Next Article in Journal
A Patent-Pending Ointment Containing Extracts of Five Different Plants Showed Antinociceptive and Anti-Inflammatory Mechanisms in Preclinical Studies
Previous Article in Journal
Synthesis and Biological Evaluation of Novel Cationic Rhenium and Technetium-99m Complexes Bearing Quinazoline Derivative for Epidermal Growth Factor Receptor Targeting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Titanium Dioxide Nanomaterials: Progress in Synthesis and Application in Drug Delivery

1
State Key Laboratory of Component-Based Chinese Medicine, Tianjin University of Traditional Chinese Medicine, Tianjin 301617, China
2
School of Traditional Chinese Medicine, Tianjin University of Traditional Chinese Medicine, Tianjin 301617, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2024, 16(9), 1214; https://doi.org/10.3390/pharmaceutics16091214
Submission received: 19 August 2024 / Revised: 13 September 2024 / Accepted: 13 September 2024 / Published: 16 September 2024

Abstract

:
Background: Recent developments in nanotechnology have provided efficient and promising methods for the treatment of diseases to achieve better therapeutic results and lower side effects. Titanium dioxide (TiO2) nanomaterials are emerging inorganic nanomaterials with excellent properties such as low toxicity and easy functionalization. TiO2 with special nanostructures can be used as delivery vehicles for drugs, genes and antigens for various therapeutic options. The exploration of TiO2-based drug delivery systems shows great promise for translating nanotechnology into clinical applications; Methods: Comprehensive data on titanium dioxide were collected from reputable online databases including PubMed, GreenMedical, Web of Science, Google Scholar, China National Knowledge Infrastructure Database, and National Intellectual Property Administration; Results: In this review, we discuss the synthesis pathways and functionalization strategies of TiO2. Recent advances of TiO2 as a drug delivery system, including sustained and controlled drug release delivery systems were introduced. Rigorous long-term systematic toxicity assessment is an extremely critical step in application to the clinic, and toxicity is still a problem that needs to be closely monitored; Conclusions: Despite the great progress made in TiO2-based smart systems, there is still a great potential for development. Future research may focus on developing dual-reaction delivery systems and single-reaction delivery systems like redox and enzyme reactions. Undertaking thorough in vivo investigations is necessary prior to initiating human clinical trials. The high versatility of these smart drug delivery systems will drive the development of novel nanomedicines for personalized treatment and diagnosis of many diseases with poor prognosis.

1. Introduction

The drug delivery system (DDS) is an important area of biomedicine. Over the past decades, nanotechnology has played an increasingly important role in drug delivery [1,2]. Conventional DDSs (tablets, capsules, ointments, etc.) suffer from poor bioavailability, fluctuating plasma drug concentrations, and make it more difficult to control drug release [3,4]. Nanomaterials have excellent properties to meet the needs of smart DDSs, which are capable of aggregating at targeted sites such as tumor tissues, reducing systemic exposure and drug clearance through enhanced permeability and retention effects, and minimizing side effects [5,6,7].
Inorganic nanoparticles are one of the mainstream directions in nano-drug delivery research [8,9,10]. Inorganic nanomaterials, such as silica [11,12], titanium dioxide [13,14], manganese oxide [15], graphene [16], etc., are ideal for drug delivery due to their ease of preparation, high drug loading rate, excellent physicochemical properties, biocompatibility, and flexibility of surface modification (Table 1). In addition, inorganic nanomaterials have a wide range of applications in biochemical sensing [17], bioimaging [18], chemical catalysis [19], and other fields. In recent years, TiO2 nanomaterials and its composite materials have shown great potential in realizing targeted drug delivery, slow drug release, and improving the solubility and bioavailability of intolerable or insoluble drugs, and have also shown good application prospects in enhancing the therapeutic effects of diseases [20,21,22].
TiO2 nanomaterials with different structures exhibit varying characteristics [23,24,25]. To achieve high effectiveness in drug delivery applications, TiO2 with a wide surface area and porosity is typically required [26]. Various preparation methods are constantly being tried to obtain TiO2 with different mesoporous structures, pore wall parameters, morphologies and compositions. This review will focus on the recent preparation and functionalization developments of TiO2 and the last advances on their application in drug delivery technologies.

2. Synthesis of TiO2 Nanomaterials

Over the past few years, researchers have worked to improve the synthesis of TiO2 nanomaterials due to the demand for certain features and the shortcomings of current approaches. Among the several synthetic routes, the commonly used methods include the sol-gel method, hydrothermal synthesis, template method, gas-state method, and so on. There are significant differences in the size, shape, morphology, composition, mesoporous structure, specific surface area, and crystal structure of TiO2 prepared by different methods, which play a crucial role in the biomedical applications of nanoparticles. In this section, synthetic methods for TiO2 nanomaterials are summarized and compared (Table 2).

2.1. Sol-Gel Method

In the 1960s, the sol-gel method was used to prepare materials with various shapes and porous structures. The sol-gel procedure included the process of hydrolysis and polycondensation, during which process M-OH-M or M-O-M bridges were established between the atoms M of the precursor molecules, resulting in oxides or hydroxides. Specifically, for TiO2, titanium alkoxides (such as titanium (IV) isopropoxide, titanium butoxide), alcohol, and acid/water were introduced into the reaction system. After stirring for several hours, densely cross-linked three-dimensional structures were built and terminated as TiO2 gel. The reactions’ mechanism was as follows [27,28]:
The hydrolysis reactions:
Ti OR n + H 2 O Ti OH OR n 1 + ROH  
Ti OH OR n 1 + H 2 O Ti OH 2 OR n 2 + ROH .
The reactions continuously go on: ⋯ Ti OH n .
The polycondensation reactions:
- Ti - OH   +   HO - Ti -     - Ti - O - Ti -   +   H 2 O
- Ti - OR + HO - Ti -     - Ti - O - Ti - + ROH .  
After the sol-gel process, aging, drying and annealing were needed to create the final TiO2 nanoparticles [29]. Figure 1 shows a schematic diagram of the steps involved in the sol-gel process.
Sol-gel synthesis is one of the simplest, fastest, and economically less expensive methods, and has advantages like low processing temperature, homogeneity of the produced material, and formation of the complex structures or composite materials [30,31]. Gwanhee Park et al. added water in the form of steam in the reaction system, and the small droplets were dispersed to collide with the titanium source, which triggered hydrolysis reaction. The additional collision between the droplets was suppressed by excess anhydrous ethanol, which resulted in the formation of smaller nanoparticles than the ones in the conventional sol-gel method [32]. Jaroenworaluck et al. prepared anatase TiO2 nanoparticles up to 10 nm in size by mixing tetraisopropylorthotitanate, methanol, and ethanol in different molar ratios [33]. The excellent properties of the prepared nanoparticles were crucial in addition to their good morphology. Among the different crystalline phases of TiO2 nanoparticles, anatase is the key phase for photocatalysis [24,34]. Calabrese et al. developed reaction parameters based on the sol-gel method to promote local alignment of anatase particles, which created TiO2 with a tunable crystalline phase and excellent photocatalytic activity [35].
Apart from pure TiO2 nanoparticles, the researchers also synthesized TiO2 doped with different compositions using the sol-gel method, which greatly improved their properties for particular applications, such as photocatalysis, antimicrobial, and photovoltaic power generation [36,37,38]. Yadav et al. prepared Cu-, Fe-, Ni-, Cr-, and Co-doped TiO2 and undoped TiO2 nanostructures and demonstrated photocatalytic dye degradation and antibacterial application. Similarly, Shahina Kader et al. synthesized TiO2, MoO3/TiO2, Ag/TiO2 and MoO3/Ag/TiO2 photocatalysts and applied them in an immobilized form on the borosilicate glass surface to construct a reactor to evaluate the performance of degrading MO dyes under UV treatment [39].

2.2. Hydrothermal Method

The hydrothermal method referred to preparing TiO2 by utilizing the highly identical and anisotropic growth mode of crystals, dissolving titanium salts, titanium containing organic compounds or titanium oxide compounds in water, and reacting with additives such as alkali and dispersant in a closed system at a high temperature and high pressure. Hydrothermal synthesis involved three key steps, crystal growth, crystal transformation, and phase equilibrium, culminating in fine-to-ultrafine crystals [40]. By regulating the ratio of different titanium sources and additives, reaction temperature and reaction time, the nucleation rate and particle size distribution could be controlled, thus obtaining TiO2 nanomaterials with different morphologies.
The hydrothermal method had been used to prepare TiO2 by many research groups after the first introduction by Kasuga et al., which could produce TiO2 with homogeneity, high purity, crystal symmetry, metastable compounds with unique properties, and narrow particle size distributions [41]. In a typical synthesis, Thapa R et al. carried out a hydrothermal reaction with titanium isopropoxide, acetylacetone, and urea at 150 °C for 18 h, ultimately yielding TiO2 nanomaterials of anatase type, which proved to be an efficient catalyst [42]. Liu et al. prepared TiO2 hollow nanospheres by dissolving Ti (SO4)2 and NH4F in deionized water and conducting a hydrothermal reaction at 160 °C for 6 h. The synthesized TiO2 hollow nanospheres have high dispersibility and large specific surface area. Subsequently, porous hollow TiO2 nanospheres could be prepared by adding corrosive additives in the hydrothermal process, and their specific surface area was 168 m2·g−1 with an average pore size of 12 nm [43].
To fulfill more application needs and improve individual flaws, hydrothermal was used to continually create TiO2 with a new morphology and varied component doping [44,45]. Sun et al. produced eggshell-like TiO2 hollow sphere nanoparticles using a one-step template-free method, and investigated their electrorheological properties at various electric field strengths [46]. The Ti3+ self-doped TiO2 nanoparticles with a fusiform-like morphology were prepared via a facile mixed solvothermal method in a mixture of water and triethylamine [47]. Rani et al. investigated the effect of anionic and nonionic surfactants on the synthesis of core–shell Fe3O4@TiO2 nanocomposites by a hydrothermal process [48].
In recent years, the microwave-assisted hydrothermal method has been generally accepted by researchers. It is an economical, fast, green and effective synthetic strategy to prepare efficient nanomaterials with controllable particle size. Majid et al. studied Fe-doped TiO2 nanomaterials by microwave hydrothermal synthesis, which improved the band gap limitation of TiO2 in applications [49]. In addition, the microwave hydrothermal method was also characterized by a fast-heating rate, sensitive reaction, uniform heating, etc. [50]. The crystal size, morphology and agglomeration could be controlled by adjusting the ratio of precursor materials, pH of the reaction system, reaction time and temperature. Sahu, K et al. prepared Cu-doped TiO2 nanomaterials and investigated the optimum Cu-doping ratio. The optics of samples prepared after optimum Cu doping and electrical properties were substantially improved [51].

2.3. Template Method

The template method is often used to produce TiO2 with mesoporous or hollow structures, as the morphology of the pores can be well controlled. Based on the properties of the template employed in the reaction, the template method can be generally divided into two categories: soft template method and hard template method [52].

2.3.1. Hard Template Method

The hard template method uses pre-synthesized organic or inorganic templates with specific shapes (morphology, surface curvature), which serve as molds for the replication of nanoporous/mesoporous TiO2 and do not involve any significant chemical interactions with the titanium precursors. The morphology of the resulting materials is pre-determined by the templates, which have well-defined nanostructures [53]. Silicon dioxide (SiO2) is widely utilized as a template; however, this has resulted in mesoporous structures that are similar to each other and unable to satisfy the increasingly demanding applications of functional replica materials. Evidently, the creation of a new template is the key to resolving this problem. Based on the successfully prepared hierarchical mesoporous SBA-15 silica microspheres with 2D and 3D porous structures [54], Wang et al. prepared hierarchical mesoporous TiO2 microspheres (HMM-TiO2-MSs) by the nanocasting method and optimized the process. The obtained HMM-TiO2-MSs reached a specific surface area and pore volume of 194 m2 g−1 and 0.68 cm3 g−1, respectively, which were more than twice those of the conventional SBA-15 templates [55]. In addition, Li et al. prepared photocatalysts with TiO2 and g-C3N4 multilayered inline-connected structures using natural montmorillonite as a hard template by two intercalation methods for the first time [56]. Simultaneously, using resorcinol/formaldehyde (RF) polymer resin as a hard template, Wang et al. synthesized hollow TiO2 spheres through the hydrolysis and carbonation of multifunctional kinetically controlled coatings [57]. The use of RF as a template reduced damage to the TiO2 structure during etching and required a few organic reagents, which was favorable in the context of green science. Furthermore, size-controllable hollow mesoporous TiO2 spheres with greater photocatalytic activity than TiO2 nanoparticles were created using carbon spheres as hard templates [58].

2.3.2. Soft Template Method

During the soft-templating process, the precursors and surfactant templates can work together to assemble an ordered mesoporous structure, which is driven by the spontaneous trend of interface energy reduction [59]. In such cases, the obtained pore structures are determined by templating agents’ molecular properties as well as the synthetic conditions, such as reactant ratios, concentration, solvents, and temperature. The soft template method can be synthesized in both aqueous and non-aqueous solutions. Since titanium precursors are highly reactive and sensitive to moisture, hydrolysis is usually too fast for the aqueous route to be controlled, making polymerization very difficult. In contrast, the non-aqueous solution synthesis method, which is often referred to as evaporation-induced self-assembly (EISA), can effectively slow down the hydrolysis and condensation of titanium precursors. Organic molecular templates, also known as structural directing agents, are important for mesoporous structures due to their composition and characteristics. To date, cationic surfactants (e.g., quaternary ammonium salts), anionic surfactants (e.g., carboxylates, sulfates, and phosphates), and non-ionic surfactants (e.g., Pluronic P123 and F127) are the most often utilized templates [60]. Hung et al. prepared ordered hexagonal mesoporous TiO2 by the EISA method using Pluronic P123 and tetrabutyl orthotitanate (Ti (OBun)4, TBOT) as the templating agent and the titanium source, respectively. They elucidated the effects of surfactant concentrations on the pore arrangement, pore size, specific surface area and structure of mesoporous TiO2 by the EISA method [61]. The variation in block copolymer templates can lead to changes in the crystal structure of synthesized nanostructures, resulting in changes in activity [62].
Although the soft template method provides excellent control over mesoscopic structures, it lacks control over atomic-scale structures. Due to the usual decomposition of surfactants below 300 °C, mesoporous metal oxides (MMOs) are often destroyed during the crystallization process. As a result, MMOs is amorphous or semi-crystalline and has low thermal stability [63]. Recently, Xiong et al. established a universal polymer-oriented evaporation-induced self-assembly strategy for synthesizing MMOs with high crystallinity. In acidic hydration environments, metal chlorides are used as precursors for metal oxides, while the cationic polymers are used as porous precursors. This method solved the drawback of the classic sol-gel-based self-assembly route using amphiphilic block copolymers as templates, which was confined to amorphous or semi-crystalline MMOs with low thermal stability [64].

