Next Article in Journal
Self-Sensing Cementitious Composites: Review and Perspective
Previous Article in Journal
Stretchable, Bio-Compatible, Antioxidant and Self-Powering Adhesives from Soluble Silk Fibroin and Vegetal Polyphenols Exfoliated Graphite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Influence of Titanium Dioxide Nanoparticles on Human Health and the Environment

by
Mohammad Mamunur Rashid
,
Petra Forte Tavčer
* and
Brigita Tomšič
*
Department of Textiles, Graphic Arts and Design, Faculty of Natural Sciences and Engineering, University of Ljubljana, Aškerčeva 12, 1000 Ljubljana, Slovenia
*
Authors to whom correspondence should be addressed.
Submission received: 20 August 2021 / Revised: 3 September 2021 / Accepted: 8 September 2021 / Published: 10 September 2021

Abstract

:
Nanotechnology has enabled tremendous breakthroughs in the development of materials and, nowadays, is well established in various economic fields. Among the various nanomaterials, TiO2 nanoparticles (NPs) occupy a special position, as they are distinguished by their high availability, high photocatalytic activity, and favorable price, which make them useful in the production of paints, plastics, paper, cosmetics, food, furniture, etc. In textiles, TiO2 NPs are widely used in chemical finishing processes to impart various protective functional properties to the fibers for the production of high-tech textile products with high added value. Such applications contribute to the overall consumption of TiO2 NPs, which gives rise to reasonable considerations about the impact of TiO2 NPs on human health and the environment, and debates regarding whether the extent of the benefits gained from the use of TiO2 NPs justifies the potential risks. In this study, different TiO2 NPs exposure modes are discussed, and their toxicity mechanisms—evaluated in various in vitro and in vivo studies—are briefly described, considering the molecular interactions with human health and the environment. In addition, in the conclusion of this study, the toxicity and biocompatibility of TiO2 NPs are discussed, along with relevant risk management strategies.

1. Introduction

TiO2 is one of the most abundant and widely used metal oxide nanomaterial in the world [1,2]. As an n-type semiconductor, with a band gap energy of 3.2–3.35 eV, depending on its crystal phase, TiO2 acts as an effective photocatalyst during the photocatalytic process for surface functionalization [3,4,5,6]. It has three crystalline structures: anatase, rutile, and brookite. Anatase is the most common type and rutile is the most stable form, while brookite is the rarest [7]. There is also amorphous TiO2, which is a non-crystalline form. TiO2 exists in various nanostructures, such as nanoparticles, nanotubes, nanorods and nanowires, nanofilms, nanosheets, and nanocoatings, with remarkable photocatalytic activity, which attract scientists to develop potential technological applications in multidisciplinary fields for industrial production [8].
TiO2 is widely used in food, paints, plastics, printing inks, papers, and biomedical and cosmetic products. Due to its unique properties, TiO2 is also advantageously used in textiles. For example, when applied to textile fibers, it confers various functionalities such as UV protection, photocatalytic self-cleaning and antimicrobial activity, electrical conductivity, and antistatic properties, as well as increased thermal stability. Moreover, it can also be used for solar energy conversion in the production of electronic textiles or wearable electronics. However, when tailoring its desired functionality, TiO2 is usually deposited on the surface of textile fibers, but due to its fairly poor adhesion to the fibers, a certain amount of TiO2 washes away, peels off, or wears off during the product’s lifetime. Therefore, such continuous leaching of TiO2 into the environment results in a potential threat to human health and the ecosystem.
Extensive use of TiO2 NPs reasonably raises significant concerns due to their potential nanotoxicity, induced by oxidative stress, which is influenced by ROS (Reactive Oxygen Species) formation on the surface of TiO2 NPs in the presence of UV light [9,10,11,12]. After the penetration of TiO2 NPs into the human body, inflammation, cytotoxicity, genotoxicity, immunotoxicity, and neurotoxicity may occur. Such nanotoxicity can damage the liver, kidneys, lungs, and skin. Special attention must be paid to production workers, in particular, who are in daily contact with high concentrations of TiO2 NPs. For instance, it has been proven that TiO2 NPs smaller than 20 nm cause a higher inflammatory response in human cells [13]. Furthermore, TiO2 NPs also pose certain damage to end users and researchers and, at lower concentrations, TiO2 NPs act as immunomodulatory agents, inducing inflammatory responses through specific interactions with immune system cells [14,15].
Undoubtedly, TiO2 is released into most ecosystems, including in the agricultural field, where its potential effects on soil properties, soil microflora, and plants are yet to be investigated. However, TiO2′s potential nanotoxicity to the aquatic ecosystem is currently under study, and has thus far revealed that TiO2 NPs negatively affect the suppression of the immune system of fish and invertebrates. Despite this, more research is needed to address the bioaccumulation profile of TiO2 and its associated biomagnification in the food web. To date, controversial results have been found when studying the toxicity of TiO2 NPs, which may be due to different particle sizes, doses, culture media, or test methods used [7,16]. Recently, green nanotechnology has been adopted by researchers to ensure the biocompatible and environmentally friendly use of TiO2 NPs by overcoming their drawbacks.
Due to the widespread use of TiO2, exposure to NPs by ingestion, inhalation, or sorption has become virtually inevitable. In the last five years, reviews have been published on the effects of TiO2 NPs in living organisms, focusing either on the toxicology of TiO2 NPs [1,2,11,17,18,19,20,21,22,23], safety concerns in various applications [24,25,26,27,28] and related impact on human health [28,29,30,31,32,33], or their effects on water/soil/environmental quality [34,35,36,37,38]. Such review studies are necessary to conclusively determine the environmental and human effects of TiO2 NPs. The purpose of this paper is to review recent advances in the potential health and environmental effects of TiO2 NPs in order to contribute to the establishment of a scientific basis for the safe application of TiO2 NP and to promote the sustainable development of nanotechnology. Accordingly, the main exposure modes of TiO2 NPs, their potential toxicity mechanisms on human cells through various signaling pathways along with the negotiable toxicity, the health effects of exposure to TiO2 NPs, the biocompatibility, and the environmental effects of TiO2 NPs are discussed in detail.

2. Modes of Exposure

2.1. Inhalation

Inhalation is the major route of nanoparticle penetration into the body [33,39]. When TiO2 NPs are inhaled, they are transported to various lung tissues, capillaries, airways, and alveoli and translocated to the heart, liver, nervous system, etc. [13]. Depending on the duration and concentration of inhalation, TiO2 NPs undergo short- and/or long-term clearance from bodily compartments. By inhalation, finer particles, such as anatase TiO2 NPs, have a more toxic effect than the comparatively coarser rutile form [40]. For a short-term, single-cycle inhalation threshold, a value of 3.5 mg/m3 is assumed for spray applications. For repeated inhalation, a threshold of 17 mg/m3 is assumed for an 8-h workday. In reality, however, the actual threshold is likely to be even higher. Nevertheless, it was shown that an inhalation concentration of up to 35 mg/m3 does not lead to chronic pulmonary overload [40].

2.2. Oral Route

TiO2 NPs can enter the blood via the oral route as they are used in foods, personal care products, sunscreens, and toothpaste. TiO2 NPs remain in the major organs for a long time and are eventually excreted in the stool. Absorbed TiO2 NPs in the liver, spleen, kidney, and lung tissues are responsible for possible nephrotoxicity and liver damage [41]. Even after the prolonged release of NPs, they can affect living organisms in the environment via biological pathways [42].

2.3. Dermal Route

Human skin has unique barrier properties that act against the penetration of TiO2 NPs into the skin. Several studies have found that TiO2 NPs cannot penetrate through the human dermis, even when their particle size is smaller than 100 nm [43]. Moreover, other studies have revealed that the TiO2 NPs that do manage to penetrate the skin do not exhibit toxicity under certain conditions [44].

2.4. Injection

TiO2 NPs can be developed as a highly efficient photothermal, medicinal, and synergistic agent to repair/remove diseased tissues such as cancer both in vitro and in vivo with negligible toxic properties, tissue damage, and kidney/liver dysfunction [45,46]. After injection of TiO2 NPs, endothelial cells can be disrupted, and they can proliferate and migrate despite the induction of toxicity and immunogenicity. It has been shown that TiO2 NPs injected into the blood during photothermal cancer therapy enrich the cancer tissue by diffusing, engulfing the cancer cells, and eventually killing them completely [47,48,49].

3. Toxicity of TiO2 NPs

There is a lack of information on the potential toxicity of TiO2 NPs. Indeed, despite numerous studies in this field, it is difficult to find separate studies using the same TiO2 NPs with the same experimental protocol to compare the results obtained. Nevertheless, the existing data support the potential toxicity of TiO2 NPs in humans, model vertebrates and invertebrates, plants, algae, and microorganisms. The mechanism of toxicity of TiO2 NPs to organisms can be outlined as follows: (i) production of reactive oxygen species (ROS) and formation of electron-hole pairs in the presence of light; (ii) binding of TiO2 NPs to the cell membrane via electrostatic interactions, resulting in cell wall damage and peroxidation of lipids in the cell membrane; and (iii) binding of TiO2 NPs to intracellular organelles and biological macromolecules [50].

3.1. Cytotoxicity

It is well known that ROS (i.e., superoxide (O2•−), hydrogen peroxide (H2O2), and the hydroxyl radical (OH)) are produced by aerobic organisms within the cell and are normally in equilibrium with antioxidant molecules (Figure 1a) [51]. The imbalance between ROS and antioxidants (AOX) caused by the excessive production of ROS or the depletion of antioxidant molecules leads to the occurrence of oxidative stress (Figure 1b,c) [51].
As for the negative biological effects, oxidative stress is the most important process involved in the formation of TiO2 NPs-induced ROS [52]. High TiO2 concentrations cause greater oxidative stress, which correlates with the increase in lipid peroxidation at the cell membrane after TiO2 NPs adsorption on the cell membrane. TiO2 NPs-induced ROS and lipid peroxidation damage the integrity of the cell wall and membrane, resulting in increased permeability [53], which allows TiO2 NPs to enter the cell. In addition to free diffusion, TiO2 NPs can also enter the cell through the process of endocytosis. While particles larger than 500 nm can be removed by phagocytes [50], smaller particles can be engulfed by a cell membrane vesicle and taken further into the cell. Accordingly, TiO2 NPs with a size of 25 nm and less have been taken up into human keratinocytes [54], as well as lung cells [52,55], lymphocytes [56], macrophages [57], keratinocytes [58], and hepatocytes [59] after endocytosis. It is also interesting to note the penetration ability of TiO2 particles with a size of 200 nm into red blood cells, which were chosen as a model for non-phagocytic cells. The results demonstrated the ability of the particles to penetrate the red blood cell membrane by a mechanism other than phagocytosis and endocytosis [60]. In this case, the penetration mechanism remained unexplained.
Upon entrance of TiO2 NP into the cell, the internalized TiO2 NPs are transported to lysosomes, where they generate lysosomal stress and release cytosol that reacts with cellular components, resulting in DNA damage, DNA rearrangement, altered gene expression, oxidative stress, and inflammation (Figure 2) [28,56,61,62,63]. TiO2 NPs impair micromolecular functions by protein adsorption, blocking signaling pathways, and binding to DNA structure [64,65]. Undoubtedly, such damage affects cell viability [66] and is dose and time dependent [67,68].
It should also be noted that the response of cells in the presence of TiO2 NPs is complex. Thus, the presence of TiO2 NPs can modulate different cell fates, including necrosis and apoptosis, which are regular cell death pathways, or autophagy, which leads to either cytoprotective mechanisms or cell death (Figure 3) [69]. TiO2 NP-induced autophagy can be exploited in new therapeutic pathway treatment of various diseases (see Section 6).

