Next Article in Journal
Comparison of Carbon-Nanoparticle-Filled Poly(Butylene Succinate-co-Adipate) Nanocomposites for Electromagnetic Applications
Next Article in Special Issue
On the Application of Calcium Phosphate Micro- and Nanoparticles as Food Additive
Previous Article in Journal
Photocatalytic and Antibacterial Activity of CoFe2O4 Nanoparticles from Hibiscus rosa-sinensis Plant Extract
Previous Article in Special Issue
Superhydrophilic Nanotextured Surfaces for Dental Implants: Influence of Early Saliva Contamination and Wet Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Medical and Dental Applications of Titania Nanoparticles: An Overview

1
Department of Dental Material Sciences, School of Dentistry, Shaheed Zulfiqar Ali Bhutto Medical University, Islamabad 44080, Pakistan
2
Department of Prosthodontics and Dental Implantology, College of Dentistry, King Faisal University, Al-Ahsa 31982, Saudi Arabia
3
Department of Dental Biomaterials, Bakhtawar Amin Medical and Dental College, Multan 60650, Pakistan
4
Islamic International Dental College, Riphah International University, Islamabad 44000, Pakistan
5
Department of Materials Engineering, NED University of Engineering & Technology, Karachi 74200, Pakistan
6
Department of Microbiology, Quaid-i-Azam University, Islamabad 45320, Pakistan
7
Center for Innovation and Research in Oral Sciences (CIROS), Faculty of Medicine, University of Coimbra, 3000-075 Coimbra, Portugal
8
Institute of Endodontics, Faculty of Medicine, University of Coimbra, 3000-075 Coimbra, Portugal
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(20), 3670; https://doi.org/10.3390/nano12203670
Submission received: 29 September 2022 / Revised: 8 October 2022 / Accepted: 9 October 2022 / Published: 19 October 2022
(This article belongs to the Special Issue Nanomaterials in Oral Science)

Abstract

:
Currently, titanium oxide (TiO2) nanoparticles are successfully employed in human food, drugs, cosmetics, advanced medicine, and dentistry because of their non-cytotoxic, non-allergic, and bio-compatible nature when used in direct close contact with the human body. These NPs are the most versatile oxides as a result of their acceptable chemical stability, lower cost, strong oxidation properties, high refractive index, and enhanced aesthetics. These NPs are fabricated by conventional (physical and chemical) methods and the latest biological methods (biological, green, and biological derivatives), with their advantages and disadvantages in this epoch. The significance of TiO2 NPs as a medical material includes drug delivery release, cancer therapy, orthopedic implants, biosensors, instruments, and devices, whereas their significance as a dental biomaterial involves dentifrices, oral antibacterial disinfectants, whitening agents, and adhesives. In addition, TiO2 NPs play an important role in orthodontics (wires and brackets), endodontics (sealers and obturating materials), maxillofacial surgeries (implants and bone plates), prosthodontics (veneers, crowns, bridges, and acrylic resin dentures), and restorative dentistry (GIC and composites).

1. Introduction

The advent of nanotechnology has led to a substantial surge in the applications of NPs for biomaterials, antibacterial substances, drug delivery systems, electronic equipment, sunscreens, and cosmetics [1,2]. Nanomaterials have previously been defined as natural, manufactured, or incidental materials that include particles as agglomerates or aggregates or in an unconfined form where 50% or a higher number of the particles depict one or more outer dimensions within the range of 1–100 nm [3]. Moreover, to consider a material as a nanoparticle one of its three dimensions should be in a range of 1–100 nm [4,5]. Unique chemical and physical properties are demonstrated by oxide NPs due to their limited size and their corner surface sites possessing higher density [6]. The physicochemical properties of fine particles (FPs) vary as compared to the NPs (NPs) that have a similar composition.
The TiO2 fine particles (FPs) are poorly soluble and have a low toxicity [7,8]. The TiO2 NPs (NPs) have an enhanced catalytic activity as compared to the fine particles (FPs), which consequently leads to their higher utilization in consumer and industrial products. The properties of the TiO2 NPs may pose a challenge to human health and are responsible for the distinctive bioactivity of these NPs [9,10]. The TiO2 NPs are prominent and versatile oxides that have a higher production due to their acceptable chemical stability, lower cost, strong oxidation properties, high refractive index, and formidable oxidation properties [11]. The TiO2 NPs have found their optical industrial applications as semiconductors due to having the critical property of the wide band gap [12,13]. They have also been used in electronic devices and as sensors due to their particular ionic and electrical properties [14]. As these NPs exist as a water insoluble powder and are white in color, they can be utilized in the paint industries as pigments [15]. Three polymorphic forms of the TiO2 NPs exist naturally; they are the rutile, brookite, and anatase forms with crystalline structures, and the gemstone industries have utilized them extensively [14]. The TiO2 NPs have been employed for decomposing water waste pollutants as photocatalysts, as food additives, and as antibacterial and antimicrobial agents [16,17,18,19]. The TiO2 NPs can help in destroying fungi, viruses, bacteria, and cancer cells [20]. The TiO2 NPs have been used in sonodynamic and photodynamic therapy (PDT) for the treatment of cancer [21]. The NPs have the ability to convert photon energy into heat through photo thermal therapy, via their super hydrophilicity, in vivo chemical and thermal stability, low toxicity, optical absorption, and thermal conductivity [22]. The nanostructure of TiO2 has been used for drug delivery systems in various anticancer drugs, such as temozolomide, cisplatin, doxorubicin, and daunorubicin [23,24,25]. Previously, polyethylene glycol-coated NPs have been used for the treatment of melanoma through photothermal therapy [26]. The (TiO2) NPs have been used in dental applications due to their ability to enhance the mechanical properties of biomaterials without altering their biocompatibility [27,28,29]. Previous studies have utilized the TiO2 NPs in order to improve the bond strength and antimicrobial properties of composites [30]. The TiO2 NPs have also been used to improve the working time, setting time, push-out bond strength, and compressive strength of cements [31]. The mechanical properties of the resins have also been optimized by incorporating titania NPs into them [32,33]. Moreover, silver-doped titania NPs have also been incorporated into dental polymers in order to produce potent bactericidal effects [34].

2. Background of Nanotechnology

Nanotechnology has been considered as a significant discipline of the latest research involving the synthesis protocols and the designing and manipulating techniques of the particle’s structure in different dimensions in the range of about 1–100 nm. Moreover, this emerging nanotechnology has specifically opened vast doors for the synthesis of nano-scaled materials and the exploration of their unique electronic, optical, physicochemical, strength, magnetic, electrical, and mechanical characteristics, together with the utilization of this rapidly growing novel technology in certain various fundamental frontiers. These fundamental areas that involve nanotechnology are electronics, multimedia, cosmetics, medicine, engineering, electronics, industries, food, environment, robotics, optics, machines, catalysts, drug delivery systems, genomics, biomedicines, material sciences, energy sciences, light emitters, solar energy processes, imaging technology, physics, chemistry, electromagnetic devices, electron transistors, energy-producing plants, water treatment plants, solar energy conversion, etc. [35,36]. Currently, nanotechnology has gained a popular status in multiple health systems, such as control, prevention, diagnosis, monitoring, and treatment; this has happened quite rapidly because of its unique and high-quality properties [37,38]. The word “nano” is taken from the Greek word nanos, meaning “dwarf”, which factually refers to 1 billionth of physical size. Furthermore, it can be inferred that the area of nanotechnology works at the level of atoms and molecules ranging between 1 and 100 nm in size [39].

2.1. History of Nanotechnology

NPs were used by artisans as far back as the 9th century in Mesopotamia for generating a glittering effect on the surface of pots. Pottery from the Middle Ages and the Renaissance often retains a distinct gold or copper-colored metallic luster [40]. The luster originated within the film itself, which contained silver and copper NPs dispersed homogeneously in the glassy matrix of the ceramic glaze [41]. The concepts that seeded nanotechnology were first discussed in 1959 by renowned physicist Richard Feynman in his talk “There’s Plenty of Room at the Bottom”, in which he described the possibility of synthesis via the direct manipulation of atoms. The term “nanotechnology” was first used by Norio Taniguchi in 1974, though it was not widely known at that time [38,40].

2.2. Significant Features of Nanotechnology

There are three major features that play a vital role in defining the versatility and superiority of nanotechnology in comparison to normal technology [42]:
  • This technology can bring changes in the original matter’s structure by converting it into a nanoscale product.
  • The newly formed nanoscale structure should contain periodic repetition (i.e., the nanoparticle must periodically repeat itself in one or more than one direction) just like the parent matter.
  • The novel properties, characteristics, and functions of a newly formed nanoscale product should be like those of the parent matter or be superior to the parent matter despite being nanometric.

2.3. Classification of NPs

NPs are categorized generally as organic, inorganic, and carbon-based in origin, as illustrated in (Figure 1). The carbon-based NPs are entirely made of carbon. The organic NPs are found to be non-toxic, biodegradable, and ideal for drug delivery. The NPs that do not consume carbon during their synthesis are called inorganic NPs and include metal and metal oxide-based NPs. The metal oxide-based NPs have extraordinary properties, such as their sizes (ranging between 10 and 100 nm); shapes (spherical, cylindrical, etc.); surface characteristics (pore size, high surface-area-to-volume ratio, surface charge dentistry, and surface charge); structures (crystalline and amorphous); color; and sensitivity to various environmental factors (heat, air sunlight, moisture, etc.) (Figure 2) [43,44].

3. The Titanium Oxide NPs

These NPs are commonly named as TiO2 or titania because they are oxides of metals [47,48,49]. Titania is available in the form of fine, white powder and/or thin film [22,50]. Naturally, there are four crystal forms of titanium oxide: TiO2 (B), brookite, anatase, and rutile [51]. However, the mixture of anatase and the rutile crystal phase form of the TiO2 NPs, which usually exists, is more inert and stable in comparison to the anatase crystal phase form, which displays more chemical reactivity. This chemical reactivity of the anatase crystal phase form could be beneficial in eliminating cancer cells efficiently [52,53].
The TiO2 NPs are famous for their well-known attractive properties, including magnetic, optical, thermal, physicochemical, electrical, rheological, mechanical, and biological properties. These NPs are highly suitable for utilization in various fields under normal conditions because of their relatively low cost, insoluble nature, and high refractive index (n = 2.4). They are absolutely extra white, which enhances their value in multiple different fields [47,48,49]. Moreover, TiO2 NPs are widely used in anti-flavor textiles, water disinfectants, food supplements, cosmetics, indoor sprays, anti-flavor textiles, household appliances, paper, plastics, food packaging materials, sunscreens, electronic products, medicines, fibers, bakery products, sensors, glazes, and coloring agents in foods due to their smaller size [54,55,56].

3.1. Basic Requirements for Employing TiO2 NPs in Medical and Dental Applications

The biocompatibility of TiO2 NPs in the medical and dental applications is the most important requirement of these NPs and is dependent on two main factors, which include its synthesis protocols and its characteristic features attained by characterization techniques.

3.1.1. Synthesis Protocols of TiO2 NPs

The synthesis of NPs is carried out by two main techniques, referred to as the conventional (top-down) and biocompatible (bottom-up) methods. The physical and chemical methods of synthesis are deployed in the conventional (top-down) techniques, whereas the biological, green, and biological derivative methods are employed in the biocompatible (bottom-up) techniques to a large extent (Figure 3) [57,58].

Conventional (Top-Down) Synthesis

The top-down technique involves the breakdown of bulk material in nano-sized particles through physical methods involving thermal decomposition, sputtering, laser ablation, nanolithography, and milling [59], whereas the chemical techniques include: co-precipitation, laser pyrolysis, the sol-gel process [60], spray pyrolysis, chemical vapor deposition [61], the aerosol process, reverse micelle [62], atomic/molecular condensation, the microemulsion process [63], and the hydrothermal method [64]. The process of the synthesis has a great impact on the transformation phase, i.e., from amorphous to anatase or rutile or brookite [65]. However, this method produces defects in the surface structure and morphology of the new material, which becomes the main limitation because the surface chemistry and physical properties of the NPs are highly dependent on the surface structure as well as the morphology [66]. The synthesis of TiO2 NPs via the chemical method is most widely used due to its ease of fabrication and the proper control of the size and shape of the NPs. On the other hand, its utilization is limited due to its over-expensive production, ecotoxicity, excessive energy, and non-availability of environmental sustainability, together with high temperature and pressure, which inhibit their mass production further, reducing their applications in certain different areas [67].
In addition, the electrophoretic fabrication of TiO2 NPs is easy as well as timesaving, but it involves the utilization of non-aqueous solvents that are harmful, as reported by Boccaccini et al. in previous research [68]. The hazardous effects, including aquatic pollution, land pollution, environmental pollution, and the destruction of wildlife as well as habitats, are caused by these non-aqueous solvents. Furthermore, these non-aqueous solvents are toxic enough to cause certain health hazards when evaporated in the surrounding air in the environment [68].
The microwave-assisted preparation of TiO2 NPs is widely accepted due to its time- saving capacity and its uniform, homogeneous, and quick heating of the reaction mixture at required the temperature, which is an advantage. On the other hand, this method is not cost-effective but rather expensive because it requires the increased power of heating. As this method is an energy-intensive procedure, it therefore lacks the production of TiO2 NPs in large amounts [69].
The sol-gel process of TiO2 NPs is extensively investigated to fabricate more crystalline NPs, followed by sol formation, gelation, and solvent removal [70]. This method is costly because it involves the utilization of expensive raw materials that lead to the cracking and volume reduction of newly formed TiO2 NPs after the drying process [71].
The solvothermal method employs both high pressure and temperature for longer duration (pressure < 1 atm and temperature > boiling point of water). This is a time-consuming process and requires high maintenance costs for autoclaves because the equipment used is quite expensive [72]. In addition, when additional surfactants are used in this process, they will not only make the reaction more complex but will also lead to the excessive formation of impurities [73].
Currently, hydrothermal methods of synthesis have been greatly employed due to their ease in fabricating TiO2 NPs, which can be used on large scale for, e.g., TiO2 nanotubes [74]. The hydrothermal method is quite expensive due to the increased requirements of high levels of energy, surfactants, pressure, and temperature. Moreover, this a time-consuming process and less yield of TiO2 NPs is attained [75]. It has also been reported that changes in the temperature, ranging between 100 and 200 °C, have a lack of control on the size, shape, and crystallinity of the nanotubes. Consequently, the morphology and other properties of the formed NPs are changed, e.g., nanotubes to nanofibers [76]. The chemical synthesis uses high temperature, high pressure, and expensive and toxic chemicals, which produce adverse effects in certain medical applications [77].

Biological (Bottom-Up) Synthesis

The bottom-up technique involves biological synthesis via microorganisms (bacteria, algae, fungi, and parasites), green synthesis through plants (roots, leaves, shrubs, herbs, and seeds), and biological synthesis from derivatives (eggshell, albumin, lignin, etc.) [78]. This technique incorporates the atoms or molecules of smaller sizes that undergo reduction or oxidation to generate NPs. These atoms or molecules aggregate biologically, without utilizing toxic chemicals or solvents, to form NPs. These obtained NPs have negligible flaws and defects in their surface morphology, topography, and physico-chemical properties, such as size, shape, crystalline phase form, functional groups, structure, and texture. Above all, TiO2 NPs produced by the biological and green routes are immensely biocompatible and biosafe due to the lack of toxic byproducts formed [79].
Similarly, advanced research reflects the successful role of microorganisms in the biological synthesis process because it is easy and less time-consuming. The TiO2 NPs formed via different microorganisms are produced in high yields which are very stable, sustainable, cost-effective, biosafe, and biocompatible. This is due to the fact that the capping and reducing agents used are the natural enzymes and proteins present in microorganisms [80,81]. Hence, biological synthesis uses different life forms and organisms from very simple prokaryotic bacterial cells to complex eukaryotic cells such as angiosperms [82]. Different green synthesis methods utilize biological derivatives, such as the polyoxometalates, polysaccharides, Tollens, and phytochemicals (phenols, terpenoids, flavonoids, and cofactors) present in plants, which act as capping and stabilizing mediators in order to improve and enhance the stability of newly formed TiO2 NPs [66,83]. The biological derivatives, such as albumin, eggshell, gelatin, lignin cellulose, starch, arginine, peptides, polyamine, lysozyme, pomelo peel, and protamine have been used to synthesize TiO2 NPs which are biocompatible, biosafe, and cost-effective (Figure 4).
The direct synthesis method is the most biocompatible technology used for fabricating the TiO2 NPs without the consumption of any additional externally introduced templates, capping, reducing agents, etc. Hence, the corresponding natural biological extracts and their derivatives used during the synthesis process can produce a strong capping layer that corresponds to the biocompatibility and stability of these NPs [84]. In addition, this method produces the cost-effective, high-yield mass production of TiO2 NPs [67], which are reliable and non-toxic because the method does not utilize high temperature, high pressure, high energy, and toxic chemicals [85].
The biocompatible technology of synthesis is the latest and advanced technique in comparison to the conventional methods. The conventional methods of TiO2 NP synthesis are costly and produce adverse effects on human health and the environment [88,89]. The comparative differences between biocompatible and conventional TiO2 NPs are given in (Table 1) [85,90,91].

Conventional and Biological Synthesis

There are multiple factors that control the synthesis of TiO2 NPs and in turn make them suitable for medical dental applications. These significant contributing factors include the total time period, the environmental conditions, the temperature, the synthesis protocols, the solution’s PH, the concentration of extract, and the raw materials used, as is demonstrated.
(i)
Solution’s medium pH:
Temperature is also an essential parameter controlling the synthesis of TiO2 NPs using three methods. The highest temperature is required by physical methods (>350 °C), while the chemical methodology requires temperature limits below 350 °C. On the other hand, below 100 °C the TiO2 NPs are synthesized microbially, and plants use C or ambient temperature. Hence, the nature of the TiO2 NPs is determined by the reaction medium’s temperature [92,93].
(ii)
Pressure:
The size and shape of biological and green TiO2 NPs are greatly influenced by the pressure enforced on the reaction medium. It has been found that the reduction rate of metal ions in biological and green TiO2 NPs is faster at ambient pressure due to the utilization of natural biological agents [92,93].
(iii)
Time:
The total period for incubating the reaction medium has an immense impact on the type, along with the quality, of the TiO2 NPs prepared by biological and green synthesis. Additionally, other factors, including light exposure, synthesis process, and storage conditions, have a great effect on the characteristics of the TiO2 NPs synthesized by these routes. The particle aggregation occurs as a result of long-term storage, subsequently resulting in the particle’s growth or shrinkage and enhancing its potency and shelf life [92,93].
(iv)
Melting point:
The melting point of TiO2 NPs has been controlled by changes in the particle’s shape and size, which eventually influence the properties of these NPs. This means that the melting point of the TiO2 NPs has been reduced when the size of these NPs has fallen into the range of the nanometer scale. The shape transformation of those TiO2 NPs which display different configurations but have similar energy levels is quite easy. These versatile properties further affect their multiple chemical properties [92,93].
(v)
Environment:
The characteristics of TiO2 NPs are greatly influenced by the surrounding environment as well. The single TiO2 NPs are converted into core-shelled TiO2 NPs either through reaction with materials or absorption by materials in the surrounding environment. This process is carried out by corrosion or oxidation occurring in the environment. The biological TiO2 NPs are thicker and larger in size due to the formation of the coating around them [38] that might be responsible for their potent characteristics. In a study, zinc sulfide NPs changed their crystalline nature instantly with a change in environment from wet to dry. In another study, cerium-nitrate NPs changed their chemical properties when peroxide was used in their suspension medium [92,93].

3.2. Characterization Techniques of TiO2 NPs

The TiO2 NPs are characterized by their size, shape, phase form, crystallinity, surface roughness, morphology, texture, topography, elemental composition, band-gap absorbance energy, and compounds and the functional groups present in them that are the key features responsible for their biocompatibility. These standardized tools render them biocompatible and enhance their potency for utilization in medical and dental applications (Figure 5) [79,94,95,96].

3.2.1. Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM) for Size and Shape

The size, shape, structure, and surface morphology of the TiO2 NPs are determined by scanning electron microscopy (SEM), which is a basic parameter for characterizing NPs. SEM is a powerful technique for obtaining magnified high-quality and high-resolution images with a good depth of field to provide information about the morphology of the TiO2 NPs [97,98]. The standardized morphological analysis tool, named the transmission electron microscope (TEM), gives a more refined knowledge regarding the crystallographic arrangement of the atoms, complex phase transitions, elemental composition, dislocations, selected area electron diffraction (SAED), and defects in the TiO2 NPs [97,98].
The studies conducted by A. Mansoor et al. revealed the different sizes and shapes of TiO2 NPs derived from Bacillus subtilis, Cassia fistula, and hydrothermal heating. These different sizes and shapes of these NPs had a great influence in developing the cytotoxicity. The TiO2 NPs prepared from Bacillus subtilis were purely natural, having a slightly large size and spherical shape that prohibited their gross adsorption and absorption in the cell surfaces coming into contact with them. Thus, these TiO2 NPs were biosafe and biocompatible for their medical and dental applications (Figure 6a and Figure 7a). On the other hand, the TiO2 NPs formed by Cassia fistula and hydrothermal heating were very much smaller in size, and the irregular shapes were easily absorbed and adsorbed in the cell surfaces, resulting in cytotoxic reactions. Therefore, these newly formed NPs could not be used in nanomedicine and nanodentistry (Figure 6b,c and Figure 7b,c) [79,99]. These NPs could be recommended for medical and dental applications only if changes were made in the time, pH, melting point, chemicals, substrates, high temperature, and pressure during the synthesis protocols in future.

3.2.2. X-ray Diffraction Powder Spectroscopy (XRD) for Crystalline Phases

The X-ray diffraction pattern (XRD) is utilized to investigate the crystallography and structural arrangement of TiO2 NPs [100]. XRD is a fast and indispensable analytical technique, which is mainly used for a material’s identification, characterization, crystal content, phase identification, quality control, lattice plane spacing, purity percentage, and epitaxial growth of crystallites [101,102]. The XRD methods utilizing powder can easily identify the crystalline, crystallite, semi-crystalline, and amorphous forms of materials [103]. Recently, the role of the crystalline phases of TiO2 NPs in the medical and dental applications were investigated by A.Mansoor et al. in 2022, where it was revealed that the mixture of the anatase–rutile phase form state attained by the TiO2 NPs formed by the Bacillus subtilis (Figure 8a) were the best for certain medical and dental applications as compared to those prepared by Cassia fistula and hydrothermal heating, where only the anatase phase form state was dominant (Figure 8b,c). This happened because of the comparatively high temperatures and pressures used in the synthesis of both of the latter TiO2 NPs. As a matter of fact, the anatase phase form state is highly thermally reactive and unstable while the rutile phase form state is the most thermally stable and is less reactive [79,99]. Efforts should be made to produce the rutile phase form state of the TiO2 NPs in order to induce them in the future nanomedical and nanodental procedures.

3.2.3. UV–Vis Diffuse Reflectance Spectroscopy

The ultraviolet (UV) visible spectrum records the reflectance and absorbance in the 300 nm to 800 nm wavelength range in order to investigate the alterations in the molecular energy levels of TiO2 NPs. Electron transfer from either π-bonded orbitals or non-bonded orbitals originates this form of alterations. Thus, this standard tool [104] obtains deep knowledge about aromatic and non-aromatic compounds, π-electron orbital systems, conjugated and non-conjugated peripheries, bonding and non-bonding attachments, and saturated and unsaturated electronic systems in TiO2 NPs. Moreover, the UV–Vis absorption spectrum depicted the two characteristic peaks of TiO2 NPs in the wavelength and absorbance ranges of 264 nm and 322 nm in previous studies [105,106]. The aqueous solutions used in UV–Vis spectroscopy are responsible for confirming the formation of TiO2 NPs [106,107]. The standard of the band gap energy for the NPs is 3.20 e.V, which means that the size of the NPs is somewhat big if the calculated value is less than the standard, but if the calculated value is greater than the standard, the size of the NPs is very small. Various researchers concluded that TiO2 NPs generated by Bacillus subtilis showed 2.7 e.V, confirming the somewhat big and safe size of TiO2 NPs for their employment in the fields of medicine and dentistry (Figure 9a). Additionally, the TiO2 NPs fabricated by Cassia fistula and hydrothermal heating depicted 3.6 e.V and 3.9 e.V values, which showed the extremely small size of these NPs (Figure 9a,b) [79,99]. It might be possible to control the large band gap energy values as compared to the standard after the advancements in the conventional synthesis processes.