2.4. Gas-State Method

The gaseous process turns basic materials into a gaseous state via sublimation, evaporation, and degradation before growing crystals under supersaturated steam through condensation and crystallization. It is commonly used to deposit TiO2 nanoparticles on various substrates or create nanofilms [65,66,67]. It is mainly divided into physical vapor deposition (PVD) and chemical vapor deposition (CVD). PVD is a method of heating raw materials by physical means, such as evaporation, ionization or sputtering, and further condensing them to form a solid material, usually without involving chemical reactions [68,69,70]. The TiO2 nanomaterials prepared by this method have the advantages of high purity, uniform distribution, small particle size and good dispersion, and it is easy to control the thickness of the film. However, the disadvantage is that the prepared materials need to be deposited under vacuum conditions, the equipment is more expensive, and the operating costs are high.
Unlike PVD, CVD forms a solid layer with conformal and fine coverage on a substrate by a chemical reaction of gases. Othman S H et al. prepared TiO2 by CVD in the deposition temperature range of 300–700 °C. The crystalline structure of TiO2 could be controlled by adjusting the deposition temperature, and products with high crystallinity, good homogeneity and high photocatalytic efficiency were obtained [71]. Many attempts have been made to directly prepare or deposit TiO2 on different substrates by the CVD method [72,73,74], and it has been confirmed that the problem of aggregation, large bandgap and rapid recombination of photo-generated electron–hole pairs can be well solved [75,76,77,78].

2.5. Solid-State Method

Solid-state reactions are generally considered to be those in which the solid participates directly in the chemistry and undergoes a chemical change, or at least, a process occurring inside or outside the solid that has a controlling effect. TiO2 is typically obtained by thoroughly mixing, grinding and calcining the formulated titanium or titanium oxide. TiO2/γ-Fe2O3 nanocomposites were synthesized as nano-adsorbent through a ball milling process by Mercyrani et al. Variations in grinding time and ball-to-powder ratio allowed for the control of the properties of the composites. Characterization such as XRD showed that spherical anatase TiO2 and cubic γ-Fe2O3 were predominantly formed under the established method [79]. Cao et al. reported a facile solid-state method to synthesize anatase TiO2 single crystals with 35% exposure of (001) facets and high purity. The shape of the obtained single crystals of TiO2 is characteristic of truncated bipyramids, and their sizes varied from 2–3 μm to 50–200 nm. They believed that with the application of this synthetic technique, TiO2 single-crystal materials may now be prepared on a massive scale [80]. In terms of the solid-state method, it is easy to operate and yields a more considerable amount of the desired product with a short processing time at ambient conditions; however, there are certain disadvantages, including easy introduction of impurities, incomplete morphology and inhomogeneous particle size. Hence, it is more suitable for large-scale industrial production [81,82].

2.6. Green Synthesis Method

Although the sol-gel, hydrothermal synthesis, template, and electrochemical methods are the most commonly used to prepare TiO2 nanomaterials, they all require high temperatures, high pressures, and harmful chemicals to complete, limiting their synthesis and potential medical applications [83]. Green synthesis, a naturally adaptable, environmentally friendly and cost-effective technique for large-scale nanoparticles synthesis, has attracted great interest in synthesizing TiO2 in a green and sustainable manner [84,85]. Figure 2 shows a schematic diagram of the preparation process of nanoparticles via the green synthesis method.
Biologically active components present in organisms such as plants and bacteria promote bio-reduction and capping processes. A range of natural reducing agents, including proteins, enzymes, and phytochemicals, are involved in the synthesis of TiO2 [83]. Kaur, H. et al. synthesized TiO2 with a mesoporous nature using Titanium (IV)-iso-propoxide as a precursor and Lagenaria siceraria leaf extract as a reducing and capping agent. They also compared the optical, morphological, structural and photocatalytic properties of green and chemically synthesized TiO2. The photocatalytic decolorization of RG-19 dye was successfully enabled by green synthesized TiO2 using L. siceraria aqueous leaf extract [86]. Furthermore, the immobilization of silver nanoparticles (Ag NPs) on the TiO2 surface was carried out using Carpobrotus acinaciformis extract as a reducing and stabilizing agent. The modified TiO2 was characterized by FT-IR spectroscopy, X-ray diffraction, field emission scanning electron microscopy and energy dispersive X-ray spectroscopy, and it was confirmed that the diameter of Ag NPs on the TiO2 surface was in the range of 20-50 nm. Photocatalytic degradation experiments of methyl orange and Congo red proved that the green synthesized composite was a highly active and recyclable photocatalyst [87]. The various green sources have been used for TiO2 synthesis, as reported in Table 3, which gives a broader view of the green approach. The synthesis of inorganic nanoparticles using biological entities is of great interest due to their unusual optical, chemical, photo electrochemical and electronic properties [88]. Bacillus sphaericus was well adapted to heavy metals and could produce unusual inorganic nanoparticles by intracellular or extracellular mechanisms [89]. Suriyaraj S. P. et al. reported a one-pot method for producing TiO2 with pure anatase phase utilizing the extremophilic bacterium Bacillus licheniformis. Without calcination, the TiO2 particles had excellent crystalline characteristics [90]. Microbial-mediated synthesis of TiO2 nanoparticles was also carried out using Aeromonas hydrophila and Bacillus subtilis [91,92]. TiO2 produced by fungi, like bacteria, has safety limits. Nonpathogenic strains, on the other hand, will eliminate the threat and may be commercially exploited [93]. Table 4 shows the synthesis of TiO2 by various bacterial species.

2.7. Other Methods

In addition to innovations in traditional preparation methods, researchers are currently also exploring several new preparation methods. Recently, a study reported a new method based on a microwave–ultrasound-assisted method for the preparation of defective TiO2 and WO3/TiO2 nanoparticles. The method required only 5 min of microwave and ultrasonic irradiation to obtain TiO2 with a microcrystalline size of about 6 nm [109]. A study comparing the ability of ZnO-TiO2 nanocomposites synthesized by ultrasound-assisted and conventional methods to degrade methylene blue dye showed that ultrasound-synthesized ZnO-TiO2 nanocomposites exhibited superior photocatalytic activity and that the degradation process followed a secondary kinetic model [110]. Studies have also attempted the preparation of TiO2 nanoparticles by inter-cation redox reaction, anode synthesis by the boosting method, current hydrodynamic transport, and one-step catalyst-free gas-phase transport [111,112,113,114].

3. Functionalization of TiO2

The physiochemical characteristics of TiO2, like biocompatibility and functional modification, offer an excellent nanoplatform for drug delivery [45]. In particular, is it a perfect building block of novel bioinorganic hybrid nanoconjugates, which show enhanced features for imaging, photocatalytic therapy and photoactivated drug release [20,115,116]. The presence of a large number of free hydroxyl groups on the surface and inside the pores facilitates the functionalization of TiO2 [117]. Several techniques have been used to attach organic or inorganic ligands to the surface of TiO2 in order to provide advantages such as slow or controlled-release molecular delivery, improved biocompatibility, increased targeting activity, and much greater stability [14,22,118,119,120]. On the basis of surface functionalization, TiO2 offers the possibility to act as nanocarriers to selectively deliver drugs to specific sites, allowing drugs to exert their therapeutic effects with fewer side effects and improved therapeutic efficacy [121]. Figure 3 shows the functionalization of TiO2 nanoparticles.

3.1. Physical Surface Modification

Physical surface modification can be achieved by covering the nanoparticle surface with an ionic or polymeric surfactant. The chemical structure of surfactants contains hydrophilic and hydrophobic ends, and the hydrophilic end is adsorbed on the surface of TiO2 through electrostatic interactions or chemical bonding for the purpose of modification, which can reduce the inter-particle interactions and minimize the influence of interfacial forces, thus improving the stability of the particles [122].
In recent years, many researchers have reported surface modification of TiO2 using surfactants. Polyethylene glycol (PEG) has become a more commonly used modified surfactant because of its FDA-approved and inexpensive nature, improved hydrophilicity, improved dispersion of nanoparticles, and characteristics that allow for favorable pharmacokinetics and tissue distribution [123,124]. Furthermore, PEG can interfere with the contacts of the mononuclear phagocyte system and therefore reduce the clearance of nanocarriers [123]. A study employing atomic and molecular dynamics simulations investigated the grafting of PEG onto the surface of TiO2 in various solvents. They discovered that spherical brushes with high grafting density in water, but not in dichloromethane, follow the Daoud–Cotton classical scaling model for polymer volume fraction dependence with distance from the center, which has been developed to describe star-shaped polymer systems [125]. Sun et al. investigated the effect of average molecular weight of PEG on the stability and re-dispersibility of TiO2 dispersion. Through hydrogen bonding, PEG molecules were adsorbed on the surface of TiO2. As the average molecular weight of PEG molecules increased, the dispersion stability and re-dispersibility of TiO2 progressively improved due to their distinct adsorption conformations. However, PEG molecules with a high molecular weight can reduce nanoparticle dispersion stability [126]. In addition to PEG, chitosan [127], polyvinyl alcohol [128], polydopamine [129] and polyethyleneimine [130] are also commonly used for modification.

3.2. Chemical Surface Modification

Chemical surface modification is an effective method for developing surface properties of nanomaterials, and this technique is based on covalent bonding between the modifier and the particle surface [122]. Numerous studies have been conducted on the chemical modification of nanoparticle surfaces, and coupling agents such as organic functional silanes [131], polymers [132], and carboxylic acids [133,134] are frequently used in the modification process.
Silica coating is a typical chemical approach for the surface modification of TiO2, providing various benefits such as long-term stability, biocompatibility, and the hydrophilicity of silane-modified TiO2. The Stöber method is the most traditional method for silica modification of TiO2. This reaction typically uses anhydrous ethanol as the reaction medium, tetraethoxysilane (TEOS) as the silica source, and ammonia as an alkaline catalyst to control the hydrolysis of TEOS and the thickness of the silica layer, resulting in the formation of TiO2@SiO2 nanoparticles with regular morphology [14]. The modification techniques are continuously optimized, and the double template method, sol-gel method, and recycled template method have also become excellent methods to consider for TiO2@SiO2 [135,136,137].
The TEOS-modified TiO2 surface has Si-O bonds, which are derived from the silica shell. However, for other modifications with specific needs, it is necessary to introduce new groups on the surface of TiO2. Through the modification of different kinds of silanes, the surface of TiO2 is introduced to different groups, which is favorable for the next step of drug-controlled-release system. Shi et al. employed (3-aminopropyl) triethoxysilane (APTMS) to react with TiO2 in a toluene medium, converting the surface groups of TiO2 to amino groups. The TiO2 with amino surface was then modified to create a unique ROS-sensitive drug delivery system, which could effectively encapsulate Docetaxel (DTX) and demonstrate good acoustic kinetic chemotherapeutic effects in both in vivo and in vitro experiments [138].
(3-Mercaptopropyl) trimethoxysilane (MPTMS) is an excellent choice for introducing sulfhydryl groups to TiO2, and it is especially well-suited to the chemical attachment of metal nanoparticles to TiO2. Recently, it was reported that Ag NPs were modified on the surface of the TiO2, which was functionalized with MPTMS. In the presence of NaBH4, Ag NPs@MPTMS-TiO2 showed ultra-high catalytic activity for 4-nitrophenol reduction, with an apparent kinetic rate constant (kapp) of up to 394 × 10−3·s−1 [139]. Furthermore, TiO2 modified with MPTMS was able to overcome the tough problem of inadvertent agglomeration of gold nanoclusters deposited on its surface, providing an effective reference for photocatalytic systems [140].

4. Sustained Drug Delivery

Sustained drug delivery systems (SDDSs) have been intensively studied for decades with some success in the pharmaceutical field. These can prolong the therapeutic effect by slowly releasing therapeutic substances after administration, thereby reducing the frequency of administration and minimizing side effects. The ideas of TiO2 carriers used for SDDS have been categorized into three groups: pore structure and morphology, interaction force with drugs, and diffusion inhibition effect.

4.1. Pore Structure and Morphology

The pore structure and morphology of the carrier have a significant impact on the sustained release rate of drugs. Therefore, in order to achieve drug sustained release, it is necessary to design TiO2 with non-connected pore structures or relatively small pore sizes. Mesoporous titanium dioxide (MTN), with pore sizes ranging from 2 to 50 nm, has controllable porosity at the mesoscale that allows for tunable release rates for drug delivery applications [141]. The drug release properties of MTN and zirconium dioxide were investigated using amoxicillin as a drug model. The drug release kinetics of both nanocarriers were governed by a diffusion control mechanism, and both could deliver the drug continuously for more than 28 days. In contrast, the slightly better drug release rate of MTN was due to its microstructure that retains the drug better [142]. Hollow nanostructures are also often studied due to their high surface area, low density and good permeability. Their inner cavities can accommodate a large number of guest molecules and the porous shells can be used as channels for drug release [143,144,145]. Liu prepared titanium dioxide hollow nanoparticles (HTNs) with controllable particle size by the hydrothermal method, and they also investigated the effects of parameters such as stirring time, precursor concentration, and hydrothermal time on the synthesis. It was found that the prepared HTNs exhibited pH-responsive slow-release behavior in vitro. The release of the loaded drug at different pH (7.4, 6.5, 5.0) conditions increased with decreasing pH value and showed a slow and constant trend, which was exemplified by the zero-order kinetic equation [143]. In addition, the plush TiO2 nanospheres obtained from the preparation had a high surface area, which was favorable for drug loading and easier for the drug to enter the active site. The drug (doxorubicin), before reaching dynamic equilibrium, was released continuously by diffusion control [45,146,147,148,149].

4.2. Interaction Force

The sustained release of the cargo could be achieved by taking advantage of the strong interaction forces such as electrostatic interactions and hydrophobic interactions between the carrier and the cargo. In the drug release behavior of mesoporous TiO2 films loaded with ibuprofen (IBU), the highly loaded films were able to release for a long period of 30 h. This result was not only related to the worm-like curved pores of the TiO2 films, but the hydrogen-bonding interactions between the IBU and the TiO2 films also played a crucial role [150]. The HTNs were bound to the drugs through electrostatic interactions. The release studies also revealed that the drugs loaded in HTNs were released for a longer period and could produce long-term antimicrobial efficacy [151]. In addition, the anticancer drug Daunorubicin (DNR) has three potential metal-binding sites capable of forming complexes with TiO2. The surface of TiO2 is negatively charged in an aqueous solution at pH 7.4, whereby the positively charged DNR self-assembles onto the TiO2 surface via electrostatic interactions. This pH-responsive novel DDS release behavior is capable of prolonging the retention time of the drug in the blood circulation, greatly reducing the side effects on normal tissues, and thus significantly improving the effectiveness of cancer therapy [152].

