3.1. Hydrothermal Method of TiO2 Preparation
TiO
2 can be obtained by high-temperature hydrolysis of various precursors directly in the autoclave or by hydrothermal treatment [
179]. For example, nanosized TiO
2 powders are obtained by adding a 0.5M solution of titanium butylate in isopropanol to deionized water ([H
2O]/[Ti]=50). Then peptization is carried out at 70°C for 1 hour in the presence of hydroxycitetraalkylammonium [
180]. This method is also widely used for the synthesis of monodisperse TiO
2 nanoparticles. In [
181], TiO
2 nanorods were obtained by hydrothermal treatment of a dilute TiCl
4 solution in the temperature range of 333–423 °K with a synthesis duration of 12 h. In [
181], TiO
2 nanotubes were synthesized by hydrothermal treatment of commercial photocatalyst P25 in a 10 M aqueous NaOH solution at 130°C and 24 h of synthesis.
The glycolate-oxo-peroxo-titanium (IV) complex was subjected to hydrothermal treatment in [
182] to obtain titanium dioxide. In an ice bath at room conditions, 20 mmol of titanium metal powder was dissolved in a combination of 40 ml of hydrogen peroxide solution and 10 ml of ammonia solution. After two hours, all of the titanium powder dissolved and a yellow solution containing the peroxo-titanium complex was formed. Then 30 mmol of glycolic acid was immediately added and the mixture was heated to 353 K to promote complexation and remove excess hydrogen peroxide and ammonia until it formed an orange gel. To obtain an aqueous solution of the ammonium salt of the glycolate-oxo-peroxo-titanium (IV) complex, it was dissolved in distilled water. Using 2, 4, 8, 12 and 16 cm
3 of stock solution and the required amount of distilled water, 40 cm3 of titanium solutions with Ti = 12.5, 25.0, 50.0, 50.0, 75.0 and 100 mM concentration were created in the next step. After the solution was placed in a 50 cm3 jar, it was sealed with a stainless steel shroud and heated in an oven for 1 to 168 hours at 473 K. The autoclaves were then allowed to cool to ambient temperature Centrifugation was used to separate the precipitate formed, which was then washed thoroughly three times with deionised water. After drying overnight at 353 K in an oven, the sample was extracted. The particle sizes of TiO
2 polymorphs in the samples that were extracted from the Ti complex solution with a concentration of 50.0 mM as a function of the procedure duration are shown in
Table 2.
The authors of the study [
183] used the hydrothermal method to efficiently grow TiO
2 nanorods. To reduce hydrolysis and condensation, an equal volume of acetylacetone was combined with tetrabutyl orthotitanate. The mixture was then gradually stirred for five minutes at room temperature by adding 40 milliliters of water. Thirty milliliters of 28-30% aqueous ammonia solution was gradually added to the mixture dropwise with continuous stirring. The solution was then transferred to a 250 mL stainless steel autoclave and immersed in a silicone oil bath. The precursor solution was then heated to 170 °C and stirred continuously at this temperature for twenty-four hours. The autoclave was then spontaneously cooled down to ambient temperature. The final product was repeatedly thoroughly purified with aqueous HCl, 2-propanol and water. It was then dried for 12 h at 120 °C. Finally, the collected samples were incinerated for one hour in a high temperature furnace at 450 °C. X-ray examination of the samples showed that they contained a large number of anatase nanorods with an average pore width of 3.1 nm and a specific surface area of about 34.82 m2/g. To enhance the charge transfer ability, the authors of another work [
184] obtained TiO
2 nanorods/nanoparticles using hydrothermal method. The obtained nanoparticles have a specific surface area of 84.83 m2/g and a pore diameter of 5.7 nm.
Using surfactant-assisted hydrothermal technology, TiO
2 with different morphologies including nanosheets, nanorods, nanotubes and nanoflowers were produced by adjusting the pH during the preparation process. The experimental results showed that the pH value is critical for controlling the shape of the generated TiO
2, as it can change the adsorption potential of surfactant on the surface of TiO
2 and its charge state in solution. The experimental protocols are summarised below: A solution resistant to hydrolysis at room temperature was created by mixing titanium isopropoxide and TEOA in the following ratio: TTIP: TEOA = 1:2. DI water was added to create solution A. Dodecanediamine was mixed with DI water to create solution B, which acts as a shape adjuster. Next, solution A and solution B were mixed. The pH of the mixed solution was adjusted by adding HClO
4 or NaOH. It was then placed in a 100 mL Teflon autoclave and incubated at 100°C for 24 h, after which the temperature was raised to 140°C for 72 h for nucleation and growth of TiO
2 particles. TiO2 nanospheres with particle sizes ranging from 30 to 60 nm were formed under acidic conditions at pH = 5.6. When the pH was increased, most of the prepared TiO
2 particles transformed from ellipsoids to nanorods. When the pH of the solution was changed to a value greater than 11, TiO
2 nanoflowers were formed. In the process of creating these structures, TiO2 nanosheets were initially formed and then folded to form the final structures at different pH values [
185].
