You are on page 1of 15

Article

Toxicology and Industrial Health


1–15
Genotoxic and oxidative stress  The Author(s) 2015
Reprints and permissions:

potential of nanosized and bulk zinc sagepub.co.uk/journalsPermissions.nav


DOI: 10.1177/0748233715599472
tih.sagepub.com
oxide particles in Drosophila melanogaster

Erico R Carmona1, Claudio Inostroza-Blancheteau2,


Laura Rubio3 and Ricard Marcos3

Abstract
Zinc oxide nanoparticles (ZnONP) are manufactured on a large scale and can be found in a variety of consumer
products, such as sunscreens, lotions, paints and food additives. Few studies have been carried out on its
genotoxic potential and related mechanisms in whole organisms. In the present study, the in vivo genotoxic
activity of ZnONP and its bulk form was assayed using the wing-spot test and comet assay in Drosophila
melanogaster. Additionally, a lipid peroxidation analysis using the thiobarbituric acid assay was also performed.
Results obtained with the wing-spot test showed a lack of genotoxic activity of both ZnO forms. However,
when both particle sizes were tested in the comet assay using larvae haemocytes, a significant increase in DNA
damage was observed for ZnONP treatments but only at the higher dose applied. In addition, the lipid per-
oxidation assay showed significant malondialdehyde (MDA) induction for both ZnO forms, but the induction of
MDA for ZnONP was higher for the ZnO bulk, suggesting that the observed DNA strand breaks could be
induced by mediated oxidative stress. The overall data suggest that the potential genotoxicity of ZnONP in
Drosophila can be considered weak according to the lack of mutagenic and recombinogenic effects and the
induction of primary DNA damage only at high toxic doses of ZnONP. This study is the first assessing the
genotoxic and oxidative stress potential of nano and bulk ZnO particles in Drosophila.

Keywords
Comet assay, DNA damage, haemocytes, malondialdehyde, wing-spot test, ZnO

Introduction NM are commonly defined as engineering materi-


als with one or more dimensions having a size in the
The development of the nanotechnology industry
range of 1–100 nm and include nanoparticles (NP),
implies the creation and use of new manufactured
nanofibres, nanotubes, composite materials and
nanomaterials (NM) with interesting physical–chem-
ical properties, such as high durability, mechanical
resistance, high conductivity and high chemical reac- 1
Grupo de Genotoxicologı́a, Núcleo de Investigación en Estudios
tivity. All of this makes NM attractive for different Ambientales, Facultad de Recursos Naturales, Escuela de Medicina
areas ranging from industry and pharmaceutical appli- Veterinaria, Universidad Católica de Temuco, Temuco, Chile
2
cations to environmental remediation and biomedi- Núcleo de Investigación en Producción Alimentaria, Facultad de
Recursos Naturales, Escuela de Agronomı́a, Universidad Católica
cine. Until now, more than 1600 globally available
de Temuco, Temuco, Chile
consumer products use manufactured NM (nanotech- 3
Grup de Mutagènesi, Departament de Genètica i de Microbiologia,
project.org), and predictions suggest that these nano- Facultat de Biociències, Universitat Autònoma de Barcelona,
based products will increase dramatically in the next Cerdanyola del Vallès, Barcelona, Spain
years (Fairbrother, 2010; Hansen et al., 2009). Hence,
Corresponding author:
the current human and environmental exposure to NM
Erico R Carmona, Núcleo de Investigación en Estudios Ambientales,
will inevitably increase in the near future (Buzea Universidad Católica de Temuco, P.O. Box 15-D, Temuco 4813302,
et al., 2007; Savolainen et al., 2010), requiring more Chile.
data on their potential secondary effects. Email: ecarmona@uct.cl