3.2.4. Fourier Transform Infrared Spectroscopy (FTIR)

FTIR is the standard characterization tool that is helpful in getting the infrared spectra of the absorption, adsorption, emission, and photoconductivity of semi-solids, gases, solids, and liquids by determining the nature of their bonds and functional groups [97]. The functional group available in a TiO2 nanoparticle is recorded between 4000 and 400 cm−1 in FTIR spectroscopy [86,108]. The stretching and bending peaks of the Ti–O–Ti vibration which appear in the range of 1000–1300 cm−1 mainly correspond to the formation of TiO2 NPs in addition to other organic compounds, including nitro compounds, ethers, esters, H-O-H bending, polyphenols, hydroxyl (OH) groups, and amide-II [109,110]. The current research concludes that the TiO2 NPs formed by Bacillus subtilis reveal only a few functional groups, such as O-H, C-O, and C-H (Figure 10a), whereas several functional groups such as O-H, C-O, C-H, C=O, C-C, C=C, and C≡C were observed in the TiO2 NPs prepared by Cassia fistula and hydrothermal heating (Figure 10b,c). These functional groups are responsible for the development of the cytotoxicity of these NPs. Thus, the presence of a few functional groups in the TiO2 NPs formed by Bacillus subtilis makes them the best option to be utilized for the medical and dental fields as a result of their non-cytotoxicity [79,99]. The TiO2 NPs generated by Cassia fistula and hydrothermal heating could be modified by using the least reactive chemicals in order to reduce the overproduction of large amounts of functional groups that could be harmful in medicine and dentistry.

3.2.5. Elemental Composition Analysis (EDX)

The energy dispersive X-ray spectrometer (EDX or EDS) is a versatile piece of equipment used to identify the spectrum and determine the elemental composition present in TiO2 NPs [111]. This elemental composition gives complete information regarding the purity and performance of these TiO2 NPs [112]. In recent studies, it was observed that TiO2 NPs prepared by Bacillus subtilis revealed a high content of titanium but a low content of oxygen (Figure 11a). On the other hand, TiO2 NPs formed by Cassia fistula and hydrothermal heating depicted a large content of oxygen but a lesser content of titanium (Figure 11b,c). The presence of large amount of available titanium in the TiO2 NPs formed by Bacillus subtilis might have enabled them to be utilized easily in the medical and dental areas as compared to the TiO2 NPs formed by Cassia fistula and hydrothermal heating, where the content of titanium was too low. This imbalance between the weight% and atomic% of titanium and the oxygen in the TiO2 NPs formed by Cassia fistula and hydrothermal heating might be the reason for their non-consumption in medicine and dentistry [79,99]. The volume and concentration (dose) of the chemicals must be controlled efficiently to produce TiO2 NPs with nearly equal or a somewhat large weight% and atomic% of titanium in comparison to the oxygen.

3.2.6. Atomic Force Microscopy Analysis (AFM)

The most useful and popular standard tool employed to image the surface topography, texture, and structure of TiO2 NPs is the atomic force microscope [113]. It gives entire, detailed information about the topographic map, including the surface roughness, bumps, ridges, and phases of these NPs [114,115]. According to the current research, the TiO2 NPs prepared by Bacillus subtilis displayed bumps of the anatase–rutile phase form states with minimum roughness (Figure 12a), as compared to the TiO2 NPs formed by Cassia fistula and hydrothermal heating (Figure 12b,c) where they showed enhanced roughness with slight or no bumps in the anatase phase form state. This minimal roughness might have allowed the least absorption and adsorption of these NPs in the cell surfaces, thus making them biocompatible for medical and dental applications. The maximum surface roughness in the TiO2 NPs formed by Cassia fistula and hydrothermal heating might have allowed the increased level of absorption and adsorption of these NPs in the cell surfaces, resulting in their cytotoxicity [79,99]. The large amount of surface roughness present in the TiO2 NPs formed by Cassia fistula and hydrothermal heating could be controlled by utilizing the large amounts of stabilizing agents. These agents could be either natural or synthetic and form a capping layer on the surface of these NPs, making them suitable for applications in the medicine and dentistry.

3.2.7. Characteristic Features of TiO2 NPs Dependent on Characterization Techniques

The unique physical, chemical, mechanical, and biological properties of TiO2 NPs make them superior to their parent bulk material from which they are formed. These ideal properties make them compatible and increase their demand in health care systems [116]. There is strong correspondence between these various properties of NPs and their utilization in the fields of medicine and dentistry. Other important parameters include dimension, durability, quality, number, quantity, mass, bulk, volume, and aggregation (Figure 13) [117,118].

Role of Size and Shape of TiO2 NPs on Their Medical and Dental Applications

The nanometer size of TiO2 NPs provides them with a high surface energy, a large fraction of surface atoms, reduced imperfections, spatial confinement, and reduced imperfections in comparison to the corresponding bulk material. At the nanometer scale, the properties of NPs become size- and shape-dependent, which involves chemical properties (reactivity), catalysis, thermal properties (melting temperature), mechanical properties (adhesion and capillary forces), optical properties (absorption and scattering of light), electrical properties (current), and magnetic properties (superparamagnetic effect) [120,121,122,123,124,125,126]. Early in 1996, Volokitin et al. [127] already confirmed that quantum size strongly influences the thermodynamic properties of metallic origin-based NPs especially TiO2 NPs. The confinement of movement of the electrons leads to the formation of discrete energy levels and electric dipoles because of the smaller size of newly formed NPs. This small size of NPs greatly influences their oxidation, reduction, and catalytic properties in a positive manner, thus reducing cytotoxicity [128]. Thus, because of these extraordinary features the TiO2 NPs are of great interest to the current researchers [129].
The size has been considered as one of the most important parameters responsible for inducing the toxic effects on human health. The much smaller-sized TiO2 NPs have a higher surface-area-to-volume ratio, with higher chemical reactivity and vice versa. Due to the large surface area of these very small-sized TiO2 NPs, more of them are absorbed in its surface, resulting in toxicity [130]. In an in vivo study, rats consumed two sizes of TiO2 NPs, with diameters of 20.27 nm and 250.10 nm; the study concluded that smaller TiO2 NPs have bit more of an adverse cytotoxic effect on lungs in comparison to large-sized TiO2 NPs of 250.10 nm [131]. A study conducted previously revealed a strong direct correlation between TiO2 NPs size and level of DNA damage [117,118]. The smaller TiO2 NPs are more easily taken up by cells, producing intense adverse effects on cell function, thereby leading to a high level of toxicity. The reactive oxygen species (ROS) production is related to nano-toxicity [132].
Moreover, particle shape also plays a dominant role in determining the toxicity of TiO2 NPs [133]. Different shapes of TiO2 NPs display different cellular uptakes, subcellular localization, and production of reactive oxygen species [134]. There exists a direct correlation between the shape of the TiO2 NPs and cytotoxicity. There are only limited studies available that have investigated biocompatibility in cell lines. Braydich-Stolle et al. [135] observed the effects of size and crystal structure on toxicity or mode of cell death. In another study, Gratton et al. investigated the influence of shape on toxicity and revealed that non-spherical TiO2 NPs (rod-shaped particles) are easier and faster to become absorbed/adsorbed on cell surfaces, thus making them cytotoxic in comparison with other TiO2 NP shapes (spherical-shaped particles) [136].

Role of Surface Area of TiO2 NPs on Their Medical and Dental Applications

The ratio of the surface atomic number to the total number of atoms increases greatly as the particle size gets smaller in NPs [121]. This decrease in particle size increases the number of atoms on the surfaces as compared to those found inside, which leads to the enhanced chemical reactivity of these NPs [137]. As a result, the surface-to-volume ratio of the NPs also increases, which has a significant effect on their physical, chemical, and mechanical properties compared to those of the corresponding bulk state in the parent material [122,123,124,125,126,137,138]. In addition, TiO2 NPs show increased catalytic activity as compared to other NPs [124,126]. The reason it became possible was because the smaller size of TiO2 NPs provides a large surface area for the catalysis to take place [126,139,140,141,142]. If the TiO2 NPs get very much smaller, then they exhibit a much greater surface area and number of particles per unit mass, which leads to much greater chemical reactivity than is even required [143]. This greater activity eventually increases the absorption of these NPs in the cell lines, elevating reactive oxygen species (ROS) production [133], oxidation [144], and DNA damage [133].

Role of Crystalline Phases of TiO2 NPs on Their Medical and Dental Applications

Owing to the many active atoms on the surface of NPs, these NPs have specific enhanced surface adsorption, dispersion, aggregation, viscosity of colloidal suspension, and rheological properties [145]. Chemically, these NPs are more active and stable than the corresponding bulk material used for their formation [146]. The three important phase states of the TiO2 NPs are brookite, anatase, and rutile [122,123]. These crystal phases are available in the form of octahedra with six oxygen anions surrounded by titanium cations. The formula of these NPs is TiO6/3, which is equal to TiO2 leading to different patterns of crystalline phase states [123]. These three distant phase form states differ in structure, texture, size, density, volume, and refractive index [122,123,124]. The rutile phase form state is well known for its high level of texture, structure, and chemical stability during the preparation methods, but its use is rare [122,123,124,125,126,147]. The anatase and brookite phase form states are meta-stable because they are thermodynamically unstable as compared to the rutile phase form state when used in the industry [125,126]. The brookite is rarely used in the industry because it is difficult to prepare [123]. The chemical reactivity of anatase is the highest when compared to the other phases [148,149]. In a study, Sayes et al. [149] reported that TiO2 NPs (80/20; anatase/rutile, 3–5 nm; 100 μg/mL) produced 6 times higher reactive oxygen species (ROS) than rutile after UV exposure. This is because the anatase phase state of TiO2 NPs produces ROS when exposed to ultraviolet light [149]. It is even suggested that TiO2 anatase has a greater toxicity potential than TiO2 rutile [150,151]. In addition, Di-Virgilio et al. evaluated the impacts of the anatase phase form states in different dose volumes, e.g., (0–300 μg/mL) on rat osteosarcoma-derived cell lines. Eventually, they reported the formation of phagocyte vesicle-induced cytotoxicity leading to osseointegration failure [152]. In another study, Martinez-Gutierrez et al. reported the effect of combining (anatase and rutile) TiO2 phase form states with silver NPs. They observed that the combination of two phases showed a superior antimicrobial activity with reduced cytotoxicity as compared to 100% rutile TiO2 NPs due to their inherent properties [153]. A recent study found that TiO2 NPs with a large percentage of anatase (anaerobic P-25 99.5% 73–85% anatase, 14–17% rutile, and 2–13% amorphous) produced genotoxicity, mutagenicity, and cytotoxicity on human fibroblasts in different doses (such as 10, 25, 50, 100, 250, 500, and 1000 μg/mL) to detect γ-H2AX expression [154].

Role of Dose of TiO2 NPs on Their Medical and Dental Applications

The TiO2 NPs can go directly into the liquid phase of the reaction system without getting attached to an inert carrier that could have adversely affected their catalytic activity, stability, and reactivity [155]. This is because metal oxide NPs, especially TiO2 NPs, show a much higher catalytic activity than any other type of NPs due to the fact that the small nano-size and the large surface area of these NPs support an exceptionally increased active catalytic activity [156,157]. The dose is directly related to the degree of exposure or concentration of the substance in the medium (such as air, food, and water). The different dosage indicators have been used in nanotoxicology, including cm2/mL, μg/mL, μg/cm2, and the number of particles/mL [158]. Previously, a study revealed that the cytotoxic effects of TiO2 NPs after inhalation were more when exposed to small-sized TiO2 NPs (20 nm diameter) with a low dose (10 mg/m3) as compared to large-sized TiO2 NPs (300 nm diameter) with a high dose (250 mg/m3) [130]. In another study, cell viability was reduced by 20–30% when HT29 cells were exposed to different concentrations of TiO2 (50, 100, 200, and 400 µg/mL) for 48 h; the study concluded that these nanoparticles affect cell viability, resulting in cytotoxicity in a dose-dependent manner [159].

Role of Surface Coating of TiO2 NPs on Their Medical and Dental Applications

The metallic NPs possess unexceptional optical properties as a result of their nano-size ranging between 1 and 100 nm and the confinement of their electrons together with the quantum effects, e.g., in solution form gold (Au) NPs are found as deep red and/or black [160]. The interaction of light and the coupling between several metallic NPs, especially TiO2 NPs, induce a field enhancement in the surroundings, resulting in increased absorption, scattering, and luminescence, thus giving them whiteness [161]. The TiO2 NPs are considered as the cosmetically most white pigments, which can easily enhance the aesthetics in medical and dental applications [162,163]. The (TiO2) NPs are white in color, with the highest refractive index of 2.50 in the anatase phase form state and 2.70 in the rutile phase form state, which is the whitest in color [122,147]. The TiO2 NPs are capable of undergoing photocatalysis because of their dual behavior of absorbing and scattering ultraviolet radiation against light [122,123,124,164].
In the cases where TiO2 NPs without surface coatings are used, they initiate the cytotoxic reactions when they come into contact with cells and other biological materials. The development of surface coatings on these NPs can make them non-toxic via the utilization of surfactants. Eventually, the physical and chemical properties of the TiO2 NPs are changed, which prohibits the cytotoxicity. The important properties in this respect are the magnetic, electrical, thermal, and optical properties and the chemical reactivity [165,166]. Moreover, the presence of oxygen, (ozone) [166], oxygen free radicals [167], and transition metals [144] on the surface of the initiating TiO2 NPs can lead to the formation of reactive oxygen species (ROS), which causes inflammation. Previously, a study was carried out which concluded that cytotoxic reactions in silica are due to reactive oxygen species and surface free radicals [130].
The metallic NPs have higher values of hardness, plasticity, specific heat, thermal expansion, conductivity, diffusivity, and lower values of sintering temperature and shrinkage as compared to the bulk material [155,168]. Other properties of metallic NPs, especially TiO2, include surface plasmon resonance and superparamagnetism, which makes them desirable as diagnostic intervention tools for medical applications [169,170,171].

3.3. Significance of TiO2 NPs in Medical and Dental Platform

Previously, the utilization of other NPs such as graphene was wide because of its versatile characteristics, including outstanding biocompatibility, physicochemical properties, and antimicrobial activity. Furthermore, it was highly conductive with a large surface area; it was wear/friction resistant and could stabilize the NPs if used as a substrate [172]. Additionally, graphene was a material of choice for various dental applications such as in bone grafting, tissue engineering, pulp and periodontium regeneration, dental cements and adhesives, resins, whitening agents, dental implants collagen membranes, and biomarkers of oral saliva. The use of graphene in dentistry resulted in enhanced positive outcomes, but its manufacturing for industrial purposes was quite challenging and time-consuming [173]. This was an initial step that opened a gateway to the incorporation of other biocompatible and cost-effective nanomaterials in the future nanodentistry. Among them, a TiO2 nanoparticle was one of the top five most common, easily available, and compatible NPs for the industry purposes [140,174,175]. There has been a persistent rise in the production of TiO2 NPs globally, which increases its mass production availability for different fields [176]. Currently, TiO2 NPs are the most biocompatible, with low toxicity and low allergic potential that makes them ideally suitable for utilization in medical and dental applications [39,40,177,178]. The TiO2 NPs are resourceful, versatile, and adaptable because of their extraordinary functionality, increased strength, light weight [179], brightness, resistance to UV light, high density [177,178,180], cost effectiveness, purely white color, high corrosion resistance [181,182,183], long-term chemical stability, and durability [184,185]. In addition, these NPs possess high mechanical resistance, high electrical conductivity, low thermal diffusivity, chemical stability, corrosion and wear resistance, and low thermal conductivity. These NPs are ductile in pure form and can be easily shaped into any form and type that makes them fatigue resistant [115].
The desired mechanical and physicochemical properties of TiO2 NPs in the context of size, shape, surface texture, morphology, topography, crystalline, and phase form can be easily attained which makes them superior when compared to other metal oxide NPs [42]. According to the previous literature, the low ion release and stability of these TiO2 NPs additionally accounts for their biocompatibility. These NPs are capable of producing a TiO2 layer on their surface after their accidental exposure to air, thus preventing the metal inside these NPs from being dissolved and getting distorted. This chemically inert layer of TiO2 plays a key role in the biosafety and biocompatibility of these NPs in the human body [181,186,187]. Owing to these versatile features, the use of TiO2 NPs in the dental and medical field has rapidly started to increase [182,188,189]. According to previous research, large quantities of pure titanium in the range of about 99.5% and 0.5% of the interstitial elements (carbon, oxygen, and nitrogen,) are even safe, biocompatible, and allergic free in both dental and medical applications. This research concluded that even significantly more large quantities of titanium oxide are required for any allergic or toxic reaction to arise in humans [182].

3.3.1. Significance of TiO2 NPs as a Medical Material

The medical applications of the TiO2 NPs is illustrated in the Figure 14a.

Use of TiO2-Based Drug Delivery Systems in Cancer Therapy

The (TiO2) NPs have potential biomedical and industrial applications due to their size, acceptable biocompatibility, lower toxicity, and cost effectiveness [190,191,192]. Nanostructured arrays comprising nanotube walls of TiO2 involving copper and silver oxide NPs have been used as surface coatings due to their antibacterial nature. Moreover, they have favorable cytocompatibility and are utilized for differentiation, cell spreading, and proliferation [193,194]. Traditional chemotherapeutic drugs have severe adverse effects on normal cells, due to which nanotechnology has gained significant importance in the diagnosis and treatment of cancer [195,196].
The role of nanotechnology in cancer treatment is related to the enhancement of therapeutic effects and the reduction in the side effects of chemotherapeutic drugs [197]. Moreover, controlled drug delivery and targeted drug release are aimed at through the use of these NPs. Previous studies investigated the TiO2 whiskers in cells for the delivery of drugs in cancer therapy and the synergistic effects on accumulation and internalization of daunorubicin [198]. Similarly, a synergistic response has been observed in cell lines involving breast cancer when the (TiO2) NPs conjugate to doxorubicin [199].
A study conducted on TiO2 nanotube surfaces containing quantum dots of zinc sulphide, which were studied with 5-flourouracil, revealed that it was of significant importance as a drug delivery system [200]. The studies revealed that NPs with a modified surface may enhance the performance of anticancer medications and lower their toxicity [195]. The biochemical and hematological results of a study revealed better anticancer performance of surface-altered paclitaxel affixed to TiO2 NPs and modified hydroxyapatite as compared to pure paclitaxel [201]. Similarly, a higher accumulation of cisplatin was observed in cancer cells when cisplatin was loaded on TiO2 NPs containing hyaluronic acid [201].
The TiO2 NPs may also serve as a photoactive carrier for drug delivery. These NPs may be a suitable carrier for various drugs due to their stability, high surface area, possibility for surface modification, and availability [202]. Previously, studies were carried out for developing skin cancer treatment and light-controlled drug release by the modification of TiO2 NPs with the help of polyethyleneimine (PEI), which may act as an agent for targeting cancer [203]. In another study, drug delivery triggered by near-infrared radiation was developed through the loading of DOX for the preparing of a porous TiO2 shell. Furthermore, hyaluronic acid capping was performed, and the release of the drugs was studied. This study also declared that at lower doses of the drug a decline in cell viability meant that it was a promising method for therapy in cancer [204].
Colloidal TiO2 NPs that are controlled through light may be used for delivering ruthenium complex to melanoma cancer cell lines. This system demonstrated a fast release profile when an ultraviolet light was used in comparison to red light. It has been pointed out that the destruction of cancer cells may be higher due to elevated oxidative stress when exogenous agents are used, as the reactive oxygen species are higher in the tumor environment [205]. Apoptosis is triggered in tumor environments when reactive oxygen species are induced, and it is an appropriate cell death type in cancers [206]. When stimulated with ultrasonics, TiO2 NPs exhibit improved anticancer activity against various anticancer cell lines. [205,207] In sonodynamic therapy, hydrophilized TiO2 NPs demonstrate higher resistance to degradation and prolonged circulating time [207]. Radiofrequency as a no noninvasive technology may help in the release of drugs from nanotechnology-based TiO2 drug-delivery systems. Drug release can be induced in titania NPs (TiO2) by transferring radiofrequency energy through efficient thermal transducers such as gold NPs [208]. This technique has displayed substantial capability in inhibiting human gastrointestinal and hepatocellular cancer cells [209].

Prevention and Treatment of Infections through TiO2 Based Antibacterial Devices

With the advancements in nanotechnology, metals and metallic oxides have gained importance in biomedical industry as they can be used as effective antimicrobial agents due to their enhanced safety and stability as compared to organically produced antimicrobial agents [210,211]. Initially, TiO2 was used as a photocatalytic microbicide against Lactobacillus acidophillus, E. coli, and Sacharromyces cerevisiae [212]. Several studies suggest that in water the photocatalytic activity of TiO2 is effective against various organisms, such as bacteria (Gram-positive and Gram-negative), viruses, algae, fungi, bacterial toxins, and protozoa [213]. In water and in response to exposure from UV light the microbial cell viability may be inhibited by a thin film of TiO2 and platinum as proven in previous studies [212].

Antibacterial Action of TiO2 in Orthopedic Implants

Osteomyelitis may be caused due to an infection with mostly bacteria or due to fungi. This infection typically occurs as a consequence of the treatment with implants required in the case of bone surgery or bone fracture. The management of this condition requires a specific and prolonged treatment with antibiotics. This infection may be prevented by the use of implants with modified surfaces such as implants coated with TiO2 nanostructures [214]. One study revealed the antibacterial properties demonstrated by titanium-based implants when silver-doped TiO2 nanotubes were used for the surface modification of these implants [215]. This study explained that the adhesion of bacteria reduced when the surface of the titanium-based implants was loaded with TiO2 as compared to non-coated implants.
Recently, micro-arc oxidation composites with incorporated (TiO2) were coated on magnesium alloy, and this was presented as a biomaterial with antibacterial properties against microorganisms such as Escherichia coli; it demonstrated sufficient bioactivity with in vivo degradation properties and acceptable cytocompatibility for applications in orthopedic surgeries [216]. Metal wires and pins are used in cases of orthopedic fractures for the purpose of fixation. These devices are susceptible to colonization with bacteria such as E. coli, Staphylococcus aureus, and Staphylococcus epidermidis [217]. The TiO2 coating of these pins and wires is effective due to its photo killing activity as it influences bacteria in comparison to the host tissue. Moreover, TiO2-coated implants and devices are biocompatible and environmentally safe [218]. Moreover, osteogenic differentiation is promoted by TiO2 NPs [219]. A study conducted on implants produced two implant surfaces (Ti-160 and Ti-120) using the electrophoretic deposition method. These treated surfaces enhanced osteoblast proliferation and reduced bacterial colonization [220].

Antibacterial Applications of TiO2 in Hospitals and Medical Devices

In medical devices, the formation of microbial biofilms poses a challenge for the health professionals. These biofilms are mostly formed by Staphylococcus aureus and Staphylococcus epidermidis and may be the source of infection. The hospital industry may employ TiO2 nanoparticle-coated surfaces due to their antibacterial properties and eliminate these sources of infection as traditional methods of disinfection are not as beneficial when compared to the photocatalytic methods [221]. The application of TiO2 NPs has been performed on ceramic tile surfaces and Petri dishes, and it has been reported that the deactivation of various microorganisms can occur in response to ultraviolet light [222]. Light-induced disinfection can be performed in TiO2-coated catheters, and these are therefore safe to use as biomedical equipment. A study conducted on TiO2-coated catheters revealed the reduction in Escherichia coli after the exposure to UV light within 1 h [223]. Bactericidal effects shown by medical tubes and silicone catheters on cells of Escherichia coli in a study were observed [224].
Self-monitoring devices for blood glucose in diabetes (SMBG) contain sterile needles known as lancets. The ability to self-sterilize is an issue with these needles. These needles have previously been sterilized with gamma ray radiation that poses a drawback of gamma ray leakage and higher cost of equipment. Therefore, the photocatalytic activity of TiO2 NPs that provides antibacterial properties to the coated surfaces has been used to resolve this issue. Under ultraviolet illumination, TiO2 coated lancets demonstrated antibacterial properties to the K-12 suspension of Escherichia coli [225]. Sulphur-doped TiO2 has also been used in medical equipment for the purpose of antimicrobial activity. In order to control hemorrhage and infection it has been proposed that metal oxides such as TiO2 may be incorporated into polymers. Hemostatic ability and antibacterial activity have been enhanced previously by the use of nanocomposite films comprising chitosan pectin containing TiO2 NPs. These could serve as basis for the production of wound healing bandages.