Hollow TiO
2 nanospheres can be obtained by hydrothermal method with the addition of aggressive chemicals, and they usually have large surface area and low density [
185,
186,
187]. For example, the hydrothermal method was used to obtain porous hollow TiO
2 aggregates with a BET surface area of 168 m2/g and an average pore size of 12 nm. Ti(SO
4)
2 and NH
4F were dissolved in DI water for Liu preparation, then the resulting mixture was stirred and placed in a Teflon-lined autoclave. The hydrothermal synthesis was carried out in an electric furnace for six hours at 160°C. The appearance of the products is shown in
Figure 7 (a-d). Despite having a greater estimated bandgap of 3.36 eV than P25 (~ 3.18 eV), the porous TiO
2 products exhibited double the activity of P25 when it came to the photodegradation of Rhodamine B [
187]. The HF produced by NH
4F during the hydrothermal process is responsible for the production of porous, hollow TiO
2 aggregates. HF, a chemical etchant that is corrosive, will erode the inside of TiO
2 to create TiF4, which will ultimately result in hollow nanospheres. Using a similar hydrothermal technique, their subsequent investigation employed the metallic Ti powder as reactants together with a specific quantity of NH
4F and H
2O
2 (30 wt.%). With a diameter of about 1 μm, a shell thickness of 150 nm and a cavity size of about 600 nm, the obtained anatase TiO
2 resembled hollow spheres. The critical factors for the creation of TiO
2 hollow spheres from metallic Ti powders were reaction time, NH
4F concentration and H
2O
2 concentration. The H
2O
2 served as both an oxidant and a bubble generator, creating O2 bubbles that served as the aggregation centre when combined with Ti particles. After that, as shown in
Figure 7(e, f) [
186], TiO
2 nanoparticles gradually aggregated at the gas-liquid interface to form hollow TiO
2 spheres.
The hydrothermal method can be used to create rutile-brucite TiO
2 nanocomposite by replacing NaOH with HCl. The obtained TiO
2 was used for photooxidation of methyl blue as well as for photocatalytic conversion of CO
2. In [
186], titanium chloride was combined with 2 mol/L HCl to obtain a clear solution. The solution was then combined with 1.0 mL of 5 V/V% Triton X-100 in ethanol. After transfer to a flask, the prepared solution was refluxed for 22 h at 100°C. The final rutile-brucite nanocomposite structures were obtained by centrifuging the obtained products, washing them in water, drying them at room temperature and then calcination at 500°C [
187].
3.2. Solvothermal Synthesis of TiO2
The solvothermal and hydrothermal methods are practically identical except that the solvothermal method uses a non-aqueous solvent. The temperature of the solvothermal process can be much higher because some organic solvents have higher boiling points [
188]. Usually, the size, shape and crystal structure of the obtained TiO
2 nanoparticles can be well controlled using the solvothermal method. The method is a versatile way to synthesize various nanoparticles with narrow size and dispersity distributions. Using the solvothermal method, TiO
2 nanoparticles with a characteristic size of less than 5 nm can be obtained [
189].
The alcohol solvothermal method was first applied to obtain TiO
2 by Kang et al. [
190]. In Kang’s study, 1,4-butanediol was used as a solvent and titanium isopropoxide was used as a source of TiO
2. The 1,4-butanediol was mixed with titanium isopropoxide at 300°C for 50 min to facilitate the synthesis. The obtained TiO
2 powder was cleaned repeatedly with acetone and then left to dry for five hours at 100°C without calcination. The obtained anatase TiO
2 nanoparticles of size 20-50 nm were highly hydrophilic and were much more efficient in photocatalytic decomposition of chloroform than TiO
2 nanoparticles obtained by sol-gel method [
190]. Subsequently, Nam and Han used the obtained TiO
2 for photodegradation of methyl orange and investigated the effect of several alcoholic solvents under the same conditions [
191]. They got ready in the following ways: Titanium isopropoxide (0.1 mol) was added to glycerol, 1-butanol, and 1,4-butanediol, in that order. The combinations were heated to 300°C and kept there for an hour under auto-generated pressure. The findings demonstrated that the kinds of solvents employed during the reaction had a significant impact on the physical characteristics of the produced TiO
2, including crystal size, shape, and structure.