Downloaded from tih.sagepub.com at University of York on January 24, 2016


2 Toxicology and Industrial Health

nanostructured surfaces (Borm et al., 2006). Particles Nam et al., 2013). All these suggest that the genotoxic
in nanoscale show particular shape, large surface area potential of ZnONP remains controversial.
and surface reactivity. However, the same characteris- Oxidative stress has been proposed as one of the
tics mentioned above could contribute to human main mechanisms involved in toxic and genotoxic
health hazards and risks because they may also be manifestations induced by nanoscale particles (Shri-
responsible for adverse biological effects (Xia et al., vastava et al., 2014; Wang et al., 2013). Therefore,
2009). To date, there is increasing evidence that NP oxidative stress analysis must be included in any gen-
manifest increased toxicity compared to micrometric otoxicity study carried out with NP to elucidate possi-
particles of the same composition (Chibber et al., ble mechanisms related to induced genetic damage.
2013). Thus, several studies reveal that NP can induce Several studies have demonstrated that ZnONP can
different adverse effects in biological systems, such as produce DNA damage mediated by the generation of
oxidative stress, cytotoxicity, DNA damage, apoptosis oxidative stress, as evidenced by an increase of reac-
and inflammatory responses (Chibber et al., 2013; Cho tive oxygen species production, oxidative DNA dam-
et al., 2010; Kumar and Dhawan, 2013). Considering age and depletion of antioxidant defence in diverse
this background, an increasing number of toxicological types of cultured human cells (Alarifi et al., 2013;
investigations are currently being carried out with NP Demir et al., 2014; Lin et al., 2009; Sharma et al.,
to determine their potential toxic effects and to eluci- 2009, 2012a). Such has also been reported for bacteria
date human and environmental health risks (Dhawan cells where glutathione depletion with an associated
and Sharma, 2010). increase in free radicals, lipid peroxidation end prod-
Among the manufactured NM more widely pro- ucts and lactate dehydrogenase activity demonstrate
duced are zinc oxide nanoparticles (ZnONP). These that ZnONP produce oxidative stress leading to geno-
may be found in a wide assortment of consumer prod- toxicity in Escherichia coli (Kumar et al., 2011b).
ucts, such as sunscreens, cosmetics, paints, pigments, Similarly, some in vivo studies suggest that exposure
food additives and disinfectants (Espitia et al., 2012; to ZnONP leads to an accumulation of NP in the liver
Raj et al., 2012). Since ZnONP provide high trans- of mice and rats, causing DNA damage and apoptosis
parency in the skin and better protection from ultra- as a result of oxidative stress (Nounou et al., 2013;
violet (UV) solar radiation than bulk or micrometric Sharma et al., 2012b). To date, studies evaluating oxi-
counterparts, this NM is being destined for use in dative stress and DNA damage after ZnONP exposure
modern sunscreens and cosmetics (Osmond and have not been carried out in Drosophila
McCall, 2010). Thus, the probability of human and melanogaster.
environmental exposure to ZnONP is increasing, and In nanotoxicology, in vitro models are usually
their potential adverse biological effects must be employed to evaluate and describe the toxicity of
adequately studied. NM. However, in vivo studies could be considered
According to various authors, ZnONP can show much more interesting in terms of risk evaluation than
genotoxic potential in several cell types and organism in vitro approaches with mammal or human cells
models, for example, it has been proven to induce since they do not completely simulate the complex
mutagenicity in bacterial assay systems (Kumar et al., cell–cell, cell–matrix interactions and hormonal effects
2011a, 2011b). Likewise, in vitro studies indicate that found in the in vivo systems (Chibber et al., 2013). In
ZnONP can induce DNA breaks; micronuclei forma- this scenario, recent studies have demonstrated that
tion; and chromosomal breakage in nasal mucosa, epi- the fruit fly D. melanogaster offers several benefits
dermal, neuronal and human blood cells (Bhattacharya as an in vivo model for the study of dietary intake and
et al., 2014; Demir et al., 2014; Hackenberg et al., tissue distribution of nano-carbon–based materials
2011; Sharma et al., 2012b; Valdiglesias et al., (Leeuw et al., 2007; Liu et al., 2009), the study of the
2013). In addition, some in vivo studies with animals potential toxicity of metal and metal oxide-based NP
have revealed genetic damage, apoptosis and tissue on reproduction and development (Gorth et al., 2011;
inflammation associated with ZnONP exposure Philbrook et al., 2011; Pompa et al., 2011; Posgai
(Nounou et al., 2013; Pasupuleti et al., 2012; Sharma et al., 2011) and the study of genotoxicity after expo-
et al., 2012b). Nevertheless, negative results have also sure to silver, cobalt, gold, titanium, zirconium and
been reported recently in the literature where ZnONP aluminium NP (Ahamed et al., 2010; Demir et al.,
were unable to exert genotoxic effects in several 2011, 2013; Sabella et al., 2011; Vales et al., 2012;
in vitro and in vitro test systems (Kwon et al., 2014; Vecchio et al., 2012). The aforementioned results

