Next Article in Journal
Hot Corrosion Behavior of Single-Layered Gd2Zr2O7, Sm2Zr2O7, and Nd2Zr2O7 Thermal Barrier Coatings Exposed to Na2SO4 + MgSO4 Environment
Previous Article in Journal
Influence of NaCl Solution External Erosion on Corrosion Resistance of RPC Reinforced with Straw Fiber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ti-Mo-O Nanotube Arrays Grown by Anodization of Magnetron Sputtered Films

1
Department of Experimental Physics, Faculty of Mathematics Physics and Informatics, Comenius University, Mlynská Dolina, 842 48 Bratislava, Slovakia
2
Department of Inorganic Chemistry, Faculty of Natural Sciences, Comenius University, Ilkovičová 6, Mlynská Dolina, 842 15 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(8), 1309; https://doi.org/10.3390/coatings13081309
Submission received: 13 June 2023 / Revised: 4 July 2023 / Accepted: 17 July 2023 / Published: 26 July 2023

Abstract

:
In this study, we introduced the method for the growth of titanium molybdenum oxide (TMO) nanotubes directly from metallic precursor solid state solution and provided their structural and chemical characterization. Precursor films with content of molybdenum from 32 to 82 at% were prepared using co-deposition magnetron sputtering. The optimization of deposition parameters allowed for the growth of a continuous nanotube array with a length up to 700 nm ± 10% by anodic oxidation. Scanning electron microscopy (SEM) combined with energy-dispersive spectroscopy (EDS) revealed nanotube formation with Ti1−xMoxO2 composition, where x can reach the value of 0.5. Scanning transmission electron microscopy combined with EDS (STEM-EDS) confirmed the incorporation of Mo into the TiO2 lattice and uniform elemental distribution across the nanotube at the submicron level. The nanobeam electron diffraction (NBD) and X-ray diffraction analyses (XRD) did not show any notable crystal phase formation for the titanium molybdenum oxide phase.

1. Introduction

The development of energy-saving and renewable technologies is currently one of the main priorities of the power generation industry. The utilization of solar energy is a great opportunity to create clean and renewable energy sources. However, the direct conversion of solar light into electric power is not the only renewable energy source currently in the spotlight. Since the discovery of the electrochemical photolysis of water in 1972 [1], many studies have been focused on improving photocatalytic efficiency. The decomposition of molecules under the influence of sunlight is used in many areas, including ecology, energy generation, and medicine [2]. One of the most promising materials for practical applications is the titanium dioxide-based photo catalyst, which can be used effectively for environmentally friendly energy generation. However, pure TiO2 shows satisfactory photocatalytic properties only under UV light due to its relatively large bandgap of 3.2 eV. Nevertheless, it possesses suitable band-edge positions for many redox reactions, a comparably high lifetime of excited electrons, and an exceptional photo corrosion resistance [3]. To broaden the use of titania nanostructures in the manufacture of chemical sensors and photocatalytic applications, a number of factors, including the conductance of TiO2 in air, the sensing signal, the response and recovery durations, must be enhanced. One of the ways to improve the sensory and photocatalytic properties of titanium dioxide nanotubes is doping with metals such as Mo [4], Mn [5], Cr [6], V [7], Ni [8], Fe [9], and valve metals [10,11]. In the form of nanotube arrays, TiO2-based materials are even more promising compared to particle-like structures due to their better incident light utilization and increased effective surface area. Therefore, they are now popular for use in electrochemical, catalytic, and sensory applications, and also as antimicrobial material [12].
As reported, metallic titanium and molybdenum form solid state solutions in a wide range of concentrations [13]. Molybdenum has also been found to be a promising dopant for improving the photocatalytic performance of TiO2. Ti4+ in the TiO2 structure can be replaced by Mo6+ because their ionic radiuses are similar (Mo6+ 0.62 Å, Ti4+ 0.68 Å) [4,14,15,16,17,18], which leads to a change in the lattice parameters and the bandgap structure. Additionally, Mo is expected to increase the mechanical and physicochemical stability of the material, allowing it to operate under harsher conditions [19].
Anodic oxidation nanotube growth is a well-developed method for aluminum and titanium [20]. The process is relatively inexpensive compared to other nanotube growth techniques. Moreover, the process can be used to create a variety of nanotube structures, such as single-walled nanotubes (SWNTs) [21], double-walled nanotubes (DWNTs) [22], and multi-walled nanotubes (MWNTs) [23]. Doping of titanium dioxide with a low percentage of various elements, including iron or molybdenum, has. Been investigated by a number of scientific groups [4,5,6,7,8,9,24,25,26,27,28,29]. However, it those publications, the value of the dopant typically did not exceed 3 at% In this paper, we used segmented target magnetron sputtering to prepare a metallic alloy with high molybdenum content. We decided to use molybdenum instead of iron, which may produce better photoreactive results (according to [30]) because the ferromagnetic properties of iron can significantly influence the geometry of the discharge and make deposition hardly repeatable. Anodic oxidation was used to prepare TiO2-based nanotubes with high molybdenum alloying of up to 30 at%. We decided not to modify the grown TiO2 nanotube but to directly prepare the alloyed nanotubes for a solid-state solution precursor. It is possible to expect changes to the bandgap value due to the significant presence of Mo in the lattice compared to that of TiO2. This will lead to more efficient utilization of solar light and an improvement of photocatalytic and antimicrobic properties [30].

