Next Article in Journal
Surface Oxidation of TiNiSn (Half-Heusler) Alloy by Oxygen and Water Vapor
Next Article in Special Issue
Development of Alumina–Mesoporous Organosilica Hybrid Materials for Carbon Dioxide Adsorption at 25 °C
Previous Article in Journal
Fundamentals of Force-Controlled Friction Riveting: Part I—Joint Formation and Heat Development
Previous Article in Special Issue
Thermal and Mechanical Properties of Silica–Lignin/Polylactide Composites Subjected to Biodegradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Titania-Based Hybrid Materials with ZnO, ZrO2 and MoS2: A Review

by
Adam Kubiak
,
Katarzyna Siwińska-Ciesielczyk
* and
Teofil Jesionowski
*
Institute of Chemical Technology and Engineering, Faculty of Chemical Technology, Poznan University of Technology, Berdychowo 4, PL-60965 Poznan, Poland
*
Authors to whom correspondence should be addressed.
Materials 2018, 11(11), 2295; https://doi.org/10.3390/ma11112295
Submission received: 16 October 2018 / Revised: 7 November 2018 / Accepted: 12 November 2018 / Published: 15 November 2018
(This article belongs to the Special Issue Multifunctional Oxide-Based Materials: From Synthesis to Application)

Abstract

:
Titania has properties that enable it to be used in a variety of applications, including self-cleaning surfaces, air and water purification systems, hydrogen evolution, and photoelectrochemical conversion. In order to improve the properties of titanium dioxide, modifications are made to obtain oxide/hybrid systems that are intended to have the properties of both components. In particular, zinc oxide, zirconia and molybdenum disulfide have been proposed as the second component of binary systems due to their antibacterial, electrochemical and photocatalytic properties. This paper presents a review of the current state of knowledge on the synthesis and practical utility of TiO2-ZnO and TiO2-ZrO2 oxide systems and TiO2-MoS2 hybrid materials. The first part focuses on the hydrothermal method; then a review is made of the literature on the synthesis of the aforementioned materials using the sol-gel method. In the last section, the literature on the electrospinning method of synthesis is reviewed. The most significant physico-chemical, structural and dispersive-morphological properties of binary hybrid systems based on TiO2 are described. A key aim of this review is to indicate the properties of TiO2-ZnO, TiO2-ZrO2 and TiO2-MoS2 hybrid systems that have the greatest importance for practical applications. The variety of utilities of titania-based hybrid materials is emphasized.

Graphical Abstract

1. Introduction

Constant scientific and technological progress, as well as the desire to create environment-friendly technologies, is leading to intensive work on obtaining new-generation, functional products with strictly designed physico-chemical and dispersive-morphological properties, dedicated to specific applications. These studies focus on changing the physico-chemical features of many well-known and widely used materials. This group of substances includes oxide and hybrid materials based on titanium dioxide, which are distinguished by specific, precisely defined physico-chemical and structural properties, determined chiefly at the stage of their synthesis. Research on the synthesis of advanced materials based on titanium dioxide is becoming oriented towards the conscious and skilled modification of those materials’ properties, achieved by selecting appropriate methods and process conditions. For every method, the selection of conditions such as the precursors used, the implementation of the process, the final processing temperature and the pH of the reaction system has a decisive impact on properties of the product such as its dispersive character, morphology, thermal and colloidal stability, parameters of porous structure, and hydrophilicity or hydrophobicity. The wide range of possible methods of synthesis enables the design of materials of this type with diverse physico-chemical and structural parameters, and is crucial in view of the constant demand for such hybrids. Furthermore, technological progress is accompanied by increased interest in and development of methods enabling better control of the process, and thus also of the properties of the synthesized materials.
Metal oxide systems have attracted a great deal of attention in recent years on account of their special electronic and chemical properties. Among the metal oxide semiconductors, compounds such as TiO2, ZrO2 and ZnO have been investigated extensively due to their chemical stability and good photocatalytic properties [1]. Titanium dioxide is the most widely used metal oxide for environmental applications, paints, electronic devices [2], gas sensors [3] and solar cells [4,5]. It is a well-known semiconductor with excellent photocatalytic properties, which has been widely used in environmental pollutant elimination [6,7], antibacterial dopes, self-cleaning surface, etc. [8]. Its unique antibacterial properties make it a candidate for applications in medical devices and sanitary ware surfaces [9]. Moreover, titania incorporated into polyester fabric [10,11] and related materials can be used as an absorber of harmful UV irradiation. Zinc oxide, due to its antibacterial [12] and photocatalytic activity [13], as well as its wide bandgap, has found applications in various areas based on its optical [14], piezoelectric and gas sensing properties [15,16], and in addition zinc compounds have generally been regarded as safe [1,17]. Zirconia is one of the most intensively studied materials owing to its technologically important applications in gas sensors [18,19], fuel cell electrolytes [20], catalysts [21] and catalytic supports [22], metal oxide-semiconductor devices [23], and in view of its superior thermal and chemical stability [24] and other properties [25].
Among synthetic hybrid oxide systems, particular attention is given to TiO2-ZrO2 materials, which thanks to the addition of zirconium dioxide have much greater surface area and mechanical strength than pure TiO2. Conjugation of zirconia together with titania leads to obtain products which are characterized with higher specific surface in comparison to pristine TiO2. Moreover, the addition of ZrO2 to TiO2 inhibits phase transformation of anatase-to-rutile, and creates more active site groups on titania surfaces [22,24]. There are many publications concerning the application of TiO2-ZrO2 hybrids in photocatalysis [26,27]. The use of TiO2-ZrO2 hybrid materials in the photo-oxidation of organic compounds or degradation of dyes originating from various industrial plants is well known. They also have applications in the photo-reduction of atmospherically harmful oxides, like CO2 and NOx, resulting for example from the combustion of fossil fuels. The advantages of the TiO2-ZrO2 hybrid are its mechanical strength, non-toxicity and corrosion resistance, and the ability to conduct photocatalytic processes using sunlight. These may be causes of increasing demand for this material in the near future [28,29,30].
Combining Titania with zinc oxide can also lead to a hybrid oxide system with good photocatalytic properties [31]. The resulting material can be used, for instance, in the degradation of organic impurities such as detergents, dyes and pesticides present in various types of wastewater. A TiO2-ZnO hybrid material can be synthesized by both physical and chemical processes, which enables enhancement of its properties, for example by widening its spectrum of light absorption. Additionally, the photocatalytic activity of oxides may help reduce the tendency of pollutants to form aggregate structures [32].
The combination of TiO2 with a semiconductor, such as molybdenum disulfide, allows the creation of a hybrid system that not only exhibits activity under the influence of visible radiation, but also allows the separation of photogenerated electron-hole pairs, which increases the catalytic capacity of the material [33,34].
The presented review focuses on recent stage on knowledge about sol-gel, hydrothermal and electrospinning synthesis of advanced, multifunctional materials based on titanium dioxide with strictly defined physico-chemical and structural properties, which significantly determine their multidirectional application. It is known that different properties of hybrid systems, such as TiO2-ZnO, TiO2-ZrO2 and TiO2-MoS2, depend on their morphology, crystallites size, and crystalline structure, which can be modified by selecting the appropriate method for their synthesis, as well as selecting the right components which together with TiO2 will create advanced hybrid materials with unique properties.

2. Method of Synthesis of Titania-Based Materials

2.1. The Hydrothermal Method

One of the most frequently used methods for obtaining powder materials with specific physico-chemical and dispersive-morphological properties is hydrothermal synthesis. This can be used to obtain oxides such as TiO2, ZrO2, Al2O3, etc. According to the definition proposed by Roy [35] and Rubenau [36], hydrothermal synthesis is a reaction that takes place in a water environment at elevated temperature (>100 °C) and pressure, carried out in an autoclave [37,38].
The physico-chemical properties of products formed using the hydrothermal method depend on many process parameters, including the temperature, pressure, time of reaction, reactor volume, type of solvent used, and ratio of reagents (Figure 1). Literature data demonstrate clearly that the temperature and time of hydrothermal treatment [39,40], as well as the relative quantities of reagents used, have an effect on the crystal structure [41,42] and crystallite size of the synthesized material. Another important parameter is the reactor volume; this conditions the creation of (hydrostatic) pressure, which has a direct impact on the size of the synthesized particles [43,44,45].
The hydrothermal method offers many advantages, including a fast reaction rate, the quality and purity of the synthesized products, and the obtaining of materials with a crystalline structure and a smaller particle size. The hydrothermal method can also be used to obtain products in intermediate oxidation states, such as chromium(IV) oxide, as well as metastable compounds such as tellurium iodide (Te2I) [39]. Furthermore, it is extremely advantageous both environmentally and economically, because of its relatively low energy costs. Compared with other well-known and widely used techniques in which the growth of crystals requires significant energy inputs (high temperature and pressure), the hydrothermal method offers conditions of crystallite growth similar to natural conditions. This is consistent with the principles of green chemistry [46,47,48].
Some unquestionable disadvantages of the hydrothermal method include the complexity and cost of the equipment (resulting from the need for the autoclaves to be resistant to high pressure). Moreover, this method does not permit direct observation of the process, and it requires the reagents used to be soluble in water, which serves as the process medium [49,50,51].
Despite these drawbacks, the hydrothermal method is often combined with other techniques (combined methods) to improve the physico-chemical properties of the synthesized materials. One such technique was described by Komarneni et al. [52] and Tompsett et al. [53], who introduced microwaves to the hydrothermal reactor, enabling it to be heated more rapidly to the required temperature. This enhancement of the hydrothermal process is made possible by the special properties of Teflon. That material has an optical gate which enables the passage of waves characteristic for the microwave range. The microwave-assisted hydrothermal method offers a high speed of crystallization (faster reaction kinetics) and leads to products with a very narrow particle size distribution and controlled morphology.
Due to the advantages of both conventional and microwave-assisted hydrothermal methods, in recent years these techniques have been used more frequently for synthesizing oxides and hybrid systems. To obtain oxide systems with diverse physico-chemical and structural properties, many researchers have used titanium dioxide (TiO2)—a compound commonly applied in many branches of industry—as a base for combination with zinc oxide (ZnO), zirconium dioxide (ZrO2) and molybdenum disulfide (MoS2) [54,55]. These materials offer good photocatalytic and electrochemical properties. For this reason, it is a problem of interest to synthesize oxide systems such as TiO2-ZnO, TiO2-ZrO2 as well as TiO2-MoS2 hybrid system using the principles of the conventional and microwave-assisted hydrothermal methods. Table 1 summarizes the most significant literature reports concerning the synthesis of TiO2-ZnO binary systems using the hydrothermal method, with or without the action of microwaves.
Many literature reports indicate that the most photo-active crystalline form of titanium dioxide is anatase [4,5]. However, its use limited by relatively large band gap 3.2 eV [6]. Therefore, the scientists have focused on modifying the physico-chemical properties TiO2 to improve its photocatalytic activity. Modification may be performed by formation hybrid materials [26,28,33]. Xu et al. [8] developed the synthesis of a TiO2-ZnO oxide system based on the principles of the sol-gel and hydrothermal methods. A key part of their work was the evaluation of the photocatalytic properties of the resulting materials in the decomposition of methyl orange (C.I. Basic Orange 10). In the first step, titanium tetrabutoxide, acetic acid and distilled water were mixed with zinc acetate, diethylamine (DEA) and ethanol, after which the solution underwent aging until it became a gel. In the second step the product was hydrothermally treated at different temperatures and for different times. X-ray diffractograms revealed the presence of characteristic diffraction reflections for anatase. Increasing the temperature of the hydrothermal treatment was shown not to affect the crystal structure of the final product. Scanning electron microscope (SEM) images showed that the TiO2-ZnO oxide systems contained particles with near-spherical shape and with a marked tendency to agglomerate. It was also found that an increase in the temperature of hydrothermal treatment, and the calcination process, did not significantly alter the morphology of the TiO2-ZnO oxide system. In photocatalytic tests, all of the synthesized materials demonstrated high photoactivity. The material subjected to hydrothermal treatment at 150 °C for 24 h and additional calcination at 350 °C for 2 h produced a 55% yield of methyl orange degradation after 3 h.
Cheng et al. [56] obtained TiO2-ZnO oxide systems which were used for energy conversion in dye-sensitized solar cells (DSSCs). The synthesis process used previously obtained TiO2 nanospheres, which were hydrothermally treated with zinc nitrate. The products were found to contain the crystal structure of anatase and wurtzite, and had the surface areas (which was determined by the multipoint Brunauere Emmette Teller method (BET)) of 348.8, 187.3 and 122.2 m2/g. It was shown that an increase in the molar fraction of ZnO in the synthesized materials led to increased intensity of the diffraction peaks corresponding to wurtzite, and deterioration of the porous structure parameters. The synthesized systems were used as photoanodes in DSSCs. The tests showed that the TiO2-ZnO oxide systems had better optical properties and higher energy conversion efficiency (8.78%) than pure titanium dioxide (6.79%).
Similar results to the aforementioned reported by Xu et al. [8] were obtained by Li [58], who observed the presence of characteristic diffraction peaks for ZnO, Ti3O5 and Ti, and found the materials to contain particles with a fusiform shape. They also demonstrated that the crystallinity of ZnO declined with an increase in the fraction of TiO2 in the final product. Tests of the photocatalytic activity of the TiO2-ZnO binary oxide system showed that it yielded complete degradation of methyl orange (C.I. Basic Orange 10) after 30 min. With pristine ZnO, by contrast, complete degradation occurred after 50 min.
With regard to the extensively described good electrochemical properties of TiO2 [67,68,69] and ZnO [70,71,72], Vlazan et al. [59] undertook work on the synthesis of a TiO2-ZnO oxide system in the form of core-shell nanoparticles, and went on to determine the electrochemical properties of the synthesized material. In the first stage of the synthesis, ZnO was obtained as a result of hydrothermal treatment of a mixture of zinc nitrate and a sodium hydroxide solution. The ZnO was then added to a suspension of poly(vinyl alcohol) and titanium tetraisopropoxide, and the reaction mixture was treated hydrothermally at 220 °C for 5 h. Analysis by techniques including X-ray diffraction (XRD) and scanning electron microscopy (SEM) showed that the obtained binary oxide system had the crystal structure of anatase and wurtzite, and that its particles were of spherical shape and measured 30 nm in diameter. Compared with the materials described by Cheng et al. [56], the obtained TiO2-ZnO oxide system had a smaller BET surface area (44.4 m2/g) and total pore volume (0.1132 cm3/g). The study also included an investigation of the electrochemical properties of the synthesized binary material. It was found that the electrical resistance of the TiO2-ZnO oxide system decreased with increasing temperature, from ρ = 1.2·× 108 Ω (−128 °C) to ρ = 7.52·× 106 Ω (29 °C). At the same time the electrical conductivity increased from σ = 3.78·× 10−8 S/cm (−128 °C) to σ = 6.04·× 10−7 S/cm (29 °C). The results demonstrate that the synthesized material offers good semiconducting properties.
The common use of titania in photocatalysis [54] motivated Zhang et al. [60] to test an oxide system based on TiO2 in a degradation process of C.I. (Colour Index) Basic Orange 10. Titanium tetrabutoxide and zinc acetate, dissolved in a poly(ethylene glycol) solution, were used as precursors. The obtained solutions were mixed and then treated hydrothermally at 120 °C for 6 h. To improve its crystallinity, the resulting oxide system was subjected to further hydrothermal treatment at 180–200 °C for a specified time. Based on X-ray diffraction results it was shown that the materials contained the crystal structure of anatase and wurtzite. It was observed that increasing the temperature and time of hydrothermal treatment improved the crystal structure of the analyzed materials. The TiO2-ZnO oxide systems contained particles of spherical shape with a tendency to agglomerate, and a range of sizes between 25 nm and 100 nm. The BET surface area was 206 m2/g and 97 m2/g respectively for materials treated at 180 °C and 200 °C. The synthesized materials were tested in the photocatalytic degradation of methyl orange (C.I. Basic Orange 10). The TiO2-ZnO oxide systems exhibited high activity in the degradation of that dye. The synthesized hybrids were also shown to have superior properties to those of pure TiO2 and ZnO.
Spherical hollow structures are receiving a great deal of attention, because they offer the lowest surface/volume ratio for aggregated products. Recent literature reports indicate the high photocatalytic activity of this type of TiO2 [73,74] and ZnO [75] structures due to their large surface area, low density and highly efficient light-harvesting abilities. Wang et al. [61] showed that the use of titanium(IV) sulphate(VI) and zinc nitrate(V) as precursors in the hydrothermal synthesis of TiO2-ZnO binary materials enables the obtaining of products with a defined morphology, spherical hollow structures (Figure 2). The synthesized oxide systems were found to contain the crystal structure of anatase and wurtzite. It was confirmed that an increase in the time of hydrothermal treatment above 6 h did not significantly affect the crystal structure. Analysis of porous structure parameters showed the BET surface area to be equal to 55.9 m2/g. To confirm the high photocatalytic activity of the products, a test was conducted using the degradation of methyl orange (C.I. Basic Orange 10). The TiO2-ZnO oxide systems produced very high rates of dye degradation (total decoloration after 25 min in UV light and after 180 min in visible light).
Core-shell [76,77] nanoparticles have a core made of one material, coated with another material on top of it. In biological applications, core-shell nanoparticles have major advantages over simple nanoparticles, which can lead to improved antibacterial properties. Rusu et al. [62] developed a simple synthesis of a TiO2-ZnO oxide system in the form of core-shell structures, with an application in a process of bacterial degradation. In the first stage of the synthesis, ammonia water was added to a zinc acetate solution. Then titanium tetraisopropoxide was added dropwise to the solution, and the resulting reaction mixture was hydrothermally treated at 220 °C for 5 h. X-ray diffraction results showed the product to contain the crystal structure of TiO2 and ZnO. SEM images revealed particles of spherical shape with a tendency to agglomerate; the particle diameters were measured at 10–20 nm for TiO2 and 30–80 nm for ZnO. Based on the results of antibacterial and antifungal tests, it was found that the material had good biological properties. The TiO2-ZnO oxide system exhibited a biostimulating effect on the biosynthesis of fungi proteases, which is very important from the point of view of biotechnological applications.
Methylene blue (C.I. Basic Blue 9) is the most commonly used substance for dyeing cotton, timber and silk. Therefore, it is of great importance to deal effectively with the pollution resulting from such processes [78,79]. Chen, Zhang, Hu and Li [63] proposed a simple synthesis of TiO2-ZnO binary oxide materials for use in methylene blue degradation. In the first stage of the synthesis, TiCl4 and ZnCl2 were dissolved in a water-ethanol mixture in the molar ratios Ti:Zn = 2:1, 1:1 and 1:2. Next, a urea solution was added with intense mixing for several hours, and then the mixture was subjected to hydrothermal treatment. The synthesized photocatalysts had similar crystal structure and morphology to the materials described in earlier studies. It was also observed that with an increase in the molar fraction of a given monoxide, there was an increase in the intensity of the diffraction peaks corresponding to its crystal structure. In the degradation of a model methylene blue solution, the synthesized TiO2-ZnO oxide systems demonstrated high photoactivity in the decomposition of the tested organic pollutant. Complete decoloration of the dye solution was observed after 180 min in the presence of the material obtained in the molar ratio TiO2:ZnO = 1:2.
The precisely defined crystal structure of this group of binary materials determines the range of their possible applications. In an investigation of the formation of specific mixed crystal structures, Wang et al. [64] carried out a synthesis of TiO2-ZnO oxide systems using the hydrothermal method, supported by an additional calcination process. Diffractograms showed that the materials obtained after the hydrothermal treatment had an amorphous structure, whereas the TiO2-ZnO oxide systems that had been calcined at 600, 700 and 900 °C showed the presence of diffraction reflections characteristic for the crystal structure of anatase (600, 700 °C), rutile (700, 900 °C), wurtzite, Zn2TiO8, Zn2TiO4 and ZnTiO3. To determine the temperature of transformation of specific crystal structures, differential thermal analysis (DTA) was conducted. It was proved that the endothermic peaks at 140, 250, 800 and 940 °C were associated with, respectively, ammonia decomposition, the transformation of Zn2TiO8 to ZnTiO3 and rutile, and the transformation of ZnTiO3 to Zn2TiO4 and rutile. Furthermore, two exothermic peaks were observed at 560 and 690 °C, indicating the creation of a Zn2TiO8 form and the formation of ZnTiO3 and rutile.