3.2. Genotoxicity

The genotoxicity of TiO2 NPs has not yet been clarified as scientists do not have sufficient evidence of genotoxicity [70,71]. Genotoxicity refers to the ability of TiO2 NPs to disrupt genetic information by causing breaks, lesions, deletions, mis-segregation, or non-disjunction in the DNA, leading to gene mutations. In vitro testing methods, such as mammalian chromosomal aberration tests, cellular gene mutation tests and bacterial reverse mutation tests, are performed to measure genotoxicity [72,73]. In contrast, the in vivo comet assay and the in vivo micronuclei/chromosome aberration assay evaluate in vivo genotoxicity [74]. The genotoxic effect of TiO2 NPs on cells is mainly studied through the circulatory or respiratory system. Despite the crystallinity of TiO2 NPs, their genotoxicity mostly depends on their particle size. Smaller TiO2 NPs possess a stronger genotoxic effect than larger ones, as they easily penetrate into the nucleus and cytoplasm of the cell [74]. Larger agglomerations of TiO2 NPs cause DNA damage [75]. Several studies have shown the genotoxic and cytotoxic effects of TiO2 NPs on human amniotic epithelial cells [76], human lung fibroblasts [77], human lymphocytes [78], and human hepatoma HepG2 cells [79]. In vitro studies of cell-induced genotoxicity caused by TiO2 NPs due to DNA breakage and gene mutations are shown in Figure 4 [22].
Genotoxicity occurs via direct or indirect genotoxicity mechanisms. Namely, TiO2 NPs can enter the nucleus, inducing direct DNA damage through direct contact with DNA and chromosome, while indirect genotoxicity of TiO2 NPs results from the increased lysosomal release of DNases, the formation of nanoaggregates that can extrude nucleus or by ROS accumulation. Moreover, TiO2 NPs can also negatively influence the repair process of DNA. In vitro studies showed the genotoxicity of TiO2 NPs during short-term exposure, which may be triggered by smaller particle sizes and mixed phases of TiO2 NPs [22].
Research has provided contradictory results regarding TiO2 NPs’ genotoxicity [7,80]. Several studies found TiO2 NPs to have no genotoxic effect [7,57,80,81]. Brandão et al. examined TiO2 NP-induced genotoxicity in human lung, liver, glial, and neuron cells [82]. They found no genotoxicity, while TiO2 NPs were successfully internalized by the experimental cells.

4. Health Effects of Exposure to TiO2 NPs

4.1. Immune System

The immune system defends the body against foreign antigens. If nanoparticles are recognized as foreign substances, they are eliminated by the immune system. If, on the other hand, the foreign substances are not recognized as a threat, they are either ignored or tolerated by the immune system [83]. The effects of TiO2 NPs on immune function are poorly documented. Underlying molecular mechanism by which TiO2 NPs influence the immune cell was studied in relation to TLRs, which are a subfamily of pattern recognition receptors, placed strategically on the cell surface and endosome of primary immune cell macrophages [84].TiO2-induced immunotoxicity was proven, which occurred through the activation of several receptors, which further activated specific signaling pathways to reduce the antioxidants through the formation of ROS. Upon increased ROS exposure, the mitochondrial membrane potential (Δψm) reduced, ultimately leading to apoptotic cell death, and inducing immunotoxicity through immune redox imbalance (Figure 5). However, in vivo experiments are essential to further validate TiO2 NPs influence on the immune system.

4.2. Neural System

TiO2 NPs can enter the brain regions by translocation across the blood–brain barrier or thenose–brain barrier, and progeny across the placental barrier, which may be the reason for the dysfunction and potential risks to the central nervous system [32,85,86,87]. Long-term exposure to TiO2 NPs can cause damages to the neurons and glial cells (U373), which may subsequently lead to neurotoxicity [32,88], also at low doses. In Figure 6, it is shown that the oral ingestion of TiO2 NPs can affect brain activities via increased oxidative stress, decreased antioxidant enzyme activity, and increased nitric oxide (NO) and ROS release (Figure 6) [89]. TiO2 NPs thus induced a neurotoxic damage accompanied by the increase in degenerated and apoptotic neurons in cerebral cortex.

4.3. Cardiovascular System

TiO2 NPs of 100 nm significantly reduce mitochondrial ‘dehydrogenase activity’ in human lymphocyte cells [81]. Mitochondria-mediated apoptosis-based cell death due to DNA damage is induced by TiO2 NPs. When TiO2 NPs react with hemoglobin, they can weaken red blood cell oxygen transport. Kongseng et al. investigated the cytotoxic effects of TiO2 NPs on human blood cells, namely peripheral blood mononuclear cells (PBMCs). TiO2 NPs incubated for 24 h significantly suppressed cell viability and increased the formation of toxic mediators (Figure 7) [10]. At high TiO2 NP concentrations (≥ 25 μg mL−1), cell apoptosis and the ‘proinflammatory cytokine secretion’ of PBMCs increased due to oxidative stress caused by ROS. The effect of TiO2 NPs and bulk material in PBMCs, neopterin formation and tryptophan degradation was studied by Becker et al. [90]. Neopterin production was increased in both unstimulated and stimulated PBMCs, while tryptophan breakdown was suppressed, thus suggesting that the total effect of TiO2 NPs was strongly pro-inflammatory. Namely, in human body fluids, such response is detected in diseases such as infections and cancer, and is also parallel to the course of atherogenesis and neurodegeneration [90].
In contrast, no significant DNA damage caused to human peripheral blood lymphocytes treated with TiO2 NPs was observed by an alkaline assay. Only extremely higher concentrations (100 µg) can show the genotoxicity of TiO2 [91].
Very little research has been completed on the effect of TiO2 NPs on the cardiovascular system. Researchers suspect that deposited TiO2 NPs in the heart may lead to inflammatory responses, system malfunction, and cardiac damage. Depending on the type and duration of exposure, ultrafine TiO2 NPs could lead to an increase in heart rate, blood pressure, and cardiac muscle damage [32].

4.4. Respiratory System

The lungs are the main target system of environmental air pollution via TiO2 NPs [92]. TiO2-induced human lung epithelial cell injury (A549) and alveolar lung inflammation have been reported after inhalation in several studies [93,94,95]. These phenomena lead to lung dysfunction, as well as irreversible changes to the cells, resulting in fibrosis and tumor development. The lung toxicity and inflammatory effects are related to the properties of TiO2 NPs, such as their size, shape, crystallinity, agglomeration, and surface coating mode [96]. At this point it should be noted that the cytotoxicity of TiO2 NPs increased in the following order: amorphous > anatase > anatase/rutile; thus, amorphous TiO2 NPs possessed greater toxic effect than anatase/rutile TiO2 NPs [95]. Fresegna et al. studied the cellular responses of human alveolar A549 and bronchial BEAS-2B cells to measure the cytotoxic and inflammatory effects of TiO2 NPs in anatase and rutile forms [7]. They found that anatase TiO2 NPs exerted greater cytotoxicity on bronchial cells compared to rutile TiO2 NPs. On the other hand, a higher level of genotoxicity was observed on alveolar and bronchial cells treated with rutile TiO2 compared to anatase TiO2. When TiO2 NPs are pre-irradiated, they have a greater cytotoxic effect on human lung cells compared to non-irradiated TiO2 NPs [93].

4.5. Digestive System

Oral ingestion of TiO2 NPs causes them to enter the digestive system. The blood passes through the stomach and also passes through the liver. Significant concentrations of TiO2 NPs can lead to liver dysfunction, liver cell damage and even liver failure, hepatocyte dysfunction, superficial staining of cytoplasm, and osteoporosis after repeated exposure [40,92]. However, these results, carried out in mice or rats, are controversial, as no toxicity was observed in other studies.

4.6. Urinary System

The kidneys are one of the major organs that filter the blood, remove metabolic wastes, control the body’s extracellular fluid balance and electrolyte composition, and return the purified blood to the body. Chen et al. demonstrated an association between TiO2 NPs and kidney toxicity, suggesting that the kidney may be a major target or organ of exposure to nano-TiO2 via various routes into the body [92]. When cultured embryonic kidney cells were incubated with TiO2 NPs, no significant induction of DNA damage was observed. In this case, only the highest concentration of TiO2 NPs, equivalent to 100 μg/mL, elicited a significant genotoxic response, but it was concluded that such a high concentration of TiO2 is not environmentally relevant [91]. However, further studies on renal toxicity in humans due to exposure to TiO2 NPs need to be conducted.

4.7. Reproductive System

Based on the research conducted to date, the toxicity of TiO2 NPs to the reproductive system such as testes, ovaries, placenta, and fetal tissues in humans is unknown. TiO2 NPs cause adverse effects on hatching and affect reproduction in zebrafish, as well as pregnancy difficulties such as nanoparticle spread to fetal brain, fetal liver, and placenta in mice [13]. However, the research data are not sufficient to conclude the development of toxicity in the reproductive system of mammals, especially humans.

4.8. Dermal System

The penetration of TiO2 NPs through human skin has not been found in most studies [97,98,99]. TiO2 NPs do not cause ROS formation, cellular glutathione content, nor apoptosis when applied to human epidermal (A431) and keratinocytic (HaCaT) skin cells [100,101]. TiO2 NPs cause dermal toxicity only when they pass through healthy or damaged skin after long-term exposure [40]. However, the opposite phenomenon has been observed in several studies. Wright et al. showed that TiO2 NPs induce superoxide formation, caspase, and cell apoptosis in human keratinocyte cells (HaCaTs) in a dose-dependent manner. This causes cytotoxicity in HaCaT cells at 10−4–10−5 mol/L [100]. In fact, human skin is not only exposed to TiO2 NPs, but also external chemicals or stressors, such as UV light [102], which also damage human dermal fibroblasts. The Scientific Committee on Consumer Safety (SCCS) therefore suggested that TiO2 NPs should not be used in sunscreen formulations with high photocatalytic activity [1].

5. TiO2 NPs in the Environment (Ecotoxicity)

Ecotoxicity occurs when biological, physical, and chemical stressors affect living organisms in the ecosystem through altered biochemistry, physiology, and cellular interactions. The ecotoxic adverse effects of TiO2 NPs have been observed in water, aquatic animals, zebrafish gills, food, and aquatic environments [67]. TiO2 NPs influence the bacterial colonies in soil, reducing microbial biomass and diversity, thereby having a negative effect by changing the bacterial composition of the ecosystem [103,104]. In addition, terrestrial plants collect TiO2 NPs from the soil and store them in stems, leaves, and fruits, which promotes germination and root expansion [105,106].