TiO2 Implants Capable of Releasing Drugs

Mesoporous surfaces can be provided to implants by combining mesoporous TiO2 with active hydrophobic or hydrophilic low molecular weight substances that promote osseointegration by providing nano-topography on the surface. A film has been developed that contains 6 nm-sized pores and 200 nm-thick TiO2 mesoporous film that is coated on titanium implants. These films were previously loaded with drugs used for osteoporosis (Raloxifene and Alendronate drugs). De novo bone growth and fixation of the implant was observed in implants with and without the drugs. The Raloxifene drug demonstrated significantly more favorable results, while improved bone implant fixation was revealed with both the drugs [226]. Similarly, these mesoporous films can be used to load antibacterial agents and elements that help in the metabolism and proliferation of osteoblast cells in the body [227,228,229]. Titanium coating enhances the chemical and mechanical properties and biocompatibility of polyetheretherketone (PEEK) implants without beam evaporation.

The TiO2-Based Biosensors

Nano-sized semiconductors have electrodes made out of TiO2 due to their large surface area, exceptional biocompatibility, and a porous structure. TiO2 may be used as an immobilizing matrix that reacts with carboxyl and amine groups in enzymes and helps them in retaining the biocatalytic activity [230,231]. The TiO2-based nanocomposites have gained interest for use in biosensors. Chitosan, with nanowires of TiO2 and graphene oxide that have been electro-reduced and are a part of a DNA-functionalized nanocomposite, has been utilized for detecting gene sequences in Vibrio parahaemolyticus. The interface area in nanocomposites can be enhanced due to TiO2 nanowires [232]. Various biomarkers, such as proteins, metabolites, DNA, lipids, RNA, etc., can be detected with TiO2-based nano-biosensors [233]. The analytical performance and desirable sensitivity of TiO2-based PEC biosensors have acquired interest in recent times [234]. However, the higher sensitivity and selectively functioning of TiO2-based nano-biosensors are required through advanced research for these biosensors to be incorporated in a clinical environment [235].

The TiO2 NPs in Sunscreens

The TiO2 NPs are excessively valuable NPs because of their potent role against the spread of various skin cancers occurring as a result of extra exposure to ultraviolet radiation [236,237]. These NPs are capable of effectively scattering the UV photons, which increases sun protection factor (SPF-1) in sun creams and thus prevents the photo damage [238,239,240]. Additionally, this photo-protective capability of TiO2 NPs provides aliments in other skin diseases as well [241,242]. The United States Food and Drug Administration (FDA) has declared the TiO2 NP-induced sunscreen safe [243] and gave it the commercial name of Eusolex-®-2000 [244]. These sunscreens are transparent and give brightness to the skin and are thus approved by the users [245,246]. A study concluded that TiO2 NPs of 62.1 nm in diameter showed the maximum protection against skin cancer and other skin diseases [247]. The European Commission’s Scientific Committee on Consumer Safety (SCCSS) declared that 25% TiO2 NPs in the sunscreen is safe with no risk of cytotoxicity when applied on sunburned skin or healthy intact skin [248,249]. Various other researchers also supported the idea that the inclusion of TiO2 NPs in sunscreen lotions is totally safe and does not penetrate deeply in the skin and thus reduces the chance of any cytotoxic effect [250,251,252,253,254].

The TiO2 NPs in Anticoagulants and Wound Dressings

The significant current medical applications of TiO2 NPs include their utilization as anticoagulant [255] wound dressings and skin tissue engineering [256,257,258]. The anticoagulant properties of these NPs inhibit the blood clot formation in the heart and brain, thus preventing their damage [255].

3.3.2. Significance of TiO2 NPs as a Dental Biomaterial

The dental applications of the TiO2 NPs are illustrated in the Figure 14b.

TiO2 NPs in Orthodontics

Recently, TiO2 NPs have been introduced in dental composite resin adhesives utilized for bonding purposes in orthodontics for increasing mechanical strength and antimicrobial properties [230]. The different percentages of TiO2 NPs in conventional composite resin revealed excellent antibacterial activity against S. sanguinis and S. mutans with maximum shear bond strength to enamel in 5% by weight [259]. The TiO2 NPs are also used in the orthodontic wires due to the increased stability, durability, biocompatibility, and non-allergic potential [179,260,261]. The (TiO2) NPs were also used for treating the surfaces of orthodontic wires. It was detected by the researchers that orthodontic wires without surface coating with TiO2 NPs showed enhanced bacterial adhesion (S. mutans) in the oral cavity in comparison to surface orthodontic wires with a surface coating of TiO2 NPs [262]. In another study, it was concluded that ceramic orthodontic brackets with a surface coating of TiO2 NPs displayed maximum antimicrobial activity against pathogenic bacteria (S. mutans and C. albicans) [263]. This shows the potent role of TiO2 NPs in the field of orthodontics [264].

The TiO2 NPs in Dental Implants

Dental implants are commonly made of titanium due to its strong mechanical and chemical properties. The dental implants made of TiO2 NPs showed both enhanced osteogenic activity and antibacterial activity but still there are a few chances of peri-implantitis, which can be overwhelmed by modifying the surface of the TiO2 implant with other materials. The TiO2 NPs are the most versatile and compatible nanomaterials, which can be used in combination with many other elements without producing any toxic effect [265,266]. The surface modification of TiO2 implants with Zinc made it not only novel and gave it cytocompatibility but also increased its antibacterial functions against S. mutans, Porphyromonas gingivalis [267].
Moreover, the modification of the TiO2 implant with multilayers of TiO2 NPs enhanced its adhesion and strength in the bone [268]. The heat-treated TiO2 implants also revealed potent osteoblastic viability and antibacterial activity against Porphyromonas gingivalis [269]. The hydrothermally treated TiO2 implant exhibited non-cytotoxic behavior against mammalian cell lines and increased deposition of calcium from the osteoblast with antimicrobial activity against methicillin resistant Staphylococcus aureus [270]. Thus, it is essential to protect the adhesion of bacteria on implant surfaces in order to prolong their longevity, stability, and durability by preventing infections [271,272,273]. The additional property of TiO2 NPs is their strong biocompatibility with blood, which can easily allow their surface modifications in the blood with compatible implants [274]. The nerve regeneration via Schwann cells to dental implants occurs through HA nanocomposite ceramic coating to the TiO2 NPs [275]. Currently, novel, multilayered (TiO2) NP-based dental implants are being introduced for their utilization [276]. Due to the improved cell adhesion, proliferation, migration, differentiation, and high contrast wettability of TiO2 NP-based surfaces, these NPs have become a promising approach for yielding high strength scaffolds of implants [277,278,279]. In a study conducted previously, TiO2 NPs were grown as nanotubes on titanium surfaces and incorporated with ZnO NPs with a diameter of 20 to 50 nm and were bonded to the walls of the TiO2 nanotubes. These dominantly TiO2 NP-based implants provided strong antibacterial activity against S. mutans and Porphyromonas gingivalis through a constant and slow release of zinc ions [280]. In another study, TiO2 NPs were used as TiO2 nanorod arrays (TNFRs) to form surface modified implants. The TiO2 nano-array modified Ti substrate promoted the adhesion, proliferation, and osteogenic differentiation of human periodontal ligament stem cells. In addition, TiO2 nanorod arrays (TNRs) exhibited superior antifungal/antibacterial properties [281,282]. The TiO2 NP-based implants coated with the hydroxyapatite nanocomposites were also successful in providing the strong bone tissue osseointegration [283].

TiO2 NPs in Oral Disinfectants and Mouth Washes

Dental caries or tooth decay is a common but pathogenic oral disease associated with the dissolution of enamel and dentin surfaces of teeth because of acid-producing bacteria. This in turn leads to white spot fabrication initially and followed finally by tooth loss. This global problem of the tooth decaying process can be easily avoided by utilizing the latest oral disinfectants with metal oxide NPs, as compared to chlorhexidine, which is a conventional oral disinfectant [284]. Previously published research depicted the potent antimicrobial activity of TiO2 NP-containing oral disinfectants against Streptococcus mutants, which are the pathogenic species in the oral cavity responsible for the tooth decay or dental caries [285]. Additionally, these TiO2 NP-containing oral disinfectants and mouthwashes are helpful in preventing the formation of white spots on the teeth surfaces after the orthodontic alignment [264].

TiO2 NPs in Restorative Materials

This success of the restorative material in dentistry is greatly dependent on its biological strength and chemical, mechanical, and physical properties [286]. The wide range applications of TiO2 NPs in bases, composites, sealants, cements, adhesives, and lining materials displayed immensely effective antimicrobial activity. Previously, Garcia-Contreras et al. [79,287,288], revealed that TiO2 NP-induced glass ionomer cement is a novel restorative material introduced in dentistry as a result of its longevity, stability, and antimicrobial activity against strong masticatory loads. In another study, a novel TiO2 NP-based GIC restorative material proved to have a long shelf life with strong antimicrobial activity against S. mutans [272]. The addition of TiO2 NPs in conventional composite restoration gave rise to the latest nanocomposites with cytocompatibility and antimicrobial activity against E. coli and S. aureus, which are also responsible for the origin of tooth decay or dental caries [271]. Moreover, Ag doped with TiO2 NPs has been also used as a restorative material due to its strong antimicrobial activity against S. aureus [289]. The addition of different concentrations of TiO2 NPs in GIC restorative materials has a great impact on mechanical and physicochemical properties. In a study, TiO2 NPs with a mean particle size of about 21 nm have been incorporated into GIC restorative material to investigate its mechanical and physicochemical properties. In the presence of TiO2 NPs, the flexural strength, hardness strength, compressive strength, enamel shear bond strength, and dentin shear bond strength of the novel titanium oxide NP-based GIC significantly increased in comparison to the conventional GIC restorative material [290]. In another study, the incorporation of a combination of 2 wt% TiO2 NPs and 1 wt% of CnC in conventional GIC restorative material enhanced the compressive strength by 18.9% and the tooth enamel shear bond strength up to 151% when compared to the conventional GIC restorative material [291].
The TiO2, ZnO, and Ag NPs were incorporated in conventional GIC 10%wt and were tested for their mechanical and physicochemical properties. The TiO2 NPs used in this study had pores and thus, they were stronger than ZnO NPs and Ag NPs. The TiO2 NPs and ZnO NPs had significantly higher compressive strength while the compressive strength of the Ag NPs was equal to the control group of the conventional GIC restorative material [292]. The addition of 3% (w/w) TiO2 NPs in the GIC restorative material improved the mechanical and antibacterial properties without interfering with the release of the fluoride capacity of GIC, which is an advantage of TiO2 NP-induced GIC [293]. The addition of TiO2 NPs significantly improved the mechanical and antibacterial activity in different previous research performed by Elaska et al. and J. Sun et al., but only up to 3% [291,294].
Recently, research conducted by A. Mansoor et al. [79,295] confirmed that the addition of 3 and 5 wt% TiO2 NPs into the conventional GIC significantly increased its flexural strength and compressive strength as compared to the conventional GIC.

The TiO2 NPs in Whitening Agents

A TiO2 nanoparticle has displayed increased aesthetics in terms of whitening in dentistry. In a study, TiO2 NPs were added into 3.5% of a H2O2 whitening agent that enhanced the aesthetics of the enamel by improving its color, as compared to the conventional 3.5% of H2O2 whitening agent without TiO2 NPs [296,297]. The TiO2 NPs with a doping of nitrogen have revealed antimicrobial activity with an increased whitening effect and reduced sensitivity [298]. In addition, the mixture of TiO2/Ag NPs in H2O2 was used as a whitening agent and then tested for its cytotoxicity through in vivo testing. The mixture of TiO2/Ag NPs showed lesser cytotoxicity when compared to the whitening agent containing only H2O2 [299].
In a previous study, TiO2 NPs were mixed with H2O2 and were applied in a gel to increase the whitening efficacy of hydrogen peroxide (H2O2). Thus, in the presence of TiO2 NPs, the H2O2 concentration could be decreased to 6% with no clinical difference in the whitening effect compared with the H2O2 of 35%, but tooth sensitivity was reduced when the H2O2 concentration was reduced to 6% and the TiO2 NPs used [300].

The (TiO2) NPs in Acrylic Resins

The poly-methyl-methacrylate (PMMA) is the best denture base resin material used for the construction of dental prosthesis, such as complete and partial dentures. This PMMA is recommended because of its biocompatibility, biosafety, aesthetics, low weight, and low cost in dental prosthesis [301]. The physical, chemical, and mechanical properties of the dental prosthesis based on PMMA are reduced, which results in defects, including voids, irregularities, pores, and cracks [302]. These defects reduce the mechanical properties [303] and increase the growth and multiplication of microorganisms on these prosthesis [304,305]. The most pathogenic microorganism that gets attached to this dental prosthesis is Candida albicans, which can cause severe infections in the oral cavity if not controlled [306,307]. The properties of PMMA can only be improved with the incorporation of nanotechnology [308,309,310].
The (TiO2) NPs are also introduced into the PMMA of acrylic resins in order to increase their mechanical strength and hardness. Furthermore, the addition of these NPs revealed superior photopolymerization, biocompatibility with low levels of cytotoxicity, mutagenicity, and genotoxicity [311]. The binding effect of TiO2 NPs in the PMMA of acrylic resins has been immensely enhanced for the dental prosthesis [312]. This became possible due to a strong bond formation between the carbonyl group of PMMA and the hydroxyl group of the (TiO2) NPs, resulting in the increased physical, chemical, and mechanical properties of these prostheses [313,314]. In addition, it also reduces the chance of the growth and proliferation of various pathogenic microorganisms, especially Candida albicans, due to the increased antimicrobial activity obtained with the addition of TiO2 NPs [315,316]. The TiO2 NPs are considered to be the best possible option to be added in PMMA because of their low investigated toxicity on living cells as compared to another NPs [317,318,319,320].
The TiO2 NPs have a large spectrum of activity against microorganisms, including both Gram-negative and Gram-positive bacteria and fungi [312]. Previously, when the ratio of TiO2 NPs was increased in PMMA, it further improved the antimicrobial behavior of PMMA by significantly reducing the bacterial adherence [311]. The PMMA incorporation with the mixture of the nanofillers of TiO2 and silicon dioxide mixture had a superior antibacterial activity under UV, which could degrade microorganisms with prolonged exposure. This study concluded that TiO2 NPs are compatible and safe when used in combination with other elements, metals, and alloys for dental purposes [321].

The TiO2 NPs in Oral Dentifrices

The TiO2 NPs are also incorporated in the oral dentifrices in 1–10% by weight for prophylaxis of sensitivity and caries [322]. These NPs are employed in oral dentifrices in the mean particle sizes ranging between 100 and 300 nm, which are easily soluble in enamel, dentine, and cementum and thus prohibit the onset of caries and tooth sensitivity [323]. The strong binding of organic molecules in dental tissues (enamel) to TiO2 NPs allows the release of a small amount fluoride from oral dentifrices which gets trapped in a tooth’s matrix. Eventually, the matrix of the tooth becomes the least soluble with the minimal fluoride release over time [324]. In addition, the previous literature also supports the idea that the increased binding of TiO2 NPs to dental tissues (enamel) reduces the size of hydroxyapatite crystals of enamel, resulting in advanced hue, value, and chroma, which maintains the color of human tooth enamel [325,326,327].

The TiO2 NPs in Dental Adhesives

The adhesives are the restorative materials used in dentistry with minimal intervention [328]. The most commonly used adhesives in dentistry are dentin adhesives [329]. The bond formed between the dentin adhesive and the dentin surface becomes unstable and gets loose over time [330]. This occurs due to the collagenous, hydrophilic nature of the dentin structure and the breakage of the strong polymeric covalent bond between them through hydrolytic degradation. Eventually, the total failure of the adhesive bond to the dentin surface occurs [331]. The incorporation of TiO2 NPs in adhesives to investigate their bonding to human teeth has been carried out [332].
Recently, research incorporating TiO2 NPs in dental adhesives concluded that they had enhanced biocompatibility with reduced solubility and water sorption when compared to the conventional adhesives used in dentistry. Thus, the usage of these NPs is highly recommended in currently available adhesive dentistry [333]. Another research work elaborated that TiO2 NPs added to self-adhesive cements revealed a maximum degree of conversion (DC) when compared to their conventional types [334]. The TiO2 NPs are also capable of improving the mechanical and physicochemical properties of adhesives when used as fillers [335]. Thus, TiO2 NPs are potent enough to be deployed as reinforcing fillers in dental adhesives due to their multiple advantages [336,337].
Previously, TiO2 NPs were mixed with a light-curable orthodontic composite paste to form TiO2 NP-based resin composite adhesives which were able to reduce enamel demineralization and provide antibacterial properties for dental adhesive systems. The shear bond strength and the adhesive remnant index for the nanocomposites with TiO2 NPs were similar to those of the control composite [259]. In another study, nanocarbons containing TiO2 NPs were added together to form biomineralized adhesive material in dentistry. The photocatalytic activity of this newly formed nano-adhesive incorporated with TiO2 proved to interfere with the bacterial acidity and was found to have a positive effect on tooth re-mineralization in simulated body fluids [338].

The TiO2 NPs in Dental Prosthesis (Veneers, Crowns, and Bridges)

The durability and appearance of the tooth is improved by replacing upper enamel layers with artificial materials such as sapphire, diamond, and ceramics because of their good biocompatibility and aesthetics. These materials are brittle and prone to fracture; therefore, NPs are added to improve their mechanical properties [339].
Currently, TiO2 NPs are incorporated in ceramics to enhance the aesthetics and mechanical properties of crowns and bridges. This addition of TiO2 NPs in ceramics not only enhanced their aesthetics but also their mechanical properties when compared to the conventional ceramics without TiO2 NPs [179,260,261,340]. This became possible due to the pure white appearance, fatigue resistance, light weight, low elastic modulus, biocompatibility, corrosion resistance, and strength [180].
Titanium is a naturally available common and pure element that shows low toxicity and no allergic reactions when compared to other metals and elements used for dental prosthesis, including cobalt, beryllium, nickel, iron, steel, stainless steel, chromium, etc. Thus, TiO2 NPs are considered as biosafe, biocompatible, easily available, and cost-effective nanomaterials to be utilized in dental prosthesis such as veneers, crowns, and bridges used in the oral cavity [48,49,341].

The TiO2 NPs in Scaffolds/Bone Grafting of Maxillofacial Regions

Naturally, bone is a nano-scaled material which is less than 100 nm in dimension. The materials used for bone grafts in the maxillo-facial regions, including the face, oral cavity, head, neck, ear, and nose, are composites composed mainly of organic compounds (mainly collagen) reinforced with inorganic ones. The nano-crystallites show a loose microstructure, with nanopores situated between the crystallites. This material structure will be completed by filling the pores in the micrometer area. By following this process, a rough surface area is formed on the boundary layer between the biomaterial and the cell, which is very important for fast cell growth. Special features that should be included in bone graft materials are osteoinduction, good processibility, non-sinter, synthetic, highly porous, nanostructure form, and ability to undergo degradation by osteoclasts [342,343].
The TiO2 NPs can be used in bone grafting because of the increased strength, biocompatibility, low modulus of elasticity, corrosion resistant nature, and the extremely light weight of titanium used in the graft composition as compared to other elements and metals [180]. All these characteristics make titanium an ideal biomaterial, especially for its insertion in the human body in the form of metal prosthesis. Moreover, titanium is the choice of material for making internal metal plates, wires, and meshes, such as fracture fixations, cranioplasty plates, cranioplasty meshes, cranioplasty wires, and coronary artery stents [344,345,346]. Additionally, titanium meshes and scaffolds, as well as prosthesis, are well known for nose, eye, maxilla, mandible, ear reconstruction procedures, and ossicular replacements [261,340,347].

The TiO2 NPs in Endodontics

The titanium metal and its metal oxide alloy, commonly known as titanium oxide, are utilized for making the maxillary and mandibular obturators, files, spreaders carvers, and reamers [260,261]. The various dental instruments made from a combination of nickel and titanium are considered ideal for endodontic procedures as a result of their flexibility that makes them fracture resistant, e.g., files, spreaders, carvers, etc. [348,349,350].

3.4. The TiO2 NPs: The Most Biocompatible Material for Medical and Dental Applications

The ideal oral biomaterial should exhibit good thermal, optical, electrical, biological, mechanical, physical, and chemical properties, along with an outstanding biocompatibility [353]. The TiO2 NPs possess almost all the properties of an ideal oral biomaterial; therefore, these NPs are used as antimicrobial agents, dental restorative material, dental prosthetics; they reduce bacterial adhesion to the tooth surface and provide protection against dental caries, bleaching agents, dental implants, and orthodontics. These TiO2 NPs revealed a long-term effect on dental implants; the surface modification provided more advantages such as less bacterial adhesion, improved hardness, biocompatible, low toxic effect, chemical stability, low density, significant adsorption ability; most importantly, they are cost-effective and have strong antimicrobial activity and enhanced mechanical properties. They also displayed extraordinary functionality, long-term stability, durability, fatigue resistance, high mechanical resistance, low thermal diffusivity, extremely light weight, corrosion and wear resistance, high electrical conductivity, and low thermal conductivity, thus enhancing their stability, sustainability, and longevity. They became the most versatile, advanced, and adaptable nanomaterials as a result of their immensely increased mechanical strength, hardness, lightness, and white color. The TiO2 NPs display a large spectrum of antimicrobial, antibacterial, antibacterial, antibactericidal, antiparasitic, and antifungal activity against various pathogenic microorganisms, including Gram-negative bacteria, parasites, Gram-positive bacteria, and fungi. The TiO2 NPs are nontoxic, and the American Food and Drug Administration (FDA) has approved TiO2 for use in human food, drugs, cosmetics, and food contact materials [354,355,356,357,358,359]. Previous research [48,49,115,179,312,360,361,362,363,364,365,366,367,368,369,370] in the context of biosafety and biocompatibility concluded no cytotoxic effect of different concentrations (1–20) mg/cm2 after oral administration. The larges dose of 1000 mg/kg BW/day did not produce any toxic effect or allergic reactions on encountering TiO2. Several studies had reported that the intake of different doses of TiO2 NPs/Kg body weight did not lead to considerable capturing of TiO2 from GIT into the blood or different internal organs, and it was also found that absorption from GIT into blood, urine, and other organs was scarce.

4. Conclusions and Futures Prospects

The TiO2 NPs are safe enough to be used in advanced medical and dental applications as a result of their biocompatibility, biosafety, and non-allergic reactions with the human tissues. Until today, size, shape, phase form, surface morphology, coating, and topography were considered as the significant features responsible for their biocompatibility. Current research on TiO2 NPs has revealed that the mode of synthesis is the most important factor responsible for the biocompatible nature of these NPs, favoring their induction in medical and dental applications. This is due to the fact that the synthesis route can be controlled by temperature, pressure, pH, melting point, environment, and stability of solution medium utilized during the synthesis of these NPs. Moreover, these TiO2 NPs are cost effective and easily available and that became the sole reason for their employment. Therefore, future studies should incorporate the most suitable synthesis mode in order to produce biosafe, biocompatible and ecofriendly TiO2 NPs for utilization the in latest nanomedicine and nanodentistry.