The microwave hydrothermal method is a recently developed technique to prepare binary materials in very short times. The advantages of this process over the conventional hydrothermal method include extremely rapid kinetics of crystallization, very rapid heating to the treatment temperature, and the possible formation of new metastable phases [80,81]. Ashok, Venkateswara and Rao [65] developed a methodology for producing a TiO2-ZnO oxide system with the use of the hydrothermal method supported by the action of microwaves. In the first stage of the synthesis, a zinc acetate solution was added to titanium tetraisopropoxide, and then the mixture was alkalized with NaOH. The product was subjected to the action of microwaves for 5 min at 180 °C, using a frequency of 2.45 GHz. In the final step, the resulting oxide system underwent a calcination process in air at temperatures of 500 and 600 °C. X-ray analysis showed the presence of diffraction peaks corresponding to the crystal structure of anatase, wurtzite, and the zinc titanates (ZnTiO3 and Zn2Ti3O4). Images obtained by the SEM technique see (Figure 3) showed the TiO2-ZnO oxide systems to contain spherical particles with a tendency to agglomerate, irrespective of the temperature of thermal treatment.
The synthesis of multi-component systems, despite its many advantages, such as the possibility of obtaining products with very good functional properties, is associated with certain problems, such as the need to optimize the process to obtain the best possible properties with substrates in given proportions. For this purpose, Divy et al. [66] used a combined method (a microwave-assisted hydrothermal process) to produce a TiO2-ZnO oxide system, which was then used as a matrix for a TiO2-ZnO-GO-Ag multi-component system. The matrix must have strictly defined parameters to produce a multi-component system with high photoactivity. For this reason, the synthesized TiO2-ZnO oxide system was analyzed with the use of techniques including XRD and BET. Diffractograms showed the presence of characteristic reflections for anatase and wurtzite. From low-temperature nitrogen sorption isotherms, the BET surface area was calculated to be 290 m2/g, the pore diameter 3.4 nm, and the pore volume 0.32 cm3/g. Because of the theoretically high photocatalytic activity of each component of the TiO2-ZnO-GO-Ag hybrid system, it was tested in the degradation of a model solution of rhodamine B (C.I. Basic Violet 10). The multi-component system exhibited high photoactivity (complete degradation of the dye after 60 min). It also demonstrated high stability, and could be used repeatedly over at least 5 catalytic cycles.
In the next part of this review a survey will be made of work on TiO2-ZrO2 binary systems, synthesized—like the materials discussed in this section—using a hydrothermal process (for review see Table 2).
In recent years, due to the increased emphasis on environmental protection, interest has grown in the use of biomass as a renewable source of fuel and organic chemicals, where fast or flash pyrolysis is used to produce bio-oils. In a study by Leahy [82], a TiO2-ZrO2 oxide material was obtained with the characteristic crystal structures of TiO2 and ZrO2. A scanning and transmission electron microscope (SEM and TEM) images (Figure 4) showed the deposition on nanorods of commercial TiO2 of ZrO2 particles with a diameter of approximately 20 nm. To enable the use of the synthesized oxide systems in the esterification of levulinic acid, surface modification with sulfate groups (SO42−) was carried out. In tests of the extraction process, the TiO2-ZrO2 system attained a high yield of 70% (for a reaction taking place at 80 °C).
Semiconductors such as titania are very important materials with potential uses in diverse applications. A significant property of semiconductors is their electron gap [91], because a prerequisite for an efficient photocatalyst is that the redox potential for the evolution of hydrogen should lie within the band gap of the semiconductor. Tomar and Chakrabarty [83] described the synthesis of TiO2-ZrO2 oxide systems and determined their band gap energies. To obtain the oxide material, titanium tetraisopropoxide was mixed with zirconium tetraisopropoxide in the molar ratios TiO2:ZrO2 = 10:90, 30:70, 40:60, 60:40, 70:30 and 90:10, and the resulting reaction mixture was subjected to hydrothermal treatment at 240 °C for 24 h. The obtained materials were further enhanced by calcination at 450 °C for 4 h. Diffractograms revealed the crystal structures of anatase, rutile, and monoclinic and tetragonal ZrO2. The molar ratio used was found to affect the crystal structure: with an increase in the zirconia content, the diffractions peaks corresponding to ZrO2 became more intense. SEM images showed the particles to have a spherical shape. The band gaps, determined by the technique of UV-Vis spectroscopy, were measured at 2.48, 1.72, 1.34, 1.7, 1.49 and 1.49 eV respectively for the oxide systems synthesized in the molar ratios TiO2:ZrO2 = 10:90, 30:70, 40:60, 60:40, 70:30 and 90:10.
Kubo et al. [84] described the effect of a calcination process on the physico-chemical properties of a TiO2-ZrO2 oxide system, using zirconium tetraisopropoxide and titanium tetraisopropoxide. In the first stage, the precursors of TiO2 and ZrO2 were mixed in the molar ratios Zr:Ti = 1:0, 9:1, 6:1 and 3:1, and then underwent hydrothermal treatment at 300 °C for 2 h in a nitrogen atmosphere. After this process, the materials were subjected to additional calcination at 500 and 800 °C. X-ray diffraction analysis showed the presence of characteristic reflections for the crystal structure of tetragonal ZrO2. SEM images showed the synthesized materials to contain particles of spherical shape with a tendency to agglomerate. The effect of fusion of particles was also observed for the materials that had undergone thermal treatment (calcination). Low-temperature nitrogen sorption measurements gave BET surface areas of 200 and 80 m2/g respectively for the systems calcined at 500 and 800 °C. The conclusion of morphological and pore structure analysis was that an increase in calcination temperature caused a decrease in BET surface area.
Many literature reports indicate the high photocatalytic activity of zirconia in the degradation of dyes such as methylene blue [92,93,94]. Hirano et al. [85] developed a simple hydrothermal synthesis of TiO2-ZrO2 binary oxide systems using titanium(IV) sulphate(VI) and zirconium sulphate(VI). Wide-ranging physico-chemical analysis confirmed the hydrothermal method as an effective process for the synthesis of binary oxide systems with defined crystallinity and morphology. The reported results indicate the presence of the crystal structure of anatase and monoclinic ZrO2. Transmission electron microscope images (Figure 5) showed that the particles of the TiO2-ZrO2 system had an almost spherical shape. In view of the presence of anatase and tetragonal ZrO2 in the synthesized material, an attempt was made to use the binary system in the process of degradation of methylene blue (C.I. Basic Blue 9). The TiO2-ZrO2 systems exhibited greater photocatalytic activity than the pure mono-oxides. The highest photoactivity in the degradation of the dye was achieved by the material obtained in the molar ratio TiO2:ZrO2 = 90:10.
At present the main focus in the field of DSSCs is the development of a photoanode that can efficiently harvest light [95]. Increased dye pickup, light scattering ability, reduced recombination reaction and improved charge transport ability are parameters that need to be engineered for high conversion efficiency. The properties of titania make it one of the most important DSSC photoanode materials. Tomar et al. [86] worked on a DSSC using a TiO2-ZrO2 system as the anode. In testing of the cell, the binary oxide material demonstrated a higher energy conversion efficiency (η = 1.97%) and fill factor (FF = 0.256) than pure TiO2 (η = 0.71%, FF = 0.061).
Yao et al. [87]—similarly to Hirano et al. [85]—described the use of a TiO2-ZrO2 oxide system in the process of photodegradation of methylene blue (C.I. Basic Blue 9). The material obtained in their study had a similar crystal structure to that described in [85]. SEM images revealed hollow microspheres of TiO2-ZrO2—this is closely linked to the high BET surface area of the material, which was determined from nitrogen adsorption/desorption isotherms to be 224 m2/g. Finally, the photocatalytic properties of the obtained binary material were tested in the degradation of methylene blue (C.I. Basic Blue 9). Use of the material led to a high level of degradation of the dye, reaching 99% after exposure to light for 60 min.
Caillot et al. [88] described the use of a synthesized TiO2-ZrO2 binary oxide system as an effective adsorbent of NH3 and SO2. The use of titanium(IV) and zirconium(IV) chlorides, and hydrothermal treatment at 220 °C, led to an amorphous structure. The absence of a defined crystal structure is linked to the high BET surface area (209 m2/g) and the unspecific particle shape. Results of energy-dispersive X-ray spectroscopy (EDX) indicated that the synthesized material contained 35% of zirconia and 65% of titania. Adsorption tests showed the material to have high adsorption capacities with respect to both NH3 and SO2; the values obtained were 573 µmol/g for NH3 and 414 µmol/g for SO2.
Ueda et al. [89] described the use of a titanium dioxide zirconium dioxide system to form bioinert foils. The synthesis of the material included hydrothermal treatment at 180 °C for 12 h. To enable use of the obtained materials in biological processes, the surface was modified using lactic acid. It was observed that while pure TiO2 had a high propensity to proliferate MC3T3-E1 cells (MC3T3 is an osteoblast precursor cell line derived from Mus musculus (mouse) calvaria), the synthesized oxide material showed a good ability to suppress proliferation of the cells, caused by a change in the surface morphology (increased roughness), which has a direct impact on the cell reproduction process. The experiment shows that the synthesized TiO2-ZrO2 material may be used for the purpose of slowing down osteointegration. The researchers also carried out a wide-ranging physico-chemical analysis to determine the morphological and crystal structure of the material. The synthesized binary oxide materials were found to contain the crystal structures of TiO2 and ZrO2, with the formation of a thin layer of ZrO2 on the surface of the TiO2.
Kim et al. [90] synthesized a TiO2-ZrO2 system for use in a DSSC. In the first stage, the material was synthesized with the use of hydrothermal treatment at 260 °C for 12 h. X-ray diffractograms showed the presence of characteristic peaks for the anatase structure. The synthesized material was then used to prepare paste electrodes. Analysis of the porous structure of the electrode material showed it to have a BET surface area of 40.6 m2/g, a pore volume of 0.18 cm3/g and a pore diameter of 5.7 nm. In electrochemical tests, the material exhibited a high energy conversion efficiency of 2.88%, compared with 0.09% for commercial TiO2 (P25).
The next part of this review will contain a survey of the literature on the synthesis and potential applications of TiO2-MoS2 hybrid systems (Table 3).
Indoles are among the most versatile and common nitrogen-based heterocyclic scaffolds and are frequently used in the synthesis of various organic compounds [106,107]. Indole-based compounds are very important among heterocyclic structures due to their biological and pharmaceutical activity. However, in order to improve their properties, it is necessary to modify them. Wang et al. [33] described a simple modification of indoles by thiocyanation, using a synthesized TiO2-MoS2 hybrid as a photocatalyst. Physico-chemical and structural analysis of this hybrid indicated a defined crystal structure and morphology. X-ray diffraction analysis showed the presence of reflections corresponding to the crystal structure of anatase TiO2 and MoS2. TEM images showed a layered structure of molybdenum disulfide with evenly distributed particles of titanium dioxide on the surface. Thiocyanation of indoles was carried out in a photocatalytic process, in which high yields were obtained. The highest photocatalytic activity (93%) was observed for the material synthesized in the molar ratio TiO2:MoS2 = 10:1.
Zhu et al. [34] synthesized a hybrid system containing the crystal structures of TiO2 and MoS2. Images obtained by scanning and transmission electron microscopy showed TiO2 particles uniformly distributed on the surface of the MoS2. Cyclic voltammetry curves showed the TiO2-MoS2 hybrid to have good electrochemical properties, with charge and discharge capacities at a current density of 100 mA/g equal to 643 and 827 mAh/g respectively in the first cycle, and 643 and 674 mAh/g in the second cycle. The capacity of the material also remained very stable over 100 subsequent cycles. It was concluded that the addition of anatase improved the electrochemical properties by facilitating the transport of electrons and ions during charging and discharging.
Ren et al. [96] used the hydrothermal method to obtain a TiO2-MoS2 hybrid material with strictly defined photoelectrochemical properties. Measurements showed the hybrids to have superior electrochemical properties to those of pure TiO2 or MoS2. The hybrid synthesized in the molar ratio TiO2:MoS2 = 20:1 had a photocurrent density of 33 µA/cm2 (compared with 8 µA/cm2 for TiO2). Photoelectrochemical results demonstrated a synergy effect between the titanium dioxide and molybdenum disulfide. A wide-ranging analysis of physico-chemical properties was also performed. The synthesized materials produced diffractions peaks corresponding to anatase and 2H-MoS2. For the hybrid obtained in the molar ratio TiO2:MoS2 = 20:1, the BET surface area was measured at 48.2 m2/g and the pore diameter at 7 nm. Morphological analysis confirmed that the particles of molybdenum disulfide were coated with particles of titanium dioxide.
Molybdenum disulfide photocatalysts are attracting wide attention as they offer a suitable band gap for visible-light harvesting, making the compound a promising earth-abundant photocatalyst for hydrogen production, environmental remediation, and photosynthesis [108,109]. A team of researchers [97] developed a two-stage method for obtaining active TiO2-MoS2 photocatalysts by hydrothermal synthesis. X-ray diffractograms revealed the crystal structure of TiO2 (for all of the materials) and MoS2 (only for the systems formed in the molar ratio Mo:Ti = 7.5%). The BET surface area was measured at 86, 87, 98 and 92 m2/g respectively for the systems with molar ratios Mo:Ti = 1%, 2.5%, 5% and 7.5%, and the respective band gaps were determined to be 3.15, 3.14, 3.1 and 2.97 eV. The synthesized TiO2-MoS2 hybrids were tested in a process of photodegradation of methylene blue (C.I. Basic Blue 9), and analysis of their photocatalytic activity showed that they produced high yields of degradation of the dye. For the system with Mo:Ti = 7.5% the solution was completely decolored after 60 min, indicating a tripling in the efficiency of photodegradation of the organic pollutant compared with pure TiO2.
Liu et al. [98] described the synthesis of multifunctional TiO2-MoS2 hybrids with photoelectrochemical and photocatalytic properties. The substances used as precursors of TiO2 and MoS2 were commercial titanium dioxide (P25 Evonik) and sodium molybdenate. These were mixed to obtain the mass ratios TiO2:MoS2 = 80:20, 60:40, 40:60 and 20:80. The systems then underwent hydrothermal processing at 180 °C for 24 h. X-ray diffractograms contained characteristic peaks for TiO2, Ti6O11 and MoS2. SEM images (Figure 6) showed that spherical particles of titanium dioxide were deposited on nanofibers of molybdenum disulfide. For the resulting systems, a determination was made of their photoelectrochemical properties and photocatalytic activity in the removal of rhodamine B (C.I. Basic Violet 10). The TiO2-MoS2 hybrid system with the mass ratio TiO2:MoS2 = 60:40 exhibited high photocurrent density (12.5 µA/cm2). An increase in the content of molybdenum disulfide in the hybrid system led to a deterioration of its photoelectrochemical properties. In the photocatalytic degradation of rhodamine B, the highest photocatalytic activity was exhibited by the system with the mass ratio TiO2:MoS2 = 60:40. Complete decoloration of the dye solution was observed after 90 min.
Zhang et al. [99] described the use of TiO2-MoS2 binary hybrids in the photocatalytic degradation of methyl orange (C.I. Basic Orange 10). In the first stage, a two-step hydrothermal method was used to synthesize the aforementioned materials. TiO2 nanoparticles were formed by a sonication method. Molybdenum disulfide was then added to the TiO2, and the resulting mixture was subjected to hydrothermal treatment at 140 °C for 3 h. Unlike other previously described hybrids, the material obtained in this study was found to contain the crystal structure of anatase, ilsemanite and molybdenite. Scanning electron microscope images showed that spherical particles of titanium dioxide were deposited on MoS2 nanofibers. In photodegradation tests, the TiO2-MoS2 hybrid system exhibited high photocatalytic efficiency in the decomposition of a model dye solution after 60 min.
Yang et al. [100] synthesized a hybrid system based on TiO2 and MoS2 using a hydrothermal method, with the aim of testing the resulting materials in a process of reduction of Cr(VI). In the first stage of synthesis, titanium dioxide was obtained from titanium in an anodic process. Next, thioacetamide and sodium molybdenate were dissolved in deionized water, the obtained titanium dioxide was added, and hydrothermal treatment was carried out at 220 °C for 24 h. X-ray diffractograms showed the synthesized materials to have the crystal structure of anatase and hexagonal MoS2. SEM images revealed hollow TiO2 nanotubes in which MoS2 particles were deposited. Photoelectrochemical tests demonstrated that the addition of MoS2 to the TiO2 structure reduced its resistance, leading to an improvement in the transport of electrons in the catalyst. A high yield of Cr(VI) reduction was obtained using the TiO2-MoS2 hybrid system, with a rate of reduction of 103.9 mg/L·min·cm2 at 0.1 V. Tests of the stability of the reaction process showed that the process efficiency did not deteriorate over five consecutive cycles.
Phenol and its derivatives have been used in many industrial processes, including the production of dyes, polymers and medicines. Due to their high production volumes and widespread application, phenols and their derivatives are released into the environment from industrial and municipal liquid wastes [110,111]. This is particularly undesirable, since phenols pose a serious hazard due to their hematotoxic and hepatotoxic actions [112]. Wang et al. [101] described the use of synthesized TiO2-MoS2 hybrid systems in the process of photodegradation of phenol. At the first stage, hollow spheres of TiO2 were obtained [108,109]. Next, sodium molybdenate was added in the molar ratios TiO2:Na2MoO4 = 30:135 (sample TM3), 60:135 (sample TM6), 90:135 (sample TM9) and 120:135 (sample TM12). Diffractograms obtained for the TiO2-MoS2 materials contained characteristic peaks for the crystal structure of anatase TiO2 and rhombohedral MoS2. The anatase structure became less visible as the MoS2 content increased. Scanning electron microscope images (see Figure 7) showed particles of titanium dioxide with close to spherical shapes, hollow inside, with diameters of 600 nm, coated with a layer of molybdenum disulfide. The synthesized materials were tested in a process of photocatalytic degradation of phenol. The hybrids exhibited good photocatalytic properties: the yield of phenol degradation was 78% after 150 min in the presence of the system obtained in the mass ratio TiO2:Na2MoO4 = 90:135. Tests of photocatalytic stability showed that the photoactivity of this material did not decrease significantly over four consecutive catalytic cycles.
Bai et al. [104] were the first to report the combination of exfoliated MoS2 with TiO2 nanoparticles and the use of the resulting material in a photocatalytic process. To establish whether the materials could be successfully used as active photocatalysts, a comprehensive physico-chemical analysis was carried out. Recorded diffraction peaks corresponding to the crystal structure of TiO2 (anatase) and MoS2. Morphological parameters were investigated by the SEM and TEM techniques. The resulting images showed the particles of molybdenum disulfide to be densely covered with particles of titanium dioxide. Distances between layers were measured at 0.27 nm and 0.35 nm, which confirmed the presence of the structure of MoS2 and TiO2 nanocrystals. In tests of photocatalytic activity in the decomposition of rhodamine B (C.I. Basic Violet 10), 100% decomposition of the dye was attained after exposure to light for 30 min in the presence of the TiO2(anatase)-MoS2(1T) hybrid system, compared with 40% in the presence of the TiO2(anatase)-MoS2(2H) system. In addition, the TiO2(anatase)-MoS2(1T) system catalysed the evolution of hydrogen at a rate of 2 mmol/g·h, compared with 0.25 mmol/g·h for TiO2(anatase)-MoS2(2H) and 0.5 mmol/g·h for pure TiO2.
Photocatalytic evolution of H2 over semiconducting materials is regarded as one of the promising solutions to the growing energy crisis, due to its potential application for clean hydrogen production from water [109,113,114]. Yuan et al. [105] described the use of a single-stage hydrothermal process to synthesize binary hybrid photocatalysts for the evolution of hydrogen. Photocatalytic hydrogen evolution experiments were mainly carried out in a Pyrex glass cell with a top window connected to a gas-closed system. The assessment of photocatalytic properties was carried out using 100 mg of H2-evolving photocatalyst powder which was suspended in 100 cm3 of an aqueous solution containing 10% methanol in volume. The resulting materials led to high yields of hydrogen: the highest—2200 µmol/g·h—was recorded for the system with a 0.5% MoS2 mass fraction (this is 36 times higher than the yield obtained with pure TiO2). For comparison, TiO2-Pt, TiO2-Pd, TiO2-Rh, TiO2-Au and TiO2-Ru were used as catalysts in the same process, giving yields of 1450, 1250 1000, 750 and 550 µmol/g·h respectively. Physico-chemical parameters were analysed by the XRD, TEM and BET techniques. X-ray diffractograms contained characteristic peaks for the crystal structure of anatase TiO2 and MoS2. Based on nitrogen sorption, a BET surface area of 105.6 m2/g was obtained. Transmission electron microscope images showed interplanar distances of 0.19 nm, characteristic for TiO2, and 0.62 nm, characteristic for hexagonal MoS2.
A comprehensive review of the literature concerning oxide systems containing titanium dioxide together with zinc oxide or zirconium dioxide, or TiO2-MoS2 hybrid systems, synthesized by the hydrothermal method confirms that by the suitable choice of process conditions (precursors, pH of the reaction system, molar ratio of reagents, temperature of thermal treatment) it is possible to achieve effective changes in the properties of the resulting materials. Appropriately selected process parameters make it possible to design the physico-chemical properties of synthesized oxide and hybrid materials, including particle shape and size, degree of crystallinity, phase and surface compositions, and porous structure. Moreover, hydrothermally synthesized materials of TiO2-ZnO, TiO2-ZrO2 and TiO2-MoS2 type are finding a wide range of applications in bacterial and fungal degradation and as photocatalysts in the decomposition of certain organic pollutants. Such systems are also used as electrode materials in lithium-ion cells and dye-sensitized solar cells.