Figure 8 (a-c) displays the SEM pictures of the TiO
2 that they produced [
191].
The authors of [
192] synthesized anatase TiO
2 nanosheets with pronounced {001} facets using an alcohol solvothermal technique. This procedure is a common method for fabricating nanosheets and can provide insight into subsequent methods. In a typical experiment, the pH was set to 1.8 and sufficient TiF
4 was added to a combination of hydrochloric acid and DI water to create a TiF
4 solution. Then 14.5 mL of the above aqueous solution of TiF
4, 13.38 mL of 2-propanol, and 0.5 mL of HF were placed in a Teflon-coated stainless steel autoclave. The autoclave was kept at 180°C for 5.5-44 h in an electric oven. Monocrystalline anatase TiO
2 nanosheets were obtained by centrifugation after the reaction, then washed three times with DI water and dried under vacuum overnight. Heat treatment at 600°C for 90 min was used to remove fluorine from the surface of anatase TiO
2. They showed that 2-propanol and HF promoted the isotropic development of single-crystalline TiO
2 nanosheets and that 2-propanol could enhance the stabilizing effect associated with fluorine adsorption on the (001) surface using first-order theoretical calculations.
Figure 8 (d, e) shows the SEM images of the anatase TiO
2 nanosheets. TiO
2 nanosheets have the potential to remove organic pollutants by photocatalysis because they can produce five times more oxidative hydroxyl radicals (
·OH) than P25 upon irradiation [
193].
Huang et al. [
193] prepared highly crystalline TiO
2 hollow spheres by alcohol solvothermal method without corrosion additives and shape controllers at 350°C, as shown in
Figure 9. Although the surface area of the TiO
2 hollow spheres was only 28.2 m2/g by BET, they possessed good ability for photocatalytic degradation of methyl orange. Titanium n-butoxide (TNB) and ethanol (EtOH) were used for their synthesis. A mixed solution of TNB and EtOH was prepared by slowly adding TNB to EtOH with a certain volume ratio.
After stirring for 30 min, the mixture was transferred to a stainless steel autoclave for the solvothermal reaction at 350 °C for 4 h. The precipitate was then washed three times with anhydrous ethanol and dried overnight. The precipitate was then washed three times with anhydrous ethanol and dried overnight. The prepared anatase TiO
2 spheres consisted of nanoparticles with an average diameter of 30.2 nm. According to their observations, pure TNB can decompose into TiO
2 nanoparticles at temperatures above 350°C to form 1-butene as in the reaction Ti(OBu)
4 →TiO
2+CH
3CH
2CHCHCHCH, which can serve as bubble templates for growing hollow TiO
2 microspheres from nanoparticles [
195].
With LA serving as a suitable coordination surfactant to promote the anisotropic crystal growth of TiO
2, nanorods would be formed. If controlling NH
4HCO
3 and LA in a desired mole ratio, TiO
2 with different morphologies could be acquired.
Figure 10 are the TEM images of the TiO
2 nanoparticles and nanorods prepared. Adding a small amount of metal chloride or nitrate into the mixed solution of NH
4HCO
3, LA, triethylamine, cyclohexane and Ti(OBu)
4, metal-doped (Fe
3 +, Co
2 +, Sn
4 +, Ni
2 +) TiO
2 nanocrystallines can be obtained [
196].
Atomically thin anatase TiO
2 wires with a diameter of 4.5 Å were prepared by the carboxylic acid solvothermal process. These diameter-tunable, ultrathin TiO
2 nanowires were obtained as a result of optimized reaction temperature and reaction time. In Yang’s synthesis [
196], a certain amount of oleic acid and cyclohexane were mixed, and then Ti(OBu)
4 was added dropwise to the mixed solution. The resulting solution was heated to 150°C for 25 hr in a Teflon-lined stainless steel autoclave. The precipitation was extracted with an excess of ethanol. The titanium complex precursor was then redispersed in a mixture of octadecene, oleic acid, and oleylamine. The solution was heated to 180°C under stirring and maintained at that temperature for 1 hr to acquire N-doped TiO
2. Their research results suggested that a uniform mixture of oleic acid and oleylamine solution favors self-assembly of the wires. In addition, an increase in reaction temperature mainly increased the wire diameter, whereas prolongation of the precursor treatment time mainly caused an increase in the wire length. Their characterization data indicated that N-doping originated from the oxidative coupling of oleylamine on the surfaces of the atomically thin wires, forming the N–O–Ti surface structures. UV–vis absorption spectra indicated that the light absorption edge was 257 nm for the atomically thin TiO
2 wires, whereas it shifted to 600 nm with N doping.
