Skip to main content
Journal of Nucleic Acids logoLink to Journal of Nucleic Acids
. 2010 Sep 21;2010:947859. doi: 10.4061/2010/947859

Current Studies into the Genotoxic Effects of Nanomaterials

Cheng-Teng Ng 1, 2,2, Jasmine J Li 1, 2,2, Boon-Huat Bay 1,*, Lin-Yue Lanry Yung 2,*
PMCID: PMC2946614  PMID: 20936181

Abstract

Nanotechnology has created opportunities for engineers to manufacture superior and more efficient devices and products. Nanomaterials (NMs) are now widely used in consumer products as well as for research applications. However, while the lists of known toxic effects of nanomaterials and nanoparticles (NPs) continue to grow, there is still a vast gap in our knowledge about the genotoxicity of NMs. In this paper, we highlight some NMs of interest and discuss the current in vivo and in vitro studies into genotoxic effects of NMs.

1. Introduction

Materials in the nanoscale are used in many commercial products and industrial practices in the new millennium. They are now increasingly found in plastic wares, clothing, cosmetics, electrical appliances, and even food products. Their applications also extend into the biomedical field and healthcare, particularly in medical imaging and diagnosis, pharmaceuticals, drug delivery, and therapy [1]. The demand for nanomaterials (NMs) in the market in the areas defined above is escalating and estimated to reach sales of up to US$1 trillion by 2015 [2]. The recent burgeoning research interest and development of NMs, nanotechnology, and nanomedicine have led to a vast potential for novel ways of rapid disease diagnosis, treatment, and enhancement of the quality of life. NMs consist of one or more components present in various forms that possess at least one-dimensional structure of diameters in the range of 1 to 100 nm [3]. Engineered NMs, including nanoparticles (NPs) and nanofibres, are generally categorized into four classes, which include carbon-based materials, metal-based materials (quantum dots, nanosilver, and nanogold), dendrimers (nanosized polymers), and composites. Their characteristic features are durability, high conductivity, and reactivity [4].

Many researchers have commented that in actuality, there is still much more to be understood about nanomaterials, especially with regard to the health risks and hazards. The Royal Society and Royal Academy of Engineering first raised this concern in 2004 [5]. This has paved the way for a rapid increase in investigational studies in the toxicity of nanobased materials, in particular, genotoxicity studies of NMs and nanoparticles (NP). A quick search through the Pubmed literature database shows that the bulk of the research articles on NM genotoxicity were published within the past 3 years. As the development of nanotechnological applications continue to grow, it is anticipated that there will be an even greater demand for safety and health and risk assessments studies in the coming years. There have been excellent reviews regarding the methodologies for studying NM-induced toxicity [68].

In this paper, we would like to briefly discuss the methodologies currently available for genotoxic studies and present a survey of the in vitro and in vivo genotoxicological studies of NMs conducted in recent years.

2. Methodologies in Genotoxicity Studies

The study of NM toxicology has its roots in ultrafine particle study, mostly starting out as particulate matter (PM10) and carbon black. The first wave of nanotoxicological studies were assessments of NM cytotoxicity which had been comprehensively outlined by Lewinski et al. [9]. Currently, there is an increasing focus on specific nanotoxic effects, and thus the advent of a subfield called “nanogenotoxicology” [10] which generally refers to the study of toxic effects of NMs on genomic stability and integrity. Common in vitro tests for measuring insults to DNA would centre on single-strand and double-strand breaks, mutations, deletions, chromosomal aberrations, impairment in DNA repair and cell-cycle while tumorigenesis and carcinogenicity are the main focus in in vivo studies. There are as many different kinds of NMs as there are elements and compounds. NMs, depending on the size, shape, elemental materials, and the surface functional groups were observed to have a range of detrimental effects on cells. Compounding the difficulties in toxicological studies, Stone et al. [6] and Landsiedel et al. [7] reiterated that based on existing knowledge, specific NMs probably induce definitive genotoxic effects. Nevertheless, some of the more common tests used in current genotoxic studies are described below.

2.1. In Vitro Techniques and Approaches

2.1.1. Ames Test (Bacterial Reversion Mutation Test)

This test is used to assess the mutagenicity of a chemical compound [11]. Various strains of the histidine dependent bacterium, Salmonella typhimurium, contain mutations in the genes that impair synthesis of histidine required for cell growth. Test substances or compounds are added to different areas on the agar plate, and the bacterium is then plated onto the minimal histidine media. The test compound is deemed to have mutagenic potential if it is able to cause mutations that allow the bacterium to revert back its histidine synthesis ability. The downside of this test is that it is difficult to translate prokaryotic data for eukaryotic genotoxicity testing, and the test is known to generate false positive results [12]. Specific to NM toxicity testing, there are doubts if the Ames test is accurately representative of genotoxicity. Some NMs are not able to cross the bacterial wall, and some kill the test organism as they are bactericidal [7]. Therefore, data should be followed up with other tests after the initial screening.

2.1.2. Comet Assay (Single-Cell Gel Electrophoresis Assay)

This is a simple, inexpensive, and sensitive technique to test for DNA damage. It was first described in 1988 by Singh et al. [13] and has since become the standard test for DNA damage. Cell samples from in vitro or in vivo experiments are first suspended in low melting point agarose and cast onto microscope slides. The cells are lysed so that only the DNA remains, which is then made to undergo-electrophoresis in order to separate the DNA strands based on molecular weight. The DNA strains are subsequently stained with, for example, SYBR green dye and viewed under a fluorescence microscope. Under specific conditions, this test is able to distinguish single-and double-strand breaks in DNA. It is a quick way to assess DNA lesions and extent of genotoxicity in individual eukaryotic cells. However, due to its sensitivity, samples should be handled appropriately to ensure reproducibility of the results.

2.1.3. Micronucleus Test (MN)/Cytokinesis Block Micronucleus Test (CBMN)

This assay is based on scoring the number of micronuclei (MNi) in treated cells [14]. MNi are formed during anaphase from chromosomal fragments or whole chromosomes that are left behind when the nucleus divides. Over time, the assay has evolved to include a pretreatment with cytochalasin-B (Cyt-B), a cytokinesis blocking agent that inhibits cell-division, thereby giving the cells a binucleated appearance. This enables more accurate scoring and the ability to sieve out the dividing cells (where MNi would be found) from the nondividing ones, thereby reducing the incidence of false positives. The CBMN method is now routinely used for measuring chromosome breakage, impairment in DNA repair, chromosome loss, nondisjunction, necrosis, apoptosis and cytostasis.

2.1.4. Hydroxy-Deoxyguanosine (8-OHdG) Analysis

Oxidative stress is considered one of the foremost reasons for DNA damage. Reactive oxygen species (ROS) generated in metabolizing cells could attack DNA base guanine forming the 8-OHdG lesions, which is known to have mutagenic potential and hence used routinely as a biomarker for carcinogenesis [15]. There are a few methods to measure the extent of 8-OHdG lesions and the most established is HPLC (high-performance lipid chromatography), which is often coupled with mass spectrometry, also known as the HPLC-MS/MS. Other methods include performing antibody probes for DNA repair proteins or posttreatment with the enzyme formamidopyrimidine DNA N-glycosylase before quantitative analysis with the comet assay to determine DNA strand breaks [16].

