and Diffuse Reflectance Spectroscopy (DRS), respectively. All TNS possess optical band-gap energy (EBG) in the range from 3.1 eV to 3.2 eV. Photocatalytic degradation of caffeine was conducted to evaluate the efficiency of TNS. Overall, nanotubular TiO2 possessed enhanced degradation efficiencies (up to 50% degradation) compared to those of porous TiO2 (up to 30% degradation). This is due to the unique properties of nanotubular TiO2, e.g., improved incident light utilization. As the anodization of large areas is, nowadays, becoming a trend, we show that both nanotubular and porous TiO2 are promising for their use in photocatalysis and could be potentially applicable in photoreactors for wastewater treatment. We believe this present work can be the foundation for future development of efficient TiO2 nanostructures for industrial applications.

1. Introduction

Electrochemical anodic oxidation (anodization) will soon celebrate 100 years since it was first patented in 1923 by Bengough and Stuart for its use on an industrial scale [1]. In general, the purpose of anodization is to passivate the surface of a material (primarily metals). Thus, the surface of the material develops a protective layer (in the form of oxide), which protects it against future corrosion. Pioneering works were conducted in 1984 [2] and 1995 [3] that allowed the preparation of not only a protective layer, but a unique nanotubular structure of TiO2 and Al2O3, respectively, for several applications in several fields including energy, environment, medicine, and technology. Ever since the anodization of metals gained significant attention from the scientific community, thousands of papers have been related to anodization [4,5,6].
In particular, anodic TiO2 nanotube layers [7] with unique properties of a 1D material are promising for photocatalysis [8], hydrogen production [9], DSSCs [10], or biomedical applications [11]. This is due to [7]: (i) high specific surface area and enhanced light absorption due to the materials’ 1D structure, (ii) a vertical transport of the photogenerated charge carriers (electron/holes), and (iii) tunable dimensions of the nanotubes, in particular, their length and diameter. In the last decade, organic-based electrolytes (e.g., glycerol and ethylene glycol) are the most popular for the preparation of TiO2 nanotube layers via anodization [7]. Indeed, in such electrolytes, a high-aspect ratio and smooth nanotubes are prone to form [12]. Nevertheless, there are still major challenges in the synthesis and application of TiO2 nanotube layers in industry. Indeed, layers of several cm2 are necessary and this relates to specific preparation challenges, i.e., when a potential is applied to areas of several cm2, the absolute current flowing between the two electrodes is much higher compared to that of smaller areas (most of the published articles report a laboratory scale of 1 cm2 area) [7]. The flowing high current leads to an increase in the electrolyte temperature and subsequently a dielectric breakdown will occur [13]. To this day, only a few reports deal with the preparation of TiO2 nanotube layers on areas >10 cm2 [9,13,14,15,16,17,18,19,20]. A field-assisted dissolution of titanium supported by fluoride ions is responsible for the formation of a nanotubular structure [7]. Thus, by using optimized anodization conditions, TiO2 nanotube layers are formed. Nevertheless, there is a thin line between the formation of a nanotubular and a porous structure of TiO2. The potential/current induced dissolution of the formed TiO2 in the presence of fluoride ions via [TiF6]2 is well known to the scientific community [21,22]. In case that the dissolution rate is higher than the formation rate, a rather porous structure is formed instead of a nanotubular one. Such a porous TiO2 is often overlooked in the literature, as it is considered an unsuccessful preparation of TiO2 nanotube layers. However, this formed porous TiO2 possesses a high specific surface area and a sufficient number of active sites necessary for, e.g., photocatalysis similarly to nanotubular TiO2. In our recent work [9], we reported that both nanotubular and porous anodic TiO2 nanostructures (TNS) possess efficient properties in the photocatalytic degradation of caffeine and hydrogen production. Therefore, we aim this work as a follow-up to our recent paper. Here, we focus solely on the degradation of caffeine (i.e., liquid phase photocatalysis) by using different types of anodic nanotubular and porous TNS prepared via anodization on a large area to gain knowledge about these two, although, different but in many ways similar materials. We believe that both nanotubular and porous anodic TNS are promising for the application in photo-induced processes including photocatalysis.
In the present work, TNS on a large area (approx. 20 cm2) were prepared via anodization in different organic-based electrolytes (i.e., glycerol and ethylene glycol) containing fluoride ions. Scanning Electron Microscopy (SEM), X-ray Diffractometry (XRD), and Diffuse Reflectance Spectroscopy (DRS) was conducted to ensure the morphology, crystal structure, and optical properties of the prepared TNS, respectively. Caffeine was employed as a model organic pollutant to evaluate the photocatalytic activity of the prepared TNS. The present work critically discusses the pros and cons of nanotubular and porous anodic TiO2 for application in liquid-phase photocatalysis. It is worthy to note that herein, we do not aim for 100% pollutant degradation, but rather for a fundamental understanding of different TNS for future development in photocatalysis.