Figure 10 (e) is a TEM image of the N-TiO
2 obtained [
197]. To further investigate the capping roles of oleic acid and oleylamine in the solvothermal system, Dinh found out that oleic acid and oleylamine had different binding strengths in controlling the growth of TiO
2 nanoparticles. By varying the ratio of them, TiO
2 with different shapes such as spherical, dog-bone, truncated and elongated rhombic was prepared. The thus-obtained TiO
2 was ascertained to be an excellent support for the synthesis of metal/TiO
2 photocatalyst in which metal clusters could be uniformly deposited on the surface of TiO
2 [
198].
Thermal decomposition of titanium alkoxides by the solvothermal reaction in inert organic solvents, such as toluene and acetone, can produce crystallized TiO
2 nanoparticles [
199,
200]. In Praserthdam’s research, nanocrystalline TiO
2 was prepared by toluene using the solvothermal method. In their synthesis, titanium n-butoxide was used as the starting material and was suspended in toluene in a test tube, which was then placed in a 300 mL autoclave. The autoclave was purged by nitrogen, after which it was heated up to 300°C and was held at 300°C for 2 hr before cooling down to room temperature. The obtained TiO
2 was washed by methanol several times and quenched in air at 77 K. The as-prepared anatase powders were of spherical shape with a size of 8–15 nm. The quenching in their research contributed to the formation of Ti
3+ surface defects due to the thermal shock effect and promoted the photocatalytic ethylene decomposition ability of TiO
2.
Figure 10 (f) displays a TEM image of the TiO
2 as prepared [
199].
The preparation of TiO
2 microspheres is usually promoted by the addition of surfactants, while [
201,
202,
203,
204] prepared TiO
2 microspheres (
Figure 11) with the solvothermal method in acetone without surfactants. Acetone itself might serve as a shape controller, but it hasn’t been confirmed yet. In their preparation, a certain volume of TiCl
4 was added dropwise to acetone under vigorous stirring at 0°C. The concentration of TiCl
4 was adjusted to 0.3 mol/L. This mixed solution was transferred into an autoclave afterwards and was maintained at 120°C for 12 hr in an oven for solvothermal treatment. After reaction, the mixture was cooled to room temperature naturally. The resulting precipitates were filtered and thoroughly washed with excessive acetone and then dried at 120°C for 12 hr under vacuum. The precipitate was calcined under 500°C for 5 hr in air. The diameter of the layered microspheric anatase TiO
2 was about 3 μm with each layer around 300 [
204]. On the contrary, the solvothermal preparation of TiO
2 in Chen’s work also used acetone as a solvent, but TiO
2 nanoparticles were formed instead of microspheres.
3.3. Sol-Gel Method of TiO2 Production
Nanoscale TiO
2 particles are synthesized by sol-gel method using hydrolysis of titanium precursors [
205]. Titanium alkoxide or titanium tetrachloride are used as precursor. At the first stage of sol-gel process hydrolysis of titanium (IV) precursor is carried out with subsequent polycondensation, which leads to the formation of colloidal solution -sol of hydroxide particles, the size of which does not exceed several tens of nanometers. Low water content (low level of hydrolysis) and excess of titanium alkoxide in the reaction mixture contribute to the development of Ti-O-Ti bond chains. Chain formation leads to the formation of a three-dimensional polymer skeleton with a near-ordered degree. The high rate of hydrolysis promotes the formation of Ti(OH)
4, which interrupts the development of the Ti-O-Ti skeleton. The presence of a large number of Ti-OH groups and insufficient development of the three-dimensional polymer skeleton leads to loose packing of particles [
205,
206,
207].
The use of low processing temperatures (<100°C) and molecular level composition uniformity makes the sol-gel technique very promising for the synthesis and manufacture of inorganic and organic-inorganic hybrid nanomaterials, as compared to the previously stated approaches [
207]. The sol-gel process makes it simple to adjust the size and form of the particles. The sol-gel technique, which is commonly used to create TiO2 materials, generates fine, spherical powders of uniform size and typically starts with an acid-catalyzed stage involving titanium (IV) alkoxides [
208,
209]. The ability to mold the resultant material into desired shapes, such as fiber, film, and monodispersed powder, is one of the most appealing aspects of the sol-gel process.