4.3. Diffusion Inhibition Effect

Modification of polymers on the surface of TiO2 increases the diffusion distance and provides hindrance, thus enabling a more sustained drug release. Polydopamine with multifunctional groups is a popular polymer modifier. Titanium dioxide nanotubes (TNTs) functionalized with polydopamine (PDA) were prepared using ultrasonication methods, and their excellent self-polymerization ability can effectively enhance drug loading and prolong drug release [22]. In addition, TNTs modified with (3-Glycidoxypropyl) trimethoxysilane (GPTMS) also exhibited prolonged release of dexamethasone (DEX). It was the electrostatic and hydrogen-bonding effects of the epoxy ring provided by GPTMS that stabilized the drug molecules on the surface of the TNTs, allowing for a slow and sustained release compared to unmodified TNTs [120].

5. Controlled Drug Delivery

To prevent the premature release of drugs in the carrier, which would result in reduced efficacy and side effects, TiO2-based controlled-release drug delivery systems have been developed to ensure the pharmacological effects while improving drug bioavailability and enhancing the target specificity of drugs. Such controlled drug delivery systems (CDDSs) are usually achieved through modifying various “gatekeepers” via physical adsorption or covalent bonding interactions, where the drug is blocked inside the pore until the CDDSs are exposed to specific stimuli such as pH, enzymes, and redox. In addition to the above endogenous stimuli, controlled release of drugs can also be achieved by relying on exogenous stimuli (light, ultrasound, magnetic, etc.). Figure 4 and Figure 5 show schematic diagrams of drug-controlled-release delivery systems with different responses and triggering the drug release process.

5.1. pH-Responsive

Due to the fast growth of cancer cells, there is a hypoxic environment and lactic acid generation via glycolysis, which lowers the extracellular pH of cancerous cells. This phenomenon, known as the Warburg effect, has led to pH-responsive drug delivery platforms [153]. Furthermore, TiO2 nanotube-based pH-controlled-release drugs and their release kinetics are frequently researched [154,155].
However, for the construction of pH-intelligent CDDSs, it is common to utilize the surface modification of the TiO2 to achieve the release in a desired set of environments. Some polymers are considered effective pH-responsive modifiers. For example, the use of alendronate sodium trihydrate (AST)-modified mesoporous titanium dioxide nanoparticles (MBTNPs) as carriers for DEX drug delivery was also able to achieve pH sensitization, which relies on hydrogen-bonding interactions between the amino groups on the surface of MBTNPs and the hydroxyl groups of DEX [156]. The poly (acrylic acid)-calcium phosphate (PAA-CaP) composite layer has pH-responsive solubilization properties, and TiO2 modified in this way accumulates much faster doxorubicin (DOX) release at pH5.6 than at pH7.4. Furthermore, the PAA-CaP layer exhibits the higher loading and encapsulation capability of DOX [157]. In another study, TNTs were first coated with dopamine, resulting in a dopamine surface with nanoparticles containing amino groups, and then polymethylsilicic acid (PMAA) was covalently attached to the surface via carboxyl–amino group interactions. PMAA has been observed to inflate at pH = 7 and collapse at pH ≤ 6 [158]. Thus, PMAA functioned as a pH-sensitive and switchable molecular “gate” for long-term and on-demand drug release [159]. Other polymers that have been used in the construction of pH-controlled-release platforms include polyvinylpyrrolidone (PVP) [160] and polyethyleneimine (PEI) [161].

5.2. Light-Responsive

A theoretical basis for light-triggered drug release is provided by the fact that the energy absorbed from the ultraviolet (UV) surpasses the band gap of TiO2, which causes the valence electrons to be excited toward the conduction band, forming an electron (e) and hole (h+) pair and further generating active free radicals (OH· and O2−) [162,163]. The novel hybrid PEI modified porous TiO2 nanomaterials were developed by researchers. PEI could form a hydrophilic coating on the surface of TiO2 and prevent premature drug release. After UV irradiation, the PEI coating was destroyed by free radicals, and the encapsulated drug molecules were thus released, realizing a UV-triggered drug delivery system [20]. In addition, it is reported that triggering chitosan (CS) can also control drug release in response to UV irradiation. CS is a natural linear polysaccharide and it has been widely used in pharmaceutical and biomedical applications due to its unique biocompatibility, biodegradability and other properties [164,165]. To improve and control drug delivery to tumors, methotrexate (MTX) loaded in CS nanoparticles has been previously developed. Al-Nemrawi et al. innovatively combined TiO2 with MTX-CS, and TiO2 was used to trigger polymer bond breaking of CS nanoparticles to enable on-demand drug release through photodegradation. Compared to pure CS, MTX, and TiO2, the system had a strong effect against MCF-7 breast cancer cells, with viabilities as low as 7% [166]. Recently, the same study prepared a triggerable system with a core–shell structure, using CS-modified TiO2 as a microcapsule loaded with oregano essential oil. It was observed that the shell molecules adsorbed TiO2 uniformly through hydrogen and ligand bonds. Since the free radicals generated by TiO2 under the action of UV light degraded the glycosidic bonds of CS, loading TiO2 on microcapsules could achieve UV-triggered release of external stimuli, and the release process followed the classical Fickian diffusion mechanism [127]. This system is an effective strategy for controlling the release by UV irradiation as needed, which can be applied to biomedical applications.
However, UV light is known to be harmful and there are some limitations in its application [167]. Near-infrared (NIR) light can penetrate deeper into tissues than UV light, reducing cell damage. As a result, an NIR-stimulated responsive drug delivery system based on TNTs has been proposed. For example, black phosphorus (BP) was introduced into the TNTs and 1-tetradecanol acted as a “gatekeeper” to stop drug leakage. Due to the photothermal properties of the composite, the temperature of the drug delivery system increased after irradiation with near-infrared light, reaching the melting point of the “gatekeeper” leading to its decomposition, and the blocked drug could be released [168]. To suit various therapeutic needs, NIR-stimulated responsive drug delivery systems are capable of transporting not only drugs but also cytokines to establish an ideal local immune microenvironment [169]. It has also been shown that the TiO2 surface was capable of attaching enzymes and light-inducing their release [170].

5.3. Microwave-Responsive

According to research, TiO2 may effectively absorb microwave energy and transform it into localized heat, creating a microwave-induced thermal effect that is sensitive and selective [171,172]. In a study, multifunctional Fe3O4@TiO2: Er3+, Yb3+-glycine nanocomposites were prepared using a hydrothermal method. Thermodynamic analyses demonstrated that the interaction between the carrier and the drug etoposide occurred through relatively weak hydrogen bonding, which was easily broken by microwave irradiation [173]. Cui’s group constructed Janus-shaped TiO2-x&mSiO2 nanoparticles using gray-black titanium dioxide (TiO2-x) and mesoporous silica (mSiO2) as carriers. The rod-shaped mSiO2 acted as an efficient drug carrier, while TiO2-x served as the main microwave absorber. DOX and Janus TiO2-x&mSiO2 interacted with each other by van der Waals forces. Under the stimulation of microwave, the carrier enhanced the drug release rate with excellent thermal conversion ability. Meanwhile, the drug release experiments showed a significant pH-dependent release behavior. Cytotoxicity experiments verified the good biocompatibility of Janus TiO2-x&mSiO2, and this drug delivery carrier was expected to pave the way for subsequent applications [174]. Recently, the authors constructed “Biped” Janus Fe3O4@nSiO2@TiO2-x&mSiO2 nanoparticles as drug carriers, which also possessed pH and microwave dual-triggering properties [175].

5.4. Ultrasound-Responsive

Ultrasound is a promising method for controlling medication release from some carriers. Ultrasound waves can be precisely targeted onto a specific area, and the interaction of ultrasound with tissues can produce a variety of beneficial bioeffects [176]. Cavitation produced by ultrasound is a well-known ultrasonic effect, leading to the development, growth, and burst of gas/vapor-filled microbubbles in liquids. During cavitation, the collapse of the bubble is adiabatic, and thus the bubble serves to concentrate acoustic energy to locally generate extreme temperature and pressure conditions in a short period [177]. Currently, there are various ultrasound responsive drug delivery vehicles based on TiO2 nanomaterials [178,179,180]. A multifunctional capsule system with a core–shell structure capable of simultaneous fluorescence imaging, magnetically guided delivery, and ultrasound-triggered release was reported [181]. The system used safe olive oil as a reservoir for oil-soluble drugs, and Fe3O4, carbon quantum dots, and bilayer porous TiO2 shells as sensitive carriers. The capsule cracked more and more as the ultrasonic duration increased, until the porous titanium shell broke and the medication gushed out. It was demonstrated that the ultrasound period might regulate the drug’s release profile.
An ideal DDS should be able to effectively encapsulate the drug before it reaches the target site [182]. As a result, it is necessary to cover the carrier surface with an effective “gatekeeper”. The β-Cyclodextrin (β-CD) was a typical “gatekeeper”, which can effectively capture a variety of payloads [183,184,185]. One study applied β-CD to drug-loaded MTN as a massive “gatekeeper” to block drug leakage from the mesopore. To enable a burst of drug release, β-CD was attached to the surface of MTN via a ROS-sensitive linker (HOOC-S-CH2-S-COOH). Once the ROS-sensitive linker was recognized and destroyed by ROS, the drug (DTX) in the carrier (MTN@DTX-CD) was instantaneously released. In vivo and in vivo studies demonstrated that this DDS showed good ultrasound-triggered effects and the feasibility of sonodynamic chemotherapy [138]. Previous studies have revealed that PEI could not only serve as “gatekeeper” by blocking mesopore pores, but also increase the cellular uptake of silica mesoporous nanoparticles (MSNs) and develop a safe and effective drug delivery system. Based on this, dendritic silica/titanium dioxide mesoporous nanoparticles (DSTNs) with PEI as a “gatekeeper” and folic acid (FA) as a targeting agent were reported. Ultrasound led to the generation of ROS from TiO2, which destroyed the “gatekeeper”, resulting in the mesoporous pores being unblocked and the curcumin (Cur) being released. The results demonstrated that the Cur@PEI-FA-DSTNs system performed better in combined chemo-acoustic-dynamic cancer therapy than in acoustic-dynamic therapy or chemotherapy alone [186].

5.5. Dual- or Multiple-Stimuli-Responsive

Nowadays, the therapeutic performance of single-responsive carriers is hardly satisfactory, and the integration of multiple stimulus-response mechanisms into one carrier is an extremely promising strategy [187]. According to previous research, Janus-shaped TiO2-x&mSiO2 nanoparticles and “Biped” Janus-Fe3O4@nSiO2@TiO2-x&mSiO2 nanoparticles could not only release drugs under pH control, but also be further triggered by microwave stimulation [175,178]. A novel composite material in which TiO2 and SiO2 nanostructures were synergistically assembled on the surface of the polymer shell of a microcapsule was capable of achieving a triple response to enzyme, UV and ultrasound. It was reported that UV and ultrasound irradiation caused the breakage of SiO2/TiO2-coated capsules into small fragments and the enzymatic degradation led to the deformation of the capsules and cargo leakage [188]. We summarized the DDSs of TiO2 in Table 5.

6. Conclusions and Outlook

In this review, we summarize research advances in TiO2 nanomaterials in terms of preparation methods, surface functionalization and drug delivery. Compared with organic nanomaterials such as liposomes, TiO2 nanomaterials have the advantages of easy functionalization, controllable particle size and morphology, stable physicochemical properties, and high loading capacity. The unique photocatalytic nature of TiO2 nanomaterials endows them with potential functions such as photothermal therapy, which can be considered a rising star in the field of biomedicine and an ideal therapeutic nanocarrier [189]. The high versatility of these smart nano-delivery systems opens up great prospects for the innovation of novel nanomedicines and provides an effective strategy for personalized therapy and diagnostics.
Although TiO2-based smart systems have made great progress, there is still much potential for development. On the one hand, the most studied stimuli for TiO2-based nanoparticles in the design of controlled-release drug delivery systems are pH, light, ultrasound, and microwave, while single-response delivery systems such as redox response and enzyme response, and dual-response delivery systems are rarely reported. These endogenous responses can be the future development direction for TiO2-based drug delivery systems. In addition, the combination of exogenous and endogenous stimuli can further enhance the spatio-temporal controllability as well as the accuracy of nanomaterials’ delivery, and multi-stimulus-responsive nanomaterials have significant advantages in vivo cycling, tumor retention, tissue penetration, cellular internalization, and endosomal escape, which will be the focus of future research.
On the other hand, it is obvious that a large amount of work still needs to be carried out before drug delivery systems achieve clinical translation. However, some TiO2 nanoparticles have been found to have several adverse effects on highly sensitive stem cells, with inflammatory responses being an important feature of their cytotoxicity [190,191]. The data emphasized the correlation of factors such as particle size, shape and crystal structure with the level of genotoxicity [192,193]. Mechanistically, the observed adverse reactions were associated with upregulation of p38, JNK and ERK protein expression [190]. The European Commission has already banned the use of TiO2 (E171) as a food additive from August 2022, citing genotoxicity. Therefore, we must pay more attention to the potential cytotoxicity of the TiO2 nanoparticles at the time of application. Long-term toxicity tests and tolerance studies need to be performed on animals to evaluate the safety of blank DDS and drug-loaded DDS in depth before proceeding with human clinical trials. More in-depth studies should be carried out by optimizing various process parameters to select TiO2 nanoparticles that are more reliable and more suitable for biomedical applications. In addition, the examination of DDS immunoreactivity is also crucial to evaluate the safety of nanoparticles. Repeated exposure to materials that are often considered biocompatible and non-immunogenic can also lead to the production of antibodies that may trigger hypersensitivity reactions [194]. Given that long-term exposure to nanoparticles may cause significant immunogenic effects, comprehensive in vivo studies are essential. Regarding the potential issues with TiO2 nanoparticles, researchers have been actively addressing them with little success. We believe that some TiO2 nanoparticles still have broad prospects in the biomedical field.