2. Materials and Methods

2.1. Synthesis of TNS

Titanium foil (Sigma-Aldrich, Darmstadt, Germany, 99.7% purity, 0.127 mm thickness, 20 cm2 area) was used as a starting substrate for the formation of TNS. According to our previous work [9], a circle of 5 cm in diameter was drawn on the square titanium foil and cut out into three corners to make a drop-like shape which was used as a working electrode (titanium foil of 5 × 2 cm2 was used as a counter electrode). One corner of the electrode was utilized as a handle to fasten the electrode. The distance between the electrodes was kept at 1.6 cm. Anodization was conducted using potential from 20 V to 80 V (20 V step) for 100 min in glycerol- (Central-Chem, Bratislava, Slovakia) and ethylene glycol-based (Central-Chem, Bratislava, Slovakia) electrolytes. To avoid overheating of the electrolyte, cooling with an ice bath was conducted to possess a constant temperature of 8 °C during anodization. Three different types of electrolytes based on either glycerol or ethylene glycol were used for the synthesis of TNS according to our previous works [9,23]. We also used three optimized electrolyte compositions, which our group often uses [24,25,26,27,28,29,30]. Moreover, we varied the concentration of fluoride ions by decreasing it to 75% and 50% of the initial amount (100%), respectively, for each electrolyte. Thus, we obtained altogether 9 different electrolytes and the electrolyte compositions along with the materials’ abbreviations and the used anodization conditions are summarized in Table 1. By these efforts, both nanotubular and porous TiO2 structures were obtained as required for the purpose of this study. After synthesis, TNS were immersed in isopropanol to remove the electrolyte leftovers from the surface and subsequently annealed in a muffle oven at 400 °C in air (2.21 °C/min heating rate). Afterward, TNS were cooled down inside the oven to room temperature.

2.2. Characterization and Photocatalytic Activity of TNS

Morphology and structure of the TNS layers was studied using Scanning Electron Microscopy (SEM, Lyra 3 Tescan, Tescan, Brno, Czech Republic, at 10 kV) equipped with Energy-Dispersive X-ray Spectroscopy (EDS) and X-ray Diffractometer (20°–60° 2 Theta; XRD, PANalytical Cu Kα radiation, λ = 1.5418 Å, Malvern-Panalytical, Malvern, United Kingdom), respectively. Optical properties were studied via Diffuse Reflectance Spectroscopy in the wavelength range from 200 nm to 800 nm (DRS, Shimadzu UV-2600, Shimadzu, Kyoto, Japan) equipped with an integrating sphere 2600 Plus.
Photocatalytic degradation of caffeine (CAF, Reagent Plus, Sigma-Aldrich, Darmstadt, Germany) aqueous solution (10 ppm, 50 mL) was evaluated in a home-made photoreactor, which consisted of a circular glass dish with a diameter of 8 cm. A sample carrier was placed at the bottom of the glass dish. Prior to the photocatalytic measurements, the whole assembly (along with TNS) was placed in the dark for 20 min to achieve the adsorption/desorption equilibrium. Afterwards, an 8W Hg lamp (Ultra-Violet Products Inc. UVP, Upland, CA, USA, λmax = 365 nm) was used as the radiation source. The degradation of CAF was monitored for 3 h at 20 min intervals. The solution was withdrawn (0.5 mL) with a micropipette and mixed with 0.5 mL of distilled water. The concentration of CAF was determined by UV-VIS spectrophotometer (Jasco V530, Kyoto, Japan), according to our previous work [9]. To confirm the good stability of all TNSs, the photocatalytic activity measurements were repeated 3 times for each sample, and the differences in the overall photocatalytic performance did not exceed ±5%.