2. Materials and Methods

A combined method of DC magnetron sputtering and anodic oxidation was employed for the preparation of Mo-alloyed TiO2 nanotubes (Figure 1). The overall process consisted of two main steps: precursor formation and anodic oxidation.

2.1. DC-Magnetron Sputtering

Precursor films were fabricated on a polished Si (100) substrate. Prior to the deposition process, the substrates underwent a cleaning procedure including ultrasonic cleaning in acetone, isopropanol, and distilled water. The base vacuum pressure inside the chamber was <5 × 10−4 Pa. During the sputtering process, argon (99.999% purity) pressure was maintained at 0.9 Pa. The deposition of Ti-Mo mixed films was performed utilizing unbalanced magnetron Torus 2” (Kurt J. Lesker, Jefferson Hills, PA, USA). Instead of using a compound Ti-Mo target, the Ti target (99.95% purity) was covered with a strip of molybdenum (99.5% purity). The amount of molybdenum in the resulting film was adjusted using strips (Figure 1A) with various widths from 10 to 16 mm.

2.2. Anodic Oxidation

The anodization was performed in an ethylene glycol-based electrolyte containing 0.5 wt% ammonium fluoride (97 wt% purity) under similar conditions to those described in [24]. The setup for anodic oxidation is shown in Figure 1B. To avoid rapid dissolution of nanotubes, at the initial phase of oxidation the current density was limited to 3 µA/cm2, while the voltage was increased to the required value. Then, the voltage was kept constant for 30 min. The optimal oxidation parameters were obtained at room temperature with a voltage of 40 V and a constant distance of 20 mm between the sample and the cathode.

2.3. Characterization

The surface morphology of all samples was characterized using field emission scanning electron microscopy (FE-SEM, LYRA III, TESCAN s.r.o., Brno, Chech Republic) and the elemental composition was estimated using energy dispersive X-ray spectroscopy (EDS) (Bruker Nano, Bruker GmBH, Mannheim, Germany).
Lamellas for transmission microscopy were milled in a TESCAN LYRA III with a Ga+ ion beam. The maximum thickness of the TEM sample was 80 nm.
Microstructural characterization (STEM, STEM-EDS, and nanobeam electron diffraction (~100 nm beam width)) was performed using a probe-corrected FEI/TFS Titan Themis 300 electron microscope operated at 200 kV accelerating voltage.
The crystalline phases were identified (The PA Nalytical X’Pert Pro (Malvern Panalytical, Malvern, UK) with Cu Kα source, incidence angle −1.5° was used).