2.2. In Vivo Approaches

There is a need for validation of animal models for studies in NM toxicity. The difficulties lie in devising the correct approach in interpreting the studies and deciding on the parameters that should be considered in examining NM toxicity in in vivo systems. Many investigators have administered NMs through inhalation exposure or orally, ingestion by feed or water supply, and direct instillation or injection into the body. Usually, the subsequent bioavailablity and translocation of the NMs are evaluated, including the organ of entry as well as in other organs where accumulation is more significant. The tests used for assessment of genotoxicity are similar to those used in the in vitro studies.

3. Nanomaterials and their Genotoxic Status

A summary of some of the current genotoxic studies in nanomaterials are shown in Tables 1 and 2, which display the in vivo and in vitro studies, respectively.

Table 1.

Selected in vivo genotoxicity studies on NMs.

Type of NP Size and form Experimental design/genotoxic tests Summary of findings References
C60
fullerenes
spheres Bone marrow micronucleus test on
ICR mice
No in vivo clastogenic ability
of C60 up to 88 mg/kg
Shinohara et al.; 2009 [17]

C60
Single-walled
carbon
nanotubes
(SWCNT)
spheres Oral administration at doses of 0.064 and 0.64 mg/kg of body weight.
8-OHdG analysis
Both NPs were associated
with increase in 8-OHdG in liver and lungs.
No impairment of DNA
repair system
Folkmann et al.; 2009 [18]

SWCNT
Multi-walled
carbon
nanotubes
(MWCNT)
nanotubes Oral administration and urinary samples collected
for Ames test
No urinary mutagenicity Szendi and Varga 2008 [19]

Carbon
black (CB)
C60
SWCNT
AuNP
Cd quantum
dots (QDs)
nanospheres Apo E knockout mice
Timepoints at 3 and 24
hours; NP administered by instillation
Increase in cytokines gene
expression. ApoE −/− mice are sensitive to particle induced inflammation.
DNA damage in order of.
QD>CB>SWCNT>C60, Au
Jacobsen et al.; 2009 [20]

TiO2 anatase/rutile 21 nm TiO2 ingested through
drinking water at
concentrations of
60, 120, 300, 600 μg/mL.
Comet assay
MN test
gamma-H2AX
immunostaining
8-OHdG analysis
Increase in 8-OHdG and
gamma-H2AX foci. indicative
of DNA double-strand
breaks. MN. shows increase
in DNA deletions.
Trouiller et al.; 2009 [21]

Ag 60 nm Oral administration
in Sprague-Dawley rats over
a period of 28 days;
doses at 30, 300 and 1000 mg/kg.
No significant genotoxicity
in bone marrow. (micronucleated
erythrocytes)
Kim et al.; 2008 [22]

Silica amorphous
37 and 83 nm
Inhalation study where mice were exposed to 3.7 × 107 and 1.8 × 108 particles/cm3 No significant pulmonary,
inflammatory, genotoxic or adverse lung
histopathological effects
Sayes et al.; 2010 [23]

Table 2.

Selected in vitro genotoxicity studies on NMs.

Type of NP Size and form Experimental design/genotoxic tests Summary of findings References
Carbons

C60 0.92 m2/g surface area Ames test No mutagenic response,
and no incidence of chromosomal aberration
Shinohara et al.; 2009 [17]

C60 polyhydroxylated CHO-K1 cells
chromosome aberration assay
CBMN test
No genotoxicity at all doses (11–221 μM) Mrdanović et al., 2009 [24]

C60 nanospheres Mouse primary embryo fibroblasts
Dihydrorhodamine 123 radical probe
Increased mutation yield and induces kilo-based pair deletion mutations in transgenic mouse cells.
Dose-dependent formation of ONOO
Xu et al.; 2009 [25]

SWCNT
-MWSCNT
nanotubes Human lymphocytes in culture CBMN test
Sister Chromatid Exchange (SCE) assay
No genotoxicity effects but SWCNT induces mitotic inhibition Szendi and Varga; 2008 [19]

MWSCNT agglomerates V79 cells treated for 18 h and 30 h at 2.5, 5 and 10 μg/mL.
Chromosome aberration test Ames test
No mutagenic or clastogenic effects Wirnitzer et al., 2009 [26]

MWSCNT nanotubes Ames test on Salmonella typhimurium TA 98 and TA 100 strains, and on Escherichia coli WP2uvrA strain, in presence and in absence of the metabolic activation system S9 No mutagenic effects Di Sotto et al.; 2009, [27]

C60
SWCNT
Carbon black
(CB)
0.7 nm (C60) 0.9–1.7 nm (SWCNT) 14 nm (CB) FE1-muta trademark mouse lung epithelial cell line comet assay
FE1-MML mutagenicity analysis
c11 mutation analysis
No cell death. Slower proliferation and cell-cycle arrest at G1 with SWCNT.
Mutant frequency unaffected by 576 h exposure
Jacobsen et al., 2008 [28]

Metals

Alumina
(Al2O3)
Cobalt
Chromium
alloy (CoCr)
bare Human primary fibroblasts over 5 days
CBMN assay
gamma-H2AX immuno staining
cytogenetic analysis (FISH)
At 24 h, Al2O3 increase micronucleus binucleated cells, chromosomal loss, gain, and polyploidy.
At 24 h, CoCr induce dose-dependent increase in micronucleus binucleated cells, chromosomal loss, gain, deletions, and polyploidy.
Tsaousi et al.; 2010, [29]

Co 20 nm
500 nm
Balb/3T3 cells at 1–100 μM dose concentrations.
CBMN test
Comet assay
Significant results for CBMN and comet assay but no dose-dependency.
Increase of type III foci
Ponti et al.; 2009 [30]

Co 100–500 nm Peripheral blood leulocytes at 24, 48 h timepoints in 10−5M and 10−4M dose concentrations
CBMN test
Comet assay
Induces DNA damage
Genotoxic effects
modulated by donor
characteristics and/or Co2+ release.
Colognato et al.; 2008 [31]

Al2O3
  TiO2
nanoparticles CHO-K1 cells
Micronucleus (MN) test
Sister chromatid exchange (SCE)
MN frequencies increase at 0.5 and 1 μg/mL TiO2 and 0.5–10 μg/mL AL2O3.
SCE higher at 1–5 μg/mL TiO2 treatment, and at 1–25 μg/mL Al2O3
Di Virgillio et al.; 2010 [32]

TiO2 rutile/anatase
fine rutile
Human bronchial epithelial cells (BEAS 2B) with 1–100 μg/cm2 at 24, 48, and 72 h.
Comet assay
MN test
Both induce DNA damage at all treatment times. Only nanosize rutile increase frequency of MN cells at 10, 60 μg/cm2, 72 h. Falck et al.; 2009 [33]

TiO2 with p,p′-DDT Human embryo L-02 hepatocyte 0.01, 0.1, 1 μg/mL treatment concentrations
Flow cytometry with DCFH-DA probe
8OHdG analysis
Comet assay
MN test
TiO2 enhances
photocatalysis. Increases oxidative stress, DNA adducts, DNA strand breaks, and chromosome damage
Shi et al.; 2010 [34]

TiO2 2–30 nm (mean at 15 nm) NIH3T3 human fibroblasts HFW cells
Short-term treatment at 24, 48 and 72 h.
Long-term treatment, cell passage every 3 days with NP media.
Flow cytometry with H2DCFDA probes
Short-term increased cell survival and growth. Long-term G2/M delay and slower cell-division with aberrant multipolar spreads. Overall disturbance in cell-cycle progression, duplicate
genome segregation, and
chromosomal instability
Huang et al.; 2009 [35]
Cell-cycle analysis
Cell-division analysis
Confocal microscopy