3. Results and Discussion

Anodization of a 20 cm2 area Ti was conducted using fluoride-containing glycerol- and ethylene glycol-based electrolytes to obtain the herein-presented materials. The anodization conditions that we used led to the formation of both nanotubular and porous TiO2 structures. Therefore, to avoid any confusion, we abbreviated the prepared materials as TNS (TiO2 nanostructures).
Representative SEM images of TNS prepared in EL1, EL2 and EL33 with different concentrations of fluoride ions (according to Table 1) at 60 V are shown in Figure 1. Using similar conditions (voltage and anodization time), both nanotubular and porous structures were formed, depending primarily on the electrolyte composition. We conducted the SEM analysis on all the prepared TNS according to Table 1; however, we did not include all the SEM images in the manuscript (not to overload the manuscript with plenty of SEM images). The surface of each prepared TNS possess similar surface morphology as seen in Figure 1. Nevertheless, we summarized the morphological features (tube/pore diameter and thickness of the layer) of each TNS in Table 2, i.e., TNS prepared in different electrolytes (glycerol- or ethylene glycol-based ones), with different concentrations of fluoride ions (according to Table 1), at different applied potentials (from 20 V to 80 V). In general, electrolyte composition and applied conditions during anodization are crucial factors that affect the final morphology of TNS. Fluoride ions present in electrolytes act as etching agents and are directly responsible for the formation of either nanotubular or porous structures [21,22,31,32,33]. Generally, when the dissolution rate is higher than the formation rate, a rather porous structure is formed instead of a nanotubular one [34]. Thus, the mechanical stability of the formed nanotubular structure is disturbed, and as a result, a collapse of the tubes’ mouth is observed. Nevertheless, both nanotubular and porous structures possess good adhesion to the underlying Ti substrate, i.e., the formed layers do not peel off the substrate after bending. Moreover, both structures are efficient in applications such as photocatalysis and hydrogen production [9].
Several trends were observed from the obtained data summarized in Table 1 and are as follows: In all types of electrolytes, the pore diameter and thickness of the nanotubular or porous TiO2 layer increased with the increased potential. This is in accordance with the plethora of reports on the effect of anodization conditions on the morphology of the layers [7,35,36,37]. In general, at higher potentials (i.e., 60 V and 80 V), rather a porous structure is formed instead of a nanotubular one. Indeed, the etching rate is higher at increased potentials; thus, a porous structure formation is present. At lower potentials (i.e., 20 V and 40 V) a nanotubular structure is prone to form. Nevertheless, in the case of EL2, the porous structure was observed at 20 V, and nanotubular structures were formed at 40 V, 60 V, and 80 V. This is due to the low content of distilled water in the electrolyte (3 mL); thus, higher potentials are required for nanotubular structure to be obtained [7]. Indeed, the viscosity of the electrolyte that contains 197 mL of ethylene glycol and 3 mL of distilled water inhibits the migration of fluoride ions present in the electrolyte during anodization.
Figure 2 shows XRD patterns of annealed TNS; XRD was performed on all the TNS (a total of 36 samples), but the applied potential during anodization (20 V–80 V) and electrolyte composition has no impact on the phase composition of the material after annealing, and all the TNS possess similar phase composition. Therefore, we included representative XRD patterns for each electrolyte type (according to Table 1) prepared at 60 V. In all TNS, two different phases were identified, i.e., tetragonal anatase TiO2 (P42/mnm; ICCD 01-086-1157) [38,39] and hexagonal metallic Ti (P63/mmc; ICCD 00-044-1294) [23]. The visible diffractions of metallic Ti stem from the underlying Ti substrate and are visible due to the penetration of X-rays. Nevertheless, the prepared nanotubular and porous TiO2 structures consist solely of anatase TiO2. The XRD patterns show typical diffractions of anodically prepared TiO2 layers [9,23].
UV-VIS DRS spectra and the corresponding Kubelka–Munk curves were recorded to determine the reflectance and the indirect optical band gap energy (EBG) of TNS (Figure 3). Although DRS measurements were performed on all TNS, we show representative spectra of TNS prepared at 60 V (similarly to SEM in Figure 1 and XRD in Figure 2). All TNS (36 different ones) possess similar optical properties with EBG in the energy range from 3.1 eV to 3.2 eV. The obtained values of the indirect EBG are in good agreement with previous reports on TiO2 materials composed solely of the anatase phase [40,41,42,43]. Indeed, the appeared reflectance edge at ~400 nm is attributed to the anatase phase with EBG ~3.2 eV [44], and no substantial differences were observed in the different TNS. Both nanotubular and porous structures contain solely anatase TiO2, thus their EBG are similar.
Photocatalytic degradation of caffeine (CAF) was explored under UVA light irradiation (λmax = 365 nm) on all TNS, and the results are summarized in Figure 4, along with the illustration of the photoreactor that was used. Photocatalytic degradation of CAF follows the first-order reaction typical for TiO2 photocatalysts and an organic pollutant [45]. Although the degradation rates were calculated, we decided to add the conversion efficiency of CAF in % for an easier and more understandable interpretation of the results (to avoid 36 different rate constants that would confuse the reader). The conversion efficiency was calculated directly from the degradation extents after 3 h, i.e., from the initial and final concentration of CAF. Overall, the most efficient CAF degradation was obtained using nanotubular structures. Indeed, the unique properties of the nanotubular TiO2 structures, such as improved charge carrier transport along the nanotube walls and overall enhanced incident light utilization is responsible for this outcome. The most efficient TNS (EL2100% at prepared 40 V) degraded approx. 50% of the pollutant. A clear trend appeared wherein by decreasing the amount of fluoride ions during anodization, the photocatalytic degradation efficiency decreased. Indeed, fluoride ions are responsible for the formation of nanotubular and porous TiO2 structures; thus, by lowering the total amount of fluorides, the etching rate during anodization is low. As a result of the low content of fluorides, thinner layers are obtained during anodization in EL1-350% as seen in Table 2 (with a maximum thickness of approx. 5 μm). With the increasing of the fluoride content in the electrolytes (EL1-375%–100%), the thickness substantially increased and reached approx. 17 μm in EL2100%. Moreover, the pore diameter is another crucial factor that directly affects the specific surface area. Herein, Brunauer–Emmett–Telle (BET) theory [46] is not applicable in the case of TNS, as it is challenging to measure BET on our kind of samples, i.e., to remove the layers from the underlying substrate and proceed with the BET measurements. Nevertheless, the obtained SEM data (Figure 1) show clear differences in the pore diameter; thus, it indirectly confirms a high surface area of the material, in general. As the degradation of CAF proceeds at the TNS/CAF interface, surface area (or porosity) is a crucial factor. The interplay of morphology (SEM, Figure 1), structure (XRD, Figure 2), and optical properties (DRS and Kubelka–Munk plots, Figure 3) are in favor of the photocatalytic efficiency of the herein-presented TNS prepared in glycerol- and ethylene glycol-based electrolytes with higher concentration of fluoride ions. The porous TiO2 structures showed approx. 2 times lower degradation efficiencies. This is due collapse/distortion of the nanotubular structure during anodization by the applied conditions. The pores are denser, thus providing fewer active sites for CAF to be adsorbed and subsequently degraded. Nevertheless, a satisfactory degradation efficiency of CAF (approx. 30%) was obtained by using porous TiO2 (EL3100% prepared at 60 V). All in all, both nanotubular and porous TiO2 structures prepared via anodization are promising for application in photoreactors for wastewater treatment. Nevertheless, there are still major challenges of synthesis such nanostructures on large area, and such synthesis is necessary for industrial applications of these nanomaterials. At last, Table 3 shows the comparison of our TNS with reports on anodic TiO2 based nanostructures for caffeine photocatalytic degradation.