Author Contributions

Conceptualization: A.M. and A.J.; methodology: A.M., M.T.K., A.J. and Z.K.; data Collection, A.M., E.M. and A.J.; data processing, A.M., M.T.K., E.M. and A.J.; data analysis and Interpretation, A.M., E.M., A.J., P.J.P., Z.K. and F.A.B.; writing—original draft, A.M., M.T.K., and A.J.; writing—review and editing, A.M., M.T.K., A.J., E.M., Z.K., P.J.P. and F.A.B.; supervision, A.J.; critical review, A.J., Z.K., P.J.P. and F.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahmadi, E.D.; Hafeji, S.; Khurshid, Z.; Imran, E.; Zafar, M.S.; Saeinasab, M.; Sefat, F. Biophotonics in Dentistry. Appl. Sci. 2022, 12, 4254. [Google Scholar] [CrossRef]
  2. Hayat, F.; Sabbir, J.; Khurshid, Z.; Zafar, M.S.; Ghabbani, H.M.; Shahbazi, M.-A.; Sefat, F. Nanoparticles in endodontics. In Biomaterials in Endodontics; Woodhead Publishers: Shaston, UK, 2022; pp. 195–209. [Google Scholar] [CrossRef]
  3. Khurshid, Z.; Zafar, M.; Qasim, S.; Shahab, S.; Naseem, M.; AbuReqaiba, A. Advances in Nanotechnology for Restorative Dentistry. Materials 2015, 8, 717–731. [Google Scholar] [CrossRef] [PubMed]
  4. Zafar, M.S.; Khurshid, Z.; Najeeb, S.; Zohaib, S.; Rehman, I.U. Therapeutic applications of nanotechnology in dentistry. Nanostruct. Oral Med. 2017, 17, 833–862. [Google Scholar] [CrossRef]
  5. Najeeb, S.; Khurshid, Z.; Ghabbani, H.; Zafar, M.S.; Sefat, F. Nano glass ionomer cement: Modification for biodental applications. In Advanced Dental Biomaterials; Woodhead Publishers: Shaston, UK, 2019; pp. 217–227. ISBN 9780081024768. [Google Scholar]
  6. Sagadevan, S.; Imteyaz, S.; Murugan, B.; Anita, L.J.; Sridewi, N.; Weldegebrieal, G.; Fatimah, I.; Oh, W.A. Comprehensive review on green synthesis of titanium dioxide nanoparticles and their diverse biomedical applications. Green Process. Synth. 2022, 11, 44–63. [Google Scholar] [CrossRef]
  7. American Conference of Governmental Industrial Hygienists (ACGIH). Threshold Limit Values and Biological Exposure Indices for 1992; American Conference of Governmental Industrial Hygienists: Cincinnati, OH, USA, 1992. [Google Scholar]
  8. ILSI Risk Science Institute. The relevance of the rat lung response to particle overload for human risk assessment: A workshop consensus report. Inhal. Toxicol. 2000, 12, 1–17. [Google Scholar] [CrossRef] [PubMed]
  9. Maynard, A.D.; Kuempel, E.D. Airborne Nanostructured Particles and Occupational Health. J. Nanopart. Res. 2005, 7, 587–614. [Google Scholar] [CrossRef]
  10. Tsuji, J.S.; Maynard, A.D.; Howard, P.C.; James, J.T.; Lam, C.W.; Warheit, D.B.; Santamaria, A.B. Research strategies for safety evaluation of nanomaterials, part IV: Risk assessment of nanoparticles. Toxicol. Sci. Off. J. Soc. Toxicol. 2006, 89, 42–50. [Google Scholar] [CrossRef]
  11. Aslam, M.; Abdullah, A.Z.; Rafatullah, M. Recent development in the green synthesis of titanium dioxide nanoparticles using plant-based biomolecules for environmental and antimicrobial applications. J. Ind. Eng. Chem. 2021, 98, 1–16. [Google Scholar] [CrossRef]
  12. Ahmed, A.A.; Afzal, N.; Devarajan, M.; Subramani, S. Structural, morphological, optical and electrical properties of NiO films prepared on Si (100) and glass substrates at different thicknesses. Mater. Res. Express 2016, 3, 116405. [Google Scholar] [CrossRef]
  13. Medhi, R.; Marquez, M.D.; Lee, T.R. Visible-light-active doped metal oxide nanoparticles: Review of their synthesis, properties, and applications. ACS Appl. Nano Mater. 2020, 3, 6156–6185. [Google Scholar] [CrossRef]
  14. Garcia, M.F.; Rodriguez, J.A. Metal Oxide Nanoparticles. In Nanomaterials: Inorganic and Bioinorganic Perspectives; BNL-79479-2007-BC; Brookhaven National Laboratory: Long Island, NY, USA, 2007. [Google Scholar]
  15. Nabi, G.; Qurat-ul-Aain, K.N.R.; Tahir, M.B.; Rafique, M.; Rizwan, M.; Hussain, S.; Iqbal, T.; Majid, A. A Review on Novel Eco-Friendly Green Approach to Synthesis Titanium dioxide (TiO2) nanoparticles Using Different Extracts. J. Inorg. Organomet. Polym. Mater. 2018, 28, 1552–1564. [Google Scholar] [CrossRef]
  16. Lazar, M.A.; Varghese, S.; Nair, S.S. Photocatalytic water treatment by Titanium dioxide: Recent updates. Catalysts 2012, 2, 572–601. [Google Scholar] [CrossRef]
  17. Zhang, W.; Li, Y.; Su, Y.; Mao, K.; Wang, Q. Effect of water composition on TiO2 photocatalytic removal of endocrine disrupting compounds (EDCs) and estrogenic activity from secondary effluent. J. Hazard Mater. 2012, 215, 252–258. [Google Scholar] [CrossRef] [PubMed]
  18. Dastjerdi, R.; Montazer, M. A review on the application of inorganic nano-structured materials in the modification of textiles: Focus on anti-microbial properties. Colloids Surf. B Biointerfaces. 2010, 79, 5–18. [Google Scholar] [CrossRef] [PubMed]
  19. Kubacka, A.; Diez, M.; Rojo, D. Understanding the antimicrobial mechanism of TiO2-based nanocomposite films in a pathogenic bacterium. Sci. Rep. 2014, 4, 4134. [Google Scholar] [CrossRef]
  20. Venkatasubbu, D.; Ramasamy, G.; Ramakrishnan, S.; Janakiraman, K. Folate targeted PEGylated titanium dioxide nanoparticles as a nanocarrier for targeted paclitaxel drug delivery. Adv. Powder Technol. 2013, 24, 947–954. [Google Scholar] [CrossRef]
  21. Yin, Z.F.; Wu, L.; Yang, H.G.; Su, Y.H. Recent progress in biomedical applications of titanium dioxide. Phys. Chem. Chem. Phys. 2013, 15, 4844–4858. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, R.; Hashimoto, K.; Fujishima, A.; Chikuni, M.; Kojima, E.; Kitamura, A.; Shimohigoshi, M.; Watanabe, T. Light-induced amphiphilic surfaces. Nature 1997, 388, 431–432. [Google Scholar] [CrossRef]
  23. Xu, P.; Wang, R.; Ouyang, J.; Chen, B. A new strategy for TiO2 whiskers mediated multi-mode cancer treatment. Nanoscale Res. Lett. 2015, 10, 94–104. [Google Scholar] [CrossRef] [PubMed]
  24. Comune, M.; Rai, A.; Palma, P.; Tondaturo, C.; Ferreira, L. Antimicrobial and pro-angiogenic properties of soluble and nanoparticle-immobilized LL37 peptides. Biomater. Sci. 2021, 9, 8153–8159. [Google Scholar] [CrossRef]
  25. Liu, E.; Zhou, Y.; Liu, Z.; Li, J.; Zhang, D.; Chen, J.; Cai, Z. Cisplatin loaded hyaluronic acid modified Titanium dioxide (TiO2) nanoparticles for neoadjuvant chemotherapy of ovarian cancer. J. Nanomater. 2015, 2015, 1–8. [Google Scholar]
  26. Behnam, M.A.; Emami, F.; Sobhani, Z.; Dehghanian, A.R. The application of titanium dioxide (TiO2) nanoparticles in the photo-thermal therapy of melanoma cancer model. Iran. J. Basic Med. Sci. 2018, 21, 1133–1139. [Google Scholar] [CrossRef]
  27. Abushowmi, T.H.; AlZaher, Z.A.; Almaskin, D.F.; Qaw, M.S.; Abualsaud, R.; Akhtar, S.; Al-Thobity, A.M.; Al-Harbi, F.A.; Gad, M.M.; Baba, N.Z. Comparative Effect of Glass Fiber and Nano-Filler Addition on Denture Repair Strength. J. Prosthodont. 2020, 29, 261–268. [Google Scholar] [CrossRef]
  28. Xia, Y.; Zhang, F.; Xie, H.; Gu, N. Nanoparticle-reinforced resin-based dental composites. J. Dent. 2008, 36, 450–455. [Google Scholar] [CrossRef]
  29. Ohkubo, C.; Hanatani, S.; Hosoi, T. Present status of titanium removable dentures—A review of the literature. J. Oral Rehabil. 2008, 35, 706–714. [Google Scholar] [CrossRef]
  30. Sodagar, A.; Akhoundi, M.; Bahador, A.; Jalali, Y.F.; Behzadi, Z.; Elhaminejad, F.; Mirhashemi, A.H. Effect of Titanium dioxide (TiO2) nanoparticles incorporation on antibacterial properties and shear bond strength of dental composite used in Orthodontics. Dent. Press J. Orthod. 2017, 22, 67–74. [Google Scholar] [CrossRef]
  31. Samiei, M.; Janani, M.; Asl-Aminabadi, N.; Ghasemi, N.; Divband, B.; Shirazi, S.; Kafili, K. Effect of the Titanium dioxide (TiO2) nanoparticles on the selected physical properties of mineral trioxide aggregate. J. Clin. Exp. Dent. 2017, 9, e191–e195. [Google Scholar]
  32. Bertani, R.; Bartolozzi, A.; Pontefisso, A.; Quaresimin, M.; Zappalorto, M. Improving the Antimicrobial and Mechanical Properties of Epoxy Resins via Nanomodification: An Overview. Mol. Ecules 2021, 26, 5426. [Google Scholar] [CrossRef]
  33. Lhotan, A.; Yates, J.; Zidan, S.; Haider, J.; Silikas, N. Assessing Fracture Toughness and Impact Strength of PMMA Reinforced with Nanoparticles and Fibre as Advanced Denture Base Materials. Materials 2021, 24, 4127. [Google Scholar] [CrossRef]
  34. Chambers, C.; Stewart, S.B.; Su, B.; Jankinson, H.E.; Sandy, J.; Ireland, A. Silver doped titanium dioxide nanoparticles as antimicrobial additives to dental polymers. Dent. Mater. 2017, 33, e115–e123. [Google Scholar] [CrossRef]
  35. Colvin, V.L.; Schlamp, M.C.; Alivisatos, A.P. Light emitting diodes made from cadmium selenide nanocrystals and a semiconducting polymer. Nature 1994, 370, 354–357. [Google Scholar] [CrossRef]
  36. Wang, Y.H.N. Nanometer-sized semiconductor clusters: Materials synthesis, quantum size effects, and photophysical properties. J. Phys. Chem. 1991, 95, 525–532. [Google Scholar] [CrossRef]
  37. Atul, R.I.; Thakare, S.R.; Khati, N.T.; Wankhade, A.V.; Burghate, D.K. Green synthesis of selenium nanoparticles under ambient condition. Chalcogenide Lett. 2010, 7, 485–489. [Google Scholar]
  38. Ozak, S.T.; Ozkan, P. Nanotechnology and dentistry. Eur. J. Dent. 2013, 7, 145–151. [Google Scholar]
  39. Saravana, K.R.; Vijayalakshmi, R. Nanotechnology in Dentistry. Ind. J. Den. Res. 2006, 17, 62–65. [Google Scholar]
  40. Ochekpe, A.N.; Olorunfemi, P.O.; Ngwuluka, N.C. Nanotechnology and Drug Delivery Part 1: Background and Applications. Trop. J. Pharm. Res. 2009, 8, 265–274. [Google Scholar] [CrossRef]
  41. Swihart, M.T. Vapor-phase synthesis of nanoparticles. Curr. Opin. Colloid. Interface Sci. 2003, 8, 127–133. [Google Scholar] [CrossRef]
  42. Narducci, D. An introduction to nanotechnologies: What’s in it for us? Vet. Res. Commun. 2007, 31 (Suppl. 1), 131–137. [Google Scholar] [CrossRef]
  43. Jayapalan, A.R.; Lee, B.Y.; Kurtis, K.E. Effect of Nano-sized Titanium Dioxide on Early Age Hydration of Portland Cement. In Nanotechnology in Construction 3; Springer: Berlin/Heidelberg, Germany, 2009; pp. 267–273. [Google Scholar]
  44. Srimaneepong, V.; Skallevold, H.E.; Khurshid, Z.; Zafar, M.S.; Rokaya, D.; Sapkota, J. Graphene for Antimicrobial and Coating Application. Int. J. Mol. Sci. 2022, 23, 499. [Google Scholar] [CrossRef]
  45. Foong, L.K.; Foroughi, M.M.; Mirhosseini, A.F.; Safaei, M.; Jahani, S.; Mostafavi, M.; Ebrahimpoor, N.; Sharifi, M.; Varma, R.S.; Khatami, M. Applications of nano-materials in diverse dentistry regimes. RSC Adv. 2020, 10, 15430–15460. [Google Scholar] [CrossRef]
  46. Yan, L.; Zhao, F.; Wang, J.; Zu, Y.; Gu, Z.; Zhao, Y. A Safe-by-Design Strategy towards Safer Nanomaterials in Nanomedicines. Adv. Mater. 2019, 31, 1805391. [Google Scholar] [CrossRef]
  47. Niinomi, M. Mechanical biocompatibilities of titanium alloys for biomedical applications. J. Mech. Behav. Biomed. Mater. 2008, 1, 30–42. [Google Scholar] [CrossRef]
  48. Burello, E.; Worth, A. Predicting toxicity of nanoparticles. Nat. Nanotechnol. 2011, 6, 138–139. [Google Scholar] [CrossRef]
  49. Tschernitschek, H.; Borchers, L.; Geurtsen, W. Nonalloyed titanium as a bioinert metal—A review. Quintessence Int. 2005, 36, 351. [Google Scholar] [CrossRef]
  50. Li, Y.; Fan, Y.; Chen, Y. A novel method for preparation of nanocrystalline rutile TiO2 powders by liquid hydrolysis of TiCl. J. Mater. Chem. 2002, 12, 1387–1390. [Google Scholar] [CrossRef]
  51. Wang, W.N.; Lenggoro, I.W.; Terashi, Y.; Kim, T.O.; Okuyama, K. One-step synthesis of titanium oxide nanoparticles by spray pyrolysis of organic precursors. Mater. Sci. Eng. 2005, 123, 194–202. [Google Scholar] [CrossRef]
  52. Warheit, D.B.; Webb, K.L.; Reed, S.; Frerichs and Sayes, C.M. Pulmonary toxicity study in rats with three forms of ultrafi ne-TiO2 particles: Diff erential responses related to surface properties. Toxicology 2007, 230, 90–104. [Google Scholar] [CrossRef]
  53. Sayes, C.M.; Wahi, R.; Kurian, P.A.; Liu, Y.; West, J.L.; Ausman, K.D.; Warheit, D.B.; Colvin, V.L. Correlating nanoscale titania structure with toxicity: A cytotoxicity and inflammatory response study with human dermal fibroblasts and human lung epithelial cells. Toxicol. Sci. 2006, 92, 174–185. [Google Scholar] [CrossRef]
  54. Balayeva, N.O.; Mamiyev, Z.; Dillert, R.; Zheng, N.; Bahnemann, D.W. Rh/TiO2-Photocatalyzed Acceptorless Dehydrogenation of N-Heterocycles upon Visible-Light Illumination. ACS Catal. 2020, 10, 5542–5553. [Google Scholar] [CrossRef]
  55. Balayeva, N.O.; Fleisch, M.; Bahnemann, D.W. Surface-grafted WO3/TiO2 photocatalysts: Enhanced visible-light activity towards indoor air purification. Catal. Today 2018, 313, 63–71. [Google Scholar] [CrossRef]
  56. Balayeva, N.O.; Zheng, N.; Dillert, R.; Bahnemann, D.W. Visible-Light-Mediated Photocatalytic Aerobic Dehydrogenation of N-heterocycles by Surface-Grafted TiO2 and 4-amino-TEMPO. ACS Catal. 2019, 9, 10694–10704. [Google Scholar] [CrossRef]
  57. Begum, S.R.; Rao, D.M.; Reddy, P.D.S. Role of Green Route Synthesized Silver Nanoparticles in Medicinal Applications with Special Reference to Cancer Therapy. Biosciences Biotechnology. Res. Asia 2018, 15, 783–790. [Google Scholar] [CrossRef]
  58. Roy, A.; Elzaki, A.; Tirth, V.; Kajoak, S.; Osman, H.; Algahtani, A.; Islam, S.; Faizo, N.L.; Khandaker, M.U.; Islam, M.N.; et al. Biological synthesis of nanocatalysts and their applications. Catalysts 2021, 11, 1494. [Google Scholar] [CrossRef]
  59. Singh, S.C.; Mishra, S.K.; Srivastava, R.K.; Gopal, R. Optical properties of selenium quantum dots produced with laser irradi- ation of water suspended Se nanoparticles. J. Phys. Chem. 2010, 114, 17374–17384. [Google Scholar]
  60. Burne, R.A. Oral streptococci... products of their environment. J. Dent. Res. 1998, 77, 445–452. [Google Scholar] [CrossRef]
  61. Ajdić, D.; McShan, W.M.; McLaughlin, R.E.; Savić, G.; Chang, J.; Carson, M.B.; Primeaux, C.; Tian, R.; Kenton, S.; Jia, S.; et al. Genome sequence of Streptococcus mutans UA159, a cariogenic dental pathogen. Proc. Natl. Acad. Sci. USA 2002, 99, 14434–14439. [Google Scholar] [CrossRef]
  62. Lemos, J.A.; Quivey, R.G.; Koo, H.; Abranches, J. Streptococcus mutans: A new Gram-positive paradigm? Microbiology 2013, 159 Pt 3, 436–445. [Google Scholar] [CrossRef]
  63. Muñoz-Sandoval, C.; Muñoz-Cifuentes, M.J.; Giacaman, R.A.; Ccahuana-Vasquez, R.A.; Cury, J.A. Effect of bovine milk on Streptococcus mutans biofilm cariogenic properties and enamel and dentin demineralization. Pediatric Dent. 2012, 34, e197–e201. [Google Scholar]
  64. Liu, C.; Niu, Y.; Zhou, X.; Zhang, K.; Cheng, L.; Li, M.; Li, Y.; Wang, R.; Yang, Y.; Xu, X. Hyperosmotic response of streptococcus mutans: From microscopic physiology to transcriptomic profile. BMC Microbiol. 2013, 13, 275. [Google Scholar] [CrossRef]
  65. Giacaman, R.A. Sugars and beyond. The role of sugars and the other nutrients and their potential impact on caries. Oral Dis. 2018, 24, 1185–1197. [Google Scholar] [CrossRef]
  66. Thakkar, K.N.; Mhatre, S.S.; Parikh, R.Y. Biological synthesis of metallic nanoparticles. Nanomedicine 2010, 6, 257–262. [Google Scholar] [CrossRef]
  67. Nadeem, M.; Tungmunnithum, D.; Hano, C.; Abbasi, B.H.; Hashmi, S.S.; Ahmad, W. The current trends in the green syntheses of titanium oxide nanoparticles and their applications. Green Chem. Lett. Rev. 2018, 11, 492–502. [Google Scholar] [CrossRef]
  68. Joshi, D.R.; Adhikari, N. An overview on common organic solvents and their toxicity. J. Pharm. Res. Int. 2019, 28, 1–18. [Google Scholar] [CrossRef]
  69. Grewal, A.S.; Kumar, K.; Redhu, S.; Bhardwaj, S. Microwave assisted synthesis: A green chemistry approach. Int. Res. J. Pharm. Appl. Sci. 2013, 3, 278–285. [Google Scholar]
  70. Sharma, R.; Sarkar, A.; Jha, R.; Kumar Sharma, A.; Sharma, D. Sol-gel–mediated synthesis of TiO2 nanocrystals: Structural, optical, and electrochemical properties. Int. J. Appl. Ceram. Technol. 2020, 17, 1400–1409. [Google Scholar] [CrossRef]
  71. Rahimi, H.-R.; Doostmohammadi, M. Nanoparticle synthesis, applications, and toxicity. In Applications of Nanobiotechnology; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef]
  72. Jensen, M.P.; Patterson, D.R. Hypnotic approaches for chronic pain management: Clinical implications of recent research findings. Am. Psychol. 2014, 69, 167–177. [Google Scholar] [CrossRef] [PubMed]
  73. Ebnalwaled, A.A.; Essai, M.H.; Hasaneen, B.M.; Mansour, H.E. Facile and surfactant-free hydrothermal synthesis of PbS nanoparticles: The role of hydrothermal reaction time. J. Mater. Sci. Mater. Electron. 2017, 28, 1958–1965. [Google Scholar] [CrossRef]
  74. Masuda, H.; Yamada, H.; Satoh, M.; Asoh, H.; Nakao, M.; Tamamura, T. Highly ordered nano channel-array architecture in anodic alumina. Appl. Phys. Lett. 1997, 71, 2770. [Google Scholar] [CrossRef]
  75. Kumar, N.; Ali, S.; Kumar, B.; Zafar, M.S.; Khurshid, Z. Hydroxyapatite and nanocomposite implant coatings. In Dental Implants: Materials, Coatings, Surf. ace Modifications and Interfaces with Oral Tissues; Woodhead Publishers: Shaston, UK, 2020; pp. 69–92. ISBN 9780128195864. [Google Scholar]
  76. Liu, N.; Chen, X.; Zhang, J.; Schwank, J.W. A review on TiO2-based nanotubes synthesized via hydrothermal method: Formation mechanism, structure modification, and photocatalytic applications. Catal. Today 2014, 225, 34–51. [Google Scholar] [CrossRef]
  77. Kumar, N.; Fareed, M.A.; Zafar, M.S.; Ghani, F.; Khurshid, Z. Influence of various specimen storage strategies on dental resin-based composite properties. Mater. Technol. 2021, 36, 54–62. [Google Scholar] [CrossRef]
  78. Singh, J.; Kumar, S.; Alok, A.; Upadhyay, S.K.; Rawat, M.; Tsang, D.C. The potential of green synthesized zinc oxide nanoparticles as nutrient source for plant growth. J. Clean. Prod. 2019, 214, 1061–1070. [Google Scholar] [CrossRef]
  79. Mansoor, A.; Khan, M.T.; Mehmood, M.; Khurshid, Z.; Ali, M.I.; Jamal, A. Synthesis and Characterization of Titanium Oxide Nanoparticles with a Novel biological Process for Dental Application. Nanomaterials 2022, 12, 1078. [Google Scholar] [CrossRef] [PubMed]
  80. Mansoor, A.; Khurshid, Z.; Mansoor, E.; Khan, M.T.; Ratanayake, J.; Jamal, A. Effect of Currently Available Nanoparticle Synthesis Routes on Their Biocompatibility with Fibroblast Cell Lines. Molecules 2022, 27, 6972. [Google Scholar] [CrossRef]
  81. Mansoor, A.; Moeen, F.; Hussain, A.; Din, S.U.; Khan, M.T.; Said, F. Age related changes in physiology of normal human tooth enamel: A review. Pak. J. Physiol. 2020, 16, 35–40. [Google Scholar]
  82. Ahmad, A.; Mukherjee, P.; Senapati, S. Extracellular biosynthesis of silver nanoparticles using the fungus Fusarium oxysporum. Colloids Surf. B Biointerfaces 2003, 28, 313–318. [Google Scholar] [CrossRef]
  83. Korbekandi, H.; Iravani, S.; Abbasi, S. Production of nanoparticles using organisms. Crit. Rev. Biotechnol. 2009, 29, 279–306. [Google Scholar] [CrossRef] [PubMed]
  84. Kharissova, O.V.; Dias, H.V.; Kharisov, B.I.; Pérez, B.O.; Pérez, V.M. The greener synthesis of nanoparticles. Trends Biotechnol. 2013, 31, 240–248. [Google Scholar] [CrossRef] [PubMed]
  85. Ramaswamy, V.; Jagtap, N.B.; Vijayanand, S.; Bhange, D.S.; Awati, P.S. Photocatalytic decomposition of methylene blue on nanocrystalline titania prepared by different methods. Mater. Res. Bull. 2008, 43, 1145–1152. [Google Scholar] [CrossRef]
  86. Sethy, N.K.; Arif, Z.; Mishra, P.K.; Kumar, P. Green synthesis of Titanium dioxide (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]
  87. 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] [PubMed]