TiO2
Fe2O3
anatase <100 nm
<100 nm
Human lung fibroblasts IMR-90 and BEAS-2B cells
Electron paramagnetic resonance (EPR)
8-OHdG analysis
TiO2 treatment showed no DNA breakage, DNA adduct nor free radical generation.  Fe2O3 had significant DNA damage after 24 h in IMR-90 cells Bhattacharya et al.; 2009 [36]

TiO2 nanoparticles
rutile
anatase
Mouse primary embryo fibroblasts
Dihydrorhodamine 123 radical probe
Increased mutation yield and induces kilo-based pair deletion mutations in transgenic mouse cells. Dose-dependent formation of ONOO Xu et al.; 2009 [25]

TiO2 100 nm Human lymphoblastoid cells. Treatment with 26, 65, 130 μg/mL at 6, 24, 48 h.
CBMN test
Comet assay
Hypoxanthine-guanine
phosphoribosyltransferase (HPRT) gene mutation assay
130  μg/mL treatment increases MNBC frequency 2-3 folds and 2.5 fold in mutation frequency.
65 μg/mL treatment induce 5 fold increase in comet tail moments
Wang et al.; 2007 [37]

ZnO nanospheres Human epidermal cell line (A431)
Treatment at 0.8, 0.008g/mL Comet assay
Significant DNA damage in comet assay. Induces oxidative stress Sharma et al.; 2009 [38]

Ag 30 nm, nanospheres Medaka fish cell lines
Treatment at 0.05, 0.1, 0.3  μg/cm2
Chromosomal aberration and aneuploidy Wise et al.; 2010 [39]

Ag 6–20 nmstarch coated IMR-90 and human glioblastoma cells U251
Comet assay
CBMN
Annexin V propidium iodide
staining
DNA aberrations more prominent in cancer cells with more chromosomal aberrations. Asharani et al.; 2009 [40]

Ag 25 nm
polysaccharide
surface functionalized and
uncoated nanospheres
Mouse embryonic stem cells and embryonic fibroblasts Immuno blot Immunoflorescence Upregulation of p53, Rad 51 and phosphorylated H2AX protein expression. Coated AgNP show more severe damage than uncoated AgNP Ahamed et al.; 2008 [41]

Au 20 nm
Serum coated
Human fetal lung fibroblasts cells (MRC-5) treated with nAu at 0,
0.5 and 1 nm concentrations.
8-OHdG analysis
Significant DNA damage in 1 nm treatment compared to control. Li et al.; 2008 [42]

Platinum (Pt NP) 5–8 nm capped with poly-vinyl alcohol Human cell line p53 activation, p21 downregulation. Increase of DNA damage, arrest at cell-cycle S phase and apoptosis Asharani et al.; 2010 [43]

Other Nanomaterials

Nanoceria (CeO2) nanoparticles Human lens epithelial cells at 5, 10 μg/mL concentrations
SCE
Comet assay (alkaline)
No DNA damage nor SCE Pierscionek et al.; 2010 [44]

Polymer NP lyophilized PELGE and PLGAnp CHO cells
MN test
SCE
No significant difference in MN assay and no cell-cycle delay. SCE found to be higher in 5 kinds of PELGE-NP than in negative controls He et al.; 2009 [45]

3.1. Carbon Fullerenes

Carbon fullerenes, which are ultrafine particulate matter, are one of the most ubiquitous NMs found [46]. They are generally present in polluted air as they are often released in soot resulting from the process of fuel combustion. Engineered carbon fullerenes are stable, soccer ball-like carbon atoms with hexagonal and pentagonal shapes. The most notable fullerene would be C60, a highly reactive biomolecules that has the ability to cross blood brain barrier (BBB) [47]. C60 fullerene is highly used in industry as catalysts, reactive oxygen species scavengers [48] and tools in drug delivery systems [49].

Since the early 1990s, there have been concerns about the potential dermal and inhalation effects of fullerenes due to their strong oxidizing and phototoxic properties [50]. In vitro experiments have shown C60 to be generally noncytotoxic with no mutagenic response [17, 24] in Chinese hamster ovary (CHO-K1) cells and mouse lung epithelial cells [28] using the Ames test and CBMN tests, respectively. Another report has found that C60 treatment also increases formamidopyrimidine [fapy]-DNA glycosylase (FPG) sensitive sites, accounting for short-term DNA strand damage. Xu et al. observed that C60 induced an increase in mutation yield in primary mouse embryo fibroblast cells and dose-dependent formation of free radical ONOO [25] using dihydrorohodamine radical probes. However, in the in vivo setting, C60 treatment was found to be associated with increased DNA damage 8-hydroxydeoxyguanosine (8-OHdG) in mouse lung and liver [18]. Not surprisingly, inflammatory cytokines such as the interleukins and MIP and MCP genes were found to be upregulated although C60 extent of damage was lower as compared to other NMs.

3.2. Carbon Nanotubes

Carbon nanotubes are the byproducts of combustion, which are commonly present in air pollution and soot. Engineered carbon nanotubes can also come in a variety of shapes and conformations, with the most common being the single-walled carbon nanotubes (SWCNTs) and the multiwalled carbon nanotubes (MWCNTs). They are also found in a wide range of applications in the industry as composites, polymers, as well as in the biomedical and pharmaceutical fields. Great physical strength, flexibility, electrical conductivity, insolubility and nonbiodegradability are among the valued properties of carbon nanotubes [51]. On the other hand, it has been postulated that these nanotubes could possess health hazards upon inhalation as their durability, biopersistency, and long and thin shape resembling asbestos fibers [52]. In addition, trace contaminations with iron and nickel have been reported to be the major cause of toxicity in carbon nanotubes [53].

There is a scarcity of information regarding SWCNTs and genotoxicity. SWCNTs have been reported to induce slower proliferation rate and cell-cycle arrest at G1 phase in mice lung epithelial cells [28] and mitotic inhibition in human lymphocyte cultures [19]. In in vivo experiments, oral administration of SWCNTs in mice is found to be associated with increase in 8-OHdG levels in liver and lung [18]. SWCNTs, compared to carbon black, only causes moderate inflammation in ApoE knockout mice [20]. However, agglomerates of MWCNTs were found to possess neither clastogenic nor mutagenic effects [19, 26, 27] when put under the Ames test and chromosome aberration test.

3.3. Titanium Dioxide and Zinc Oxide Nanoparticles (TiO2 and ZnO NPs)

TiO2 and ZnO NPs, which have the properties of high refractive index and brightness, are regularly used as whitening pigments or reflective optical coats [54]. These specific properties lead to the application in commercial products such as paint and whitening agents in food products [55]. Nanoparticulate suspensions of ZnO and TiO2 also appear transparent in air and liquid under visible light. As such, ultrafine TiO2 is also extensively used in cosmetics, skin care, and sunscreen products, as their application does not leave unsightly white residue on skin unlike bulk TiO2 [56]. ZnO is quite well known to be cytotoxic to cells in culture [57], while the toxicity of TiO2NP is rapidly gaining attention due to the increased use and applications in many accessible medical and cosmetic products. TiO2NP comes in two common shapes, namely, the rutile and anatase forms. Although both are found to be genotoxic, one study showed that the anatase form induced greater DNA damage in human bronchial epithelium [33]. TiO2NP could also increase cell sensitivity to phototoxicity [34], as well as induce more DNA adducts, strand breaks, base-pair mutations and chromosomal damage [21, 25, 37]. Interestingly, Huang et al. reported that while long-term exposure to TiO2NP slowed down cell-division and induced aberrant multipolar spreads, chromatin alignment, and segregation, short-term exposure increased cell survival and growth and number of multinucleated cells [35]. Another group of investigators did not observe DNA breakage under TiO2NP treatment but found positive DNA adduct formation and free radical generation [36].