2.2. The Sol-Gel Method

The consistent development of technology has led to increased awareness of the importance of chemical synthesis methods, particularly those involving the use of organic precursors. One of the basic types of synthesis in which metallo-organic precursors are used to synthesize binary hybrid systems is the sol-gel method [115,116,117,118,119,120]. In recent years many definitions of the sol-gel process have been put forward, including that of Dislich and Hinz [121], which covers only multicomponent oxides. According to the most popular and widespread definition, such a process is based on a hydrolysis reaction of metallo-organic precursors of metals (usually alkoxides) in a reaction mixture [31,122,123,124,125].
A sol-gel process consists of several stages. In the first, a solution of metal alkoxides is mixed. This is followed by a process of hydrolysis and condensation, after which a sol is obtained [126,127,128,129,130,131,132]. In the next stage, gelation takes place—that is, the sol is transformed into a gel. The final stage involves the removal of residues of the liquid phase via a drying process. A block diagram of the sol-gel process appears in Figure 8 [133,134,135,136,137].
Sol-gel processes may be used in many branches of industry; for example, in the production of catalysts or catalyst supports [138], optical and electrical sensors [139], and implant coatings [139,140] Recent reports concern the use of the sol-gel method in the process of immobilization of a wide range of organic materials, including enzymes [141,142,143,144].
Compared with other conventional methods of synthesis, the sol-gel method has a number of advantages which enable its use in a wide range of technological processes. Some of these advantages are the ability to control the microstructure of the product, the use of low temperatures and the resulting low energy costs, the high homogeneity of the product, the absence of a need for complex apparatus, and the possibility of synthesizing new materials with designed properties [145,146].
The most significant drawbacks of the sol-gel process include the high costs of the metal precursors, the need to use reagents of high purity, the absence of total control over the process (the evolution of various oxides due to differences in the reactivity of the precursors), and the fusion of the product during calcination [122,138,139].
The sol-gel method may be used to obtain oxide systems containing TiO2. Titanium dioxide, because of its photocatalytic [147,148,149] and photoelectrochemical properties [150,151], represents an interesting base for the production of multifunctional oxide systems. There has recently been a marked growth in interest in such materials as TiO2-ZnO and TiO2-ZrO2 oxide systems, in view of the useful properties of zinc oxide (photocatalytic, photoelectrochemical and photoluminescent) and zirconium dioxide (photocatalytic and photoelectrochemical).
In Table 4, an overview is given of the current stage of knowledge concerning the synthesis of TiO2-ZnO oxide systems using the sol-gel process.
Stoyanova et al. investigated the effect of the type of precursor used, the TiO2:ZnO molar ratio and the calcination temperature on the physico-chemical properties of TiO2-ZnO oxide systems synthesized by the sol-gel method [1]. At the first stage, TiO2-ZnO systems were obtained using zinc acetate (sample A) and zinc chloride (sample B) as the ZnO precursor, and titanium tetrachloride as the precursor of TiO2. The TiO2:ZnO molar ratio was 90:10 (sample A) or 50:50 (sample B). The materials prepared from zinc acetate were calcined at 500 °C for 6 h, and those made using zinc chloride were calcined at 200, 400, 550 and 600 °C for 2 h. The diffractograms contained peaks corresponding to anatase in sample A, and to anatase, rutile and zinc titanate in sample B. It was demonstrated that the crystal structure of the final product was affected by the precursor type, the molar ratio of the reagents and the temperature of thermal processing. SEM images showed that the synthesized materials consisted of particles with a spherical shape and a marked tendency to agglomerate. Authors applied the obtained materials in a process of photodegradation of malachite green (C.I. Basic Green 4). They also determined the antibacterial properties of the materials against the Gram-positive E. coli (ATCC 25922). The results of the dye photodegradation experiment showed that both the material obtained from zinc acetate and that made using zinc chloride exhibited high photocatalytic activity: total degradation occurred after 40 min in the presence of sample B and after 170 min in the presence of sample A. In the tests with E. coli, the TiO2-ZnO system obtained using zinc chloride demonstrated 100% antibacterial activity after 120 min, while the system made using zinc acetate achieved only a 30% yield in the same amount of time. Under UV light, it was found that the bacteria were removed completely after 15 min in the presence of sample A and after 30 min in the presence of sample B.
In Section 2.1 it was described how a TiO2-ZnO system synthesized by the hydrothermal method could be used in dye-sensitized solar cells [56]. Shanmugam et al. [152] also described the use of a binary oxide system in a DSSC, but applied the sol-gel method of synthesis. Electrochemical characterization of the cell showed it to have a current density of 13.1 mA/cm3 and a photoconversion efficiency of 32% (in darkness). The maximum equivalent quantum efficiency (EQE) of the DSSC built from a TiO2-ZnO oxide system was measured at 85%, compared with 67% for a cell based on TiO2 alone.
In recent years there have been reports concerning the potential use of non-toxic and inexpensive titania nanoparticles for imparting multifunctional properties to various textile materials [166,167,168]. Chu et al. [153] described the effect of the addition of a titania-based binary material on the antistatic properties of polyester fabrics. It was shown by physico-chemical analysis that the resulting materials had good stability in water, and that the stability improved as the molar fraction of ZnO was increased. SEM images of polyester fibres covered with TiO2-ZnO showed that the fabric was covered uniformly by a layer of the oxide material. The best antistatic properties were exhibited by materials obtained with TiO2:ZnO molar ratios of 5:5 and 7:3. Mechanical characterization of the materials showed that they offered greater permeability to gases than pure TiO2 or ZnO. A whiteness quality test showed that the white colouring of the obtained TiO2-ZnO systems was 10–30% lower than that of the reference samples.
Recent literature reports indicate that both titania [169,170] and zinc oxide [171] can be used as anti-corrosion coatings. Yu et al. [154] presented a simple synthesis of TiO2-ZnO oxide systems using the sol-gel method and described their properties as anti-corrosion coatings. In the first stage, TiO2 sol was obtained from titanium tetrabutoxide, ethanol, glacial acetic acid and water, and ZnO sol from zinc acetate, diethylamine, ethanol and water. These sols were mixed together in the molar ratios TiO2:ZnO = 0:1, 1:3, 1:1, 3:1, 1:0. The materials were applied to carbon steel using an immersion technique. The steel with the oxide layer was then dried at 70 °C for 10 h. Finally the materials were calcined at 350 and 500 °C for 2 h. The procedure was repeated two or four times to obtain coatings consisting of two, four or eight layers. The carbon steel plates coated with oxide materials were subjected to comprehensive physico-chemical analysis. The materials were shown to have the crystal structure of Fe, TiO2, Fe3O4, Fe2O3 and ZnFe2O4. X-ray diffractograms contained peaks originating from the iron present in the carbon steel. The material obtained at a TiO2:ZnO molar ratio of 1:0 was characterized by an anatase crystal structure. For the product calcined at 500 °C the forms Fe3O4 and Fe2O3 were observed, which indicates oxidation of the steel. SEM images showed that the materials calcined at 350 °C had numerous cracks, creating a non-uniform layer on the surface of the steel. An increase in the ZnO content in the oxide system was found to improve the uniformity of the coating. Potentiodynamic polarization curves were obtained to determine the anticorrosive properties of the coatings: the results showed that significant protection against corrosion was provided in contact with sea water. The best anticorrosive properties were obtained for the systems with TiO2:ZnO ratios of 1:1 and 1:3, thermally treated at 500 °C.
To obtain a specific crystal structure of materials synthesized by the sol-gel method, it is necessary to carry out a calcination process. Many literature reports refer to the effect of calcination on photocatalytic properties [172,173,174]. Wang et al. [155] conducted a simple synthesis of a TiO2-ZnO oxide system using the sol-gel method and a calcination process. A TiO2 sol was obtained from titanium tetrabutoxide, anhydrous ethanol and hydrochloric acid, and a ZnO sol was obtained from zinc acetate, anhydrous ethanol and deionized water. Oxide systems were obtained by mixing the TiO2 and ZnO sols at the molar ratios ZnO:TiO2 = 0.1:0.15, 0.2:0.25, 0.3:0.35. In the final stage the oxide materials were calcined at 450, 480, 500, 550 and 600 °C for 1, 1.5, 2, 2.5 and 3 h. Since the physico-chemical properties of titania-based materials determine their photocatalytic ability, analysis was performed using such techniques as X-ray diffraction and scanning electron microscopy. The results of both XRD and SEM indicate the formation of the crystalline structure of anatase and wurtzite, and the spherical shape of the particles of the materials obtained. Their photocatalytic activity was evaluated with respect to the photodegradation of methyl orange (C.I. Basic Orange 10). All of the materials exhibited good degradation efficiency. The highest yield of dye degradation (93% after 5 h) was achieved by the material obtained at a ZnO:TiO2 molar ratio of 0.2:0.25.
Due to the increase in the number of cars and the emission of harmful gases in city centers, it is apparent that there is a need to remove pollutants such as nitrogen oxides (NOx). Many researchers, such as Dalton et al. [175], Karapati et al. [176] and Todorova et al. [177], have reported the use of titanium dioxide as an effective NOx photocatalyst. Giannakopoulou et al. [156] described the synthesis of an effective binary oxide photocatalyst of nitrogen oxides using the sol-gel method. At the first stage of synthesis, TiO2 sol was obtained by mixing titanium tetrabutoxide with anhydrous ethanol, nitric acid and distilled water. The ZnO sol was produced using zinc acetate, ethylene glycol, glycerine, TEA and anhydrous ethanol. The sols were mixed together at the molar ratios TiO2:ZnO = 100:0, 75:25, 50:50, 25:75 and 0:100. Thin foils of the resulting binary oxide photocatalysts were produced using an immersion technique. The final step was calcination at 500 °C for 1 h. Analysis of X-ray diffractograms showed that the TiO2:ZnO molar ratio affected the parameters of the crystal structure. An increase in the ZnO content leads to the vanishing of the diffraction reflections characteristic of TiO2. The oxide materials with TiO2:ZnO molar ratios of 100:0 and 75:25 contained a single diffraction peak corresponding to anatase. When the ratio was 50:50 or 25:75 an amorphous structure was observed, while the system with 100% ZnO had the crystal structure of wurtzite. A test of photocatalytic activity showed that the highest activity in the oxidation of NOx was achieved by the materials obtained with TiO2:ZnO molar ratios of 100:0 and 75:25. An increase in the content of ZnO led to a significant drop in photocatalytic activity, possibly due to the weak crystallinity of the materials obtained.
The surface wettability of a material is one of the properties that determine its potential applications. Chen et al. [157] reported the synthesis of a superhydrophilic TiO2-ZnO oxide system in the form of a thin foil. It was produced using a TiO2 sol made from titanium tetrabutoxide, anhydrous ethanol, diethylamine (DEA) and nitric acid, and a ZnO sol obtained using zinc acetate, anhydrous ethanol and DEA. The precursor solutions were mixed together in the molar ratios TiO2:ZnO = 90:10, 80:20, 70:30, 60:40, and then underwent calcination at 500 °C for 30 min. The results relating to the crystal structure were similar to those reported by Giannakopoulou et al. [156]. An increase in the molar fraction of ZnO led to the vanishing of the anatase structure, and in the case of the materials with TiO2:ZnO ratios of 70:30 and 60:40, this resulted in the presence of an amorphous structure. Atomic force microscope (AFM) images (Figure 9) showed that the obtained composite foils had a microporous structure. Contact angle measurements showed all of the materials to be superhydrophilic, with contact angles in the range of 1.8–3.2° and diffusion of a drop of water on the surface of the material within 800 ms. A test involving the photodegradation of methylene blue (C.I. Basic Blue 9) showed that the addition of a small quantity (10%) of ZnO improved the material’s photocatalytic activity: the degradation yield was 80% after exposure to light for 280 min.
Pournuroz et al. [158] analyzed the degradation of the dyes C.I. Reactive Red 195 (RR195) and C.I. Reactive Blue 19 (RB19) in the presence of a TiO2-ZnO oxide system. In the synthesis process sols of TiO2 and ZnO were obtained, which were mixed together in the molar ratio TiO2:ZnO = 10:3. The resulting material exhibited high photocatalytic activity. Complete degradation of the RR195 and RB19 dyes was observed after exposure to light for 40 and 60 min respectively, in the presence of the synthesized material.
The sol-gel auto-combustion synthesis method (also called low-temperature self-combustion, auto-ignition or self-propagation, as well as gel-thermal decomposition) combines the chemical sol-gel and combustion processes [178,179]. This method has shown great potential in the preparation of novel nanomaterials. Al-Mayman et al. [159] described the use of a sol-gel auto-ignition technique to synthesize a TiO2-ZnO oxide system with defined photocatalytic properties. At the first stage, TiO2 and ZnO precursors were dissolved in deionized water, and citric acid was added. The reaction mixture was then heated to 100 °C to evaporate off water. The resulting gel was ignited by heating to a temperature of approximately 300 °C. In the final stage, the TiO2-ZnO system was calcined in air at 500 °C for 5 h. Materials were obtained with the general formula Ti1−xZnxO2, for x = 0%, 1%, 5%, 10%, 12% and 30%. Their diffractograms contained peaks corresponding to the crystal structure of anatase. Microstructural and morphological analysis indicated particles with a spherical shape and a marked tendency to agglomerate. The synthesized TO2-ZnO oxide materials had surface areas of 52, 63.4, 69.9, 64 and 36.6 m2/g when the mass fraction of ZnO was 1%, 5%, 10%, 12% and 30% respectively. The wide band gaps of titanium dioxide and zinc oxide make them promising materials for applications in the process of photocatalytic hydrogen production. All of the obtained materials were shown to be capable of evolving hydrogen under UV light. The highest yield was obtained from the material with the formula Ti90Zn10O2 (1.048 mmol/h of hydrogen after 9 h of irradiation).
Zulkiflee and Hussin [160] investigated the effect of the temperature of thermal processing on the physico-chemical properties of a TiO2-ZnO oxide system obtained by the sol-gel method. At the first stage a TiO2 sol was obtained from titanium tetrabutoxide, acetic acid and butanol, and a ZnO sol from zinc acetate, diethylamine and propan-2-ol. To obtain thin layers of the oxide material, glass plates were immersed in the titanium dioxide sol and calcined at 400 °C. The plates were then immersed in ZnO sol and calcined at 500 and 600 °C for 2 h. X-ray diffraction results showed that the materials obtained had the crystal structure of anatase and wurtzite. Atomic force microscope (AFM) observations demonstrated that an increase in the calcination temperature caused an increase in the roughness of the material surface. The band gaps of the synthesized materials were measured at 3.6 and 3.7 eV for the samples TiO2(400 °C)-ZnO(500 °C) and TiO2(400 °C)-ZnO(600 °C).
Prasannalakshmi and Shanmugam [161] used a TiO2-ZnO oxide system in the process of photocatalytic degradation of brilliant green (C.I. Basic Green 1) and methylene blue (C.I. Basic Blue 9). In the first stage of the synthesis, a TiO2 sol was obtained by hydrolysis of titanium tetraisopropoxide, and a ZnO sol was prepared from zinc acetate and 0.5 M sodium hydroxide. The sols were mixed together in the molar ratios TiO2:ZnO = 0.5:0.25, 1:0.5, 1.5:0.75, 2:1, to produce samples labelled TZO1 TZO2, TZO3 and TZO4. The final stage of the process was calcination at 550 °C for 4 h. In analysis of the crystal structure, all of the oxide systems gave diffraction reflections corresponding to anatase and wurtzite. The XRD results were confirmed using HRTEM analysis. HRTEM images showed distances between planes characteristic of (101) TiO2 and (101) ZnO. The surface areas, determined by low-temperature nitrogen adsorption and calculated based on BET theory, were 74.7, 32.2, 12.58 and 8.82 m2/g for the samples TZO1 TZO2, TZO3 and TZO4 respectively. All of the synthesized materials exhibited high photocatalytic activity, and each of the binary samples (TZO1, TZO2, TZO3, TZO4) had superior photocatalytic properties to those of pure TiO2 or ZnO.
Materials such as titania [180,181] and zinc oxide [182] are successfully used for the construction of p-silicon heterojunction photodiodes. Al-Hazami and Yakuphanoglu [162] used a binary combination of these materials to build a p-type diode. A layer of the TiO2-ZnO oxide system was applied to a prepared silicone base, after which the material was dried and then calcined in air at 450 °C for 1 h. SEM results showed particles with a spherical shape with no tendency to agglomerate. The band gap was measured at 3.41 eV. Photoelectrochemical tests showed that the p-type diode covered with a thin foil of the TiO2-ZnO system had photovoltaic and photoconductive properties. Tests under different luminous intensities confirmed a rapid reaction to changes in the studied parameters and good process repeatability.
Taking account of the known antibacterial properties of titanium dioxide and zinc oxide, Armin et al. [163] carried out a wide-ranging analysis of possible applications of a TiO2-ZnO binary system. The obtained material was used in electrochemical, photocatalytic and antibacterial tests. Dynamic potentiometric curves showed that TiO2-ZnO coatings offered high resistance at low currents. The curves are contained within a very wide range of potentials, from −1.4 to 1.1 V. Based on the measured parameters, a corrosion index was computed for coated and uncoated materials. The material coated with TiO2-ZnO had a corrosion protection index of 98.6%, compared with only 82% for the uncoated material. Antibacterial tests showed that the TiO2-ZnO material suppressed the growth of populations of the organisms C. albicans (1.5·× 105) and E. coli (1.5·× 107). In an experiment involving the photodegradation of methylene blue (C.I. Basic Blue 9), the TiO2-ZnO oxide system produced a high yield of dye decomposition (70% after exposure to UV light for 6 h).
Chamanzadeh et al. [164] described the synthesis of a TiO2-ZnO oxide system with potential applications in DSSCs. Similar properties of a TiO2-ZnO binary oxide material were reported by Shanmugam et al. [152]. A DSSC was built using the commercial dye N719 (di-tetrabutylammonium cis-bis(isothiocyanato) bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II)). Based on obtained electrochemical curves it was shown that the ZnO layer improves adsorption of the dye and reduces the electron recombination rate. It was also found that the materials offered high photovoltaic efficiency (PCE = 8.3%) and charge accumulation at a level of 99% (for the material TiO2-(1)ZnO).
The combination of several catalytic techniques can provide greater possibilities of removing dangerous contaminants. The observed improvement in catalytic activity may be linked to a synergy effect, as has been widely described in literature [183,184,185]. The combination of techniques such as photocatalysis and sonocatalysis can significantly improve the efficiency of degradation of organic impurities using TiO2 [186,187]. Fatimah and Novirasari [165] described a simple synthesis of a binary photoactive material based on titanium dioxide. A combination of different catalytic techniques was studied with the aim of improving the catalytic properties. The material obtained was found to exhibit high photoactivity. A synergy effect was observed between photocatalytic degradation (yield 47%) and sonocatalytic degradation (yield 37%): the yield of phenol degradation in a sonophotocatalytic process was as high as 99%.
The following part of this review will discuss binary oxide systems based on TiO2 and ZrO2, again synthesized using the sol-gel method. A summary is given in Table 5.
The use of polymer matrices such as the commercial Pluronic P123 or F127 (poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol)—PEG-PPG-PEG) is becoming more and more popular in the synthesis of titania-based materials [169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198]. Fan et al. [29] developed a method of synthesizing a TiO2-ZrO2 oxide system using Pluronic123 (P123) and Macrogol20000 (PEG) matrices. In the first stage, Pluronic P123 and Macrogol20000 were added to titanium tetrabutoxide. Then a zirconium precursor was added to the mixture, to achieve the molar ratio Ti:Zr = 1:0.1. The product underwent aging for 24 h at room temperature, and was then dried and calcined at 800 °C for 5 h. Diffractograms showed characteristic peaks for anatase, rutile and tetragonal zirconium dioxide. BET surface areas of 136.9, 138.5 and 148.9 m2/g were measured for the systems TiO2-ZrO2(P123 + PEG), TiO2-ZrO2(PEG) and TiO2-ZrO2(P123) respectively. The synthesized oxide materials were tested in the photodegradation of rhodamine B (C.I. Basic Violet 10) in UV light and in darkness. The materials demonstrated high photocatalytic efficiency: 91.93%, 91% and 90% for TiO2-ZrO2(P123 + PEG), TiO2-ZrO2(PEG) and TiO2-ZrO2(P123) respectively.
Environmental pollution by harmful organic compounds is a severe problem for both humans and animals. Acetone is one of the most serious air pollutants in indoor environments, and its catalytic decomposition to less harmful compounds is a problem of interest to many research centres worldwide. Hernandez-Alonso et al. [188] obtained a TiO2-ZrO2 oxide system in the form of thin films, which they tested for photocatalytic activity in the degradation of acetone. Two methods of synthesis were investigated: sample M1, in which Ti0.9Zr0.1O2 was obtained by simultaneous hydrolysis of titanium tetraisopropoxide and zirconium tetrapropoxide; and sample M2, in which Ti0.9Zr0.1O2 was produced by polymerization of metal alkoxides in n-propanol. Thin layers of the binary oxide materials may be used successfully to coat Raschig rings, for example, and applied in various catalytic processes. The results of photo-oxidation of acetone confirmed the good photocatalytic properties of the synthesized materials. The method of synthesis (M1 or M2) was not found to affect the photocatalytic activity.
Kraleva et al. [189] described a simple method of obtaining a TiO2-ZrO2 oxide system using metal alkoxides (titanium tetraisopropoxide and zirconium tetrapropoxide). In the first stage the TiO2 and ZrO2 precursors underwent hydrolysis in a solution of propan-2-ol and ammonia, and the resulting sols were subjected to aging for 24 h. The TiO2-ZrO2 system was formed by mixing the sols, with the molar contribution of zirconium dioxide equal to 3, 6, 13 and 37 (the samples were denoted TiZrn, where n was the molar contribution of ZrO2). The resulting systems were dried at 100 °C for 24 h, and calcined at 550 and 700 °C for 5 h. The final binary oxide materials had a similar crystal structure to those described in earlier work. X-ray diffractograms contained characteristic peaks for anatase and rutile. For the system TiZr37 calcined at 700 °C, the presence of srilankite (Ti0.63Zr0.37Ox) was also bserved. The calcination temperature was found to have a significant effect on the crystaline structure, with anatase-rutile transformation at 700 °C. TEM images (Figure 10) showed particles with a spherical shape, having a marked tendency to agglomerate.
The BET surface areas of the TiO2-ZrO2 systems calcined at different temperatures were 26, 37, 40 and 172 m2/g for 550 °C and 26, 29, 30 and 36 m2/g for 700 °C, respectively for the systems denoted TiZr3, TiZr6, TiZr13 and TiZr37. The calcination temperature was found to affect the porous structure parameters: with an increase in temperature, the BET surface area decreased (this is related to the effect of sintering).
Atmospheric pollution is one of the most urgent present-day problems. A priority task in the synthesis of new nanomaterials is the possibility of controlling exhaust emissions. Titania-based gas sensors for the quantitative detection of various toxic and harmful gases have been widely developed in view of their high response, outstanding selectivity, excellent repeatability, and good stability. Mohammadi and Fray [190] described the use of a system combining titanium dioxide and zirconium dioxide in detecting nitrogen(IV) oxide and carbon(II) oxide. Results obtained for the sensoric properties of the materials showed that the highest response to the applied concentrations of NO2 (2 ppm) and CO (100 ppm) was given by the system that had been calcined at 500 °C (the response was 9.1 for CO and 5.1 for NO2 at a temperature of 150 °C). Detailed analysis of the morphological and crystal structures was carried out using SEM and XRD. In the crystalline structure analysis, diffraction peaks for anatase and zirconium dioxide were obtained. Based on the X-ray analysis, the average crystallite size was computed to be 5.8 nm (500 °C) and 8.5 nm (900 °C). The SEM images showed particles with a spherical shape, having a marked tendency to agglomerate. The average particle size was 20 nm (500 °C) and 36 nm (900 °C).
Naumenko et al. [191] described the synthesis of a TiO2-ZrO2 oxide system using a sol-gel method, with titanium tetraisopropoxide and zirconium tetrapropoxide as precursors of the oxides. The first stage of the synthesis was the hydrolysis of Ti[OCH(CH3)2]4 and Zr[OCH(CH3)2]4 in a mixture of deionized water and hydrochloric acid. The prepared sols were then mixed, with the Ti[OCH(CH3)2]4:Zr[OCH(CH3)2]4 molar ratio ranging from 1:0 to 0.7:0.3. The resulting materials underwent aging for 12 h, and were then calcined at 400 and 500 °C in air. The products were found to contain only the crystal structure of anatase. Detailed analysis of the results obtained for crystal structure parameters showed a change in the parameters when the molar fraction of ZrO2 was 0.3, with values of a = 3.82 Å, c = 9.726 Å (ASTM 84-1286: a = 3.7822 Å, c = 9.5023 Å). This may indicate that the ZrO2 was incorporated into the TiO2 structure to form Ti1−xZrxO2. AFM images showed the TiO2 particles to have diameters of 10–15 nm. An increase in the molar fraction of ZrO2 led to the appearance of particles with diameter 20 nm.
It is well known that in order to obtain a binary oxide material based on titania with strictly defined properties, it is necessary to choose appropriate synthesis conditions. Parameters such as the molar ratio or the temperature of thermal treatment can have a decisive influence on the resulting material. Kokporka et al. [27] investigated the effect of the molar ratio of the oxides and the temperature of calcination on the physico-chemical properties of a TiO2-ZrO2 system. They also determined the photocatalytic activity of the materials in the evolution of hydrogen. Diffractograms showed the presence of characteristic peaks for the crystal structures of anatase and rutile. The crystal structure parameters were found to be affected by the temperature of thermal treatment and by the molar ratio. Scanning electron microscope images showed particles of spherical shape with a marked tendency to agglomerate. The BET surface area, determined by low-temperature nitrogen sorption, was 130 m2/g for the systems 0.95TiO2-0.05ZrO2 and 0.2TiO2-0.8ZrO2, and 190 m2/g for 0.4TiO2-0.6ZrO2 and 0.6TiO2-0.4ZrO2. In the tests of photocatalytic activity, high yields of hydrogen were obtained in the presence of TiO2-ZrO2. The highest yield (0.61 mL/h·g) was obtained using the oxide system 0.95TiO2-0.05ZrO2 calcined at 800 °C.
Due to their high chemical and thermal stability, both Titania [199] and Zirconia [200] are used as matrices for the application of nanoparticles. Karthika et al. [192] described an application of a TiO2-ZrO2 binary material as a matrix for Sm3+ and CdS nanoparticles. TEM images showed a crystallite size of 7.8 nm and an interplanar distance of 3.533 Å, which corresponds to the CdS (100) plane. FTIR spectra contained bands for Ti–O–Ti at the wavenumber 440 cm−1, Zr–O–Ti at 663 cm−1 and Zr–O–Zr at 841 cm−1. Optical absorption and emission spectra confirmed the presence of CdS nanoparticles along with Sm3+ ions in the TiO2-ZrO2 matrices.
Due to the ever faster development of industry, it is necessary to obtain effective membranes for gas separation processes, such as the separation of CO2 and purification of organic gases under normal temperature and pressure conditions [201]. Therefore, interest in inexpensive inorganic materials such as Titania [202] and zirconia [201,203] is increasing. Fukumato et al. [193] described the use of the chelating complexes DEA and isoeugenol (ISOH) in the synthesis of a TiO2-ZrO2 oxide system. The goal of the study was to obtain amorphous microporous membranes. The single-stage synthesis involved the hydrolysis of titanium tetraisopropoxide and zirconium tetrapropoxide in a water solution of propan-1-ol. Two forms of material were obtained: with the addition of diethylamine (TiO2-ZrO2-DEA) and with the addition of isoeugenol (TiO2-ZrO2-ISOH). The synthesis product was used to prepare a membrane on a base made from Al2O3 and SiO2-ZrO2. The final stage of the process was calcination of the TiO2-ZrO2 system in membrane form at 500 and 600 °C in air and at 350 and 400 °C in a nitrogen atmosphere. The TiO2-ZrO2 membranes were tested for permeability to the gases He, N2, CO2 and N2. The oxide systems calcined in air exhibited higher permeability than those calcined in nitrogen. This is probably due to residues of the chelating compounds present in the nitrogen-calcined systems. The permeabilities of the TiO2-ZrO2-ISOH membranes were higher than those of the TiO2-ZrO2-DEA membranes, and were equal to 110, 48 and 50 [10−7 mol/m2·s·Pa] for He, CO2 and N2 respectively. Figure 11 shows SEM images of the membrane.
Anodized aluminium oxide (AAO), due to properties such as its nano-sized, self-organized, hexagonal porous structure [204], is an interesting material as a substrate for titania-based materials. Such materials can be used, for example, as sensors for hydrogen detection [204,205,206]. Qu et al. [26] investigated the use of anodic aluminium oxide as a matrix to obtain a photoactive TiO2-ZrO2 oxide system. In the first stage, AAO was immersed in zirconium nitrate for 12 h at 80 °C, after which it was dried for 1 h at 80 °C. The ZrO2-grafted AAO was then immersed in TiF4 for 9 min at 60 °C. In the final stage, the product was calcined in air for 3 h at 600 °C. Diffractograms showed the presence of characteristic diffraction peaks for anatase and for monoclinic zirconium dioxide. SEM images showed particles in the form of densely packed nanotubes with a tendency to agglomerate. Analysis of porous structure parameters showed the BET surface area of the oxide system to be 47.4 m2/g. The band gap of the system was measured at 3.22 eV. The synthesized material was tested as a catalyst in the photodegradation of methyl orange (C.I. Basic Orange 10). A high yield of photodegradation of the dye was obtained, amounting to 96% after 180 min, compared with 83% for TiO2 and 45% for ZrO2.
Another team of researchers [194] described the use of a TiO2-ZrO2 oxide system to obtained hydroxymethylfufural (HMF) from biomass. HMF is a precursor of materials such as polyesters and polymers based on furfural, as well as diesel fuels [207,208]. The binary oxide system was obtained in a single-stage sol-gel process, with the simultaneous hydrolysis of titanium tetrabutoxide and zirconium tetrabutoxide. The material was then dried at 120 °C for 12 h and calcined at 500 °C for 12 h. Diffractograms showed the presence of peaks corresponding to the structures of anatase and of monoclinic and tetragonal zirconium dioxide. When the molar fraction of ZrO2 was increased, the anatase structure became less visible. Measurements of the porous structure parameters gave BET surface areas of 68, 62 and 60 m2/g and pore volumes of 0.27, 0.18 and 0.13 cm3/g respectively for the systems TiO2-ZrO2(3/1), TiO2-ZrO2(1/1) and TiO2-ZrO2(1/3). Tests of glucose conversion gave yields of 72–86%, the main product being HMF; its maximum yield was obtained using the system TiO2-ZrO2(1/1). The use of tetrahydrofuran (THF) as a co-solvent and NaCl as the aqueous phase caused a tripling of the HMF conversion yield. The HMF yield varied depending on the co-catalysts: Amberlyst 70 (85.6%) > Nafion NR50 (73.4%) > Cs3.5SiW (35.2%) > Cs2.5PW (31.8%).
One of the pollutants of greatest importance in drinking water treatment is humic acid [209]. Many attempts have been made to use low-pressure driven membrane systems for the removal of the largest fractions of humic acid. The recently widely described photocatalytic membrane reactor is an interesting alternative to conventional techniques [210,211,212]. Khan et al. [195] described the synthesis of a TiO2-ZrO2 binary oxide system as an active photocatalyst for removing humic acid in membrane reactors. The TiO2-ZrO2 system achieved 85% efficiency in the removal of total organic carbon. The effect of double-positive ions (Ca2+) on the efficiency of the process was also investigated: addition of these ions caused an increase in the efficiency of removal of total organic carbon, possibly due to the reduced electrostatic repulsion between humic acid and the TiO2-ZrO2 system. The last part of the study concerned the effect of UV radiation, which has oxidizing properties, on a PVDF membrane placed in a reactor. FTIR spectra demonstrated that the structure of the PVDF membrane did not undergo an oxidation process under irradiation with UV light.
Literature reports on the synthesis of TiO2-ZnO and TiO2-ZrO2 oxide systems by sol-gel routes confirm that this technique provides the possibility of controlling the process through appropriate selection of process conditions, and thus enables control of the physico-chemical properties of the final product (well-defined particle size, crystal structure and surface activity). A suitable choice of process conditions (applied reagents such as organic and inorganic precursors, solvent type, type of hydrolysing and/or nucleating agents, molar ratio of oxides, pH of the reaction system, rate and direction of reagent dosing, time and temperature of the process, conditions of final heat treatment, etc.) can lead to oxide materials based on titanium dioxide together with zinc oxide or zirconium dioxide with precisely designed structure and texture. This has a significant impact on the possible applications of the produced systems in such fields as photocatalysis, electrochemistry and adsorption.

