Next Article in Journal
Can a Fraction of Flour and Sugar Be Replaced with Fruit By-Product Extracts in a Gluten-Free and Vegan Cookie Recipe?
Next Article in Special Issue
Functional Photocatalysts: Material Design, Synthesis and Applications
Previous Article in Journal
Novel Scaffold Agonists of the α2A Adrenergic Receptor Identified via Ensemble-Based Strategy
Previous Article in Special Issue
Synthesis, Structural Characterization, Hirschfeld Surface Analysis and Photocatalytic CO2 Reduction Activity of a New Dinuclear Gd(III) Complex with 6-Phenylpyridine-2-Carboxylic Acid and 1,10-Phenanthroline Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization

1
Institute of Physics and Technology, Almaty 050032, Kazakhstan
2
Institute of Nuclear Physics, Almaty 050032, Kazakhstan
3
Department of Materials Science, Nanotechnology and Engineering Physics, Satbaev University, Almaty 050032, Kazakhstan
4
Department of Mechanics and Mechanical Engineering, M.Kh. Dulaty Taraz Regional University, Taraz 080000, Kazakhstan
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(5), 1101; https://doi.org/10.3390/molecules29051101
Submission received: 24 January 2024 / Revised: 23 February 2024 / Accepted: 23 February 2024 / Published: 29 February 2024

Abstract

:
Layers of TiO2 nanotubes formed by the anodization process represent an area of active research in the context of innovative energy conversion and storage systems. Titanium nanotubes (TNTs) have attracted attention because of their unique properties, especially their high surface-to-volume ratio, which makes them a desirable material for various technological applications. The anodization method is widely used to produce TNTs because of its simplicity and relative cheapness; the method enables precise control over the thickness of TiO2 nanotubes. Anodization can also be used to create decorative and colored coatings on titanium nanotubes. In this study, a combined structure including anodic TiO2 nanotubes and SrTiO3 particles was fabricated using chemical synthesis techniques. TiO2 nanotubes were prepared by anodizing them in ethylene glycol containing NH4F and H2O while applying a voltage of 30 volts. An anode nanotube array heat-treated at 450 °C was then placed in an autoclave filled with dilute SrTiO3 solution. Scanning electron microscopy (SEM) analysis showed that the TNTs were characterized by clear and open tube ends, with an average outer diameter of 1.01 μm and an inner diameter of 69 nm, and their length is 133 nm. The results confirm the successful formation of a structure that can be potentially applied in a variety of applications, including hydrogen production by the photocatalytic decomposition of water under sunlight.