Although ZnONPs are probably the less studied of the two, there is also evidence to suggest that they may also induce significant DNA damage through oxidative stress, albeit with less obvious effects than in TiO2NPs [38].

3.4. Aluminium Oxide Nanoparticles (Al2O3NPs)

Al2O3NP, or alumina NP as it is commonly known, belongs to a class of materials known as nanoceramics. It is widely used in industrial and medical product such as orthopaedic parts and composite repellants. However, the toxic and genotoxic effects of Al2O3NP are not well known, and there are very few research studies on the toxicity of this material. Thus far, Al2O3NPs were found to significantly increase micronucleus frequencies, chromosomal loss, and gain mutations as well as polyploidy but no sister chromatid exchanges were found to take place [29, 32].

3.5. Cobalt and Cobalt-Chromium Nanoparticles (CoNPs and CoCrNPs)

Cobalt and its alloy are commonly used in hip joint replacements and other orthopedic joint replacements. Unfortunately, the friction produced in movement of the replacement joints generate NPs of the metal which could reach out and affect the surrounding tissue and even lymphocytes, thereby lead to some concerns regarding the genotoxicity observed from clinical studies [58]. Hence, much interest was generated to study the effects of these wear particles and a significant amount of research into Co and CoCr NPs are centered around these issues. The results, although not surprising, are generally aligned to positive indications of genotoxicity. Analysis of peripheral blood leukocytes of patients with cobalt alloy joint replacements showed positive DNA damage in comet assays [31]. However, it was also suggested that these results could possibly be modulated by donor characteristics and may be due to Co2+ release instead of CoNPs per se. Recent studies show that by 24 h, CoCrNPs induced a dose-dependant increase in micronucleus containing cells as well as chromosomal loss, gains, deletions, and polyploidy [29]. In a separate study with CoNPs on Balb/3T3 cells, there were significant results in micronuclei and comet assay for NP induced DNA damage but the results were not dose-dependent [30].

3.6. Quantum Dots

Quantum dots are crystalline semi-conducting NPs. They are comprised of a metalloid crystalline core and a “cap” or “shell” that shields the core or renders the dots biologically compatible [4]. The metalloid crystalline core is normally made up of heavy metals like cadmium and lead or sometimes from other semiconductor, noble, and transition metals. These are also quantum dots that are coated with materials such as polyethylene glycol, zinc sulphide, or polyacrate [59] Quantum dots are used in composites, paints, inks, solar cells, and optoelectronics [4]. Due to their bright fluorescence, narrow emission, broad UV excitation, and photostability, they have been used as alternative fluorescent dyes for labeling cell structure in vitro and for fluorescence imaging in vivo [60].

They are considered one of the most toxic of substances and there are many studies showing the acute cytotoxicity of quantum dots [61]. The cadmium and lead metals themselves are considered potent human carcinogens. Cadmium induces DNA damage and mutation through ROS production and inhibition of DNA repair and methylation [62]. It also incites disruption of E-cadherin cell-to-cell adhesion which could lead to tumor formation. Lead and its compounds are listed under group 2B of possible human carcinogens in IARC reports [63], as they are found to induce lipid peroxidation and inhibit enzymes and antioxidants thereby putting the cell under an environment of oxidative stress [64]. However, few have ventured into exploring the genotoxicity of such QDs. One experiment with Apo E knockout mice showed that such mice were more sensitized to QD-induced inflammation, upregulating gene expressions of cytokines, IL-6, Mip 2 and Mcp signaling molecules [20].

3.7. Silver and Gold Nanoparticles (AgNPs and AuNPs)

AgNPs and AuNPs are the most marketable NPs and widely used in consumer products. AgNPs are particulary known for their antimicrobial qualities, while AuNPs are used in bioimaging and diagnosis applications. They are also easily synthesized from their salt compounds and are convenient to handle, which also makes them another popular choice of NMs to work with. What is of concern is that several studies have found AgNPs to be toxic in aquatic animals [65] and AuNPs to possess some degree of toxicity in vitro [66]. Many researchers have focused on AgNPs because of the acute toxicity shown in vitro experiments. AgNPs were found to induce DNA damage in human glioblastoma cells as well as chromosomal aberrations in human fibroblast cells [40]. Other genotoxic reactions include upregulation of p53 and DNA repair protein Rad51 observed in mouse embryonic stem cells and fibroblasts [41]. In the same study, AgNP when functionalized with polysaccharide on its surface was more DNA damaging than uncoated AgNPs. In long-term rodent studies, oral administration of high-dose AgNPs for 28 days resulted in liver damage but no significant genotoxicity in erythrocytes and bone marrow [22]. A number of studies have also shown that AgNP treatment induced DNA damaging effects on aquatic and plant cells with impairment of cell-division [39, 65]. Although less dramatic than AgNPs, AuNPs are also able to induce DNA damage in the form of single-strand lesions in human lung fibroblasts [42].

3.8. Other Nanoparticles

There are a few research groups working with new types of NPs. The rare earth metal cerium oxide NPs (nanoceria) is one example. Researchers have found nanaoceria to be a radical scavenger with antiinflammatory effects [67] which causes no DNA damage [44]. They are currently being developed for application in human lens epithelium. Although this is a promising NM for future applications, it has also been reported that nanoceria exerts differential growth in soybean seedlings [68]. Silica NPs, or often known as mesoporous silica, are also popular materials for development of drug delivery and cell-imaging systems [69, 70]. There are few genotoxicity studies on silica NPs but a notable one by Sayes et al. [23] has shown that there are no significant inflammatory or genotoxic effects in mouse lungs on short-term exposure. Metal NPs such as platinum NPs (PtNPs) and iron oxide Fe2O3NPs are also popular alternatives. There is one report on PtNP toxicity which showed an increase in DNA damage concurrent with p53, p21 downregulation, and cell-cycle arrest at the S-phase [43]. Fe and Fe2O3NPs are also known to be toxic and can cause significant DNA damage [36].

Other particles of note are the nanopolymers. Although there is a wide variety of such nanopolymers, they are generally known as a family of compounds that consists of chain units, which could be fashioned into nano-sized particles. These are also largely being developed for use in drug delivery [71]. Current genotoxicity studies suggest that some of these nanopolymers show antiinflammatory properties and also non or limited DNA damage [45, 72]. However, a recent report has implicated long-term nanopolymer exposure to pulmonary fibrosis and granuloma formation, resulting in two fatal deaths [73]. This case cannot be taken in isolation and others have raised the concern that the workplace condition as well as health or other pre-dispositions of the workers involved should be considered [74].