2.3. The Electrospinning Method

Another developing method of synthesis of oxide systems containing titanium dioxide is the electrospinning method, which represents a continuation of research into electrosprays. The first work on an electrospray was reported by Rayleigh, who observed that a droplet subjected to high voltage disintegrates due to the potential gradient. This line of research was taken up by Zeylen and Dole [213,214], who described droplets dissociated in an electric field to form an aerosol. The physical principles behind the two processes are similar, the only difference being the form of the product obtained: nanofibers in the case of electrospinning, and nanodroplets in the case of electrosprays. The apparatus used to obtain nanofibers by electrospinning consists of a high voltage source in the range 1–30 kV, a solution feeder (a syringe and needle), a pump and a collector [215,216,217,218]. Schematic diagrams of vertical and horizontal electrospinning processes appear in Figure 12.
In electrospinning, a solution is placed in a syringe, and then a high voltage is applied to the needle or nozzle, producing electrostatic forces which cause movement of the charges in the solution towards the collector. A dosing pump, as well as exceeding of the critical value of the field intensity, initiates the outward flow of the solution towards the collector. Under the influence of the applied voltage, the liquid is stretched into a thin stream. As the stream falls, the solvent evaporates, and the resulting nanofibers are accumulated on the collector [220,221,222].
As in the case of the other methods that we have described, such as the hydrothermal and sol-gel methods, the parameters of the electrospinning process may affect the physico-chemical properties of the final product [223,224,225,226]. Table 6 illustrates the process parameters impacting the result of electrospinning, relating to the solution, the apparatus and the external environment.
The chief advantages of the electrospinning method include the relatively simple apparatus and the fast production time. Another advantage of this method is the ability to produce nanofibers, for both laboratory and industrial use. The greatest drawbacks, however, are the instability of the stream, which affects the diameter of the resulting nanofibers, the fact that individual fibers may stick together, and the need for the polymer used in electrospinning to be soluble in the reaction medium [227,228,229,230,231].
Because of its advantages, electrospinning is becoming an interesting alternative method for obtaining oxide systems containing titanium dioxide. There has recently been a marked increase in interest in the use of electrospinning to produce oxide systems such as TiO2-ZnO and TiO2-ZrO2, or TiO2-MoS2 hybrid systems.
In the next part of this review, a survey will be made of work in which TiO2-ZnO oxide systems were obtained by the electrospinning method (Table 7).
Ammonia is a useful chemical material, widely used in many industries. However, it is a common material in industrial effluents and very dangerous to the environment and humans [239,240]. It attacks the respiratory system, skin, and eyes, and at a concentration higher than 300 ppm may lead to death [241,242]. Therefore, it is necessary to develop new materials to detect this type of impurity. Nanofibers possess favorable properties such as small diameters and large specific surface areas. For gas sensors, surface area is one of the most important factors influencing sensing performance [243]. Wang et al. [232] described the use of the electrospinning method for the saturation of a binary TiO2-ZnO material with sensory properties. The first stage was the preparation of a solution of precursors of the oxides and acetic acid with the addition of poly(vinylpyrrolidone). The solution was then transferred to a syringe pump and a voltage of +7 kV was applied. The spinning rate was 0.3 cm3/h. The oxide material was calcined at 500 °C for 3 h, and then immersed in a 0.1 mol/L solution of FeCl3 for 30 min, dried, and subjected to the action of polypyrrole (PPy) for 5 h, to produce the oxide system TiO2-ZnO-PPy. Analysis of XPS spectra showed the presence of Ti2p, O1s and Zn2p bands. SEM images of the PPy-modified materials showed the oxide system to have the form of black nanofibers with diameter 180 nm and lengths of several millimeters. The unmodified fibers had a uniform smooth surface. TEM images showed the oxide material to have a core-shell structure, where the thickness of the PPy shell was 7 nm. The TiO2-ZnO-PPy system was tested as an ammonia sensor. It was found to have a fast response rate and a detection limit of 60 ppb. The results obtained were probably influenced by the reduction of diffusion resistance caused by the porous PPy shell.
Park et al. [233] described a combination of the methods of electrospinning and atomic layer deposition (ALD) in the synthesis of a TiO2-ZnO binary oxide system. In the first stage, TiO2 nanofibers were obtained by electrospinning from a solution of titanium tetraisopropoxide and poly(vinyl acetate), with an applied voltage of +10 kV and a constant spinning rate of 0.2 cm3/h. The resulting nanofibers were calcined at 600 °C for 8 h. The TiO2 fibers were then coated with ZnO using the ALD method. The number of ALD cycles was in the range 50–400. Diffractograms showed the TiO2-ZnO system to contain the crystal structure of anatase and wurtzite. The HRTEM results were consistent with those of XRD. SEM images showed that the increment in the ZnO layer per ALD cycle was 0.66 nm. Based on EDS (Energy-dispersive X-ray spectroscopy) mapping a core-shell structure was demonstrated, with a core of TiO2 and a shell of ZnO. As in the work of Wang et al. [232], the obtained TiO2-ZnO nanofibers were tested as a chemical sensor. Park et al. [233] assessed the sensory properties of their product using oxygen. It was found to have good properties for the detection of oxygen, and the sensor’s response varied with changes in the oxygen concentration. The tested chemical sensor demonstrated high stability and repeatability.
Surveys of the subject literature have already shown that binary oxide materials such as TiO2-ZnO obtained by the hydrothermal method [62] and by the sol-gel method [156,161,163] have good photocatalytic properties in the degradation of methylene blue (C.I. Basic Blue 9). Materials of the same type synthesized by other methods can be expected to exhibit similar properties. A team of researchers [234] described the synthesis of a TiO2-ZnO system by electrospinning and its use to catalyze the photodegradation of C.I. Basic Blue 9. X-ray diffraction results showed the presence of reflection corresponding to the structure of anatase, wurtzite and zinc titanate. From TEM images, the mean particle diameter was determined to be 20 nm. Using low-temperature nitrogen sorption, the BET surface area was measured at 203 m2/g and the pore volume at 0.24 cm3/g. The TiO2-ZnO oxide system synthesized by electrospinning demonstrated high photocatalytic activity, at a similar level to the same material produced using other methods.
Rhodamine B is widely used as a colorant in textiles and foodstuffs, and is also a well-known water tracer fluorescent [244]. Many scientific studies have demonstrated the harmful effects of rhodamine B on humans and the environment. Its carcinogenicity, reproductive and developmental toxicity, neurotoxicity and chronic toxicity to humans and animals have been experimentally proven [245,246,247]. Due to the harmful effects of this material, it is necessary to develop new nanomaterials enabling its effective degradation [248,249]. Liu et al. [235] used cellulose acetate as a matrix for the synthesis of TiO2-ZnO oxide systems with TiO2:ZnO mass ratios of 87.6:12.4, 84.2:15.8 and 77.9:22.1. XRD results showed the presence of diffraction peaks corresponding to anatase, wurtzite and zinc titanate. Based on TEM images, the interplanar distance was measured at 3.56 Å, corresponding to dimension d of the (101) plane of the anatase crystal structure. XPS spectra showed the presence of Ti2p, Zn2p and O1s bands. The obtained materials were used in the process of photodegradation of rhodamine B (C.I. Basic Violet 10) and phenol. All of the oxide systems demonstrated higher photocatalytic activity than pure TiO2. The highest photodegradation yield was obtained in the presence of the material synthesized in the molar ratio TiO2:ZnO = 84.2:15.8. The researchers suggested that excessive ZnO may restrict access to recombination sites, thus reducing the photocatalytic performance.
Kanjwal et al. [236] compared TiO2-ZnO binary oxide materials synthesized by the hydrothermal and electrospinning methods. In the electrospinning process, vinyl acetate was dissolved in DMF, and then titanium tetraisopropoxide, zinc oxide and acetic acid were added. The solution was transferred to a syringe, and a voltage of +12 kV was applied. The material was dried at 80 °C for 24 h and calcined at 600 °C for 1 h. Hydrothermal synthesis was carried out at 150 °C for 1 h, using nanofibers of TiO2 and zinc nitrate. Diffractograms revealed the characteristic crystal structures of wurtzite (after electrospinning) and anatase (after hydrothermal synthesis). SEM and TEM images showed that the electrospun oxide system consisted of nanofibers with a smooth surface, while in the case of the material obtained hydrothermally, ZnO nanoparticles were observed on the TiO2 surface. Physico-chemical analysis of the binary oxide materials showed marked differences in their morphological and crystal structure. Their photocatalytic properties were assessed in the degradation of rhodamine B (C.I. Basic Violet 10). Both the electrospun and hydrothermally synthesized materials demonstrated superior photocatalytic activity to pure TiO2 or ZnO. The hydrothermally treated materials produced higher degradation yields (100% after 105 min of exposure to light) than the electrospun materials; this may be linked to differences in the crystal structures of the oxide systems.
Today, various chemical pollutants such as dyes are attracting increased attention owing to their significant effects on public health and the environment. The efficient degradation of this type of pollution from wastewater is becoming an urgent and challenging problem. Many centers are conducting research on the use of electrospinning for the synthesis of titania-based materials and their use in wastewater photodegradation [250,251,252]. Li et al. [237] developed a synthesis process based on electrospinning for a photocatalytic TiO2-ZnO oxide system. The resulting system demonstrated very good photocatalytic properties in the degradation of rhodamine B (C.I. Basic Violet 10), producing a higher degradation yield than pure TiO2 or ZnO. It also exhibited high photocatalytic stability over five consecutive cycles. More detailed physico-chemical analysis was performed using the XRD, SEM and TEM techniques. X-ray diffraction results demonstrated the presence of characteristic diffraction peaks corresponding to anatase and wurtzite. The SEM and TEM images showed smooth nanofibers consisting of single nanoparticles.
Araújo et al. [238] synthesized a TiO2-ZnO oxide material using the methods of electrospinning and atomic layer deposition. In the first stage, a solution of TiO2 (P25) and poly(methacrylic acid-co-methyl methacrylate) 1:1 block copolymer in ethanol was subjected to a constant voltage of +15 kV for 5 min. Next, zinc acetate was dissolved in water and mixed for 15 min at 85 °C, after which triethylamine was added and the mixture incubated for 3 h (solution 1). At the same time, zinc acetate was dissolved in deionized water with the addition of hexamethyltetramine (solution 2). The obtained TiO2 nanofibers were immersed in solution 1, dried at 120 °C for 1 h, then added to solution 2 and kept at 85 °C for 24 h. The synthesized oxide material was calcined for one hour at 500 °C. Diffractograms revealed the presence of characteristic reflections for the crystalline structure of anatase, rutile and wurtzite. SEM images (Figure 13a–d) showed that the TiO2 nanofibers were coated with ZnO nanorods. The EDX spectrum (Figure 13e–f) showed characteristic bands for Ti, O and Zn. The materials were tested as a sensor for the detection of moisture. Materials with one and three layers of ZnO produced a high response to changes in humidity. The three-layer material exhibited significantly greater sensitivity in the humidity range 40–100%. It also demonstrated significantly better parameters in moisture detection than the sensors described in a previous study [253]. The researchers suggest that the increase in ZnO content improves the porous structure, and thus increases the sensitivity [253].
Table 8 contains a review of the literature concerning the synthesis of TiO2-ZrO2 oxide systems using the electrospinning method.
Humidity is a very common component in our environment that can affect not only human life, but also many industrial processes. Therefore, the measurement and control of humidity are of great importance. The design of a successful humidity sensor is a complex process, as the material used to build the sensor must offer properties including linear response, high sensitivity, fast response time, chemical and physical stability, wide operating range and low cost. Su et al. [254] developed a simple synthesis method involving electrospinning of a TiO2-ZrO2 oxide system with defined moisture detection properties. The impedance was shown to differ by four orders of magnitude, from 105 to 101 kΩ, in the humidity range 11–97%, which indicates correct operation of the sensor. Good repeatability was demonstrated for the obtained results, and the influence of temperature on the sensor was shown to be small. To determine the physico-chemical characteristics of the sensor material, analysis was performed using XRD, SEM and TEM. The XRD results indicated the crystal structure of anatase and tetragonal zirconium dioxide. Based on the SEM images it was concluded that the binary oxide material contained fibers originating from both ZrO2 and TiO2. The diameters of the nanofibers were found to be in the range 240–400 nm in the case of ZrO2, and 60–200 nm in the case of TiO2. Transmission electron microscopy results (Figure 14) showed the nanofibers to be built of numerous small particles.
Recent years have seen a marked increase in interest in proton exchange membrane fuel cells (PEMFCs). The development of the ionomer membrane is crucial for the reliability of such cells and their high-volume commercialization. This type of membrane must attain a high proton conductivity and a low gas permeability, offering at the same time high mechanical, chemical and thermal stability, a low degree of deformation due to water absorption, and low costs of production so that it can be used globally [256,257,258]. Lee et al. [255] described the use of a TiO2-ZrO2 binary oxide system to prepare aquivion (KNF Neuberger GmbH) membrane for use in a single fuel cell. In the course of comprehensive physico-chemical analysis, diffractograms were obtained which indicated an amorphous structure for the material calcined at 500 °C, and the crystal structure of srilankite (Ti0.5Zr0.5O2) for the oxide systems calcined at 700 and 1000 °C. In view of these results, further analysis was carried out only on the material with an amorphous structure, since that material had the highest specific surface area among the synthesized oxide systems. Based on SEM images, the average fiber diameter was measured at 497 nm before calcination and 344 nm after calcination. The synthesized TiO2-ZrO2 oxide system calcined at 500 °C was modified with 85% phosphoric acid to increase its proton conductivity. The XPS spectrum before modification showed two bands at 530.35 and 532.8 eV, corresponding to O–M bonds (M = Ti, Zr) and O–H bonds. On the spectrum of the modified material a new band appeared at 531.36 eV, originating from P–O bonds. The use of the oxide material to build the membrane eliminated the effect of cracking, and the membrane modified with phosphate groups showed improved proton conductivity compared with pure TiO2-ZrO2 and aquivion. The composite membrane demonstrated good durability, as confirmed by the results of accelerated lifetime (ALT) measurements.
The above review of the literature on the formation of TiO2-ZnO and TiO2-ZrO2 oxide systems by the electrospinning method confirms that this is a new and promising method which leads to the effective synthesis of materials in the form of nanofibers with defined chemical composition, porosity, geometry and fiber dimensions. Moreover, the method does not place excessive limitations on researchers. The sole problem is the selection of optimum process conditions, including the type and concentration of solvent, temperature, solution viscosity, average molecular weight of polymer, applied voltage, and flow rate. It has been shown that electrospun TiO2-ZnO and TiO2-ZrO2 fibers may be used, among other things, as sensors for the detection of various types of gas, as photocatalysts in the degradation of certain types of organic pollutants, and as membranes for the production of fuel cells.

3. Conclusions

The literature review that has been presented here reflects the search for innovative combinations of TiO2 with other materials, which may cause morphological changes in the crystalline phases of titanium dioxide and changes in its electron structure, leading to—among other things—improvement in its photocatalytic activity and spectral sensitivity. Studies being carried out at multiple research centers worldwide are largely focused on the choice of appropriate methods for the synthesis or modification of titanium dioxide, and on the selection of suitable components to be combined with TiO2 to produce advanced multifunctional hybrid materials.
It is widely known that the physico-chemical as well as the morphological-structural properties of materials based on titanium dioxide, such as particle shape and size, degree of crystallinity, surface composition, pore size distribution, specific surface area, etc., are strongly dependent on the suitable specification of synthesis conditions, such as the nature and composition of the precursors, solvent and complexing/templating agent, and the conditions of hydrolysis and calcination. Suitably chosen process conditions for the synthesis of oxide and hybrid materials based on TiO2 make it possible to design their physico-chemical properties.
Issues relating to the creation of a new generation of modified forms of TiO2 are of huge importance, particularly in view of the growing need for active photocatalysts functioning in visible light, as well as high-performance electrode materials.

Author Contributions

A.K.—described a review on obtaining hybrid systems using the hydrothermal method and electrospinning; K.S.-C.—prepared a review on the synthesis of oxide systems using the sol-gel method; T.J.—performed critical revision and supervised all research aspects.