1. Introduction

Rapid growth in the world population has increased the demand for energy, the bulk of which is provided by fossil fuels for power generation, industrial needs, and transportation [1,2,3,4]. However, in addition to limited availability, the use of fossil fuels has a negative impact on the environment, creating by-products such as carbon, nitrogen, and sulfur oxides [5,6]. Therefore, there is an urgent need to develop cleaner alternative energy sources that are sustainable and have a minimal impact on the environment [7,8]. Hydrogen stands out as a clean and efficient energy source, and its production is becoming an important challenge in the field of sustainable energy sources.
Water is an abundant source of hydrogen, but given the need to introduce energy to overcome the energy barrier associated with chemical stability, it is difficult to separate water into stoichiometric hydrogen and oxygen on an industrial scale. Nevertheless, one method of hydrogen production is the photocatalytic splitting of water. Photocatalysis can efficiently utilize solar energy to split water into its individual elements [9,10,11]. This is a unique and promising method of hydrogen production based on the use of solar energy to convert water into hydrogen and oxygen [12]. This process could be an important step toward sustainable energy sources, as it combines the efficiency of solar panels with the ability to produce clean hydrogen. In particular, photocatalysis has been shown to be a more efficient form of wastewater treatment because of the impressive efficiency of photocatalytic removal, rapid oxidation process, lower costs, and a lack of toxicity [13]. Photocatalytic water splitting involves the use of semiconductors as photocatalysts. The most studied photocatalysts are TiO2, ZnO, CdS, and SrTiO3, which are used for various photocatalytic applications including photocatalytic water splitting [14,15,16]. To achieve efficient photocatalytic water splitting, a sophisticated photocatalyst is required that can overcome problems in the water oxidation process.
Titanium dioxide (titanium, TiO2) is considered the most promising and versatile material. Over the past decades, TiO2 has been extensively investigated in various fields because of its unique properties such as outstanding corrosion resistance, high biocompatibility, suitable bandgap for water splitting, and stable physicochemical characteristics [17,18]. The narrow– bandgap facilitates the more efficient collection of solar energy, making it an ideal material for creating an electron–hole pair. This pair is actively involved in redox reactions and finds applications in various fields, such as dyes, food, biomedicine, photocatalysis, photodegradation of water, photosensitive materials, dye-sensitized solar cells, and gas-sensitive devices. In recent years, significant research efforts have been devoted to the development of new nanomaterials, including nanostructured titanium obtained via anodization, sol–gel, hydrothermal treatment, and vapor deposition techniques [19]. Nowadays, a wide range of materials are required to develop and research advanced devices suitable for various commercial applications. Nanomaterials play a key role in emerging technologies, enabling the creation of high-performance devices [20,21]. The performance of such devices is largely determined by the geometry, shape, and morphology of the nanostructures [7]. The exponential growth in the literature indicates that interest in the nanoscale began in the 1990s. Interest in the nanoscale is driven by the commercial availability of tools used to manipulate and measure nanoscale characteristics for several reasons: (1) the anticipation of the novel physical, chemical, and biological properties of nanostructures; (2) the assumption that nanostructures will provide new building blocks for innovative materials with unique properties; (3) the miniaturization of the semiconductor industry to the nanoscale; and (4) the recognition that molecular mechanisms in biological cells function at the nanoscale [22].