  88. Irshad, M.A.; Nawaz, R.; Rehman, M.Z.U.; Adrees, M.; Rizwan, M.; Ali, S.; Ahmad, S.; Tasleem, S. Synthesis, characterization and advanced sustainable applications of titanium dioxide nanoparticles: A review. Ecotoxicol. Environ. Saf. 2021, 212, 111978. [Google Scholar] [CrossRef] [PubMed]
  89. Ealias, A.M.; Saravanakumar, M.P. A review on the classification, characterisation, synthesis of nanoparticles and their application. IOP Conf. Ser. Mater. Sci. Eng. 2017, 263, 032019. [Google Scholar]
  90. Kawai-Nakamura, A.; Sato, T.; Sue, K.; Tanaka, S.; Saitoh, K.; Aida, K.; Hiaki, T. Rapid and continuous hydrothermal synthesis of metal and metal oxide nanoparticles with a microtube-reactor at 523 K and 30 MPa. Mater. Lett. 2008, 62, 3471–3473. [Google Scholar] [CrossRef]
  91. Li, H.; Duan, X.; Liu, G.; Jia, X.; Liu, X. Morphology controllable synthesis of TiO2 by a facile hydrothermal process. Mater. Lett. 2008, 62, 4035–4037. [Google Scholar] [CrossRef]
  92. Patra, J.K.; Baek, K.H. Green Nanobiotechnology: Factors Affecting Synthesis and Characterization Techniques. J. Nanomater. 2014, 2014, 1–12. [Google Scholar] [CrossRef]
  93. Jameel, M.S.; Aziz, A.A.; Dheyab, M.A. Green synthesis: Proposed mechanism and factors influencing the synthesis of platinum nanoparticles. Green Process. Synth. 2020, 9, 386–398. [Google Scholar] [CrossRef]
  94. Logeswari, P.; Silambarasan, S.; Abraham, J. Synthesis of silver nanoparticles using plants extract and analysis of their antimicrobial property. J. Saudi. Chem. Soc 2015, 19, 311317. [Google Scholar] [CrossRef]
  95. Sadeghi, B.; Gholamhoseinpoor, F. A study on the stability and green synthesis of silver nanoparticles using Ziziphora tenuior (Zt) extract at room temperature. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 134, 310–315. [Google Scholar] [CrossRef] [PubMed]
  96. Ulug, B.; Turkdemir, M.H.; Cicek, A.; Mete, A. Role of irradiation in the green synthesis of silver nanoparticles mediated by fig (Ficus carica) leaf extract. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 135, 153–161. [Google Scholar] [CrossRef]
  97. Gopinath, V.; Mubarak, A.D.; Priyadarshini, S.; Priyadharsshini, N.M.; Thajuddin, N.; Velusamy, P. Biosynthesis of silver nanoparticles from Tribulus terrestris and its antimicrobial activity: A novel biological approach. Colloids Surf. B Biointerfaces 2012, 96, 69–74. [Google Scholar] [CrossRef]
  98. Hodoroaba, V.; Rades, S.; Unger, W.E.S. Inspection of morphology and elemental imaging of single nanoparticles by high- resolution SEM/EDX in transmission mode. Surf. Interface Anal. 2014, 46, 945–948. [Google Scholar] [CrossRef]
  99. Kumar, N.; Ghani, F.; Fareed, M.A.; Riaz, S.; Khurshid, Z.; Zafar, M.S. Bi-axial flexural strength of resin based dental composites–Influence and reliability of the testing method configuration. Mater. Technol. 2021, 37, 2166–2172. [Google Scholar] [CrossRef]
  100. Yano, F.; Hiraoka, A.; Itoga, T.; Kojima, H.; Kanehori, K.; Mitsui, Y. Influence of ion implantation on native oxidation of Si in a clean-room atmosphere. Appl. Surf. Sci. 1996, 100–101, 138–142. [Google Scholar] [CrossRef]
  101. Abisharani, J.M.; Devikala, S.; Kumar, R.D.; Arthanareeswari, M.; Kamaraj, P. Green synthesis of Titanium dioxide (TiO2) nanoparticles using Cucurbita pepo seeds extract. Mater. Today Proc. 2019, 14, 302–307. [Google Scholar] [CrossRef]
  102. Goutam, S.P.; Saxena, G.; Singh, V.; Yadav, A.K.; Bharagava, R.N.; Thapa, K.B. Green synthesis of Titanium dioxide (TiO2) nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. Chem. Eng. J. 2018, 336, 386–396. [Google Scholar] [CrossRef]
  103. Zafar, M.S.; Alnazzawi, A.A.; Alrahabi, M.; Fareed, M.A.; Najeeb, S.; Khurshid, Z. Nanotechnology and Nanomaterials in Dentistry; Elsevier Ltd.: Basel, Switzerland, 2019. [Google Scholar]
  104. Zakrzewski, W.; Dobrzyński, M.; Zawadzka-Knefel, A.; Lubojański, A.; Dobrzyński, W.; Janecki, M.; Kurek, K.; Szymonowicz, M.; Wiglusz, R.J.; Rybak, Z. Nanomaterials Application in Endodontics. Materials 2021, 14, 5296. [Google Scholar] [CrossRef]
  105. Solano, R.A.; Herrera, A.P.; Maestre, D.; Cremades, A. Fe-Titanium dioxide (TiO2) nanoparticles synthesized by green chemistry for potential application in waste water photocatalytic treatment. J. Nanotechnol. 2019, 2019, 1–11. [Google Scholar] [CrossRef]
  106. Maar, R.R.; Zhang, R.; Stephens, D.G.; Ding, Z.; Gilroy, J.B. Near-infrared photoluminescence and electrochemiluminescence from a remarkably simple boron difluoride formazanate dye. Angew. Chem. Int. 2019, 58, 1052–1056. [Google Scholar] [CrossRef]
  107. Yeganeh-Faal, A.; Bordbar, M.; Negahdar, N.; Nasrollahzadeh, M. Green synthesis of the Ag/ZnO nanocomposite using Valeriana officinalis L. root extract: Application as a reusable catalyst for the reduction of organic dyes in a very short time. IET Nanobiotechnol 2017, 11, 669–676. [Google Scholar] [CrossRef]
  108. Irshad, M.A.; Nawaz, R.; ur Rehman, M.Z.; Imran, M.; Ahmad, M.J.; Ahmad, S.; Ali, S. Synthesis and characterization of titanium dioxide nanoparticles by chemical and green methods and their antifungal activities against wheat rust. Chemosphere 2020, 258, 127352. [Google Scholar] [CrossRef]
  109. Jayaseelan, C.; Elango, G. Evaluation of Catharanthus roseus leaf extract mediated biosynthesis of titanium dioxide nanoparticles against Hippobosca maculata and Bovicola ovis. Parasitol. Res. 2012, 111, 2329–2337. [Google Scholar]
  110. Ambika, S.; Sundrarajan, M. [EMIM] BF4 ionic liquid-mediated synthesis of iO2 nanoparticles using Vitex negundo Linn extract and its antibacterial activity. J. Mol. Liq. 2016, 221, 986–992. [Google Scholar]
  111. Sarwar, M.S.; Huang, Q.; Ghaffar, A.; Abid, M.A.; Zafar, M.S.; Khurshid, Z.; Latif, M. A smart drug delivery system based on biodegradable chitosan/poly(Allylamine hydrochloride) blend films. Pharmaceutics 2020, 12, 131. [Google Scholar] [CrossRef]
  112. Sharma, V.; Rao, L.J. An overview on chemical composition, bioactivity and processing of leaves of Cinnamomum tamala. Crit. Rev. Food Sci. Nutr. 2014, 54, 433–448. [Google Scholar] [CrossRef]
  113. Jalili, N.; Laxminarayana, K. A review of atomic force microscopy imaging systems: Application to molecular metrology and biological sciences. Mechatronics 2004, 14, 907–945. [Google Scholar] [CrossRef]
  114. Butt, H.-J.; Cappella, B.; Kappl, M. Force measurements with the atomic force microscope: Technique, interpretation and applications. Surf. Sci. Rep. 2005, 59, 1–152. [Google Scholar] [CrossRef]
  115. Yang, H.; Wang, Y.; Lai, S.; An, H.; Li, Y.; Chen, F. Application of Atomic Force Microscopy as a Nanotechnology Tool in Food Science. J. Food Sci. 2007, 72, R65–R75. [Google Scholar] [CrossRef]
  116. Cao, Z.; Zhu, W.; Wang, W.; Zhang, C.; Xu, M.; Liu, J.; Feng, S.T.; Jiang, Q.; Xie, X. Stable cerasomes for simultaneous drug delivery and magnetic resonance imaging. Int. J. Nanomed. 2014, 9, 5103–5116. [Google Scholar] [CrossRef] [PubMed]
  117. Louro, H.; Saruga, A.; Santos, J.; Pinhão, M.; Silva, M.J. Biological impact of metal nanomaterials in relation to their physicochemical characteristics. Toxicol. Vitr. 2019, 56, 172–183. [Google Scholar] [CrossRef]
  118. Louro, H. Relevance of physicochemical characterization of nanomaterials for understanding nano-cellular interactions. Adv. Exp. Med. Biol. 2018, 1048, 123–142. [Google Scholar]
  119. Jeyaraj, M.; Gurunathan, S.; Qasim, M.; Kang, M.H.; Kim, J.H. A Comprehensive Review on the Synthesis, Characterization, and Biomedical Application of Platinum Nanoparticles. Nanomaterials 2019, 9, 1719. [Google Scholar] [CrossRef]
  120. Jacobs, P.W.; Wind, S.J.; Ribeiro, F.H.; Somorjai, G.A. Nanometer size platinum particle arrays: Catalytic and surface chemical properties. Surf. Sci. 1997, 372, L249–L253. [Google Scholar] [CrossRef]
  121. Issa, B.; Obaidat, I.M.; Albiss, B.A.; Haik, Y. Magnetic nanoparticles: Surface effects and properties related to biomedicine applications. Int. J. Mol. Sci. 2013, 14, 21266–21305. [Google Scholar] [CrossRef]
  122. Braun, J.H.; Baidins, A.; Marganski, R.E. TiO2 pigment technology: A review. Prog. Org. Coat. 1992, 20, 105–138. [Google Scholar] [CrossRef]
  123. Shand, M.; Anderson, A.J. Aqueous phase photocatalytic nitrate destruction using titania-based materials: Routes to enhanced performance and prospects for visible light activation. Catal. Sci. Technol. 2013, 3, 879–899. [Google Scholar] [CrossRef]
  124. Gopinath, K.P.; Madhav, N.V.; Krishnan, A.; Malolan, R.; Rangarajan, G. Present applications of titanium dioxide for the photocatalytic removal of pollutants from water: A review. J. Environ. Manag. 2020, 270, 110906. [Google Scholar] [CrossRef]
  125. Cihlar, J.; Navarro, L.K.T.; Kasparek, V.; Michalicka, J.; Kastyl, J.; Castkova, K.; Celko, L. Inflfluence of LA/Ti molar ratio on the complex synthesis of anatase/brookite nanoparticles and their hydrogen production. Int. J. Hydrog. Energy 2021, 46, 8578–8593. [Google Scholar] [CrossRef]
  126. Jing, L.; Zhou, W.; Tian, G.; Fu, H. Surface tuning for oxide-based nanomaterials as effificient photocatalysts. Chem. Soc. Rev. 2013, 42, 9509–9549. [Google Scholar] [CrossRef]
  127. Spampinato, V.; Ceccone, G.; Giordani, S. Surface analysis of zinc-porphyrin functionalized carbon nano-onions. Biointerphases 2015, 10, 019006. [Google Scholar] [CrossRef]
  128. Volokitin, Y.; Sinzig, J.; Dejongh, L.J.; Schmid, G.; Vargaftik, M.N.; Moiseev, I.I. Quantum-size effects in the thermodynamic properties of metallic nanoparticles. Nature 1996, 384, 621–623. [Google Scholar] [CrossRef]
  129. Ziental, D.; Czarczynska-Goslinska, B.; Mlynarczyk, D.T.; Glowacka-Sobotta, A.; Stanisz, B.; Goslinski, T.; Sobotta, L. Titanium dioxide nanoparticles: Prospects and applications in medicine. Nanomaterials 2020, 10, 387. [Google Scholar] [CrossRef]
  130. Hoet, P.H.; Brüske-Hohlfeld, I.; Salata, O.V. Nanoparticles-known and unknown health risks. J. Nanobiotechnol. 2004, 2, 12. [Google Scholar] [CrossRef]
  131. Wang, Z.; Haidry, A.A.; Xie, L.; Zavabeti, A.; Li, Z.; Yin, W.; Fomekong, R.L.; Saruhan, B. Acetone sensing applications of Ag modified TiO2 porous nanoparticles synthesized via facile hydrothermal method. Appl. Sur. Sci. 2020, 533, 147383. [Google Scholar] [CrossRef]
  132. Song, B.; Zhou, T.; Yang, W.L.; Liu, J.; Shao, L.Q. Contribution of oxidative stress to TiO2 nanoparticle-induced toxicity. Environ. Toxicol. Pharmacol. 2016, 48, 130–140. [Google Scholar] [CrossRef]
  133. Risom, L.; Møller, P.; Loft, S. Oxidative stress-induced DNA damage by particulate air pollution. Mutat. Res. 2005, 592, 119–137. [Google Scholar] [CrossRef]
  134. Xia, T.; Kovochich, M.; Brant, J.; Hotze, M.; Sempf, J.; Oberley, T.; Sioutas, C.; Yeh, J.I.; Wiesner, M.R.; Nel, A.E. Comparison of the abilities of ambient and manufactured nanoparticles to induce cellular toxicity according to an oxidative stress paradigm. Nano Lett. 2006, 6, 1794–1807. [Google Scholar] [CrossRef]
  135. Braydich-Stolle, L.K.; Schaeublin, N.M.; Murdock, R.C.; Jiang, J.; Biswas, P.; Schlager, J.J.; Hussain, S.M. Crystal structure mediates mode of cell death in TiO2 nanotoxicity. J. Nanopart. Res. 2009, 11, 1361–1374. [Google Scholar] [CrossRef]
  136. Gratton, S.E.; Ropp, P.A.; Pohlhaus, P.D.; Luft, J.C.; Madden, V.J.; Napier, M.E.; DeSimone, J.M. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad. Sci. USA 2008, 105, 11613–11618. [Google Scholar] [CrossRef]
  137. Ekimov, A.I.; Efros, A.L.; Onushchenko, A.A. Quantum size effect in semiconductor microcrystals. Solid State Commun. 1993, 88, 947–950. [Google Scholar] [CrossRef]
  138. Buzea, C.; Pacheco, I.I.; Robbie, K. Nanomaterials and nanoparticles: Sources and toxicity. Biointerphases 2007, 2, MR17–MR71. [Google Scholar] [CrossRef]
  139. Fröschl, T.; Hörmann, U.; Kubiak, P.; Kučerová, G.; Pfanzelt, M.; Weiss, C.K.; Behm, R.J.; Hüsing, N.; Kaiser, U.; Landfester, K.; et al. High surface area crystalline titanium dioxide: Potential and limits in electrochemical energy storage and catalysis. Chem. Soc. Rev. 2012, 41, 5313–5360. [Google Scholar] [CrossRef]
  140. Shukla, R.; Sharma, V.; Pandey, A.K.; Singh, S.; Sultana, S.; Dhawan, A. ROS-mediated genotoxicity induced by titanium dioxide nanoparticles in human epidermal cells. Toxicol. Vitr. 2011, 25, 231–241. [Google Scholar] [CrossRef]
  141. Balachandran, K.; Mageswari, S.; Preethi, A. Photocatalytic decomposition of A549-lung cancer cells by Titanium dioxide (TiO2) nanoparticles. Mater. Today Proc. 2021, 37, 1071–1074. [Google Scholar] [CrossRef]
  142. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modififications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  143. Nasirian, M.; Mehrvar, M. Photocatalytic degradation of aqueous Methyl Orange using nitrogen-doped TiO2 photocatalyst prepared by novel method of ultravioletassisted thermal synthesis. J. Environ. Sci. 2018, 66, 81–93. [Google Scholar] [CrossRef]
  144. Donaldson, K.; Stone, V. Current hypotheses on the mechanisms of toxicity of ultrafine particles. Ann. Dell’istituto Super. Sanita 2003, 39, 405–410. [Google Scholar]
  145. Luo, X.; Aoife, M.; Anthony, J.K.; Malcolm, R.S. Application of nanoparticles in electrochemical sensors and biosensors. Electroanalysis 2006, 18, 319–326. [Google Scholar] [CrossRef]
  146. Zhang, H.; Huang, X.J.; Liu, X.Y.; Zhu, Q.M.; Liu, Y.; Lin, X.F. Research advances in safety of nanometer materials and their risk assessment system. Chin. J. Process. Eng. 2013, 13, 893–900. [Google Scholar]
  147. Bichowsky, F.V. Titanium White—A new method for its preparation. Ind. Eng. Chem. 1929, 21, 1061–1063. [Google Scholar] [CrossRef]
  148. Warheit, D.; Webb, T.; Reed, K. Pulmonary toxicity screening studies in male rats with TiO2 particulates substantially encapsulated with pyrogenically deposited, amorphous silica. Part. Fibre Toxicol. 2006, 3, 3. [Google Scholar] [CrossRef]
  149. Khorrami, S.; Zarrabi, A.; Khaleghi, M.; Danaei, M.; Mozafari, M.R. Selective cytotoxicity of green synthesized silver nanoparticles against the MCF-7 tumor cell line and their enhanced antioxidant and antimicrobial properties. Int. J. Nanomed. 2018, 13, 8013–8024. [Google Scholar] [CrossRef]
  150. Xue, C.; Wu, J.; Lan, F.; Liu, W.; Yang, X.; Zeng, F.; Xu, H. Nano titanium dioxide induces the generation of ROS and potential damage in HaCaT cells under UVA irradiation. J. Nanosci. Nanotechnol. 2010, 10, 8500–8507. [Google Scholar] [CrossRef]
  151. Petković, J.; Zegura, B.; Stevanović, M.; Drnovšek, N.; Uskoković, D.; Novak, S.; Filipič, M. DNA damage and alterations in expression of DNA damage responsive genes induced by Titanium dioxide (TiO2) nanoparticles in human hepatoma HepG2 cells. Nanotoxicology 2011, 5, 341–353. [Google Scholar] [CrossRef]
  152. Di Virgilio, A.L.; Reigosa, M.; de Mele, M.F. Response of UMR 106 cells exposed to titanium oxide and aluminum oxide nanoparticles. J. Biomed. Mater. Res. Part A 2010, 92, 80–86. [Google Scholar] [CrossRef]
  153. Martinez-Gutierrez, F.; Olive, P.L.; Banuelos, A.; Orrantia, E.; Nino, N.; Sanchez, E.M.; Ruiz, F.; Bach, H.; Av-Gay, Y. Synthesis, characterization, and evaluation of antimicrobial and cytotoxic effect of silver and titanium nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 681–688. [Google Scholar] [CrossRef]
  154. Setyawati, M.I.; Khoo, P.K.; Eng, B.H.; Xiong, S.; Zhao, X.; Das, G.K.; Tan, T.T.; Loo, J.S.; Leong, D.T.; Ng, K.W. Cytotoxic and genotoxic characterization of titanium dioxide, gadolinium oxide, and poly(lactic-co-glycolic acid) nanoparticles in human fibroblasts. J. Biomed. Mater. Res. Part A 2013, 101, 633–640. [Google Scholar] [CrossRef]
  155. Lee, W.G.; Kim, Y.G.; Chung, B.G.; Demirci, U.; Khademhosseini, A. Nano/Microfluidics for diagnosis of infectious diseases in developing countries. Adv. Drug Deliv. Rev 2010, 62, 449–457. [Google Scholar] [CrossRef]
  156. Trovarelli, A. ChemInform Abstract: Catalytic Properties of Ceria and CeO2-Containing Materials. Cheminform 1997, 38, 439–520. [Google Scholar] [CrossRef]
  157. Rokaya, D.; Kongkiatkamon, S.; Heboyan, A.; Dam, V.V.; Amornvit, P.; Khurshid, Z.; Srimaneepong, V.; Zafar, M.S. 3D-Printed Biomaterials in Biomedical Application. In Functional Biomaterials; Springer: Singapore, 2022; pp. 319–339. [Google Scholar]
  158. Lison, D.; Vietti, G.; Brule, S.V.D. Paracelsus in nanotoxicology. Part Fibre Toxicol. 2014, 11, 35. [Google Scholar] [CrossRef]
  159. Kukia, N.R.; Rasmi, Y.; Abbasi, A.; Koshoridze, N.; Shirpoor, A.; Burjanadze, G.; Saboory, E. Bio-effffects of Titanium dioxide (TiO2) nanoparticles on human colorectal cancer and umbilical vein endothelial cell lines. Asian Pac. J. Cancer Prev. 2018, 19, 2821–2829. [Google Scholar]
  160. Berginc, G. Optical properties of nanostructured materials: A review. J. Nanophoton. 2011, 5, 052502. [Google Scholar]
  161. Chen, Y.W.; Liou, G.G.; Pan, H.B.; Tseng, H.H.; Hung, Y.T.; Chou, C.P. Specific detection of CD133-positive tumor cells with iron oxide nanoparticles labeling using noninvasive molecular magnetic resonance imaging. Int. J. Nanomed. 2015, 10, 6997–7018. [Google Scholar]
  162. Golasik, M.; Herman, M.; Piekoszewski, W. Toxicological aspects of soluble titanium—A review of in vitro and in vivo studies. Metallomics 2016, 8, 1227–1242. [Google Scholar] [CrossRef] [PubMed]
  163. Wang, Y.; Sauvat, A.; Lacrouts, C.; Lebeau, J.; Grall, R.; Hullo, M.; Nesslany, F.; Chevillard, S. TiO2 nanomaterials non-controlled contamination could be hazardous for normal cells located in the fifield of radiotherapy. Int. J. Mol. Sci. 2020, 21, 940. [Google Scholar] [CrossRef] [PubMed]
  164. Li, W.; Elzatahry, A.; Aldhayan, D.; Zhao, D. Core–shell structured titanium dioxide nanomaterials for solar energy utilization. Chem. Soc. Rev. 2018, 47, 8203–8237. [Google Scholar] [CrossRef] [PubMed]
  165. Yin, H.; Too, H.P.; Chow, G.M. The effects of particle size and surface coating on the cytotoxicity of nickel ferrite. Biomaterials 2005, 26, 5818–5826. [Google Scholar] [CrossRef] [PubMed]
  166. Gupta, A.K.; Gupta, M. Cytotoxicity suppression and cellular uptake enhancement of surface modified magnetic nanoparticles. Biomaterials 2005, 26, 1565–1573. [Google Scholar] [CrossRef]
  167. Sayes, C.M.; Fortner, J.D.; Guo, W.; Lyon, D.; Boyd, A.M.; Ausman, K.D.; Tao, Y.J.; Sitharaman, B.; Wilson, L.J.; Highes, J.B.; et al. The differential cytotoxicity of water-soluble fullerenes. Nano Lett. 2004, 4, 1881–1887. [Google Scholar] [CrossRef]
  168. Li, J.; Singh, V.V.; Sattayasamitsathit, S.; Orozco, J.; Kaufmann, K.; Dong, R.; Gao, W.; Jurado-Sanchez, B.; Fedorak, Y.; Wang, J. Water-driven micromotors for rapid photocatalytic degradation of biological and chemical warfare agents. ACS Nano 2014, 8, 11118–11125. [Google Scholar] [CrossRef]
  169. Estelrich, J.; Sanchez-Martin, M.J.; Busquets, M.A. Nanoparticles in magnetic resonance imaging: From simple to dual contrast agents. Int. J. Nanomed. 2015, 10, 1727–1741. [Google Scholar]
  170. Palma, P.J.; Messias, A.; Cerqueira, A.R.; Tavares, L.D.; Caramelo, F.; Roseiro, L.; Santos, J.M. Cyclic fatigue resistance of three rotary file systems in a dynamic model after immersion in sodium hypochlorite. Odontology 2019, 107, 324–332. [Google Scholar] [PubMed]
  171. Skopalik, J.; Polakova, K.; Havrdova, M.; Justan, I.; Magro, M.; Milde, D.; Knopfova, L.; Smarda, J.; Polakova, H.; Gabrielova, E.; et al. Mesenchymal stromal cell labeling by new uncoated superparamagnetic maghemite nanoparticles in comparison with commercial Resovist--an initial in vitro study. Int. J. Nanomed. 2014, 9, 5355–5372. [Google Scholar] [CrossRef]