Funding

Presented research was financed within Poznan University of Technology research grant No. 03/32/DSPB/0806/2018.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Stoyanova, A.; Hitkova, H.; Bachvarova-Nedelcheva, A.; Iordanova, R.; Ivanova, N.; Sredkova, M. Synthesis and antibacterial activity of TiO2/ZnO nanocomposites prepared via nonhydrolytic route. J. Chem. Technol. Metall. 2013, 48, 154–161. [Google Scholar]
  2. Bach, U.; Corr, D.; Lupo, D.; Pichot, F.; Ryan, M. Nanomaterials-based electrochromics for paper-quality displays. Adv. Mater. 2002, 14, 845–848. [Google Scholar] [CrossRef]
  3. Gouma, P.I.; Mills, M.J.; Sandhage, K.H. Fabrication of free-standing titania-based gas sensors by the oxidation of metallic titanium foils. J. Am. Ceram. Soc. 2000, 83, 1007–1009. [Google Scholar] [CrossRef]
  4. Oey, C.C.; Djurišić, A.B.; Wang, H.; Man, K.K.Y.; Chan, W.K.; Xie, M.H.; Leung, Y.H.; Pandey, A.; Nunzi, J.-M.; Chui, P.C. Polymer-TiO2 solar cells: TiO2 interconnected network for improved cell performance. Nanotechnology 2006, 17, 706–713. [Google Scholar] [CrossRef]
  5. Polleux, J.; Gurlo, A.; Barsan, N.; Weimar, U.; Antonietti, M.; Niederberger, M. Template-free synthesis and assembly of single-crystalline tungsten oxide nanowires and their gas-sensing properties. Angew. Chem. Int. Ed. 2006, 45, 261–265. [Google Scholar] [CrossRef] [PubMed]
  6. McCullagh, C.; Robertson, J.M.C.; Bahnemann, D.W.; Robertson, P.K.J. The application of TiO2 photocatalysis for disinfection of water contaminated with pathogenic micro-organisms: A review. Res. Chem. Intermed. 2007, 33, 359–375. [Google Scholar] [CrossRef]
  7. Sobczyński, A.; Dobosz, A. Water purification by photocatalysis on semiconductors. Pol. J. Environ. Stud. 2001, 10, 195–205. [Google Scholar]
  8. Xu, X.; Wang, J.; Tian, J.; Wang, X.; Dai, J.; Liu, X. Hydrothermal and post-heat treatments of TiO2/ZnO composite powder and its photodegradation behavior on methyl orange. Ceram. Int. 2011, 37, 2201–2206. [Google Scholar] [CrossRef]
  9. Liao, D.L.; Badour, C.A.; Liao, B.Q. Preparation of nanosized TiO2/ZnO composite catalyst and its photocatalytic activity for degradation of methyl orange. J. Photochem. Photobiol. A 2008, 194, 11–19. [Google Scholar] [CrossRef]
  10. Siwinska-Stefanska, K.; Kubiak, A.; Kurc, B.; Moszynski, D.; Goscianska, J.; Jesionowski, T. An active anode material based on titania and zinc oxide hybrids fabricated via hydrothermal route: Comperehensive physicochemical and electrochemical evaluations. J. Electrochem. Soc. 2018, 165, A3056–A3066. [Google Scholar] [CrossRef]
  11. Siwinska-Stefanska, K.; Paukszta, D.; Piasecki, A.; Jesionowski, T. Synthesis and physicochemical characteristics of titanium dioxide doped with selected metals. Physicochem. Probl. Miner. Process. 2014, 50, 265–276. [Google Scholar] [CrossRef]
  12. Raghupathi, K.R.; Koodali, R.T.; Manna, A.C. Size-dependent bacterial growth inhibition and mechanism of antibacterial activity of zinc oxide nanoparticles. Langmuir 2011, 27, 4020–4028. [Google Scholar] [CrossRef] [PubMed]
  13. Xia, Y.; Wang, J.; Chen, R.; Zhou, D.; Xiang, L. A review on the fabrication of hierarchical ZnO nanostructures for photocatalysis application. Crystals 2016, 6, 148. [Google Scholar] [CrossRef]
  14. Wang, J.; Gao, L. Hydrothermal synthesis and photoluminescence properties of ZnO nanowires. Solid State Commun. 2004, 132, 269–271. [Google Scholar] [CrossRef]
  15. Zhang, J.; Wang, S.; Xu, M.; Wang, Y.; Zhu, B.; Zhang, S.; Huang, W.; Wu, S. Hierarchically porous ZnO architectures for gas sensor application. Cryst. Growth Des. 2009, 9, 3532–3537. [Google Scholar] [CrossRef]
  16. Wang, L.; Kang, Y.; Liu, X.; Zhang, S.; Huang, W.; Wang, S. ZnO nanorod gas sensor for ethanol detection. Sens. Actuators B Chem. 2012, 162, 237–243. [Google Scholar] [CrossRef]
  17. Kolodziejczak-Radzimska, A.; Jesionowski, T. Zinc oxide-from synthesis to application: A review. Materials 2014, 7, 2833–2881. [Google Scholar] [CrossRef] [PubMed]
  18. Takenaka, S.; Shimizu, T.; Otsuka, K. Complete removal of carbon monoxide in hydrogen-rich gas stream through methanation over supported metal catalysts. Int. J. Hydrogen Energy 2004, 29, 1065–1073. [Google Scholar] [CrossRef]
  19. Yang, Y.; Yang, H.; Yang, M.; Ilu, Y.; Shen, G.; Yu, R. Amperometric glucose biosensor based on a surface treated nanoporous ZrO2/Chitosan composite film as immobilization matrix. Anal. Chim. Acta 2004, 525, 213–220. [Google Scholar] [CrossRef]
  20. Liu, Q.; Long, S.; Lv, H.; Wang, W.; Niu, J.; Huo, Z.; Chen, J.; Liu, M. Controllable growth of nanoscale conductive filaments in solid-electrolyte-based ReRAM by using a metal nanocrystal covered bottom electrode. ACS Nano 2010, 4, 6162–6188. [Google Scholar] [CrossRef] [PubMed]
  21. Apostolescu, N.; Geiger, B.; Hizbullah, K.; Jan, M.T.; Kureti, S.; Reichert, D.; Schott, F.; Weisweiler, W. Selective catalytic reduction of nitrogen oxides by ammonia on iron oxide catalysts. Appl. Catal. B 2006, 62, 104–114. [Google Scholar] [CrossRef]
  22. Wysokowski, M.; Szalaty, T.J.; Jesionowski, T.; Motylenko, M.; Rafaja, D.; Koltsov, I.; Stöcker, H.; Bazhenov, V.V.; Ehrlich, H.; Stelling, A.L.; et al. Extreme biomimetic approach for synthesis of nanocrystalline chitin-(Ti,Zr)O2 multiphase composites. Mater. Chem. Phys. 2017, 188, 115–124. [Google Scholar] [CrossRef]
  23. Emeline, A.; Kataeva, G.V.; Litke, A.S.; Rudakova, A.V.; Ryabchuk, V.K.; Serpone, N. Spectroscopic and photoluminescence studies of a wide band gap insulating material: Powdered and colloidal ZrO2 sols. Laungmir 1998, 7463, 5011–5022. [Google Scholar] [CrossRef]
  24. Siwińska-Stefańska, K.; Kurc, B. A composite TiO2-SiO2-ZrO2 oxide system as a high-performance anode material for lithium-ion batteries. J. Electrochem. Soc. 2017, 164, A728–A734. [Google Scholar] [CrossRef]
  25. Jangra, S.L.; Stalin, K.; Dilbaghi, N.; Kumar, S.; Tawale, J.; Singh, S.P.; Pasricha, R. Antimicrobial activity of zirconia (ZrO2) nanoparticles and zirconium complexes. J. Nanosci. Nanotechnol. 2012, 12, 7105–7112. [Google Scholar] [CrossRef] [PubMed]
  26. Qu, X.; Xie, D.; Cao, L.; Du, F. Synthesis and characterization of TiO2/ZrO2 coaxial core-shell composite nanotubes for photocatalytic applications. Ceram. Int. 2014, 40, 12647–12653. [Google Scholar] [CrossRef]
  27. Kokporka, L.; Onsuratoom, S.; Puangpetch, T.; Chavadej, S. Sol-gel-synthesized mesoporous-assembled TiO2-ZrO2 mixed oxide nanocrystals and their photocatalytic sensitized H2 production activity under visible light irradiation. Mater. Sci. Semicond. Process. 2013, 16, 667–678. [Google Scholar] [CrossRef]
  28. Zhou, W.; Liu, K.; Fu, H.; Pan, K.; Zhang, L.; Wang, L.; Sun, C. Multi-modal mesoporous TiO2-ZrO2 composites with high photocatalytic activity and hydrophilicity. Nanotechnology 2008, 19, 035610. [Google Scholar] [CrossRef] [PubMed]
  29. Fan, M.; Hu, S.; Ren, B.; Wang, J.; Jing, X. Synthesis of nanocomposite TiO2/ZrO2 prepared by different templates and photocatalytic properties for the photodegradation of Rhodamine B. Powder Technol. 2013, 235, 27–32. [Google Scholar] [CrossRef]
  30. Liu, H.; Su, Y.; Hu, H.; Cao, W.; Chen, Z. An ionic liquid route to prepare mesoporous ZrO2-TiO2 nanocomposites and study on their photocatalytic activities. Adv. Powder Technol. 2013, 24, 683–688. [Google Scholar] [CrossRef]
  31. Siwińska-Stefańska, K.; Kubiak, A.; Piasecki, A.; Goscianska, J.; Nowaczyk, G.; Jurga, S.; Jesionowski, T. TiO2-ZnO binary oxide systems: Comprehensive characterization and tests of photocatalytic activity. Materials 2018, 11, 841. [Google Scholar] [CrossRef]
  32. Tian, J.; Chen, L.; Dai, J.; Wang, X.; Yin, Y.; Wu, P. Preparation and characterization of TiO2, ZnO, and TiO2/ZnO nanofilms via sol-gel process. Ceram. Int. 2009, 35, 2261–2270. [Google Scholar] [CrossRef]
  33. Wang, L.; Wang, C.; Liu, W.; Chen, Q.; He, M. Visible-light-induced aerobic thiocyanation of indoles using reusable TiO2/MoS2 nanocomposite photocatalyst. Tetrahedron Lett. 2016, 57, 1771–1774. [Google Scholar] [CrossRef]
  34. Zhu, X.; Yang, C.; Xiao, F.; Wang, J.; Su, X. Synthesis of nano-TiO2-decorated MoS2 nanosheets for lithium ion batteries. New J. Chem. 2015, 39, 683–688. [Google Scholar] [CrossRef]
  35. Roy, R. Accelerating the kinetics of low-temperature inorganic syntheses. J. Solid State Chem. 1994, 111, 11–17. [Google Scholar] [CrossRef]
  36. Rabenau, A. The role of hydrothermal synthesis in preparative chemistry. Angew. Chem. Int. 1985, 24, 1026–1040. [Google Scholar] [CrossRef]
  37. Wang, X.; Li, Y. Selected-control hydrothermal synthesis of α- and β-MnO2 single crystal nanowires. J. Am. Chem. Soc. 2002, 124, 2880–2881. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, N.; Cai, Y.; Zhang, R.Q. Growth of nanowires. Mater. Sci. Eng. R Rep. 2008, 60, 1–51. [Google Scholar] [CrossRef]
  39. Yu, J.; Wang, G.; Cheng, B.; Zhou, M. Effects of hydrothermal temperature and time on the photocatalytic activity and microstructures of bimodal mesoporous TiO2 powders. Appl. Catal. B Environ. 2007, 69, 171–180. [Google Scholar] [CrossRef]
  40. Jung, J.; Perrut, M. Particle design using supercritical fluids: Literature and patent survey. J. Supercrit. Fluids 2001, 20, 179–219. [Google Scholar] [CrossRef]
  41. Li, W.J.; Shi, E.W.; Zhong, W.Z.; Yin, Z.W. Growth mechanism and growth habit of oxide crystals. J. Cryst. Growth 1999, 203, 186–196. [Google Scholar] [CrossRef]
  42. Bavykin, D.V.; Parmon, V.N.; Lapkin, A.A.; Walsh, F.C. The effect of hydrothermal conditions on the mesoporous structure of TiO2 nanotubes. J. Mater. Chem. 2004, 14, 3370–3377. [Google Scholar] [CrossRef]
  43. Byrappa, K.; Yoshimura, M. Hydrothermal technology—Principles and applications. In Handbook of Hydrothermal Technology, 2nd ed.; William Andrew Publishing LLC.: Norwich, NY, USA, 2013; pp. 1–52. [Google Scholar]
  44. Byrappa, K.; Yoshimura, M. Physical chemistry of hydrothermal growth of crystals. In Handbook of Hydrothermal Technology, 2nd ed.; William Andrew Publishing LLC.: Norwich, NY, USA, 2013; pp. 139–175. [Google Scholar]
  45. Shen, J.; Yan, B.; Shi, M.; Ma, H.; Li, N.; Ye, M. One step hydrothermal synthesis of TiO2-reduced graphene oxide sheets. J. Mater. Chem. 2011, 21, 3415–3421. [Google Scholar] [CrossRef]
  46. Riman, R.E.; Suchanek, W.L.; Lencka, M.M. Hydrothermal crystallization of ceramics. Ann. Chim. Sci. Mater. 2002, 27, 15–36. [Google Scholar] [CrossRef]
  47. Zhang, Y.X.; Li, G.H.; Jin, Y.X.; Zhang, Y.; Zhang, J.; Zhang, L.D. Hydrothermal synthesis and photoluminescence of TiO2 nanowires. Chem. Phys. Lett. 2002, 365, 300–304. [Google Scholar] [CrossRef]
  48. Yuan, Z.Y.; Su, B.L. Titanium oxide nanotubes, nanofibers and nanowires. Colloids Surf. A Physicochem. Eng. Asp. 2004, 241, 173–183. [Google Scholar] [CrossRef]
  49. Shandilya, M.; Rai, R.; Singh, J. Review: Hydrothermal technology for smart materials. Adv. Appl. Ceram. 2016, 115, 354–376. [Google Scholar] [CrossRef]
  50. Zhang, L.; Jeem, M.; Okamoto, K.; Watanabe, S. Photochemistry and the role of light during the submerged photosynthesis of zinc oxide nanorods. Sci. Rep. 2018, 8, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Ding, S.; Luan, D.; Boey, F.Y.C.; Chen, J.S.; Lou, X.W. SnO2 nanosheets grown on graphene sheets with enhanced lithium storage properties. Chem. Commun. 2011, 47, 7155–7157. [Google Scholar] [CrossRef] [PubMed]
  52. Komarneni, S.; Roy, R.; Li, Q.H. Microwave-hydrothermal synthesis of ceramic powders. Mater. Res. Bull. 1992, 27, 1393–1405. [Google Scholar] [CrossRef]
  53. Tompsett, G.A.; Conner, W.C.; Yngvesson, K.S. Microwave synthesis of nanoporous materials. Chem. Phys. Chem. 2006, 7, 296–319. [Google Scholar] [CrossRef] [PubMed]
  54. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  55. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  56. Cheng, P.; Wang, Y.; Xu, L.; Sun, P.; Su, Z.; Jin, F.; Liu, F.; Sun, Y.; Lu, G. 3D TiO2/ZnO composite nanospheres as an excellent electron transport anode for efficient dye-sensitized solar cells. RSC Adv. 2016, 6, 51320–51326. [Google Scholar] [CrossRef]
  57. Cheng, P.; Du, S.; Cai, Y.; Liu, F.; Sun, P.; Zheng, J.; Lu, G. Tripartite layered photoanode from hierarchical anatase TiO2 urchin-like spheres and P25: A candidate for enhanced efficiency dye sensitized solar cells. J. Phys. Chem. C 2013, 117, 24150–24156. [Google Scholar] [CrossRef]
  58. Li, Y. Synthesis and characterization of TiO2 doped ZnO microtubes. Chin. J. Chem. Phys. 2010, 23, 358–362. [Google Scholar] [CrossRef]
  59. Vlazan, P.; Ursu, D.H.; Irina-Moisescu, C.; Miron, I.; Sfirloaga, P.; Rusu, E. Structural and electrical properties of TiO2/ZnO core–shell nanoparticles synthesized by hydrothermal method. Mater. Charact. 2015, 101, 153–158. [Google Scholar] [CrossRef]
  60. Zhang, M.; An, T.; Liu, X.; Hu, X.; Sheng, G.; Fu, J. Preparation of a high-activity ZnO/TiO2 photocatalyst via homogeneous hydrolysis method with low temperature crystallization. Mater. Lett. 2010, 64, 1883–1886. [Google Scholar] [CrossRef]
  61. Wang, Y.; Zhu, S.; Chen, X.; Tang, Y.; Jiang, Y.; Peng, Z.; Wang, H. One-step template-free fabrication of mesoporous ZnO/TiO2 hollow microspheres with enhanced photocatalytic activity. Appl. Surf. Sci. 2014, 307, 263–271. [Google Scholar] [CrossRef]
  62. Rusu, E.; Ursaki, V.; Gutul, T.; Vlazan, P.; Siminel, A. Characterization of TiO2 nanoparticles and ZnO/TiO2 composite obtained by hydrothermal method. In 3rd International Conference on Nanotechnologies and Biomedical Engineering; Sontea, V., Tiginyanu, I., Eds.; Springer: Singapore, 2016; Volume 55, pp. 93–96. [Google Scholar] [CrossRef]
  63. Chen, D.; Zhang, H.; Hu, S.; Li, J. Preparation and enhanced photoelectrochemical performance of coupled bicomponent. J. Phys. Chem. C 2008, 112, 117–122. [Google Scholar] [CrossRef]
  64. Wang, C.; Hwang, W.; Chang, K.; Ko, H.; Hsi, C.; Huang, H.; Wang, M. Formation and morphology of Zn2Ti3O8 powders using hydrothermal process without dispersant agent or mineralizer. Int. J. Mol. Sci. 2011, 12, 935–945. [Google Scholar] [CrossRef] [PubMed]
  65. Ashok, C.; Venkateswara Rao, K. ZnO/TiO2 nanocomposite rods synthesized by microwave-assisted method for humidity sensor application. Superlattices Microstruct. 2014, 76, 46–54. [Google Scholar] [CrossRef]
  66. Divya, K.S.; Marilyn, M.X.; Vandana, P.V.; Rethu, V.N.; Suresh, M. A quaternary TiO2 /ZnO/RGO/Ag nanocomposite with enhanced visible light photocatalytic performance. New J. Chem. 2017, 41, 6445–6454. [Google Scholar] [CrossRef]
  67. Cho, I.S.; Chen, Z.; Forman, A.J.; Kim, D.R.; Rao, P.M.; Jaramillo, T.F.; Zheng, X. Branched TiO2 nanorods for photoelectrochemical hydrogen production. Nano Lett. 2011, 11, 4978–4984. [Google Scholar] [CrossRef] [PubMed]
  68. Macák, J.M.; Tsuchiya, H.; Schmuki, P. High-aspect-ratio TiO2 nanotubes by anodization of titanium. Angew. Chem. 2005, 44, 2100–2102. [Google Scholar] [CrossRef] [PubMed]
  69. Hu, Y.S.; Kienle, L.; Guo, Y.G.; Maier, J. High lithium electroactivity of nanometer-sized rutile TiO2. Adv. Mater. 2006, 18, 1421–1426. [Google Scholar] [CrossRef]
  70. Yang, P.; Xiao, X.; Li, Y.; Ding, Y.; Qiang, P.; Tan, X.; Mai, W.; Lin, Z.; Wu, W.; Li, T.; et al. Hydrogenated ZnO core-shell nanocables for flexible supercapacitors and self-powered systems. ACS Nano 2013, 7, 2617–2626. [Google Scholar] [CrossRef] [PubMed]
  71. Keis, K.; Magnusson, E.; Lindstr, H.; Lindquist, S.; Hagfeldt, A.; Lindström, H.; Lindquist, S.; Hagfeldt, A. A 5% efficient photoelectrochemical solar cell based on nanostructured ZnO electrodes. Sol. Energy Mater. Sol. Cells 2002, 73, 51–58. [Google Scholar] [CrossRef]
  72. Zhang, L.; Jiang, Y.; Ding, Y.; Povey, M.; York, D. Investigation into the antibacterial behaviour of suspensions of ZnO nanoparticles (ZnO nanofluids). J. Nanopart. Res. 2007, 9, 479–489. [Google Scholar] [CrossRef]
  73. Wang, C.; Ao, Y.; Wang, P.; Hou, J.; Qian, J. Preparation, characterization and photocatalytic activity of the neodymium-doped TiO2 hollow spheres. Appl. Surf. Sci. 2010, 257, 227–231. [Google Scholar] [CrossRef]
  74. Chen, J.; Nie, X.; Shi, H.; Li, G.; An, T. Synthesis of TiO2 hollow sphere multimer photocatalyst by etching titanium plate and its application to the photocatalytic decomposition of gaseous styrene. Chem. Eng. J. 2013, 228, 834–842. [Google Scholar] [CrossRef]
  75. Sui, Y.; Yang, H.; Fu, W.; Xu, J.; Chang, L.; Zhu, H.; Yu, Q.; Li, M.; Zou, G. Preparation and characterization of hollow glass microspheres/ZnO composites. J. Alloys Compd. 2009, 469, 2–6. [Google Scholar] [CrossRef]
  76. Das, S.; Chatterjee, S.; Pramanik, S.; Devi, P.S.; Kumar, G.S. A new insight into the interaction of ZnO with calf thymus DNA through surface defects. J. Photochem. Photobiol. B Biol. 2018, 178, 339–347. [Google Scholar] [CrossRef] [PubMed]
  77. Ghosh Chaudhuri, R.; Paria, S. Core/shell nanoparticles: Classes, properties, synthesis mechanisms, characterization, and applications. Chem. Rev. 2012, 112, 2373–2433. [Google Scholar] [CrossRef] [PubMed]
  78. Rafatullah, M.; Sulaiman, O.; Hashim, R.; Ahmad, A. Adsorption of methylene blue on low-cost adsorbents: A review. J. Hazard. Mater. 2010, 177, 70–80. [Google Scholar] [CrossRef] [PubMed]
  79. Garg, V.K.; Amita, M.; Kumar, R.; Gupta, R. Basic dye (methylene blue) removal from simulated wastewater by adsorption using Indian rosewood sawdust: A timber industry waste. Dyes Pigments 2004, 63, 243–250. [Google Scholar] [CrossRef]
  80. Murugan, A.V.; Samuel, V.; Ravi, V. Synthesis of nanocrystalline anatase TiO2 by microwave hydrothermal method. Mater. Lett. 2006, 60, 479–480. [Google Scholar] [CrossRef]
  81. Cabello, G.; Davoglio, R.A.; Pereira, E.C. Microwave-assisted synthesis of anatase-TiO2 nanoparticles with catalytic activity in oxygen reduction. J. Electroanal. Chem. 2017, 794, 36–42. [Google Scholar] [CrossRef]
  82. Leahy, J.J. ZrO2-modified TiO2 nanorod composite: Hydrothermal synthesis, characterization and application in esterification of organic acid. Mater. Chem. Phys. 2014, 145, 82–89. [Google Scholar] [CrossRef]
  83. Tomar, L.J.; Chakrabarty, B.S. Synthesis, structural and optical properties of TiO2-ZrO2 nanocomposite by hydrothermal method. Adv. Mater. Lett. 2013, 4, 64–67. [Google Scholar] [CrossRef]
  84. Kubo, K.; Hosokawa, S.; Furukawa, S.; Inoue, M. Synthesis of ZrO2-TiO2 solid solutions by various synthetic methods in the region of high zirconium contents. J. Mater. Sci. 2008, 2198–2205. [Google Scholar] [CrossRef]
  85. Hirano, M.; Nakahara, C.; Ota, K.; Tanaike, O.; Inagaki, M. Photoactivity and phase stability of ZrO2-doped anatase-type TiO2 directly formed as nanometer-sized particles by hydrolysis under hydrothermal conditions. J. Solid State Chem. 2003, 170, 39–47. [Google Scholar] [CrossRef]
  86. Tomar, L.J.; Bhatt, P.J.; Desai, R.K.; Chakrabarty, B.S.; Panchal, C.J. Improved conversion efficiency of dye sensitized solar cell using Zn doped TiO2-ZrO2 nanocomposite. AIP Conf. Proc. 2016, 1731, 50132. [Google Scholar] [CrossRef]
  87. Yao, B.; Han, X.; Ying, L.; Peng, C.; Zhang, C. Hydrothermal synthesis and photocatalytic activity of TiO2-ZrO2 hybrid composite microspheres. Mater. Sci. Forum 2016, 852, 257–263. [Google Scholar] [CrossRef]
  88. Caillot, T.; Salama, Z.; Chanut, N.; Cadete Santos Aires, F.J.; Bennici, S.; Auroux, A. Hydrothermal synthesis and characterization of zirconia based catalysts. J. Solid State Chem. 2013, 203, 79–85. [Google Scholar] [CrossRef]
  89. Ueda, M.; Sasaki, Y.; Ikeda, M.; Ogawa, M.; Fujitani, W.; Nakano, T. Chemical-hydrothermal synthesis of bioinert ZrO2-TiO2 films on pure Ti substrates and proliferation of osteoblast-like cells. Mater. Trans. 2009, 50, 2147–2153. [Google Scholar] [CrossRef]
  90. Kim, D.; Kim, J.; Kim, K.; Cho, S.; Seo, M.; Kim, M.; Lee, J.; Kim, J.; Kim, K.; Cho, S.; et al. Preparations of titanium composite electrodes from commercial inorganic pigment and its application to light scattering layers on dye-sensitized solar cells. Mol. Cryst. Liq. Cryst. 2011, 539, 156–165. [Google Scholar] [CrossRef]
  91. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. [Google Scholar] [CrossRef] [PubMed]
  92. Kristianto, Y.; Taufik, A.; Saleh, R. Preparation and catalytic performance of ZrO2- nanographene platelets composites. J. Phys. Conf. Ser. 2016, 776, 012040. [Google Scholar] [CrossRef]
  93. Moazami, A.; Montazer, M. A novel multifunctional cotton fabric using ZrO2 NPs/urea/CTAB/MA/SHP: Introducing flame retardant, photoactive and antibacterial properties. J. Text. Inst. 2016, 107, 1253–1263. [Google Scholar] [CrossRef]
  94. Sudrajat, H.; Babel, S.; Sakai, H.; Takizawa, S. Rapid enhanced photocatalytic degradation of dyes using novel N-doped ZrO2. J. Environ. Manag. 2016, 165, 224–234. [Google Scholar] [CrossRef] [PubMed]
  95. Shakeel Ahmad, M.; Pandey, A.K.; Abd Rahim, N. Advancements in the development of TiO2 photoanodes and its fabrication methods for dye sensitized solar cell (DSSC) applications. A review. Renew. Sustain. Energy Rev. 2017, 77, 89–108. [Google Scholar] [CrossRef]
  96. Ren, X.; Qi, X.; Shen, Y.; Xiao, S.; Xu, G.; Zhang, Z.; Huang, Z.; Zhong, J. 2D co-catalytic MoS2 nanosheets embedded with 1D TiO2 nanoparticles for enhancing photocatalytic activity. J. Phys. D Appl. Phys. 2016, 49, 315304. [Google Scholar] [CrossRef]
  97. Zhang, J.; Huang, L.; Lu, Z.; Jin, Z.; Wang, X.; Xu, G.; Zhang, E.; Wang, H.; Kong, Z.; Xi, J.; et al. Crystal face regulating MoS2/TiO2 (001) heterostructure for high photocatalytic activity. J. Alloys Compd. 2016, 688, 840–848. [Google Scholar] [CrossRef]
  98. Liu, H.; Lv, T.; Zhu, C.; Su, X.; Zhu, Z. Efficient synthesis of MoS2 nanoparticles modified TiO2 nanobelts with enhanced visible-light-driven photocatalytic activity. J. Mol. Catal. A Chem. 2015, 396, 136–142. [Google Scholar] [CrossRef]
  99. Zhang, W.; Xiao, X.; Zheng, L.; Wan, C. Fabrication of TiO2/MoS2 composite photocatalyst and its photocatalytic mechanism for degradation of methyl orange under visible light. Can. J. Chem. Eng. 2015, 93, 1594–1602. [Google Scholar] [CrossRef]
  100. Yang, L.; Zheng, X.; Liu, M.; Luo, S.; Luo, Y.; Li, G. Fast photoelectro-reduction of Cr(VI) over MoS2/TiO2 nanotubes on Ti wire. J. Hazard. Mater. 2017, 329, 230–240. [Google Scholar] [CrossRef] [PubMed]
  101. Wang, J.; Wei, B.; Xu, L.; Gao, H.; Sun, W.; Che, J. Multilayered MoS2 coated TiO2 hollow spheres for efficient photodegradation of phenol under visible light irradiation. Mater. Lett. 2016, 179, 42–46. [Google Scholar] [CrossRef]
  102. Yang, H.G.; Zeng, H.C. Preparation of hollow anatase TiO2 nanospheres via Ostwald ripening. J. Phys. Chem. B. 2004, 108, 3492–3495. [Google Scholar] [CrossRef]
  103. Yan, C.; Rosei, F. Hollow micro/nanostructured materials prepared by ion exchange synthesis and their potential applications. New. J. Chem. 2014, 5, 1883–1904. [Google Scholar] [CrossRef]
  104. Bai, S.; Wang, L.; Chen, X.; Du, J.; Xiong, Y. Chemically exfoliated metallic MoS2 nanosheets: A promising supporting co-catalyst for enhancing the photocatalytic performance of TiO2 nanocrystals. Nano Res. 2015, 8, 175–183. [Google Scholar] [CrossRef]
  105. Yuan, Y.; Ye, Z.; Lu, H.; Hu, B.; Li, Y.; Chen, D.; Zhong, J.; Yu, Z.; Zou, Z. Constructing anatase TiO2 nanosheets with exposed (001) facets/layered MoS2 two-dimensional nanojunctions for enhanced solar hydrogen generation. ACS Catal. 2016, 6, 532–541. [Google Scholar] [CrossRef]