3. Results and Discussion

In our previous study [31], the optimal deposition parameters for pure titanium films were discussed, which served as the starting point for the Mo-alloyed precursor films deposition. These parameters included a magnetron voltage of 650 V and a current of 100 µA, with a target-to-sample distance of 20 mm. Initial Ti-Mo films prepared in the mentioned conditions exhibit a columnar structure that is typical for metal films produced by magnetron sputtering [32]. However, such a structure is not ideal for the subsequent anodic oxidation process. In the anodic oxidation of titanium and molybdenum, fluorine-based electrolytes were utilized, as described in Section 2.2. It is known that fluorine tends to travel along grain boundaries [33], which can interfere with the reaction between the metal (Ti, Mo) and oxygen. To mitigate the formation of columnar grains and promote a more favorable structure for anodic oxidation, we followed Anders’ structure zone model [34] and increased the energy flux during deposition. In this context, we refer to the well-known effect of nanopores suppression due to the local rearrangement of atoms by ion impact in metal layers [34,35,36].
To achieve this, we employed a combination of two methods: reducing the target-to-sample distance and increasing the magnetron power. Additionally, a bias voltage was applied. Figure 2A illustrates the influence of the sample–target distance on the deposition rate, as well as the effect of voltage bias on the Ti-Mo ratio. Optimization was performed on the setup with a 12 mm wide molybdenum strip.
By reducing the target-to-sample distance (Figure 2A,B) and increasing the energy of Ar ions accelerating towards the growing films, the deposition rate of the films can be adjusted. This allows for the control of film thickness and promotes the formation of a more favorable structure for subsequent anodic oxidation. Moreover, the application of a bias voltage during deposition (Figure 2C) can further influence the composition of the Ti-Mo films. These modifications in deposition parameters and energy flux management are crucial for tailoring the structure and composition of the Ti-Mo films to enhance their suitability for the desired applications, such as anodic oxidation and subsequent nanotube array growth.
It is possible to see a significant increase of deposition rate at short target–sample distances (Figure 2A) but even the highest deposition rates do not sufficiently inhibit the formation of columnar structures (see Figure 3A). As a result, a shapeless oxide—instead of nanotube array—was obtained later during oxidation.
The application to the sample of a negative bias up to 100 V, while the other sputtering parameters were kept unaltered, changed both the structure and composition (Figure 2C). The relatively light Ti atoms are more prone to re-sputtering [31,33] by impinging ions than the heavier Mo atoms; therefore, increasing the bias voltage could reduce the relative amount of Ti in the coating. However, the main purpose of bias application is to suppress the columnar structure, as described above. The optimal combination of composition, deposition rate, and film structure was obtained at −75 V bias. An indication of the difference in structure is shown in Figure 3A,B, where the layers without biasing show a sharper boundary between individual columnar grains compared to those with applied bias. Moreover, in the following text, we show that we were able to prepare nanotubes on precursor layers prepared with bias, which provides additional indirect evidence of nanopores suppression. The samples deposited in the set-up with a 12 mm Mo strip with applied bias shows a composition of Ti0.5Mo0.5 with negligible oxygen content, according to the EDS analysis.
Figure 3D shows the grazing incidence X-ray diffraction (GIXRD) pattern of the precursor film, where several peaks are visible. Blue vertical lines indicate the database position of peaks for molybdenum (COD 1512521) and the green lines for titanium (COD 96-151-2548). In both cases, it is a cubic system with space group Im–3m. The measured peaks are in between these two phases; this indicates the presence of a single phase of TixMo1−x, probably a solid solution, which corresponds with the binary phase diagram [11]. No observable peak splitting indicates the presence of a single phase. From the position of peaks and lattice parameters for Mo (a = 3.1472 Å) and Ti (a = 3.3065 Å), the lattice parameter of TixMo1−x can be determined according to Vegard’s law as:
a T i x M o 1 x = x a T i + 1 x a M o = 0.5 × 3.3065 + ( 1 0.5 ) × 3.1472 = 3.2269   Å .
This is in fair correspondence with the EDS analysis shown in Figure 3C and confirms the formation of the single phase in the binary alloy. On this basis, we deposited a set of precursors with various compositions for anodic oxidation. The modification of the metals ratio was achieved by changing the width of the molybdenum strip. For our particular system, the width range was 10 to 16 mm. The rest of the deposition parameters were kept unchanged. The precursors’ concentrations are shown in Table 1.
Precursor samples with concentrations of molybdenum of 34.2 at% (A), 52.7 at% (B), 63.7 at% (C), and 73.2 at% (D) showed nanotube formation after oxidation under the conditions described in Section 2.2. The top view and cross-sections are shown in Figure 4. The precursor films were not oxidized down to the substrate to avoid possible delamination; therefore, some residual Ti-Mo layer was still present under the nanotube array. As can be seen from Table 1, the Ti-Mo ratio remained almost unchanged after oxidation and the samples consisted of well-defined nanotubes with a diameter of 50 nm ± 5% (A, B, C). On the other hand, samples with high concentrations of Mo (D, E) showed significant changes. The mechanism of Ti-O nanotube formation is well known [37,38], and similar behavior is expected for nanotubes with lower concentrations of Mo. The growth includes two competitive processes:
Formation of oxide layer:
Ti + 2H2O → TiO2 + 4H
Dissolution of oxide layer:
TiO2 + 6F + 4H → [TiF6]2− + 2H2O
F-ions are attracted to the anode by applied positive voltage; therefore, reactions proceed in a defined direction, forming a nanotubular array. According to Figure 4, the Ti-driven mechanism works up to a concentration of approximately 60 at% of molybdenum. At higher concentrations of molybdenum, reactions with Mo become prevalent.
Mo + 3F2 → MoF6.
MoF6 + 3H2O → MoO3 + 6HF
In the second case, a significant amount of volatile oxide MoO3 was created, resulting in a decrease of Mo content, and well-defined nanotubes were not formed.
For a detailed investigation of nanotube structure, thin lamellas were prepared on sample B (with a thickness of less than 80 nm, aiming for single-nanotube thickness) using the FIB tool. Both micrographs in Figure 5A,B present a homogenous structure without any apparent grain formation The nanobeam electron diffraction (NBD) pattern, probing only the tube array, is dominated by the broad diffuse maxima typical of amorphous material (Figure 5C). The amorphous nature of the as-grown/oxidized nanotube array is also confirmed by the GIXRD pattern (Figure 6, black). The XRD pattern does not show any other notable maxima different from the precursor pattern (Figure 6, blue). A detailed study using STEM-EDS showed the uniform distribution of Ti and Mo across the nanotube, even at then a no meter scale (Figure 7). The presence of fluorine at the inner edges of the nanotubes is also visible. So, it is possible to assume that the influence of the bottom not-oxidized layer does not significantly contribute to the results of the SEM-EDS analysis.