5.1. TiO2 NPs in the Plant and Soil Environment

After the use of TiO2 NPs in various products, this inorganic nanomaterial is mostly abandoned, and a large quantity of TiO2 NPs is distributed to the environment, reaching the air, soil, water, and living organisms [54]. Figure 8 represents the use, release pathways, distribution, and interaction of TiO2 NPs to plants and the surrounding environment. TiO2 NPs used in pigments, food additives, and personal care products are released into the soil (13.8%), water (18.5%), and air (2.2%). Therefore, plants come into direct contact with TiO2 NPs through the soil, water, and air.
After the entry of TiO2 NPs via several pathways, soil properties—such as soil enzymes, microbial communities, nutritional elements, and pH—affect their behavior, mobility, and bioavailability, thereby determining the fate of TiO2 NPs. The interconnected soil properties/factors, such as soil type, pH, ionic strength, and organic matter, affect the transport of TiO2 NPs by changing their zeta potential, aggregation, surface charge, and van der Waals force [108,109]. The existence of microbes in the soil is vitally important for the decomposition and recycling of organic material. As a result, TiO2 NPs may alter the microbial population, diversity, and activity by changing the soil properties [103,110]. Low concentrations of TiO2 NPs also increase urea activity [111]. On the other hand, at extremely higher concentrations (1000 mg/L), TiO2 NPs decrease urea activity [112,113], affect the level of bacterial nitrogen, and reduce catalase, phosphatase, invertase, and peroxidase activities [103]. Reports suggests that TiO2 NPs also disrupt the gene expression of bacteria, resulting in decreasing nitrogen fixation and methane oxidation, which are essential for the decomposition of proteins and organic pollutants [81].
The interaction between plants and TiO2 NPs depends on their particle size, crystal phase, and surface coating [107]. Smaller TiO2 NPs (less than 30 nm) can enter into plant cells by reducing the size of the pores and the flow of water in corn [114] and wheat [115]. Investigations show that TiO2 NPs of 12, 22, and 25 nm can be translocated from the roots to the leaves [115]. The antioxidant stress is interrupted by TiO2 NPs in duckweed [116] and tomato [117]. TiO2 NPs also interrupt different parameters in raceme elm [118], onion [119], soybean [120], rice, spinach [121], and parsley [122]. As TiO2 NPs show photocatalytic activity under light irradiation, they can alter photosynthesis, metabolism, and gene expression within plants. Some investigations reported negative effects caused by TiO2 NPs on plant growth, whereas it was concluded that TiO2 NPs may cause some sensitive plant growth-promoting bacteria to disappear from soil. Accordingly, such impairment of the soil bacterial community composition may further affect ecosystem functioning [123,124].

5.2. TiO2 NPs in the Aquatic Environment

The significantly increased leakage of TiO2 NPs into surface and marine water environments has a great impact on aquatic ecosystems. The related studies have been mainly concentrated in determining the behavior of TiO2 NP in marine environments, focusing mostly on marine plankton and benthos, with comparable contribution of papers, i.e., 42.1% and 44.7%, respectively, as well as in marine fish, with the smallest share of the research, i.e., 13.2% [125]. As TiO2 NPs are very reactive, they follow different transformation processes when they are released into the aquatic ecosystem. These transformation processes involve physical (agglomeration, aggregation, and sedimentation) interactions with TiO2 (adsorption), chemical (photochemical) interactions, and biological (biomodification) interactions. The ability of TiO2 NPs to enter in the aquatic organisms is worrying, as it leads to bioaccumulation in their cellular tissue. Accordingly, NPs negatively affect environmental food webs in three different major ecosystems—freshwater, marine, and terrestrial [24,126]. Bearing in mind that mammalians are at the top of the ecological food chain, the ecotoxicity induced by TiO2 NPs in the environment is easily converted into cytotoxicity in humans (Figure 9) [24].
Phytoplankton is the most dominant factor in marine ecosystems and the food web [127]. The interaction of TiO2 NPs with marine phytoplankton has therefore been investigated [127,128]. In an aquatic environment, TiO2 NPs may be adsorbed or diffused by the phytoplankton surface. TiO2 NP-mediated ROS may be diffused by the cell wall when the TiO2–plankton complex generates ‘ligand-to-metal charge transfer reactions’ [126]. After the aggregation and settling of TiO2 in marine environments, there is still a small fraction of TiO2 NP nanoparticles in the water column that may be hazardous to the living organisms in that water column. Thus, at higher TiO2 concentrations (>/20 mg/L), TiO2 NPs can significantly reduce P. tricornutum growth, one of the most widely used model organisms used in marine ecotoxicology studies [127]. Miller et al. demonstrated that, at low UV levels, TiO2 NPs’ photocatalytic activity can induce toxicity in marine phytoplankton [128]. However, in the absence of UV light, no effects were observed.
The toxicity of TiO2 NPs in marine mussels has been investigated in several studies [38,129,130,131,132,133,134]. Due to the filtrating behavior and bioaccumulation tendency of bivalve mollusks, they accumulate various pollutants, such as microalgae, sediments, bacteria, and contaminants, within their tissues [38]. However, no acute TiO2 NP toxicity was found in marine abalone at TiO2 NP concentrations from 0.1 to 10.0 mg/L [135]. In spite of the absence of toxicity, minor oxidative stress was induced. In another study, the combined effect of TiO2 NPs and ocean acidification was assessed in mussels [136]. A low pH increases the toxicity of TiO2 NPs and the impairment of feeding and metabolism was observed in mussels at different pH and concentrations.
In a different study, the TiO2 NP–fish interaction was examined at different TiO2 NP concentrations and exposure conditions [137]. TiO2 NPs of 0, 1, 10, and 100 mg/L were applied to fish for 96 h, resulting in no mortality or sublethal effects. Very few studies associated with TiO2 NPs have been conducted on fish. As such, more studies need to be conducted in order to understand the potentially cytotoxic effects of TiO2 NPs on fish.
Moreover, in aquatic environments, TiO2 NPs can interact with heavy metals (Cu, Zn, Cd, As, etc.) and toxic organics, resulting in the formation of a harmful environment that can alter the bioavailability of aquatic organisms [38,138,139]. If TiO2 NPs are exposed to arsenic (As), this can increase arsenic accumulation in aquatic animals and the human food web [139]. However, TiO2 NP–Cd interactions do not result in toxicity in Mediterranean mussels [140]. The immunotoxicity, genotoxicity, and neurotoxicity of TiO2 NPs in marine living organisms were reported by several researchers, with varying results depending on the particle size, exposure duration, exposure type, and stress factors [38,134,141,142,143,144,145].

6. Biocompatibility of TiO2 NPs

The potential toxicity of the TiO2 NPs discussed above does not suggest that they are unsafe for humans or the environment. Depending on the size and shape of the nanoparticles, TiO2 NPs may be safe, due to their very low toxicity. When it comes to the issue of skin permeation, long-term skin exposure to TiO2 NPs can be harmful to humans if they overdose, which is almost impossible in everyday life [146]. Biocompatibility refers to the testing of TiO2 NPs for cytotoxicity, genotoxicity, immunotoxicity, systemic toxicity, hemocompatibility, pyrogenicity, and implantation, evaluated by ISO, ISO/TR 10993-22:2017 via in vitro and in vivo studies [47]. The evaluation of the toxicity and biocompatibility of TiO2 NPs is crucial to understanding the deleterious biological responses of the properties of TiO2 NPs, their functionalities, and their contact surfaces (Figure 10a).
In terms of the biocompatibility of TiO2 NPs, they are used in drug carrier biosensing, implants, and antibacterial activity [148]. Due to the spherical shape of TiO2, TiO2–cell interactions, TiO2′s biocompatibility, excellent drug release properties, and lower toxicity (than Al2O3 and SiO2) [148,149,150], TiO2 NPs are used in antitumor treatments (Figure 10b). Due to the cytotoxic ROS (O2, OH, H2O2)-based interaction between TiO2 NPs and specific cancer cell membranes, surface-functionalized biocompatible TiO2 NPs have recently been used for targeted cancer therapy [48]. TiO2 NPs have also been used in the biomedical field in photothermal therapy (PTT), photodynamic therapy (PDT), and sonodynamic therapy (SDT) for cancer treatment via the targeted, controlled, stimulus-driven delivery and release of cytotoxic anti-cancer agents [148]. Due to their low phototoxicity, biocompatibility and stable structure, TiO2 NPs have potential applications in phototherapy for the treatment of cancer cells [49,151]. When TiO2 NPs are exposed to light, an oxidative radical (ROS) is generated, which subsequently destroys the cell and cellular components, such as the lipids, proteins, carbohydrates, and nucleic acids in cancer cells [147,152].
Furthermore, several authors have reported that the toxicity of TiO2 NPs arises due to a ‘particle effect’ rather than a ‘chemical effect’ [40]. However, further studies are required to assess the nature, mechanisms, and effects of the toxicity of TiO2 NPs on humans and the surrounding environment.

7. Risk Management

7.1. Risk Following Oral Exposure

Considering the published studies, TiO2 NPs have no acute toxic effects after oral exposure. There are insufficient data on the repeated-dose toxicity of TiO2 NPs, so there is not believed to be any significant risk from oral exposure.

7.2. Risk Following Dermal Exposure

Based on the current data, short-term dermal exposure to TiO2 NPs has very little or no toxic effect on healthy skin. Further studies on long-term dermal exposure, as well as on damaged skin, are required in order to evaluate TiO2 NPs’ nanotoxicity on the skin [40].

7.3. Risk Following Inhalation

It is well established that smaller TiO2 NPs are more toxic than comparatively coarser TiO2 NPs. Anatase TiO2 is more toxic than rutile TiO2, with its crystalline structure. In terms of inhalation, particle size, surface area, crystalline structure, agglomeration, and exposure, time plays a crucial role in the toxicity of TiO2 NPs. Inhalation is a major problem in workplaces due to high concentrations of TiO2 NPs, as well as the fact that TiO2 NPs enter the body mainly through respiration.
The establishment of risk management strategies is therefore crucial (Figure 11), particularly in relation to short-/long-term exposure, as well as the frequency and level of exposure; however, such strategies have not yet been developed. Regulations and legalization for controlling TiO2 NPs and other engineered NPs (ENPs) are unsure, living the interpretation of the scientific data challenging to the government, agencies, industry, and consumers. Recently, suggestions for the safe handling of TiO2 NPs and other engineered NPs (ENPs) were proposed by Besha et al. [153], highlighting the mitigation approaches to curtail the possible hazard effects of ENPs.

8. Conclusions and Future of TiO2 NPs in Textile Applications

TiO2 NPs are important photocatalysts that are implemented in various fields of application. The toxicity of TiO2 NPs to human health and the environment is still controversial. TiO2 NPs are classified as possibly carcinogenic to humans by the International Agency for Research on Cancer (IARC), and as a chemical risk to humans by the Workplace Hazardous Materials Information System (WHMIS) (in group D2A). At lower TiO2 NP concentrations, recent studies have found that TiO2 NPs are almost non-toxic, as no remarkable toxicity was observed [99,154,155]. However, TiO2/textile composites, for example, distinguished by remarkable functional properties [156], have no adverse effects on human skin, as superoxide and hydroxyl radicals are not able to damage human cells from the outside as they cannot enter the inner layer of the human skin [155,157]. As TiO2/rGO-coated cotton fabric is biocompatible, it does not cause cell membrane damage and cell proliferation, as textiles coated with TiO2 NPs do not exhibit cytotoxicity over 24 h of incubation [158]. With the exception of the other previously mentioned study, to the best of our current knowledge, no experiments have been conducted with TiO2 NP-modified textiles on human skin. Considering the toxicity and ecotoxic pathways described in in vivo and in vitro studies, TiO2 NPs are the safer option when compared to other metal oxide nanomaterials [148]. The nanotoxicity and contradictive biocompatibility of TiO2 NPs should be further investigated by researchers, whereas studies on long-term or accelerated effects of TiO2 need to be carried out in order to estimate precise implications of TiO2 NPs for human health and the surrounding environment. Therefore, careful consideration should be given to the benefits of TiO2, and the associated potential risks based on its intended use. Moreover, mitigating the toxicity of TiO2 NPs transferred from consumer products to the environment requires appropriate strategies and regulatory frameworks to protect humans and the environment. Accordingly, the establishment of international standard methods for the exact evaluation of risk–benefit assessments is a prerequisite to allowing the safe use of TiO2-functionalized materials.