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 3

would reinforce the usefulness of the Drosophila performed. Lipid peroxidation is a consequence of
model as a first-tier in vivo test for genotoxicity test- free radical attacks in the membrane system and
ing of NM. could be considered an indirect measure of oxidative
In this context, the aim of the present study was to stress in the cells. This study employed the thiobar-
explore the in vivo genotoxic activity and the levels bituric acid reactive substance (TBARS) assay
of oxidative stress associated with the exposure of described by Tironi et al. (2007). This biochemical
ZnONP as compared to the effects induced by its bulk test is able to detect the extent of peroxidation of
form using genotoxic and biochemical approaches in lipids by measuring the malondialdehyde (MDA)
D. melanogaster. concentration in tissue extracts of exposed larvae.
The wing-spot test was used as a short-test system MDA is one of the low-molecular-weight end prod-
based on the loss of heterozygosity (LOH) in normal ucts formed via the decomposition of certain primary
genes, and the corresponding expressions of recessive and secondary lipid peroxidation products (Halliwell
markers are called multiple wing hairs (mwh) and and Chirico, 1993).
flare-3 (flr3) in the wing blade of adult flies (Graf
et al., 1984). Thus, the induced genotoxic effects have
been microscopically observed as an increase in the Materials and methods
frequency of mutant clones markers (mwh or flr3) in
wing slides preparation. This assay can detect a wide Chemicals
range of mutational events such as point mutations, ZnONP (!100 nm average particle size, surface area
deletions, certain types of chromosome aberrations 15–25 m2/g, >99% purity, CAS no. 1314-13-2, ref.
(non-disjunction) and mitotic recombination (Graf no. 544906), ZnO bulk (<5 mm, "99.0% purity, CAS
et al., 1984). It should be noted that the quantification no. 1314-13-2, ref. no. 96479), ethyl methane sulpho-
of mitotic recombination in somatic cells is consid- nate (EMS, 100% purity, CAS no. 62-50-0), Triton X-
ered a relevant process for genotoxicity screening 100, Trizma base, sodium hydroxide (NaOH), sodium
because aberrant recombination activity is commonly chloride (NaCl) and disodium salt of ethylenediami-
associated with carcinogenesis (Luo et al., 2000; netetraacetic acid (Na2-EDTA) were obtained from
Sengstag, 1994). Sigma-Aldrich (St Louis, Missouri, USA); low-
An in vivo comet assay was also applied using hae- melting-point agarose (LMA), normal melting point
mocytes of D. melanogaster larvae to detect the induc- agarose (NMA), phosphate-buffered saline (PBS)
tion of primary DNA damage (Carmona et al., 2011b; and 4,6-diamidine-2-phenylindole (DAPI) were
Marcos and Carmona, 2013). Haemocytes of Droso- obtained from Life Technologies Corporation
phila, present in the haemolymph, may be directly (Carlsbad, California, USA); phenylthiourea (PTU),
exposed to toxic substances circulating in the haemo- N-lauroylsarcosine sodium salt, potassium chloride,
lymph and are, therefore, an interesting target tissue trichloroacetic acid (TCA) and 2-thiobarbituric acid
for genotoxicity assessment (Villela et al., 2006). The (TBA) were obtained from Merck Company (Darm-
comet assay is a sensitive tool for the detection of stadt, Germany).
several kinds of DNA damage, such as double- and
single-strand DNA breaks, DNA alkali-labile sites and
incomplete repair sites in individual cells (Tice et al., ZnONP preparation
2000). This assay is a reliable technique to assess DNA Various concentrations of ZnONP were prepared with
damage in different model organisms, including D. distilled water. Dispersion was carried out by sonica-
melanogaster (Dhawan et al., 2009). The alkaline tion using an ultrasonic bath (37 kHz, Elmasonic S,
comet assay version has been proven as a useful and Elma, Germany) for 30 min at room temperature. ZnO
sensitive method for detecting DNA damage caused bulk was used to compare the genotoxic effects
by fullerenes, carbon nanotubes, quantum dots, between micrometric and nanoparticulated forms.
metal-based and metal oxide NP; therefore, it is highly The compound was prepared with distilled water
recommended for the genotoxic risk assessment of NM and diluted through magnetic stirring for 10 min
(Karlsson, 2010). at room temperature. Distilled water was used for
Finally, to determine the induction of oxidative negative control, whilst the mutagenic agent EMS
stress in treated Drosophila larvae with ZnONP and was used as positive control in each experiment car-
bulk forms, a lipid peroxidation assay was also ried out with both the wing-spot and comet assays.

Downloaded from tih.sagepub.com at University of York on January 24, 2016


4 Toxicology and Industrial Health

NP characterization periods in culture bottles containing the standard


To confirm the physical characteristics of ZnONP, the medium. The resulting 3-day-old larvae (third instar
following techniques were performed: transmission larvae) were then placed in plastic vials containing
electron microscopy (TEM), dynamic light scattering 4.5 g of Drosophila instant medium (Carolina Bio-
(DLS) and laser doppler velocimetry (LDV). TEM logical Supply, Burlington, North Carolina, USA)
was carried out with a JEOL JEM-2011 instrument prepared with 10 ml of various non-toxic concen-
(Japan) to determine the size and shape of ZnONP trations of ZnONP and ZnO bulk (6, 12, 18 and
in dry form. DLS and LDV were performed with a 24 mM, respectively). Larvae were fed this medium
Malvern Zetasizer Nano-ZS zen3600 instrument until pupation. The surviving adults were collected
(UK) to measure the hydrodynamic diameter and zeta and stored in 70% ethanol. Afterwards, their wings
potential in aqueous suspension, respectively. For were removed with fine tweezers and mounted in
TEM analyses, ZnONP were measured at a concentra- Faure’s solution on microscope slides. The wings
tion of 2.56 mg/ml. For DLS and LDV techniques, were scored at 400$ magnification for the presence
ZnONP samples were measured at a concentration of small single spots, large single spots and twin spots
of 10 mg/ml. (Figure 1). In each series, 60 wings were scored (from
30 individuals). Scoring of flies and data evaluation
were conducted following the standard procedures for
Strains the wing-spot test, as used in recent investigations
The following mutant Drosophila strains were used (Arossi et al., 2010).
for the wing-spot test: the multiple wing hairs strain The conditional binomial test was applied to assess
with the genetic constitution y; mwhj; and the flare- differences between the frequencies of each type of
3 strain with the genetic constitution flr3/ln (3LR) spot in treated and concurrent negative control with
TM3, Bds. The multiple wing hairs marker (mwh, significant levels ! ¼ " ¼ 0.05 (Kastenbaum and
3–0.3) is a completely recessive homozygous viable Bowman, 1970). The multiple-decision procedure
mutation, which is kept in homozygous condition. It was used to judge the overall response of an agent
produces multiple trichomes per cell instead of the as positive, negative or inconclusive (Frei and Würg-
normally unique trichome in the wing cells. The ler, 1988). The treatment was considered positive if
flare-3 marker (flr3, 3–38.8) is a recessive mutation the frequency of mutant clones in the treated series
that affects the shape of wing hairs, producing mal- was at least m (multiplication factor) times greater
formed wing hairs that have a flare shape. Given their than in the control series (Frei and Würgler, 1988).
zygotic lethality, flare alleles have to be kept in stocks Since small single spots and total spots have a com-
over balancer chromosomes carrying multiple inver- paratively high spontaneous frequency, m was fixed
sions and a dominant marker that is a lethal homo- at a value of 2 (testing for a doubling of the sponta-
zygous (TM3, Bds). More detailed information on neous frequency). For large single spots and twin
genetic markers and descriptions of the phenotypes spots, which have a low spontaneous frequency, m
are given by Lindsley and Zimm (1992). ¼ 5 was used. The frequency of clone formation was
In this study, the flr3 strain was also used for the calculated, without size correction, by dividing the
comet assay experiments with Drosophila haemo- number of mwh clones per wing by 24,400, which is
cytes. This strain was chosen according to the high the approximate number of cells inspected in one
female fertility and low background level of DNA wing (Alonso-Moraga and Graf, 1989).
damage in blood cells from untreated third instar lar-
vae. Both strains were cultured in glass bottles with
the standard medium for Drosophila (i.e. agar, corn
Comet assay
flour and yeast) at a temperature of 25 + 1# C and a Third instar larvae at 72 + 4-h-old were placed in
relative humidity of approximately 60%. plastic vials containing 4.5 g of Drosophila instant
medium prepared with 10 ml of various non-toxic
concentrations of ZnONP and ZnO bulk (6, 12 and
Wing-spot test 24 mM, respectively). Haemocytes from Drosophila
Virgin females of the flr3 strain were mated to mwh haemolymph were then collected according to the
males as previously described (Marcos and Carmona, method described by Marcos and Carmona (2013).
2013). Eggs from this cross were collected during 8 h Briefly, chilled larvae at 96 + 4-h-old were removed