2. Results and Discussion

The morphology of the obtained SrTiO3 samples was studied using scanning (SEM) and transmission (TEM) electron microscopes at different resolutions. The scanning electron microscopy results (Figure 1a–c) show that the SrTiO3 particles that calcined at 900 °C possess cubic shapes and have sizes ranging from 150 nm to 300 nm. Calcination at 800 °C leads to the formation of finer particles but with more significant numbers of impurities such as SrCO3. Based on the literature and experimental data [23,24], the optimal calcination temperature is 900 °C, which is followed by treatment in 1 m nitric acid solution to remove residual SrCO3. However, it is worth noting that the particle sizes are highly heterogeneous. Given studies in the literature, doping SrTiO3 with other elements, such as Al or Mn, can contribute to the size reduction and distortion in SrTiO3‘s crystal shape.
In the case of TEM, clearly formed cubes of SrTiO3 with anisotropic structures with an average size of about 200 nm are clearly visible, as shown in Figure 1d–f. An important feature of these particles is the anisotropic structure, which creates a difference in energy at different faces, leading to the formation of p–n junctions. This allows the charge within each photocatalyst particle to be separated using an inter-domain electric field. Thus, electrons are concentrated on some faces and holes on other faces, which provides for the separation of photocatalytic reduction and oxidation processes on different faces. Given the anisotropic crystal structure, the selective deposition of catalysts takes place, which leads to the release of hydrogen and oxygen on the faces of the cubic photocatalyst.
Among various nanostructured oxide materials, TiO2 nanotubes have been emphasized because of their improved properties, economical design, and higher surface-to-volume ratio [25]. TNTs with high specific surface areas, ion exchange abilities, and photocatalytic properties have been considered for various potential applications and can be excellent candidates as catalysts in photocatalysis [26]. Figure 2 shows images of the top surfaces of the anodized TiO2 nanotube samples before and after the deposition of SrTiO3 on their surfaces. The top surfaces of the anodized and annealed TNTs at 450 °C shown in Figure 2a show well-defined tubes with open ends that form a hexagonal order. This is typical of anodization, as previously noted in [27]. After applying SrTiO3, pronounced morphological changes are observed with the presence of interface regions between SrTiO3 and TNT, which indicates the success of the combination (Figure 2b). During 6 h of autoclave treatment, the surface showed a tendency to be coated with nanoparticles, and uneven deposition was also found. Agglomerates are formed on the surface, and round holes corresponding to TNT are still visible. Note that increasing the treatment time to 6 h significantly affects the surface morphology, leading to the formation of larger agglomerates and the blocking of the tube tops [28]. Figure 2c shows that the initial TNTs have an average outer diameter of 1 μm, while Figure 2a shows an inner diameter of 69 nm. The length of the tubes is 133 nm, as seen in the inset. A cross-sectional view of a freestanding titanium dioxide membrane with an average thickness of more than 50 nm is shown in Figure 2d, mechanically collapsed for visualization. In a related study [25], Paulose et al. obtained nanotubes measuring 360 μm in length over a 96 h period, utilizing a voltage of 60 V. They employed a titanium foil with a thickness of 0.25 mm, immersed in a solution comprising 0.3 wt% NH4F and 2% H2O in ethylene glycol. Our results—derived from anodization in a solution comprising 0.7 wt% NH4F and 3.5 wt% distilled water at 30 V—revealed the length of the TNT nanotubes to be 133 nm at the nanoscale. The SEM images also demonstrate that the obtained nanotubes are ordered and have clear open ends. Despite the low voltage (30 V), we compensated for this by increasing the concentrations of NH4F and H2O in the anode solution. Given the higher mass percentage of NH4F, compensation is accomplished by increasing the concentration of H2O, resulting in faster growth and, hence, longer nanotube lengths.
XRD analysis of the TNT@SrTiO3 samples was performed on an X-ray diffractometer with detection unit rotation angles ranging from 20° to 80° and a minimum detection unit movement step of 0.01, as shown in Figure 3a. The characteristic peaks of the TNT samples appear at 2θ 25.4°, 37.9°, and 53.4°, 71.5°, indicating the polycrystalline structure of the anatase, in good agreement with the standard map for TNT (JCPDS map 1286) [20]. In addition, the appearance of new peaks at 32.2°, 46.9°, and 57.8° in the X-ray diffraction spectrum of the TNT@SrTiO3 samples indicates the combination of two components in the composite, which additionally proves the successful connection and interaction between the components. This confirmation is based on a comparison of the diffraction spectra of the composite with TNT, which makes it possible to determine whether changes have occurred in the crystal structure during their combination. It is particularly important to note that the peak at 2θ, equal to 71.23°, has a high intensity, indicating the high crystallinity of the semiconductor. This is significant because the transport efficiency of charged carriers generated during photogeneration can be strongly dependent on the crystallinity of the material. Low crystallinity can lead to the inefficient migration of charged particles. In addition, semi-quantitative elemental analysis of the particles confirmed the composition of the obtained samples. The presence of the elements Ti and Sr was confirmed without detecting other impurities. According to the atomic percentages in the index (Figure 3b), it can be established that Ti/Sr is 81.10%/18.90%, respectively. These results confirm that the designs contain the expected elements and have no significant impurities.
Low-temperature electron paramagnetic resonance (EPR) spectra were determined on the SrTiO3/TiO2 samples to confirm the presence of Ti3+ and oxygen vacancies. The initial SrTiO3/TiO2 (Figure 4c, marked in red), containing mainly Ti4+ 3d0 states, exhibits a weak EPR signal, which may be due to the surface adsorption of O2 from air. For SrTiO3 and TiO2, a strong signal from Ti3+ spins (marked in blue and black) is also observed. It is generally believed that photoelectrons can be captured by Ti4+ and lead to the reduction of Ti4+ cations to the Ti3+ state, which is usually accompanied by the loss of oxygen from the surface of TiO2 and SrTiO3. Thus, these data clearly confirm that Ti3+ and oxygen vacancies were formed in all SrTiO3/TiO2, TiO2, and SrTiO3 samples.
TiO2 nanotubes can be produced in various ways [29], among which, the most widely studied is the use of electrochemical anodization. The advantage of anodic TiO2 nanotubes over TiO2 nanotubes produced by other methods is their availability and cost-effectiveness. Also, one of the advantages of this method is that the anodic TiO2 nanotubes grow vertically on the Ti substrate with nanotube holes on top and closed nanotube bottoms attached to the Ti substrate. Thus, no further immobilization on the substrate is required. The TNT layers are highly ordered, which favors a direct diffusion pathway. In addition, the nanotube layers can be removed from the Ti substrate and used as powders if required. Another advantage is that the nanotube layer thickness and nanotube diameter can be controlled by adjusting the anodization electrolyte, potential, and time [30].