Figure 12 illustrates how a sol-gel technique, as proposed by Mehrotra and Singh [
208], applies a number of variables and phases to regulate the final morphology.
Metal oxides and metal chlorides are common precursors. The compound M─O─R, where M is metal, O is oxygen, and R is an alkyl group, is a metal alkoxide. The M─O bond is polarised, making it vulnerable to nucleophilic attack. Hydrolysis is a process in which an alkoxide in the presence of water undergoes a nucleophilic substitution reaction in which hydroxyl groups from water replace alkoxy groups (OR). Condensation is a process in which metal hydroxide groups combine with each other to form a hydrated metal-oxide network in which tiny nuclei are eventually formed.
The reactivity of metal alkoxides used in the sol-gel process needs to be controlled to obtain sols and gels with desired properties. This can be done by adding chelating ligands such as β-diketones, carboxylic acids or other complex ligands, or by using modifiers. To improve the control of the hydrolysis-condensation process in sol-gel fabrication, modifiers react with alkoxides to form new molecular precursors. These new precursors reduce functionality and reactivity, inhibit condensation and induce smaller species. The potential of acetylacetone to improve sol-gel processing of metal alkoxides was studied by Livage et al. in 1988 [
209]. Susceptibility to hydrolysis is reduced when modifiers change the number of M-OR bonds available for hydrolysis. Since β-diketones are surface capping reagents and polymerisation fixatives, their use reduces nuclearity and results in the formation of fine particles. Acetic acid and other carboxylate ligands mainly act as bridging chelating ligands.
The sol-gel method has several advantages, which include [
210]: (I) low-temperature preparation; (II) easy and efficient control of particle size, shape and properties; (III) improve the homogeneity of the raw materials; (IV) increase the purity of the starting material; and (IV) create the structure and properties of the material by appropriate choice of precursor.
3.4. Sonochemical and Microwave-Assisted Methods of TiO2 Synthesis
The sonochemical strategy has been applied to deliver exceptionally photoactive TiO
2 nanoparticles by the hydrolysis of titanium tetraisopropoxide (TTIP) in unadulterated water or in an ethanol/water blend under ultrasonic radiation [
211]. Acoustic cavitation, or the formation, growth, and collapse of bubbles within a liquid medium, is the basis for sonochemistry. Heat (~5000 K) and high tensions (~1000 atm) are delivered by cavitational breakdown [
212]. Microwaves, which are electromagnetic waves with wavelengths ranging from 1 mm to 1 m and frequencies ranging from 0.3 to 300 GHz, are utilized in microwave-assisted techniques. As indicated by Zhu and Chen [
213], microwave warming includes two principal systems to be specific dipolar polarization and ionic conduction. Any materials that contain versatile electric charges, for example, polar particles or leading particles are by and large intensity by microwaves. When polar molecules attempt to align themselves with the rapidly shifting alternating electric field in the microwave, they generate heat through rotation, friction, and collision. Assuming particles are available in arrangement, they will travel through the arrangement and continually taking an alternate route in view of the direction of the electric field bringing about nearby temperature climb because of erosion and impact [
214].
Microwave warming is as an elective intensity hotspot for quick warming with more limited response time and higher response rate, selectivity and yield when contrasted with the regular warming techniques [
213]. Pulsed microwave heating and continuous microwave heating are the two types of microwave heating. In 1995, Jacob et al. came up with two models for how microwaves increase reaction rates. The main instrument expects to be that, albeit the response time is vigorously abbreviated for a microwave-instigated response, the energy or component of the compound response isn’t modified suggesting that the upgrade of the response rate is because of the warm warming impact [
215]. The second proposed system makes a supposition that there are “nonthermal microwave impacts” notwithstanding the warm impacts thus the impacts of microwave light in substance responses are because of both warm impacts and nonthermal impacts [
216]. The nonthermal impacts are because of direct communication of microwaves with specific particles in the response medium.