Author Contributions

Conceptualization, F.Z., J.M., H.O. and X.Q.; data curation, C.L.; formal analysis, Y.Z., C.L., Y.L. and K.Z.; funding acquisition, J.H.; investigation, T.W., K.Z. and Y.L.; supervision, X.Q. and J.H.; visualization, T.W., X.W. and H.O.; methodology, X.W. and J.M.; writing—original draft, F.Z. and Y.Z.; writing—review and editing, X.Q. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (82274091).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Bonifácio, B.V.; da Silva, P.B.; Ramos, M.A.D.; Negri, K.M.S.; Bauab, T.M.; Chorilli, M. Nanotechnology-based drug delivery systems and herbal medicines: A review. Int. J. Nanomed. 2014, 9, 1–15. [Google Scholar]
  2. Sahu, T.; Ratre, Y.K.; Chauhan, S.; Bhaskar, L.; Nair, M.P.; Verma, H.K. Nanotechnology based drug delivery system: Current strategies and emerging therapeutic potential for medical science. J. Drug Deliv. Sci. Technol. 2021, 63, 102487. [Google Scholar] [CrossRef]
  3. Al Ragib, A.; Chakma, R.; Dewan, K.; Islam, T.; Kormoker, T.; Idris, A.M. Current advanced drug delivery systems: Challenges and potentialities. J. Drug Deliv. Sci. Technol. 2022, 76, 103727. [Google Scholar] [CrossRef]
  4. Ezike, T.C.; Okpala, U.S.; Onoja, U.L.; Nwike, C.P.; Ezeako, E.C.; Okpara, O.J.; Okoroafor, C.C.; Eze, S.C.; Kalu, O.L.; Odoh, E.C.; et al. Advances in drug delivery systems, challenges and future directions. Heliyon 2023, 9, e17488. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, W.; Lu, K.J.; Yu, C.H.; Huang, Q.L.; Du, Y.Z. Nano-drug delivery systems in wound treatment and skin regeneration. J. Nanobiotechnol. 2019, 17, 82. [Google Scholar] [CrossRef]
  6. Liu, J.Y.; Li, S.Q.; Wang, J.; Li, N.N.; Zhou, J.N.; Chen, H.X. Application of Nano Drug Delivery System (NDDS) in Cancer Therapy: A Perspective. Recent Pat. Anti-Cancer Drug Discov. 2023, 18, 125–132. [Google Scholar] [CrossRef]
  7. Liu, J.; Huang, Y.R.; Kumar, A.; Tan, A.; Jin, S.B.; Mozhi, A.; Liang, X.J. pH-Sensitive nano-systems for drug delivery in cancer therapy. Biotechnol. Adv. 2014, 32, 693–710. [Google Scholar] [CrossRef]
  8. Parra-Nieto, J.; del Cid, M.A.G.; de Cárcer, I.A.; Baeza, A. Inorganic Porous Nanoparticles for Drug Delivery in Antitumoral Therapy. Biotechnol. J. 2021, 16, 2000150. [Google Scholar] [CrossRef]
  9. Luther, D.C.; Huang, R.; Jeon, T.; Zhang, X.Z.; Lee, Y.W.; Nagaraj, H.; Rotello, V.M. Delivery of drugs, proteins, and nucleic acids using inorganic nanoparticles. Adv. Drug Deliv. Rev. 2020, 156, 188–213. [Google Scholar] [CrossRef]
  10. Elzoghby, A.O.; Hemasa, A.L.; Freag, M.S. Hybrid protein-inorganic nanoparticles: From tumor-targeted drug delivery to cancer imaging. J. Control Release 2016, 243, 303–322. [Google Scholar] [CrossRef]
  11. Gupta, J.; Quadros, M.; Momin, M. Mesoporous silica nanoparticles: Synthesis and multifaceted functionalization for controlled drug delivery. J. Drug Deliv. Sci. Technol. 2023, 81, 104305. [Google Scholar] [CrossRef]
  12. Manzano, M.; Vallet-Regí, M. Mesoporous Silica Nanoparticles for Drug Delivery. Adv. Funct. Mater. 2020, 30, 1902634. [Google Scholar] [CrossRef]
  13. Saha, S.; Ali, M.R.; Khaleque, M.A.; Bacchu, M.S.; Aly, M.A.S.; Khan, M.Z.H. Metal oxide nanocarrier for targeted drug delivery towards the treatment of global infectious diseases: A review. J. Drug Deliv. Sci. Technol. 2023, 86, 104728. [Google Scholar] [CrossRef]
  14. Haghighi, F.H.; Mercurio, M.; Cerra, S.; Salamone, T.A.; Bianymotlagh, R.; Palocci, C.; Spica, V.R.; Fratoddi, I. Surface modification of TiO2 nanoparticles with organic molecules and their biological applications. J. Mater. Chem. B 2023, 11, 2334–2366. [Google Scholar] [CrossRef]
  15. Malehmir, S.; Abedini, A.; Sobhani-Nasab, A.; Eshraghi, R.; Akbari, M.; Atapour, A.; Hasan-Abad, A.M. A review of biogenic routes for the manufacture of manganese oxide nanostructures and its anti-cancer, drug delivery, anti-bacterial, and bioimaging potentials. Inorg. Chem. Commun. 2023, 156, 111306. [Google Scholar] [CrossRef]
  16. Sattari, S.; Adeli, M.; Beyranvand, S.; Nemati, M. Functionalized Graphene Platforms for Anticancer Drug Delivery. Int. J. Nanomed. 2021, 16, 5955–5980. [Google Scholar] [CrossRef]
  17. An, N.; Qin, C.Y.; Li, Y.W.; Tan, T.; Yuan, Z.Y.; Zhang, H.; Wu, Y.; Yao, B.C.; Rao, Y.J. Graphene-Fiber Biochemical Sensors: Principles, Implementations, and Advances. Photonic Sens. 2021, 11, 123–139. [Google Scholar] [CrossRef]
  18. Pourmadadi, M.; Rajabzadeh-Khosroshahi, M.; Eshaghi, M.M.; Rahmani, E.; Motasadizadeh, H.; Arshad, R.; Rahdar, A.; Pandey, S. TiO2-based nanocomposites for cancer diagnosis and therapy: A comprehensive review. J. Drug Deliv. Sci. Technol. 2023, 82, 104370. [Google Scholar] [CrossRef]
  19. Sun, Z.; Huang, X.B.; Zhang, G. TiO2-based catalysts for photothermal catalysis: Mechanisms, materials and applications. J. Clean. Prod. 2022, 381, 135156. [Google Scholar] [CrossRef]
  20. Wang, T.Y.; Jiang, H.T.; Wan, L.; Zhao, Q.F.; Jiang, T.Y.; Wang, B.; Wang, S.L. Potential application of functional porous TiO2 nanoparticles in light-controlled drug release and targeted drug delivery. Acta Biomater. 2015, 13, 354–363. [Google Scholar] [CrossRef]
  21. Torres, C.C.; Campos, C.H.; Diáz, C.; Jiménez, V.A.; Vidal, F.; Guzmán, L.; Alderete, J.B. PAMAM-grafted TiO2 nanotubes as novel versatile materials for drug delivery applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 65, 164–171. [Google Scholar] [CrossRef] [PubMed]
  22. Li, L.F.; Xie, C.L.; Xiao, X.F. Polydopamine modified TiO2 nanotube arrays as a local drug delivery system for ibuprofen. J. Drug Deliv. Sci. Technol. 2020, 56, 135156. [Google Scholar] [CrossRef]
  23. Li, N.; Zhang, Q.; Joo, J.B.; Lu, Z.; Dahl, M.; Gan, Y.; Yin, Y. Water-assisted crystallization of mesoporous anatase TiO2 nanospheres. Nanoscale 2016, 8, 9113–9117. [Google Scholar] [CrossRef] [PubMed]
  24. Joo, J.B.; Zhang, Q.; Dahl, M.; Lee, I.; Goebl, J.; Zaera, F.; Yin, Y. Control of the nanoscale crystallinity in mesoporous TiO2 shells for enhanced photocatalytic activity. Energy Environ. Sci. 2012, 5, 6321–6327. [Google Scholar] [CrossRef]
  25. Jiang, H.; Wang, T.; Wang, L.; Sun, C.; Jiang, T.; Cheng, G.; Wang, S. Development of an amorphous mesoporous TiO2 nanosphere as a novel carrier for poorly water-soluble drugs: Effect of different crystal forms of TiO2 carriers on drug loading and release behaviors. Microporous Mesoporous Mater. 2012, 153, 124–130. [Google Scholar] [CrossRef]
  26. Chen, S.; Shen, L.L.; Huang, D.; Du, J.; Fan, X.X.; Wei, A.L.; Jia, L.; Chen, W.Y. Facile synthesis, microstructure, formation mechanism, in vitro biocompatibility, and drug delivery property of novel dendritic TiO2 nanofibers with ultrahigh surface area. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 115, 111100. [Google Scholar] [CrossRef]
  27. Gupta, S.M.; Tripathi, M. A review on the synthesis of TiO2 nanoparticles by solution route. Cent. Eur. J. Chem. 2012, 10, 279–294. [Google Scholar] [CrossRef]
  28. Li, W.; Yang, J.P.; Wu, Z.X.; Wang, J.X.; Li, B.; Feng, S.S.; Deng, Y.H.; Zhang, F.; Zhao, D.Y. A Versatile Kinetics-Controlled Coating Method To Construct Uniform Porous TiO2 Shells for Multifunctional Core-Shell Structures. J. Am. Chem. Soc. 2012, 134, 11864–11867. [Google Scholar] [CrossRef]
  29. Celzard, A.; Mareche, J.F. Applications of the Sol-gel Process Using Well-Tested Recipes. Am. Chem. Soc. 2002, 79, 854–859. [Google Scholar] [CrossRef]
  30. Li, Y.; White, T.; Lim, S.H. Structure control and its influence on photoactivity and phase transformation of TiO2 nano-particles. Rev. Adv. Mater. Sci. 2003, 5, 211–215. [Google Scholar]
  31. Li, B.R.; Wang, X.H.; Yan, M.Y.; Li, L.T. Preparation and characterization of nano-TiO2 powder. Mater. Chem. Phys. 2003, 78, 184–188. [Google Scholar] [CrossRef]
  32. Park, G.; Kim, M.; Lee, H. Steam-modified sol-gel method for TiO2-based nanosized particles. Ceram. Int. 2022, 48, 25656–25660. [Google Scholar] [CrossRef]
  33. Jaroenworaluck, A.; Sunsaneeyametha, W.; Kosachan, N.; Stevens, R. Characteristics of silica-coated TiO2 and its UV absorption for sunscreen cosmetic applications. Surf. Interface Anal. 2006, 38, 473–477. [Google Scholar] [CrossRef]
  34. Réti, B.; Major, Z.; Szarka, D.; Boldizsár, T.; Horváth, E.; Magrez, A.; Forró, L.; Dombi, A.; Hernádi, K. Influence of TiO2 phase composition on the photocatalytic activity of TiO2/MWCNT composites prepared by combined sol-gel/hydrothermal method. J. Mol. Catal. A Chem. 2016, 414, 140–147. [Google Scholar] [CrossRef]
  35. Calabrese, C.; Maertens, A.; Piras, A.; Aprile, C.; Liotta, L.F. Novel Sol-Gel Synthesis of TiO2 Spherical Porous Nanoparticles Assemblies with Photocatalytic Activity. Nanomaterials 2023, 13, 1928. [Google Scholar] [CrossRef]
  36. Wu, J.H.; Tu, J.L.; Hu, K.; Xiao, X.J.; Li, L.; Yu, S.Z.; Xie, Y.C.; Wu, H.; Yang, Y.Y. Sol-gel-derived bayberry-like SiO2@TiO2 multifunctional antireflective coatings for enhancing photovoltaic power generation. Colloid. Surf. A 2022, 654, 130173. [Google Scholar] [CrossRef]
  37. Usman, A.K.; Cursaru, D.L.; Branoiu, G.; Somoghi, R.; Manta, A.M.; Matei, D.; Mihai, S. A Modified Sol-Gel Synthesis of Anatase {001}-TiO2/Au Hybrid Nanocomposites for Enhanced Photodegradation of Organic Contaminants. Gels 2022, 8, 728. [Google Scholar] [CrossRef]
  38. Sondezi, N.; Njengele-Tetyana, Z.; Matabola, K.P.; Makhetha, T.A. Sol-Gel-Derived TiO2 and TiO2/Cu Nanoparticles: Synthesis, Characterization, and Antibacterial Efficacy. ACS Omega 2024, 9, 15959–15970. [Google Scholar] [CrossRef]
  39. Kader, S.; Al-Mamun, M.R.; Suhan, M.B.K.; Shuchi, S.B.; Islam, M.S. Enhanced photodegradation of methyl orange dye under UV irradiation using MoO3 and Ag doped TiO2 photocatalysts. Environ. Technol. Innov. 2022, 27, 102476. [Google Scholar] [CrossRef]
  40. Byrappa, K.; Adschiri, T. Hydrothermal technology for nanotechnology. Prog. Cryst. Growth Charact. Mater. 2007, 53, 117–166. [Google Scholar] [CrossRef]
  41. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Formation of titanium oxide nanotube. Langmuir 1998, 14, 3160–3163. [Google Scholar] [CrossRef]
  42. Thapa, R.; Maiti, S.; Rana, T.H.; Maiti, U.N.; Chattopadhyay, K.K. Anatase TiO2 nanoparticles synthesis via simple hydrothermal route: Degradation of Orange II, Methyl Orange and Rhodamine B. J. Mol. Catal. A Chem. 2012, 363, 223–229. [Google Scholar] [CrossRef]
  43. Liu, Z.Y.; Sun, D.D.; Guo, P.; Leckie, J.O. One-step fabrication and high photocatalytic activity of porous TiO2 hollow aggregates by using a low-temperature hydrothermal method without templates. Chem. A Eur. J. 2007, 13, 1851–1855. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, R.; Bao, Y.; Kang, Q.; Liu, C.; Zhang, W.; Zhu, Q. Solvent-controlled synthesis and photocatalytic activity of hollow TiO2 microspheres prepared by the solvothermal method. Colloids Surf. A Physicochem. Eng. Asp. 2022, 633, 127931. [Google Scholar] [CrossRef]
  45. Cui, X.-G.; Chen, H.; Ye, Q.-B.; Cui, X.-Y.; Cui, X.-J.; Cui, H.-J.; Shen, G.-Z.; Li, M.-J.; Lin, J.-T.; Sun, Y.-X. Porous Titanium Dioxide Spheres for Drug Delivery and Sustained Release. Front. Mater. 2021, 8, 649237. [Google Scholar] [CrossRef]