  106. Yasumitsu, T.; Sundo, K.; Moon Woo, C.; Masazumi, I. Tiocyanation of indoles. J. Heterocycl. Chem. 1978, 3, 425–427. [Google Scholar] [CrossRef]
  107. Mohammadi Ziarani, G.; Moradi, R.; Ahmadi, T.; Lashgari, N. Recent advances in the application of indoles in multicomponent reactions. RSC Adv. 2018, 8, 12069–12103. [Google Scholar] [CrossRef] [Green Version]
  108. Shi, X.; Fujitsuka, M.; Majima, T. Electron transfer dynamics of quaternary sulfur semiconductor/MoS2 layer-on-layer for efficient visible-light H2 evolution. Appl. Catal. B 2018, 235, 9–16. [Google Scholar] [CrossRef]
  109. Li, Z.; Meng, X.; Zhang, Z. Recent development on MoS2-based photocatalysis: A review. J. Photochem. Photobiol. C Photochem. Rev. 2018, 35, 39–55. [Google Scholar] [CrossRef]
  110. Kang, X.; Song, X.Z.; Han, Y.; Cao, J.; Tan, Z. Defect-engineered TiO2 hollow spiny nanocubes for phenol degradation under visible light irradiation. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef] [PubMed]
  111. Kermani, M.; Kakavandi, B.; Farzadkia, M.; Esrafili, A.; Jokandan, S.F.; Shahsavani, A. Catalytic ozonation of high concentrations of catechol over TiO2@Fe3O4 magnetic core-shell nanocatalyst: Optimization, toxicity and degradation pathway studies. J. Clean. Prod. 2018, 192, 597–607. [Google Scholar] [CrossRef]
  112. Norman, M.; Żółtowska-Aksamitowska, S.; Zgoła-Grześkowiak, A.; Ehrlich, H.; Jesionowski, T. Iron(III) phthalocyanine supported on a spongin scaffold as an advanced photocatalyst in a highly efficient removal process of halophenols and bisphenol A. J. Hazard. Mater. 2018, 347, 78–88. [Google Scholar] [CrossRef] [PubMed]
  113. Song, S.; Wang, J.; Peng, T.; Fu, W.; Zan, L. MoS2-MoO3-x hybrid cocatalyst for effectively enhanced H2 production photoactivity of AgIn5S8 nano-octahedrons. Appl. Catal. B 2018, 228, 39–46. [Google Scholar] [CrossRef]
  114. Rozenfeld, S.; Teller, H.; Schechter, M.; Farber, R.; Krichevski, O.; Schechter, A.; Cahan, R. Exfoliated molybdenum di-sulfide (MoS2) electrode for hydrogen production in microbial electrolysis cell. Bioelectrochemistry 2018, 123, 201–210. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, A.; Holt-hindle, P. Platinum-based nanostructured materials: Synthesis, properties, and applications. Chem. Rev. 2010, 110, 3767–3804. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, X.Y.; Li, H.P.; Cui, X.L.; Lin, Y. Graphene/TiO2 nanocomposites: Synthesis, characterization and application in hydrogen evolution from water photocatalytic splitting. J. Mater. Chem. 2010, 20, 2801–2806. [Google Scholar] [CrossRef]
  117. Klapiszewski, Ł.; Królak, M.; Jesionowski, T. Silica synthesis by the sol-gel method and its use in the preparation of multifunctional biocomposites. Cent. Eur. J. Chem. 2014, 12, 173–184. [Google Scholar] [CrossRef]
  118. Ciesielczyk, F.; Przybysz, M.; Zdarta, J.; Piasecki, A.; Paukszta, D.; Jesionowski, T. The sol-gel approach as a method of synthesis of xMgO·ySiO2 powder with defined physicochemical properties including crystalline structure. J. Sol-Gel Sci. Technol. 2014, 71, 501–513. [Google Scholar] [CrossRef]
  119. Baccile, N.; Fischer, A.; Julián-López, B.; Grosso, D.; Sanchez, C. Core-shell effects of functionalized oxide nanoparticles inside long-range meso-ordered spray-dried silica spheres. J. Sol-Gel Sci. Technol. 2008, 47, 119–123. [Google Scholar] [CrossRef] [Green Version]
  120. Faustini, M.; Grosso, D.; Boissière, C.; Backov, R.; Sanchez, C. “Integrative sol-gel chemistry”: A nanofoundry for materials science. J. Sol-Gel Sci. Technol. 2014, 70, 216–226. [Google Scholar] [CrossRef]
  121. Dislich, H.; Hinz, P. History and principles of the sol-gel process, and some new multicomponent oxide coatings. J. Non-Cryst. Solids 1982, 48, 11–16. [Google Scholar] [CrossRef]
  122. Carter, B.; Norton, G. Sols, Gels, and Organic Chemistry. In Ceramic Materials Science and Engineering, 2nd ed.; Springer: New York, NY, USA, 2007; pp. 400–411. ISBN 9780387462707. [Google Scholar]
  123. Ciesielczyk, F.; Szczekocka, W.; Siwińska-Stefańska, K.; Piasecki, A.; Paukszta, D.; Jesionowski, T. Evaluation of the photocatalytic ability of a sol-gel-derived MgO-ZrO2 oxide material. Open Chem. 2017, 15, 7–18. [Google Scholar] [CrossRef]
  124. Letailleur, A.A.; Ribot, F.; Boissière, C.; Teisseire, J.; Barthel, E.; Desmazières, B.; Chemin, N.; Sanchez, C. Sol-gel derived hybrid thin films: The chemistry behind processing. Chem. Mater. 2011, 23, 5082–5089. [Google Scholar] [CrossRef]
  125. Faustini, M.; Nicole, L.; Ruiz-Hitzky, E.; Sanchez, C. History of organic-inorganic hybrid materials: Prehistory, art, science, and advanced applications. Adv. Funct. Mater. 2018, 28, 1704158. [Google Scholar] [CrossRef]
  126. Gleiter, H. Nanocrystalline materials. Prog. Mater. Sci. 1989, 33, 223–315. [Google Scholar] [CrossRef]
  127. Livage, J.; Henry, M.; Sanchez, C. Sol-gel chemistry of transition metal oxides. Prog. Solid State Chem. 1988, 18, 259–341. [Google Scholar] [CrossRef]
  128. Schottner, G. Hybrid sol-gel-derived polymers: Applications of multifunctional materials. Chem. Mater. 2001, 13, 3422–3435. [Google Scholar] [CrossRef]
  129. Akpan, U.G.; Hameed, B.H. Parameters affecting the photocatalytic degradation of dyes using TiO2-based photocatalysts: A review. J. Hazard. Mater. 2009, 170, 520–529. [Google Scholar] [CrossRef] [PubMed]
  130. Miao, Z.; Xu, D.; Ouyang, J.; Guo, G.; Zhao, X.; Tang, Y. Electrochemically induced sol-gel preparation of single-crystalline TiO2 nanowires. Nano Lett. 2002, 2, 717–720. [Google Scholar] [CrossRef]
  131. Siwinska-Stefanska, K.; Zdarta, J.; Paukszta, D.; Jesionowski, T. The influence of addition of a catalyst and cheating agent on the properties of titanium dioxide synthesized via the sol-gel method. J. Sol-Gel Sci. Technol. 2015, 75, 264–278. [Google Scholar] [CrossRef]
  132. Ciesielczyk, F.; Bartczak, P.; Zdarta, J.; Jesionowski, T. Active MgO-SiO2 hybrid material for organic dye removal: A mechanism and interaction study of the adsorption of C.I. Acid Blue 29 and C.I. Basic Blue 9. J. Environ. Manag. 2017, 204, 123–135. [Google Scholar] [CrossRef] [PubMed]
  133. Fu, G.; Vary, P.S.; Lin, C.-T. Anatase TiO2 Nanocomposites for antimicrobial coatings. J. Phys. Chem. B 2005, 109, 8889–8898. [Google Scholar] [CrossRef] [PubMed]
  134. Nakashima, T.; Kimizuka, N. Interfacial synthesis of hollow TiO2 microspheres in ionic liquids. J. Am. Chem. Soc. 2003, 125, 6386–6387. [Google Scholar] [CrossRef] [PubMed]
  135. Liu, X.; Chu, P.K.; Ding, C. Surface modification of titanium, titanium alloys, and related materials for biomedical applications. Mater. Sci. Eng. R Rep. 2004, 47, 49–121. [Google Scholar] [CrossRef] [Green Version]
  136. Lettmann, C.; Hildenbrand, K.; Kisch, H.; Macyk, W.; Maier, W.F. Visible light photodegradation of 4-chlorophenol with a coke-containing titanium dioxide photocatalyst. Appl. Catal. B 2001, 32, 215–227. [Google Scholar] [CrossRef]
  137. Li, G.; Li, L.; Boerio-Goates, J.; Woodfield, B.F. High purity anatase TiO2 nanocrystals: Near room-temperature synthesis, grain growth kinetics, and surface hydration chemistry. J. Am. Chem. Soc. 2005, 127, 8659–8666. [Google Scholar] [CrossRef] [PubMed]
  138. Pierre, A.C. Introduction to Sol-Gel Processing, 1st ed.; Springer: New York, NY, USA, 1998; pp. 205–247. ISBN 9780792381211. [Google Scholar]
  139. Klein, L.; Aparicio, M.; Jitianu, A. Handbook of Sol-Gel Science and Technology, 2nd ed.; Springer: Cham, Switzerland, 2018; ISBN 9783319320991. [Google Scholar]
  140. Srikanth, B.; Goutham, R.; Badri Narayan, R.; Ramprasath, A.; Gopinath, K.P.; Sankaranarayanan, A.R. Recent advancements in supporting materials for immobilised photocatalytic applications in waste water treatment. J. Environ. Manag. 2017, 200, 60–78. [Google Scholar] [CrossRef] [PubMed]
  141. Long, J.; Zhang, B.; Li, X.; Zhan, X.; Xu, X.; Xie, Z.; Jin, Z. Effective production of resistant starch using pullulanase immobilized onto magnetic chitosan/Fe3O4 nanoparticles. Food Chem. 2018, 239, 276–286. [Google Scholar] [CrossRef] [PubMed]
  142. Cui, H.F.; Wu, W.W.; Li, M.M.; Song, X.; Lv, Y.; Zhang, T.T. A highly stable acetylcholinesterase biosensor based on chitosan-TiO2-graphene nanocomposites for detection of organophosphate pesticides. Biosens. Bioelectron. 2018, 99, 223–229. [Google Scholar] [CrossRef] [PubMed]
  143. Kołodziejczak-Radzimska, A.; Zdarta, J.; Jesionowski, T. Physicochemical and catalytic properties of acylase I from aspergillus melleus immobilized on amino- and carbonyl-grafted Stöber silica. Biotechnol. Prog. 2018, 34, 767–777. [Google Scholar] [CrossRef] [PubMed]
  144. Kolodziejczak-Radzimska, A. Functionalized Stöber silica as a support in immobilization process of lipase from Candida rugosa. Physicochem. Probl. Miner. Process. 2017, 53, 878–892. [Google Scholar] [CrossRef]
  145. Brinker, C.J.; Scherer, G.W. Particulate Sol and Gels. In Sol-Gel Science, the Physics and Chemistry of Sol-Gel Processing, 1st ed.; Academic Press, INC.: New York, NY, USA, 1990; pp. 235–297. ISBN 9780080571034. [Google Scholar]
  146. Nogami, M. Semiconductor-doped sol-gel optics. In Sol-Gel Optics: Processing and Applications, 1st ed.; Klein, L.C., Ed.; Springer: New York, NY, USA, 1994; pp. 329–344. [Google Scholar]
  147. Challagulla, S.; Tarafder, K.; Ganesan, R.; Roy, S. Structure sensitive photocatalytic reduction of nitroarenes over TiO2. Sci. Rep. 2017, 7, 8783. [Google Scholar] [CrossRef] [PubMed]
  148. Ye, Y.; Feng, Y.; Bruning, H.; Yntema, D.; Rijnaarts, H.H.M. Photocatalytic degradation of metoprolol by TiO2 nanotube arrays and UV-LED: Effects of catalyst properties, operational parameters, commonly present water constituents, and photo-induced reactive species. Appl. Catal. B Environ. 2018, 220, 171–181. [Google Scholar] [CrossRef]
  149. MiarAlipour, S.; Friedmann, D.; Scott, J.; Amal, R. TiO2/porous adsorbents: Recent advances and novel applications. J. Hazard. Mater. 2018, 341, 404–423. [Google Scholar] [CrossRef] [PubMed]
  150. Sopha, H.; Krbal, M.; Ng, S.; Prikryl, J.; Zazpe, R.; Yam, F.K.; Macak, J.M. Highly efficient photoelectrochemical and photocatalytic anodic TiO2 nanotube layers with additional TiO2 coating. Appl. Mater. Today 2017, 9, 104–110. [Google Scholar] [CrossRef]
  151. Kmentova, H.; Kment, S.; Wang, L.; Pausova, S.; Vaclavu, T.; Kuzel, R.; Han, H.; Hubicka, Z.; Zlamal, M.; Olejnicek, J.; et al. Photoelectrochemical and structural properties of TiO2 nanotubes and nanorods grown on FTO substrate: Comparative study between electrochemical anodization and hydrothermal method used for the nanostructures fabrication. Catal. Today 2017, 287, 130–136. [Google Scholar] [CrossRef]
  152. Shanmugam, M.; Bills, B.; Baroughi, M.F.; Galipeau, D. Electron transport in dye sensitized solar cells with TiO2/ZnO core- shell photoelectrode. In Proceedings of the 35th IEEE Photovoltaic Specialists Conference, Honolulu, HI, USA, 20–25 June 2016. [Google Scholar] [CrossRef]
  153. Chu, Y.; Wang, Q.; Cui, S. TiO2 and ZnO water sol preparation by sol-gel method and application on polyester fabric of antistatic finishing. Adv. Mater. Res. 2011, 331, 270–274. [Google Scholar] [CrossRef]
  154. Yu, D.; Wang, J.; Tian, J.; Xu, X.; Dai, J.; Wang, X. Preparation and characterization of TiO2/ZnO composite coating on carbon steel surface and its anticorrosive behavior in seawater. Compos. Part B 2013, 46, 135–144. [Google Scholar] [CrossRef]
  155. Wang, L.; Fu, X.; Han, Y.; Chang, E.; Wu, H.; Wang, H.; Li, K.; Qi, X.; Wang, L.; Fu, X.; et al. Preparation, characterization, and photocatalytic activity of TiO2/ZnO nanocomposites. J. Nanomater. 2013, 2013, 321459. [Google Scholar] [CrossRef]
  156. Giannakopoulou, T.; Todorova, N.; Giannouri, M.; Yu, J.; Trapalis, C. Optical and photocatalytic properties of composite TiO2/ZnO thin films. Catal. Today 2014, 230, 174–180. [Google Scholar] [CrossRef]
  157. Chen, Y.; Zhang, C.; Huang, W.; Yang, C.; Huang, T.; Situ, Y.; Huang, H. Synthesis of porous ZnO/TiO2 thin films with superhydrophilicity and photocatalytic activity via a template-free sol-gel method. Surf. Coat. Technol. 2014, 258, 531–538. [Google Scholar] [CrossRef]
  158. Pournuroz, Z.; Khosravi, M.; Giahi, M.; Sohrabi, M. Compare removal of Reactive Red 195 and Blue 19 by nano TiO2-ZnO. Orient. J. Chem. 2015, 31, 2–6. [Google Scholar] [CrossRef]
  159. Al-Mayman, S.I.; Al-Johani, M.S.; Mokhtar, M.M.; Al-Zeghayer, Y.S.; Ramay, S.M.; Al-Awadi, A.S.; Soliman, M.A. TiO2-ZnO photocatalysts synthesized by sol-gel auto-ignition technique for hydrogen production. Int. J. Hydrogen Energy 2016, 42, 5016–5025. [Google Scholar] [CrossRef]
  160. Zulkiflee, N.S.; Hussin, R. Effect of temperature on TiO2/ZnO nanostructure thin films. Mater. Sci. Forum 2016, 840, 262–266. [Google Scholar] [CrossRef]
  161. Prasannalakshmi, P.; Shanmugam, N. Fabrication of TiO2/ZnO nanocomposites for solar energy driven photocatalysis. Mater. Sci. Semicond. Process. 2017, 61, 114–124. [Google Scholar] [CrossRef]
  162. Al-Hazami, F.E.; Yakuphanoglu, F. Photoconducting and photovoltaic properties of ZnO:TiO2 composite/p-silicon heterojunction photodiode. Silicon 2017, 10, 781–787. [Google Scholar] [CrossRef]
  163. Armin, S.; Estekhraji, Z.; Amiri, S. Synthesis and characterization of anti-fungus, anti-corrosion and self-cleaning hybrid nanocomposite coatings based on sol-gel process. J. Inorg. Organomet. Polym. Mater. 2017, 27, 883–891. [Google Scholar] [CrossRef]
  164. Chamanzadeh, Z.; Noormohammadi, M.; Zahedifar, M. Enhanced photovoltaic performance of dye sensitized solar cell using TiO2 and ZnO nanoparticles on top of free standing TiO2 nanotube arrays. Mater. Sci. Semicond. Process. 2017, 61, 107–113. [Google Scholar] [CrossRef]
  165. Fatimah, I. Preparation of TiO2-ZnO and its activity test in sonophotocatalytic degradation of phenol. IOP Conf. Ser. Mater. Sci. Eng. 2016, 107, 012003. [Google Scholar] [CrossRef]
  166. Bozzi, A.; Yuranova, T.; Guasaquillo, I.; Laub, D.; Kiwi, J. Self-cleaning of modified cotton textiles by TiO2 at low temperatures under daylight irradiation. J. Photochem. Photobiol. A Chem. 2005, 174, 156–164. [Google Scholar] [CrossRef]
  167. Bozzi, A.; Yuranova, T.; Kiwi, J. Self-cleaning of wool-polyamide and polyester textiles by TiO2-rutile modification under daylight irradiation at ambient temperature. J. Photochem. Photobiol. A Chem. 2005, 172, 27–34. [Google Scholar] [CrossRef]
  168. Mihailović, D.; Šaponjić, Z.; Radoičić, M.; Radetić, T.; Jovančić, P.; Nedeljković, J.; Radetić, M. Functionalization of polyester fabrics with alginates and TiO2 nanoparticles. Carbohydr. Polym. 2010, 79, 526–532. [Google Scholar] [CrossRef]
  169. Rashvand, M.; Ranjbar, Z.; Rastegar, S. Preserving anti-corrosion properties of epoxy based coatings imultaneously exposed to humidity and UV-radiation using nano zinc oxide. J. Electrochem. Soc. 2012, 159, 129–132. [Google Scholar] [CrossRef]
  170. Li, Q.; Yang, X.; Zhang, L.; Wang, J.; Chen, B. Corrosion resistance and mechanical properties of pulse electrodeposited Ni-TiO2 composite coating for sintered NdFeB magnet. J. Alloys Compd. 2009, 482, 339–344. [Google Scholar] [CrossRef]
  171. Yu, Z.; Di, H.; Ma, Y.; He, Y.; Liang, L.; Lv, L.; Ran, X.; Pan, Y.; Luo, Z. Preparation of graphene oxide modified by titanium dioxide to enhance the anti-corrosion performance of epoxy coatings. Surf. Coat. Technol. 2015, 276, 471–478. [Google Scholar] [CrossRef]
  172. You, X.; Chen, F.; Zhang, J. Effects of calcination on the physical and photocatalytic properties of TiO2 powders prepared by sol-gel template method. J. Sol-Gel Sci. Technol. 2005, 34, 181–187. [Google Scholar] [CrossRef]
  173. Chen, Y.; Dionysiou, D.D. Effect of calcination temperature on the photocatalytic activity and adhesion of TiO2 films prepared by the P-25 powder-modified sol-gel method. J. Mol. Catal. A Chem. 2006, 244, 73–82. [Google Scholar] [CrossRef]
  174. Wu, N.L.; Lee, M.S.; Pon, Z.J.; Hsu, J.Z. Effect of calcination atmosphere on TiO2 photocatalysis in hydrogen production from methanol/water solution. J. Photochem. Photobiol. A Chem. 2004, 163, 277–280. [Google Scholar] [CrossRef]
  175. Dalton, J.S.; Janes, P.A.; Jones, N.G.; Nicholson, J.A.; Hallam, K.R.; Allen, G.C. Photocatalytic oxidation of NOx gases using TiO2: A surface spectroscopic approach. Environ. Pollut. 2002, 120, 415–422. [Google Scholar] [CrossRef]
  176. Karapati, S.; Giannakopoulou, T.; Todorova, N.; Boukos, N.; Antiohos, S.; Papageorgiou, D.; Chaniotakis, E.; Dimotikali, D.; Trapalis, C. TiO2 functionalization for efficient NOx removal in photoactive cement. Appl. Surf. Sci. 2014, 319, 29–36. [Google Scholar] [CrossRef]
  177. Todorova, N.; Giannakopoulou, T.; Karapati, S.; Petridis, D.; Vaimakis, T.; Trapalis, C. Composite TiO2/clays materials for photocatalytic NOx oxidation. Appl. Surf. Sci. 2014, 319, 113–120. [Google Scholar] [CrossRef]
  178. Thompson, S.; Shirtcliffe, N.J.; O’Keefe, E.S.; Appleton, S.; Perry, C.C. Synthesis of SrCoxTixFe(12-2x)O19 through sol-gel auto-ignition and its characterisation. J. Magn. Magn. Mater. 2005, 292, 100–107. [Google Scholar] [CrossRef]
  179. Sutka, A.; Mezinskis, G. Sol-gel auto-combustion synthesis of spinel-type ferrite nanomaterials. Front. Mater. Sci. 2012, 6, 128–141. [Google Scholar] [CrossRef]
  180. Hendi, A.A.; Yakuphanoglu, F. Graphene doped TiO2/p-silicon heterojunction photodiode. J. Alloys Compd. 2016, 665, 418–427. [Google Scholar] [CrossRef]
  181. Yu, H.; Li, X.; Quan, X.; Chen, S.; Zhang, Y. Effective utilization of visible light (including λ > 600 nm) in phenol degradation with p-silicon nanowire/TiO2 core/shell heterojunction array cathode. Environ. Sci. Technol. 2009, 43, 7849–7855. [Google Scholar] [CrossRef] [PubMed]
  182. Kang, H.; Park, J.; Choi, T.; Jung, H.; Lee, K.H.; Im, S.; Kim, H. N-ZnO:N/p-Si nanowire photodiode prepared by atomic layer deposition. Appl. Phys. Lett. 2012, 100, 2010–2014. [Google Scholar] [CrossRef]
  183. Silva, C.G.; Faria, J.L. Photocatalytic oxidation of benzene derivatives in aqueous suspensions: Synergic effect induced by the introduction of carbon nanotubes in a TiO2 matrix. Appl. Catal. B Environ. 2010, 101, 81–89. [Google Scholar] [CrossRef]
  184. Liu, G.; Zhao, Y.; Sun, C.; Li, F.; Lu, G.Q.; Cheng, H.M. Synergistic effects of B/N doping on the visible-light photocatalytic activity of mesoporous TiO2. Angew. Chem. Int. Ed. 2008, 47, 4516–4520. [Google Scholar] [CrossRef] [PubMed]
  185. Hoang, S.; Berglund, S.P.; Hahn, N.T.; Bard, A.J.; Mullins, C.B. Enhancing visible light photo-oxidation of water with TiO2 nanowire arrays via cotreatment with H2 and NH3: Synergistic effects between Ti3+ and N. J. Am. Chem. Soc. 2012, 134, 3659–3662. [Google Scholar] [CrossRef] [PubMed]
  186. Zhang, K.; Zhang, F.J.; Chen, M.L.; Oh, W.C. Comparison of catalytic activities for photocatalytic and sonocatalytic degradation of methylene blue in present of anatase TiO2-CNT catalysts. Ultrason. Sonochem. 2011, 18, 765–772. [Google Scholar] [CrossRef] [PubMed]
  187. Wang, J.; Guo, B.; Zhang, X.; Zhang, Z.; Han, J.; Wu, J. Sonocatalytic degradation of methyl orange in the presence of TiO2 catalysts and catalytic activity comparison of rutile and anatase. Ultrason. Sonochem. 2005, 12, 331–337. [Google Scholar] [CrossRef] [PubMed]
  188. Hernández-Alonso, M.D.; Tejedor-Tejedor, I.; Coronado, J.M.; Soria, J.; Anderson, M.A. Sol-gel preparation of TiO2-ZrO2 thin films supported on glass rings: Influence of phase composition on photocatalytic activity. Thin Solid Films 2006, 502, 125–131. [Google Scholar] [CrossRef]
  189. Kraleva, E.; Saladino, M.L.; Matassa, R.; Caponetti, E.; Enzo, S.; Spojakina, A. Phase formation in mixed TiO2-ZrO2 oxides prepared by sol-gel method. J. Struct. Chem. 2011, 52, 330–339. [Google Scholar] [CrossRef]
  190. Mohammadi, M.R.; Fray, D.J. Synthesis and characterisation of nanosized TiO2-ZrO2 binary system prepared by an aqueous sol-gel process: Physical and sensing properties. Sens. Actuators B Chem. 2011, 155, 568–576. [Google Scholar] [CrossRef]
  191. Naumenko, A.; Gnatiuk, I.; Smirnova, N.; Eremenko, A. Characterization of sol-gel derived TiO2/ZrO2 films and powders by Raman spectroscopy. Thin Solid Films 2012, 520, 4541–4546. [Google Scholar] [CrossRef]
  192. Karthika, S.; Prathibha, V.; Ann, M.K.A.; Viji, V.; Biju, P.R.; Unnikrishnan, N.V. Structural and spectroscopic studies of Sm3+/CdS nanocrystallites in sol-gel TiO2-ZrO2 matrix. J. Electron. Mater. 2014, 43, 447–451. [Google Scholar] [CrossRef]
  193. Fukumoto, T.; Yoshioka, T.; Nagasawa, H.; Kanezashi, M.; Tsuru, T. Development and gas permeation properties of microporous amorphous TiO2-ZrO2-organic composite membranes using chelating ligands. J. Membr. Sci. 2014, 461, 96–105. [Google Scholar] [CrossRef]
  194. Atanda, L.; Silahua, A.; Mukundan, S.; Shrotri, A.; Torres-Torres, G.; Beltramini, J. Catalytic behaviour of TiO2-ZrO2 binary oxide synthesized by sol-gel process for glucose conversion to 5-hydroxymethylfurfural. RSC Adv. 2015, 5, 80346–80352. [Google Scholar] [CrossRef]
  195. Khan, S.; Kim, J.; Sotto, A.; Van der Bruggen, B. Humic acid fouling in a submerged photocatalytic membrane reactor with binary TiO2-ZrO2 particles. J. Ind. Eng. Chem. 2015, 21, 779–786. [Google Scholar] [CrossRef]
  196. Zukalová, M.; Zukal, A.; Kavan, L.; Nazeeruddin, M.K.; Liska, P.; Grätzel, M. Organized mesoporous TiO2 films exhibiting greatly enhanced performance in dye-sensitized solar cells. Nano Lett. 2005, 5, 1789–1792. [Google Scholar] [CrossRef] [PubMed]
  197. Marien, C.B.D.; Marchal, C.; Koch, A.; Robert, D.; Drogui, P. Sol-gel synthesis of TiO2 nanoparticles: Effect of Pluronic P123 on particle’s morphology and photocatalytic degradation of paraquat. Environ. Sci. Pollut. Res. 2017, 24, 12582–12588. [Google Scholar] [CrossRef] [PubMed]
  198. Choi, S.Y.; Mamak, M.; Coombs, N.; Chopra, N.; Ozin, G.A. Thermally stable two-dimensional hexagonal mesoporous nanocrystalline anatase, meso-nc-TiO2: Bulk and crack-free thin film morphologies. Adv. Funct. Mater. 2004, 14, 335–344. [Google Scholar] [CrossRef]
  199. Vatanpour, V.; Madaeni, S.S.; Khataee, A.R.; Salehi, E.; Zinadini, S.; Monfared, H.A. TiO2 embedded mixed matrix PES nanocomposite membranes: Influence of different sizes and types of nanoparticles on antifouling and performance. Desalination 2012, 292, 19–29. [Google Scholar] [CrossRef]
  200. Ramachandra, M.; Abhishek, A.; Siddeshwar, P.; Bharathi, V. Hardness and wear resistance of ZrO2 nano particle reinforced Al nanocomposites produced by powder metallurgy. Procedia Mater. Sci. 2015, 10, 212–219. [Google Scholar] [CrossRef]
  201. Shojai, F.; Mäntylä, T.A. Structural stability of yttria doped zirconia membranes in acid and basic aqueous solutions. J. Eur. Ceram. Soc. 2001, 21, 37–44. [Google Scholar] [CrossRef]
  202. Puhlfürß, P.; Voigt, A.; Weber, R.; Morbé, M. Microporous TiO2 membranes with a cut off <500 Da. J. Membr. Sci. 2000, 174, 123–133. [Google Scholar] [CrossRef]
  203. Fan, J.; Ohya, H.; Suga, T.; Ohashi, H.; Yamashita, K.; Tsuchiya, S.; Aihara, M.; Takeuchi, T.; Negishi, Y. High flux zirconia composite membrane for hydrogen separation at elevated temperature. J. Membr. Sci. 2000, 170, 113–125. [Google Scholar] [CrossRef]