  172. Kolahalam, L.A.; Kasi Viswanath, I.V.; Diwakar, B.S.; Govindh, B.; Reddy, V.; Murthy, Y.L.N. Review on nanomaterials: Synthesis and applications. Mater. Today Proc. 2019, 18, 2182–2190. [Google Scholar] [CrossRef]
  173. Li, X.; Liang, X.; Wang, Y.; Wang, D.; Teng, M.; Xu, H.; Zhao, B.; Han, L. Graphene-Based Nanomaterials for Dental Applications: Principles, Current Advances, and Future Outlook. Front. Bioeng. Biotechnol. 2022, 10, 804201. [Google Scholar] [CrossRef] [PubMed]
  174. Piccinno, F.; Gottschalk, F.; Seeger, S.; Nowack, B. Industrial production quantities and uses of ten engineered Nanomaterials in Europ and the word. J. Nanoparts. Res. 2012, 14, 1109. [Google Scholar] [CrossRef]
  175. Sharma, S.; Sharma, R.K.; Gaur, K.; Cátala Torres, J.F.; Loza-Rosas, S.A.; Torres, A.; Saxena, M.; Julin, M.; Tinoco, A.D. Fueling a hot debate on the application of Titanium dioxide (TiO2) nanoparticles in sunscreen. Materials 2019, 12, 2317. [Google Scholar] [CrossRef]
  176. Rokaya, D.; Bohara, S.; Srimaneepong, V.; Kongkiatkamon, S.; Khurshid, Z.; Heboyan, A.; Zafar, M.S.; Sapkota, J. Metallic Biomaterials for Medical and Dental Prosthetic Applications. In Functional Biomaterials; Springer: Singapore, 2022; pp. 503–522. [Google Scholar]
  177. Tomisa, A.P.; Launey, M.E.; Lee, J.S.; Mankani, M.H.; Wegst, U.G.; Saiz, E. Nanotechnology approaches to improve dental implants. Int. J. Oral Maxillofac. Implant. 2011, 26, 25–49. [Google Scholar]
  178. Kumar, P.; Metha, D.S.; Guvva, S. Nanotechnological impact on future clinical dental prospects: An insight. Eur. J. Gen. Dent. 2013, 2, 86–87. [Google Scholar] [CrossRef]
  179. Tahir, A.; Moeen, F.; Mehmood, M.; Mansoor, A.; Abbas, Z.; Hussain, A.; Kashif, M. Compressive Strength and Flexural Strength of Titanium Nano-Enriched Gic at Different Percentages an in Vitro Study. Ann. Dent. Spec. 2019, 7, 1. [Google Scholar]
  180. Lautenschlager, E.P.; Monaghan, P. Titanium and titanium alloys as dental materials. Int. Dent. J. 1993, 43, 245–253. [Google Scholar] [PubMed]
  181. Kovvuru, S.K.; Mahita, V.N.; Manjunatha, B.; Babu, B.S.; Mahita, V.N. Nanotechnology: The Emerging Science in Dentistry. J. Orofac. Res. 2019, 2, 33. [Google Scholar]
  182. Moghimi, S.M.; Hunter, A.C.; Murray, J.C. Nanomedicine: Current status and future prospects. FASEB J. 2005, 19, 311–330. [Google Scholar] [CrossRef] [PubMed]
  183. Uskoković, V.; Bertassoni, L.E. Nanotechnology in Dental Sciences: Moving towards a Finer Way of Doing Dentistry. Materials 2010, 3, 1674–1691. [Google Scholar] [CrossRef]
  184. Tamirat, Y. The Role of Nanotechnology in Semiconductor Industry: Review Article. J. Mater. Sci. Nanotechnol. 2017, 5, 202. [Google Scholar]
  185. Mantri, S.S.; Mantri, S.P. The nano era in dentistry. J. Nat. Sci. Biol. Med. 2013, 4, 39–44. [Google Scholar] [CrossRef] [PubMed]
  186. Bhuvaneswarri, J.; Alam, M.; Chandrasekaran, S.; Sathya, M. Future impact of nanotechnology in dentistry—A review. Int. J. Nanotechnol. Appl. (IJNA) 2013, 3, 15–20. [Google Scholar]
  187. Chandki, R.; Kala, M.; Kumar, K.N.; Brigit, B.; Banthia, P.; Banthia, R. ‘Nanodentistry’: Exploring the beauty of miniature. J. Clin. Exp. Dent. 2012, 4, e119–e124. [Google Scholar] [CrossRef]
  188. Khan, M.T.; Moeen, F.; Safi, S.Z.; Said, F.; Mansoor, A.; Khan, A. The Structural, Physical, and In Vitro Biological Performance of Freshly Mixed and Set Endodontic Sealers. Eur. Endod. J. 2021, 6, 98–109. [Google Scholar] [CrossRef] [PubMed]
  189. Said, F.; Moeen, F.; Khan, M.T.; Mansoor, A.; Uzbek, U.H.; Alam, M.K.; Siddiqui, A.A. Cytotoxicity, morphology and chemical composition of two luting cements: An in vitro study. Pesqui. Bras. Odontopediatria E Clin. Integr. 2020, 20, 1–8. [Google Scholar] [CrossRef]
  190. Naseri, N.; Janfaza, S.; Irani, R. Visible light switchable br/TiO2 nanostructured photoanodes for bio-inspired solar energy conversion. RSC Adv. 2015, 5, 18642–18646. [Google Scholar] [CrossRef]
  191. Molaeirad, A.; Janfaza, S.; Karimi-Fard, A.; Mahyad, B. Photocurrent generation by adsorption of two main pigments of halobacterium salinarum on TiO2 nanostructured electrode. Biotechnol Appl. Biochem. 2015, 62, 121–125. [Google Scholar] [CrossRef] [PubMed]
  192. McCullagh, C.; Robertson, J.M.; Bahnemann, D.W.; Robertson, P.K. The application of TiO2 photocatalysis for disinfection of water contaminated with pathogenic micro-organisms: A review. Res. Chem. Intermed. 2007, 33, 359–375. [Google Scholar] [CrossRef]
  193. Gao, A.; Hang, R.; Huang, X.; Zaho, L.; Zhang, X.; Wang, L.; Tang, B.; Ma, S.; Chu, P.K. The effects of titania nanotubes with embedded silver oxide nanoparticles on bacteria and osteoblasts. Biomaterials 2014, 35, 4223–4235. [Google Scholar] [CrossRef] [PubMed]
  194. Hang, R.; Gao, A.; Huang, X.; Zhang, X.; Qin, L.; Tang, B. Antibacterial activity and cytocompatibility of cu–ti–o nanotubes. J. Biomed. Mater. Res. A 2014, 102, 1850–1858. [Google Scholar] [CrossRef] [PubMed]
  195. Darvishi, M.H.; Nomani, A.; Hashemzadeh, H.; Amini, M.; Shokrgozar, M.A.; Dinarvand, R. Targeted DNA delivery to cancer cells using a biotinylated chitosan carrier. Biotechnol Appl. Biochem. 2017, 64, 423–432. [Google Scholar] [CrossRef]
  196. Zabarjadi, A.; Hashemzadeh, H.; Taravat, E. Telomere, chromosome end, and telomerase enzyme as a cancer biomarker. Genet Third Millenn. 2013, 11, 3018–3027. [Google Scholar]
  197. Hashemzadeh, H.; Javadi, H.; Darvishi, M.H. Study of structural stability and formation mechanisms in dspc and dpsm liposomes: A coarse-grained molecular dynamics simulation. Sci. Rep. 2020, 10, 1837. [Google Scholar] [CrossRef]
  198. Li, Q.; Wang, X.; Lu, X.; Tian, H.; Jiang, H.; Lv, G.; Guo, D.; Wu, C.; Chen, B. The incorporation of daunorubicin in cancer cells through the use of titanium dioxide whiskers. Biomaterials 2009, 30, 4708–4715. [Google Scholar] [CrossRef] [PubMed]
  199. Akram, M.W.; Raziq, F.; Fakhar-e-Alam, M.; Aziz, K.; Alimgeer, M.; Atif, M.; Amir, A.; Hanif, W.; Farooq, A. Tailoring of au-titanium dioxide (TiO2) nanoparticles conjugated with doxorubicin for their synergistic response and photodynamic therapy applications. J. Photochem. Photobiol. A Chem. 2019, 384, 112040. [Google Scholar] [CrossRef]
  200. Hashemzadeh, H.; Allahverdi, A.; Ertl, P.; Naderi-Manesh, H. Comparison between three-dimensional spheroid and two-dimensional monolayer in a549 lung cancer and pc9 normal cell lines under treatment of silver nanoparticles. Modares J. Biotechnol. 2019, 10, 573–580. [Google Scholar]
  201. Shea, C.M. Future management research directions in nanotechnology: A case study. J. Eng. Technol. Manag. 2005, 22, 185–200. [Google Scholar] [CrossRef]
  202. Nešić, M.D.; Žakula, J.; Korićanac, L.; Stepić, M.; Radoičić, M.; Popović, I.; Šaponjić, Z.V.; Petković, M. Light controlled metallo-drug delivery system based on the TiO2-nanoparticles and Ru-complex. J. Photochem. Photobiol. A-Chem. 2017, 347, 55–66. [Google Scholar] [CrossRef]
  203. Wang, T.; Jiang, H.; Wan, L.; Zhao, Q.; Jiang, T.; Wang, B.; Wang, S. Potential application of functional porous titanium dioxide (TiO2) nanoparticles in light-controlled drug release and targeted drug delivery. Acta Biomater. 2015, 13, 354–363. [Google Scholar] [CrossRef] [PubMed]
  204. Yin, M.; Ju, E.; Chen, Z.; Li, Z.; Ren, J.; Qu, X. Upconverting nanoparticles with a mesoporous TiO2 shell for near-infrared-triggered drug delivery and synergistic targeted cancer therapy. Chem. Eur. J. 2014, 20, 14012–14017. [Google Scholar] [CrossRef]
  205. Chu, X.; Mao, L.; Johnson, O.; Li, K.; Phan, J.; Yin, Q.; Li, L.; Zhang, J.; Chen, W.; Zhang, Y. Exploration of TiO2 nanoparticle mediated microdynamic therapy on cancer treatment. Nanomedicine 2019, 18, 272–281. [Google Scholar] [CrossRef]
  206. Jafari, S.; Lavasanifar, A.; Hejazi, M.S.; Maleki-Dizaji, N.; Mesgari, M.; Molavi, O. Stat3 inhibitory stattic enhances immunogenic cell death induced by chemotherapy in cancer cells. DARU J. Pharm. Sci. 2020, 28, 159–169. [Google Scholar] [CrossRef]
  207. Harada, A.; Ono, M.; Yuba, E.; Kono, K. Titanium dioxide nanoparticle-entrapped polyion complex micelles generate singlet oxygen in the cells by ultrasound irradiation for sonodynamic therapy. Biomater. Sci. 2013, 1, 65–73. [Google Scholar] [CrossRef]
  208. Wang, Q.; Huang, J.Y.; Li, H.Q.; Chen, Z.; Zhao, A.Z.; Wang, Y.; Zhang, K.Q.; Sun, H.T.; Al-Deyab, S.S.; Lai, Y.K. TiO2 nanotube platforms for smart drug delivery: A review. Int. J. Nanomed. 2016, 11, 4819–4834. [Google Scholar] [CrossRef]
  209. Gannon, C.J.; Patra, C.R.; Bhattacharya, R.; Mukherjee, P.; Curley, S.A. Intracellular gold nanoparticles enhance non-invasive radiofrequency thermal destruction of human gastrointestinal cancer cells. J. Nanobiotechnol. 2008, 6, 2. [Google Scholar] [CrossRef] [PubMed]
  210. Calamak, S.; Shahbazi, R.; Eroglu, I.; Gultekinoglu, M.; Ulubayram, K. An overview of nanofiber-based antibacterial drug design. Expert Opin. Drug Discov. 2017, 12, 391–406. [Google Scholar] [CrossRef] [PubMed]
  211. Kaviyarasu, K.; Geetha, N.; Kanimozhi, K.; Maria Magdalane, C.; Siveranjani, S.; Ayeshamariam, A.; Kenned, J.; Mazza, M. In vivo cytotoxicity effect and antibacterial performance of human lung epithelial cells a549 activity of zinc oxide doped TiO2 nanocrystals: Investigation of bio-medical application by chemical method. Mater. Sci. Eng. C 2017, 74, 325–333. [Google Scholar] [CrossRef] [PubMed]
  212. Matsunaga, T.; Tomoda, R.; Nakajima, T.; Wake, H. Photoelectrochemical sterilization of microbial cells by semiconductor powders. FEMS Microbiol. Lett. 1985, 29, 211–214. [Google Scholar] [CrossRef]
  213. Foster, H.A.; Ditta, I.B.; Varghese, S.; Steele, A. Photocatalytic disinfection using titanium dioxide: Spectrum and mechanism of antimicrobial activity. Appl. Microbiol. Biotechnol. 2011, 90, 1847–1868. [Google Scholar] [CrossRef]
  214. Parnia, F.; Yazdani, J.; Javaherzadeh, V.; Dizaj, S.M. Overview of nanoparticle coating of dental implants for enhanced osseointegration and antimicrobial purposes. J. Pharm. Pharm. Sci. 2017, 20, 148–160. [Google Scholar] [CrossRef]
  215. Hou, X.; Mao, D.; Ma, H.; Ai, Y.; Zhao, X.; Deng, J.; Li, D.; Liao, B. Antibacterial ability of ag–TiO2 nanotubes prepared by ion implantation and anodic oxidation. Mater. Lett. 2015, 161, 309–312. [Google Scholar] [CrossRef]
  216. Bakhsheshi-Rad, H.R.; Hamzah, E.; Ismail, A.F.; Aziz, M.; Daroonparvar, M.; Saebnoori, E.; Chami, A. In vitro degradation behavior, antibacterial activity and cytotoxicity of TiO2-MAO/ZnHA composite coating on Mg alloy for orthopedic implants. Surf. Coat. Technol. 2018, 334, 450–460. [Google Scholar] [CrossRef]
  217. Roy, A.S.; Parveen, A.; Koppalkar, A.R.; Prasad, M.A. Effect of nano-titanium dioxide with different antibiotics against methicillin-resistant staphylococcus aureus. J. Biomater. Nanobiotechnol. 2010, 1, 37. [Google Scholar] [CrossRef]
  218. Liu, W.; Su, P.; Chen, S.; Wang, N.; Wang, J.; Liu, Y.; Ma, Y.; Li, H.; Zhang, Z.; Webster, T.J. Antibacterial and osteogenic stem cell differentiation properties of photoinduced TiO2 nanoparticle-decorated TiO2 nanotubes. Nanomedicine 2015, 10, 713–723, retracted. [Google Scholar] [CrossRef]
  219. Jin, Z.; Yan, X.; Shen, K.; Fang, X.; Zhang, C.; Ming, Q.; Lai, M.; Cai, K. Tio2 nanotubes promote osteogenic differentiation of mesenchymal stem cells via regulation of lncrna ccl3-as. Colloids Surf. B Biointerfaces 2019, 181, 416–425. [Google Scholar] [CrossRef]
  220. Bhardwaj, G.; Webster, T.J. Reduced bacterial growth and increased osteoblast proliferation on titanium with a nanophase TiO2 surface reatment. Int. J. Nanomed. 2017, 12, 363. [Google Scholar] [CrossRef]
  221. Velasco, E.; Baldovino-Medrano, V.G.; Gaigneaux, E.M.; Giraldo, S.A. Development of an efficient strategy for coating tio 2 on polyester– cotton fabrics for bactericidal applications. Top Catal. 2016, 59, 378–386. [Google Scholar] [CrossRef]
  222. Chawengkijwanich, C.; Hayata, Y. Development of TiO2 powder-coated food packaging film and its ability to inactivate Escherichia coli in vitro and in actual tests. Int. J. Food Microbiol. 2008, 3, 288–292. [Google Scholar] [CrossRef] [PubMed]
  223. Sekiguchi, Y.; Yao, Y.; Ohko, Y.; Tanaka, K.; Ishido, T.; Fujishima, A.; Kubota, Y. Self-sterilizing catheters with titanium dioxide photocatalyst thin films for clean intermittent catheterization: Basis and study of clinical use. Int. J. Urol. Off. J. Jpn. Urol. Assoc. 2007, 14, 426–430. [Google Scholar] [CrossRef] [PubMed]
  224. Ohko, Y.; Utsumi, Y.; Niwa, C.; Tatsuma, T.; Kobayakawa, K.; Satoh, Y.; Kubota, Y.; Fujishima, A. Self-sterilizing and self-cleaning of silicone catheters coated with TiO(2) photocatalyst thin films: A preclinical work. J. Biomed. Mater. Res. 2001, 58, 97–101. [Google Scholar] [CrossRef]
  225. Nakamura, H.; Tanaka, M.; Shinohara, S.; Gotoh, M.; Karube, I. Development of a self-sterilizing lancet coated with a titanium dioxide photocatalytic nano-layer for self-monitoring of blood glucose. Biosens. Bioelectron. 2007, 22, 1920–1925. [Google Scholar] [CrossRef]
  226. Harmankaya, N.; Karlsson, J.; Palmquist, A.; Halvarsson, M.; Igawa, K.; Andersson, M.; Tengvall, P. Raloxifene and alendronate containing thin mesoporous titanium oxide films improve implant fixation to bone. Acta Biomater. 2013, 9, 7064–7073. [Google Scholar] [CrossRef]
  227. Park, S.W.; Lee, D.; Choi, Y.S.; Jeon, H.B.; Lee, C.H.; Moon, J.H. Mesoporous TiO2 implants for loading high dosage of antibacterial agent. Appl. Surf. Sci. 2014, 303, 140–146. [Google Scholar] [CrossRef]
  228. Galli, S.; Naito, Y.; Karlsson, J.; He, W.; Andersson, M.; Wennerberg, A.; Jimbo, R. Osteoconductive Potential of Mesoporous Titania Implant Surfaces Loaded with Magnesium: An Experimental Study in the Rabbit. Clin. Implant. Dent. Relat.Res. 2015, 17, 1048–1059. [Google Scholar] [CrossRef]
  229. Cecchinato, F.; Xue, Y.; Karlsson, J.; He, W.; Wennerberg, A.; Mustafa, K.; Andersson, M.; Jimbo, R. In vitro evaluation of human fetal osteoblast response to magnesium loaded mesoporous TiO2 coating. J. Biomed. Mater. Res. Part A 2014, 102, 3862–3871. [Google Scholar] [CrossRef]
  230. Kennedy, J.F.; Kay, I.M. Hydrous titanium oxides-new supports for the simple immobilisation of enzymes. J. Chem. Soc. Perkin 1976, 3, 329–335. [Google Scholar] [CrossRef]
  231. Kurokawa, Y.; Sano, T.; Ohta, H.; Nakagawa, Y. Immobilization of enzyme onto cellulose–titanium oxide composite fiber. Biotechnol. Bioeng. 1993, 42, 394–397. [Google Scholar] [CrossRef]
  232. Wang, X.; Li, G.; Liu, L.; Cheng, Y.; Zheng, W.; Wu, S.; Wu, F.; Sun, W. Application of titanium dioxide nanowires and electroreduced graphene oxide modified electrodes for the electrochemical detection of specific tlh gene sequence from vibrio parahaemolyticus. Anal. Methods 2015, 7, 2623–2629. [Google Scholar] [CrossRef]
  233. Atchudan, R.; Muthuchamy, N.; Edison, T.; Perumal, S.; Vinodh, R.; Park, K.H.; Lee, Y.R. An ultrasensitive photoelectrochemical biosensor for glucose based on bio-derived nitrogen-doped carbon sheets wrapped titanium dioxide nanoparticles. Biosens. Bioelectron. 2019, 126, 160–169. [Google Scholar] [CrossRef]
  234. Ge, L.; Liu, Q.; Hao, N.; Kun, W. Recent developments of photoelectrochemical biosensors for food analysis. J. Mater. Chem. B. 2019, 7, 7283–7300. [Google Scholar] [CrossRef]
  235. Zhang, X.; Peng, J.; Song, Y.; Chen, Y.; Lu, F.; Gao, W. Porous hollow carbon nanobubbles@ZnCdS multi-shelled dodecahedral cages with enhanced visible-light harvesting for ultrasensitive photoelectrochemical biosensors. Biosens. Bioelectron. 2019, 133, 125–132. [Google Scholar] [CrossRef]
  236. De la Calle, I.; Menta, M.; Klein, M.; Séby, F. Screening of TiO2 and Au nanoparticles in cosmetics and determination of elemental impurities by multiple techniques (DLS, SP-ICP-MS, ICP-MS and ICP-OES). Talanta 2017, 171, 291–306. [Google Scholar] [CrossRef]
  237. Altsitsiadis, E.; Undheim, T.; de Vries, E.; Hinrichs, B.; Stockflfleth, E.; Trakatelli, M.; on behalf of the EPIDERM Group. Health literacy, sunscreen and sunbed use: An uneasy association. Br. J. Dermatol. 2012, 167, 14–21. [Google Scholar] [CrossRef]
  238. Barnard, A.S. One-to-one comparison of sunscreen effificacy, aesthetics and potential nanotoxicity. Nat. Nanotechnol. 2010, 5, 271–274. [Google Scholar] [CrossRef]
  239. Veierød, M.B.; Weiderpass, E.; Thörn, M.; Hansson, J.; Lund, E.; Armstrong, B.; Adami, H.-O. A prospective study of pigmentation, sun exposure, and risk of cutaneous malignant melanoma in women. JNCI J. Natl. Cancer Inst. 2003, 95, 1530–1538. [Google Scholar] [CrossRef]
  240. Rodrigues, N.D.; Stavros, V.G. From fundamental science to product: A bottom-up approach to sunscreen development. Sci. Prog. 2018, 101, 8–31. [Google Scholar] [CrossRef]
  241. Chen, Z.; Han, S.; Zheng, P.; Zhang, J.; Zhou, S.; Jia, G. Landscape of lipidomic metabolites in gut-liver axis of Sprague-Dawley rats after oral exposure to titanium dioxide nanoparticles. Part Fibre Toxicol. 2022, 19, 53. [Google Scholar] [CrossRef]
  242. Kiss, B.; Br, T.; Czifra, G.; Tóth, B.I.; Kertész, Z.; Szikszai, Z.; Kiss, Á.Z.; Juhász, I.; Zouboulis, C.C.; Hunyadi, J. Investigation of micronized titanium dioxide penetration in human skin xenografts and its effect on cellular functions of human skin-derived cells. Exp. Dermatol. 2008, 17, 659–667. [Google Scholar] [CrossRef]
  243. Forestier, S. Rationale for sunscreen development. J. Am. Acad. Dermatol. 2008, 58, S133–S138. [Google Scholar] [CrossRef]
  244. Rodrigues, N.D.; Staniforth, M.; Stavros, V.G. Photophysics of sunscreen molecules in the gas phase: A stepwise approach towards understanding and developing next-generation sunscreens. Proc. Math. Phys. Eng. Sci. 2016, 472, 20160677. [Google Scholar] [CrossRef]
  245. Bocca, B.; Caimi, S.; Senofonte, O.; Alimonti, A.; Petrucci, F. ICP-MS based methods to characterize nanoparticles of TiO2 and ZnO in sunscreens with focus on regulatory and safety issues. Sci. Total Environ. 2018, 630, 922–930. [Google Scholar] [CrossRef]
  246. Jugan, M.L.; Barillet, S.; Simon-Deckers, A.; Sauvaigo, S.; Douki, T.; Herlin, N.; Carrière, M. Cytotoxic and genotoxic impact of TiO2 nanoparticles on A549 cells. J. Biomed. Nanotechnol. 2011, 7, 22–23. [Google Scholar] [CrossRef]
  247. Popov, A.P.; Priezzhev, A.V.; Lademann, J.; Myllylä, R. Titanium dioxide (TiO2) nanoparticles as an effffective UV-B radiation skin-protective compound in sunscreens. J. Phys. D Appl. Phys. 2005, 38, 2564–2570. [Google Scholar] [CrossRef]
  248. European Commission; Directorate General for Health & Consumers. Opinion on Titanium Dioxide (Nano Form): COLIPA n◦ S75; European Commission: Luxembourg, 2013; ISBN 978-92-79-30114-8. [Google Scholar]
  249. European Commission; Directorate-General for Health and Food Safety. Opinion on Titanium Dioxide (Nano Form) as UV-Filter in Sprays; European Commission: Luxembourg, 2018; ISBN 978-92-76-00201-7. [Google Scholar]
  250. Filipe, P.; Silva, J.N.; Silva, R.; de Castro, C.J.L.; Marques Gomes, M.; Alves, L.C.; Santus, R.; Pinheiro, T. Stratum corneum is an effffective barrier to TiO2 and ZnO nanoparticle percutaneous absorption. Ski. Pharm. Physiol. 2009, 22, 266–275. [Google Scholar] [CrossRef]
  251. Sadrieh, N.; Wokovich, A.M.; Gopee, N.V.; Zheng, J.; Haines, D.; Parmiter, D.; Siitonen, P.H.; Cozart, C.R.; Patri, A.K.; McNeil, S.E.; et al. Lack of signifificant dermal penetration of titanium dioxide from sunscreen formulations containing nano-and submicron-size TiO2 particles. Toxicol. Sci. 2010, 115, 156–166. [Google Scholar] [CrossRef]