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 5

Figure 1. Typical manifestations of mutant spots affecting trichomes number and shapes on wing surface of Drosophila
melanogaster. (a) Small single mwh spot; (b) large single mwh spot; and (c) twin spot. Images were taken with $400
magnification.

from food media, washed in water, sterilized in 5% were mounted in each slide and covered with a cov-
bleach and dried. The cuticles from 50 larvae were erslip. Immediately after agarose solidification (for
then disrupted with two fine forceps under a stereo- 10 min at 4# C), the coverslips were removed and the
scopic microscope. The haemolymph and circulating slides were immersed in a cold fresh lysis solution
haemocytes were directly collected into a drop of cold (2.5 M NaCl, 100 mM Na2-EDTA, 10 mM Tris, 1%
PBS solution containing 0.07% PTU and separated in Triton X-100 and 1% N-lauroylsarcosine, pH 10) for
1.5 ml microcentrifuge tube. Pooled haemolymph was 2 h at 4# CC in a dark chamber. To prevent additional
centrifuged at 300g for 10 min, the supernatant was DNA damage, the following steps were performed
discarded, and the pellet was resuspended in 20 ml under dim light: the slides were placed for 25 min
of cold PBS. in a horizontal gel electrophoresis tank filled with
The comet assay was performed as previously cold electrophoretic buffer (1 mM Na2-EDTA and
described by Singh et al. (1988) with minor modifica- 300 mM NaOH, pH 13) to allow DNA unwinding.
tions. Cell samples (approximately 40,000 cells in 20 Electrophoresis was carried out in the same buffer for
ml) were carefully resuspended in 75 ml of 0.75% 20 min at 25 V and 300 mA. Unwinding and electro-
LMA and layered onto microscope slides pre-coated phoresis processes were done at 4# C. After electro-
with 1% NMA (dried at room temperature). Two gels phoresis, slides were neutralized with two washes of