3. Materials and Methods

3.1. Materials

Ti foil (99.9%; thickness, 0.1 mm; China), ethanol (45%), ethylene glycol (99.9%, Russia), ammonium fluoride, and sodium nitrate (70%) were used without further purification. Distilled water was used as a solvent in all experiments.

3.2. Synthesis of SrTiO3

SrTiO3 was obtained using a chemical precipitation method [24,31,32,33]. For this purpose, 2.54 g of Sr (NO3)2 was mixed with 100 mL of distilled water; then, 0.958 g of TiO2 was added in a 1:1 ratio of Ti and SrTiO3 to this solution. The solution was then treated for 30 min in an ultrasonic bath. The solution was gradually added while maintaining vigorous stirring, and the pH of the mixture was brought to 6–7 using 10% NH3OH solution. The suspension was washed several times with distilled water. The resulting powder was dried at 60 °C overnight and then calcined at 900 °C for 1 h.

3.3. Nanotube Synthesis

TiO2 was obtained using an anodization method. The 0.1 mm thick Ti foil was initially cut into 1 cm × 6 cm samples and mechanically polished with P150 sandpaper. The sheets were then ultrasonically treated in sodium nitrate, ethanol, and distilled water for final cleaning. Electrochemical anodization experiments were carried out in a two-electrode electrochemical cell, where titanium foil served as the working electrode and a sheet of nickel foil as the counter electrode at constant potential and room temperature (≈22 °C). Figure 1 shows a schematic of the titanium nanotube formation process. A constant current power supply unit model, UNI-T UTP3315TPL from UNI-TREND Technology, China, was used. This unit was used as a voltage source to control the anodization. The electrolyte for anodizing consisted of ethylene glycol with 0.7 wt% NH4F and 3.5 wt% distilled water added. The anodization process was carried out at 30 V for 96 h at room temperature. The anodized titanium nanotube samples were then placed in ethylene glycol and subjected to ultrasonic stirring until the nanotube film separated from the titanium substrate. The suspension was filtered; the residue was washed several times with distilled water. The resulting powder was dried at 60 °C for 3 h and then calcined at 450 °C for 1 h.

3.4. Synthesis of TNT@SrTiO3

To create the combined TNT@ SrTiO3 structure, powders of 0.2 g of TNT and 0.1 g of SrTiO3 were taken, mixed with 40 mL of distilled water, and placed in a stainless autoclave. The sealed autoclave was heated to 90 °C and incubated for 6 h. At the end of the experiment, the autoclave was cooled to room temperature. The samples were then washed with distilled water and dried in an oven for 5 h at 60 °C.

3.5. Material Characterization Techniques

The morphologies of the TNT and the combined TNT@SrTiO3 were analyzed using a JSM-6490LA scanning electron microscope from JEOL, Tokyo, Japan. TESCAN MAIA3 XMU scanning transmission electron microscopy (STEM) was used to further investigate the morphology at high resolution. The crystal structure of the samples was studied using a Drone-8 X-ray diffractometer. An EPR spectrometer “JEOL” (JES-FA200, Japan) was also used. Measurements were in ranges of ~9.4 GHz (X-Band) and ~35 GHz (Q-Band). Microwave frequency stability—~10− 6. Sensitivity—7 × 109/10− 4 Tl. Resolution—2.35 μT. Output power—from 200 mW to 0.1 μW. Quality factor (Q-factor)—18,000.