  252. Crosera, M.; Prodi, A.; Mauro, M.; Pelin, M.; Florio, C.; Bellomo, F.; Adami, G.; Apostoli, P.; De Palma, G.; Bovenzi, M.; et al. Titanium dioxide nanoparticle penetration into the skin and effffects on HaCaT cells. Int. J. Environ. Res. Public Health 2015, 12, 9282–9297. [Google Scholar] [CrossRef]
  253. Xie, G.; Lu, W.; Lu, D. Penetration of titanium dioxide nanoparticles through slightly damaged skin in vitro and in vivo. J. Appl. Biomater. Funct. Mater. 2015, 13, 356–361. [Google Scholar] [CrossRef]
  254. Pelclova, D.; Navratil, T.; Kacerova, T.; Zamostna, B.; Fenclova, Z.; Vlckova, S.; Kacer, P. NanoTiO2 sunscreen does not prevent systemic oxidative stress caused by UV radiation and a minor amount of nanoTiO2 is absorbed in humans. Nanomaterials 2019, 9, 888. [Google Scholar] [CrossRef] [PubMed]
  255. Lingaraju, K.; Basavaraj, R.; Jayanna, K.; Bhavana, S.; Devaraja, S.; Swamy, H.K.; Nagaraju, G.; Nagabhushana, H.; Naika, H.R. Biocompatible fabrication of Titanium dioxide (TiO2) nanoparticles: Antimicrobial, anticoagulant, antiplatelet, direct hemolytic and cytotoxicity properties. Inorg. Chem. Commun. 2021, 127, 108505. [Google Scholar] [CrossRef]
  256. Babitha, S.; Korrapati, P.S. Biodegradable zein-polydopamine polymeric scaffold impregnated with Titanium dioxide (TiO2) nanoparticles for skin tissue engineering. Biomed. Mater. 2017, 12, 055008. [Google Scholar] [CrossRef] [PubMed]
  257. Stan, M.S.; Nica, I.C.; Dinischiotu, A.; Varzaru, E.; Iordache, O.G.; Dumitrescu, I.; Popa, M.; Chififiriuc, M.C.; Pircalabioru, G.G.; Lazar, V.; et al. Photocatalytic, antimicrobial and biocompatibility features of cotton knit coated with Fe-N- Doped titanium dioxide nanoparticles. Materials 2016, 9, 789. [Google Scholar] [CrossRef]
  258. Seisenbaeva, G.A.; Fromell, K.; Vinogradov, V.V.; Terekhov, A.N.; Pakhomov, A.V.; Nilsson, B.; Ekdahl, K.N.; Vinogradov, V.V.; Kessler, V.G. Dispersion of Titanium dioxide (TiO2) nanoparticles improves burn wound healing and tissue regeneration through specifific interaction with blood serum proteins. Sci. Rep. 2017, 7, 15448. [Google Scholar] [CrossRef]
  259. Poosti, M.; Ramazanzadeh, B.; Zebarjad, M.; Javadzadeh, P.; Naderinasab, M.; Shakeri, M.T. Shear bond strength and antibacterial effects of orthodontic composite containing Titanium dioxide (TiO2) nanoparticles. Eur. J. Orthod. 2013, 35, 676–679. [Google Scholar] [CrossRef] [PubMed]
  260. Tahir, A.; Moeen, F.; Mehmood, M.; Mansoor, A.; Abbas, Z.; Shahzad, M.H.; Kashif, M. Evaluation of Shear Bond Strength to Human Teeth and Microhardness of GIC Containing Titanium Nano Particles: An In-Vitro Study. Int. Med. J. 2020, 25, 1477–1488. [Google Scholar]
  261. Schmeidl, K.; Janiszewska-Olszowska, J.; Grocholewicz, K. Clinical Features and Physical Properties of Gummetal Orthodontic Wire in Comparison with Dissimilar Archwires: A Critical Review. BioMed Res. Int. 2021, 2021, 6611979. [Google Scholar] [CrossRef]
  262. Chun, M.-J.; Shim, E.; Kho, E.-H.; Park, K.-J.; Jung, J.; Kim, J.-M.; Kim, B.; Lee, K.-H.; Cho, D.-L.; Bai, D.-H.; et al. Surface Modifification of Orthodontic Wires with Photocatalytic Titanium Oxide for its Antiadherent and Antibacterial Properties. Angle Orthod. 2007, 77, 483–488. [Google Scholar] [CrossRef]
  263. Özyıldız, F.; Güden, M.; Uzel, A.; Karaboz, I.; Akil, O.; Bulut, H. Antimicrobial activity of TiO2-coated orthodontic ceramic brackets against Streptococcus mutans and Candida albicans. Biotechnol. Bioprocess Eng. 2010, 15, 680–685. [Google Scholar] [CrossRef]
  264. Amin, F.; Rahman, S.; Khurshid, Z.; Zafar, M.S.; Sefat, F.; Kumar, N. Effect of Nanostructures on the Properties of Glass Ionomer Dental Restoratives/Cements: A Comprehensive Narrative Review. Materials 2021, 14, 6260. [Google Scholar] [CrossRef] [PubMed]
  265. Azzawi, Z.; Hamad, T.I.; Kadhim, S.A.; Naji, G.A. Osseointegration evaluation of laser-deposited titanium dioxide nanoparticles on commercially pure titanium dental implants. J. Mater. Sci. Mater. Med. 2018, 29, 96. [Google Scholar] [CrossRef] [PubMed]
  266. Souza, J.; Sordi, M.B.; Kanazawa, M.; Ravindran, S.; Henriques, B.; Silva, F.S.; Aparicio, C.; Cooper, L.F. Nano-scale modification of titanium implant surfaces to enhance osseointegration. Acta Biomater. 2019, 94, 112–131. [Google Scholar] [CrossRef] [PubMed]
  267. Li, H.; Zhang, S.; Nie, B.; Du, Z.; Long, T.; Yue, B. The antimicrobial peptide KR-12 promotes the osteogenic differentiation of human bone marrow stem cells by stimulating BMP/SMAD signaling. RSC Adv. 2018, 8, 15547–15557. [Google Scholar] [CrossRef]
  268. Choi, S.H.; Jang, Y.S.; Jang, J.H.; Bae, T.S.; Lee, S.J.; Lee, M.H. Enhanced antibacterial activity of titanium by surface modification with polydopamine and silver for dental implant application. J. Appl. Biomater. Funct. Mater. 2019, 17, 2280800019847067. [Google Scholar] [CrossRef]
  269. Ji, M.K.; Oh, G.; Kim, J.W.; Park, S.; Yun, K.D.; Bae, J.C.; Lim, H.P. Effects on Antibacterial Activity and Osteoblast Viability of Non-Thermal Atmospheric Pressure Plasma and Heat Treatments of TiO2 Nanotubes. J. Nanosci. Nanotechnol. 2017, 17, 2312–2315. [Google Scholar] [CrossRef] [PubMed]
  270. Vishnu, J.; Manivasagam, V.K.; Gopal, V.; Bartomeu Garcia, C.; Hameed, P.; Manivasagam, G.; Webster, T.J. Hydrothermal treatment of etched titanium: A potential surface nano-modification technique for enhanced biocompatibility. Nanomed. Nanotechnol. Biol. Med. 2019, 20, 102016. [Google Scholar] [CrossRef] [PubMed]
  271. Chen, S.; Yang, J.; Li, K.; Lu, B.; Ren, L. Carboxylic acid-functionalized TiO2 nanoparticle-loaded PMMA/PEEK copolymer matrix as a dental resin for 3D complete denture manufacturing by stereolitographic technique. Int. J. Food Prop. 2018, 21, 2557–2565. [Google Scholar] [CrossRef]
  272. Shirai, R.; Miura, T.; Yoshida, A.; Yoshino, F.; Ito, T.; Yoshinari, M.; Yajima, Y. Antimicrobial effect of titanium dioxide after ultraviolet irradiation against periodontal pathogen. Dent. Mater. J. 2016, 35, 511–516. [Google Scholar] [CrossRef] [PubMed]
  273. Luttrell, T.; Halpegamage, S.; Tao, J.; Kramer, A.; Sutter, E.; Batzill, M. Why is anatase a better photocatalyst than rutile?—Model studies on epitaxial TiO2 films. Sci. Rep. 2014, 4, 4043. [Google Scholar] [CrossRef] [PubMed]
  274. Cheng, Y.; Yang, H.; Yang, Y.; Huang, J.; Wu, K.; Chen, Z.; Wang, X.; Lin, C.; Lai, Y. Progress in tio 2 nanotube coatings for biomedical applications: A review. J. Mater. Chem. B 2018, 6, 1862–1886. [Google Scholar] [CrossRef]
  275. Yuan, Q.; He, G.; Tan, Z.; Gong, P.; Li, X.Y.; Chen, Z.Q.; Ban, Y.; Tang, H.; Wang, F.F.; Man, Y.; et al. Biocompatibility of nano-TiO2/ha bioceramic coating for nerve regeneration around dental implants. In Key Engineering Materials; Trans Tech Publications Ltd.: Wollerau, Switzerland, 2007; Volume 330, pp. 1393–1396. [Google Scholar]
  276. Yang, W.-E.; Huang, H.-H. Multiform TiO2 nano-network enhances biological response to titanium surface for dental implant applications. Appl. Surf. Sci. 2019, 471, 1041–1052. [Google Scholar] [CrossRef]
  277. Li, H.; Lai, Y.; Huang, J.; Tang, Y.; Yang, L.; Chen, Z.; Zhang, K.; Wang, X.; Tan, L.P. Multifunctional wettability patterns prepared by laser processing on superhydrophobic TiO2 nanostructured surfaces. J. Mater. Chem. 2015, 3, 342–347. [Google Scholar] [CrossRef]
  278. Lai, Y.; Lin, L.; Pan, F.; Hujng, J.; Song, R.; Huang, Y.; Lin, C.; Fuchs, H.; Chi, L. Bioinspired patterning with extreme wettability contrast on TiO2 nanotube array surface: A versatile plat form for biomedical applications. Small 2013, 9, 2945–2953. [Google Scholar] [CrossRef]
  279. Zafar, M.S.; Alnazzawi, A.A.; Alrahabi, M.; Fareed, M.A.; Najeeb, S.; Khurshid, Z. 18-Nanotechnology and nanomaterials in dentistry. In Advanced Dental Biomaterials; Woodhead Publishing: Shaston, UK, 2019; pp. 477–505. ISBN 9780081024768. [Google Scholar] [CrossRef]
  280. Liu, W.; Su, P.; Su, C.; Wang, N.; Ma, Y.; Liu, Y.; Wang, J.; Zhang, Z.; Li, H.; Webste, T.J. Synthesis of TiO2 nanotubes with ZnO nanoparticles to achieve antibacterial properties and stem cell compatibility. Nanoscale 2014, 6, 9050–9062. [Google Scholar] [CrossRef] [PubMed]
  281. Li, Z.; Qiu, J.; Du, L.Q.; Jia, L.; Liu, H.; Ge, S. TiO2 nanorod arrays modified Ti substrates promote the adhesion, proliferation and osteogenic differentiation of human periodontal ligament stem cells. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 76, 684–691. [Google Scholar] [CrossRef] [PubMed]
  282. Jia, L.; Qiu, J.; Du, L.; Li, Z.; Liu, H.; Ge, S. TiO2 nanorod arrays as a photocatalytic coating enhanced antifungal and antibacterial efficiency of Ti substrates. Nanomedicine 2017, 12, 761–776. [Google Scholar] [CrossRef]
  283. Faria, H.A.M.; de Queiroz, A.A.A. A novel drug delivery of 5-fluorouracil device based on TiO2/zns nanotubes. Mater. Sci. Eng. C 2015, 56, 260–268. [Google Scholar] [CrossRef]
  284. Panpaliya, N.P.; Dahake, P.T.; Kale, Y.J.; Dadpe, M.V.; Kendre, S.B.; Siddiqi, A.G.; Maggavi, U.R. In vitro evaluation of antimicrobial property of silver nanoparticles and chlorhexidine against five different oral pathogenic bacteria. Saudi Dent. J. 2019, 31, 76–83. [Google Scholar] [CrossRef] [PubMed]
  285. Besinis, A.; De Peralta, T.; Handy, R.D. The antibacterial effects of silver, titanium dioxide and silica dioxide nanoparticles compared to the dental disinfectant chlorhexidine on Streptococcus mutans using a suite of bioassays. Nanotoxicology 2014, 8, 1–16. [Google Scholar] [CrossRef]
  286. Pacheco, R.R.; Vitti, R.P.; Andrada, A.C.; MenMartins, C.; Catelan, C.; Khurshid, Z. Restorative materials used in endodontics. Biomater. Endod. 2022, 2, 99–122. [Google Scholar]
  287. Garcia-Contreras, R.; Scougall-Vilchis, R.J.; Contreras-Bulnes, R.; Sakagami, H.; Morales-Luckie, R.A.; Nakajima, H. Mechanical, antibacterial and bond strength properties of nano-titanium-enriched glass ionomer cement. J. Appl. Oral Sci. Rev. FOB 2015, 23, 321–328. [Google Scholar] [CrossRef]
  288. Najeeb, S.; Khurshid, Z.; Zafar, M.S.; Khan, A.S.; Zohaib, S.; Martí, J.M.N.; Sauro, S.; Matinlinna, J.P.; Rehman, I.U. Modifications in glass ionomer cements: Nano-sized fillers and bioactive nanoceramics. Int. J. Mol. Sci. 2016, 17, 1134. [Google Scholar] [CrossRef] [PubMed]
  289. Bilek, O.; Fialova, T.; Otahal, A.; Adam, V.; Smerkova, K.; Fohlerova, Z. Antibacterial activity of AgNPs-TiO2 nanotubes: Influence of different nanoparticle stabilizers. RSC Adv. 2020, 10, 44601–44610. [Google Scholar] [CrossRef] [PubMed]
  290. Martín, J.; Vildósola, P.; Bersezio, C.; Herrera, A.; Bortolatto, J.; Saad, J.R.; Oliveira, O.B., Jr.; Fernández, E. Effectiveness of 6% hydrogen peroxide concentration for tooth bleaching—A double-blind, randomized clinical trial. J. Dent. 2015, 43, 965–972. [Google Scholar] [CrossRef] [PubMed]
  291. Sun, J.; Xu, Y.; Zhu, B.; Gao, G.; Ren, J.; Wang, H.; Lin, Y.; Cao, B. Synergistic effects of titanium dioxide and cellulose on the properties of glassionomer cement. Dent. Mater. J. 2019, 38, 41–51. [Google Scholar] [CrossRef] [PubMed]
  292. Gjorgievska, E.; Van Tendeloo, G.; Nicholson, J.W.; Coleman, N.J.; Slipper, I.J.; Booth, S. The incorporation of nanoparticles into conventional glass-ionomer dental restorative cements. Microsc. Microanal. Off. J. Microsc. Soc. Am. Microbeam Anal. Soc. Microsc. Soc. Can. 2015, 21, 392–406. [Google Scholar] [CrossRef]
  293. Yang, J.; Mei, S.; Ferreira, J.M.F. Hydrothermal synthesis of nanosized titania powders: Influence of tetraalkyl ammonium hydroxides on particle characteristics. J. Am. Ceram. Soc. 2004, 4, 1696–1702. [Google Scholar] [CrossRef]
  294. Elsaka, S.E.; Hamouda, I.M.; Swain, M.V. Titanium dioxide nanoparticles addition to a conventional glass-ionomer restorative: Inflfluence on physical and antibacterial properties. J. Dent. 2011, 39, 589–598. [Google Scholar] [CrossRef]
  295. Hibino, Y.; Kuramochi, K.; Harashima, A.; Honda, M.; Yamazaki, A.; Nagasawa, Y. Correlation between the strength of glass ionomer cements and their bond strength to bovine teeth. Dent. Mater. J. 2004, 23, 656–660. [Google Scholar] [CrossRef]
  296. Monteiro, N.R.; Basting, R.T.; Amaral, F.L.B.D.; França, F.M.G.; Turssi, C.P.; Gomes, O.P.; Lisboa Filho, P.N.; Kantovitz, K.R.; Basting, R.T. Titanium dioxide nanotubes incorporated into bleaching agents: Physicochemical characterization and enamel color change. J. Appl. Oral Sci. 2020, 28. [Google Scholar] [CrossRef]
  297. Komatsu, O.; Nishida, H.; Sekino, T.; Yamamoto, K. Application of titanium dioxide nanotubes to tooth whitening. Nano Biomed. 2014, 6, 63–72. [Google Scholar]
  298. Argueta-Figueroa, L.; Torres-Gómez, N.; Scougall-Vilchis, R.J.; García-Contreras, R. Biocompatibility and antibacterial activity of nitrogen-doped titanium dioxide nanoparticles for use in dental resin formulations. Int. J. Nanomed. 2016, 11, 6459. [Google Scholar]
  299. Kurzmann, C.; Verheyen, J.; Coto, M.; Kumar, R.V.; Divitini, G.; Shokoohi-Tabrizi, H.A.; Verheyen, P.; De Moor, R.J.G.; Moritz, A.; Agis, H. In vitro evaluation of experimental light activated gels for tooth bleaching. Photochem. Photobiol. Sci. 2019, 18, 1009–1019. [Google Scholar] [CrossRef]
  300. Endo, K.; Ueno, T.; Kondo, S.; Wakisaka, N.; Murono, S.; Ito, M.; Kataoka, K.; Kato, Y.; Yoshizaki, T. Tumor-targeted chemotherapy with the nanopolymer-based drug NC-6004 for oral squamous cell carcinoma. Cancer Sci. 2013, 104, 369–374. [Google Scholar] [CrossRef]
  301. Ladha, K.; Shah, D. An In-Vitro Evaluation of the Flexural Strength of Heat-Polymerized Poly (Methyl Methacrylate) Denture Resin Reinforced with Fibers. J. Indian Prosthodont. Soc. 2011, 11, 215. [Google Scholar] [CrossRef]
  302. Kasina, S.P.; Ajaz, T.; Attili, S.; Surapaneni, H.; Cherukuri, M.; Srinath, H.P. To evaluate and compare the porosities in the acrylic mandibular denture bases processed by two different polymerization techniques, using two different brands of commercially available denture base resins—An in vitro study. J. Int. Oral Health 2014, 6, 72–77. [Google Scholar]
  303. Sakaguchi, R.; Power, J. Craig’s Restorative Dental Materials, 13th ed.; Elsevier Health Sciences: Amsterdam, The Netherlands, 2012; ISBN 9780323081085. [Google Scholar]
  304. Ayaz, E.A.; Durkan, R.; Koroglu, A.; Bagis, B. Comparative Effect of Different Polymerization Techniques on Residual Monomer and Hardness Properties of PMMA-Based Denture Resins. J. Appl. Biomater. Funct. Mater. 2014, 12, 228–233. [Google Scholar] [CrossRef]
  305. Arenas-Arrocena, M.C.; Argueta-Figueroa, L.; García-Contreras, R.; Martínez-Arenas, O.; Camacho-Flores, B.; del RodriguezTorres, M.P.; de la Fuente-Hernández, J.; Acosta-Torres, L.S. New Trends for the Processing of Poly(Methyl Methacrylate) Biomaterial for Dental Prosthodontics. In Acrylic Polymers in Healthcare; InTech: London, UK, 2017. [Google Scholar]
  306. Moura, J.S.; Jose Da Silva, W.; Pereira, T.; Del, A.A.; Cury, B.; Cunha, R.; Rodrigues Garcia, M. Inflfluence of Acrylic Resin Polymerization Methods and Saliva on the Adherence of Four Candida Species. J. Prosthet. Dent. 2006, 96, 205–211. [Google Scholar] [CrossRef]
  307. Mehendale, A.V. Adherence of Candida Albicans on Polyamides in Comparison with Conventional Acrylic Surfaces—A Short Study. J. Indian Prosthodont. Soc. 2018, 18 (Suppl. 2), S71. [Google Scholar] [CrossRef]
  308. Aeran, H.; Kumar, V.; Uniyal, S.; Tanwer, P. Nanodentistry: Is just a Fiction or Future. J. Oral Biol. Craniofac. Res. 2015, 5, 207–211. [Google Scholar] [CrossRef]
  309. Mansoor, A.; Moeen, F.; Mehmood, M.; Hassan, S.M.U.; Ubaidullah, M.; Said, F. Age related changes in bleached enamel through scanning electron microscopy and atomic force microscopy. Pak. J. Physiol. 2019, 15, 43–48. [Google Scholar]
  310. Ali Sabri, B.; Satgunam, M.; Abreeza, N.; Abed, A. A Review on Enhancements of PMMA Denture Base Material with Different Nano-Fillers. Cogent Eng. 2021, 8, 1875968. [Google Scholar] [CrossRef]
  311. Alrahlah, A.; Fouad, H. Titanium Oxide (TiO2)/Polymethylmethacrylate (PMMA) Denture Base Nanocomposites: Mechanical, Viscoelastic and Antibacterial Behavior. Materials 2018, 11, 1096. [Google Scholar] [CrossRef]
  312. Totu, E.E.; Nechifor, A.C. Poly (methyl methacrylate) with Titanium dioxide (TiO2) nanoparticles inclusion for stereolitographic complete denture manufacturing—The fututre in dental care for elderly edentulous patients? J. Dent. 2017, 59, 68–77. [Google Scholar] [CrossRef]
  313. Cierech, M.; Szerszeń, M.; Wojnarowicz, J.; Łojkowski, W.; Kostrzewa-Janicka, J.; Mierzwińska-Nastalska, E. Preparation and Char acterisation of Poly(methyl metacrylate)-Titanium Dioxide Nanocomposites for Denture Bases. Polymers 2020, 12, 2655. [Google Scholar] [CrossRef]
  314. Shirkavand, S.; Moslehifard, E. Effect of Titanium dioxide (TiO2) nanoparticles on Tensile Strength of Dental Acrylic Resins. Dent. Clin. Dent. Prospect. 2014, 8, 197–203. [Google Scholar]
  315. Yoshijima, Y.; Murakami, K.; Kayama, S.; Liu, D.; Hirota, K.; Ichikawa, T.; Miyake, Y. Effect of Substrate Surface Hydrophobicity on the Adherence of Yeast and Hyphal Candida. Mycoses 2010, 53, 221–226. [Google Scholar] [CrossRef]
  316. Cannon, R.D.; Chaffifin, W.L. Oral Colonization by Candida Albicans. Oral Biol. Med. 1999, 10, 359–383. [Google Scholar] [CrossRef]
  317. De Matteis, V.; Cascione, M.; Toma, C.C.; Leporatti, S. Morphomechanical and Organelle Perturbation Induced by Silver Nanoparticle Exposure. J. Nanopart. Res. 2018, 20, 273. [Google Scholar] [CrossRef]
  318. Zaidi, S.J.A.; Shaikh, A.H.; Kazmi, S.; Ahmad, M.; Afaq, A.; Mansoor, A. Emergency Management Readiness of Pakistani Dentists at Public Sector Teaching Hospitals. Pak. J. Med. Dent. 2022, 11, 78–84. [Google Scholar] [CrossRef]
  319. Attarilar, S.; Yang, J.; Ebrahimi, M.; Wang, Q.; Liu, J.; Tang, Y.; Yang, J. The Toxicity Phenomenon and the Related Occurrence in Metal and Metal Oxide Nanoparticles: A Brief Review from the Biomedical Perspective. Front. Bioeng. Biotechnol. 2020, 8, 822. [Google Scholar] [CrossRef]
  320. Naz, S.; Gul, A.; Zia, M. Toxicity of Copper Oxide Nanoparticles: A Review Study. IET Nanobiotechnol. 2020, 14, 1–13. [Google Scholar] [CrossRef]
  321. Xu, H.H.; Moreau, J.L.; Sun, L.; Chow, L.C. Nanocomposite containing amorphous calcium phosphate nanoparticles for caries inhibition. Dent. Mater. Off. Publ. Acad. Dent. Mater. 2011, 27, 762–769. [Google Scholar] [CrossRef]
  322. Weir, A.; Westerhoff, P.; Fabricius, L.; Hristovski, K.; von Goetz, N. Titanium dioxide nanoparticles in food and personal care products. Environ. Sci. Technol. 2012, 46, 2242–2250. [Google Scholar] [CrossRef]
  323. Afonso, R.L.; Pessan, J.P.; Igreja, B.B.; Cantagallo, C.F.; Danelon, M.; Delbem, A.C. In situ protocol for the determination of dose-response effect of low-fluoride dentifrices on enamel remineralization. J. Appl. Oral Sci. Rev. FOB 2013, 21, 525–532. [Google Scholar] [CrossRef]
  324. Koray, M.; Oner-Iyidogan, Y.; Soyman, M.; Gürdöl, F. The effects of fluorides and/or trace elements on the solubilities of enamel and cementum. J. Trace Elem. Med. Biol. Organ Soc. Miner. Trace Elem. (GMS) 1996, 10, 255–259. [Google Scholar] [CrossRef]
  325. Huang, J.; Best, S.M.; Bonfield, W.; Buckland, T. Development and characterization of titanium-containing hydroxyapatite for medical applications. Acta Biomater. 2011, 6, 241–249. [Google Scholar] [CrossRef]
  326. Eimar, H.; Marelli, B.; Nazhat, S.N.; Abi Nader, S.; Amin, W.M.; Torres, J.; de Albuquerque, R.F., Jr.; Tamimi, F. The role of enamel crystallography on tooth shade. J. Dent. 2011, 39 (Suppl. 3), e3–e10. [Google Scholar] [CrossRef]