4. Conclusions

Anodization of an approx. 20 cm2 Ti area was conducted to prepare nanotubular and porous TiO2 structures for comparison in the photocatalytic degradation of caffeine under UVA. A variety of conditions were conducted to obtain a total of 36 different TNS: (i) three different electrolytes based on glycerol and ethylene glycol, (ii) three different concentrations of fluoride ions in each type of electrolyte, and (iii) different applied potentials from 20 V to 80 V for each type of electrolyte. Depending on the different conditions, nanotubular and porous TiO2 structures were formed with thickness up to 17 μm and pore diameter up to 100 nm, and subsequently annealed at 400 °C to obtain crystalline anatase TiO2. All TNS possess EBG in the range from 3.1 eV to 3.2 eV. Overall, both nanotubular and porous TNS lead to efficient degradation extents of caffeine after 3 h, although nanotubular TNS showed enhanced photocatalytic activity (reaching degradation up to 50%) compared to that of porous TNS. Indeed, the unique properties of 1D nanotubular structures are responsible for the improved photocatalytic activity. The degradation efficiency of both nanostructures is due to their high surface area associated with a sufficient number of active sites for caffeine to adsorb on the surface of TNS and subsequently degrade the pollutant. As the anodization of large areas is, nowadays, becoming a trend, we show that both nanotubular and porous TiO2 are promising for their use in photocatalysis and could, potentially, be applicable in photoreactors for wastewater treatment. We believe this present work can be the foundation for future development of efficient TiO2 nanostructures for industrial applications.