4. Conclusions

In this paper, the synthesis of arrays of TiO2 nanotubes using an electrochemical anodization method was successfully demonstrated. The obtained nanotubes have clear and open ends and are 133.9 nm long, and their membranes are more than 1 μm thick. The anodization process of 0.1 mm thick Ti foil at 450 °C can easily produce such nanotube arrays. SEM analysis showed that the TNTs are characterized by clear and open tube ends, with an average outer diameter of 1 μm and an inner diameter of 69 nm, and their length is 133 nm. In addition, a combined structure of TNT@SrTiO3 was fabricated in this study using chemical autoclave synthesis techniques. X-ray phase analysis confirmed the high crystallinity and orientation of crystallites along the preferential growth direction, indicating the successful formation of the structure. The results obtained here have potential significance for various fields including the sunlight-induced photocatalytic decomposition of water and other applications in energy conversion and storage. Further research and development in this area can contribute to the development of innovative technologies and improve the efficiency of energy systems.

Author Contributions

Conceptualization, methodology, writing—original draft preparation, investigation: M.B.; resources, visualization, project administration, funding acquisition: A.U., K.M., A.M. and Y.Y.; formal analysis: A.S.; supervision, data curation, writing—review and editing: Z.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science Committee of the Ministry of Science and Higher Education and of the Republic of Kazakhstan (Grant No AP19680604).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Zhang, Z.; Wang, Q.; Xu, H.; Zhang, W.; Zhou, Q.; Zeng, H.; Yang, J.; Zhu, J.; Zhu, X. TiO2 Nanotube Arrays with a Volume Expansion Factor Greater than 2.0: Evidence against the Field-Assisted Ejection Theory. Electrochem. Commun. 2020, 114, 106717. [Google Scholar] [CrossRef]
  2. Prikhodko, N.; Yeleuov, M.; Abdisattar, A.; Askaruly, K.; Taurbekov, A.; Tolynbekov, A.; Rakhymzhan, N.; Daulbayev, C. Enhancing Supercapacitor Performance through Graphene Flame Synthesis on Nickel Current Collectors and Active Carbon Material from Plant Biomass. J. Energy Storage 2023, 73, 108853. [Google Scholar] [CrossRef]
  3. Askaruly, K.; Yeleuov, M.; Taurbekov, A.; Sarsembayeva, B.; Tolynbekov, A.; Zhylybayeva, N.; Azat, S.; Abdisattar, A.; Daulbayev, C. A Facile Synthesis of Graphite-Coated Amorphous SiO2 from Biosources as Anode Material for Libs. Mater. Today Commun. 2023, 34, 105136. [Google Scholar] [CrossRef]
  4. Yeleuov, M.; Daulbayev, C.; Taurbekov, A.; Abdisattar, A.; Ebrahim, R.; Kumekov, S.; Prikhodko, N.; Lesbayev, B.; Batyrzhan, K. Synthesis of Graphene-like Porous Carbon from Biomass for Electrochemical Energy Storage Applications. Diam. Relat. Mater. 2021, 119, 108560. [Google Scholar] [CrossRef]
  5. Galstyan, V.; Macak, J.M.; Djenizian, T. Anodic TiO2 Nanotubes: A Promising Material for Energy Conversion and Storage. Appl. Mater. Today 2022, 29, 101613. [Google Scholar] [CrossRef]