  204. Lu, C.; Chen, Z. High-temperature resistive hydrogen sensor based on thin nanoporous rutile TiO2 film on anodic aluminum oxide. Sens. Actuators B Chem. 2009, 140, 109–115. [Google Scholar] [CrossRef]
  205. Mather, G.C.; Marques, F.M.B.; Frade, J.R. Detection mechanism of TiO2 -based ceramic H2 sensors. J. Eur. Ceram. Soc. 1999, 19, 887–891. [Google Scholar] [CrossRef]
  206. Tang, H.; Prasad, K.; Sanjines, R.; Levy, F. TiO2 anatase thin-films as gas sensors. Sens. Actuators B-Chem. 1995, 26, 71–75. [Google Scholar] [CrossRef]
  207. Kuster, B.F.M. 5-Hydroxymethylfurfural (HMF). A review focussing on its manufacture. Starch-Stärke 1990, 42, 314–321. [Google Scholar] [CrossRef]
  208. Rosatella, A.A.; Simeonov, S.P.; Frade, R.F.M.; Afonso, C.A.M. 5-Hydroxymethylfurfural (HMF) as a building block platform: Biological properties, synthesis and synthetic applications. Green Chem. 2011, 13, 754–793. [Google Scholar] [CrossRef]
  209. Ho, D.P.; Vigneswaran, S.; Ngo, H.H. Photocatalysis-membrane hybrid system for organic removal from biologically treated sewage effluent. Sep. Purif. Technol. 2009, 68, 145–152. [Google Scholar] [CrossRef]
  210. Yang, N.; Wen, X.; Waite, T.D.; Wang, X.; Huang, X. Natural organic matter fouling of microfiltration membranes: Prediction of constant flux behavior from constant pressure materials properties determination. J. Membr. Sci. 2011, 366, 192–202. [Google Scholar] [CrossRef]
  211. Katrib, A.; Benadda, A.; Sobczak, J.W.; Maire, G. XPS and catalytic properties of the bifunctional supported MoO2(Hx)ac on TiO2 for the hydroisomerization reactions of hexanes and 1-hexene. Appl. Catal. A Gen. 2003, 242, 31–40. [Google Scholar] [CrossRef]
  212. Fu, J.; Ji, M.; Wang, Z.; Jin, L.; An, D. A new submerged membrane photocatalysis reactor (SMPR) for fulvic acid removal using a nano-structured photocatalyst. J. Hazard. Mater. 2006, 131, 238–242. [Google Scholar] [CrossRef] [PubMed]
  213. Zeleny, J. Instability of electrified liquid surfaces. Phys. Rev. 1917, 10, 1–6. [Google Scholar] [CrossRef]
  214. Dole, M.; Hines, R.L.; Mack, L.L.; Mobley, R.C.; Ferguson, L.D.; Alice, M.B. Gas phase macroions. Macromolecules 1968, 1, 96–97. [Google Scholar] [CrossRef]
  215. Ramaseshan, R.; Sundarrajan, S.; Jose, R.; Ramakrishna, S. Nanostructured ceramics by electrospinning. J. Appl. Phys. 2007, 102, 111101. [Google Scholar] [CrossRef]
  216. Dai, Y.; Cobley, C.M.; Zeng, J.; Sun, Y.; Xia, Y. Synthesis of anatase TiO2 nanocrystals with exposed (001) facets. Nano Lett. 2009, 9, 1–5. [Google Scholar] [CrossRef] [PubMed]
  217. Formo, E.; Lee, E.; Campbell, D.; Xia, Y. Functionalization of electrospun TiO2 nanofibers with Pt nanoparticles and nanowires for catalytic applications. Nano Lett. 2008, 8, 668–672. [Google Scholar] [CrossRef] [PubMed]
  218. Inagaki, M.; Yang, Y.; Kang, F. Carbon nanofibers prepared via electrospinning. Adv. Mater. 2012, 24, 2547–2566. [Google Scholar] [CrossRef] [PubMed]
  219. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef] [PubMed]
  220. Liu, Z.; Sun, D.D.; Guo, P.; Leckie, J.O. An efficient bicomponent TiO2/SnO2 nanofiber photocatalyst fabricated by electrospinning with a side-by-side dual spinneret method. Nano Lett. 2007, 7, 1081–1085. [Google Scholar] [CrossRef] [PubMed]
  221. Lu, X.; Wang, C.; Wei, Y. One-dimensional composite nanomaterials: Synthesis by electrospinning and their applications. Small 2009, 5, 2349–2370. [Google Scholar] [CrossRef] [PubMed]
  222. Thavasi, V.; Singh, G.; Ramakrishna, S. Electrospun nanofibers in energy and environmental applications. Energy Environ. Sci. 2008, 1, 205–221. [Google Scholar] [CrossRef]
  223. Anselme, K.; Davidson, P.; Popa, A.M.; Giazzon, M.; Liley, M.; Ploux, L. The interaction of cells and bacteria with surfaces structured at the nanometre scale. Acta Biomater. 2010, 6, 3824–3846. [Google Scholar] [CrossRef] [PubMed]
  224. Ding, B.; Wang, M.; Wang, X.; Yu, J.; Sun, G. Electrospun nanomaterials for ultrasensitive sensors. Mater. Today 2010, 13, 16–27. [Google Scholar] [CrossRef]
  225. Agarwal, S.; Greiner, A.; Wendorff, J.H. Functional materials by electrospinning of polymers. Prog. Polym. Sci. 2013, 38, 963–991. [Google Scholar] [CrossRef]
  226. Long, Y.Z.; Li, M.M.; Gu, C.; Wan, M.; Duvail, J.L.; Liu, Z.; Fan, Z. Recent advances in synthesis, physical properties and applications of conducting polymer nanotubes and nanofibers. Prog. Polym. Sci. 2011, 36, 1415–1442. [Google Scholar] [CrossRef]
  227. Huang, Z.M.; Zhang, Y.Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  228. Spasova, M.; Mincheva, R.; Paneva, D.; Manolova, N.; Rashkov, I. Perspectives on: Criteria for complex evaluation of the morphology and alignment of electrospun polymer nanofibers. J. Bioact. Compat. Polym. 2006, 21, 465–479. [Google Scholar] [CrossRef]
  229. Reneker, D.H.; Chun, I. Nanometre diameter fibres of polymer, produced by electrospinning. Nanotechnology 1996, 7, 216–223. [Google Scholar] [CrossRef]
  230. Fong, H.; Liu, W.; Wang, C.S.; Vaia, R.A. Generation of electrospun fibers of nylon 6 and nylon 6-montmorillonite nanocomposite. Polymer 2001, 43, 775–780. [Google Scholar] [CrossRef]
  231. Mitarai, T.; Shander, A.; Tight, M.; Fabrics, N.; Martin, F.N.; English, J.T. An introduction to electrospinning and nanofibres. J. Cardiothorac. Vasc. Anesth. 2013, 27, 1–2. [Google Scholar] [CrossRef]
  232. Wang, Y.; Jia, W.; Strout, T.; Schempf, A.; Zhang, H.; Li, B.; Cui, J.; Lei, Y. Ammonia gas sensor using polypyrrole-coated TiO2/ZnO nanofibers. Electroanalysis 2009, 21, 1432–1438. [Google Scholar] [CrossRef]
  233. Park, J.Y.; Choi, S.W.; Lee, J.W.; Lee, C.; Kim, S.S. Synthesis and gas sensing properties of TiO2-ZnO core-shell nanofibers. J. Am. Ceram. Soc. 2009, 92, 2551–2554. [Google Scholar] [CrossRef]
  234. Wang, H.Y.; Yang, Y.; Li, X.; Li, L.J.; Wang, C. Preparation and characterization of porous TiO2/ZnO composite nanofibers via electrospinning. Chin. Chem. Lett. 2010, 21, 1119–1123. [Google Scholar] [CrossRef]
  235. Liu, R.; Ye, H.; Xiong, X.; Liu, H. Fabrication of TiO2/ZnO composite nanofibers by electrospinning and their photocatalytic property. Mater. Chem. Phys. 2010, 121, 432–439. [Google Scholar] [CrossRef]
  236. Kanjwal, M.A.; Barakat, N.A.M.; Sheikh, F.A.; Park, S.J.; Kim, H.Y. Photocatalytic activity of ZnO-TiO2 hierarchical nanostructure prepared by combined electrospinning and hydrothermal techniques. Macromol. Res. 2010, 18, 233–240. [Google Scholar] [CrossRef]
  237. Li, J.; Yan, L.; Wang, Y.; Kang, Y.; Wang, C.; Yang, S. Fabrication of TiO2/ZnO composite nanofibers with enhanced photocatalytic activity. J. Mater. Sci. Mater. Electron. 2016, 27, 7834–7838. [Google Scholar] [CrossRef]
  238. Araújo, E.S.; Libardi, J.; Faia, P.M.; de Oliveira, H.P. Humidity-sensing properties of hierarchical TiO2:ZnO composite grown on electrospun fibers. J. Mater. Sci. Mater. Electron. 2017, 28, 1–9. [Google Scholar] [CrossRef]
  239. Karunagaran, B.; Uthirakumar, P.; Chung, S.J.; Velumani, S.; Suh, E.K. TiO2 thin film gas sensor for monitoring ammonia. Mater. Charact. 2007, 58, 680–684. [Google Scholar] [CrossRef]
  240. Shavisi, Y.; Sharifnia, S.; Hosseini, S.N.; Khadivi, M.A. Application of TiO2/perlite photocatalysis for degradation of ammonia in wastewater. J. Ind. Eng. Chem. 2014, 20, 278–283. [Google Scholar] [CrossRef]
  241. Saha, D.; Deng, S. Characteristics of ammonia adsorption on activated alumina. J. Chem. Eng. Data 2010, 55, 5587–5593. [Google Scholar] [CrossRef]
  242. Yuzawa, H.; Mori, T.; Itoh, H.; Yoshida, H. Reaction mechanism of ammonia decomposition to nitrogen and hydrogen over metal loaded titanium oxide photocatalyst. J. Phys. Chem. C 2012, 116, 4126–4136. [Google Scholar] [CrossRef]
  243. Pang, Z.; Yang, Z.; Chen, Y.; Zhang, J.; Wang, Q.; Huang, F.; Wei, Q. A room temperature ammonia gas sensor based on cellulose/TiO2/PANI composite nanofibers. Colloids Surf. A Physicochem. Eng. Asp. 2016, 494, 248–255. [Google Scholar] [CrossRef]
  244. Nagaraja, R.; Kottam, N.; Girija, C.R.; Nagabhushana, B.M. Photocatalytic degradation of Rhodamine B dye under UV/solar light using ZnO nanopowder synthesized by solution combustion route. Powder Technol. 2012, 215–216, 91–97. [Google Scholar] [CrossRef]
  245. Kornbrust, D.; Barfknecht, T. Testing of 24 food, drug, cosmetic, and fabric dyes in the in vitro and the in vivo/in vitro rat hepatocyte primary culture/DNA repair assays. Environ. Mutagen 1985, 7, 101–120. [Google Scholar] [CrossRef] [PubMed]
  246. Snawder, J.E.; Lipscomb, J.C. Interindividual variance of cytochrome P450 forms in human hepatic microsomes: Correlation of individual forms with xenobiotic metabolism and implications in risk assessment. Regul. Toxicol. Pharm. 2000, 32, 200–209. [Google Scholar] [CrossRef] [PubMed]
  247. Mirsalis, J.C.; Tyson, C.K.; Steinmetz, K.L.; Loh, E.K.; Hamilton, C.M.; Bakke, J.P.; Spalding, J.W. Measurement of unscheduled DNA synthesis and S-phase synthesis in rodent hepatocytes following in vivo treatment: Testing of 24 compounds. Environ. Mol. Mutagen 1989, 14, 155–164. [Google Scholar] [CrossRef] [PubMed]
  248. Ding, L.; Zou, B.; Gao, W.; Liu, Q.; Wang, Z.; Guo, Y.; Wang, X.; Liu, Y. Adsorption of Rhodamine-B from aqueous solution using treated rice husk-based activated carbon. Colloids Surf. A Physicochem. Eng. Asp. 2014, 446, 1–7. [Google Scholar] [CrossRef]
  249. Jain, R.; Mathur, M.; Sikarwar, S.; Mittal, A. Removal of the hazardous dye rhodamine B through photocatalytic and adsorption treatments. J. Environ. Manag. 2007, 85, 956–964. [Google Scholar] [CrossRef] [PubMed]
  250. Liang, Z.; Li, Q.; Li, F.; Zhao, S. Microstructure of SiO2/TiO2 hybrid electrospun nanofibers and their application in dye degradation. Res. Chem. Intermed. 2016, 42, 7017–7029. [Google Scholar] [CrossRef]
  251. Hou, C.; Jiao, T.; Xing, R.; Chen, Y.; Zhou, J.; Zhang, L. Preparation of TiO2 nanoparticles modified electrospun nanocomposite membranes toward efficient dye degradation for wastewater treatment. J. Taiwan Inst. Chem. Eng. 2017, 78, 118–126. [Google Scholar] [CrossRef]
  252. Doh, S.J.; Kim, C.; Lee, S.G.; Lee, S.J.; Kim, H. Development of photocatalytic TiO2 nanofibers by electrospinning and its application to degradation of dye pollutants. J. Hazard. Mater. 2008, 154, 118–127. [Google Scholar] [CrossRef] [PubMed]
  253. Faia, P.M.; Furtado, C.S. Effect of composition on electrical response to humidity of TiO2:ZnO sensors investigated by impedance spectroscopy. Sens. Actuators B Chem. 2013, 181, 720–729. [Google Scholar] [CrossRef]
  254. Su, M.; Wang, J.; Du, H.; Yao, P.; Zheng, Y.; Li, X. Characterization and humidity sensitivity of electrospun ZrO2:TiO2 hetero-nanofibers with double jets. Sens. Actuators B Chem. 2012, 161, 1038–1045. [Google Scholar] [CrossRef]
  255. Lee, C.; Park, J.H.; Jeon, Y.; Park, J.I.; Einaga, H.; Truong, Y.B.; Kyratzis, I.L.; Mochida, I.; Choi, J.; Shul, Y.G. Phosphate modified TiO2/ZrO2 nanofibrous web composite membrane for enhanced performance and durability of high temperature PEM fuel cells. Energy Fuels 2017, 31, 4–11. [Google Scholar] [CrossRef]
  256. Boaretti, C.; Pasquini, L.; Sood, R.; Giancola, S.; Donnadio, A.; Roso, M.; Modesti, M.; Cavaliere, S. Mechanically stable nanofibrous sPEEK/Aquivion® composite membranes for fuel cell applications. J. Membr. Sci. 2018, 545, 66–74. [Google Scholar] [CrossRef]
  257. Xiao, P.; Li, J.; Tang, H.; Wang, Z.; Pan, M. Physically stable and high performance Aquivion/ePTFE composite membrane for high temperature fuel cell application. J. Membr. Sci. 2013, 442, 65–71. [Google Scholar] [CrossRef]
  258. Marrony, M.; Beretta, D.; Ginocchio, S.; Nedellec, Y.; Subianto, S.; Jones, D.J. Lifetime prediction approach applied to the aquivionTM short side chain perfluorosulfonic acid ionomer membrane for intermediate temperature proton exchange membrane fuel cell application. Fuel Cells 2013, 13, 1146–1154. [Google Scholar] [CrossRef]
Figure 1. Process parameters determining the physico-chemical and structural properties of products obtained by the hydrothermal method.
Figure 1. Process parameters determining the physico-chemical and structural properties of products obtained by the hydrothermal method.
Materials 11 02295 g001
Figure 2. (a,b) SEM; (c) high-resolution transmission electron microscopy (HRTEM) and (d) selected area electron diffraction (SAED) images of the TiO2-ZnO system (created based on [61] with permission from Elsevier Publisher).
Figure 2. (a,b) SEM; (c) high-resolution transmission electron microscopy (HRTEM) and (d) selected area electron diffraction (SAED) images of the TiO2-ZnO system (created based on [61] with permission from Elsevier Publisher).
Materials 11 02295 g002
Figure 3. SEM images of a TiO2-ZnO oxide system subjected to calcination at temperatures of (a) 500 °C and (b) 600 °C (created based on [65] with permission from Elsevier Publisher).
Figure 3. SEM images of a TiO2-ZnO oxide system subjected to calcination at temperatures of (a) 500 °C and (b) 600 °C (created based on [65] with permission from Elsevier Publisher).
Materials 11 02295 g003
Figure 4. TEM images of (a) TiO2; (b) TiO2-ZrO2; HRTEM images of (c) TiO2 and (d) TiO2-ZrO2 (created based on [82] with permission from Elsevier Publisher).
Figure 4. TEM images of (a) TiO2; (b) TiO2-ZrO2; HRTEM images of (c) TiO2 and (d) TiO2-ZrO2 (created based on [82] with permission from Elsevier Publisher).
Materials 11 02295 g004
Figure 5. TEM images of the oxide system 90%TiO2-10%ZrO2 (created based on [85] with permission from Elsevier Publisher).
Figure 5. TEM images of the oxide system 90%TiO2-10%ZrO2 (created based on [85] with permission from Elsevier Publisher).
Materials 11 02295 g005
Figure 6. SEM images of TiO2-MoS2 hybrid systems with MoS2 contents of (a,b) 20%; (c,d) 40%; (e,f) 60%; and (g,h) 80% (created based on [98] with permission from Elsevier Publisher).
Figure 6. SEM images of TiO2-MoS2 hybrid systems with MoS2 contents of (a,b) 20%; (c,d) 40%; (e,f) 60%; and (g,h) 80% (created based on [98] with permission from Elsevier Publisher).
Materials 11 02295 g006
Figure 7. SEM images of (a) TiO2; (b) MoS2; (c) TM3; (d) TM6; (e) TM9 and (f) TM12 (created based on [101] with permission from Elsevier Publisher).
Figure 7. SEM images of (a) TiO2; (b) MoS2; (c) TM3; (d) TM6; (e) TM9 and (f) TM12 (created based on [101] with permission from Elsevier Publisher).
Materials 11 02295 g007
Figure 8. A block diagram of the sol-gel process (created based on [122] with permission from Springer).
Figure 8. A block diagram of the sol-gel process (created based on [122] with permission from Springer).
Materials 11 02295 g008
Figure 9. AFM images of the oxide system (90)TiO2-(10)ZnO (created based on [157] with permission from Elsevier Publisher).
Figure 9. AFM images of the oxide system (90)TiO2-(10)ZnO (created based on [157] with permission from Elsevier Publisher).
Materials 11 02295 g009
Figure 10. HRTEM images of the oxide system TiZr6 calcined at 700 °C (created based on [189] with permission from Springer).
Figure 10. HRTEM images of the oxide system TiZr6 calcined at 700 °C (created based on [189] with permission from Springer).
Materials 11 02295 g010
Figure 11. SEM images of surface and cross-section views of a TiO2–ZrO2–ISOH membrane calcined at 350 °C under N2: (a) surface and cross-section views at 20,000×; (b) cross-section view at 200,000× (created based on [193] with permission from Elsevier Publisher).
Figure 11. SEM images of surface and cross-section views of a TiO2–ZrO2–ISOH membrane calcined at 350 °C under N2: (a) surface and cross-section views at 20,000×; (b) cross-section view at 200,000× (created based on [193] with permission from Elsevier Publisher).
Materials 11 02295 g011
Figure 12. Schematic diagrams of (a) vertical and (b) horizontal electrospinning processes (created based on [219] with permission from Elsevier Publisher).
Figure 12. Schematic diagrams of (a) vertical and (b) horizontal electrospinning processes (created based on [219] with permission from Elsevier Publisher).
Materials 11 02295 g012
Figure 13. Images of TiO2-ZnO oxide materials: (a,b) SEM for samples before calcination; (c,d) SEM for calcined materials; (e) EDS mapping and (f) EDX spectrum (created based on [238] with permission from Springer).
Figure 13. Images of TiO2-ZnO oxide materials: (a,b) SEM for samples before calcination; (c,d) SEM for calcined materials; (e) EDS mapping and (f) EDX spectrum (created based on [238] with permission from Springer).
Materials 11 02295 g013
Figure 14. (a) SEM image and (b) TEM image of the TiO2-ZrO2 oxide system (created based on [254] with permission from Elsevier Publisher).
Figure 14. (a) SEM image and (b) TEM image of the TiO2-ZrO2 oxide system (created based on [254] with permission from Elsevier Publisher).
Materials 11 02295 g014
Table 1. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO oxide systems obtained by the hydrothermal method.
Table 1. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO oxide systems obtained by the hydrothermal method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O,
ethanol, deionized water
hydrothermal treatment: 120, 180, 200 °C,
reaction time: 1, 12, 24, 48 h,
drying: 60 °C,
calcination: 350 °C for 1 h
anatase crystal structure, spherical shaped particles with a tendency to agglomerate and particle sizes in the range 413–527 nm[8]
TiO2 microspheres (P25 Degussa),
Zn(NO3)2·6H2O,
hexamethylenetetramine, distilled water
hydrothermal treatment: 100 °C,
reaction time: 1–5 h,
drying: 60 °C
crystal structure of anatase and wurtzite, spherical shaped particles coated with nanospindles, particle diameter 1.5–2 μm, BET surface area 122.2–34.8 m2/g (reduction of specific surface area with increase of ZnO fraction)[56,57]
Ti(SO4)2, ZnCl2, NH3·H2O, cetyltrimethylammonium bromide,
distilled water
hydrothermal treatment: 100 °C,
reaction time: 4 h,
drying: 60 °C
crystal structure of TiO, Ti3O5 and ZnO, particles with a smooth surface with a tendency to agglomerate[58]
Ti[OCH(CH3)2]4,
Zn(NO3)2·6H2O, NaOH,
distilled water, poly(vinyl alcohol)
hydrothermal treatment: 220 °C,
reaction time: 5 h,
drying: 80 °C for 2 h
crystal structure of anatase and wurtzite, particles with spherical and cubic shapes with a tendency to agglomerate, particle diameter 30 nm[59]
Ti(OC4H9)4, Zn(CH3COO)2·2H2O,
acetic acid, ethylene glycol, deionized water
hydrothermal treatment: 180, 190, 200 °C,
reaction time: 8, 12, 15, 24 h,
drying: 60 °C
crystal structure of anatase and wurtzite, particle size distribution in the range 25–100 nm with average particle diameter 64 nm, BET surface area 97 m2/g and 206 m2/g for systems obtained at 180 and 200 °C respectively[60]
Ti(SO4)2,
Zn(NO3)2·6H2O, NH4F, CO(NH2)2, ethanol, distilled water
hydrothermal treatment: 120-180 °C,
reaction time: 2–24 h
drying: 80 °C for 24 h
crystal structure anatase and wurtzite, spherical particles with a tendency to agglomerate, nanoparticles of diameter 0.5–2 µm[61]
Ti[OCH(CH3)2]4,
Zn(CH3COO)2·2H2O,
NH3·H2O, ethylene glycol
hydrothermal treatment: 220 °C,
reaction time: 5 h,
drying: 60 °C
crystal structure of anatase and wurtzite, particles with a tendency to agglomerate, porous microstructure[62]
TiCl4, ZnCl2, urea, ethanol, distilled waterhydrothermal treatment: 180 °C,
reaction time: 16 h,
calcination: 450 °C for 2 h
crystal structure of anatase and wurtzite, particles taking the shape of nanorods with a tendency to agglomerate[63]
TiCl4, Zn(NO3)2·6H2O, NH3*H2O, ethanol, distilled waterhydrothermal treatment: 150 °C,
reaction time: 1 h,
freeze-drying: −55 °C in vacuum,
calcination: 600, 700, 900 °C for 1 h
crystal structure of rutile, wurtzite, Zn2TiO3O8, spherical particles with a tendency to agglomerate, particle diameters in the range 140–270 μm[64]
Ti[OCH(CH3)2]4,
Zn(CH3COO)2·2H2O,
sodium hydroxide, deionized water
microwave treatment: 180 °C for 5 min,
frequency: 2.45 GHz,
calcination: 500, 600 °C
crystal structure of anatase, wurtzite and zinc titanates (ZnTiO3, Zn2Ti3O4), spherical and hexagonal particles with a tendency to agglomerate, mean particle size 36 and 31 nm for samples calcined at 500 and 600 °C[65]
Ti(OC4H9)4, Zn(CH3COO)2,
deionized water
microwave treatment: 80 °C for 30 min,
drying: 80 °C
crystal structure of anatase and wurtzite, band gap energy 3.15 eV, spherical shaped particles with a tendency to agglomerate, BET surface area 290 m2/g, pore diameter 3.4 nm, pore volume 0.32 cm3/g[66]
Table 2. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the hydrothermal method.
Table 2. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the hydrothermal method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
TiO2, ZrOCl2·8H2O, NaOH, distilled waterhydrothermal treatment: 180 °C,
reaction time: 48 h,
drying: 60 °C,
calcination: 500 °C for 3 h
crystal structure of TiO2 and tetragonal ZrO2, nanorod particles up to 5 μm in length[82]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, propan-2-olhydrothermal treatment: 240 °C,
reaction time: 24 h,
drying: 100 °C for 2 h,
calcination: 450 °C for 4 h
crystal structure of anatase, rutile and tetragonal ZrO2, spherical particles with a tendency to agglomerate[83]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, 1,3-butanediol, ethylene glycolhydrothermal treatment: 300 °C,
reaction time: 2 h,
drying: 60 °C,
calcination: 500 and 800 °C
crystal structure of tetragonal ZrO2, spherical particles with diameter up to 1 μm[84]
TiOSO4, Zr(SO4)2, distilled waterhydrothermal treatment: 200, 240 °C,
reaction time: 24 h,
drying: 60 °C,
calcination: 400–1000 °C for 1 h
anatase crystal structure (for systems obtained at a molar ratio of TiO2 > 50%) and monocrystalline ZrO2 (for all synthesized systems), spherical particles with diameter 13 nm[85]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, Zn(NO3)2·6H2O, propan-2-ol, nitric acidhydrothermal treatment: 240 °C,
reaction time: 24 h,
drying: 100 °C for 2 h,
calcination: 450 °C for 4 h
crystal structure of anatase, rutile and monolithic ZrO2, average crystallite size 17.4 nm, band gap for TiO2-ZrO2 = 2.15 eV[86]
Ti[OCH(CH3)2]4, ZrOCl2·8H2O, poly(methyl methacrylate), ethanolhydrothermal treatment: 160 °C,
reaction time: 10 h,
drying: 100 °C,
calcination: 800 °C
crystal structure of anatase and monolithic ZrO2 (for a calcined system), particles with a structure of hollow microspheres, diameters in the range 0.7–2 μm, BET surface area 224.6 m2/g[87]
TiCi4, ZrOCl2·8H2O, NH3·H2O, distilled waterhydrothermal treatment: 220 °C,
reaction time: 4 h,
drying: 120 °C,
calcination: 500 °C for 10 h
amorphous structure, high thermal stability (10% mass loss), BET surface area 209 m2/g[88]
commercial Ti, ZrOCl2·8H2O, NH3·H2O, H2O2, HNO3, lactic acidhydrothermal treatment: 180 °C,
reaction time: 12 h,
drying: 50 °C for 12 h
crystal structure of anatase and ZrO2, spherical and spindle shaped particles[89]
Ti[OCH(CH3)2]4, ZrCl2, HNO3, methanolhydrothermal treatment: 260 °C,
reaction time: 12 h,
drying: 80 °C for 24 h
crystal structure of anatase, BET surface area 40.6 m2/g, pore volume 0.183 cm3/g, pore diameter 5.7 nm[90]
Table 3. Resources, synthesis conditions and physico-chemical properties of TiO2-MoS2 hybrid systems obtained by the hydrothermal method.
Table 3. Resources, synthesis conditions and physico-chemical properties of TiO2-MoS2 hybrid systems obtained by the hydrothermal method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