Downloaded from tih.sagepub.com at University of York on January 24, 2016


6 Toxicology and Industrial Health

0.4 mM Tris (pH 7.5) for 5 min each. The slides were USA). TBARS levels were converted to MDA and
stained with 20 ml of DAPI (1 mg/ml) per gel. The expressed in milligram of MDA/kilogram of sample
images were examined at 400$ magnification with using a molar extinction coefficient of 1.56 $ 105 M&1.
a Nikon Eclipse E200 fluorescence microscope ANOVA was used to analyse differences in MDA
(Japan), coupled with a CMOS digital camera. One levels with different treatments and compounds.
hundred randomly selected cells (50 cells on each one Tukey’s post hoc test was performed to compare neg-
of the two replicate slides) were analysed per treat- ative controls versus different treatments and ZnO
ment. The percentage of DNA in the tail (%DNA tail) particles. Results were considered statistically signif-
was used to measure DNA damage since this is the icant at p ! 0.05. All data were presented as arith-
most widely used and recommended parameter for metic mean + standard deviation.
comet data analysis (Kumaravel et al., 2009). The
%DNA tail was computed using CometScore version
1.5 image analysis system (TriTek Corp., Sumerduck, Results
Virginia, USA). Physical characterization of ZnONP
A generalized linear model (GLM) was used to
analyse differences in the %DNA tail with different TEM was used to characterize the morphology and
treatments and compounds. The conservative Scheffe size of ZnONP. Most of the NP analysed were sphe-
post hoc test was performed to compare negative con- rical, triangle, rod and rectangle shapes (i.e. hexago-
trols versus different treatments. The GLM is analo- nal shapes) and showed low levels of agglomeration
gous to traditional analysis of variance (ANOVA), (Figure 2(a)). The size of the ZnONP agglomerates
but it allows the use of non-parametric and heterosce- ranged from 25.06 to 360.97 nm, and the average +
dastic data, which was our case (Bolker et al., 2009). SD size was 106.55 + 64.79 nm (Figure 2(b)). TEM
Before analysis with GLM, the homogeneity of var- images and analyses of representative NP (n ¼ 200)
iance and normality assumption of data were tested indicate that a significant number of ZnONP (40%
with the Bartlett and Kolmogorov–Smirnov tests, approximately) show sizes higher than 100 nm, dif-
respectively. Results were considered statistically sig- fering from the manufacturer indications (less than
nificant at p ! 0.05. All data were presented as arith- 100 nm).
metic mean + standard error, and 95% confidence DLS and LDV techniques were used to measure
intervals were constructed. hydrodynamic diameter and zeta potential of ZnONP
in water suspension with previous dispersion by soni-
cation. The diameter average in water suspension was
Lipid peroxidation assay different and higher than TEM analyses, reaching the
mean + SD values of 225.40 + 0.93 nm (Figure 3(a)).
Third instar larvae of Drosophila (72 + 4-h-old) were
Finally, the average + SD of zeta potential was
treated with three different concentrations of nano and
&21.00 + 0.80 for ZnONP (Figure 3(b)), indicating
bulk-ZnO (6, 12 and 24 mM, respectively). Control lar-
a good stability and dispersion of this nano compound
vae received untreated instant medium for Drosophila
in aqueous solution for genotoxic experiments with
rehydrated with distilled water. Larvae were exposed
Drosophila larvae.
to the compounds during approximately 24 h. Lipid
The physical properties of ZnO bulk were described
peroxidation or TBARS assay was performed accord-
in detail in a previous work by Lin et al. (2009). In
ing to the modified method of Tironi et al. (2007).
summary, most of the ZnO particles showed hexagonal
Approximately 0.5 g of larvae were used in each expo-
shapes. The primary size average was 4.2 mm (mea-
sure and control groups. Three replications and one
sured by TEM) and the hydrodynamic size was 9 mm
independent experiment were assessed. Control and
(with DLS methodology). The ZnO particles showed
ZnONP-treated larvae were homogenized with 0.5%
a specific surface area of 8.61 m2/g Brunauer-
w/v of TCA. The larvae homogenates were maintained
Emmett-Teller analysis (BET).
for 30 min in ice and then were filtered. A mixture of 0.5
ml filtered tissue homogenates and 0.5 ml of 0.5% w/v
TBA solution was incubated for 30 min to 70# C. The Toxicity of ZnO compounds
absorbance was measured at 532 nm using a Thermo- The different concentrations of nano and bulk forms of
Scientific SpectronicGenesys 10 UV-visible (Vis) ZnO used in the genotoxicity experiments with D. mel-
scanning spectrophotometer (Madison, Wisconsin, anogaster strains were selected according to previous

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 7

Figure 2. Physical characterization of ZnONP with TEM. (a) TEM image with high magnification (bar scales representing
200 nm) and (b) histogram showing the size distribution of ZnONP. ZnONP: zinc oxide nanoparticles; TEM: transmission
electron microscopy.

to choose the final concentrations was the number of


emerging larvae and adults after treatments were high
enough to perform genotoxic experiments with wing-
spot and comet tests (Marcos and Carmona, 2013).

Genotoxicity analysis with the wing-spot test


Both sizes of ZnO were supplied to 72-h-old larvae (third
instar) at concentrations ranging from 6 mM to 24 mM.
The results indicate that no concentration of ZnONP and
ZnO bulk, administered by ingestion, induced signifi-
cant increases in the frequency of mutant spots (i.e. small
single, large single and twin spots), compared with the
negative control (Tables 1 and 2). In this study, the neg-
ative control frequencies (0.30 and 0.28) were in accor-
dance with the normal background range observed in our
laboratory and were not significantly different from pre-
Figure 3. Characterization of ZnONP in water suspen- vious results (Carmona et al., 2011a; Gürbüzel et al.,
sion. (a, b) Hydrodynamic size and zeta potential measured 2013). The positive control carried out with 1 mM of
with DLS and LDV techniques, respectively. ZnONP: zinc EMS showed a clear response, and the mutant spot fre-
oxide nanoparticle; DLS: dynamic light scattering; LDV: quencies scored were also in agreement with previous
laser doppler velocimetry. and recent studies (Demir and Kaya, 2013; Gürbüzel
et al., 2012; Karadeniz et al., 2011), which supports the
toxicity and viability studies carried out in the labora- validity of the negative results found for the nanoparticu-
tory. Hence, a range of doses from 6 to 24 mM was lated and micrometric compounds of ZnO.
established for both particle sizes of ZnO. Viability was
higher than 70% in all concentrations, except to 24 mM
where the viability was approximately 50%, and no dif- DNA damage evaluated with the comet assay
ferences in toxicity were observed between nanosized The results obtained in the comet assay to test the pri-
and bulk forms of ZnO for both mwh and flr3 strains mary DNA damage (expressed as percentage of DNA
of D. melanogaster (data not shown). The main criterion in tail) of NP and bulk compounds of ZnO on larval