  6. Taurbekov, A.; Abdisattar, A.; Atamanov, M.; Yeleuov, M.; Daulbayev, C.; Askaruly, K.; Kaidar, B.; Mansurov, Z.; Castro-Gutierrez, J.; Celzard, A.; et al. Biomass Derived High Porous Carbon via CO2 Activation for Supercapacitor Electrodes. J. Compos. Sci. 2023, 7, 444. [Google Scholar] [CrossRef]
  7. Saddique, Z.; Imran, M.; Javaid, A.; Kanwal, F.; Latif, S.; dos Santos, J.C.S.; Kim, T.H.; Boczkaj, G. Bismuth-Based Nanomaterials-Assisted Photocatalytic Water Splitting for Sustainable Hydrogen Production. Int. J. Hydrogen Energy 2024, 52, 594–611. [Google Scholar] [CrossRef]
  8. Yergaziyeva, G.; Kuspanov, Z.; Mambetova, M.; Khudaibergenov, N.; Makayeva, N.; Daulbayev, C. Advancements in Catalytic, Photocatalytic, and Electrocatalytic CO2 Conversion Processes: Current Trends and Future Outlook. J. CO2 Util. 2024, 80, 102682. [Google Scholar] [CrossRef]
  9. Wang, S.-J.; Su, D.; Zhu, Y.-F.; Lu, C.-H.; Zhang, T. The State-of-the-Art Review on Rational Design for Cavitation Assisted Photocatalysis. Mater. Des. 2023, 234, 112377. [Google Scholar] [CrossRef]
  10. Kuspanov, Z.; Bakbolat, B.; Baimenov, A.; Issadykov, A.; Yeleuov, M.; Daulbayev, C. Photocatalysts for a Sustainable Future: Innovations in Large-Scale Environmental and Energy Applications. Sci. Total Environ. 2023, 885, 163914. [Google Scholar] [CrossRef]
  11. Megbenu, H.K.; Daulbayev, C.; Nursharip, A.; Tauanov, Z.; Poulopoulos, S.; Busquets, R.; Baimenov, A. Photocatalytic and Adsorption Performance of MXene@Ag/Cryogel Composites for Sulfamethoxazole and Mercury Removal from Water Matrices. Environ. Technol. Innov. 2023, 32, 103350. [Google Scholar] [CrossRef]
  12. Baimenov, A.; Montagnaro, F.; Inglezakis, V.J.; Balsamo, M. Experimental and Modeling Studies of Sr2+ and Cs+ Sorption on Cryogels and Comparison to Commercial Adsorbents. Ind. Eng. Chem. Res. 2022, 61, 8204–8219. [Google Scholar] [CrossRef]
  13. Lee, D.-E.; Kim, M.-K.; Danish, M.; Jo, W.-K. State-of-the-Art Review on Photocatalysis for Efficient Wastewater Treatment: Attractive Approach in Photocatalyst Design and Parameters Affecting the Photocatalytic Degradation. Catal. Commun. 2023, 183, 106764. [Google Scholar] [CrossRef]
  14. 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-Corros. 2015, 1, 28. [Google Scholar] [CrossRef]
  15. Kuspanov, Z.; Umirzakov, A.; Serik, A.; Baimenov, A.; Yeleuov, M.; Daulbayev, C. Multifunctional Strontium Titanate Perovskite-Based Composite Photocatalysts for Energy Conversion and Other Applications. Int. J. Hydrogen Energy 2023, 48, 38634–38654. [Google Scholar] [CrossRef]
  16. Bakbolat, B.; Daulbayev, C.; Sultanov, F.; Beissenov, R.; Umirzakov, A.; Mereke, A.; Bekbaev, A.; Chuprakov, I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials 2020, 10, 1790. [Google Scholar] [CrossRef] [PubMed]
  17. Moon, S.; Nagappagari, L.R.; Lee, J.; Lee, H.; Lee, W.; Lee, K. Electrochemical Detection of 2,4,6-Trinitrotoluene Reduction in Aqueous Solution by Using Highly Ordered 1D TiO2 Nanotube Arrays. Mater. Today Commun. 2020, 25, 101389. [Google Scholar] [CrossRef]