  46. Sun, W.J.; Zheng, H.N.; Ma, J.B.; Xi, Z.Y.; Wang, B.X.; Hao, C.C. Preparation and electrorheological properties of eggshell-like TiO2 hollow spheres via one step template-free solvothermal method. Colloid. Surf. A 2020, 601, 125055. [Google Scholar] [CrossRef]
  47. Song, Y.D.; Li, B.; Shi, Y.Y.; Wang, Q.Y.; Jin, R.C.; Huang, B.B.; Dai, Y.; Gao, S.M. Preparation of fusiform Ti3+ self-doped TiO2 nanoparticles by mixed solvothermal method and its photoelectrochemical properties. Mater. Lett. 2019, 252, 134–137. [Google Scholar] [CrossRef]
  48. Rani, N.; Dehiya, B.S. Influence of anionic and non-ionic surfactants on the synthesis of core-shell Fe3O4@TiO2 nanocomposite synthesized by hydrothermal method. Ceram. Int. 2020, 46, 23516–23525. [Google Scholar] [CrossRef]
  49. Jahdi, M.; Mishra, S.B.; Nxumalo, E.N.; Mhlanga, S.D.; Mishra, A.K. Synergistic effects of sodium fluoride (NaF) on the crystallinity and band gap of Fe-doped TiO2 developed via microwave-assisted hydrothermal treatment. Opt. Mater. 2020, 104, 109844. [Google Scholar] [CrossRef]
  50. Cao, Z.Z.; Yin, Q.; Zhang, Y.X.; Li, Y.; Yu, C.J.; Zhang, M.K.; Fan, B.B.; Shao, G.; Wang, H.L.; Xu, H.L.; et al. Heterostructure composites of TiO2 and CdZnS nanoparticles decorated on Ti3C2TX nanosheets and their enhanced photocatalytic performance by microwave hydrothermal method. J. Alloys Compd. 2022, 918, 165681. [Google Scholar] [CrossRef]
  51. Sahu, K.; Dhonde, M.; Murty, V.V.S. Microwave-assisted hydrothermal synthesis of Cu-doped TiO2 nanoparticles for efficient dye-sensitized solar cell with improved open-circuit voltage. Int. J. Energy Res. 2020, 45, 5423–5432. [Google Scholar] [CrossRef]
  52. Shi, Y.F.; Wan, Y.; Zhao, D.Y. Ordered mesoporous non-oxide materials. Chem. Soc. Rev. 2011, 40, 3854–3878. [Google Scholar] [CrossRef] [PubMed]
  53. Deng, X.H.; Chen, K.; Tüysüz, H. Protocol for the Nanocasting Method: Preparation of Ordered Mesoporous Metal Oxides. Chem. Mater. 2017, 29, 40–52. [Google Scholar] [CrossRef]
  54. Wang, W.; Li, T.; Ye, K.; Long, H.B.; Wang, X.; Ru, H.Q. Preparation of hierarchically mesoporous silica microspheres with ordered mesochannels and ultra-large intra-particulate pores. Microporous Mesoporous Mater. 2016, 234, 267–276. [Google Scholar] [CrossRef]
  55. Wang, W.; Sun, R.; Zhu, L.; Wu, C.; Ru, H. Hierarchical mesoporous silica microspheres as unique hard-template for preparation of hierarchical mesoporous TiO2 microspheres with trimodal mesoporosities via nanocasting. Ceram. Int. 2019, 45, 16521–16529. [Google Scholar] [CrossRef]
  56. Li, Z.; He, J.; Ma, H.; Zang, L.; Li, D.; Guo, S.; Ci, Y. Preparation of heterogeneous TiO2/g-C3N4 with a layered mosaic stack structure by use of montmorillonite as a hard template approach: TC degradation, kinetic, mechanism, pathway and DFT investigation. Appl. Clay Sci. 2021, 207, 106107. [Google Scholar] [CrossRef]
  57. Wang, J.L.; Wang, X.Q.; Chen, Q.H.; Xu, H.L.; Dai, M.; Zhang, M.; Wang, W.Y.; Song, H. Microstructural modification of hollow TiO2 nanospheres and their photocatalytic performance. Appl. Surf. Sci. 2021, 535, 147641. [Google Scholar] [CrossRef]
  58. Kanjana, N.; Maiaugree, W.; Poolcharuansin, P.; Laokul, P. Size controllable synthesis and photocatalytic performance of mesoporous TiO2 hollow spheres. J. Mater. Sci. Technol. 2020, 48, 105–113. [Google Scholar] [CrossRef]
  59. Monnier, A.; Schüth, F.; Huo, Q.; Kumar, D.; Margolese, D.; Maxwell, R.S.; Stucky, G.D.; Krishnamurty, M.; Petroff, P.; Firouzi, A.; et al. Cooperative Formation of Inorganic-Organic Interfaces in the Synthesis of Silicate Mesostructures. Science 1993, 261, 1299–1303. [Google Scholar] [CrossRef]
  60. Lan, K.; Liu, Y.; Zhang, W.; Liu, Y.; Elzatahry, A.; Wang, R.C.; Xia, Y.Y.; Al-Dhayan, D.; Zheng, N.F.; Zhao, D.Y. Uniform Ordered Two-Dimensional Mesoporous TiO2 Nanosheets from Hydrothermal-Induced Solvent-Confined Monomicelle Assembly. J. Am. Chem. Soc. 2018, 140, 4135–4143. [Google Scholar] [CrossRef]
  61. Hung, I.M.; Wang, Y.; Huang, C.-F.; Fan, Y.-S.; Han, Y.-J.; Peng, H.-W. Effects of templating surfactant concentrations on the mesostructure of ordered mesoporous anatase TiO2 by an evaporation-induced self-assembly method. J. Eur. Ceram. Soc. 2010, 30, 2065–2072. [Google Scholar] [CrossRef]
  62. Lee, J.-H.; Chakraborty, D.; Chatterjee, S.; Cho, E.-B. Role of polymer template in crystal structure and photoactivity of Cu-TiO2 heterojunction nanostructures towards environmental remediation. Environ. Res. 2023, 232, 116352. [Google Scholar] [CrossRef]
  63. Fan, J.; Boettcher, S.W.; Stucky, G.D. Nanoparticle assembly of ordered multicomponent mesostructured metal oxides via a versatile sol-gel process. Chem. Mater. 2006, 18, 6391–6396. [Google Scholar] [CrossRef]
  64. Xiong, H.; Zhou, H.; Qi, C.; Liu, Z.; Zhang, L.; Zhang, L.; Qiao, Z.-A. Polymer-oriented evaporation induced self-assembly strategy to synthesize highly crystalline mesoporous metal oxides. Chem. Eng. J. 2020, 398, 125527. [Google Scholar] [CrossRef]
  65. Wisutiratanamanee, A.; Poochinda, K.; Poompradub, S. Low-temperature particle synthesis of titania/silica/natural rubber composites for antibacterial properties. Adv. Powder Technol. 2017, 28, 1263–1269. [Google Scholar] [CrossRef]
  66. Yang, K.; Zhong, S.; Zhou, X.M.; Tang, S.Y.; Qiao, K.; Ma, K.; Song, L.; Yue, H.R.; Liang, B. Controllable Al2O3 coating makes TiO2 photocatalysts active under visible light by pulsed chemical vapor deposition. Chem. Eng. Sci. 2023, 277, 118792. [Google Scholar] [CrossRef]
  67. Lang, J.H.; Takahashi, K.; Kubo, M.; Shimada, M. Ag-Doped TiO2 Composite Films Prepared Using Aerosol-Assisted, Plasma-Enhanced Chemical Vapor Deposition. Catalysts 2022, 12, 365. [Google Scholar] [CrossRef]
  68. Hudandini, M.; Kusdianto, K.; Kubo, M.; Shimada, M. Gas-Phase Fabrication and Photocatalytic Activity of TiO2 and TiO2-CuO Nanoparticulate Thin Films. Materials 2024, 17, 1149. [Google Scholar] [CrossRef]
  69. Hasan, M.M.; Haseeb, A.; Saidur, R.; Masjuki, H.H.; Hamdi, M. Influence of substrate and annealing temperatures on optical properties of RF-sputtered TiO2 thin films. Opt. Mater. 2010, 32, 690–695. [Google Scholar] [CrossRef]
  70. González-García, L.; González-Valls, I.; Lira-Cantu, M.; Barranco, A.; González-Elipe, A.R. Aligned TiO2 nanocolumnar layers prepared by PVD-GLAD for transparent dye sensitized solar cells. Energy Environ. Sci. 2011, 4, 3426–3435. [Google Scholar] [CrossRef]
  71. Othman, S.H.; Rashid, S.A.; Ghazi, T.I.M.; Abdullah, N. Effect of Postdeposition Heat Treatment on the Crystallinity, Size, and Photocatalytic Activity of TiO2 Nanoparticles Produced via Chemical Vapour Deposition. J. Nanomater. 2010, 2010, 512785. [Google Scholar] [CrossRef]
  72. Li, G.; Bai, R.B.; Zhao, X.S. Coating of TiO2 Thin Films on the Surface of SiO2 Microspheres: Toward Industrial Photocatalysis. Ind. Eng. Chem. Res. 2008, 47, 8228–8232. [Google Scholar] [CrossRef]
  73. Evans, P.; Sheel, D.W. Photoactive and antibacterial TiO2 thin films on stainless steel. Surf. Coat. Technol. 2007, 201, 9319–9324. [Google Scholar] [CrossRef]
  74. Zhang, Q.; Li, C. High Temperature Stable Anatase Phase Titanium Dioxide Films Synthesized by Mist Chemical Vapor Deposition. Nanomaterials 2020, 10, 911. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, W.; Tadé, M.O.; Shao, Z.P. Nitrogen-doped simple and complex oxides for photocatalysis: A review. Prog. Mater. Sci. 2018, 92, 33–63. [Google Scholar] [CrossRef]
  76. Alotaibi, A.M.; Williamson, B.A.D.; Sathasivam, S.; Kafizas, A.; Alqahtani, M.; Sotelo-Vazquez, C.; Buckeridge, J.; Wu, J.; Nair, S.P.; Scanlon, D.O.; et al. Enhanced Photocatalytic and Antibacterial Ability of Cu-Doped Anatase TiO2 Thin Films: Theory and Experiment. ACS Appl. Mater. Interfaces 2020, 12, 15348–15361. [Google Scholar] [CrossRef]
  77. Su, K.; Tan, L.; Liu, X.M.; Cui, Z.D.; Zheng, Y.F.; Li, B.; Han, Y.; Li, Z.Y.; Zhu, S.L.; Liang, Y.Q.; et al. Rapid Photo-Sonotherapy for Clinical Treatment of Bacterial Infected Bone Implants by Creating Oxygen Deficiency Using Sulfur Doping. Acs Nano 2020, 14, 2077–2089. [Google Scholar] [CrossRef]
  78. Xia, P.F.; Cao, S.W.; Zhu, B.C.; Liu, M.J.; Shi, M.S.; Yu, J.G.; Zhang, Y.F. Designing a 0D/2D S-Scheme Heterojunction over Polymeric Carbon Nitride for Visible-Light Photocatalytic Inactivation of Bacteria. Angew. Chem. Int. Ed. 2020, 59, 5218–5225. [Google Scholar] [CrossRef]
  79. Babudurai, M.; Nwakanma, O.; Romero-Nuñez, A.; Manisekaran, R.; Subramaniam, V.; Castaneda, H.; Jantrania, A. Mechanical activation of TiO2/Fe2O3 nanocomposite for arsenic adsorption: Effect of ball-to-powder ratio and milling time. J. Nanostruct. Chem. 2021, 11, 619–632. [Google Scholar] [CrossRef]
  80. Cao, F.; Li, Y.; Tang, C.; Qian, X.; Bian, Z. Fast synthesis of anatase TiO2 single crystals by a facile solid-state method. Res. Chem. Intermed. 2016, 42, 5975–5981. [Google Scholar] [CrossRef]
  81. Hu, J.W.; Geng, X.Z.; Duan, Y.F.; Zhao, W.M.; Zhu, M.Q.; Ren, S.J. Effect of Mechanical-Chemical Modification Process on Mercury Removal of Bromine Modified Fly Ash. Energy Fuels 2020, 34, 9829–9839. [Google Scholar] [CrossRef]
  82. Do, J.L.; Friscic, T. Mechanochemistry: A Force of Synthesis. ACS Cent. Sci. 2017, 3, 13–19. [Google Scholar] [CrossRef] [PubMed]
  83. Verma, V.; Al-Dossari, M.; Singh, J.; Rawat, M.; Kordy, M.G.M.; Shaban, M. A Review on Green Synthesis of TiO2 NPs: Photocatalysis and Antimicrobial Applications. Polymers 2022, 14, 1444. [Google Scholar] [CrossRef] [PubMed]
  84. Chandra, A.; Bhattarai, A.; Yadav, A.K.; Adhikari, J.; Singh, M.; Giri, B. Green Synthesis of Silver Nanoparticles Using Tea Leaves from Three Different Elevations. Chemistryselect 2020, 5, 4239–4246. [Google Scholar] [CrossRef]
  85. Syahin Firdaus Aziz Zamri, M.; Sapawe, N. Effect of pH on Phenol Degradation Using Green Synthesized Titanium Dioxide Nanoparticles. Mater. Today Proc. 2019, 19, 1321–1326. [Google Scholar] [CrossRef]
  86. Kaur, H.; Kaur, S.; Kumar, S.; Singh, J.; Rawat, M. Eco-friendly Approach: Synthesis of Novel Green TiO2 Nanoparticles for Degradation of Reactive Green 19 Dye and Replacement of Chemical Synthesized TiO2. J. Clust. Sci. 2021, 32, 1191–1204. [Google Scholar] [CrossRef]
  87. Rostami-Vartooni, A.; Nasrollahzadeh, M.; Salavati-Niasari, M.; Atarod, M. Photocatalytic degradation of azo dyes by titanium dioxide supported silver nanoparticles prepared by a green method using Carpobrotus acinaciformis extract. J. Alloys Compd. 2016, 689, 15–20. [Google Scholar] [CrossRef]
  88. Mohanpuria, P.; Rana, N.K.; Yadav, S.K. Biosynthesis of nanoparticles: Technological concepts and future applications. J. Nanopart. Res. 2008, 10, 507–517. [Google Scholar] [CrossRef]
  89. Nies, D.H. Microbial heavy-metal resistance. Appl. Microbiol. Biotechnol. 1999, 51, 730–750. [Google Scholar] [CrossRef]
  90. Suriyaraj, S.P.; Selvakumar, R. Room temperature biosynthesis of crystalline TiO2 nanoparticles using Bacillus licheniformis and studies on the effect of calcination on phase structure and optical properties. RSC Adv. 2014, 4, 39619–39624. [Google Scholar] [CrossRef]