  327. Ghadimi, E.; Eimar, H.; Marelli, B.; Nazhat, S.N.; Asgharian, M.; Vali, H.; Tamimi, F. Trace elements can influence the physical properties of tooth enamel. SpringerPlus 2013, 2, 499. [Google Scholar] [CrossRef]
  328. FDI World Dental Federation. FDI policy statement on Minimal Intervention Dentistry (MID) for managing dental caries: Adapted by the General Assembly: September 2016, Poznan, Poland. Int. Dent. J. 2017, 67, 6–7. [Google Scholar] [CrossRef]
  329. Sofan, E.; Sofan, A.; Palaia, G.; Tenore, G.; Romeo, U.; Migliau, G. Classifification review of dental adhesive systems: From the IV generation to the universal type. Ann. Stomatol. 2017, 8, 1–17. [Google Scholar]
  330. Ferracane, J.L. Models of Caries Formation around Dental Composite Restorations. J. Dent. Res. 2017, 96, 364–371. [Google Scholar] [CrossRef]
  331. Amin, F.; Fareed, M.A.; Zafar, M.S.; Khurshid, Z.; Palma, P.J.; Kumar, N. Degradation and Stabilization of Resin-Dentine Interfaces in Polymeric Dental Adhesives: An Updated Review. Coatings 2022, 12, 1094. [Google Scholar] [CrossRef]
  332. Sun, J.; Petersen, E.J.; Watson, S.S.; Sims, C.M.; Kassman, A.; Frukhtbeyn, S.; Skrtic, D.; Ok, M.T.; Jacobs, D.S.; Reipa, V.; et al. Biophysical characterization of functionalized titania nanoparticles and their application in dental adhesives. Acta Biomater. 2017, 53, 585–597. [Google Scholar] [CrossRef]
  333. Esteban Florez, F.L.; Kraemer, H.; Hiers, R.D.; Sacramento, C.M.; Rondinone, A.J.; Silverio, K.G.; Khajotia, S.S. Sorption, solubility and cytotoxicity of novel antibacterial nanofifilled dental adhesive resins. Sci. Rep. 2020, 10, 13503. [Google Scholar] [CrossRef]
  334. Ramos-Tonello, C.M.; Lisboa-Filho, P.N.; Arruda, L.B.; Tokuhara, C.K.; Oliveira, R.C.; Furuse, A.Y.; Rubo, J.H.; Borges, A.F.S. Titanium dioxide nanotubes addition to self-adhesive resin cement: Effect on physical and biological properties. Dent. Mater. 2017, 33, 866–875. [Google Scholar] [CrossRef]
  335. Sturmer, M.; Garcia, I.M.; Souza, V.S.; Visioli, F.; Scholten, J.D.; Samuel, S.M.W.; Leitune, V.C.B.; Collares, F.M. Titanium dioxide nanotubes with triazine-methacrylate monomer to improve physicochemical and biological properties of adhesives. Dent. Mater. 2021, 37, 223–235. [Google Scholar]
  336. Ibrahim, M.A.; Meera, P.B.; Neo, J.; Fawzy, A.S. Characterization of Chitosan/TiO2 Nano-Powder Modified Glass-Ionomer Cement for Restorative Dental Applications. J. Esthet. Restor. Dent. 2017, 29, 146–156. [Google Scholar] [CrossRef]
  337. Magni, E.; Ferrari, M.; Hickel, R.; Ilie, N. Evaluation of the mechanical properties of dental adhesives and glass-ionomer cements. Clin Oral Investig. 2010, 14, 79–87. [Google Scholar] [CrossRef]
  338. Srinivasan, S.; Jayasree, R.; Chennazhi, K.P.; Nair, S.V.; Jayakumar, R. Biocompatible alginate/nano bioactive glass ceramic composite scaffolds for periodontal tissue regeneration. Carbohydr. Polym. 2012, 87, 274–283. [Google Scholar] [CrossRef]
  339. Huang, L.; Dai, T.; Xuan, Y.; Tegos, G.P.; Hamblin, M.R. Synergistic combination of chitosan acetate with nanoparticle silver as a topical antimicrobial: Efficacy against bacterial burn infections. Antimicrob. Agents Chemother. 2011, 55, 3432–3438. [Google Scholar] [CrossRef]
  340. Wang, J.; Fan, Y. Lung Injury Induced by TiO2 Nanoparticles Depends on Their Structural Features: Size, Shape, Crystal Phases, and Surface Coating. Int. J. Mol. Sci. 2014, 15, 22258–22278. [Google Scholar] [CrossRef]
  341. Chang, H.P.; Tseng, Y.C. Miniscrew implant applications in contemporary orthodontics. Kaohsiung J. Med. Sci. 2014, 30, 111–115. [Google Scholar] [CrossRef]
  342. Takahashi, N.; Nyvad, B. The role of bacteria in the caries process: Ecological perspectives. J. Dent. Res. 2011, 90, 294–303. [Google Scholar] [CrossRef] [PubMed]
  343. Usha, C.; Ru, S. Dental caries—A complete changeover (Part I). J. Conserv. Dent. JCD 2009, 12, 46–54. [Google Scholar] [CrossRef]
  344. Agnihotry, A.; Fedorowicz, Z.; Nasser, M.; Gill, K.S. Resorbable versus titanium plates for orthognathic surgery. Cochrane Database Syst. Rev. 2017, 10, CD006204. [Google Scholar] [CrossRef]
  345. Costa, F.W.; Bezerra, M.F.; Ribeiro, T.R.; Pouchain, E.C.; Sabóia, V.; Soares, E.C. Biomechanical analysis of titanium plate systems in mandibular condyle fractures: A systematized literature review. Acta Cir. Bras. 2012, 27, 424–429. [Google Scholar] [CrossRef]
  346. Hill, C.S.; Luoma, A.M.; Wilson, S.R.; Kitchen, N. Titanium cranioplasty and the prediction of complications. Br. J. Neurosurg. 2012, 26, 832–837. [Google Scholar] [CrossRef]
  347. Sefat, F.; Raja, T.I.; Zafar, M.S.; Khurshid, Z.; Najeeb, S.; Zohaib, S.; Ahmadi, E.D.; Rahmati, M.; Mozafari, M. Nanoengineered Biomaterials for Cartilage Repair; Wood Head Publishers: Shastion, UK, 2019; Chapter 3; ISBN 9780128133552. [Google Scholar] [CrossRef]
  348. Himel, V.T.; Levitan, M.E. Use of nickel titanium instruments for cleaning and shaping root canal systems. Tex. Dent. J. 2003, 120, 262–268. [Google Scholar] [PubMed]
  349. Tomson, P.L.; Simon, S.R. Contemporary Cleaning and Shaping of the Root Canal System. Prim. Dent. J. 2016, 5, 46–53. [Google Scholar] [CrossRef] [PubMed]
  350. Mansoor, A.; Moeen, F.; Mehmood, M.; Hassan, S.M.; Ullah, M.U.; Khan, M.T. Correlation between micro-hardness and mineral content in the healthy tooth enamel of humans belonging to different age groups. Pak. Armed Forces Med. J. 2019, 69, 1204–1209. Available online: https://www.pafmj.org/index.php/PAFMJ/article/view/3636 (accessed on 28 September 2022).
  351. Jafari, S.; Mahyad, B.; Hashemzadeh, H.; Janfaza, S.; Gholikhani, T.; Tayebi, L. Biomedical Applications of TiO2 Nanostructures: Recent Advances. Int. J. Nanomed. 2020, 15, 3447–3470. [Google Scholar]
  352. Liu, F.; Hong, T.; Xie, J.; Zhan, X.; Wang, Y. Application of reactive oxygen species-based nanomaterials in dentistry: A review. Crystals 2021, 11, 266. [Google Scholar] [CrossRef]
  353. Zafar, M.S.; Ullah, R.; Qamar, Z.; Fareed, M.A.; Amin, F.; Khurshid, Z.; Sefat, F. 2-Properties of Dental Biomaterials; Wood Head Publishers: Shastion, UK, 2019; pp. 7–35. ISBN 9780081024768. [Google Scholar] [CrossRef]
  354. Warheit, D.B.; Donner, E.M. Risk assessment strategies for nanoscale and fine-sized titanium dioxide particles: Recognizing hazard and exposure issues. Food Chem. Toxicol. 2015, 85, 138–147. [Google Scholar] [CrossRef] [PubMed]
  355. Baan, R.A. Carcinogenic hazards from inhaled carbon black, titanium dioxide, and talc not containing asbestos or asbestiform fibers: Recent evaluations by an IARC monographs working group. Inhal. Toxicol. 2007, 19 (Suppl. 1), 213–222. [Google Scholar] [CrossRef] [PubMed]
  356. Dabrowska, S.; Chudoba, T.; Wojnarowicz, J.; Łojkowski, W. Current trends in the development of microwave reactors for the synthesis of nanomaterials in laboratories and industries: A Review. Crystals 2018, 8, 379. [Google Scholar]
  357. Thakur, B.; Kumar, A.; Kumar, D. Green synthesis of titanium dioxide nanoparticles using Azadirachta indica leaf extract and evaluation of their antibacterial activity. S. Afr. J. Bot. 2019, 124, 223–227. [Google Scholar]
  358. Namita Ashish, S. Nanotechnology definitions, research, industry and property rights. In Nanoscience in Food and Agriculture 1, Sustainable Agriculture Reviews 20; Ranjan, S., Dasgupta, N., Lichtfouse, E., Eds.; Springer International Publishing: Basel, Switzerland, 2016; pp. 43–64. [Google Scholar]
  359. US Food and Drug Administration. FDA Advances New Proposed Regulation to Make Sure That Sunscreens Are Safe and Effffective. Available online: http://www.fda.gov/news-events/press-announcements/fda-advances-new-proposed-regulation-make-sure-sunscreens-are-safe-and-effffective (accessed on 6 November 2019).
  360. Li, W.; Zeng, T. Preparation of TiO2 anatase nanocrystals by TiCl4 hydrolysis with additive H2SO. PLoS ONE 2011, 6, e21082. [Google Scholar] [CrossRef]
  361. Wang, R.R.; Fenton, A. Titanium for prosthodontic applications: A review of the literature. Quintessence 1996, 27, 401–408. [Google Scholar]
  362. Khan, M.T.; Mansoor, A.; Hussain, A.; Hussain, M.W.; Ahmad, F.; Mansoor, E. Effectiveness of bcg vaccination against sars-cov-2. Pak. J. Physiol. 2021, 17, 40–44. [Google Scholar]
  363. Mueller, N.C.; Nowack, B. Exposure modeling of engineered nanoparticles in the environment. Environ. Sci Technol. 2008, 42, 4447–4453. [Google Scholar] [CrossRef]
  364. Gupta, K.; Singh, R.P.; Pandey, A.; Pandey, A. Photocatalytic antibacterial performance of TiO2 and Agdoped TiO2 against S. aureus. P. aeruginosa. Beilsten J. Nanotechnol. 2013, 4, 345–351. [Google Scholar] [CrossRef] [PubMed]
  365. Bonetta, S.; Bonetta, S.; Motta, F.; Strini, A.; Carraro, E. Photocatalytic bacterial inactivation by TiO2-coated surfaces. AMB Express 2013, 3, 59. [Google Scholar] [CrossRef] [PubMed]
  366. Othman, S.H.; Abd Salam, N.R.; Zainal, N.; Kadir Basha, R.; Talib, R.A. Antimicrobial activity of TiO2 nanoparticle-coated film for potential food packaging applications. Int. J. Photoenergy 2014, 2014, 945930. [Google Scholar] [CrossRef]
  367. Du, S.; Wu, J.; Alshareedah, O.; Shi, X. Nanotechnology in cement-based materials: A review of durability, modeling, and advanced characterization. Nanomaterials 2019, 9, 1213. [Google Scholar] [CrossRef]
  368. Sengul, A.B.; Asmatulu, E. Toxicity of metal and metal oxide nanoparticles: A review. Environ. Chem. Lett. 2020, 18, 1659–1683. [Google Scholar] [CrossRef]
  369. Ikarashi, Y.; Momma, J.; Tsuchiya, T.; Nakamura, A. Evaluation of skin sensitization potential of nickel, chromium, titanium and zirconium salts using guinea-pigs and mice. Biomaterials 1996, 17, 2103–2108. [Google Scholar] [CrossRef]
  370. Jones, K.; Morton, J.; Smith, I.; Jurkschat, K.; Harding, A.H.; Evans, G. Human in vivo and in vitro studies on gastrointestinal absorption of titanium dioxide nanoparticles. Toxicol. Lett. 2015, 233, 95–101. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Classification of NPs: (1) organic-origin NPs, containing polymeric NPs and lipid-based NPs; (2) inorganic-origin NPs, containing metal NPs, metal oxide NPs, ceramic NPs, and semiconductor NPs; (3) carbon-origin NPs, containing fullerene NPs, graphene NPs, and carbon NPs (idea adapted from ref [45].
Figure 1. Classification of NPs: (1) organic-origin NPs, containing polymeric NPs and lipid-based NPs; (2) inorganic-origin NPs, containing metal NPs, metal oxide NPs, ceramic NPs, and semiconductor NPs; (3) carbon-origin NPs, containing fullerene NPs, graphene NPs, and carbon NPs (idea adapted from ref [45].
Nanomaterials 12 03670 g001
Figure 2. Size and shape of different types of NPs: (a) polymers, micelles, lipids, and dendrimers belonging to organic-origin NPs; (b) metals, metals oxide, ceramics, semiconductors belonging to inorganic-origin NPs; (c) fullerenes, copolymer, carbon nanotubes, carbon nanofibers, and carbon black belonging to carbon-origin NPs (idea adapted from the ref [46]).
Figure 2. Size and shape of different types of NPs: (a) polymers, micelles, lipids, and dendrimers belonging to organic-origin NPs; (b) metals, metals oxide, ceramics, semiconductors belonging to inorganic-origin NPs; (c) fullerenes, copolymer, carbon nanotubes, carbon nanofibers, and carbon black belonging to carbon-origin NPs (idea adapted from the ref [46]).
Nanomaterials 12 03670 g002
Figure 3. Schematic representation of protocols employed for the synthesis of TiO2 NPs: (a) chemical and physical methods involved in the conventional methods category and biological methods involved in the latest methods category, showing (b) chemical, physical, and biological synthesis involving both the bottom-up and top-down techniques (idea adapted from the ref [57,58].
Figure 3. Schematic representation of protocols employed for the synthesis of TiO2 NPs: (a) chemical and physical methods involved in the conventional methods category and biological methods involved in the latest methods category, showing (b) chemical, physical, and biological synthesis involving both the bottom-up and top-down techniques (idea adapted from the ref [57,58].
Nanomaterials 12 03670 g003
Figure 4. Role of synthesis protocols on advantages and disadvantages of conventional and biological TiO2 NPs: (a) micro-organisms, plant tissues, and natural derivatives utilized by the latest biological synthesis; (b) chemical salts and reducing agents utilized by conventional chemical synthesis; and (c) bulk form and reducing agents utilized by conventional physical synthesis (idea adapted from the ref) [86,87].
Figure 4. Role of synthesis protocols on advantages and disadvantages of conventional and biological TiO2 NPs: (a) micro-organisms, plant tissues, and natural derivatives utilized by the latest biological synthesis; (b) chemical salts and reducing agents utilized by conventional chemical synthesis; and (c) bulk form and reducing agents utilized by conventional physical synthesis (idea adapted from the ref) [86,87].
Nanomaterials 12 03670 g004
Figure 5. Different characterization techniques utilized by TiO2 NPs for their particular features: (a) X-ray diffraction pattern spectroscopy (XRD) for investigating the size, shape, and phase form; (b) energy-dispersive X-ray spectroscopy (EDS) for investigating the elemental composition; (c) scanning electron microscope (SEM) for investigating the size, shape, and surface morphology; (d) transmission electron microscope (TEM) for investigating the size and shape; (e) atomic force microscope (AFM) for investigating the surface roughness and topography; (f) Fourier transform infrared spectroscope (FTIR) for investigating the functional groups; and (g) diffuse reflectance spectroscope (DRS) for investigating the size, shape, and crystal type [79].
Figure 5. Different characterization techniques utilized by TiO2 NPs for their particular features: (a) X-ray diffraction pattern spectroscopy (XRD) for investigating the size, shape, and phase form; (b) energy-dispersive X-ray spectroscopy (EDS) for investigating the elemental composition; (c) scanning electron microscope (SEM) for investigating the size, shape, and surface morphology; (d) transmission electron microscope (TEM) for investigating the size and shape; (e) atomic force microscope (AFM) for investigating the surface roughness and topography; (f) Fourier transform infrared spectroscope (FTIR) for investigating the functional groups; and (g) diffuse reflectance spectroscope (DRS) for investigating the size, shape, and crystal type [79].
Nanomaterials 12 03670 g005
Figure 6. SEM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 6. SEM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g006
Figure 7. TEM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 7. TEM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g007
Figure 8. XRD images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 8. XRD images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g008
Figure 9. DRS images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 9. DRS images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g009
Figure 10. FTIR images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 10. FTIR images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g010
Figure 11. EDX images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 11. EDX images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g011
Figure 12. AFM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Figure 12. AFM images of different sizes of TiO2 NPs attained from (a) Bacillus subtilis [79], (b) Cassia fistula, and (c) hydrothermal heating [80].
Nanomaterials 12 03670 g012
Figure 13. Characteristic features of TiO2 NPs dependent on characterization, showing the surface functionality, size distribution, particle shape, porosity, surface area, particle size, composition, surface, crystallinity, and agglomeration state [119].
Figure 13. Characteristic features of TiO2 NPs dependent on characterization, showing the surface functionality, size distribution, particle shape, porosity, surface area, particle size, composition, surface, crystallinity, and agglomeration state [119].
Nanomaterials 12 03670 g013
Figure 14. Medical and dental applications of TiO2 NPs: (a) drug delivery in cancer therapy, photocatalytic activity, antibacterial activity, and bio sensors in medical field; (b) orthodontics, implants, oral disinfectants, restorations, tooth whitening, acrylic dentures, oral dentifrices, dental adhesives, veneers, crowns, bridges, maxillofacial surgery, and endodontics in dental field (idea adapted from the ref [351,352]).
Figure 14. Medical and dental applications of TiO2 NPs: (a) drug delivery in cancer therapy, photocatalytic activity, antibacterial activity, and bio sensors in medical field; (b) orthodontics, implants, oral disinfectants, restorations, tooth whitening, acrylic dentures, oral dentifrices, dental adhesives, veneers, crowns, bridges, maxillofacial surgery, and endodontics in dental field (idea adapted from the ref [351,352]).
Nanomaterials 12 03670 g014
Table 1. Comparative differences between conventional and biological TiO2 NPs [40,51,85,90,91].
Table 1. Comparative differences between conventional and biological TiO2 NPs [40,51,85,90,91].
PropertiesConventional TiO2Biological TiO2References
NatureExpensive, toxicCost effective, nontoxic[40]
Reducing agentDimethylformamide, ethylene glycol, hydrazine hydrate, sodium borohydride, polyol, sodium citrate, and N,N-dimethylformamideBiomolecules include phenolics, polysaccharides, flavones, terpenoids, alkaloids, proteins, amino acids, enzymes, predominantly, and nitrate reductase[51]
MethodStabilizer (surfactant) is added to the first solution to prevent the agglomeration of NPsThere is no need to add a stabilizing agent[85]
Environmental impactEnvironmental pollution is a disadvantage of the chemical method and the chemical reduction methods are energy-intensiveSynthesis carried out in environmental conditions, and they are safe enough and consume no energy[90]
Antibacterial
activity
The chemically synthesized NPs show comparatively lower antimicrobial activity against pathogenic bacteriaThe NPs synthesized from biological means are showing better antimicrobial activity against the pathogenic bacteria[91]
StabilityNPs are not stableNPs are highly stable[85]
Byproducts
formation
Harmful byproducts formedNo harmful byproducts formed[91]
Biocompatible
coating formation
No biocompatible coating is formedThe additional safe and effective surface area for interaction in the biological environment is produced due to the biocompatible coating on the nanoparticle’s surface[51]
ProcessNon-economicalEconomical[40]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mansoor, A.; Khurshid, Z.; Khan, M.T.; Mansoor, E.; Butt, F.A.; Jamal, A.; Palma, P.J. Medical and Dental Applications of Titania Nanoparticles: An Overview. Nanomaterials 2022, 12, 3670. https://doi.org/10.3390/nano12203670

AMA Style

Mansoor A, Khurshid Z, Khan MT, Mansoor E, Butt FA, Jamal A, Palma PJ. Medical and Dental Applications of Titania Nanoparticles: An Overview. Nanomaterials. 2022; 12(20):3670. https://doi.org/10.3390/nano12203670

Chicago/Turabian Style

Mansoor, Afsheen, Zohaib Khurshid, Muhammad Talal Khan, Emaan Mansoor, Faaz Ahmad Butt, Asif Jamal, and Paulo J. Palma. 2022. "Medical and Dental Applications of Titania Nanoparticles: An Overview" Nanomaterials 12, no. 20: 3670. https://doi.org/10.3390/nano12203670

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