Microwave radiations can likewise be applied to deliver different TiO
2 nanomaterials [
217]. This method has the advantage of rapid heat transfer and selective heating for industrial processing. This procedure gives uniform conveyance of energy inside the example, better reproducibility and amazing control of exploratory boundaries. When compared to the several hours required for the conventional methods of forced hydrolysis at high temperatures (195°C), the colloidal TiO
2 nanoparticles can be prepared in a short amount of time (within 5–60 minutes) [
218]. TiO
2 nanotubes which are unassuming and multi-walled with breadths of 8-12 nm and lengths somewhere in the range of 200 and 1000 nm were additionally pre-arranged utilizing this technique [
219]. TiO
2 nanoparticles in the anatase stage were ready by Baldassari et al. [
220] utilizing microwave-helped hydrolysis of titanium tetrachloride (TiCl
4) in a weaken acidic watery medium. Under microwave-hydrothermal conditions, they discovered that the product nearly crystallized within 30 minutes. Because the sulfate prevented brookite from crystallizing, they used H
2SO
4 as the acid to produce a pure anatase phase. In another review, they likewise pre-arranged TiO
2 nanoparticles in the rutile stage from TiCl
4 by a microwave-aqueous cycle at various temperatures somewhere in the range of 100 and 160°C for 5-120 min [
220]. The morphology and size of the subsequent nanoparticles can be fluctuated by changing the hour of response, microwave power and reactant fixation.
In [
221], titanium slags were converted into rutile TiO
2 powder by microwave activation (
Figure 13). Then, the effects of the Na
2CO
3 additive on the calcined product’s surface functional groups, crystallinity, phase transformation, and surface microstructure were examined. The following is the makeup of titanium slag: 9.72% Fe, 5.87% Al
2O
3, 5.23% SiO
2, 1.23% MgO, 1.81% CaO, 75.34% TiO
2, and other trace elements, such S and P.Using a planetary ball mill (model: QM-3SP4), the material was first processed into a powder for 180 minutes in order to improve the specific surface area of the slag. The obtained titanium slag sample, weighing 100 g, was then equally divided into five pieces, each of which was mixed with Na
2CO
3 in an agate mortar for ten minutes.
For the mixes, the mass ratios of Na2CO3 to titanium slag were 0.2, 0.3, 0.4, 0.5, and 0.6, respectively. After that, the mixture was put in a corundum crucible and heated to 850°C for 30 minutes using a 1 kW microwave heating power in a microwave box reactor. With the use of a magnetic stirrer, 10 g of calcined slag was leached for 4 hours at 92–95°C using 20% HCl (mass ratio of liquid/solid: 4:1). Following three rounds of water washing, the residue from the leaching process was collected and put in a corundum crucible for high-temperature annealing in a microwave box reactor set at 900°C for 60 minutes with a 1 kW microwave heating power. The calcined product was then cooled and put to use in an analysis. The findings demonstrated that the ideal mass ratio of Na2CO3 was 0.4, at which point the average size of the crystallites was 43.5 nm and the rutile TiO2 crystallinity attained its maximum value of 99.21 percent.
3.7. Green Synthesis of TiO2
An ecologically benign substitute for the chemical method of creating nanomaterials is the “green” synthesis of TiO
2 functional materials. Using biological agents including bacteria, fungus, actinomycetes, yeast, and plants, the biological technique offers a multitude of resources for the production of nanoparticles [
231,
232]. The pace at which metal ions are reduced with the aid of biological agents is substantially quicker than it is because of the surrounding pressure and temperature. Significant advancements in “green” synthesis techniques for the creation of many nanoparticles have resulted from the extraction of TiO
2 from plant extracts [
233]. Green tea extract was used by the authors in study [
234] to create mesoporous TiO
2 nanoparticles using the sol-gel technique. A mixture of 9 ml of titanium isopropoxide and 60 ml of isopropanol was stirred continuously with a magnetic stirrer at room temperature for 1 hour. Then green tea extract was added in various ratios (0.5, 1, and 1.5 g in 30 ml of distilled water) and stirred slowly for 3 hours in order to obtain a colloidal solution. It was found that the pH of the solution was 6.0 during the TiO
2 nanoparticles synthesis. The resulting sol was kept at rest for 10 hours to obtain a gel. Then the gel was filtered, dried at 110°C for 3 hours and calcined at 500°C for 10 hours. The calcined samples were designated as NTG0.5, NTG1, and NTG1.5, which corresponded to mass ratios in samples 1:0.06, 1:0.12, and 1:0.18 TiO