Author Contributions

Conceptualization; investigation, and writing—original draft preparation, M.M.R.; writing—review and editing and supervision, P.F.T.; and conceptualization and writing—review and editing, B.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Slovenian Research Agency (Programs P2-0213 and Infrastructural Centre RIC UL-NTF).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, X.; Li, W.; Yang, Z. Toxicology of nanosized titanium dioxide: An update. Arch. Toxicol. 2015, 89, 2207–2217. [Google Scholar] [CrossRef]
  2. Shabbir, S.; Kulyar, M.F.-E.-A.; Bhutta, Z.A.; Boruah, P.; Asif, M. Toxicological Consequences of Titanium Dioxide Nanoparticles (TiO2NPs) and Their Jeopardy to Human Population. BioNanoScience 2021, 11, 621–632. [Google Scholar] [CrossRef]
  3. Scharnberg, A.R.D.A.; de Loreto, A.C.; Wermuth, T.B.; Alves, A.K.; Arcaro, S.; dos Santos, P.A.M.; Rodriguez, A.D.A.L. Porous ceramic supported TiO2 nanoparticles: Enhanced photocatalytic activity for Rhodamine B degradation. Boletín Soc. Española Cerám. Vidrio 2020, 59, 230–238. [Google Scholar] [CrossRef]
  4. Do, H.H.; Nguyen, D.L.T.; Nguyen, X.C.; Le, T.-H.; Nguyen, T.P.; Trinh, Q.T.; Ahn, S.H.; Vo, D.-V.N.; Kim, S.Y.; Van Le, Q. Recent progress in TiO2-based photocatalysts for hydrogen evolution reaction: A review. Arab. J. Chem. 2020, 13, 3653–3671. [Google Scholar] [CrossRef]
  5. Hiremath, V.; Deonikar, V.G.; Kim, H.; Gil Seo, J. Hierarchically assembled porous TiO2 nanoparticles with enhanced photocatalytic activity towards Rhodamine-B degradation. Colloids Surf. A Physicochem. Eng. Asp. 2020, 586, 124199. [Google Scholar] [CrossRef]
  6. Bai, X.; Jia, J.; Du, Y.; Hu, X.; Li, J.; Liu, E.; Fan, J. Multi-level trapped electrons system in enhancing photocatalytic activity of TiO2 nanosheets for simultaneous reduction of Cr (VI) and RhB degradation. Appl. Surf. Sci. 2020, 503, 144298. [Google Scholar] [CrossRef]
  7. Fresegna, A.M.; Ursini, C.L.; Ciervo, A.; Maiello, R.; Casciardi, S.; Iavicoli, S.; Cavallo, D. Assessment of the Influence of Crystalline Form on Cyto-Genotoxic and Inflammatory Effects Induced by TiO2 Nanoparticles on Human Bronchial and Alveolar Cells. Nanomaterials 2021, 11, 253. [Google Scholar] [CrossRef]
  8. Alotaibi, A.M.; Williamson, B.A.D.; Sathasivam, S.S.; Kafizas, A.; Alqahtani, M.; Sotelo-Vazquez, C.; Buckeridge, J.; Wu, J.; Nair, S.P.; Scanlon, D.O.; et al. Enhanced Photocatalytic and Antibacterial Ability of Cu-Doped Anatase TiO2 Thin Films: Theory and Experiment. ACS Appl. Mater. Interfaces 2020, 12, 15348–15361. [Google Scholar] [CrossRef] [Green Version]
  9. Zdravkovic, B.; Zdravkovic, T.P.; Zdravkovic, M.; Strukelj, B.; Ferk, P. The influence of nano-TiO2 on metabolic activity, cytotoxicity and ABCB5 mRNA expression in WM-266-4 human metastatic melanoma cell line. J. Balk. Union Oncol. 2019, 24, 338–346. [Google Scholar]
  10. Kongseng, S.; Yoovathaworn, K.; Wongprasert, K.; Chunhabundit, R.; Sukwong, P.; Pissuwan, D. Cytotoxic and inflammatory responses of TiO2nanoparticles on human peripheral blood mononuclear cells. J. Appl. Toxicol. 2016, 36, 1364–1373. [Google Scholar] [CrossRef] [PubMed]
  11. Shakeel, M.; Jabeen, F.; Shabbir, S.; Asghar, M.S.; Khan, M.S.; Chaudhry, A.S. Toxicity of Nano-Titanium Dioxide (TiO2-NP) Through Various Routes of Exposure: A Review. Biol. Trace Elem. Res. 2016, 172, 1–36. [Google Scholar] [CrossRef] [PubMed]
  12. Aliakbari, F.; Hosseinali, S.H.; Sarokhalil, Z.K.; Shahpasand, K.; Saboury, A.A.; Akhtari, K.; Falahati, M. Reactive oxygen species generated by titanium oxide nanoparticles stimulate the hemoglobin denaturation and cytotoxicity against human lymphocyte cell. J. Biomol. Struct. Dyn. 2019, 37, 4875–4881. [Google Scholar] [CrossRef]
  13. Dar, G.I.; Saeed, M.; Wu, A. Toxicity of TiO2 Nanoparticles. In TiO2 Nanoparticles: Applications in Nanobiotechnology and Nanomedicine; Wiley: Hoboken, NJ, USA, 2020; pp. 67–103. [Google Scholar] [CrossRef]
  14. Madhubala, V.; Pugazhendhi, A.; Thirunavukarasu, K. Cytotoxic and immunomodulatory effects of the low concentration of titanium dioxide nanoparticles (TiO2 NPs) on human cell lines—An in vitro study. Process. Biochem. 2019, 86, 186–195. [Google Scholar] [CrossRef]
  15. Lappas, C.M. The immunomodulatory effects of titanium dioxide and silver nanoparticles. Food Chem. Toxicol. 2015, 85, 78–83. [Google Scholar] [CrossRef] [PubMed]
  16. Kazimirova, A.; El Yamani, N.; Rubio, L.; García-Rodríguez, A.; Barancokova, M.; Marcos, R.; Dusinska, M. Effects of Titanium Dioxide Nanoparticles on the Hprt Gene Mutations in V79 Hamster Cells. Nanomaterials 2020, 10, 465. [Google Scholar] [CrossRef] [Green Version]
  17. Zhu, Y.; Wu, J.; Chen, M.; Liu, X.; Xiong, Y.; Wang, Y.; Feng, T.; Kang, S.; Wang, X. Recent advances in the biotoxicity of metal oxide nanoparticles: Impacts on plants, animals and microorganisms. Chemosphere 2019, 237, 124403. [Google Scholar] [CrossRef]
  18. 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]
  19. Tarrahi, R.; Mahjouri, S.; Khataee, A. A review on in vivo and in vitro nanotoxicological studies in plants: A headlight for future targets. Ecotoxicol. Environ. Saf. 2021, 208, 111697. [Google Scholar] [CrossRef]
  20. Hayes, A.W.; Sahu, S.C. Genotoxicity of engineered nanomaterials found in the human environment. Curr. Opin. Toxicol. 2020, 19, 68–71. [Google Scholar] [CrossRef]
  21. Cox, A.; Venkatachalam, P.; Sahi, S.; Sharma, N. Silver and titanium dioxide nanoparticle toxicity in plants: A review of current research. Plant Physiol. Biochem. 2016, 107, 147–163. [Google Scholar] [CrossRef]
  22. Ling, C.; An, H.; Li, L.; Wang, J.; Lu, T.; Wang, H.; Hu, Y.; Song, G.; Liu, S. Genotoxicity Evaluation of Titanium Dioxide Nanoparticles In Vitro: A Systematic Review of the Literature and Meta-analysis. Biol. Trace Elem. Res. 2021, 199, 2057–2076. [Google Scholar] [CrossRef]
  23. Markowska-Szczupak, A.; Endo-Kimura, M.; Paszkiewicz, O.; Kowalska, E. Are Titania Photocatalysts and Titanium Implants Safe? Review on the Toxicity of Titanium Compounds. Nanomaterials 2020, 10, 2065. [Google Scholar] [CrossRef]
  24. Sharma, S.; Sharma, R.K.; Gaur, K.; Torres, J.F.C.; Loza-Rosas, S.A.; Torres, A.; Saxena, M.; Julin, M.; Tinoco, A.D. Fueling a Hot Debate on the Application of TiO2 Nanoparticles in Sunscreen. Materials 2019, 12, 2317. [Google Scholar] [CrossRef] [Green Version]
  25. Adler, B.L.; DeLeo, V.A. Sunscreen Safety: A Review of Recent Studies on Humans and the Environment. Curr. Dermatol. Rep. 2020, 9, 1–9. [Google Scholar] [CrossRef]
  26. Akın, N.; Danacı, H.M. An investigation into the architectural use of nanotechnology in the context of the titanium dioxide. Environ. Sci. Pollut. Res. 2021, 93, 1–7. [Google Scholar] [CrossRef]
  27. Chen, Z.; Han, S.; Zhou, S.; Feng, H.; Liu, Y.; Jia, G. Review of health safety aspects of titanium dioxide nanoparticles in food application. NanoImpact 2020, 18, 100224. [Google Scholar] [CrossRef]
  28. Luo, Z.; Li, Z.; Xie, Z.; Sokolova, I.M.; Song, L.; Peijnenburg, W.J.G.M.; Hu, M.; Wang, Y. Rethinking Nano-TiO2 Safety: Overview of Toxic Effects in Humans and Aquatic Animals. Small 2020, 16, 2002019. [Google Scholar] [CrossRef] [PubMed]
  29. Tibau, A.V.; Grube, B.D.; Velez, B.J.; Vega, V.M.; Mutter, J. Titanium exposure and human health. Oral Sci. Int. 2019, 16, 15–24. [Google Scholar] [CrossRef] [Green Version]
  30. Bischoff, N.S.; De Kok, T.M.; Sijm, D.T.; Van Breda, S.G.; Briedé, J.J.; Castenmiller, J.J.; Opperhuizen, A.; Chirino, Y.I.; Dirven, H.; Gott, D.; et al. Possible Adverse Effects of Food Additive E171 (Titanium Dioxide) Related to Particle Specific Human Toxicity, Including the Immune System. Int. J. Mol. Sci. 2020, 22, 207. [Google Scholar] [CrossRef] [PubMed]
  31. Cao, Y.; Gong, Y.; Liao, W.; Luo, Y.; Wu, C.; Wang, M.; Yang, Q. A review of cardiovascular toxicity of TiO2, ZnO and Ag nanoparticles (NPs). BioMetals 2018, 31, 457–476. [Google Scholar] [CrossRef] [PubMed]
  32. Baranowska-Wójcik, E.; Szwajgier, D.; Oleszczuk, P.; Winiarska-Mieczan, A. Effects of Titanium Dioxide Nanoparticles Exposure on Human Health—A Review. Biol. Trace Elem. Res. 2020, 193, 118–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Praphawatvet, T.; Peters, J.I.; Williams, R.O. Inhaled nanoparticles—An updated review. Int. J. Pharm. 2020, 587, 119671. [Google Scholar] [CrossRef] [PubMed]
  34. Kabir, E.; Kumar, V.; Kim, K.-H.; Yip, A.C.; Sohn, J. Environmental impacts of nanomaterials. J. Environ. Manag. 2018, 225, 261–271. [Google Scholar] [CrossRef] [PubMed]
  35. Martínez, G.; Merinero, M.; Pérez-Aranda, M.; Pérez-Soriano, E.; Ortiz, T.; Villamor, E.; Begines, B.; Alcudia, A. Environmental Impact of Nanoparticles’ Application as an Emerging Technology: A Review. Materials 2020, 14, 166. [Google Scholar] [CrossRef] [PubMed]