4. NMs and Carcinogenesis

While it has been shown in many in vitro experiments that NMs are able to induce DNA damage and some form of mutagenesis, there is still a lack of evidence for tumorigenicity of NPs. Of note, in vivo studies involving MWCNT has demonstrated formation of mesotheliomas in rodents in works by Takagi and colleagues [75] and Sakamoto et al. [76]. Wide spread deposition of MWCNTs were observed in the peritoneal cavity where the nanotubes were injected. In the study by Sakamoto et al., they have even found mesotheliomas in the peritoneal cavity away from the original site of injection, suggesting that MWCNTs may easily translocate and also exert effects away from organ of exposure. Both studies emphasized on the persistency, size and shape on the carcinogenic potential of MWCNTs. While such studies may provide some insight into the outcome of NM toxicity, one must take into account the differences in how the nanotubes were prepared as well as the experimental design. Muller et al. conducted similar tests on MWCNTs but reported no carcinogenicity after a 2 year period of exposure [77]. They speculate that tumor formation could be dependent on size and length of the nanotubes administered and the p53 knockout mice used in the Takagi study produced a more sensitive carcinogenic reaction. However, NMs can induce oxidative stress and trigger inflammatory responses, which could form the starting point for carcinogenesis to occur. NMs that are highly reactive are also more likely to absorb endogenous substances, react with proteins and enzymes, trigger cytokine release. This would mediate inflammatory responses and potentially initiate a series of toxic responses far from the initial site of deposition [78, 79]. C60 fullerene, for example, was reported to cause photo-induced DNA damage by interacting with biological reducing agents such as NADH to cleave supercoiled DNA [80]. Similarly, exposure to carbon nanotubes in atmospheric air pollution has been associated with adverse cardiovascular effects by causing aortic DNA damage, platelet aggregation and enhances vascular thrombosis through inflammatory events [81].

Biopersistence of NMs pose a certain degree of adverse health effect. For instance, when the clearance rate is slower than the accumulative rate, the NMs will remain in the lungs; and those containing mutagenic substances will increase the risk of developing cancer. To address this concern, Sera et al. conducted a mutagenicity test using 3 different strains of Salmonella and found C60 Fullerene to exert mutagenic activity due to the oxidized phospholipids in rat liver microsoms [82].

There are also certain shortcomings in the current research field. The short-term nature of toxicology tests in the treatment period for NMs generally lasts only up to three days, which implies that testing is limited to acute toxicity. In vitro and in vivo genotoxicity testing will have to be conducted for longer periods to observe if there are long-term effects of NMs such as tumour formation and carcinogenesis. Treatment intervals will have to go beyond days to weeks or even months in animal studies. It will also be useful to look at the clearance of NMs from the body and to study if there is a preference for accumulation in certain organs and any effect from biopersistence of such NMs.

On the public front, safety measures have been implemented to safeguard the public health. The International Agency for Research on Cancer (IARC) recently classified TiO2 as a potential Group 2B human carcinogen. This decision was made on the experimental animal carcinogenicity data [83]. There had been four previous epidemiological studies conducted among the male production workers at TiO2 industry from Western Europe and North America. After comparing the risk for lung and kidney cancer with the general population, they concluded that these data were not supportive enough to conclude the association between occupational exposure of TiO2 and cancer risk. Hence, data collected were inadequate in classifying TiO2 as potential carcinogen. However, there was sufficient animal carcinogenicity data that provided evidence of TiO2-induced carcinogenicity. Several TiO2 exposure routes were chosen for experimental animal studies. These include oral, inhalation, intratracheal, subcutaneous injection, and intraperitoneal administration. Researchers observed an increase in tumor incidence in these experimental animals upon TiO2 exposure. After considering other relevant data such as clearance kinetics of TiO2 and micronucleus formation, a conclusion that TiO2 possess possible carcinogenicity to human was made.

5. Conclusion

The field of nanotoxicology, besides investigations on the adverse effects of NMs, also include continuous monitoring and risk assessment of NMs. Despite the many nanotoxicological studies that are ongoing, there are questions that need to be answered and addressed. There is difficulty in interpreting data in view of variable parameters utilized in the study, for example, the sizes of the NMs and its composing materials. The most critical research gap is the lack of studies on real-time NM exposure. Moreover, there is a need for long-term nanomaterials exposure assessment for studies on tumourigenesis. At the industry level, close monitoring and followup on the levels of emissions from NM production industries are essential in protecting public health and our environment. However, there still exists a lack of appropriate epidemiological studies and equipment for accurate collection of data in assessing the real risk of NM exposure in the workplace. Despite the promising applications of NMs, there are still doubts regarding their safety. There is some certainty that NMs do pose a certain degree of health risk that would require further investigation. A proper guideline on NM usage is imperative to ensure the safety of NMs for consumer usage and environment.

Acknowledgments

The paper was supported by the Singapore Ministry of Education Academic Research Fund Tier 2 via grant MOE2008-T2-1-046 and the National University of Singapore (NUS) Environmental Research Institute (NERI) via Grant no. R706000002646.