2:GTE (extract), respectively. TiO
2 nanoparticles prepared without extract were monitored and were designated as NT. The mild, non-toxic, and inexpensive green tea extract, which hass active organic components, limited agglomeration, and promoted the growth of TiO
2 nanoparticles. In work [
235], the authors have synthesized titanium dioxide nanoparticles by an improved hydrothermal method using Morinda citrifolia leaf extract. 50 ml of M. citrifolia leaf extract was added to 0.1 M TiCl
4 solution. The solution was transferred to a 100 ml stainless steel autoclave at 120 °C for 8 hours and then cooled to room temperature. A white suspension was obtained, which was centrifuged at 5,000 rpm/min for 10 minutes to remove unreacted chemicals. The resulting suspension was filtered and washed several times with deionized water and ethanol. The filtered suspension was dried in an oven at 100°C for 5 hours. Titanium hydroxide was calcined at 400°C for 4 hours in a muffle furnace, resulting in quasi-microspheres of TiO
2 nanoparticles. X-ray diffraction patterns showed the presence of rutile phase TiO
2 and confirmed an average crystallites size of 10 nm. The authors [
236] synthesized TiO
2 nanoparticles by the hydrothermal method using Aloe Veragel for use as a photocatalyst in the degradation of picric acid. Aloe Verawas peeled and the gel was washed seven times under running water. 10 ml of the gel was added to 100 ml of deionized water and stirred for 1 hour. To this aqueous solution was added dropwise 0.1M titanium (IV) isopropoxide. The reaction mixture was stirred continuously for one hour at 20°C. The solution was kept in an autoclave at a temperature of 180 °C for 4 hours. Then the solution was heated on a hot plate at a temperature of 80 °C. The resulting product was ground and calcined in a muffle furnace at a temperature of 500°C for 5 hours. The size of the synthesized TiO2 nanoparticles ranged from 6 to 13 nm. In a study [
237], titanium dioxide nanoparticles were efficiently synthesized using aqueous extracts of Parthenium hysterophorus leaves by microwave irradiation. The collected leaves were washed with distilled water to remove dust particles and contaminants. About 20 g of leaves were weighed and crushed into small pieces with a mortar and pestle. The samples were added to 100 ml of distilled water and boiled for 10 minutes at 60 °C in a microwave oven. After boiling, the extract left to cool at room temperature.
3.8. Electrodeposition and Ionic Liquid-Assisted Methods
Electrodeposition is a plating process in which ions in a solution migrate under the influence of an electric field (electrophoresis) and are deposited onto an electrode. In a normal process, components containing one or more dissolved metal salts are immersed in electrolytes, and the metallic ions are attracted to the cathode to be deposited. Electrodeposition is easily controlled, can produce tight coating with uniform thickness and is able to coat complex fabricated objects [
238,
239]. Electrodeposition of TiO
2 nanoparticles onto multiwalled carbon nanotube arrays was conducted in an electrolyte consisting of 3 mol/L KCl solution, 10 mmol/L H
2O
2, and 10 mmol/L Ti(SO
4)
2. Multiwalled carbon nanotube arrays were used as the working electrode, an Ag/AgCl electrode as the reference electrode, and Pt as the counter electrode. The working potential was − 0.10 V and the deposition time was 30 min [
240].
Ionic liquids refer to salts in the liquid state. Actually, when heated to a high temperature, almost all salts can become ionic liquids. The ionic liquid referred to here is a kind of salt that is in liquid states at low temperatures (< 100°C) or even room temperature. The ion size of these liquids is usually large and poorly coordinated, resulting in a low bounding force and a loose structure, thus forming a liquid rather than a solid at a relatively low temperature. Ionic liquids have many merits, they exhibit excellent thermal stability, powerful solubility, good electrical conducting ability, low viscosity, and have almost no vapor pressure. Usually, the low temperature ionic liquids have at least an organic cation (such as methylimidazolium and pyridinium ions) and an inorganic canion (such as halide, tetrafluoro-borate, and hexafluoro-phosphoric ions). Despite the fact that the ionic liquid is usually poisonous, it has found its way into the research of pharmaceuticals, gas treatment, cellulose processing, solar thermal energy, etc., and is recently used to modify the preparation process of TiO
2 [
241,
242,
243].