  36. Heilgeist, S.; Sekine, R.; Sahin, O.; Stewart, R. Finding Nano: Challenges Involved in Monitoring the Presence and Fate of Engineered Titanium Dioxide Nanoparticles in Aquatic Environments. Water 2021, 13, 734. [Google Scholar] [CrossRef]
  37. Lehutso, R.; Tancu, Y.; Maity, A.; Thwala, M. Aquatic toxicity of transformed and product-released engineered nanomaterials: An overview of the current state of knowledge. Process. Saf. Environ. Prot. 2020, 138, 39–56. [Google Scholar] [CrossRef]
  38. Abdel-Latif, H.M.; Dawood, M.; Menanteau-Ledouble, S.; El-Matbouli, M. Environmental transformation of n-TiO2 in the aquatic systems and their ecotoxicity in bivalve mollusks: A systematic review. Ecotoxicol. Environ. Saf. 2020, 200, 110776. [Google Scholar] [CrossRef]
  39. Yang, W.; Peters, J.I.; Williams, R.O., III. Inhaled nanoparticles—A current review. Int. J. Pharm. 2008, 356, 239–247. [Google Scholar] [CrossRef]
  40. Christensen, F.M.; Johnston, H.J.; Stone, V.; Aitken, R.J.; Hankin, S.; Peters, S.; Aschberger, K. Nano-TiO2—Feasibility and challenges for human health risk assessment based on open literature. Nanotoxicology 2010, 5, 110–124. [Google Scholar] [CrossRef]
  41. Kreyling, W.G.; Holzwarth, U.; Schleh, C.; Kozempel, J.; Wenk, A.; Haberl, N.; Hirn, S.; Schäffler, M.; Lipka, J.; Semmler-Behnke, M.; et al. Quantitative biokinetics of titanium dioxide nanoparticles after oral application in rats: Part 2. Nanotoxicology 2017, 11, 443–453. [Google Scholar] [CrossRef]
  42. Heyder, J.; Gebhart, J.; Rudolf, G.; Schiller, C.; Stahlhofen, W. Deposition of particles in the human respiratory tract in the size range 0.005–15 μm. J. Aerosol Sci. 1986, 17, 811–825. [Google Scholar] [CrossRef]
  43. Hongbo, S.; Magaye, R.; Castranova, V.; Zhao, J. Titanium dioxide nanoparticles: A review of current toxicological data. Part. Fibre Toxicol. 2013, 10, 1–33. [Google Scholar] [CrossRef] [Green Version]
  44. Kermanizadeh, A.; Balharry, D.; Wallin, H.; Loft, S.; Møller, P. Nanomaterial translocation–the biokinetics, tissue accumulation, toxicity and fate of materials in secondary organs—A review. Crit. Rev. Toxicol. 2015, 45, 837–872. [Google Scholar] [CrossRef] [PubMed]
  45. Ren, W.; Yan, Y.; Zeng, L.; Shi, Z.; Gong, A.; Schaaf, P.; Wang, D.; Zhao, J.; Zou, B.; Yu, H.; et al. A Near Infrared Light Triggered Hydrogenated Black TiO2 for Cancer Photothermal Therapy. Adv. Health Mater. 2015, 4, 1526–1536. [Google Scholar] [CrossRef] [PubMed]
  46. Ren, W.; Iqbal, M.Z.; Zeng, L.; Chen, T.; Pan, Y.; Zhao, J.; Yin, H.; Zhang, L.; Zhang, J.; Li, A.; et al. Black TiO2 based core–shell nanocomposites as doxorubicin carriers for thermal imaging guided synergistic therapy of breast cancer. Nanoscale 2017, 9, 11195–11204. [Google Scholar] [CrossRef] [PubMed]
  47. Kyriakides, T.R.; Raj, A.; Tseng, T.H.; Xiao, H.; Nguyen, R.; Mohammed, F.S.; Halder, S.S.; Xu, M.; Wu, M.J.; Bao, S.; et al. Biocompatibility of nanomaterials and their immunological properties. Biomed. Mater. 2021, 16, 042005. [Google Scholar] [CrossRef]
  48. Thevenot, P.; Cho, J.; Wavhal, D.; Timmons, R.B.; Tang, L. Surface chemistry influences cancer killing effect of TiO2 nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2008, 4, 226–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Jukapli, N.M.; Bagheri, S. Recent developments on titania nanoparticle as photocatalytic cancer cells treatment. J. Photochem. Photobiol. B Biol. 2016, 163, 421–430. [Google Scholar] [CrossRef]
  50. Hou, J.; Wang, L.; Wang, C.; Zhang, S.; Liu, H.; Li, S.; Wang, X. Toxicity and mechanisms of action of titanium dioxide nanoparticles in living organisms. J. Environ. Sci. 2019, 75, 40–53. [Google Scholar] [CrossRef]
  51. Scandalios, J.G. Oxidative stress responses—What have genome-scale studies taught us? Genome Biol. 2002, 3, 1–6. [Google Scholar] [CrossRef]
  52. Bhattacharya, K.; Davoren, M.; Boertz, J.; Schins, R.P.; Hoffmann, E.; Dopp, E. Titanium dioxide nanoparticles induce oxidative stress and DNA-adduct formation but not DNA-breakage in human lung cells. Part. Fibre Toxicol. 2009, 6, 17. [Google Scholar] [CrossRef] [Green Version]
  53. Vevers, W.; Jha, A.N. Genotoxic and cytotoxic potential of titanium dioxide (TiO2) nanoparticles on fish cells in vitro. Ecotoxicology 2008, 17, 410–420. [Google Scholar] [CrossRef] [PubMed]
  54. 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] [PubMed]
  55. Aderem, A.; Underhill, D. Mechanisms of Phagocytosis in Macrophages. Annu. Rev. Immunol. 1999, 17, 593–623. [Google Scholar] [CrossRef]
  56. Ghosh, M.; Bandyopadhyay, M.; Mukherjee, A. Genotoxicity of titanium dioxide (TiO2) nanoparticles at two trophic levels: Plant and human lymphocytes. Chemosphere 2010, 81, 1253–1262. [Google Scholar] [CrossRef]
  57. Morishige, T.; Yoshioka, Y.; Tanabe, A.; Yao, X.; Tsunoda, S.-I.; Tsutsumi, Y.; Mukai, Y.; Okada, N.; Nakagawa, S. Titanium dioxide induces different levels of IL-1β production dependent on its particle characteristics through caspase-1 activation mediated by reactive oxygen species and cathepsin B. Biochem. Biophys. Res. Commun. 2010, 392, 160–165. [Google Scholar] [CrossRef]
  58. Kocbek, P.; Teskač, K.; Kreft, M.E.; Kristl, J. Toxicological Aspects of Long-Term Treatment of Keratinocytes with ZnO and TiO2 Nanoparticles. Small 2010, 6, 1908–1917. [Google Scholar] [CrossRef] [PubMed]
  59. Sha, B.; Gao, W.; Wang, S.; Gou, X.; Li, W.; Liang, X.; Qu, Z.; Xu, F.; Lu, T.J. Oxidative stress increased hepatotoxicity induced by nano-titanium dioxide in BRL-3A cells and Sprague-Dawley rats. J. Appl. Toxicol. 2013, 34, 345–356. [Google Scholar] [CrossRef] [PubMed]
  60. Rothen-Rutishauser, B.M.; Schürch, S.; Haenni, B.; Kapp, N.; Gehr, P. Interaction of Fine Particles and Nanoparticles with Red Blood Cells Visualized with Advanced Microscopic Techniques. Environ. Sci. Technol. 2006, 40, 4353–4359. [Google Scholar] [CrossRef] [PubMed]
  61. Di Virgilio, A.L.; Reigosa, M.; de Mele, M.F.L. 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] [PubMed]
  62. Carriere, M.; Arnal, M.-E.; Douki, T. TiO2 genotoxicity: An update of the results published over the last six years. Mutat. Res. Toxicol. Environ. Mutagen. 2020, 854–855, 503198. [Google Scholar] [CrossRef]
  63. Kansara, K.; Kumar, A.; Karakoti, A.S. Combination of humic acid and clay reduce the ecotoxic effect of TiO2 NPs: A combined physico-chemical and genetic study using zebrafish embryo. Sci. Total Environ. 2020, 698, 134133. [Google Scholar] [CrossRef]
  64. Li, N.; Ma, L.; Wang, J.; Zheng, L.; Liu, J.; Duan, Y.; Liu, H.; Zhao, X.; Wang, S.; Wang, H.; et al. Interaction Between Nano-Anatase TiO2 and Liver DNA from Mice In Vivo. Nanoscale Res. Lett. 2010, 5, 108–115. [Google Scholar] [CrossRef] [Green Version]
  65. Kim, K.-P.; Lee, S.-J.; Kim, D.-H.; Hwang, D.-K. Effect of anodic aluminum oxide template imprinting on TiO2 blocking layer of flexible dye-sensitized solar cell. J. Nanosci. Nanotechnol. 2013, 13, 1888–1890. [Google Scholar] [CrossRef]
  66. Aslantürk, Ö.S. In Vitro Cytotoxicity and Cell Viability Assays: Principles, Advantages, and Disadvantages. In Genotoxicity A Predictable Risk to Our Actual World; InTech: Rijeka, Croatia, 2018. [Google Scholar]
  67. Sha, B.; Gao, W.; Wang, S.; Xu, F.; Lu, T. Cytotoxicity of titanium dioxide nanoparticles differs in four liver cells from human and rat. Compos. Part B Eng. 2011, 42, 2136–2144. [Google Scholar] [CrossRef]
  68. Sha, B.; Gao, W.; Cui, X.; Wang, L.; Xu, F. The potential health challenges of TiO2 nanomaterials. J. Appl. Toxicol. 2015, 35, 1086–1101. [Google Scholar] [CrossRef] [PubMed]
  69. Mohammadinejad, R.; Moosavi, M.A.; Tavakol, S.; Vardar, D.Ö.; Hosseini, A.; Rahmati, M.; Dini, L.; Hussain, S.; Mandegary, A.; Klionsky, D.J. Necrotic, apoptotic and autophagic cell fates triggered by nanoparticles. Autophagy 2019, 15, 4–33. [Google Scholar] [CrossRef] [Green Version]
  70. Wang, J.; Chen, C.; Liu, Y.; Jiao, F.; Li, W.; Lao, F.; Li, Y.-F.; Li, B.; Ge, C.; Zhou, G. Potential neurological lesion after nasal instillation of TiO2 nanoparticles in the anatase and rutile crystal phases. Toxicol. Lett. 2008, 183, 72–80. [Google Scholar] [CrossRef] [PubMed]
  71. Kang, J.L.; Moon, C.; Lee, H.S.; Lee, H.W.; Park, E.-M.; Kim, H.S.; Castranova, V. Comparison of the Biological Activity between Ultrafine and Fine Titanium Dioxide Particles in RAW 264.7 Cells Associated with Oxidative Stress. J. Toxicol. Environ. Health Part A 2008, 71, 478–485. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, T.; Yan, J.; Li, Y. Genotoxicity of titanium dioxide nanoparticles. J. Food Drug Anal. 2014, 22, 95–104. [Google Scholar] [CrossRef]