  18. An, X.; Hua, W.; Rui, L.; Liu, P.; Li, X. Application of a New Nano-TiO2 Composite Antibacterial Agent in Nursing Management of Operating Room: Based on Real-Time Information Push Assistant System. Prev. Med. 2023, 172, 107541. [Google Scholar] [CrossRef]
  19. Yin, H.; Liu, H.; Shen, W.Z. The Large Diameter and Fast Growth of Self-Organized TiO2 Nanotube Arrays Achieved via Electrochemical Anodization. Nanotechnology 2009, 21, 035601. [Google Scholar] [CrossRef]
  20. Ghani, T.; Mujahid, M.; Mehmood, M.; Zhang, G.; Naz, S. Highly Ordered Combined Structure of Anodic TiO2 Nanotubes and TiO2 Nanoparticles Prepared by a Novel Route for Dye-Sensitized Solar Cells. J. Saudi Chem. Soc. 2019, 23, 1231–1240. [Google Scholar] [CrossRef]
  21. Jandosov, J.; Alavijeh, M.; Sultakhan, S.; Baimenov, A.; Bernardo, M.; Sakipova, Z.; Azat, S.; Lyubchyk, S.; Zhylybayeva, N.; Naurzbayeva, G.; et al. Activated Carbon/Pectin Composite Enterosorbent for Human Protection from Intoxication with Xenobiotics Pb(II) and Sodium Diclofenac. Molecules 2022, 27, 2296. [Google Scholar] [CrossRef]
  22. Zhang, W.; Sun, Y.; Tian, R.; Gao, Q.; Wang, J.; Liu, Y.; Yang, F. Anodic Growth of TiO2 Nanotube Arrays: Effects of Substrate Curvature and Residual Stress. Surf. Coat. Technol. 2023, 469, 129783. [Google Scholar] [CrossRef]
  23. Roy, P.K.; Bera, J. Formation of SrTiO3 from Sr-Oxalate and TiO2. Mater. Res. Bull. 2005, 40, 599–604. [Google Scholar] [CrossRef]
  24. Kudaibergen, A.D.; Kuspanov, Z.B.; Issadykov, A.N.; Beisenov, R.E.; Mansurov, Z.A.; Yeleuov, M.A.; Daulbayev, C.B. Synthesis, Structure, and Energetic Characteristics of Perovskite Photocatalyst SrTiO3: An Experimental and DFT Study. Eurasian Chem.-Technol. J. 2023, 25, 139–146. [Google Scholar] [CrossRef]
  25. Paulose, M.; Prakasam, H.E.; Varghese, O.K.; Peng, L.; Popat, K.C.; Mor, G.K.; Desai, T.A.; Grimes, C.A. TiO2 Nanotube Arrays of 1000 μm Length by Anodization of Titanium Foil:  Phenol Red Diffusion. J. Phys. Chem. C 2007, 111, 14992–14997. [Google Scholar] [CrossRef]
  26. Moulis, F.; Krýsa, J. Photocatalytic Degradation of Several VOCs (n-Hexane, n-Butyl Acetate and Toluene) on TiO2 Layer in a Closed-Loop Reactor. Catal. Today 2013, 209, 153–158. [Google Scholar] [CrossRef]
  27. Hou, X.; Lund, P.D.; Li, Y. Controlling Anodization Time to Monitor Film Thickness, Phase Composition and Crystal Orientation during Anodic Growth of TiO2 Nanotubes. Electrochem. Commun. 2022, 134, 107168. [Google Scholar] [CrossRef]
  28. Sopha, H.; Samoril, T.; Palesch, E.; Hromadko, L.; Zazpe, R.; Skoda, D.; Urbanek, M.; Ng, S.; Prikryl, J.; Macak, J.M. Ideally Hexagonally Ordered TiO2 Nanotube Arrays. ChemistryOpen 2017, 6, 480–483. [Google Scholar] [CrossRef] [PubMed]
  29. Beketova, D.; Motola, M.; Sopha, H.; Michalicka, J.; Cicmancova, V.; Dvorak, F.; Hromadko, L.; Frumarova, B.; Stoica, M.; Macak, J.M. One-Step Decoration of TiO2 Nanotubes with Fe3O4 Nanoparticles: Synthesis and Photocatalytic and Magnetic Properties. ACS Appl. Nano Mater. 2020, 3, 1553–1563. [Google Scholar] [CrossRef]
  30. Tak, M.; Tomar, H.; Mote, R.G. Synthesis of Titanium Nanotubes (TNT) and Its Influence on Electrochemical Micromachining of Titanium. Procedia CIRP 2020, 95, 803–808. [Google Scholar] [CrossRef]