  91. Jayaseelan, C.; Rahuman, A.A.; Roopan, S.M.; Kirthi, A.V.; Venkatesan, J.; Kim, S.K.; Iyappan, M.; Siva, C. Biological approach to synthesize TiO2 nanoparticles using Aeromonas hydrophila and its antibacterial activity. Spectrochim. Acta A 2013, 107, 82–89. [Google Scholar] [CrossRef] [PubMed]
  92. Dhandapani, P.; Maruthamuthu, S.; Rajagopal, G. Bio-mediated synthesis of TiO2 nanoparticles and its photocatalytic effect on aquatic biofilm. J. Photochem. Photobiol. B Biol. 2012, 110, 43–49. [Google Scholar] [CrossRef] [PubMed]
  93. Haider, A.J.; Jameel, Z.N.; Al-Hussaini, I.H.M. Review on: Titanium Dioxide Applications. Energy Procedia 2019, 157, 17–29. [Google Scholar] [CrossRef]
  94. Saini, R.; Kumar, P. Green synthesis of TiO2 nanoparticles using Tinospora cordifolia plant extract & its potential application for photocatalysis and antibacterial activity. Inorg. Chem. Commun. 2023, 156, 111221. [Google Scholar]
  95. Nabi, G.; Majid, A.; Riaz, A.; Alharbi, T.; Kamran, M.A.; Al-Habardi, M. Green synthesis of spherical TiO2 nanoparticles using Citrus Limetta extract: Excellent photocatalytic water decontamination agent for RhB dye. Inorg. Chem. Commun. 2021, 129, 108618. [Google Scholar] [CrossRef]
  96. Sethy, N.K.; Arif, Z.; Mishra, P.K.; Kumar, P. Green synthesis of TiO2 nanoparticles from Syzygium cumini extract for photo-catalytic removal of lead (Pb) in explosive industrial wastewater. Green Process. Synth. 2020, 9, 171–181. [Google Scholar] [CrossRef]
  97. Anbumani, D.; Dhandapani, K.V.; Manoharan, J.; Babujanarthanam, R.; Bashir, A.K.H.; Muthusamy, K.; Alfarhan, A.; Kanimozhi, K. Green synthesis and antimicrobial efficacy of titanium dioxide nanoparticles using Luffa acutangula leaf extract. J. King Saud Univ. Sci. 2022, 34, 101896. [Google Scholar] [CrossRef]
  98. Ahmad, W.; Jaiswal, K.K.; Soni, S. Green synthesis of titanium dioxide (TiO2) nanoparticles by using Mentha arvensis leaves extract and its antimicrobial properties. Inorg. Nano-Met. Chem. 2020, 50, 1032–1038. [Google Scholar] [CrossRef]
  99. Ansari, A.; Siddiqui, V.U.; Rehman, W.U.; Akram, M.K.; Siddiqi, W.A.; Alosaimi, A.M.; Hussein, M.A.; Rafatullah, M. Green Synthesis of TiO2 Nanoparticles Using Acorus calamus Leaf Extract and Evaluating Its Photocatalytic and In Vitro Antimicrobial Activity. Catalysts 2022, 12, 181. [Google Scholar] [CrossRef]
  100. Singh, S.; Maurya, I.C.; Tiwari, A.; Srivastava, P.; Bahadur, L. Green synthesis of TiO2 nanoparticles using Citrus limon juice extract as a bio-capping agent for enhanced performance of dye-sensitized solar cells. Surf. Interfaces 2022, 28, 101652. [Google Scholar] [CrossRef]
  101. Rahmawati, T.; Butburee, T.; Sangkhun, W.; Wutikhun, T.; Padchasri, J.; Kidkhunthod, P.; Phromma, S.; Eksangsri, T.; Kangwansupamonkon, W.; Leeladee, P.; et al. Green synthesis of Ag-TiO2 nanoparticles using turmeric extract and its enhanced photocatalytic activity under visible light. Colloid. Surf. A 2023, 665, 131206. [Google Scholar] [CrossRef]
  102. Panneerselvam, A.; Velayutham, J.; Ramasamy, S. Green synthesis of TiO2 nanoparticles prepared from Phyllanthus niruri leaf extract for dye adsorption and their isotherm and kinetic studies. IET Nanobiotechnol. 2021, 15, 164–172. [Google Scholar] [CrossRef] [PubMed]
  103. Rajeswari, V.D.; Eed, E.M.; Elfasakhany, A.; Badruddin, I.A.; Kamangar, S.; Brindhadevi, K. Green synthesis of titanium dioxide nanoparticles using Laurus nobilis (bay leaf): Antioxidant and antimicrobial activities. Appl. Nanosci. 2021, 13, 1477–1484. [Google Scholar] [CrossRef]
  104. Taran, M.; Rad, M.; Alavi, M. Biosynthesis of TiO2 and ZnO nanoparticles by Halomonas elongata IBRC-M 10214 in different conditions of medium. Bioimpacts 2018, 8, 81–89. [Google Scholar] [CrossRef]
  105. Jha, A.K.; Prasad, K. Biosynthesis of metal and oxide nanoparticles using Lactobacilli from yoghurt and probiotic spore tablets. Biotechnol. J. 2010, 5, 285–291. [Google Scholar] [CrossRef]
  106. Kirthi, A.V.; Rahuman, A.A.; Rajakumar, G.; Marimuthu, S.; Santhoshkumar, T.; Jayaseelan, C.; Elango, G.; Zahir, A.A.; Kamaraj, C.; Bagavan, A. Biosynthesis of titanium dioxide nanoparticles using bacterium Bacillus subtilis. Mater. Lett. 2011, 65, 2745–2747. [Google Scholar] [CrossRef]
  107. Kulal, D.; Shetty Kodialbail, V. Visible light mediated photocatalytic dye degradation using Ag2O/AgO-TiO2 nanocomposite synthesized by extracellular bacterial mediated synthesis-An eco-friendly approach for pollution abatement. J. Environ. Chem. Eng. 2021, 9, 105389. [Google Scholar] [CrossRef]
  108. Bansal, V.; Poddar, P.; Ahmad, A.; Sastry, M. Room-temperature biosynthesis of ferroelectric barium titanate nanoparticles. Journal of the American Chem. Soc. 2006, 128, 11958–11963. [Google Scholar] [CrossRef]
  109. Cantera, J.F.; Wang, J.; Carrea, S.P.P.; Chen, L.; Salmones, J.; González, J. A microwave-ultrasound assisted synthesis of defective TiO2 and WO3/TiO2 nanoparticles for ultralow sulfur diesel production. Mater. Lett. 2024, 360, 136030. [Google Scholar] [CrossRef]
  110. Giram, D.; Das, A.; Bhanvase, B. Comparative study of ZnO-TiO2 nanocomposites synthesized by ultrasound and conventional methods for the degradation of methylene blue dye. Indian J. Chem. Technol. 2023, 30, 693–704. [Google Scholar]
  111. Sivaprakash, V.; Narayanan, R. Anodic synthesis of TiO2 nanotubes by step-up voltages. J. Compos. Mater. 2021, 55, 3775–3784. [Google Scholar] [CrossRef]
  112. Pillai, K. Single crystalline rutile TiO2 nanorods synthesis by onestep catalyst-free vapor transport method. Solid State Commun. 2021, 333, 114342. [Google Scholar] [CrossRef]
  113. Kang, S.; Choi, J.; Park, G.Y.; Kim, H.R.; Hwang, J. A novel and facile synthesis of Ag-doped TiO2 nanofiber for airborne virus/bacteria inactivation and VOC elimination under visible light. Appl. Surf. Sci. 2022, 599, 153930. [Google Scholar] [CrossRef]
  114. Cho, S.; Yim, G.; Park, J.T.; Jang, H. Surfactant-free one-pot synthesis of Au-TiO2 core-shell nanostars by inter-cation redox reaction for photoelectrochemical water splitting. Energy Convers. Manag. 2022, 252, 115038. [Google Scholar] [CrossRef]
  115. Garcia-Munoz, P.; Zussblatt, N.P.; Pliego, G.; Zazo, J.A.; Fresno, F.; Chmelka, B.F.; Casas, J.A. Evaluation of photoassisted treatments for norfloxacin removal in water using mesoporous Fe2O3-TiO2 materials. J. Environ. Manag. 2019, 238, 243–250. [Google Scholar] [CrossRef]
  116. Hu, Y.; Cai, K.Y.; Luo, Z.; Xu, D.W.; Xie, D.C.; Huang, Y.R.; Yang, W.H.; Liu, P. TiO2 nanotubes as drug nanoreservoirs for the regulation of mobility and differentiation of mesenchymal stem cells. Acta Biomater. 2012, 8, 439–448. [Google Scholar] [CrossRef]
  117. Wen, J.Q.; Li, X.; Liu, W.; Fang, Y.P.; Xie, J.; Xu, Y.H. Photocatalysis fundamentals and surface modification of TiO2 nanomaterials. Chin. J. Catal. 2015, 36, 2049–2070. [Google Scholar] [CrossRef]
  118. Wang, Y.; Chen, L.; Liu, P. Biocompatible triplex Ag@SiO2@mTiO2 core-shell nanoparticles for simultaneous fluorescence-SERS bimodal imaging and drug delivery. Chemistry 2012, 18, 5935–5943. [Google Scholar] [CrossRef]
  119. Monreal-Pérez, P.; Isasi, J.R.; González-Benito, J.; Olmos, D.; González-Gaitano, G. Cyclodextrin-Grafted TiO2 Nanoparticles: Synthesis, Complexation Capacity, and Dispersion in Polymeric Matrices. Nanomaterials 2018, 8, 642. [Google Scholar] [CrossRef]
  120. Khoshnood, N.; Zamanian, A.; Massoudi, A. Tailoring in vitro drug delivery properties of titania nanotubes functionalized with (3-Glycidoxypropyl) trimethoxysilane. Mater. Chem. Phys. 2017, 193, 290–297. [Google Scholar] [CrossRef]
  121. Qiao, Y.T.; Wan, J.Q.; Zhou, L.Q.; Ma, W.; Yang, Y.Y.; Luo, W.X.; Yu, Z.Q.; Wang, H.X. Stimuli-responsive nanotherapeutics for precision drug delivery and cancer therapy. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2019, 11, e1527. [Google Scholar] [CrossRef] [PubMed]
  122. Rong, M.Z.; Zhang, M.Q.; Ruan, W.H. Surface modification of nanoscale fillers for improving properties of polymer nanocomposites: A review. Mater. Sci. Technol. 2006, 22, 787–796. [Google Scholar] [CrossRef]
  123. Hussain, Z.; Khan, S.; Imran, M.; Sohail, M.; Shah, S.W.A.; de Matas, M. PEGylation: A promising strategy to overcome challenges to cancer-targeted nanomedicines: A review of challenges to clinical transition and promising resolution. Drug Deliv. Transl. Res. 2019, 9, 721–734. [Google Scholar] [CrossRef] [PubMed]
  124. Gamucci, O.; Bertero, A.; Gagliardi, M.; Bardi, G. Biomedical Nanoparticles: Overview of Their Surface Immune-Compatibility. Coatings 2014, 4, 139–159. [Google Scholar] [CrossRef]
  125. Selli, D.; Motta, S.; Di Valentin, C. Impact of surface curvature, grafting density and solvent type on the PEGylation of titanium dioxide nanoparticles. J. Colloid. Interf. Sci. 2019, 555, 519–531. [Google Scholar] [CrossRef]
  126. Sun, X.Y.; Wang, H.D.; Lai, Z.Y.; Zhang, Z.H.; Zhang, Z.C.; Hang, J.Z.; Shi, L.Y. High stability, high solid content, low viscosity titanium dioxide dispersion. J. Coat. Technol. Res. 2024, 1–10. [Google Scholar] [CrossRef]
  127. Xu, L.A.; Xu, X.L.; Xu, Y.J.; Huang, M.Y.; Li, Y.L. Fabrication and immediate release characterization of UV responded oregano essential oil loaded microcapsules by chitosan-decorated titanium dioxide. Food Chem. 2023, 400, 133965. [Google Scholar] [CrossRef]
  128. Yousefi, E.; Javadpour, S.; Ansari, M.; Eslami, H. Sonodynamic therapy of cancer using a novel TiO2-based nanoparticles. Mater. Technol. 2021, 36, 521–528. [Google Scholar]
  129. Wu, L.Z.; Wu, X.; Wu, L.Z.; Chen, D.D.; Zhang, T.; Zheng, H.; Xiao, X.F. Polydopamine-Modified Titanium Dioxide Nanotube Arrays Doped with Calcium as a Sustained Drug Delivery System. ACS Omega 2024, 9, 4949–4956. [Google Scholar] [CrossRef]
  130. Matiyani, M.; Rana, A.; Karki, N.; Garwal, K.; Pal, M.; Sahoo, N.G. Development of multi-functionalized graphene oxide based nanocarrier for the delivery of poorly water soluble anticancer drugs. J. Drug Deliv. Sci. Technol. 2023, 83, 104412. [Google Scholar] [CrossRef]
  131. Cheng, F.; Sajedin, S.M.; Kelly, S.M.; Lee, A.F.; Kornherr, A. UV-stable paper coated with APTES-modified P25 TiO2 nanoparticles. Carbohydr. Polym. 2014, 114, 246–252. [Google Scholar] [CrossRef] [PubMed]
  132. Khabibullin, A.; Bhangaonkar, K.; Mahoney, C.; Lu, Z.; Schmitt, M.; Sekizkardes, A.K.; Bockstaller, M.R.; Matyjaszewski, K. Grafting PMMA Brushes from α-Alumina Nanoparticles via SI-ATRP. Acs Appl. Mater. Interfaces 2016, 8, 5458–5465. [Google Scholar] [CrossRef] [PubMed]
  133. Deshmukh, S.; Kandasamy, G.; Upadhyay, R.K.; Bhattacharya, G.; Banerjee, D.; Maity, D.; Deshusses, M.A.; Roy, S.S. Terephthalic acid capped iron oxide nanoparticles for sensitive electrochemical detection of heavy metal ions in water. J. Electroanal. Chem. 2017, 788, 91–98. [Google Scholar] [CrossRef]
  134. Jia, X.Q.; Ma, J.P.; Xia, F.; Xu, Y.M.; Gao, J.; Xu, J. Carboxylic acid-modified metal oxide catalyst for selectivity-tunable aerobic ammoxidation. Nat. Commun. 2018, 9, 933. [Google Scholar] [CrossRef]