  73. Sv, R. Genotoxicity: Mechanisms, Testing Guidelines and Methods. Glob. J. Pharm. Pharm. Sci. 2017, 1, 133–138. [Google Scholar] [CrossRef] [Green Version]
  74. Yazdi, A.S.; Guarda, G.; Riteau, N.; Drexler, S.K.; Tardivel, A.; Couillin, I.; Tschopp, J. Nanoparticles activate the NLR pyrin domain containing 3 (Nlrp3) inflammasome and cause pulmonary inflammation through release of IL-1α and IL-1β. Proc. Natl. Acad. Sci. USA 2010, 107, 19449–19454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Long, T.C.; Tajuba, J.; Sama, P.; Saleh, N.; Swartz, C.; Parker, J.; Hester, S.; Lowry, G.V.; Veronesi, B. Nanosize Titanium Dioxide Stimulates Reactive Oxygen Species in Brain Microglia and Damages Neurons in Vitro. Environ. Health Perspect. 2007, 115, 1631–1637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Jugan, M.-L.; Barillet, S.; Simon-Deckers, A.; Herlin-Boime, N.; Sauvaigo, S.; Douki, T.; Carriere, M. Titanium dioxide nanoparticles exhibit genotoxicity and impair DNA repair activity in A549 cells. Nanotoxicology 2011, 6, 501–513. [Google Scholar] [CrossRef] [PubMed]
  77. Saquib, Q.; Al-Khedhairy, A.A.; Siddiqui, M.A.; Abou-Tarboush, F.M.; Azam, A.; Musarrat, J. Titanium dioxide nanoparticles induced cytotoxicity, oxidative stress and DNA damage in human amnion epithelial (WISH) cells. Toxicol. Vitr. 2012, 26, 351–361. [Google Scholar] [CrossRef] [PubMed]
  78. Trouiller, B.; Reliene, R.; Westbrook, A.; Solaimani, P.; Schiestl, R.H. Titanium Dioxide Nanoparticles Induce DNA Damage and Genetic Instability in vivo in Mice. Cancer Res. 2009, 69, 8784–8789. [Google Scholar] [CrossRef] [Green Version]
  79. Falck, G.C.M.; Lindberg, H.K.; Suhonen, S.; Vippola, M.; Vanhala, E.; Catalán, J.; Savolainen, K.; Norppa, H. Genotoxic effects of nanosized and fine TiO2. Hum. Exp. Toxicol. 2009, 28, 339–352. [Google Scholar] [CrossRef]
  80. Suzuki, T.; Miura, N.; Hojo, R.; Yanagiba, Y.; Suda, M.; Hasegawa, T.; Miyagawa, M.; Wang, R.-S. Genotoxicity assessment of titanium dioxide nanoparticle accumulation of 90 days in the liver of gpt delta transgenic mice. Genes Environ. 2020, 42, 1–6. [Google Scholar] [CrossRef] [Green Version]
  81. Ghosh, M.; Chakraborty, A.; Mukherjee, A. Cytotoxic, genotoxic and the hemolytic effect of titanium dioxide (TiO2) nanoparticles on human erythrocyte and lymphocyte cells in vitro. J. Appl. Toxicol. 2013, 33, 1097–1110. [Google Scholar] [CrossRef]
  82. Brandão, F.; Fernández-Bertólez, N.; Rosário, F.; Bessa, M.J.; Fraga, S.; Pásaro, E.; Teixeira, J.P.; Laffon, B.; Valdiglesias, V.; Costa, C. Genotoxicity of TiO2 Nanoparticles in Four Different Human Cell Lines (A549, HEPG2, A172 and SH-SY5Y). Nanomaterials 2020, 10, 412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Luo, Y.-H.; Chang, L.W.; Lin, P. Metal-Based Nanoparticles and the Immune System: Activation, Inflammation, and Potential Applications. BioMed Res. Int. 2015, 2015, 1–12. [Google Scholar] [CrossRef]
  84. Dhupal, M.; Oh, J.-M.; Tripathy, D.R.; Kim, S.-K.; Koh, S.B.; Park, K.-S. Immunotoxicity of titanium dioxide nanoparticles via simultaneous induction of apoptosis and multiple toll-like receptors signaling through ROS-dependent SAPK/JNK and p38 MAPK activation. Int. J. Nanomed. 2018, 13, 6735–6750. [Google Scholar] [CrossRef] [Green Version]
  85. Peters, A.; Veronesi, B.; Calderón-Garcidueñas, L.; Gehr, P.; Chen, L.C.; Geiser, M.; Reed, W.; Rothen-Rutishauser, B.; Schürch, S.; Schulz, H. Translocation and potential neurological effects of fine and ultrafine particles a critical update. Part. Fibre Toxicol. 2006, 3, 13. [Google Scholar] [CrossRef] [Green Version]
  86. Song, B.; Liu, J.; Feng, X.; Wei, L.; Shao, L. A review on potential neurotoxicity of titanium dioxide nanoparticles. Nanoscale Res. Lett. 2015, 10, 1–17. [Google Scholar] [CrossRef] [Green Version]
  87. Li, X.; Xu, S.; Zhang, Z.; Schluesener, H.J. Apoptosis induced by titanium dioxide nanoparticles in cultured murine microglia N9 cells. Chin. Sci. Bull. 2009, 54, 3830–3836. [Google Scholar] [CrossRef]
  88. Huerta-García, E.; Pérez-Arizti, J.A.; Márquez-Ramírez, S.G.; Delgado-Buenrostro, N.L.; Chirino, Y.I.; Iglesias, G.G.; López-Marure, R. Titanium dioxide nanoparticles induce strong oxidative stress and mitochondrial damage in glial cells. Free. Radic. Biol. Med. 2014, 73, 84–94. [Google Scholar] [CrossRef]
  89. Grissa, I.; ElGhoul, J.; Mrimi, R.; El Mir, L.; Ben Cheikh, H.; Horcajada, P. In deep evaluation of the neurotoxicity of orally administered TiO2 nanoparticles. Brain Res. Bull. 2020, 155, 119–128. [Google Scholar] [CrossRef] [PubMed]
  90. Becker, K.; Schroecksnadel, S.; Geisler, S.; Carriere, M.; Gostner, J.; Schennach, H.; Herlin, N.; Fuchs, D. TiO2 nanoparticles and bulk material stimulate human peripheral blood mononuclear cells. Food Chem. Toxicol. 2014, 65, 63–69. [Google Scholar] [CrossRef] [PubMed]
  91. Demir, E.; Burgucu, D.; Turna, F.; Aksakal, S.; Kaya, B. Determination of TiO2, ZrO2, and Al2O3 Nanoparticles on Genotoxic Responses in Human Peripheral Blood Lymphocytes and Cultured Embyronic Kidney Cells. J. Toxicol. Environ. Health Part A 2013, 76, 990–1002. [Google Scholar] [CrossRef]
  92. Chang, X.; Zhang, Y.; Tang, M.; Wang, B. Health effects of exposure to nano-TiO2: A meta-analysis of experimental studies. Nanoscale Res. Lett. 2013, 8, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kose, O.; Tomatis, M.; Turci, F.; Belblidia, N.-B.; Hochepied, J.-F.; Pourchez, J.; Forest, V. Short Preirradiation of TiO2 Nanoparticles Increases Cytotoxicity on Human Lung Coculture System. Chem. Res. Toxicol. 2021, 34, 733–742. [Google Scholar] [CrossRef] [PubMed]
  94. Kim, I.-S.; Baek, M.; Choi, S.-J. Comparative cytotoxicity of Al2O3, CeO2, TiO2 and ZnO nanoparticles to human lung cells. J. Nanosci. Nanotechnol. 2010, 10, 3453–3458. [Google Scholar] [CrossRef] [PubMed]
  95. Hsiao, I.-L.; Huang, Y.-J. Effects of various physicochemical characteristics on the toxicities of ZnO and TiO2 nanoparticles toward human lung epithelial cells. Sci. Total Environ. 2011, 409, 1219–1228. [Google Scholar] [CrossRef] [PubMed]
  96. 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]
  97. 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 Effects on HaCaT Cells. Int. J. Environ. Res. Public Health 2015, 12, 9282–9297. [Google Scholar] [CrossRef]
  98. 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] [PubMed] [Green Version]
  99. Geppert, M.; Schwarz, A.; Stangassinger, L.M.; Wenger, S.; Wienerroither, L.M.; Ess, S.; Duschl, A.; Himly, M. Interactions of TiO2 Nanoparticles with Ingredients from Modern Lifestyle Products and Their Effects on Human Skin Cells. Chem. Res. Toxicol. 2020, 33, 1215–1225. [Google Scholar] [CrossRef] [Green Version]
  100. Wright, C.; Iyer, A.K.V.; Wang, L.; Wu, N.; Yakisich, J.S.; Rojanasakul, Y.; Azad, N. Effects of titanium dioxide nanoparticles on human keratinocytes. Drug Chem. Toxicol. 2016, 40, 90–100. [Google Scholar] [CrossRef] [Green Version]
  101. 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]
  102. Smijs, T.G.; Pavel, S. Titanium dioxide and zinc oxide nanoparticles in sunscreens: Focus on their safety and effectiveness. Nanotechnol. Sci. Appl. 2011, 4, 95. [Google Scholar] [CrossRef] [Green Version]
  103. Ge, Y.; Priester, J.H.; Van De Werfhorst, L.C.; Schimel, J.P.; Holden, P.A. Potential Mechanisms and Environmental Controls of TiO2 Nanoparticle Effects on Soil Bacterial Communities. Environ. Sci. Technol. 2013, 47, 14411–14417. [Google Scholar] [CrossRef]
  104. Binh, C.T.T.; Tong, T.; Gaillard, J.-F.; Gray, K.A.; Kelly, J.J. Common freshwater bacteria vary in their responses to short-term exposure to nano-TiO2. Environ. Toxicol. Chem. 2013, 33, 317–327. [Google Scholar] [CrossRef]
  105. Song, U.; Jun, H.; Waldman, B.; Roh, J.; Kim, Y.; Yi, J.; Lee, E.J. Functional analyses of nanoparticle toxicity: A comparative study of the effects of TiO2 and Ag on tomatoes (Lycopersicon esculentum). Ecotoxicol. Environ. Saf. 2013, 93, 60–67. [Google Scholar] [CrossRef] [PubMed]
  106. Song, U.; Shin, M.; Lee, G.; Roh, J.; Kim, Y.; Lee, E.J. Functional Analysis of TiO2 Nanoparticle Toxicity in Three Plant Species. Biol. Trace Elem. Res. 2013, 155, 93–103. [Google Scholar] [CrossRef] [PubMed]
  107. Tan, W.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Interaction of titanium dioxide nanoparticles with soil components and plants: Current knowledge and future research needs—A critical review. Environ. Sci. Nano 2018, 5, 257–278. [Google Scholar] [CrossRef]
  108. Lin, D.; Tian, X.; Wu, F.; Xing, B. Fate and Transport of Engineered Nanomaterials in the Environment. J. Environ. Qual. 2010, 39, 1896–1908. [Google Scholar] [CrossRef] [Green Version]