4. Conclusions

In this study, we presented a new method for the fabrication of an Mo-alloyed TiO nanotube array on the basis of precursors deposited by magnetron sputtering with partial target covering. In our method, nanotubes were grown directly from a Ti-Mo solid state solution—contrary to the methods of other publications—which makes a high dopant content in the lattice possible.
As a precursor, 500 nm thick Ti-Mo films with Mo concentrations in the range of 32 at%–83 at% were prepared. Although the columnar character of the precursor was suppressed by applying a negative bias during deposition, X-ray analysis of the films confirmed the formation of a crystalline BCC phase.
Ti1−xMoxO2 nanotubes with an average diameter of 50 nm ± 5% and an average length of up to 700 nm ± 10% were obtained during an anodization process that lasted 30 min and used ethylene glycol electrolyte containing 0.5 wt% NH4F at 40 V.
The grown nanotubes kept the Ti-Mo ratio close to that of precursor values, up to Ti0.5Mo0.5O2. Precursor samples with higher concentrations of Mo did not form nanotubular structures after the anodic oxidation process.
A deeper examination of the nanostructure showed a homogeneous distribution of Ti and Mo in the formed nanotubes, which confirmed the well-controlled synthesis processes. In addition, the mutual interaction between both metals and oxygen led to the formation of amorphous oxide phases.
In the case of films with a higher concentration of molybdenum (>72 at%), anodic oxidation processes did not lead to the formation of nanotubes.
NBD and XRD analysis did not show any notable formation of crystal structures of titanium molybdenum oxide phase.