  135. Pal, S.; Taurino, A.; Catalano, M.; Licciulli, A. Block Copolymer and Cellulose Templated Mesoporous TiO2-SiO2 Nanocomposite as Superior Photocatalyst. Catalysts 2022, 12, 770. [Google Scholar] [CrossRef]
  136. Joseph, C.G.; Taufiq-Yap, Y.H.; Letshmanan, E.; Vijayan, V. Heterogeneous Photocatalytic Chlorination of Methylene Blue Using a Newly Synthesized TiO2-SiO2 Photocatalyst. Catalysts 2022, 12, 156. [Google Scholar] [CrossRef]
  137. Li, L.; Chen, X.H.; Xiong, X.; Wu, X.P.; Xie, Z.A.; Liu, Z.H. Synthesis of hollow TiO2@SiO2 spheres via a recycling template method for solar heat protection coating. Ceram. Int. 2021, 47, 2678–2685. [Google Scholar] [CrossRef]
  138. Shi, J.; Chen, Z.; Wang, B.; Wang, L.; Lu, T.; Zhang, Z. Reactive Oxygen Species-Manipulated Drug Release from a Smart Envelope-Type Mesoporous Titanium Nanovehicle for Tumor Sonodynamic-Chemotherapy. ACS Appl. Mater. Interfaces 2015, 7, 28554–28565. [Google Scholar] [CrossRef]
  139. Meng, L.; Liu, Z.H.; Lan, C.W.; Xu, N. In-Situ Fabricating Ag Nanoparticles on TiO2 for Unprecedented High Catalytic Activity of 4-Nitrophenol Reduction. Catal. Lett. 2022, 152, 912–920. [Google Scholar] [CrossRef]
  140. Motamedisade, A.; Johnston, M.R.; Alotaibi, A.E.H.; Andersson, G.A. Au9 nanocluster adsorption and agglomeration control through sulfur modification of mesoporous TiO2. Phys. Chem. Chem. Phys. 2024, 26, 9500–9509. [Google Scholar] [CrossRef]
  141. Mirzaei, M.; Zarch, M.B.; Darroudi, M.; Sayyadi, K.; Keshavarz, S.T.; Sayyadi, J.; Fallah, A.; Maleki, H. Silica Mesoporous Structures: Effective Nanocarriers in Drug Delivery and Nanocatalysts. Appl. Sci. 2020, 10, 7533. [Google Scholar] [CrossRef]
  142. Mabrouk, M.; Moaness, M.; Beherei, H.H. Fabrication of mesoporous zirconia and titania nanomaterials for bone regeneration and drug delivery applications. J. Drug Deliv. Sci. Technol. 2022, 78, 103957. [Google Scholar] [CrossRef]
  143. Liu, D.; Bi, Y.-G. Controllable fabrication of hollow TiO2 spheres as sustained release drug carriers. Adv. Powder Technol. 2019, 30, 2169–2177. [Google Scholar] [CrossRef]
  144. Mai, Z.; Chen, J.; Cao, Q.; Hu, Y.; Dong, X.; Zhang, H.; Huang, W.; Zhou, W. Rational design of hollow mesoporous titania nanoparticles loaded with curcumin for UV-controlled release and targeted drug delivery. Nanotechnology 2021, 32, 205604. [Google Scholar] [CrossRef] [PubMed]
  145. Chen, T.R.; Wang, Y.Y.; Zhang, P.; Alomar, T.S.; Wang, G.Q.; Gao, Y.A.; Liu, M.; AlMasoud, N.; El-Bahy, Z.M.; Ren, J.N.; et al. Targeted synthesis of hollow titania microspheres with sustained release behaviors of 1,2-benzisothiazolin-3-one (BIT) for good marine antifouling performance. Adv. Compos. Hybrid Mater. 2023, 6, 151. [Google Scholar] [CrossRef]
  146. Ye, J.; Liu, W.; Cai, J.; Chen, S.; Zhao, X.; Zhou, H.; Qi, L. Nanoporous anatase TiO2 mesocrystals: Additive-free synthesis, remarkable crystalline-phase stability, and improved lithium insertion behavior. J. Am. Chem. Soc. 2011, 133, 933–940. [Google Scholar] [CrossRef]
  147. Pakdel, E.; Zhao, H.; Wang, J.; Tang, B.; Varley, R.J.; Wang, X. Superhydrophobic and photocatalytic self-cleaning cotton fabric using flower-like N-doped TiO2/PDMS coating. Cellulose 2021, 28, 8807–8820. [Google Scholar] [CrossRef]
  148. Wang, H.-X.; Li, X.-X.; Tang, L. Effects of surfactants on the morphology and properties of TiO2. Appl. Phys. A 2020, 126, 448. [Google Scholar] [CrossRef]
  149. Pakdel, E.; Wang, J.; Allardyce, B.J.; Rajkhowa, R.; Wang, X. Functionality of nano and 3D-microhierarchical TiO2 particles as coagulants for sericin extraction from the silk degumming wastewater. Sep. Purif. Technol. 2016, 170, 92–101. [Google Scholar] [CrossRef]
  150. Chao, C.S.; Liu, K.H.; Tung, W.L.; Chen, S.Y.; Liu, D.M.; Chang, Y.P. Bioactive TiO2 ultrathin film with worm-like mesoporosity for controlled drug delivery. Microporous Mesoporous Mater. 2012, 152, 58–63. [Google Scholar] [CrossRef]
  151. Seyed-Talebi, S.M.; Kazeminezhad, I.; Motamedi, H. TiO2 hollow spheres as a novel antibiotic carrier for the direct delivery of gentamicin. Ceram. Int. 2018, 44, 13457–13462. [Google Scholar] [CrossRef]
  152. Zhang, H.J.; Wang, C.L.; Chen, B.A.; Wang, X.M. Daunorubicin-TiO2 nanocomposites as a “smart” pH-responsive drug delivery system. Int. J. Nanomed. 2012, 7, 235–242. [Google Scholar] [CrossRef]
  153. Malekimusavi, H.; Ghaemi, A.; Masoudi, G.; Chogan, F.; Rashedi, H.; Yazdian, F.; Omidi, M.; Javadi, S.; Haghiralsadat, B.F.; Teimouri, M.; et al. Graphene oxide-l-arginine nanogel: A pH-sensitive fluorouracil nanocarrier. Biotechnol. Appl. Biochem. 2019, 66, 772–780. [Google Scholar] [CrossRef] [PubMed]
  154. Wang, Y.L.; Yuan, L.L.; Yao, C.J.; Fang, J.; Wu, M.H. Cytotoxicity Evaluation of pH-Controlled Antitumor Drug Release System of Titanium Dioxide Nanotubes. J. Nanosci. Nanotechnol. 2015, 15, 4143–4148. [Google Scholar] [CrossRef]
  155. Jia, H.Y.; Kerr, L.L. Kinetics of Drug Release from Drug Carrier of Polymer/TiO2 Nanotubes Composite-pH Dependent Study. J. Appl. Polym. Sci. 2015, 132, 41570. [Google Scholar] [CrossRef]
  156. Pour, M.M.; Moghbeli, M.R.; Larijani, B.; Javar, H.A. pH-Sensitive mesoporous bisphosphonate-based TiO2 nanoparticles utilized for controlled drug delivery of dexamethasone. Chem. Pap. 2022, 76, 439–451. [Google Scholar] [CrossRef]
  157. Han, J.; Jang, E.K.; Ki, M.R.; Son, R.G.; Kim, S.; Choe, Y.; Pack, S.P.; Chung, S. pH-responsive phototherapeutic poly(acrylic acid)-calcium phosphate passivated TiO2 nanoparticle-based drug delivery system for cancer treatment applications. J. Ind. Eng. Chem. 2022, 112, 258–270. [Google Scholar] [CrossRef]
  158. Yan, S.J.; Shi, H.C.; Song, L.J.; Wang, X.H.; Liu, L.; Luan, S.F.; Yang, Y.M.; Yin, J.H. Nonleaching Bacteria-Responsive Antibacterial Surface Based on a Unique Hierarchical. Acs Appl. Mater. Interfaces 2016, 8, 24471–24481. [Google Scholar] [CrossRef]
  159. Chen, J.; Shi, X.; Zhu, Y.; Chen, Y.; Gao, M.; Gao, H.; Liu, L.; Wang, L.; Mao, C.; Wang, Y. On-demand storage and release of antimicrobial peptides using Pandora’s box-like nanotubes gated with a bacterial infection-responsive polymer. Theranostics 2020, 10, 109–122. [Google Scholar] [CrossRef]
  160. Pourmadadi, M.; Yazdian, F.; Koulivand, A.; Rahmani, E. Green synthesized polyvinylpyrrolidone/titanium dioxide hydrogel nanocomposite modified with agarose macromolecules for sustained and pH-responsive release of anticancer drug. Int. J. Biol. Macromol. 2023, 240, 124345. [Google Scholar] [CrossRef]
  161. Zhang, H.J.; Ji, Y.D.; Chen, Q.Q.; Zhu, X.; Zhang, X.G.; Tan, Z.Y.; Tian, Q.Q.; Yang, X.B.; Zhang, Z.Z. Invitro and invivo chemo-phototherapy of magnetic TiO2 drug delivery system formed by pH-sensitive coordination bond. J. Biomater. Appl. 2016, 31, 568–581. [Google Scholar] [CrossRef] [PubMed]
  162. Bagwasi, S.; Tian, B.Z.; Zhang, J.L.; Nasir, M. Synthesis, characterization and application of bismuth and boron Co-doped TiO2: A visible light active photocatalyst. Chem. Eng. J. 2013, 217, 108–118. [Google Scholar] [CrossRef]
  163. Low, W.; Boonamnuayvitaya, V. Enhancing the photocatalytic activity of TiO2 co-doping of graphene-Fe3+ ions for formaldehyde removal. J. Environ. Manag. 2013, 127, 142–149. [Google Scholar] [CrossRef] [PubMed]
  164. Al-Nemrawi, N.K.; Alshraiedeh, N.A.H.; Zayed, A.L.; Altaani, B.M. Low Molecular Weight Chitosan-Coated PLGA Nanoparticles for Pulmonary Delivery of Tobramycin for Cystic Fibrosis. Pharmaceuticals 2018, 11, 28. [Google Scholar] [CrossRef] [PubMed]
  165. Al-Nemrawi, N.K.; Alsharif, S.S.M.; Alzoubi, K.H.; Alkhatib, R.Q. Preparation and characterization of insulin chitosan-nanoparticles loaded in buccal films. Pharm. Dev. Technol. 2019, 24, 967–974. [Google Scholar] [CrossRef]
  166. Al-Nemrawi, N.; Hameedat, F.; Al-Husein, B.; Nimrawi, S. Photolytic Controlled Release Formulation of Methotrexate Loaded in Chitosan/TiO2 Nanoparticles for Breast Cancer. Pharmaceuticals 2022, 15, 149. [Google Scholar] [CrossRef]
  167. Sargazi, S.; Er, S.; Sacide Gelen, S.; Rahdar, A.; Bilal, M.; Arshad, R.; Ajalli, N.; Farhan Ali Khan, M.; Pandey, S. Application of titanium dioxide nanoparticles in photothermal and photodynamic therapy of cancer: An updated and comprehensive review. J. Drug Deliv. Sci. Technol. 2022, 75, 103605. [Google Scholar] [CrossRef]
  168. Shen, Y.; Xie, C.; Xiao, X. Black phosphorus-incorporated titanium dioxide nanotube arrays for near-infrared–triggered drug delivery. J. Drug Deliv. Sci. Technol. 2022, 72, 103400. [Google Scholar] [CrossRef]
  169. Chen, B.; Liang, Y.; Song, Y.; Liang, Y.; Jiao, J.; Bai, H.; Li, Y. Photothermal-Controlled Release of IL-4 in IL-4/PDA-Immobilized Black Titanium Dioxide (TiO2) Nanotubes Surface to Enhance Osseointegration: An In Vivo Study. Materials 2022, 15, 5962. [Google Scholar] [CrossRef]
  170. Song, Y.Y.; Hildebrand, H.; Schmuki, P. Photoinduced release of active proteins from TiO2 surfaces. Electrochem. Commun. 2009, 11, 1429–1433. [Google Scholar] [CrossRef]
  171. Kumar, A.; Agarwala, V.; Singh, D. Microwave absorbing behavior of metal dispersed TiO2 nanocomposites. Adv. Powder Technol. 2014, 25, 483–489. [Google Scholar] [CrossRef]
  172. Liu, J.W.; Che, R.C.; Chen, H.J.; Zhang, F.; Xia, F.; Wu, Q.S.; Wang, M. Microwave Absorption Enhancement of Multifunctional Composite Microspheres with Spinel Fe3O4 Cores and Anatase TiO2 Shells. Small 2012, 8, 1214–1221. [Google Scholar] [CrossRef] [PubMed]
  173. Peng, H.; Cui, B.; Zhao, W.; Wang, Y.; Chang, Z.; Wang, Y. Glycine-functionalized Fe3O4@TiO2:E3+,Yb3+ nanocarrier for microwave-triggered controllable drug release and study on mechanism of loading release process using microcalorimetry. Expert Opin. Drug Deliv. 2015, 12, 1397–1409. [Google Scholar] [CrossRef] [PubMed]
  174. Liu, Y.; Si, Y.Y.; Di, M.Y.; Tang, D.J.; Meng, L.; Cui, B. A novel microwave stimulus remote-controlled anticancer drug release system based on Janus TiO2−x&mSiO2 nanocarriers. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 123, 111968. [Google Scholar]
  175. Liu, Y.; Hui, Z.; Zhan, Z.; Cui, L.; Liu, X.; Cui, B. “Biped” Janus Fe3O4@nSiO2@TiO2−x&mSiO2 Nanoparticles for Drug Delivery and Microwave-Triggered Drug Release. Nano 2023, 18, 2350057. [Google Scholar]
  176. Deckers, R.; Rome, C.; Moonen, C.T.W. The role of ultrasound and magnetic resonance in local drug delivery. J. Magn. Reson. Imaging 2008, 27, 400–409. [Google Scholar] [CrossRef]
  177. Ye, L.Z.; Zhu, X.J.; Liu, Y. Numerical study on dual-frequency ultrasonic enhancing cavitation effect based on bubble dynamic evolution. Ultrason. Sonochem. 2019, 59, 104744. [Google Scholar] [CrossRef]
  178. Zhou, J.; Frank, M.A.; Yang, Y.; Boccaccini, A.R.; Virtanen, S. A novel local drug delivery system: Superhydrophobic titanium oxide nanotube arrays serve as the drug reservoir and ultrasonication functions as the drug release trigger. Mater. Sci. Eng. C 2018, 82, 277–283. [Google Scholar] [CrossRef]
  179. Rehman, F.U.; Rauf, M.A.; Ullah, S.; Shaikh, S.; Qambrani, A.; Muhammad, P.; Hanif, S. Ultrasound-activated nano-TiO2 loaded with temozolomide paves the way for resection of chemoresistant glioblastoma multiforme. Cancer Nanotechnol. 2021, 12, 17. [Google Scholar] [CrossRef]