  31. Daulbayev, C.; Sultanov, F.; Korobeinyk, A.V.; Yeleuov, M.; Azat, S.; Bakbolat, B.; Umirzakov, A.; Mansurov, Z. Bio-Waste-Derived Few-Layered Graphene/SrTiO3/PAN as Efficient Photocatalytic System for Water Splitting. Appl. Surf. Sci. 2021, 549, 149176. [Google Scholar] [CrossRef]
  32. Sultanov, F.; Daulbayev, C.; Azat, S.; Kuterbekov, K.; Bekmyrza, K.; Bakbolat, B.; Bigaj, M.; Mansurov, Z. Influence of Metal Oxide Particles on Bandgap of 1D Photocatalysts Based on SrTiO3/PAN Fibers. Nanomaterials 2020, 10, 1734. [Google Scholar] [CrossRef] [PubMed]
  33. Sultanov, F.; Daulbayev, C.; Bakbolat, B.; Daulbayev, O.; Bigaj, M.; Mansurov, Z.; Kuterbekov, K.; Bekmyrza, K. Aligned Composite SrTiO3/PAN Fibers as 1D Photocatalyst Obtained by Electrospinning Method. Chem. Phys. Lett. 2019, 737, 136821. [Google Scholar] [CrossRef]
Figure 1. SEM (ac) and TEM (df) images at different magnifications of cubic SrTiO3 obtained by a chemical precipitation method.
Figure 1. SEM (ac) and TEM (df) images at different magnifications of cubic SrTiO3 obtained by a chemical precipitation method.
Molecules 29 01101 g001
Figure 2. (a) SEM images showing the top surface of the TNT anode array; (b) the TNT@SrTiO3 array; (c) the top view and (d) side view of samples with magnification of the surface of the anode array prepared at 30 V.
Figure 2. (a) SEM images showing the top surface of the TNT anode array; (b) the TNT@SrTiO3 array; (c) the top view and (d) side view of samples with magnification of the surface of the anode array prepared at 30 V.
Molecules 29 01101 g002
Figure 3. (a) X-ray diffraction analysis of combined TNT@ SrTiO3; (b) semi-quantitative elemental analysis of TNT@SrTiO3 particles; (c) EPR spectra of pristine TiO2 and SrTiO3 and pristine SrTiO3/TiO2 nanotube arrays after hydrothermal reaction of 5 h duration.
Figure 3. (a) X-ray diffraction analysis of combined TNT@ SrTiO3; (b) semi-quantitative elemental analysis of TNT@SrTiO3 particles; (c) EPR spectra of pristine TiO2 and SrTiO3 and pristine SrTiO3/TiO2 nanotube arrays after hydrothermal reaction of 5 h duration.
Molecules 29 01101 g003
Figure 4. Schematic illustration of the stages of obtaining TNT.
Figure 4. Schematic illustration of the stages of obtaining TNT.
Molecules 29 01101 g004
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

Bissenova, M.; Umirzakov, A.; Mit, K.; Mereke, A.; Yerubayev, Y.; Serik, A.; Kuspanov, Z. Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization. Molecules 2024, 29, 1101. https://doi.org/10.3390/molecules29051101

AMA Style

Bissenova M, Umirzakov A, Mit K, Mereke A, Yerubayev Y, Serik A, Kuspanov Z. Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization. Molecules. 2024; 29(5):1101. https://doi.org/10.3390/molecules29051101

Chicago/Turabian Style

Bissenova, Madina, Arman Umirzakov, Konstantin Mit, Almaz Mereke, Yerlan Yerubayev, Aigerim Serik, and Zhengisbek Kuspanov. 2024. "Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization" Molecules 29, no. 5: 1101. https://doi.org/10.3390/molecules29051101

Article Metrics

Back to TopTop