Author Contributions

Experimental part H.M.; design of experiment and methodology L.S.: experimental M.V.; XRD analysis T.R.; methodology G.P.; deposition technics and funding acquisition M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by an Operational Program Integrated Infrastructure (Project No. ITMS 313011AUH4) and this publication is the result of support under the Operational Program Integrated Infrastructure for the project: Up Scale of Comenius University Capacities and Competence in Research, Development and Innovation (USCCCORD, ITMS2014+: 313021BUZ3), co-financed by the European Regional Development Fund.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Lofrano, G.; Ubaldi, F.; Albarano, L.; Carotenuto, M.; Vaiano, V.; Valeriani, F.; Libralato, G.; Gianfranceschi, G.; Fratoddi, I.; Meric, S.; et al. Antimicrobial Effectiveness of Innovative Photocatalysts: A Review. Nanomaterials 2022, 12, 2831. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, Y.; Ling, Y.; Wang, G.; Liu, T.; Wang, F.; Zhai, T.; Tong, Y.; Li, Y. Photohole Induced Corrosion of Titanium Dioxide: Mechanism and Solutions. Nano Lett. 2015, 15, 7051–7057. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, S.; Bai, L.N.; Sun, H.M.; Jiang, Q.; Lian, J.S. Structure and photocatalytic property of Mo-doped TiO2 nanoparticles. Powder Technol. 2013, 244, 9–15. [Google Scholar] [CrossRef]
  5. Guo, M.; Gao, Y. ShaoIntrinsic properties of high-aspect ratio single- and double-wall anodic TiO2 nanotube layers annealed at different temperatures. Phys. Chem. Chem. Phys. 2016, 18, 2818–2829. [Google Scholar] [CrossRef]
  6. Xue, C.; Hu, S.; Chang, Q.; Li, Y.; Li, N.; Li, C.; Yang, J. TiCr alloy anodization for Cr-doped TiO2 nanotube array with improved photocatalytic activity. Mater. Res. Express 2019, 6, 075014. [Google Scholar] [CrossRef]
  7. Motola, M.; Satrapinskyy, L.; Čaplovicová, M. Intrinsic properties of high-aspect ratio single- and double-wall anodic TiO2 nanotube layers annealed at different temperatures. Appl. Surf. Sci. 2018, 434, 1257–1265. [Google Scholar] [CrossRef]
  8. Li, X.; Wu, Y.; Shen, Y.; Sun, Y.; Yang, Y.; Xie, A. A novel bifunctional Ni-doped TiO2 inverse opal with enhanced SERS performance and excellent photocatalytic activity. Appl. Surf. Sci. 2018, 427, 739–744. [Google Scholar] [CrossRef]
  9. Tu, Y.-F.; Huang, S.-Y.; Sang, J.-P.; Zou, X.-W. Preparation of Fe-doped TiO2 nanotube arrays and their photocatalytic activities under visible light. Mater. Res. Bull. 2010, 45, 224–229. [Google Scholar] [CrossRef]
  10. Zaleska, A. Doped-TiO2: A review. Recent. Pat. Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  11. Ghicov, A.; Schmuki, P. Self-ordering electrochemistry: A review on growth and functionality of TiO2nanotubes and other self-aligned MOx structures. Chem. Commun. 2009, 20, 2791–2808. [Google Scholar] [CrossRef] [PubMed]