References

  • 1.Nowack B, Bucheli TD. Occurrence, behavior and effects of nanoparticles in the environment. Environmental Pollution. 2007;150(1):5–22. doi: 10.1016/j.envpol.2007.06.006. [DOI] [PubMed] [Google Scholar]
  • 2.Roco MC. Environmentally responsible development of nanotechnology. Environmental Science and Technology. 2005;39(5):106A–112A. doi: 10.1021/es053199u. [DOI] [PubMed] [Google Scholar]
  • 3.Warheit DB, Sayes CM, Reed KL, Swain KA. Health effects related to nanoparticle exposures: environmental, health and safety considerations for assessing hazards and risks. Pharmacology and Therapeutics. 2008;120(1):35–42. doi: 10.1016/j.pharmthera.2008.07.001. [DOI] [PubMed] [Google Scholar]
  • 4.Powell MC, Kanarek MS. Nanomaterial health effects—part 1: background and current knowledge. Wisconsin Medical Journal. 2006;105(2):16–20. [PubMed] [Google Scholar]
  • 5.The Royal Society and the Royal Academy of Engineering. Environmental applications and impacts of nanotechnology: summary of evidence presented to nanotechnology working group. 2003.
  • 6.Stone V, Johnston H, Schins RPF. Development of in vitro systems for nanotoxicology: methodological considerations in vitro methods for nanotoxicology Vicki Stone et al. Critical Reviews in Toxicology. 2009;39(7):613–626. doi: 10.1080/10408440903120975. [DOI] [PubMed] [Google Scholar]
  • 7.Landsiedel R, Kapp MD, Schulz M, Wiench K, Oesch F. Genotoxicity investigations on nanomaterials: methods, preparation and characterization of test material, potential artifacts and limitations-many questions, some answers. Mutation Research. 2009;681(2-3):241–258. doi: 10.1016/j.mrrev.2008.10.002. [DOI] [PubMed] [Google Scholar]
  • 8.Singh N, Manshian B, Jenkins GJS, et al. NanoGenotoxicology: the DNA damaging potential of engineered nanomaterials. Biomaterials. 2009;30(23-24):3891–3914. doi: 10.1016/j.biomaterials.2009.04.009. [DOI] [PubMed] [Google Scholar]
  • 9.Lewinski N, Colvin V, Drezek R. Cytotoxicity of nanopartides. Small. 2008;4(1):26–49. doi: 10.1002/smll.200700595. [DOI] [PubMed] [Google Scholar]
  • 10.Donaldson K, Stone V, Tran CL, Kreyling W, Borm PJA. Nanotoxicology. Occupational and Environmental Medicine. 2004;61(9):727–728. doi: 10.1136/oem.2004.013243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Mortelmans K, Zeiger E. The Ames Salmonella/microsome mutagenicity assay. Mutation Research. 2000;455(1-2):29–60. doi: 10.1016/s0027-5107(00)00064-6. [DOI] [PubMed] [Google Scholar]
  • 12.Khandoudi N, Porte P, Chtourou S, Nesslany F, Marzin D, Le Curieux F. The presence of arginine may be a source of false positive results in the Ames test. Mutation Research. 2009;679(1-2):65–71. doi: 10.1016/j.mrgentox.2009.03.010. [DOI] [PubMed] [Google Scholar]
  • 13.Singh NP, McCoy MT, Tice RR, Schneider EL. A simple technique for quantitation of low levels of DNA damage in individual cells. Experimental Cell Research. 1988;175(1):184–191. doi: 10.1016/0014-4827(88)90265-0. [DOI] [PubMed] [Google Scholar]
  • 14.Fenech M. Cytokinesis-block micronucleus cytome assay. Nature Protocols. 2007;2(5):1084–1104. doi: 10.1038/nprot.2007.77. [DOI] [PubMed] [Google Scholar]
  • 15.Karihtala P, Soini Y, Vaskivuo L, Bloigu R, Puistola U. DNA adduct 8-hydroxydeoxyguanosine, a novel putative marker of prognostic significance in ovarian carcinoma. International Journal of Gynecological Cancer. 2009;19(6):1047–1051. doi: 10.1111/IGC.0b013e3181ad0f0d. [DOI] [PubMed] [Google Scholar]
  • 16.Valavanidis A, Vlachogianni T, Fiotakis C. 8-hydroxy-2’ -deoxyguanosine (8-OHdG): a critical biomarker of oxidative stress and carcinogenesis. Journal of Environmental Science and Health. Part C. 2009;27(2):120–139. doi: 10.1080/10590500902885684. [DOI] [PubMed] [Google Scholar]
  • 17.Shinohara N, Matsumoto K, Endoh S, Maru J, Nakanishi J. In vitro and in vivo genotoxicity tests on fullerene C60 nanoparticles. Toxicology Letters. 2009;191(2-3):289–296. doi: 10.1016/j.toxlet.2009.09.012. [DOI] [PubMed] [Google Scholar]
  • 18.Folkmann JK, Risom L, Jacobsen NR, Wallin H, Loft S, Møller P. Oxidatively damaged DNA in rats exposed by oral gavage to C60 fullerenes and single-walled carbon nanotubes. Environmental Health Perspectives. 2009;117(5):703–708. doi: 10.1289/ehp.11922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Szendi K, Varga C. Lack of genotoxicity of carbon nanotubes in a pilot study. Anticancer Research. 2008;28(1):349–352. [PubMed] [Google Scholar]
  • 20.Jacobsen NR, Møller P, Jensen KA, et al. Lung inflammation and genotoxicity following pulmonary exposure to nanoparticles in ApoE-/-mice. Particle and Fibre Toxicology. 2009;6, article 2 doi: 10.1186/1743-8977-6-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Trouiller B, Reliene R, Westbrook A, Solaimani P, Schiestl RH. Titanium dioxide nanoparticles induce DNA damage and genetic instability in vivo in mice. Cancer Research. 2009;69(22):8784–8789. doi: 10.1158/0008-5472.CAN-09-2496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Kim YS, Kim JS, Cho HS, et al. Twenty-eight-day oral toxicity, genotoxicity, and gender-related tissue distribution of silver nanoparticles in Sprague-Dawley rats. Inhalation Toxicology. 2008;20(6):575–583. doi: 10.1080/08958370701874663. [DOI] [PubMed] [Google Scholar]
  • 23.Sayes CM, Reed KL, Glover KP, et al. Changing the dose metric for inhalation toxicity studies: short-term study in rats with engineered aerosolized amorphous silica nanoparticles. Inhalation Toxicology. 2010;22(4):348–354. doi: 10.3109/08958370903359992. [DOI] [PubMed] [Google Scholar]
  • 24.Mrdanović J, Šolajić S, Bogdanović V, Stankov K, Bogdanović G, Djordjevic A. Effects of fullerenol C60(OH)24 on the frequency of micronuclei and chromosome aberrations in CHO-K1 cells. Mutation Research. 2009;680(1-2):25–30. doi: 10.1016/j.mrgentox.2009.08.008. [DOI] [PubMed] [Google Scholar]
  • 25.Xu A, Chai Y, Nohmi T, Hei TK. Genotoxic responses to titanium dioxide nanoparticles and fullerene in gpt delta transgenic MEF cells. Particle and Fibre Toxicology. 2009;6, article 3 doi: 10.1186/1743-8977-6-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Wirnitzer U, Herbold B, Voetz M, Ragot J. Studies on the in vitro genotoxicity of baytubes®, agglomerates of engineered multi-walled carbon-nanotubes (MWCNT) Toxicology Letters. 2009;186(3):160–165. doi: 10.1016/j.toxlet.2008.11.024. [DOI] [PubMed] [Google Scholar]
  • 27.Di Sotto A, Chiaretti M, Carru GA, Bellucci S, Mazzanti G. Multi-walled carbon nanotubes: lack of mutagenic activity in the bacterial reverse mutation assay. Toxicology Letters. 2009;184(3):192–197. doi: 10.1016/j.toxlet.2008.11.007. [DOI] [PubMed] [Google Scholar]