Author Contributions

Conceptualization, M.M.; methodology, M.S., M.B.H., V.L., O.M. and M.M.; software, M.S. and M.B.H.; validation, M.S., M.B.H. and M.M.; formal analysis, M.S., H.M., M.B.H. and V.L.; investigation M.S., M.B.H. and H.M.; resources, M.M.; data curation, M.B.H.; writing—original draft preparation, M.S., M.B.H., O.M. and M.M.; writing—review and editing, M.S., M.B.H., V.L., O.M. and M.M.; visualization, M.S. and M.B.H.; supervision, M.M.; project administration, M.M.; funding acquisition, M.S., O.M., M.B.H. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was co-financed by (i) Slovak Research and Development Agency (APVV): contracts No. APVV-21-0039 and No. APVV-21-0053, (ii) Scientific Grant Agency of the Slovak Ministry of Education, Sciences, Research and Sport (VEGA): project No. 1/0062/22, (iii) Grant of the Comenius University Bratislava for Young Scientists (UK/3/2022), (iv) the ESF in “Science without borders 2.0” project, reg. nr. CZ.02.2.69/0.0/0.0/18_053/0016985 within the Operational Programme Research, Development and Education, and (v) ERDF “Institute of Environmental Technology-Excellent Research” (No. CZ.02.1.01/0.0/0.0/16_019/0000853). This work has also been partially supported by the project USCCCORD (ŽoNFP:NFP313020BUZ3), co-financed by the European Regional Development Fund within the Operational Programme Integrated Infrastructure.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bengough, G.D.; Stuart, J.M. Improved Process of Protecting Surfaces of Aluminium of Aluminium Alloys. UK Pat. 1923, 223, 36. [Google Scholar]
  2. Assefpour-Dezfuly, M.; Vlachos, C.; Andrews, E.H. Oxide Morphology and Adhesive Bonding on Titanium Surfaces. J. Mater. Sci. 1984, 19, 3626–3639. [Google Scholar] [CrossRef]
  3. Masuda, H.; Fukuda, K. Ordered Metal Nanohole Arrays Made by a Two-Step Replication of Honeycomb Structures of Anodic Alumina. Science 1995, 268, 1466–1468. [Google Scholar] [CrossRef] [PubMed]
  4. El-Said Shehata, O.; Abdel-karim, A.M.; Abdel Fatah, A.H. New Trends in Anodizing and Electrolytic Coloring of Metals. Egypt. J. Chem. 2022, 65, 229–241. [Google Scholar] [CrossRef]
  5. David, T.M.; Dev, P.R.; Wilson, P.; Sagayaraj, P.; Mathews, T. A Critical Review on the Variations in Anodization Parameters toward Microstructural Formation of TiO2 Nanotubes. Electrochem. Sci. Adv. 2021, e202100083. [Google Scholar] [CrossRef]
  6. Minagar, S.; Berndt, C.C.; Wang, J.; Ivanova, E.; Wen, C. A Review of the Application of Anodization for the Fabrication of Nanotubes on Metal Implant Surfaces. Acta Biomater. 2012, 8, 2875–2888. [Google Scholar] [CrossRef]
  7. Lee, K.; Mazare, A.; Schmuki, P. One-Dimensional Titanium Dioxide Nanomaterials: Nanotubes. Chem. Rev. 2014, 114, 9385–9454. [Google Scholar] [CrossRef] [Green Version]
  8. Macak, J.M.; Zlamal, M.; Krysa, J.; Schmuki, P. Self-Organized TiO2 Nanotube Layers as Highly Efficient Photocatalysts. Small 2007, 3, 300–304. [Google Scholar] [CrossRef]
  9. Šihor, M.; Hanif, M.B.; Thirunavukkarasu, G.K.; Liapun, V.; Edelmannova, M.F.; Roch, T.; Satrapinskyy, L.; Plecenik, T.; Rauf, S.; Hensel, K.; et al. Anodization of Large Area Ti: Versatile Material for Caffeine Photodegradation and Hydrogen Production. Catal. Sci. Technol. 2022. [Google Scholar] [CrossRef]
  10. O’Regan, B.; Grätzel, M. A Low-Cost, High-Efficiency Solar Cell Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  11. Motola, M.; Capek, J.; Zazpe, R.; Bacova, J.; Hromádko, L.; Bruckova, L.; Ng, S.; Handl, J.; Spotz, Z.; Knotek, P.; et al. Thin TiO2 Coatings by ALD Enhance the Cell Growth on TiO2 Nanotubular and Flat Substrates. ACS Appl. Bio Mater. 2020, 3, 6447–6456. [Google Scholar] [CrossRef] [PubMed]
  12. Macak, J.M.; Tsuchiya, H.; Taveira, L.; Aldabergerova, S.; Schmuki, P. Smooth Anodic TiO2 Nanotubes. Angew. Chemie Int. Ed. 2005, 44, 7463–7465. [Google Scholar] [CrossRef] [PubMed]
  13. Sopha, H.; Baudys, M.; Krbal, M.; Zazpe, R.; Prikryl, J.; Krysa, J.; Macak, J.M. Scaling up Anodic TiO2 Nanotube Layers for Gas Phase Photocatalysis. Electrochem. Commun. 2018, 97, 91–95. [Google Scholar] [CrossRef]
  14. Motola, M.; Satrapinskyy, L.; Roch, T.; Šubrt, J.; Kupčík, J.; Klementová, M.; Jakubičková, M.; Peterka, F.; Plesch, G. Anatase TiO2 nanotube Arrays and Titania Films on Titanium Mesh for Photocatalytic NOX removal and Water Cleaning. Catal. Today 2017, 287, 59–64. [Google Scholar] [CrossRef]
  15. Mena, E.; Martín de Vidales, M.J.; Mesones, S.; Marugán, J. Influence of Anodization Mode on the Morphology and Photocatalytic Activity of TiO2-NTs Array Large Size Electrodes. Catal. Today 2018, 313, 33–39. [Google Scholar] [CrossRef]
  16. Szkoda, M.; Trzciński, K.; Zarach, Z.; Roda, D.; Łapiński, M.; Nowak, A.P. Scaling Up the Process of Titanium Dioxide Nanotube Synthesis and Its Effect on Photoelectrochemical Properties. Materials 2021, 14, 5686. [Google Scholar] [CrossRef]