  109. Pachapur, V.L.; Larios, A.D.; Cledón, M.; Brar, S.K.; Verma, M.; Surampalli, R. Behavior and characterization of titanium dioxide and silver nanoparticles in soils. Sci. Total Environ. 2016, 563–564, 933–943. [Google Scholar] [CrossRef]
  110. Cai, L.; Tong, M.; Wang, X.; Kim, H. Influence of Clay Particles on the Transport and Retention of Titanium Dioxide Nanoparticles in Quartz Sand. Environ. Sci. Technol. 2014, 48, 7323–7332. [Google Scholar] [CrossRef]
  111. Du, W.; Sun, Y.; Ji, R.; Zhu, J.; Wu, J.; Guo, H. TiO2 and ZnO nanoparticles negatively affect wheat growth and soil enzyme activities in agricultural soil. J. Environ. Monit. 2011, 13, 822–828. [Google Scholar] [CrossRef]
  112. You, T.; Liu, D.; Chen, J.; Yang, Z.; Dou, R.; Gao, X.; Wang, L. Effects of metal oxide nanoparticles on soil enzyme activities and bacterial communities in two different soil types. J. Soils Sediments 2018, 18, 211–221. [Google Scholar] [CrossRef]
  113. Chai, H.; Yao, J.; Sun, J.; Zhang, C.; Liu, W.; Zhu, M.; Ceccanti, B. The Effect of Metal Oxide Nanoparticles on Functional Bacteria and Metabolic Profiles in Agricultural Soil. Bull. Environ. Contam. Toxicol. 2015, 94, 490–495. [Google Scholar] [CrossRef]
  114. Asli, S.; Neumann, P.M. Colloidal suspensions of clay or titanium dioxide nanoparticles can inhibit leaf growth and transpiration via physical effects on root water transport. Plant Cell Environ. 2009, 32, 577–584. [Google Scholar] [CrossRef]
  115. Larue, C.; Laurette, J.; Herlin-Boime, N.; Khodja, H.; Fayard, B.; Flank, A.-M.; Brisset, F.; Carriere, M. Accumulation, translocation and impact of TiO2 nanoparticles in wheat (Triticum aestivum spp.): Influence of diameter and crystal phase. Sci. Total Environ. 2012, 431, 197–208. [Google Scholar] [CrossRef]
  116. Song, G.; Gao, Y.; Wu, H.; Hou, W.; Zhang, C.; Ma, H. Physiological effect of anatase TiO2 nanoparticles on Lemna minor. Environ. Toxicol. Chem. 2012, 31, 2147–2152. [Google Scholar] [CrossRef]
  117. Morteza, E.; Moaveni, P.; Farahani, H.A.; Kiyani, M. Study of photosynthetic pigments changes of maize (Zea mays L.) under nano TiO2 spraying at various growth stages. SpringerPlus 2013, 2, 247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Gao, J.; Xu, G.; Qian, H.; Liu, P.; Zhao, P.; Hu, Y. Effects of nano-TiO2 on photosynthetic characteristics of Ulmus elongata seedlings. Environ. Pollut. 2013, 176, 63–70. [Google Scholar] [CrossRef] [PubMed]
  119. Pakrashi, S.; Jain, N.; Dalai, S.; Jayakumar, J.; Chandrasekaran, P.T.; Raichur, A.; Chandrasekaran, N.; Mukherjee, A. In Vivo Genotoxicity Assessment of Titanium Dioxide Nanoparticles by Allium cepa Root Tip Assay at High Exposure Concentrations. PLoS ONE 2014, 9, e87789. [Google Scholar] [CrossRef] [PubMed]
  120. Burke, D.J.; Zhu, S.; Pablico-Lansigan, M.P.; Hewins, C.R.; Samia, A.C.S. Titanium oxide nanoparticle effects on composition of soil microbial communities and plant performance. Biol. Fertil. Soils 2014, 50, 1169–1173. [Google Scholar] [CrossRef]
  121. Yang, F.; Liu, C.; Gao, F.; Su, M.; Wu, X.; Zheng, L.; Hong, F.; Yang, P. The Improvement of Spinach Growth by Nano-anatase TiO2 Treatment Is Related to Nitrogen Photoreduction. Biol. Trace Elem. Res. 2007, 119, 77–88. [Google Scholar] [CrossRef]
  122. Dehkourdi, E.H.; Mosavi, M. Effect of Anatase Nanoparticles (TiO2) on Parsley Seed Germination (Petroselinum crispum) In Vitro. Biol. Trace Elem. Res. 2013, 155, 283–286. [Google Scholar] [CrossRef]
  123. Chavan, S.; Sarangdhar, V.; Nadanathangam, V. Toxicological effects of TiO2 nanoparticles on plant growth promoting soil bacteria. Emerg. Contam. 2020, 6, 87–92. [Google Scholar] [CrossRef]
  124. Simonin, M.; Martins, J.; LE Roux, X.; Uzu, G.; Calas, A.; Richaume, A. Toxicity of TiO2 nanoparticles on soil nitrification at environmentally relevant concentrations: Lack of classical dose–response relationships. Nanotoxicology 2017, 11, 247–255. [Google Scholar] [CrossRef]
  125. Corsi, I.; Bergami, E.; Grassi, G. Behavior and Bio-Interactions of Anthropogenic Particles in Marine Environment for a More Realistic Ecological Risk Assessment. Front. Environ. Sci. 2020, 8, 60. [Google Scholar] [CrossRef]
  126. Auffan, M.; Santaella, C.; Thiéry, A.; Pailles, C.; Rose, J.; Achouak, W.; Thill, A.; Masion, A.; Wiesner, M.R.; Bottero, J.Y. Ecotoxicity of inorganic nanoparticles: From unicellular organisms to invertebrates. In Encyclopedia of Nanotechnology; Springer: Dordrecht, The Netherlands, 2012; pp. 623–636. [Google Scholar] [CrossRef]
  127. Wang, Y.; Zhu, X.; Lao, Y.; Lv, X.; Tao, Y.; Huang, B.; Wang, J.; Zhou, J.; Cai, Z. TiO2 nanoparticles in the marine environment: Physical effects responsible for the toxicity on algae Phaeodactylum tricornutum. Sci. Total Environ. 2016, 565, 818–826. [Google Scholar] [CrossRef]
  128. Miller, R.J.; Bennett, S.; Keller, A.A.; Pease, S.; Lenihan, H.S. TiO2 Nanoparticles Are Phototoxic to Marine Phytoplankton. PLoS ONE 2012, 7, e30321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Della Torre, C.; Balbi, T.; Grassi, G.; Frenzilli, G.; Bernardeschi, M.; Smerilli, A.; Guidi, P.; Canesi, L.; Nigro, M.; Monaci, F.; et al. Titanium dioxide nanoparticles modulate the toxicological response to cadmium in the gills of Mytilus galloprovincialis. J. Hazard. Mater. 2015, 297, 92–100. [Google Scholar] [CrossRef] [PubMed]
  130. Banni, M.; Sforzini, S.; Balbi, T.; Corsi, I.; Viarengo, A.; Canesi, L. Combined effects of n-TiO2 and 2,3,7,8-TCDD in Mytilus galloprovincialis digestive gland: A transcriptomic and immunohistochemical study. Environ. Res. 2016, 145, 135–144. [Google Scholar] [CrossRef]
  131. Gornati, R.; Longo, A.; Rossi, F.; Maisano, M.; Sabatino, G.; Mauceri, A.; Bernardini, G.; Fasulo, S. Effects of titanium dioxide nanoparticle exposure in Mytilus galloprovincialis gills and digestive gland. Nanotoxicology 2016, 10, 807–817. [Google Scholar] [CrossRef] [PubMed]
  132. Jimeno-Romero, A.; Oron, M.; Cajaraville, M.P.; Soto, M.; Marigómez, I. Nanoparticle size and combined toxicity of TiO2 and DSLS (surfactant) contribute to lysosomal responses in digestive cells of mussels exposed to TiO2 nanoparticles. Nanotoxicology 2016, 10, 1168–1176. [Google Scholar] [CrossRef]
  133. Sureda, A.; Capó, X.; Busquets-Cortés, C.; Tejada, S. Acute exposure to sunscreen containing titanium induces an adaptive response and oxidative stress in Mytillus galloprovincialis. Ecotoxicol. Environ. Saf. 2018, 149, 58–63. [Google Scholar] [CrossRef]
  134. Auguste, M.; Lasa, A.; Pallavicini, A.; Gualdi, S.; Vezzulli, L.; Canesi, L. Exposure to TiO2 nanoparticles induces shifts in the microbiota composition of Mytilus galloprovincialis hemolymph. Sci. Total Environ. 2019, 670, 129–137. [Google Scholar] [CrossRef]
  135. Zhu, X.; Zhou, J.; Cai, Z. The toxicity and oxidative stress of TiO2 nanoparticles in marine abalone (Haliotis diversicolor supertexta). Mar. Pollut. Bull. 2011, 63, 334–338. [Google Scholar] [CrossRef]
  136. Shang, Y.; Wu, F.; Wei, S.; Guo, W.; Chen, J.; Huang, W.; Hu, M.; Wang, Y. Specific dynamic action of mussels exposed to TiO2 nanoparticles and seawater acidification. Chemosphere 2020, 241, 125104. [Google Scholar] [CrossRef]
  137. Clemente, Z.; Castro, V.L.; Feitosa, L.; Lima, R.; Jonsson, C.; Maia, A.; Fraceto, L. Fish exposure to nano-TiO2 under different experimental conditions: Methodological aspects for nanoecotoxicology investigations. Sci. Total Environ. 2013, 463–464, 647–656. [Google Scholar] [CrossRef] [PubMed]
  138. Naasz, S.; Altenburger, R.; Kühnel, D. Environmental mixtures of nanomaterials and chemicals: The Trojan-horse phenomenon and its relevance for ecotoxicity. Sci. Total Environ. 2018, 635, 1170–1181. [Google Scholar] [CrossRef] [PubMed]
  139. Qian, W.; Chen, C.C.; Zhou, S.; Huang, Y.; Zhu, X.; Wang, Z.; Cai, Z. TiO2 Nanoparticles in the Marine Environment: Enhancing Bioconcentration, While Limiting Biotransformation of Arsenic in the Mussel Perna viridis. Environ. Sci. Technol. 2020, 54, 12254–12261. [Google Scholar] [CrossRef] [PubMed]
  140. Balbi, T.; Smerilli, A.; Fabbri, R.; Ciacci, C.; Montagna, M.; Grasselli, E.; Brunelli, A.; Pojana, G.; Marcomini, A.; Gallo, G.; et al. Co-exposure to n-TiO2 and Cd2+ results in interactive effects on biomarker responses but not in increased toxicity in the marine bivalve M. galloprovincialis. Sci. Total Environ. 2014, 493, 355–364. [Google Scholar] [CrossRef] [PubMed]
  141. Wang, T.; Huang, X.; Jiang, X.; Hu, M.; Huang, W.; Wang, Y. Differential in vivo hemocyte responses to nano titanium dioxide in mussels: Effects of particle size. Aquat. Toxicol. 2019, 212, 28–36. [Google Scholar] [CrossRef]