Different binary ionic liquids were applied to synthesize TiO
2 hollow spheres. It was found out that the shape, size and crystallinity were different by varying the binary ionic liquid composition, which is a result of different interface interactions. In a typical process, 3.6 mL of binary ionic liquids were mixed with 0.4 mL of anhydrous toluene solution containing 0.2 mol/L titanium isopropoxide. 6 mL methanol was then added and centrifuged. The final mesoporous TiO
2 was obtained after filtration and calcination at 500°C. They tested all the binary mixtures of six different ionic liquids, which were 1-butyl-3-methylimidazolium hexafluorophosphate ([Bmim][PF
6]), 1-hexyl-3-methylimidazolium hexafluorophosphate ([Hmim][PF
6]), 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim][BF
4]), 1-octyl-3-methylimidazolium hexafluorophosphate ([Omim][PF
6]), 1-hexyl-3-methylimidazolium tetrafluoroborate ([Hmim][BF
4]), and 1-octyl-3-methylimidazolium tetrafluoroborate ([Omim][BF
4]) and discovered that [Bmim][BF
4] + [Omim][PF
6] mixtures were the most effective group and would make anatase TiO
2 with a surface area of about 100 m
2/g after calcination [
244]. Ionic liquid-assisted hydrothermal synthesis was reported by using 3-carboxymethyl-1-methylimidazoliurn bisulfate ([CMIM][HSO
4]), titanium isopropoxide, concentrated HCl and H
2O to fabricate rutile TiO
2 nanorod films. In a typical process, 0.5 mL titanium isopropoxide was added to a mixed solution of DI water, concentrated HCl and 0.5 mL [CMIM][HSO
4]. The resulting transparent mixture was then transferred to a hydrothermal autoclave with a piece of glass immersed in the solution. The autoclave was heated to 180°C for 3 hr. When ionic liquids weren’t used in the process, the diameter of TiO
2 was about 250 nm, but when [CMIM][HSO
4] was used, the diameter of TiO
2 nanorods decreased to 62 nm. This was not only due to the effective prevention of the gathering of the nanoparticles by the surfactant-acted ionic liquids, but also because of the extended hydrogen bonding and ionic strength which favored the formation of small crystals.
Figure 14 (a) is an FESEM image of TiO
2 nanorods in Mali’s work [
245].
3.9. Synthesis of Nanoscale and Thin Film Structures of TiO2
It is obvious that the use of nanosized TiO
2 particles leads to a significant increase in its photocatalytic activity. An undoubted advantage of nanoparticles compared to microparticles is the greater probability of charge release on the catalyst surface. Due to the fact that the penetration depth of UV light of TiO
2 particles is limited (~100 nm), only the outer surface is active [
246].
Figure 14 (b) shows a diagram of light absorption by TiO
2 nano- and microparticles.
As can be seen from
Figure 14 (b), reducing the size of the photocatalyst particles to nanoscale values promotes light absorption by the entire volume of particles. In this regard, the use of TiO
2 in heterogeneous photocatalysis processes is associated with the need to obtain nanosized particles. To date, TiO
2 nanoparticles are obtained with various morphologies, mainly nanotubes, nanowires, nanorods and mesoporous structures [
247].
Thus, the properties and applications of titanium oxide (TiO
2) nanostructures largely depend on the particle size, structure, effective surface area, and surface properties. Since these properties are in turn influenced with synthesis methods, in this section we will have an overview of different methods of synthesis of nanoparticles Titanium oxide (TiO
2) thin films [
248,
249,
250]. Chemical reactions for the synthesis of substances can take place in gaseous, liquid or solid forms. The rate of penetration of reactants in the gas or liquid phase is several times faster than in the solid phase. Therefore, the synthesis methods for titanium dioxide-based nanostructures are mainly divided into liquid-phase synthesis and vapour-phase synthesis [
251,
252,
253]. Moreover, nanostructured TiO
2 particles can be obtained by oxidation of metallic titanium using various chemical oxidants [
254]. TiO
2 nanorods were obtained by this method (by oxidation of a titanium metal plate with hydrogen peroxide). In [
255], the authors showed that anodic oxidation of titanium in a fluoride-containing electrolyte allows obtaining nanostructured coatings consisting of TiO
2 tubes, the properties of which can be controlled by varying the oxidation conditions. However, the authors showed that such coatings with a nanotube length of more than 1 μm have low adhesion, and their application is still very limited. Then the authors in [
256] proved that poor adhesion is due to the low packing density of the nanotubes (
Figure 14 (c, d)).
According to modern concepts, the growth model of TiO
2 nanotubes during potentiostatic anodizing consists of several stages, and the reactions responsible for the formation of porous aluminum oxide and TiO
2 nanotubes [
254,
255,
256,
257] are identical. Despite the similarity of the processes occurring during the anodization of titanium and aluminum, the morphology of the resulting oxides differs greatly. For example, during the anodization of aluminum, a mesoporous structure is formed, whereas during the anodization of titanium, both mesoporous and nanotubular structures can be obtained [
257]. At the same time, the use of porous carriers active only under the influence of UV light, inside the pores of which there are particles activated by light with a wavelength of 400–700 nm, seems very promising. Such an approach will make it possible to use both visible and UV radiation in photocatalytic processes.