  17. Kim, H.-I.; Kim, D.; Kim, W.; Ha, Y.-C.; Sim, S.-J.; Kim, S.; Choi, W. Anodic TiO2 Nanotube Layer Directly Formed on the Inner Surface of Ti Pipe for a Tubular Photocatalytic Reactor. Appl. Catal. A Gen. 2016, 521, 174–181. [Google Scholar] [CrossRef]
  18. Xiang, C.; Sun, L.; Wang, Y.; Wang, G.; Zhao, X.; Zhang, S. Large-Scale, Uniform, and Superhydrophobic Titania Nanotubes at the Inner Surface of 1000 Mm Long Titanium Tubes. J. Phys. Chem. C 2017, 121, 15448–15455. [Google Scholar] [CrossRef]
  19. Ghosh, J.P.; Achari, G.; Langford, C.H. Design and Evaluation of a UV LED Photocatalytic Reactor Using Anodized TiO2 Nanotubes. Water Environ. Res. 2016, 88, 785–791. [Google Scholar] [CrossRef]
  20. Franz, S.; Perego, D.; Marchese, O.; Bestetti, M. Photoelectrochemical Advanced Oxidation Processes on Nanostructured TiO2 Catalysts: Decolorization of a Textile Azo-Dye. J. Water Chem. Technol. 2015, 37, 108–115. [Google Scholar] [CrossRef] [Green Version]
  21. Zhang, W.; Liu, Y.; Guo, F.; Liu, J.; Yang, F. Kinetic Analysis of the Anodic Growth of TiO2 Nanotubes: Effects of Voltage and Temperature. J. Mater. Chem. C 2019, 7, 14098–14108. [Google Scholar] [CrossRef]
  22. Indira, K.; Mudali, U.K.; Nishimura, T.; Rajendran, N. A Review on TiO2 Nanotubes: Influence of Anodization Parameters, Formation Mechanism, Properties, Corrosion Behavior, and Biomedical Applications. J. Bio-Tribo-Corrosion 2015, 1, 1–22. [Google Scholar] [CrossRef] [Green Version]
  23. Motola, M.; Hromadko, L.; Prikryl, J.; Sopha, H.; Krbal, M.; Macak, J.M. Intrinsic Properties of High-Aspect Ratio Single- and Double-Wall Anodic TiO2 Nanotube Layers Annealed at Different Temperatures. Electrochim. Acta 2020, 352, 136479. [Google Scholar] [CrossRef]
  24. Michalkova, H.; Skubalova, Z.; Sopha, H.; Strmiska, V.; Tesarova, B.; Dostalova, S.; Svec, P.; Hromadko, L.; Motola, M.; Macak, J.M.; et al. Complex Cytotoxicity Mechanism of Bundles Formed from Self-Organised 1-D Anodic TiO2 Nanotubes Layers. J. Hazard. Mater. 2020, 388, 122054. [Google Scholar] [CrossRef]
  25. Motola, M.; Čaplovičová, M.; Krbal, M.; Sopha, H.; Thirunavukkarasu, G.K.; Gregor, M.; Plesch, G.; Macak, J.M. Ti3+ Doped Anodic Single-Wall TiO2 Nanotubes as Highly Efficient Photocatalyst. Electrochim. Acta 2020, 331, 135374. [Google Scholar] [CrossRef]
  26. Krbal, M.; Ng, S.; Motola, M.; Hromadko, L.; Dvorak, F.; Prokop, V.; Sopha, H.; Macak, J.M. Sulfur Treated 1D Anodic TiO2 Nanotube Layers for Significant Photo- and Electroactivity Enhancement. Appl. Mater. Today 2019, 17, 104–111. [Google Scholar] [CrossRef]
  27. Motola, M.; Sopha, H.; Krbal, M.; Hromádko, L.; Zmrhalová, Z.O.; Plesch, G.; Macak, J.M. Comparison of Photoelectrochemical Performance of Anodic Single- and Double-Walled TiO2 nanotube Layers. Electrochem. Commun. 2018, 97, 1–5. [Google Scholar] [CrossRef]
  28. Motola, M.; Baudys, M.; Zazpe, R.; Krbal, M.; Michalička, J.; Rodriguez-Pereira, J.; Pavliňák, D.; Přikryl, J.; Hromádko, L.; Sopha, H.; et al. 2D MoS 2 Nanosheets on 1D Anodic TiO 2 Nanotube Layers: An Efficient Co-Catalyst for Liquid and Gas Phase Photocatalysis. Nanoscale 2019, 11, 23126–23131. [Google Scholar] [CrossRef]
  29. Motola, M.; Satrapinskyy, L.; Čaplovicová, M.; Roch, T.; Gregor, M.; Grančič, B.; Greguš, J.; Čaplovič, Ľ.; Plesch, G. Enhanced Photocatalytic Activity of Hydrogenated and Vanadium Doped TiO2 nanotube Arrays Grown by Anodization of Sputtered Ti Layers. Appl. Surf. Sci. 2018, 434, 1257–1265. [Google Scholar] [CrossRef]
  30. Motola, M.; Zazpe, R.; Hromadko, L.; Prikryl, J.; Cicmancova, V.; Rodriguez-Pereira, J.; Sopha, H.; Macak, J.M. Anodic TiO2 Nanotube Walls Reconstructed: Inner Wall Replaced by ALD TiO2 Coating. Appl. Surf. Sci. 2021, 549, 149306. [Google Scholar] [CrossRef]
  31. Cha, G.; Schmuki, P.; Altomare, M. Free-Standing Membranes to Study the Optical Properties of Anodic TiO2 Nanotube Layers. Chem. Asian J. 2016, 11, 789–797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Fraoucene, H.; Sugiawati, V.A.; Hatem, D.; Belkaid, M.S.; Vacandio, F.; Eyraud, M.; Pasquinelli, M.; Djenizian, T. Optical and Electrochemical Properties of Self-Organized TiO2 Nanotube Arrays from Anodized Ti-6Al-4V Alloy. Front. Chem. 2019, 7, 66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Pishkar, N.; Jedi-soltanabadi, Z.; Ghoranneviss, M. Reduction in the Band Gap of Anodic TiO2 Nanotube Arrays by H2 Plasma Treatment. Results Phys. 2018, 10, 466–468. [Google Scholar] [CrossRef]
  34. Varghese, O.K.; Gong, D.; Paulose, M.; Grimes, C.A.; Dickey, E.C. Crystallization and High-Temperature Structural Stability of Titanium Oxide Nanotube Arrays. J. Mater. Res. 2003, 18, 156–165. [Google Scholar] [CrossRef]