  142. Marisa, I.; Matozzo, V.; Martucci, A.; Franceschinis, E.; Brianese, N.; Marin, M.G. Bioaccumulation and effects of titanium dioxide nanoparticles and bulk in the clam Ruditapes philippinarum. Mar. Environ. Res. 2018, 136, 179–189. [Google Scholar] [CrossRef]
  143. Guan, X.; Tang, Y.; Zha, S.; Han, Y.; Shi, W.; Ren, P.; Yan, M.; Pan, Q.; Hu, Y.; Fang, J.; et al. Exogenous Ca2+ mitigates the toxic effects of TiO2 nanoparticles on phagocytosis, cell viability, and apoptosis in haemocytes of a marine bivalve mollusk, Tegillarca granosa. Environ. Pollut. 2019, 252, 1764–1771. [Google Scholar] [CrossRef] [PubMed]
  144. Shi, W.; Han, Y.; Guo, C.; Zhao, X.; Liu, S.; Su, W.; Zha, S.; Wang, Y.; Liu, G. Immunotoxicity of nanoparticle nTiO2 to a commercial marine bivalve species, Tegillarca granosa. Fish Shellfish. Immunol. 2017, 66, 300–306. [Google Scholar] [CrossRef]
  145. Gnatyshyna, L.; Falfushynska, H.; Horyn, O.; Khoma, V.; Martinyuk, V.; Mishchuk, O.; Mishchuk, N.; Stoliar, O. Biochemical responses of freshwater mussel Unio tumidus to titanium oxide nanoparticles, Bisphenol A, and their combination. Ecotoxicology 2019, 28, 923–937. [Google Scholar] [CrossRef]
  146. 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] [PubMed] [Green Version]
  147. Hou, Z.; Zhang, Y.; Deng, K.; Chen, Y.; Li, X.; Deng, X.; Cheng, Z.; Lian, H.; Li, C.; Lin, J. UV-Emitting Upconversion-Based TiO2 Photosensitizing Nanoplatform: Near-Infrared Light Mediated in vivo Photodynamic Therapy via Mitochondria-Involved Apoptosis Pathway. ACS Nano 2015, 9, 2584–2599. [Google Scholar] [CrossRef]
  148. 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] [CrossRef] [PubMed]
  149. Suh, W.H.; Suslick, K.; Stucky, G.D.; Suh, Y.-H. Nanotechnology, nanotoxicology, and neuroscience. Prog. Neurobiol. 2009, 87, 133–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. De Angelis, I.; Barone, F.; Zijno, A.; Bizzarri, L.; Russo, M.T.; Pozzi, R.; Franchini, F.; Giudetti, G.; Uboldi, C.; Ponti, J.; et al. Comparative study of ZnO and TiO2 nanoparticles: Physicochemical characterisation and toxicological effects on human colon carcinoma cells. Nanotoxicology 2012, 7, 1361–1372. [Google Scholar] [CrossRef] [PubMed]
  151. Zhang, A.-P.; Sun, Y.-P. Photocatalytic killing effect of TiO2 nanoparticles on Ls-174-t human colon carcinoma cells. World J. Gastroenterol. 2004, 10, 3191–3193. [Google Scholar] [CrossRef] [PubMed]
  152. Fang-Yin, C.; Lin, Y. Effect of C Content and Calcination Temperatures on the Characteristics of C-doped TiO2 Photocatalyst. NSTI-Nanotech 2010, 1, 491–493. [Google Scholar]
  153. Besha, A.T.; Liu, Y.; Bekele, D.N.; Dong, Z.; Naidu, R.; Gebremariam, G.N. Sustainability and environmental ethics for the application of engineered nanoparticles. Environ. Sci. Policy 2020, 103, 85–98. [Google Scholar] [CrossRef]
  154. Horie, M.; Sugino, S.; Kato, H.; Tabei, Y.; Nakamura, A.; Yoshida, Y. Does photocatalytic activity of TiO2 nanoparticles correspond to photo-cytotoxicity? Cellular uptake of TiO2 nanoparticles is important in their photo-cytotoxicity. Toxicol. Mech. Methods 2016, 26, 1–11. [Google Scholar] [CrossRef]
  155. Zhang, L.W.; Monteiro-Riviere, N.A. Toxicity assessment of six titanium dioxide nanoparticles in human epidermal keratinocytes. Cutan. Ocul. Toxicol. 2019, 38, 66–80. [Google Scholar] [CrossRef]
  156. Rashid, M.M.; Simončič, B.; Tomšič, B. Recent advances in TiO2-functionalized textile surfaces. Surf. Interfaces 2021, 22, 100890. [Google Scholar] [CrossRef]
  157. Miyani, V.A.; Hughes, M.F. Assessment of the in vitro dermal irritation potential of cerium, silver, and titanium nanoparticles in a human skin equivalent model. Cutan. Ocul. Toxicol. 2017, 36, 145–151. [Google Scholar] [CrossRef] [PubMed]
  158. Stan, M.S.; Nica, I.C.; Popa, M.; Chifiriuc, M.C.; Iordache, O.; Dumitrescu, I.; Diamandescu, L.; Dinischiotu, A. Reduced graphene oxide/TiO2 nanocomposites coating of cotton fabrics with antibacterial and self-cleaning properties. J. Ind. Text. 2018, 49, 277–293. [Google Scholar] [CrossRef]
Figure 1. Equilibrium of ROS and antioxidants (AOX) (a) and their disequilibrium causing oxidative stress, either by an excess of ROS (b) or a deficiency of AOX (c).
Figure 1. Equilibrium of ROS and antioxidants (AOX) (a) and their disequilibrium causing oxidative stress, either by an excess of ROS (b) or a deficiency of AOX (c).
Nanomaterials 11 02354 g001
Figure 2. Toxicity mechanism of the reaction of TiO2 NPs with cells. Adapted with permission from [28]. Copyright, 2020 John Wiley & Sons, Inc.
Figure 2. Toxicity mechanism of the reaction of TiO2 NPs with cells. Adapted with permission from [28]. Copyright, 2020 John Wiley & Sons, Inc.
Nanomaterials 11 02354 g002
Figure 3. Influence of TiO2 NPs on the cell fate.
Figure 3. Influence of TiO2 NPs on the cell fate.
Nanomaterials 11 02354 g003
Figure 4. Mechanism of possible genotoxicity of TiO2 NPs in cells. Adapted with permission from [22]. Copyright, 2021 Sringer Nature.
Figure 4. Mechanism of possible genotoxicity of TiO2 NPs in cells. Adapted with permission from [22]. Copyright, 2021 Sringer Nature.
Nanomaterials 11 02354 g004
Figure 5. Mechanism of TiO2 NP-induced immunotoxicity in cells (TLR = Toll-like receptor; MAPK = mitogen-associated protein kinase; JNK = c-jun N-terminal kinase; NRF2 = nuclear factor erythroid 2 factor 2; and SOD = super oxide dismutase). Adapted with permission from [84]. Copyright, 2018 Dove Medical Press Limited.
Figure 5. Mechanism of TiO2 NP-induced immunotoxicity in cells (TLR = Toll-like receptor; MAPK = mitogen-associated protein kinase; JNK = c-jun N-terminal kinase; NRF2 = nuclear factor erythroid 2 factor 2; and SOD = super oxide dismutase). Adapted with permission from [84]. Copyright, 2018 Dove Medical Press Limited.
Nanomaterials 11 02354 g005
Figure 6. Mechanism of neurotoxicity of TiO2 NPs in cells. Adapted with permission from [89]. Copyright, 2020 Elsevier B.V.
Figure 6. Mechanism of neurotoxicity of TiO2 NPs in cells. Adapted with permission from [89]. Copyright, 2020 Elsevier B.V.
Nanomaterials 11 02354 g006
Figure 7. Cytotoxic mechanisms of peripheral blood mononuclear cells (PBMCs) investigated with TiO2 NPs (b). TEM images of apoptosis (a) and necrosis (c) of PBMCs. (COX-2 = cyclooxygenase-2; IL = interleukin; and TNF = tumor necrosis factor). Adapted with permission from [10]. Copyright, 2016 John Wiley & Sons, Inc.
Figure 7. Cytotoxic mechanisms of peripheral blood mononuclear cells (PBMCs) investigated with TiO2 NPs (b). TEM images of apoptosis (a) and necrosis (c) of PBMCs. (COX-2 = cyclooxygenase-2; IL = interleukin; and TNF = tumor necrosis factor). Adapted with permission from [10]. Copyright, 2016 John Wiley & Sons, Inc.
Nanomaterials 11 02354 g007
Figure 8. Uses and dispersion of TiO2 NPs into the environment and interaction with plants. Adapted with permission from [107]. Copyright, 2018 The Royal Society of Chemistry.
Figure 8. Uses and dispersion of TiO2 NPs into the environment and interaction with plants. Adapted with permission from [107]. Copyright, 2018 The Royal Society of Chemistry.
Nanomaterials 11 02354 g008
Figure 9. Environmental and biological effects of the TiO2 NPs. Adapted with permission from [24]. Copyright, 2019 MDPI AG.
Figure 9. Environmental and biological effects of the TiO2 NPs. Adapted with permission from [24]. Copyright, 2019 MDPI AG.
Nanomaterials 11 02354 g009
Figure 10. Biocompatibility of TiO2 NPs (a) and antitumor treatment by TiO2 NP ROS formation (b). Adapted with permission from [147]. Copyright, 2015 American Chemical Society.
Figure 10. Biocompatibility of TiO2 NPs (a) and antitumor treatment by TiO2 NP ROS formation (b). Adapted with permission from [147]. Copyright, 2015 American Chemical Society.
Nanomaterials 11 02354 g010
Figure 11. Risk management strategies for human exposure to TiO2 NPs via different exposure routes.
Figure 11. Risk management strategies for human exposure to TiO2 NPs via different exposure routes.
Nanomaterials 11 02354 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rashid, M.M.; Forte Tavčer, P.; Tomšič, B. Influence of Titanium Dioxide Nanoparticles on Human Health and the Environment. Nanomaterials 2021, 11, 2354. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092354

AMA Style

Rashid MM, Forte Tavčer P, Tomšič B. Influence of Titanium Dioxide Nanoparticles on Human Health and the Environment. Nanomaterials. 2021; 11(9):2354. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092354

Chicago/Turabian Style

Rashid, Mohammad Mamunur, Petra Forte Tavčer, and Brigita Tomšič. 2021. "Influence of Titanium Dioxide Nanoparticles on Human Health and the Environment" Nanomaterials 11, no. 9: 2354. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11092354

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