Downloaded from tih.sagepub.com at University of York on January 24, 2016


8 Toxicology and Industrial Health

Table 1. Induction of mutant spots in Drosophila larvae treated with ZnONP.a


Small single Large single
spots (1–2 cells) spots (>2 cells) Twin spots Total mwh Total spots
(m ¼ 2) (m ¼ 5) (m ¼ 5) spots (m ¼ 2) (m ¼2) Frequency of clone
Compounds formation per
concentration (mM) N Fr D N Fr D N Fr D N Fr D N Fr D 105 cells
ZnONP
Control 15 0.25 1 0.02 1 0.02 17 0.28 18 0.30 1.22
6 10 0.17 & 3 0.05 i 1 0.02 & 14 0.23 & 15 0.25 & 1.02
12 13 0.22 & 1 0.02 & 0 0.00 14 0.23 & 15 0.25 & 1.02
18 20 0.33 i 0 0.00 0 0.00 20 0.33 i 20 0.33 i 1.35
24 15 0.38 & 3 0.05 i 1 0.02 & 19 0.32 i 19 0.32 i 1.31
EMS
1 162 2.70 þ 49 0.82 þ 33 0.55 þ 244 4.06 þ 260 4.33 þ 17.75
N: number of clones; Fr: frequency; D: statistical diagnosis according to Frei and Würgler (1988): þ: positive; &: negative; i: inconclusive;
m: multiplication factor; ZnONP: zinc oxide nanoparticles; EMS: ethyl methane sulphonate.
a
Results obtained from mwh/flr3 wings. Probability level ! ¼ " ¼ 0.05; 60 wings were analysed for each concentration (30 individuals).
Ultrapure water was used as negative control, and EMS as positive control.

Table 2. Mutant spots induced in Drosophila larvae treated with ZnO bulk.a
Small single Large single
spots (1–2 cells) spots (>2 cells) Twin spots Total mwh Total spots
(m ¼ 2) (m ¼ 5) (m ¼ 5) spots (m ¼ 2) (m ¼ 2) Frequency of clone
Compounds formation per
concentration (mM) N Fr D N Fr D N Fr D N Fr D N Fr D 105 cells
ZnO bulk
Control 15 0.25 1 0.02 1 0.02 17 0.28 17 0.28 1.14
6 12 0.20 & 2 0.03 i 1 0.02 & 15 0.25 & 16 0.26 & 1.06
12 14 0.23 & 1 0.02 & 1 0.02 & 16 0.26 & 16 0.26 & 1.06
18 14 0.23 & 2 0.03 i 0 0.00 16 0.26 & 16 0.26 & 1.06
24 18 0.30 i 2 0.03 i 0 0.00 20 0.33 i 20 0.33 i 1.35
EMS
1 195 3.25 þ 40 0.66 þ 34 0.56 þ 269 4.48 þ 284 4.73 þ 19.39
N: number of clones; Fr: frequency; D: statistical diagnosis according to Frei and Würgler (1988); þ: positive; &: negative; i: inconclusive;
m: multiplication factor; ZnO: zinc oxide; EMS: ethyl methane sulphonate.
a
Results obtained from mwh/flr3 wings. Probability level ! ¼ " ¼ 0.05; 60 wings were analysed for each concentration (30 individuals).
Distilled water was used as negative control, and EMS as positive control.

haemocytes of D. melanogaster are summarized in Contrary to the results found with nanosized ZnO,
Table 3. Both compounds were administered by feeding the treatments carried out with different concentra-
during 24 + 4 h to third instar larvae in the same doses as tions of micrometric ZnO did not induce significant
in the wing-spot test, ranging from 6 mM to 24 mM. DNA damage in haemolymph cells of D. melanoga-
Concurrent negative and positive controls were also per- ster as compared to the negative control. The negative
formed for each comet experiment. Afterwards, the hae- and positive control values in %DNA in tail were in
molymph was extracted from the larvae, and circulating agreement with the background range observed in
haemocytes were isolated by centrifugation for comet recent studies with the same test (Demir, 2012; Demir
testing and data analysis of primary DNA damage. and Kaya, 2013; Mukhopadhyay et al., 2004).
The data obtained indicate that the nanoparticu-
lated form of ZnO can induce significant increases Lipid peroxidation of ZnONP in Drosophila
of DNA damage in larval hemocytes, but only at the
Third instar Drosophila larvae were exposed to nano
higher concentrations tested, that is, 24 mM (Figure 4).
and bulk forms of ZnO by feeding during 24 + 4 h at