  35. Hamlekhan, A.; Butt, A.; Patel, S.; Royhman, D.; Takoudis, C.; Sukotjo, C.; Mathew, M.T.; Shokuhfar, T. Optimization of Anodization and Annealing Condition Enhances TiO2 Nanotubular Surface Hydrophilicity. In TMS 2014: 143rd Annual Meeting & Exhibition; Springer: Cham, Germany, 2014; pp. 221–228. [Google Scholar] [CrossRef]
  36. Gulati, K.; Santos, A.; Findlay, D.; Losic, D. Optimizing Anodization Conditions for the Growth of Titania Nanotubes on Curved Surfaces. J. Phys. Chem. C 2015, 119, 16033–16045. [Google Scholar] [CrossRef]
  37. Nogueira, R.P.; Uchoa, J.D.; Hilario, F.; Santana-Melo, G.D.F.; de Vasconcellos, L.M.R.; Marciano, F.R.; Roche, V.; Junior, A.M.J.; Lobo, A.O. Characterization of Optimized TiO2 Nanotubes Morphology for Medical Implants: Biological Activity and Corrosion Resistance. Int. J. Nanomed. 2021, 16, 667–682. [Google Scholar] [CrossRef]
  38. Zhang, H.; Banfield, J.F. Phase Transformation of Nanocrystalline Anatase-to-Rutile via Combined Interface and Surface Nucleation. J. Mater. Res. 2011, 15, 437–448. [Google Scholar] [CrossRef]
  39. Gouma, P.I.; Mills, M.J. Anatase-to-Rutile Transformation in Titania Powders. J. Am. Ceram. Soc. 2001, 84, 619–622. [Google Scholar] [CrossRef]
  40. Liu, H.Y.; Hsu, Y.L.; Su, H.Y.; Huang, R.C.; Hou, F.Y.; Tu, G.C.; Liu, W.H. A Comparative Study of Amorphous, Anatase, Rutile, and Mixed Phase TiO2 Films by Mist Chemical Vapor Deposition and Ultraviolet Photodetectors Applications. IEEE Sens. J. 2018, 18, 4022–4029. [Google Scholar] [CrossRef]
  41. Siah, W.R.; Lintang, H.O.; Shamsuddin, M.; Yuliati, L. High Photocatalytic Activity of Mixed Anatase-Rutile Phases on Commercial TiO2 Nanoparticles. IOP Conference Series: Materials Science and Engineering, 10th Joint Conference on Chemistry 8–9 September 2015, Solo, Indonesia; IOP Publishing: Bristol, UK, 2016; Volume 107, p. 012005. [Google Scholar] [CrossRef]
  42. Fu, W.; Li, G.; Wang, Y.; Zeng, S.; Yan, Z.; Wang, J.; Xin, S.; Zhang, L.; Wu, S.; Zhang, Z. Facile Formation of Mesoporous Structured Mixed-Phase (Anatase/Rutile) TiO2 with Enhanced Visible Light Photocatalytic Activity. Chem. Commun. 2017, 54, 58–61. [Google Scholar] [CrossRef] [PubMed]
  43. Luo, Z.; Poyraz, A.S.; Kuo, C.H.; Miao, R.; Meng, Y.; Chen, S.Y.; Jiang, T.; Wenos, C.; Suib, S.L. Crystalline Mixed Phase (Anatase/Rutile) Mesoporous Titanium Dioxides for Visible Light Photocatalytic Activity. Chem. Mater. 2015, 27, 6–17. [Google Scholar] [CrossRef]
  44. Rambabu, Y.; Jaiswal, M.; Roy, S.C. Effect of Annealing Temperature on the Phase Transition, Structural Stability and Photo-Electrochemical Performance of TiO2 Multi-Leg Nanotubes. Catal. Today 2016, 278, 255–261. [Google Scholar] [CrossRef]
  45. Zlamal, M.; Macak, J.M.; Schmuki, P.; Krýsa, J. Electrochemically Assisted Photocatalysis on Self-Organized TiO2 Nanotubes. Electrochem. Commun. 2007, 9, 2822–2826. [Google Scholar] [CrossRef]
  46. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 2002, 60, 309–319. [Google Scholar] [CrossRef]
  47. Suhadolnik, L.; Marinko, Ž.; Ponikvar-Svet, M.; Tavčar, G.; Kovač, J.; Čeh, M. Influence of Anodization-Electrolyte Aging on the Photocatalytic Activity of TiO2 Nanotube Arrays. J. Phys. Chem. C 2020, 124, 4073–4080. [Google Scholar] [CrossRef]
  48. Thirunavukkarasu, G.K.; Gowrisankaran, S.; Caplovicova, M.; Satrapinskyy, L.; Gregor, M.; Lavrikova, A.; Gregus, J.; Halko, R.; Plesch, G.; Motola, M.; et al. Contribution of Photocatalytic and Fenton-Based Processes in Nanotwin Structured Anodic TiO2 Nanotube Layers Modified by Ce and V. Dalt. Trans. 2022. [Google Scholar] [CrossRef]
  49. Arfanis, M.K.; Adamou, P.; Moustakas, N.G.; Triantis, T.M.; Kontos, A.G.; Falaras, P. Photocatalytic Degradation of Salicylic Acid and Caffeine Emerging Contaminants Using Titania Nanotubes. Chem. Eng. J. 2017, 310, 525–536. [Google Scholar] [CrossRef]
  50. Marinko, Ž.; Suhadolnik, L.; Šetina Batič, B.; Šelih, V.S.; Majaron, B.; Kovač, J.; Čeh, M. Toward a Flexible and Efficient TiO2 Photocatalyst Immobilized on a Titanium Foil. ACS Omega 2021, 6, 23233–23242. [Google Scholar] [CrossRef]
Figure 1. Representative SEM images of TNS prepared in different electrolytes (according to Table 1) at 60 V: (ac) EL1100%, EL175%, and EL150%, respectively, (df) EL2100%, EL275%, and EL250%, respectively, and (gi) EL3100%, EL375%, and EL350%, respectively.
Figure 1. Representative SEM images of TNS prepared in different electrolytes (according to Table 1) at 60 V: (ac) EL1100%, EL175%, and EL150%, respectively, (df) EL2100%, EL275%, and EL250%, respectively, and (gi) EL3100%, EL375%, and EL350%, respectively.
Coatings 12 01002 g001
Figure 2. XRD patterns of TNS prepared in different electrolytes (according to Table 1) at 60 V. A—anatase; T—titanium.
Figure 2. XRD patterns of TNS prepared in different electrolytes (according to Table 1) at 60 V. A—anatase; T—titanium.
Coatings 12 01002 g002