  • 28.Jacobsen NR, Pojana G, White P, et al. Genotoxicity, cytotoxicity, and reactive oxygen species induced by single-walled carbon nanotubes and C60 fullerenes in the FE1-Muta™mouse lung epithelial cells. Environmental and Molecular Mutagenesis. 2008;49(6):476–487. doi: 10.1002/em.20406. [DOI] [PubMed] [Google Scholar]
  • 29.Tsaousi A, Jones E, Case CP. The in vitro genotoxicity of orthopaedic ceramic (Al2O3 ) and metal (CoCr alloy) particles. Mutation Research. 2010;697(1-2):1–9. doi: 10.1016/j.mrgentox.2010.01.012. [DOI] [PubMed] [Google Scholar]
  • 30.Ponti J, Sabbioni E, Munaro B, et al. Genotoxicity and morphological transformation induced by cobalt nanoparticles and cobalt chloride: an in vitro study in Balb/3T3 mouse fibroblasts. Mutagenesis. 2009;24(5):439–445. doi: 10.1093/mutage/gep027. [DOI] [PubMed] [Google Scholar]
  • 31.Colognato R, Bonelli A, Ponti J, et al. Comparative genotoxicity of cobalt nanoparticles and ions on human peripheral leukocytes in vitro. Mutagenesis. 2008;23(5):377–382. doi: 10.1093/mutage/gen024. [DOI] [PubMed] [Google Scholar]
  • 32.Di Virgilio AL, Reigosa M, Arnal PM, Fernández Lorenzo de Mele M. Comparative study of the cytotoxic and genotoxic effects of titanium oxide and aluminium oxide nanoparticles in Chinese hamster ovary (CHO-K1) cells. Journal of Hazardous Materials. 2010;177(1–3):711–718. doi: 10.1016/j.jhazmat.2009.12.089. [DOI] [PubMed] [Google Scholar]
  • 33.Falck GCM, Lindberg HK, Suhonen S, et al. Genotoxic effects of nanosized and fine TiO2. Human and Experimental Toxicology. 2009;28(6-7):339–352. doi: 10.1177/0960327109105163. [DOI] [PubMed] [Google Scholar]
  • 34.Shi Y, Zhang J-H, Jiang M, Zhu L-H, Tan H-Q, Lu B. Synergistic genotoxicity caused by low concentration of titanium dioxide nanoparticles and p,p’-DDT in human hepatocytes. Environmental and Molecular Mutagenesis. 2010;51(3):192–204. doi: 10.1002/em.20527. [DOI] [PubMed] [Google Scholar]
  • 35.Huang S, Chueh PJ, Lin Y-W, Shih T-S, Chuang S-M. Disturbed mitotic progression and genome segregation are involved in cell transformation mediated by nano-TiO2 long-term exposure. Toxicology and Applied Pharmacology. 2009;241(2):182–194. doi: 10.1016/j.taap.2009.08.013. [DOI] [PubMed] [Google Scholar]
  • 36.Bhattacharya K, Davoren M, Boertz J, Schins RPF, Hoffmann E, Dopp E. Titanium dioxide nanoparticles induce oxidative stress and DNA-adduct formation but not DNA-breakage in human lung cells. Particle and Fibre Toxicology. 2009;6, article 17 doi: 10.1186/1743-8977-6-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Wang JJ, Sanderson BJS, Wang H. Cyto- and genotoxicity of ultrafine TiO2 particles in cultured human lymphoblastoid cells. Mutation Research. 2007;628(2):99–106. doi: 10.1016/j.mrgentox.2006.12.003. [DOI] [PubMed] [Google Scholar]
  • 38.Sharma V, Shukla RK, Saxena N, Parmar D, Das M, Dhawan A. DNA damaging potential of zinc oxide nanoparticles in human epidermal cells. Toxicology Letters. 2009;185(3):211–218. doi: 10.1016/j.toxlet.2009.01.008. [DOI] [PubMed] [Google Scholar]
  • 39.Wise Sr. JP, Goodale BC, Wise SS, et al. Silver nanospheres are cytotoxic and genotoxic to fish cells. Aquatic Toxicology. 2010;97(1):34–41. doi: 10.1016/j.aquatox.2009.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.AshaRani PV, Mun GLK, Hande MP, Valiyaveettil S. Cytotoxicity and genotoxicity of silver nanoparticles in human cells. ACS Nano. 2009;3(2):279–290. doi: 10.1021/nn800596w. [DOI] [PubMed] [Google Scholar]
  • 41.Ahamed M, Karns M, Goodson M, et al. DNA damage response to different surface chemistry of silver nanoparticles in mammalian cells. Toxicology and Applied Pharmacology. 2008;233(3):404–410. doi: 10.1016/j.taap.2008.09.015. [DOI] [PubMed] [Google Scholar]
  • 42.Li JJ, Zou L, Hartono D, Ong C-N, Bay B-H, Yung L-YL. Gold nanoparticles induce oxidative damage in lung fibroblasts in vitro. Advanced Materials. 2008;20(1):138–142. [Google Scholar]
  • 43.Asharani PV, Xinyi N, Hande MP, Valiyaveettil S. DNA damage and p53-mediated growth arrest in human cells treated with platinum nanoparticles. Nanomedicine. 2010;5(1):51–64. doi: 10.2217/nnm.09.85. [DOI] [PubMed] [Google Scholar]
  • 44.Pierscionek BK, Li Y, Yasseen AA, Colhoun LM, Schachar RA, Chen W. Nanoceria have no genotoxic effect on human lens epithelial cells. Nanotechnology. 2010;21(3) doi: 10.1088/0957-4484/21/3/035102. Article ID 035102. [DOI] [PubMed] [Google Scholar]
  • 45.He L, Yang L, Zhang Z-R, et al. In vitro evaluation of the genotoxicity of a family of novel MeO-PEG-poly(D,L-lactic-co-glycolic acid)-PEG-OMe triblock copolymer and PLGA nanoparticles. Nanotechnology. 2009;20(45) doi: 10.1088/0957-4484/20/45/455102. Article ID 455102. [DOI] [PubMed] [Google Scholar]
  • 46.Nielsen GD, Roursgaard M, Jensen KA, Poulsen SS, Larsen ST. In vivo biology and toxicology of fullerenes and their derivatives. Basic and Clinical Pharmacology and Toxicology. 2008;103(3):197–208. doi: 10.1111/j.1742-7843.2008.00266.x. [DOI] [PubMed] [Google Scholar]
  • 47.Lao F, Chen L, Li W, et al. Fullerene nanoparticles selectively enter oxidation-damaged cerebral microvessel endothelial cells inhibit JNK-related apoptosis. ACS Nano. 2009;3(11):3358–3368. doi: 10.1021/nn900912n. [DOI] [PubMed] [Google Scholar]
  • 48.Hu Z, Guan W, Wang W, Huang L, Xing H, Zhu Z. Synthesis of β-alanine C60 derivative and its protective effect on hydrogen peroxide-induced apoptosis in rat pheochromocytoma cells. Cell Biology International. 2007;31(8):798–804. doi: 10.1016/j.cellbi.2007.01.013. [DOI] [PubMed] [Google Scholar]
  • 49.Sitharaman B, Zakharian TY, Saraf A, et al. Water-soluble fullerene (C60) derivatives as nonviral gene-delivery vectors. Molecular Pharmaceutics. 2008;5(4):567–578. doi: 10.1021/mp700106w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Foote CS, Diederich FN, Whetten R, Wudl F. Buckminsterfullerene. Chemical and Engineering News. 1990;68(51):p. 2. [Google Scholar]
  • 51.Aitken RC, Creely KS, Tran CL. Nanoparticles: An Occupational Hygiene Review. Suffolk, UK: HSE Books; 2004. [Google Scholar]
  • 52.Moreira S, Silva NB, Almeida-Lima J, et al. BC nanofibres: in vitro study of genotoxicity and cell proliferation. Toxicology Letters. 2009;189(3):235–241. doi: 10.1016/j.toxlet.2009.06.849. [DOI] [PubMed] [Google Scholar]
  • 53.Pulskamp K, Diabaté S, Krug HF. Carbon nanotubes show no sign of acute toxicity but induce intracellular reactive oxygen species in dependence on contaminants. Toxicology Letters. 2007;168(1):58–74. doi: 10.1016/j.toxlet.2006.11.001. [DOI] [PubMed] [Google Scholar]
  • 54.Allen NS, Edge M, Sandoval G, Verran J, Stratton J, Maltby J. Photocatalytic coatings for environmental applications. Photochemistry and Photobiology. 2005;81(2):279–290. doi: 10.1562/2004-07-01-RA-221. [DOI] [PubMed] [Google Scholar]
  • 55.Jackson MJ, Waqar A. Titanium Dioxide Coatings in Medical Devices. New York, NY, USA: Springer; 2007. [Google Scholar]