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 9

Table 3. Primary DNA damage induction, measured as the


percentage of DNA in the tail, obtained in larval hemocytes
with the comet assay.
Compound 95% CI of %DNA
concentration (mM) %DNA in taila in tailb
Pooled negative control 11.41 + 0.58 10.28–12.54
Pooled positive control 21.61 + 0.93c 19.80–23.42
ZnONP
6 10.32 + 0.86 8.63–12.01
12 9.69 + 0.87 7.98–11.4
24 18.39 + 1.49c 15.47–21.31
ZnO bulk
6 12.82 + 0.99 10.87–14.77
12 11.30 + 0.90 9.54–13.06
24 10.75 + 0.94 8.91–12.59
Figure 5. Lipid peroxidation, measured as TBARS accumu-
CI: confidence interval; ZnONP: zinc oxide nanoparticles; EMS:
lation in Drosophila melanogaster when treated with various
ethyl methane sulphonate.
a
Mean + standard error based on experiments. concentrations of ZnONP and ZnO bulk. Bars show MDA
b
95% CI for % of DNA in tail. average (mg/kg FW) standard deviation (n ¼ 3). Different
c
p < 0.01 versus negative control. lowercase letters indicate differences between the same
treatments in the different concentrations (p ! 0.05).
Asterisks indicate significant differences (p ! 0.05) between
treatments at the same concentrations. ZnONP: zinc
oxide nanoparticle; ZnO: zinc oxide; MDA: malondialde-
hyde; TBARS: thiobarbituric acid reactive substance.

response relationship. When MDA induction was


compared between both ZnO forms at the same con-
centration, ZnONP induced higher MDA levels than
ZnO microparticles for all evaluated concentrations
(Figure 5, p ! 0.05).

Discussion
ZnONP is one of the manufactured NM more widely
produced and may be found in a wide variety of con-
Figure 4. Comet assay images showing genetic damage sumer products available in the market, especially in
induced by ZnONP in haemocytes of Drosophila ($400 cosmetics and sunscreens. Thus, the probability of
magnification). ZnONP: zinc oxide nanoparticle. human and environmental exposure to ZnONP is
increasing, and its genotoxic potential risk must be
concentrations ranging from 6 mM to 24 mM. After exhaustively studied. Despite the high production and
the treatments, larval tissue of D. melanogaster was wide use of ZnONP, there are few studies covering
used to evaluate the MDA marker for oxidative stress. their in vivo potential genotoxicity (Kumar and Dha-
MDA values after ZnONP treatments increased sig- wan, 2013).
nificantly at all concentrations in comparison with the In the present study, we used commercially avail-
values observed in the negative control (p ! 0.05). able, manufactured ZnONP with the average particle
However, this increase in MDA values does not fol- size, given by the manufacturer (Sigma-Aldrich), of
low a linear dose–response because a notable reduc- approximately 100 nm. This particle size was slightly
tion of MDA concentration was observed at the lower to that was obtained in the TEM analysis car-
higher concentration of ZnONP evaluated (24 mM; ried out in this study, where ZnO nanopowder reached
Figure 5). ZnO bulk showed a significant increase an average of approximately 107 nm. According to
of MDA levels in treated larvae with a linear dose– the DLS results, the hydrodynamic size of ZnONP

Downloaded from tih.sagepub.com at University of York on January 24, 2016


10 Toxicology and Industrial Health

was larger than TEM analyses, with an average of stress assessment of ZnONP in Drosophila. The
approximately 225 nm, suggesting a certain grade of results obtained with the wing-spot test after chronic
agglomeration in aqueous media. These results are exposure to NP and microparticles of ZnO at the same
in agreement with other similar in vivo genotoxic concentrations clearly indicate that both ZnO forms
studies, where ZnONP showed sizes greater than were unable to induce genotoxic effects, as
200 nm in aqueous media and sizes less than 100 an insignificant increase of mutant clone formations
nm in dry state with TEM analyses (Nounou et al., were observed for all concentrations assayed in wing
2013; Sharma et al., 2012b). These differences are preparations. Unlike the absence of genotoxicity
commonly explained by the tendency of NP to within the wing-spot test, the results obtained with the
agglomerate in aqueous medium, making them larger in vivo comet assay in haemocytes of Drosophila
than in dry state (Dhawan and Sharma, 2010; Sharma showed that ZnONP treatments can induce DNA
et al., 2012b). Taking into account the importance of strand breaks in circulating blood cells, but at the
size in the biological properties of NP, the size char- highest concentration tested. In addition, in the lipid
acterization by DLS technique may be considered peroxidation assay, ZnONP showed higher MDA
indispensable and more relevant compared to TEM induction than its bulk form, suggesting that DNA
analyses, as it can measure size under conditions that damage observed with the comet assay could be
more closely resemble the exposure conditions in mediated by oxidative stress induction. The contra-
Drosophila experiments (Sharma et al., 2009). dictory results obtained in our study with both geno-
The zeta potential of ZnONP measured with toxic tests in Drosophila could be explained in part
LDV technique reached an average of approximately by the weak or low genotoxic potential of ZnONP
&21 in water suspension. This measure indicates the and/or the differences of genotoxic end points used
moderate colloidal stability (i.e. resistance to agglom- in the wing-spot and comet tests for measuring DNA
eration in water) of ZnONP, indicating suitable damage (mutation/recombination vs. primary DNA
exposure conditions for nanogenotoxic assays with damage).
Drosophila. In general, there are few genotoxic studies specifi-
To increase our knowledge of the potential geno- cally for ZnONP, on whole organisms (i.e. in vivo
toxic risk associated with ZnONP, we used two approaches). From the available literature, the geno-
well-known in vivo genotoxic assays using somatic toxicity of nanosized ZnO could be considered weak,
cells of Drosophila: the wing-spot test and the comet as mainly negative and weak genotoxic potential has
assay on haemocytes. The wing-spot assay can detect been reported in the literature (Chibber et al., 2013).
a wide range of mutational events, as well as mitotic Although most of the in vitro studies have indicated
recombination, a relevant process related with carci- genotoxic effects associated with ZnONP exposure
nogenesis (Luo et al., 2000). In addition, this system (such as DNA strand breaks, micronuclei formation
assay has already been demonstrated to be an excel- and chromosomes breaks in human cells in culture
lent tool to assess genotoxic effects of different NM, (Alarifi et al., 2013; Bhattacharya et al., 2014; Demir
including silver and cobalt NP (Demir et al., 2011; et al., 2014; Hackenberg et al., 2011; Lin et al., 2009;
Vales et al., 2012); titanium, zirconium and alumi- Sharma et al., 2009, 2012b; Valdiglesias et al., 2013)),
nium metal oxide NP (Demir et al., 2013); and most of the studies carried out with bacteria indicate
multi-walled carbon nanotubes (de Andrade et al., negative or weak mutagenic effects (Kwon et al.,
2014; Machado et al., 2013). 2014; Nam et al., 2013; Pan et al., 2010). In addition,
The comet assay is one of the most widely recent in vivo studies with rodents indicate clearly that
employed methods of in vivo genotoxicity screening ZnONP with different sizes (20 and 70 nm) and elec-
because it can be applied to any type of cell, it can tric charges (i.e. anion and cationic NP) were unable
detect low levels of DNA damage and it can detect to induce DNA strand breaks and micronucleus for-
different kinds of damage (Hartmann et al., 2003). mation in rats and mice orally exposed at different
Thus, given these advantages, this assay is being rou- concentrations and at times of nanosized ZnO (Kwon
tinely used to assess DNA damage of NM in different et al., 2014). In this context, the absence of genotoxic
cell and organism models (Karlsson, 2010), including effects found in this recent study agree with our neg-
D. melanogaster (Sabella et al., 2011). ative data for the wing-spot test, suggesting that
To the best of our knowledge, our results are the ZnONP does not show significant genotoxic potential
first to be reported on the genotoxic and oxidative on whole organisms.