  12. Lalabadi, M.A.; Ehsani, A.; Divband, B.; Sani, M.A. Antimicrobial activity of Titanium dioxide and Zinc oxide nanoparticles supported in 4A zeolite and evaluation the morphological characteristic. Sci. Rep. 2019, 9, 17439. [Google Scholar] [CrossRef] [Green Version]
  13. Baker, H. Alloy Phase Diagrams ASM Handbook; ASM International: Materials Park, OH, USA, 1992; Volume 3. [Google Scholar]
  14. Luo, S.-Y.; Yan, B.-X.; Shen, J. Enhancement of photoelectric and photocatalytic activities: Mo doped TiO2 thin films deposited by sputtering. Thin Solid. Film. 2012, 522, 361–365. [Google Scholar] [CrossRef]
  15. Bakhtchadjian, R.; Manucharova, L.A.; Tavadyan, L.A. Selective Oxidation of DDT by Dioxygen on the Dioxo-Mo(VI) Complex Anchored on a TiO2 under UV-Irradiation. Catal. Commun. 2015, 69, 193. [Google Scholar]
  16. Devi, L.G.; Kumar, S.G.; Murthy, B.N.; Kottam, N. Influence of Mn2+ and Mo6+ dopants on the phase transformations of TiO2 lattice and its photo catalytic activity under solar illumination. Catal. Commun. 2009, 10, 794–798. [Google Scholar] [CrossRef]
  17. Devi, L.G.; Murthy, B.N. Mechanism of Charge Transfer in the Transition Metal Ion Doped TiO2 with Bicrystalline Framework of Anatase and Rutile: Photocatalytic and Photoelectrocatalytic Activity. Catal. Lett. 2008, 125, 320–330. [Google Scholar]
  18. Huang, J.; Guo, X.; Wang, B.; Li, L.; Zhao, M.; Dong, L.; Liu, X.; Huang, Y. Synthesis and Photocatalytic Activity of Mo-Doped TiO2 Nanoparticles. J. Spectrosc. 2015, 1–8. [Google Scholar] [CrossRef] [Green Version]
  19. Houng, B.; Liu, C.; Hung, M.T. Structural, electrical and optical properties of molybdenum-doped TiO2 thin films. Ceram. Int. 2013, 39, 3669–3676. [Google Scholar] [CrossRef]
  20. Masuda, H.; Yasui, K.; Sakamoto, Y.; Nakao, M.; Tamamura, T.; Nishio, K. Ideally Ordered Anodic Porous Alumina Mask Prepared by Imprinting of Vacuum-Evaporated Al on Si. Jpn. J. Appl. Phys. 2001, 40, L1267. [Google Scholar] [CrossRef]
  21. Motola, M.; Sopha, H.; Krbal, M.; Hromádko, L.; OlmrováZmrhalová, Z.; 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]
  22. 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]
  23. Xia, X.-H.; Jia, Z.-J.; Yu, Y.; Wang, Y.L.Z.; Ma, L.-L. Preparation of multi-walled carbon nanotube supported TiO2 and its photocatalytic activity in the reduction of CO2 with H2O. Carbon 2007, 45, 717–721. [Google Scholar] [CrossRef]
  24. Roy, P.; Das, C.; Lee, K.; Hahn, R.; Ruff, T.; Moll, M.; Schmuki, P. Oxide nanotubes on Ti-Ru alloys: Strongly enhanced and stable photoelectrochemical activity for water splitting. J. Am. Chem. Soc. 2011, 133, 5629–5631. [Google Scholar] [CrossRef] [PubMed]
  25. Morandi, S.; Paganini, M.C.; Giamello, E.; Bini, M.; Capsoni, D.; Massarotti, V.; Ghiotti, G.J. Structural and spectroscopic characterization of Mo1−xWxO3−δ mixed oxides. Solid. State Chem. 2009, 18, 3342–3352. [Google Scholar] [CrossRef]
  26. Elbanna, A.M.; Salem, K.E.; Mokhtar, A.M.; Ramadan, M.; Motaweh, H.A.; Tourk, H.M.; Gepreel, M.A.-H.; AllamTernary, N.K. Ternary Ti–Mo–FeNanotubes as Efficient Photoanodes for Solar-Assisted Water Splitting. J. Phys. Chem. C 2021, 125, 12504–12517. [Google Scholar] [CrossRef]