  • 56.Powell MC, Kanarek MS. Nanomaterial health effects—part 2: uncertainties and recommendations for the future. Wisconsin Medical Journal. 2006;105(3):18–23. [PubMed] [Google Scholar]
  • 57.Kim I-S, Baek M, Choi S-J. Comparative Cytotoxicity of Al2O3 , CeO2 , TiO2 and ZnO nanoparticles to human lung cells. Journal of Nanoscience and Nanotechnology. 2010;10(5):3453–3458. doi: 10.1166/jnn.2010.2340. [DOI] [PubMed] [Google Scholar]
  • 58.Case CP. Chromosomal changes after surgery for joint replacement. The Journal of Bone and Joint Surgery—Series B. 2001;83(8):1093–1095. doi: 10.1302/0301-620x.83b8.12755. [DOI] [PubMed] [Google Scholar]
  • 59.Hardman R. A toxicologic review of quantum dots: toxicity depends on physicochemical and environmental factors. Environmental Health Perspectives. 2006;114(2):165–172. doi: 10.1289/ehp.8284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Derfus AM, Chan WCW, Bhatia SN. Probing the cytotoxicity of semiconductor quantum dots. Nano Letters. 2004;4(1):11–18. doi: 10.1021/nl0347334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Li KG, Chen JT, Bai SS, et al. Intracellular oxidative stress and cadmium ions release induce cytotoxicity of unmodified cadmium sulfide quantum dots. Toxicology in Vitro. 2009;23(6):1007–1013. doi: 10.1016/j.tiv.2009.06.020. [DOI] [PubMed] [Google Scholar]
  • 62.Waisberg M, Joseph P, Hale B, Beyersmann D. Molecular and cellular mechanisms of cadmium carcinogenesis. Toxicology. 2003;192(2-3):95–117. doi: 10.1016/s0300-483x(03)00305-6. [DOI] [PubMed] [Google Scholar]
  • 63.Rousseau M-C, Straif K, Siemiatycki J. IARC carcinogen update. Environmental Health Perspectives. 2005;113(9):A580–A581. doi: 10.1289/ehp.113-1280416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Flora SJS, Mittal M, Mehta A. Heavy metal induced oxidative stress & its possible reversal by chelation therapy. Indian Journal of Medical Research. 2008;128(4):501–523. [PubMed] [Google Scholar]
  • 65.Kumari M, Mukherjee A, Chandrasekaran N. Genotoxicity of silver nanoparticles in Allium cepa. Science of the Total Environment. 2009;407(19):5243–5246. doi: 10.1016/j.scitotenv.2009.06.024. [DOI] [PubMed] [Google Scholar]
  • 66.Pan Y, Leifert A, Ruau D, et al. Gold nanoparticles of diameter 1.4 nm trigger necrosis by oxidative stress and mitochondrial damage. Small. 2009;5(18):2067–2076. doi: 10.1002/smll.200900466. [DOI] [PubMed] [Google Scholar]
  • 67.Hirst SM, Karakoti AS, Tyler RD, Sriranganathan N, Seal S, Reilly CM. Anti-inflammatory properties of cerium oxide nanoparticles. Small. 2009;5(24):2848–2856. doi: 10.1002/smll.200901048. [DOI] [PubMed] [Google Scholar]
  • 68.López-Moreno ML, de la Rosa G, Hernandez-Viezcas JA. Evidence of the differential biotransformation and genotoxicity of ZnO and CeO(2) nanoparticles on Soybean (Glycine max) plants. doi: 10.1021/es903891g. Environmental Science & Technology. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.He Q, Shi J, Chen F, Zhu M, Zhang L. An anticancer drug delivery system based on surfactant-templated mesoporous silica nanoparticles. Biomaterials. 2010;31(12):3335–3346. doi: 10.1016/j.biomaterials.2010.01.015. [DOI] [PubMed] [Google Scholar]
  • 70.Liong M, Lu J, Kovochich M, et al. Multifunctional inorganic nanoparticles for imaging, targeting, and drug delivery. ACS Nano. 2008;2(5):889–896. doi: 10.1021/nn800072t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Piddubnyak V, Kurcok P, Matuszowicz A, et al. Oligo-3-hydroxybutyrates as potential carriers for drug delivery. Biomaterials. 2004;25(22):5271–5279. doi: 10.1016/j.biomaterials.2003.12.029. [DOI] [PubMed] [Google Scholar]
  • 72.Chauhan AS, Diwan PV, Jain NK, Tomalia DA. Unexpected in vivo anti-inflammatory activity observed for simple, surface functionalized poly(amidoamine) dendrimers. Biomacromolecules. 2009;10(5):1195–1202. doi: 10.1021/bm9000298. [DOI] [PubMed] [Google Scholar]
  • 73.Song Y, Li X, Du X. Exposure to nanoparticles is related to pleural effusion, pulmonary fibrosis and granuloma. European Respiratory Journal. 2009;34(3):559–567. doi: 10.1183/09031936.00178308. [DOI] [PubMed] [Google Scholar]
  • 74.Brain JD, Kreyling W, Gehr P. To the editors: express concern about the recent paper by Song et al. The European Respiratory Journal. 2010;35(1):226–227. doi: 10.1183/09031936.00159909. [DOI] [PubMed] [Google Scholar]
  • 75.Takagi A, Hirose A, Nishimura T, et al. Induction of mesothelioma in p53+/- mouse by intraperitoneal application of multi-wall carbon nanotube. Journal of Toxicological Sciences. 2008;33(1):105–116. doi: 10.2131/jts.33.105. [DOI] [PubMed] [Google Scholar]
  • 76.Sakamoto Y, Nakae D, Fukumori N, et al. Induction of mesothelioma by a single intrascrotal administration of multi-wall carbon nanotube in intact male Fischer 344 rats. Journal of Toxicological Sciences. 2009;34(1):65–76. doi: 10.2131/jts.34.65. [DOI] [PubMed] [Google Scholar]
  • 77.Muller J, Delos M, Panin N, Rabolli V, Huaux F, Lison D. Absence of carcinogenic response to multiwall carbon nanotubes in a 2-year bioassay in the peritoneal cavity of the rat. Toxicological Sciences. 2009;110(2):442–448. doi: 10.1093/toxsci/kfp100. [DOI] [PubMed] [Google Scholar]
  • 78.Bergamaschi E, Bussolati O, Magrini A, et al. Nanomaterials and lung toxicity: interactions with airways cells and relevance for occupational health risk assessment. International Journal of Immunopathology and Pharmacology. 2006;19(supplement 4):3–10. [PubMed] [Google Scholar]
  • 79.Borm PJA, Kreyling W. Toxicological hazards of inhaled nanoparticles—potential implications for drug delivery. Journal of Nanoscience and Nanotechnology. 2004;4(5):521–531. doi: 10.1166/jnn.2004.081. [DOI] [PubMed] [Google Scholar]
  • 80.Wang D, Sun L, Liu W, Chang W, Gao X, Wang Z. Photoinduced DNA cleavage by α-, β-, and γ-cyclodextrin- bicapped C60 supramolecular complexes. Environmental Science and Technology. 2009;43(15):5825–5829. doi: 10.1021/es900709p. [DOI] [PubMed] [Google Scholar]
  • 81.Radomski A, Jurasz P, Alonso-Escolano D, et al. Nanoparticle-induced platelet aggregation and vascular thrombosis. British Journal of Pharmacology. 2005;146(6):882–893. doi: 10.1038/sj.bjp.0706386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Sera N, Tokiwa H, Miyata N. Mutagenicity of the fullerene C60-generated singlet oxygen dependent formation of lipid peroxides. Carcinogenesis. 1996;17(10):2163–2169. doi: 10.1093/carcin/17.10.2163. [DOI] [PubMed] [Google Scholar]
  • 83.Baan R, Straif K, Grosse Y, Secretan B, El Ghissassi F, Cogliano V. Carcinogenicity of carbon black, titanium dioxide, and talc. The Lancet Oncology. 2006;7(4):295–296. doi: 10.1016/s1470-2045(06)70651-9. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Nucleic Acids are provided here courtesy of Wiley

RESOURCES