Figure 3. UV-VIS DRS spectra and the corresponding Kubelka–Munk curves (as insets) of TNS prepared in different electrolytes (according to Table 1) at 60 V: (a) EL150%–EL1100%, (b) EL250%–EL2100%, and (c) EL350%–EL3100%.
Figure 3. UV-VIS DRS spectra and the corresponding Kubelka–Munk curves (as insets) of TNS prepared in different electrolytes (according to Table 1) at 60 V: (a) EL150%–EL1100%, (b) EL250%–EL2100%, and (c) EL350%–EL3100%.
Coatings 12 01002 g003
Figure 4. Photocatalytic degradation of caffeine using TNS prepared in different electrolytes (according to Table 1): (a) EL1-3100%, (b) EL1-375%, (c) EL1-350%, and (d) illustration of the photocatalytic reactor. T—nanotubular; P—porous.
Figure 4. Photocatalytic degradation of caffeine using TNS prepared in different electrolytes (according to Table 1): (a) EL1-3100%, (b) EL1-375%, (c) EL1-350%, and (d) illustration of the photocatalytic reactor. T—nanotubular; P—porous.
Coatings 12 01002 g004
Table 1. Summary: electrolyte abbreviation, composition, and applied potential.
Table 1. Summary: electrolyte abbreviation, composition, and applied potential.
ElectrolyteElectrolyte CompositionElectrolyteElectrolyte CompositionElectrolyteElectrolyte CompositionApplied Potential (V)
EL1100%2 g NH4F; 100 mL glycerol; 100 mL H2OEL2100%1.3 g NH4F; 197 mL ethylene glycol; 3 mL H2OEL3100%1.1045 g NH4F; 180 mL ethylene glycol; 20 mL H2O20
40
60
80
EL175%1.5 g NH4F; 100 mL glycerol; 100 mL H2OEL275%0.975 g NH4F; 197 mL ethylene glycol; 3 mL H2OEL375%0.83 g NH4F; 180 mL ethylene glycol; 20 mL H2O20
40
60
80
EL150%1 g NH4F; 100 mL glycerol; 100 mL H2OEL250%0.65 g NH4F; 197 mL ethylene glycol; 3 mL H2OEL350%0.55 g NH4F; 180 mL ethylene glycol; 20 mL H2O20
40
60
80
Table 2. Morphological features of the prepared TNS in different electrolytes.
Table 2. Morphological features of the prepared TNS in different electrolytes.
ElectrolyteApplied Potential (V)Morphological Features
Structure TypePore Diameter (nm)Layer Thickness (μm)
EL1100%20Nanotubular30–603–4
40Nanotubular40–706–8
60Porous50–1007–9
80Porous50–1009–12
EL175%20Nanotubular20–406–8
40Nanotubular50–806–8
60Porous50–1006–8
80Porous50–1008–10
EL150%20Nanotubular10–202–4
40Nanotubular30–502–4
60Porous50–1004–5
80Porous50–1005–6
EL2100%20Porous10–307–8
40Nanotubular20–5012–17
60Nanotubular30–4013–17
80Nanotubular40–6014–17
EL275%20Porous10–204–6
40Nanotubular30–504–8
60Nanotubular40–508–9
80Nanotubular40–509–12
EL250%20Porous5–201–2
40Nanotubular20–401–2
60Nanotubular30–402–3
80Nanotubular30–403–4
EL3100%20Nanotubular30–503–5
40Porous30–703–5
60Porous50–804–7
80Porous60–806–9
EL375%20Nanotubular20–402–3
40Porous40–702–3
60Porous50–804–6
80Porous60–804–8
EL350%20Nanotubular10–201–2
40Porous10–301–3
60Porous50–802–5
80Porous50–904–7
Table 3. The comparison of anodic TiO2 based nanostructure for caffeine photocatalytic degradation. Anodization condition, target pollutant (and its concentration), rate constant, and time duration were taken from the literature.
Table 3. The comparison of anodic TiO2 based nanostructure for caffeine photocatalytic degradation. Anodization condition, target pollutant (and its concentration), rate constant, and time duration were taken from the literature.
MaterialAnodization
Conditions
Caffeine
Concentration
Rate constant (k)/
Conversion Efficiency (%)
Time
(min)
Ref
TiO2Glycol electrolyte with 0.3 wt % NH4F and 2 vol % Dl-H2O
(60 V for 6 h)
50 mg L–144%180 [47]
TiO22g NH4F + 100 mL DI-H2O in glycerol-based
electrolyte
(20 V for 100 min).
20 ppm0.0069 min−1120 [9]
V- TiO2NH4F (1.3 g) in EG (197 mL) and Dl-H2O (3 mL).20 ppm46%120 [48]
TiO298% ethylene glycol, 2% DI-H2O and 0.3 wt % NH4F.15 mg L−151%240 [49]
TiO2EG and a solution of 0.3 wt % NH4F in 2 vol % DI-H2O
(60 V for 3 h)
10 ppm2.12 ± 0.13 × 10–2 min–1180 min [50]
EL2100%1.3 g NH4F; 197 mL EG; 3 mL H2O
(40 V for 100 min)
10 ppm50%180 minThis study
EL275%0.975 g NH4F; 197 mL EG; 3 mL H2O
(80 V for 100 min)
10 ppm50%180 minThis study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hanif, M.B.; Sihor, M.; Liapun, V.; Makarov, H.; Monfort, O.; Motola, M. Porous vs. Nanotubular Anodic TiO2: Does the Morphology Really Matters for the Photodegradation of Caffeine? Coatings 2022, 12, 1002. https://doi.org/10.3390/coatings12071002

AMA Style

Hanif MB, Sihor M, Liapun V, Makarov H, Monfort O, Motola M. Porous vs. Nanotubular Anodic TiO2: Does the Morphology Really Matters for the Photodegradation of Caffeine? Coatings. 2022; 12(7):1002. https://doi.org/10.3390/coatings12071002

Chicago/Turabian Style

Hanif, Muhammad Bilal, Marcel Sihor, Viktoriia Liapun, Hryhorii Makarov, Olivier Monfort, and Martin Motola. 2022. "Porous vs. Nanotubular Anodic TiO2: Does the Morphology Really Matters for the Photodegradation of Caffeine?" Coatings 12, no. 7: 1002. https://doi.org/10.3390/coatings12071002

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