  27. Hamedani, H.A.; Allam, N.K.; Garmestani, H.; El-Sayed, M.A. Electrochemical Fabrication of Strontium-Doped TiO2 Nanotube Array Electrodes and Investigation of Their Photoelectrochemical Properties. J. Phys. Chem. C 2011, 115, 13480–13486. [Google Scholar] [CrossRef]
  28. Ghicov, A.; Aldabergenova, S.; Tsuchyia, H.; Schmuki, P. TiO2-Nb2O5 nanotubes with electrochemically tunable morphologies. Angew. Chem. Int. Ed. 2006, 45, 6993–6996. [Google Scholar] [CrossRef]
  29. Choi, W.; Termin, A.; Hoffmann, M.R. The Role of Metal Ion Dopants in Quantum-Sized TiO2: Correlation between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 1994, 98, 13669–13679. [Google Scholar] [CrossRef]
  30. Samu, G.F.; Veres, A.; Tallósy, S.z.P.; Janovák, L.; Dékány, I.; Yepez, A.; Luque, R. Janáky, Photocatalytic, photoelectrochemical, and antibacterial activity of benign-by-design mechanochemically synthesized metal oxide nanomaterials. Catal. Today 2017, 284, 1–3. [Google Scholar] [CrossRef]
  31. Petrisková, P.; Monfort, O.; Satrapinskyy, L.; Dobročka, E.; Plecenik, T.; Plesch, G.; Papšík, G.; Bermejo, R.; Lenčéš, Z. Preparation and photocatalytic activity of TiO2 nanotube arrays prepared on transparent spinel substrate. Ceram. Int. 2021, 47, 12970–12980. [Google Scholar] [CrossRef]
  32. Krishna, S. Handbook of Thin Film Deposition, 3rd ed.; William Andrew Elsevier: Waktham, MA, USA, 2012; pp. 55–88. [Google Scholar]
  33. Habazaki, H.; Fushimi, K.; Shimizu, K.; Skeldon, P.; Thompson, G.E. Fast migration of fluoride ions in growing anodic titanium oxide. Electrochem. Commun. 2007, 9, 1222–1227. [Google Scholar] [CrossRef] [Green Version]
  34. Petrov, I.; Barna, P.B.; Hultman, L.; Greene, J.E. Microstructural evolution during film growth. J. Vac. Sci. Technol. A Vac. Surf. Film. 2003, 21, S117–S128. [Google Scholar] [CrossRef]
  35. Anders, A. A structure zone diagram including plasma-based deposition and ion etching. Thin Solid. Film. 2009, 518, 4087–4090. [Google Scholar] [CrossRef] [Green Version]
  36. Müller, K.-H. Role of incident kinetic energy of adatoms in thin film growth. Surf. Sci. 1987, 184, L375–L382. [Google Scholar] [CrossRef]
  37. 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, 28, 59–64. [Google Scholar]
  38. Roy, P.; Berger, S.; Schmuki, P. TiO2 Nanotubes: Synthesis and Applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of preparation of Mo-doped TiO2 nanotubes: (A) DC-magnetron sputtering, (B) anodic oxidation, (C) annealing.
Figure 1. Schematic diagram of preparation of Mo-doped TiO2 nanotubes: (A) DC-magnetron sputtering, (B) anodic oxidation, (C) annealing.
Coatings 13 01309 g001
Figure 2. The influence of sample–target distance on: (A) deposition rate; (B) Ti-Mo ratio; (C) the influence of bias on the Ti-Mo ratio.
Figure 2. The influence of sample–target distance on: (A) deposition rate; (B) Ti-Mo ratio; (C) the influence of bias on the Ti-Mo ratio.
Coatings 13 01309 g002