  180. Wang, Z.; Wang, M.; Qian, Y.; Xie, Y.; Sun, Q.; Gao, M.; Li, C. Dual-targeted nanoformulation with Janus structure for synergistic enhancement of sonodynamic therapy and chemotherapy. Chin. Chem. Lett. 2023, 34, 107853. [Google Scholar] [CrossRef]
  181. Jing, Y.; Zhu, Y.; Yang, X.; Shen, J.; Li, C. Ultrasound-Triggered Smart Drug Release from Multifunctional Core−Shell Capsules One-Step Fabricated by Coaxial Electrospray Method. Langmuir 2010, 27, 1175–1180. [Google Scholar] [CrossRef] [PubMed]
  182. Zhao, Q.; Wang, C.; Liu, Y.; Wang, J.; Gao, Y.; Zhang, X.; Jiang, T.; Wang, S. PEGylated mesoporous silica as a redox-responsive drug delivery system for loading thiol-containing drugs. Int. J. Pharm. 2014, 477, 613–622. [Google Scholar] [CrossRef] [PubMed]
  183. Wu, Y.; Xu, Z.; Sun, W.; Yang, Y.; Jin, H.; Qiu, L.; Chen, J.; Chen, J. Co-responsive smart cyclodextrin-gated mesoporous silica nanoparticles with ligand-receptor engagement for anti-cancer treatment. Mater. Sci. Eng. C 2019, 103, 109831. [Google Scholar] [CrossRef] [PubMed]
  184. Luo, Z.; Hu, Y.; Cai, K.; Ding, X.; Zhang, Q.; Li, M.; Ma, X.; Zhang, B.; Zeng, Y.; Li, P.; et al. Intracellular redox-activated anticancer drug delivery by functionalized hollow mesoporous silica nanoreservoirs with tumor specificity. Biomaterials 2014, 35, 7951–7962. [Google Scholar] [CrossRef] [PubMed]
  185. Liu, J.; Li, Y.; Zhao, M.; Lei, Z.; Guo, H.; Tang, Y.; Yan, H. Redox-responsive hollow mesoporous silica nanoparticles constructed via host–guest interactions for controllable drug release. J. Biomater. Sci. Polym. Ed. 2019, 31, 472–490. [Google Scholar] [CrossRef]
  186. Malekmohammadi, S.; Hadadzadeh, H.; Rezakhani, S.; Amirghofran, Z. Design and Synthesis of Gatekeeper Coated Dendritic Silica/Titania Mesoporous Nanoparticles with Sustained and Controlled Drug Release Properties for Targeted Synergetic Chemo-Sonodynamic Therapy. Acs Biomater. Sci. Eng. 2019, 5, 4405–4415. [Google Scholar] [CrossRef]
  187. Hu, X.X.; Hao, X.H.; Wu, Y.; Zhang, J.C.; Zhang, X.N.; Wang, P.C.; Zou, G.Z.; Liang, X.J. Multifunctional hybrid silica nanoparticles for controlled doxorubicin loading and release with thermal and pH dual response. J. Mater. Chem. B 2013, 1, 1109–1118. [Google Scholar] [CrossRef]
  188. Timin, A.S.; Muslimov, A.R.; Lepik, K.V.; Saprykina, N.N.; Sergeev, V.S.; Afanasyev, B.V.; Vilesov, A.D.; Sukhorukov, G.B. Triple-responsive inorganic-organic hybrid microcapsules as a biocompatible smart platform for the delivery of small molecules. J. Mater. Chem. B 2016, 4, 7270–7282. [Google Scholar] [CrossRef]
  189. Innocenzi, P.; Malfatti, L. Mesoporous ordered titania films: An advanced platform for photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2024, 58, 100646. [Google Scholar] [CrossRef]
  190. Yu, S.; Mu, Y.; Zhang, X.; Li, J.; Lee, C.; Wang, H. Molecular mechanisms underlying titanium dioxide nanoparticles (TiO2 NP) induced autophagy in mesenchymal stem cells (MSC). J. Toxicol. Environ. Health Part A 2019, 82, 997–1008. [Google Scholar] [CrossRef]
  191. Ren, Y.; Feng, X.; Lang, X.; Wang, J.; Du, Z.; Niu, X.; Li, X. Evaluation of Osteogenic Potentials of Titanium Dioxide Nanoparticles with Different Sizes and Shapes. J. Nanomater. 2020, 2020, 8887323. [Google Scholar] [CrossRef]
  192. Kose, O.; Tomatis, M.; Leclerc, L.; Belblidia, N.-B.; Hochepied, J.-F.; Turci, F.; Pourchez, J.; Forest, V. Impact of the Physicochemical Features of TiO2 Nanoparticles on Their In Vitro Toxicity. Chem. Res. Toxicol. 2020, 33, 2324–2337. [Google Scholar] [CrossRef] [PubMed]
  193. Danielsen, P.H.; Knudsen, K.B.; Štrancar, J.; Umek, P.; Koklič, T.; Garvas, M.; Vanhala, E.; Savukoski, S.; Ding, Y.; Madsen, A.M.; et al. Effects of physicochemical properties of TiO2 nanomaterials for pulmonary inflammation, acute phase response and alveolar proteinosis in intratracheally exposed mice. Toxicol. Appl. Pharmacol. 2020, 386, 114830. [Google Scholar] [CrossRef] [PubMed]
  194. Padin-Gonzalez, E.; Lancaster, P.; Bottini, M.; Gasco, P.; Tran, L.; Fadeel, B.; Wilkins, T.; Monopoli, M.P. Understanding the Role and Impact of Poly (Ethylene Glycol) (PEG) on Nanoparticle Formulation: Implications for COVID-19 Vaccines. Front. Bioeng. Biotechnol. 2022, 10, 882363. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of steps involved in sol-gel process.
Figure 1. Schematic diagram of steps involved in sol-gel process.
Pharmaceutics 16 01214 g001
Figure 2. Schematic diagram of the preparation process of nanoparticles via the green synthesis method.
Figure 2. Schematic diagram of the preparation process of nanoparticles via the green synthesis method.
Pharmaceutics 16 01214 g002
Figure 3. Functionalization of TiO2 nanoparticles.
Figure 3. Functionalization of TiO2 nanoparticles.
Pharmaceutics 16 01214 g003
Figure 4. Schematic diagram of drug-controlled-release delivery systems with different responses.
Figure 4. Schematic diagram of drug-controlled-release delivery systems with different responses.
Pharmaceutics 16 01214 g004
Figure 5. Schematic diagram of drug release process triggered by stimulation.
Figure 5. Schematic diagram of drug release process triggered by stimulation.
Pharmaceutics 16 01214 g005
Table 1. Advantages and disadvantages of different types of inorganic nanomaterials.
Table 1. Advantages and disadvantages of different types of inorganic nanomaterials.
TypeAdvantagesDisadvantagesApplication
TiO2Easy surface functionalization, antibacterial properties, photocatalytic degradation propertiesWide band gap, fast hydrolysis rate and lack of biosafety evaluationNano-biosensing, medical implantation, drug delivery, and antimicrobial agents
SiO2Low toxicity, easy surface functionalizationLack of biosafety evaluationBioimaging, drug delivery
Fe3O4Magnetic, biodegradabilityEasy to aggregate and oxidizeBiological separation and detection, magnetic resonance imaging, and magnetic hyperthermia
Table 2. Characteristics of different synthesis methods.
Table 2. Characteristics of different synthesis methods.
MethodsClassificationCharacteristic
Sol-gel method-Simple, fast, economically less expensive, low processing temperature, homogeneity of the produced material
Hydrothermal method-High purity, good dispersibility, low safety and high energy consumption
Template methodHard template methodSimple operation and various structures
Soft template method
Gas-state methodPhysical vapor depositionHigh purity, uniform distribution, small particle size and good dispersion, expensive
Chemical vapor depositionHigh purity, high reaction temperature
Solid-state method-Easy to operate, short processing time, impurity, incomplete morphology and inhomogeneous particle size
Green synthesis method-Naturally adaptable, environmentally friendly and cost-effective
Table 3. TiO2 prepared by utilizing a variety of plants.
Table 3. TiO2 prepared by utilizing a variety of plants.
S/NPlant ExtractReactantShapeAverage Particle Size (nm)ApplicationRef.
1Tinospora cordifoliaTitanium (IV) isopropoxideTriangular and irregularly shape7–21Photocatalysis[94]
2Citrus LimettaTitanium butoxide, waterSpherical80–100Photocatalysis[95]
3Syzygium cuminiTitanium-isopropoxide, waterSpherical15–22Photocatalysis[96]
4Luffa acutangulaTitanium sulfate, waterAggregated10–49Antibacterial[97]
5Mentha arvensisTitanium tetra-isopropoxide, ethanolSpherical20–70Antibacterial[98]
6Acorus calamusTitanium (IV) isopropoxide, water, aqueous ammoniaSpherical and interconnected11–30Photocatalysis and antibacterial[99]
7Citrus limonTitanium (IV) butoxide, isopropanol, glacial acetic acidSpherical9–18Dye-sensitized solar cells[100]
8Curcuma Longa L.Methanol, titanium tetra-isopropoxide, waterSpherical, cubic and hexagonal20.8–40.1Photocatalysis[101]
9Phyllanthus niruriTitanium isopropoxide, waterSpherical20Dye adsorption[102]
10Laurus nobilisTitanium tetra isopropoxide, waterSpherical80–120Antibacterial and antioxidant[103]
Table 4. TiO2 produced by several bacterial communities.
Table 4. TiO2 produced by several bacterial communities.
S/NBacterial SpeciesReactantShapeAverage Particle Size (nm)ApplicationRef.
1Aeromonas hydrophilaMetatitanic acid (TiO (OH)2)Spherical40–50Antibacterial[104]
2LactobacilliTiO (OH)2Irregular10–25-[105]
3Bacillus subtilisTiO (OH)2, waterSpherical66–77-[106]
4Bacillus subtilisTi4+ ions, waterSpherical10–30Photocatalysis[92]
5Alcaligenes aquatilisK2TiF6, water, silver nitrate solutionSpherical10–90Photocatalysis[107]
6Fusarium oxysporumAqueous solution, (CH3COO)2Ba, K2TiF6Spherical10-[108]
Table 5. TiO2 drug delivery systems with different response types.
Table 5. TiO2 drug delivery systems with different response types.
TiO2-TypeSynthesis MethodDrugStimulusRef.
MTNPolymer sacrificial methodAmoxicillinSustained release[142]
HMTNHydrothermal methodDoxorubicinSustained release[143]
HMTNStöber methodCurcuminUV[144]
HMTNSurface-layer-absorption template1,2-benzisothiazolin-3-oneSustained release[145]
plush TiO2Hydrothermal methodDoxorubicinSustained release[45]
TiO2 filmEISA methodIbuprofen and vancomycinSustained release[150]
HMTNSol-gel methodGentamycinSustained release[151]
DNR-TiO2-DaunorubicinpH[152]
PDA-TNTs-IbuprofenSustained release[22]
GPTMS-TNTsHydrothermal methodDexamethasoneSustained release[120]
TNTs-DoxorubicinpH[154]
TNTsTwo-step anodization methodMethotrexatepH[155]
MBTNPsTemplate methodDexamethasonepH[156]
TiO2@PAA-CaP NPsHydrothermal methodDoxorubicinpH[157]
PVP-Ag-TiO2-DoxorubicinpH[160]
TiO2@Fe3O4-PEISol-gel methodDoxorubicinpH[161]
CS-NPsIonic gelation methodMethotrexateLight[166]
BP-TNTs-IbuprofenNIR[168]
BP-TNTsElectrochemical reduction methodIL-4NIR[169]
TiO2-ProteinsUV[170]
Fe3O4@TiO2Hydrothermal methodEtoposideMicrowave[173]
TiO2-x&mSiO2Sol-gel methodDoxorubicinMicrowave and pH[174]
Fe3O4@nSiO2@TiO2-x&mSiO2-DoxorubicinMicrowave[175]
Superhydrophobic -TNTsTetracycline hydrochlorideElectrochemical anoDization methodUltrasound[178]
TiO2 nanosticks-GlioblasToma multiformeUltrasound[179]
TiO2-x @NaGdF4-IR780 iodineUltrasound[180]
TiO2 Core-Shell CapsulesCoaxial electrospray methodPaclitaxelUltrasound[181]
PEI-FA-DSTNs-CurcuminUltrasound[186]
MTN-CDSol-gel methodDocetaxelUltrasound[138]
SiO2/TiO2Template methodRh-BUV, ultrasound and enzymatic treatment[188]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zuo, F.; Zhu, Y.; Wu, T.; Li, C.; Liu, Y.; Wu, X.; Ma, J.; Zhang, K.; Ouyang, H.; Qiu, X.; et al. Titanium Dioxide Nanomaterials: Progress in Synthesis and Application in Drug Delivery. Pharmaceutics 2024, 16, 1214. https://doi.org/10.3390/pharmaceutics16091214

AMA Style

Zuo F, Zhu Y, Wu T, Li C, Liu Y, Wu X, Ma J, Zhang K, Ouyang H, Qiu X, et al. Titanium Dioxide Nanomaterials: Progress in Synthesis and Application in Drug Delivery. Pharmaceutics. 2024; 16(9):1214. https://doi.org/10.3390/pharmaceutics16091214

Chicago/Turabian Style

Zuo, Fanjiao, Yameng Zhu, Tiantian Wu, Caixia Li, Yang Liu, Xiwei Wu, Jinyue Ma, Kaili Zhang, Huizi Ouyang, Xilong Qiu, and et al. 2024. "Titanium Dioxide Nanomaterials: Progress in Synthesis and Application in Drug Delivery" Pharmaceutics 16, no. 9: 1214. https://doi.org/10.3390/pharmaceutics16091214

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