TiCl4, Na2MoO4·2H2O, thioacetamide, deionized waterhydrothermal treatment: 240 °C,
reaction time: 24 h,
drying: 60 °C
crystal structure of TiO2 (anatase) and MoS2, layered structure of MoS2 particles, molybdenum disulfide particles coated with TiO2 nanoparticles (3–5 nm)[33]
TiCl4, (NH4)6Mo7O24·4H2O, thiourea, ethanol, deionized waterhydrothermal treatment: 220 °C,
reaction time: 24 h,
drying: 60 °C for 12 h,
calcination: 500 °C for 4 h
crystal structure characteristic for TiO2 (anatase) and MoS2, 8% loss of mass when heated to 700 °C, particles in the form of a single sheet, 20 nm TiO2 particles evenly distributed on the surface of MoS2[34]
Ti(OC4H9)4, MoS2, ethylene glycol, HCl, deionized waterhydrothermal treatment: 180 °C,
reaction time: 12 h
crystal structure of anatase and 2H-MoS2, BET surface area = 48.2 m2/g, MoS2 nanoparticles coated with TiO2 particles[96]
Ti(OC4H9)4, Na2MoO4·2H2O, thiourea, HF, deionized waterpreparation of TiO2 hydrothermal treatment: 180 °C,
reaction time: 24 h
preparation of TiO2-MoS2:
hydrothermal treatment: 200 °C,
reaction time: 24 h,
drying: 105 °C for 24 h,
calcination: 400 °C for 4 h
crystal structure of TiO2 (anatase) and MoS2 (for the molar ratio Mo:Ti = 7.5%), BET surface area = 86, 87, 98, 92 m2/g for systems obtained with molar ratios Mo:Ti = 2.5, 5, 7.5% respectively, TiO2 coated with MoS2 particles[97]
TiO2-P25 (Degussa), Na2MoO4·2H2O, thiourea, HCl, NaOH, deionized waterpreparation of TiO2: hydrothermal treatment: 180 °C,
reaction time: 24 h,
preparation of TiO2-MoS2:
hydrothermal treatment: 180 °C,
reaction time: 24 h,
drying: 80 °C for 12 h
crystal structure characteristic for TiO2 and MoS2, MoS2 particles in the form of nanofibres with embedded spherical TiO2 agglomerates[98]
TiCl4, MoS2, ethanol, glycerol, sodium hexametaphosphateobtaining of TiO2: sonication
preparation of TiO2-MoS2:
hydrothermal treatment: 140 °C,
reaction time: 3 h
crystal structure of TiO2 (anatase), M3O8·2H2O (ilsemanite) and 3R-MoS2 (molybdenite), particle diameter 46 nm, MoS2 particles coated with spherical TiO2 particles, 23% mass loss of the TiO2-MoS2 hybrid when heated to 900 °C[99]
Ti, Na2MoO4·2H2O, thioacetamide, deionized water, acetone, ethanolpreparation of TiO2: anodic process on steel substrates in DSMO/HF electrolyte, calcination: 500 °C for 4 h
preparation of TiO2-MoS2:
hydrothermal treatment: 220 °C,
reaction time: 24 h,
drying: 80 °C for 12 h
crystal structure of anatase and hexagonal MoS2, particles in the form of nanotubes coated with MoS2[100]
TiF4, Na2MoO4·2H2O, thioacetamide, deionized waterpreparation of TiO2:
hydrothermal treatment: 140 °C,
reaction time: 1.5 h
preparation of TiO2-MoS2:
hydrothermal treatment: 220 °C,
reaction time: 24 h,
drying: 60 °C for 12 h
crystal structure of TiO2 (anatase) and rhombohedral MoS2, disappearance of anatase with an increase in the fraction of MoS2, TiO2 spheres covered with MoS2 nanoparticles[101,102,103]
Ti(OC4H9)4, MoS2, H2SO4, N,N-dimethylformamide, ethanol, deionized water,hydrothermal treatment: 200 °C,
reaction time: 20 h
crystal structure of TiO2 (anatase) and MoS2, dense covering of molybdenum disulfide particles with titanium dioxide particles of 10 nm diameter[104]
Ti(OC4H9)4, Na2MoO4·2H2O, thioacetamide, HFhydrothermal treatment: 200 °C,
reaction time: 24 h,
drying: 80 °C for 12 h,
calcination: 400 °C for 1 h
crystal structure of TiO2 (anatase) and MoS2, number of MoS2 layers in the range 6–9, TiO2 particles applied to MoS2 nanoparticles[105]
Table 4. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO hybrid systems obtained by the sol-gel method.
Table 4. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO hybrid systems obtained by the sol-gel method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
TiCl4, Zn(CH3COO)2·2H2O, ZnCl2, benzyl alcohol, propan-2-ol, deionized waterZnO sol: Zn(CH3COO)2·2H2O
(sample A) or ZnCl2 (sample B), propan-2-ol
TiO2-ZnO system: ZnO sol, TiCl4 mixed in the ratio TiO2:ZnO = 9:1 (sample A);
5:5 (sample B)
calcination: 500 °C for 6 h (sample A);
200; 400; 550; 600 °C for 2 h (sample B)
crystal structure of anatase (sample A) and anatase, rutile and zinc titanate (sample B), spherical shaped particles with agglomeration[1]
TiO2, Zn(CH3COO)2·2H2O, diethylamine, propan-2-ol, deionized waterTiO2 sol: TiO2 compact layer was deposited on an ITO (indium tin oxide) coated glass substrates by RF-sputtering technique with RF-power of 150 W,
calcination: 450 °C for 30 min
ZnO sol: Zn(CH3COO)2·2H2O, propan-2-ol
TiO2-ZnO system: the TiO2 layer was immersed in ZnO sol solution
bands on XPS spectra—Ti2s, Ti3s, Ti2p, Zn2p, Zn3p, Zn3s, O1s, spherical particles, TiO2 particles coated with ZnO (core-shell)[152]
Ti(OC4H9)4, Zn(CH3COO)2·2H2O,
ethylamine, glacial acetic acid, ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, ethanol, acetic acid
ZnO sol: Zn(CH3COO)2·2H2O, ethanol, ethylamine, acetic acid
TiO2-ZnO system: the sols of the respective oxides were mixed at molar ratio TiO2:ZnO = 7:3, 5:5, 3:7
antistatic properties, homogeneous surface of polyester material coated with a TiO2-ZnO oxide system[153]
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O,
glacial acetic acid, diethylamine, ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, ethanol, acetic acid
ZnO sol: Zn(CH3COO)2·2H2O, ethanol
TiO2-ZnO system: obtained sols were mixed at molar ratio TiO2:ZnO = 1:0; 3:1; 1:3; 0:1, the obtained materials were deposited on carbon steel by immersion in solution,
drying: 70 °C for 10 h
calcination: 350, 500 °C for 2 h
crystal structure of Fe, TiO2, Fe3O4, ZnFe2O4,
morphology – rough surfaces with numerous cracks,
EDX spectrum showed bands characteristic
for Ti, O, Fe, Zn
[154]
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O,
hydrochloric acid, anhydrous ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, ethanol, HCl
ZnO sol: Zn(CH3COO)2·2H2O, ethanol, deionized water
TiO2-ZnO system: materials obtained at molar ratio ZnO:TiO2 = 0.1:0.15; 0.2:0.25; 0.3:0.35,
calcination: 450; 480; 500; 550; 600 °C
for 1; 1.5; 2; 2.5; 3 h
crystal structure of anatase and wurtzite, particles
with spherical shape with a tendency to agglomerate, BET surface area 76.258 m2/g, pore volume 0.0361 cm3/g, pore diameter 6.6 nm
[155]
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O,
nitric acid, ethylene glycol, glycerol, trimethylamine (TEA), anhydrous ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, nitric acid, ethanol
ZnO sol: Zn(CH3COO)2·2H2O, ethylene glycol, glycerol, TEA, ethanol
TiO2-ZnO system: obtained sols were mixed at molar ratio TiO2:ZnO = 100:0, 75:25, 50:50, 25:75, 0:100,
calcination: 500 °C for 1 h
anatase crystal structure (TiO2:ZnO = 100:0, 75:25), amorphous structure (TiO2:ZnO = 50:50, 25:75), wurtzite structure (TiO2:ZnO = 0:100), XPS shows the presence of Ti2p, Zn2p, O1s bands[156]
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O, nitric acid, acetic acid, diethylamine, acetylacetone, anhydrous ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, ethanol, acetylacetone, acetic acid, diethylamine
ZnO sol: Zn(CH3COO)2·2H2O, ethanol, diethylamine
TiO2-ZnO system: obtained sols were mixed at molar ratio TiO2:ZnO = 90:10, 80:20, 70:30, 60:40,
calcination: 500 °C for 30 min
anatase crystal structure (TiO2:ZnO = 90:10, 80:20), amorphous structure (TiO2:ZnO = 70:30, 60:40), spherical shaped particles and porous structure, XPS shows bands characteristic for Ti2p, Zn2p, O1s[157]
Ti[OCH(CH3)2]4,
Zn(NO3)2·6H2O, glacial acetic acid, ethanol, deionized water
TiO2 sol: Ti(OC4H9)4, ethanol
ZnO sol: Zn(NO3)2·6H2O, ethanol, glacial acetic acid
TiO2-ZnO: obtained sols were mixed at molar ratio TiO2:ZnO = 10:3,
calcination: 400 °C for 2 h
crystal structure of anatase and zincite, particles with spherical shape and a tendency to agglomerate, EDX shows the presence of bands derived from Ti, Zn, O[158]
Ti(OC4H9)4, Zn(NO3)2·6H2O,
citric acid, deionized water
TiO2-ZnO system: sol-gel auto-ignition method, citric acid (catalyst) Ti(OC4H9)4, Zn(NO3)2·6H2O was heated to a temperature of 300 °C, auto-ignition,
calcination: 500 °C for 5 h
crystal structure of anatase, BET surface area = 52, 63.4, 69.9, 64, 36.6 m2/g respectively for systems prepared with a ZnO mass fraction of 1, 5, 10, 12, 30%, particles of irregular shape with a tendency to agglomerate, EDX shows bands characteristic for Ti, O, Zn[159]
Ti[OCH(CH3)2]4, Zn(CH3COO)2·2H2O, acetic acid, diethylamine, butanol, propan-2-ol, deionized waterTiO2 sol: Ti(OC4H9)4, butanol, acetic acid
ZnO sol: Zn(CH3COO)2·2H2O, propan-2-ol, diethylamine, deionized water
TiO2-ZnO system: the glass plate was covered with a layer of TiO2 sol,
calcination: 400 °C for 2 h,
covering by ZnO layer,
calcination: 500, 600 °C for 2 h
crystal structure of anatase and wurtzite, increase of surface roughness with increasing calcination temperature[160]
Ti(OC4H9)4,
Zn(CH3COO)2·2H2O, sodium hydroxide, propan-2ol, deionized water
sol of TiO2: Ti(OC4H9)4, propan-2-ol
sol of ZnO: Zn(CH3COO)2·2H2O, sodium hydroxide (0.5 M), deionized water
TiO2-ZnO system: obtained sols was mixed at molar ratio TiO2:ZnO = 0.5:0.25; 1:0.5;
1.5:0.75; 2:1, are designated respectively as TZO1 TZO2, TZO3 and TZO4,
calcination: 550 °C for 4 h
crystalline structure of anatase and wurtzite, Raman spectroscopy - bands characteristic for TiO2 and ZnO, spherical shaped particles and rods with tendency to agglomerate, EDX—strands derived from Ti, Zn, O, BET surface area equal to 74.7, 32.2, 12.58.
8.82 m2/g respectively for samples
TZO1 TZO2, TZO3, and TZO4
[161]
Ti(OC4H9)4, Zn(CH3COO)2·2H2O, nitric acid, ethylene glycol, ethanol, deionized waterTiO2 sol: Ti(OC4′H9)4, ethanol, nitric acid
ZnO sol: Zn(CH3COO)2·2H2O, ethylene glycol
TiO2-ZnO system: TiO2 and ZnO sols were mixed at molar ratio TiO2:ZnO = 2:1,
calcination: 450 °C for 1 h
particles of spherical shape with no tendency to agglomerate, EDX showed characteristic bands for Ti, Zn, O, energy band gap 3.41 eV[162]
Ti(OC4H9)4, Zn(CH3COO)2·2H2O, nitric acid, diethylamine, acetylacetone, propan-2-ol,
deionized water
TiO2 sol: Ti(OC4H9)4, acetylacetone, propan-2-ol, nitric acid
ZnO sol: Zn(CH3COO)2·2H2O, propan-2-ol, diethylamine
TiO2-ZnO system: obtained sols were mixed in equimolar ratio
irregular shaped particles, EDX showed bands characteristic for Ti, Zn, O[163]
TiCl4, Zn(CH3COO)2·2H2O, ammonium fluoride, acetonitrile, butanol, acetone, diethylamine, ethanol, deionized waterTiO2: electrode method
ZnO sol: Zn(CH3COO)2·2H2O, ethanol diethylamine
TiO2-ZnO system: TiO2 (anodic method) was immersed in a solution of TiCl4,
calcination: 450 °C for 15 min,
TiO2 was covered by ZnO layer,
drying: 180 °C for 10 min,
calcination: 500 °C for 1 h
particles with shapes similar to nanotubes and spherical, EDX showed characteristic bands for Ti, Zn, O[164]
Ti[OCH(CH3)2]4,
Zn(CH3COO)2·2H2O, acetic acid, propan-2-ol
TiO2-ZnO system: to Ti[OCH(CH3)2]4 was added Zn(CH3COO)2·2H2O dissolved in propan-2-ol (molar ratio Ti:Zn = 5:1),
calcination: 400 °C for 4 h
crystal structure of anatase and wurtzite, particles with a spherical shape with agglomeration, BET surface area = 91 m2/g, pore size 1.49 nm, pore volume 0.343 cm3/g[165]
Table 5. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the sol-gel method.
Table 5. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the sol-gel method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
Ti(OC4H9)4, ZrOCl2·8H2O, nitric acid, polyethylene glycol (PEG), (PEO)20(PPO)70(PEO)20 (Pluronic P123, MW = 5800, Aldrich), ethanol, deionized waterTiO2-ZrO2 oxide system: single-step synthesis, mixed with Ti(C4H9)4: PEG:P123 with ZrOCl2·8H2O (molar ratio Ti:Zr = 1:0.1),
aging: 24 h,
calcination: 800 °C for 5 h
crystal structure of anatase, tetragonal zirconia rutile, BET surface area = 148.9; 138.5; 136.9 m2/g for
TiO2-ZrO2(P123 + PEG); TiO2-ZrO2(PEG); TiO2-ZrO2(P123)
[29]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, nitric acid, deionized waterTiO2-ZrO2 oxide systems: M1 and M2
M1: Ti0.9Zr0.1O2, hydrolysis of precursors in an aqueous environment
M2: Ti0.9Zr0.1O2, polymerization of precursors in propan-2-ol
drying: 100 °C,
calcination: 350 °C for 2 h
crystal structure of anatase and brookite, Raman spectroscopy showed anatase-specific bands, BET surface area = 313 m2/g (M1); 269 m2/g (M2) [188]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, ammonia water, propan-2-ol, distilled waterTiO2 sol: Ti[OCH(CH3)2]4, propan-2-ol, ammonia water
ZrO2 sol: Zr[OCH(CH3)2]4, propan-2-ol, ammonia water
TiO2-ZrO2 oxide systems: sol was mixed with ZrO2 in a molar fraction of 3%, 6%, 13%, 37%,
drying: 100 °C for 24 h,
calcination: 550, 700 °C for 5 h
crystal structure of anatase and rutile, spherical particle shape with agglomeration, BET surface area = 26, 37, 40, 172 m2/g for materials calcined at 550 °C and 26, 29, 30, 36 m2/g for materials calcined at 700 °C, for systems with ZrO2 molar fractions 3%, 6%, 13%, 37% respectively[189]
Ti[OCH(CH3)2]4, Zr(CH3COO)2·2H2O, hydrochloric acid, hydroxypropylcellulose, deionized waterZrO2 sol: Zr(CH3COO)2·2H2O, hydroxypropylcellulose
TiO2 sol: Ti[OCH(CH3)2]4, HCl, water
TiO2-ZrO2: the sol was mixed maintaining the ratio Ti:Zr = 0.5:0.5,
drying: 150 °C for 1 h,
calcination: 300, 500, 700, 900 °C for 1 h
crystal structure of anatase and zirconium dioxide, crystallite size 5.8 and 8.5 nm for calcination temperatures of 500 °C and 900 °C, particles with a spherical shape with a tendency to agglomerate[190]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, poly(ethylene oxide), (PEO)20(PPO)70(PEO)20 (Pluronic P123, BASF), acetylacetone, hydrochloric acid, deionized waterTiO2 sol: Ti[OCH(CH3)2]4, HCl, water
ZrO2 sol: Zr[OCH(CH3)2]4, HCl, water
TiO2-ZrO2: the sol was mixed,
aging: 12 h,
calcination: 400, 500 °C for 2 h
crystal structure of anatase, incorporation of zirconium dioxide into titania network forming Ti1−xZrxO2 with anatase crystal structure, Raman spectroscopy showed anatase and ZrO2 characteristic bands, spherical shaped particles densely packed with diameter 10–15 nm[191]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, lauramine hydrochloride, acetyl acetone, diethylamine, deionized waterTiO2-ZrO2 oxide system: single-step synthesis, mixed Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4 acetylacetone and laurylamine,
aging: 7 days,
calcination: 500–900 °C
crystal structure of anatase, rutile and tetragonal zirconia, particles with spherical shape with agglomeration[27]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, nitric acid, ethanol, deionized waterTiO2-ZrO2: one-stage synthesis, hydrolysis of Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4 in ethanol:waterFourier-transform infrared spectroscopy (FTIR) spectroscopy showed bands characteristic for TiO2 and ZrO2, distances between planes 3.533 Å [192]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, diethylamine (DEA), isoeugenol (ISOH), propan-1-ol, deionized waterTiO2-ZrO2 oxide system: one-stage synthesis, Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4 mixed with a chelating complex (diethylamine or isoeugenol),
calcination: 500, 650 °C in air and 350, 400 °C in a nitrogen atmosphere
amorphous structure (materials calcined at 500 °C), crystal structure of anatase and rhombic TiO2-ZrO2 (materials calcined at 650 °C), BET surface area = 7.8 m2/g (TiO2-ZrO2-ISOH) and 8.3 m2/g
(TiO2-ZrO2-DEA) for samples calcined in a nitrogen atmosphere, 240 m2/g (TiO2-ZrO2-ISOH) and 186 m2/g (TiO2-ZrO2-DEA) for samples calcined in air
[193]
TiF4, Zr(NO3)4, anodic aluminium oxide (AAO), ethanol, sulphuric acid, deionized waterTiO2-ZrO2 oxide system: single-step synthesis, AAO was immersed in Zr(NO3)4,
drying: 80 °C for 1 h,
immersed in TiF4 for 9 min,
calcination: 600 °C for 3 h
crystal structure of anatase and monoclinic zirconia, nanotube-shaped particles densely packed, BET surface area = 47.4 m2/g[26]
Ti(OC4H9)4, Zr(OC4H9)4, butan-1-ol, ammonium hydroxide, deionized waterTiO2-ZrO2 oxide system: single-step synthesis, hydrolysis of Ti(OC4H9)4, Zr(OC4H9)4 in butan-1-ol:water, ammonium hydroxide was added to neutralize the pH,
drying: 120 °C for 12 h,
calcination: 500 °C for 12 h
crystal structure of anatase, monoclinic and tetragonal zirconia, Raman spectroscopy showed anatase-specific bands, BET surface area = 68, 62 and 60 m2/g for systems obtained at molar ratio TiO2:ZrO2 = 3:1, 1:1, 1:3[194]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, ethanol, deionized waterTiO2-ZrO2: single-step synthesis, hydrolysis of Ti(OC4H9)4, Zr(OC4H9)4
in ethanol:water,
drying: 100 °C for 6 h,
calcination: 550 °C for 2 h
crystal structure of anatase, monoclinic and tetragonal zirconia, spherical shaped particles with a pronounced tendency to agglomerate, BET surface area 32.5, 42.5, 36 m2/g for 67%TiO2-33%ZrO2, 50%TiO2-50%ZrO2, 33%TiO2-67%ZrO2 [195]
Table 6. Process parameters affecting the electrospinning method.
Table 6. Process parameters affecting the electrospinning method.
Process Parameters
SolutionApparatusExternal Environment
type of solventapplied voltagetemperature
concentrationdistance of nozzle from collectorhumidity
-diameter and length of nozzle-
Table 7. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO hybrid systems obtained by the electrospinning method.
Table 7. Resources, synthesis conditions and physico-chemical properties of TiO2-ZnO hybrid systems obtained by the electrospinning method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
Ti[OCH(CH3)2]4,
Zn(CH3COO)2·2H2O, polyvinylpyrrolidone, acetic acid, ethanol, deionized water
solution of TiO2 and ZnO: Zn(CH3COO)2·2H2O, acetic acid and Ti[OCH(CH3)2]4 dissolved in a mixture of ethanol and polyvinylpyrrolidone
TiO2-ZnO electrospinning:
voltage: +7 kV, spinning speed 0.3 cm3/h
calcination: 500 °C for 3 h
nanofibre shaped particles, X-ray photoelectron spectroscopy (XPS) shows bands derived from Ti2p, O1s, Zn2p[232]
Ti[OCH(CH3)2]4,
Zn(C2H5)2, dimethylformamide (DMF), acetic acid, poly(vinyl acetate), deionized water
preparation of TiO2 solution: Ti[OCH(CH3)2]4, acetic acid, dimethylformamide, polyvinyl acetate
TiO2 electrospinning: voltage: +10 kV,
spinning speed 0.2 cm3/h,
calcination: 600 °C for 8 h
TiO2-ZnO oxide system: ZnO was applied
by atomic layer deposition (ALD)
crystal structure of anatase and wurtzite, nanofibre particles of core-shell structure, average nanofibre diameter 250 nm, increase of ZnO coating thickness 0.66 nm per ALD cycle[233]
Ti[OCH(CH3)2]4, ZnCl2, polyvinylpyrrolidone, (PEO)20(PPO)70(PEO)20 (Pluronic P123), ethanol, deionized waterpreparation of TiO2 and ZnO precursors: polyvinylpyrrolidone dissolved in ethanol,
Ti(OCH(CH3)2)4 solution added
to aqueous ZnCl2 solution,
calcination: 500 °C for 4 h
TiO2-ZnO electrospinning: voltage: +20 kV
crystal structure of anatase, wurtzite and zinc titanate (ZnTiO3), particles with diameter 20 nm, BET specific surface area = 203 m2/g, pore volume 0.24 cm3/g[234]
Ti[OCH(CH3)2]4, Zn(CH3COO)2·2H2O, cellulose acetate, dimethylformamide (DMF), acetic acid, acetone, deionized waterpreparation of TiO2 and ZnO precursors:
Ti[OCH(CH3)2]4, Zn(CH3COO)2·2H2O
cellulose acetate was dissolved in a mixture of DMF:acetone = 1:2 (v/v)
TiO2-ZnO electrospinning: voltage:
+10 kV, spinning speed 10 μL/min,
calcination: 500 °C for 5 h
crystal structure of anatase, wurtzite and zinc titanate (ZnTiO3), diameter of nanofibres 85–200 nm, cylindrical nanofibres made of granular nanoparticles, XPS showed characteristic bands Ti2p, O1s and Zn2p[235]
Ti[OCH(CH3)2]4, ZnO, poly(vinyl acetate) (PVA), dimethylformamide (DMF), acetic acid, deionized waterpreparation of TiO2 and ZnO precursors: polyvinyl acetate was dissolved in DMF, added Ti[OCH(CH3)2]4 and ZnO
TiO2-ZnO electrospinning: voltage: +12 kV,
drying: 80 °C for 12 h,
calcination: 600 °C for 1 h
crystal structure of wurtzite, nanofibres with a smooth surface, EDX showed presence of elements Ti, Zn, O[236]
Ti[OCH(CH3)2]4, Zn(CH3COO)2·2H2O, polyvinylpyrrolidone, dimethylformamide (DMF), ethanol, deionized waterpreparation of precursors TiO2 and ZnO: Ti[OCH(CH3)2]4 and Zn(CH3COO)2·2H2O dissolved in DMF, polyvinylpyrrolidone was added
TiO2-ZnO electrospinning: voltage: +10 kV,
calcination: 400 °C for 2 h
crystal structure of anatase and wurtzite, nanofibres composed of individual nanoparticles, XPS showed presence of specific bands for Ti2p, Zn2p, O1s[237]
TiO2 (P25), Zn(CH3COO)2·2H2O, hexamethyltetramine, ethanol, demineralized waterelectrospinning of TiO2 fibres: P25 solution
and polymer in ethanol, applied voltage: +15 eV, TiO2 fibres mixed with solution 1
heating: 120 °C for 1 h
TiO2 nanofibres were immersed in solution 2 (aqueous zinc acetate solution and hexamethyl tetramine) and maintained at 85 °C for 24 h,
calcination: 500 °C for 1 h
crystal structure of anatase, rutile and wurtzite, TiO2 fibres coated with ZnO nanorods, EDX spectrum contained specific bands characteristic for Ti, O, Zn, Raman spectrum contained bands characteristic for TiO2, ZnO and ZnTiO3[238]
Table 8. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the electrospinning method.
Table 8. Resources, synthesis conditions and physico-chemical properties of TiO2-ZrO2 oxide systems obtained by the electrospinning method.
ResourcesSynthesis ConditionsPhysico-Chemical Properties of the Final ProductReferences
Ti(O4H9)4, ZrOCl2,
polyvinylpyrrolidone,
acetic acid, ethanol, deionized water
solution of ZrO2:ZrOCl2 was dissolved
in ethanol:water, poly(vinylpyrrolidone) was added, stirred for 12 h
at room temperature
TiO2:Ti(O4H9)4 solution was hydrolysed in ethanol:acetic acid, polyvinylpyrrolidone solution was added, stirred for 6 h at room temperature
TiO2-ZrO2 electrospinning:
voltage for the ZrO2 stream: +5 kV,
voltage for TiO2: −7 kV,
calcination: 600 °C for 3 h
crystal structure of anatase and tetragonal zirconia, presence of nanofibers with different diameters,
(130 nm and 70 nm for ZrO2 and TiO2 respectively)
[254]
Ti[OCH(CH3)2]4, Zr[OCH(CH3)2]4, polyvinylpyrrolidone, phosphoric acid, acetic acid, ethanol, deionized waterTiO2 and ZrO2 precursor solution:
polyvinylpyrrolidone dissolved in ethanol, mixed for 1 h, acetic acid, Ti [CH(CH3)2]4,
Zr[OCH(CH3)2]4 (atomic ratio Ti:Zr = 1:1)
TiO2-ZrO2 electrospinning: voltage 14 kV, spinning speed 30 μL/min,
calcination: 500, 700, 1000 °C for 6 h
amorphous structure for material calcined at 500 °C, crystal structure of srilankite (Ti0.5Zr0.5O2) for calcination temperature 700 and 1000 °C, BET surface area = 75 m2/g for material calcined at 500 °C, average fibre diameter 497 nm (before calcination) and 344 nm (after calcination at 500 °C)[255]

Share and Cite

MDPI and ACS Style

Kubiak, A.; Siwińska-Ciesielczyk, K.; Jesionowski, T. Titania-Based Hybrid Materials with ZnO, ZrO2 and MoS2: A Review. Materials 2018, 11, 2295. https://doi.org/10.3390/ma11112295

AMA Style

Kubiak A, Siwińska-Ciesielczyk K, Jesionowski T. Titania-Based Hybrid Materials with ZnO, ZrO2 and MoS2: A Review. Materials. 2018; 11(11):2295. https://doi.org/10.3390/ma11112295

Chicago/Turabian Style

Kubiak, Adam, Katarzyna Siwińska-Ciesielczyk, and Teofil Jesionowski. 2018. "Titania-Based Hybrid Materials with ZnO, ZrO2 and MoS2: A Review" Materials 11, no. 11: 2295. https://doi.org/10.3390/ma11112295

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