Figure 3. (A) Structure of a films deposited by magnetron sputtering without an applied bias; (B) Structure of a film deposited by magnetron sputtering with applied bias −75 V; (C) EDS (spectrum corresponds to image (B); (D) XRD pattern (corresponds to image (B)).
Figure 3. (A) Structure of a films deposited by magnetron sputtering without an applied bias; (B) Structure of a film deposited by magnetron sputtering with applied bias −75 V; (C) EDS (spectrum corresponds to image (B); (D) XRD pattern (corresponds to image (B)).
Coatings 13 01309 g003
Figure 4. Top view SEM images of samples (A1D1) and cross-sections (A2D2). Concentrations are shown in Table 1.
Figure 4. Top view SEM images of samples (A1D1) and cross-sections (A2D2). Concentrations are shown in Table 1.
Coatings 13 01309 g004
Figure 5. Overview of STEM micrographs of the lamella (sample B in Table 1) in bright field mode (A) and HAADF mode (B); nanobeam diffraction pattern from the tube array (C).
Figure 5. Overview of STEM micrographs of the lamella (sample B in Table 1) in bright field mode (A) and HAADF mode (B); nanobeam diffraction pattern from the tube array (C).
Coatings 13 01309 g005
Figure 6. Comparison of XRD analyses of precursor (blue) as-anodized MTO (black) and annealed (red) (sample with concentration of molybdenum 53 at%).
Figure 6. Comparison of XRD analyses of precursor (blue) as-anodized MTO (black) and annealed (red) (sample with concentration of molybdenum 53 at%).
Coatings 13 01309 g006
Figure 7. STEM-EDS spectrum and distribution of Ti, Mo, O, and F across the nanotube.
Figure 7. STEM-EDS spectrum and distribution of Ti, Mo, O, and F across the nanotube.
Coatings 13 01309 g007
Table 1. Composition of precursors and oxidized and nanotube arrays measured by SEM-EDS.
Table 1. Composition of precursors and oxidized and nanotube arrays measured by SEM-EDS.
Ti
at%
Mo
at%
O
at%
C
at%
F
at%
Ti/Mo Ratio
sample Aprecursor65.834.2 1.9
oxidized19.09.657.96.57.02.0
sample Bprecursor47.352.7 0.9
oxidized13.614.555.812.73.10.9
sample Cprecursor36.363.7 0.6
oxidized17.216.556.16.33.61.0
sample Dprecursor26.873.2 0.4
oxidized14.125.149.79.41.20.6
Sample Eprecursor15.484.5 0.2
oxidized10.417.960.09.32.40.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Makarov, H.; Satrapinskyy, L.; Vidiš, M.; Roch, T.; Plesch, G.; Mikula, M. Ti-Mo-O Nanotube Arrays Grown by Anodization of Magnetron Sputtered Films. Coatings 2023, 13, 1309. https://doi.org/10.3390/coatings13081309

AMA Style

Makarov H, Satrapinskyy L, Vidiš M, Roch T, Plesch G, Mikula M. Ti-Mo-O Nanotube Arrays Grown by Anodization of Magnetron Sputtered Films. Coatings. 2023; 13(8):1309. https://doi.org/10.3390/coatings13081309

Chicago/Turabian Style

Makarov, Hryhorii, Leonid Satrapinskyy, Marek Vidiš, Tomáš Roch, Gustáv Plesch, and Marian Mikula. 2023. "Ti-Mo-O Nanotube Arrays Grown by Anodization of Magnetron Sputtered Films" Coatings 13, no. 8: 1309. https://doi.org/10.3390/coatings13081309

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