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 11

The results obtained with the alkaline comet assay (e.g. superoxide dismutase and glutathione) have been
in haemolymph cells of Drosophila indicate that proposed as one of the main mechanisms involved in
ZnONP exposure can induce genetic damage. Never- toxic and genotoxic manifestations induced by nanos-
theless, the low induction and the fact that this effect cale particles (Shrivastava et al., 2014; Wang et al.,
was only observed at the highest concentration evalu- 2013). NP can generate ROS in the cells that may
ated does not support a genotoxic potential for cause indirect oxidative damage to DNA through free
ZnONP. Similar results have been reported with the radical attack. ZnONP were found to generate free
same test system in whole animals and have shown radicals and oxidative stress, which lead to oxidative
that ZnONP can exert DNA damage in kidney and DNA damage in vivo and suggest a probable mechan-
liver cells of mice (Sharma et al., 2012b) and induce ism of genotoxicity in liver and lung cells of rodents
DNA fragmentation and DNA breakages in lung cells (Nounou et al., 2013; Sharma et al., 2012b). This
of rats exposed orally to different concentrations of effect has also been observed in several studies with
ZnONP (Nounou et al., 2013). Likewise, another in human and mammals cells in culture, where ZnONP
vivo study on earthworms (Eisenia fetida) indicates exposure induced high ROS production; oxidative
that nanosized ZnO show significant primary DNA DNA lesions; and depletion of glutathione, superox-
damage in coelomocytes at the highest dose tested ide dismutase and catalase antioxidants (Alarifi
(Hu et al., 2010). Thus, it seems that ZnONP is able et al., 2013; Demir et al., 2014; Lin et al., 2009;
to induce primary DNA damage in different cells and Sharma et al., 2012a; Valdiglesias et al., 2013). In our
living organisms, but this happens under specific con- study, we found that although ZnONP can induce
ditions of toxicity (Magdolenova et al., 2014). This is lipid peroxidation, this effect was also observed after
in agreement with our data from the comet assay. exposure to bulk ZnO, but at a lesser level. Thus, this
Some studies indicate that ZnONP dissolution in effect is not an intrinsic characteristic of ZnONP.
aqueous solutions, and the released Zn ions can play It is important to note that in our research we used
a significant role in their toxicity (Wang et al., two genotoxic assays with different genetic end points
2014; Xia et al., 2008). However, other studies indi- to assess the genotoxicity of nanoparticulated ZnO in
cate that the role of zinc (Zn) ions release in the toxi- Drosophila. It is also important to state that the
city of ZnO materials may be irrelevant. For example, genetic damage detected in the alkaline comet assay
Alarifi et al. (2013) found that Zn ions can induce toxi- is primary DNA damage that can be easily repaired
city and DNA damage in human cells A375, but these in the cells as to avoid mutations (Azqueta and Col-
effects were less significant in comparison with similar lins, 2013; Bajpayee et al., 2013). On the contrary, the
doses of ZnONP. Sharma et al. (2012a) suggested that Drosophila wing-spot test determines fixed DNA
although the ZnONP released Zn ions into the media damage (mutation and/or recombination) generated
test, this amount was insufficient to cause toxicity after DNA lesions are misrepaired by DNA repair sys-
in human liver cells. Similarly, Valdiglesias et al. tems in somatic cells (Graf et al., 1984; Garcı́a Sar
(2013) found low Zn ions release and the cytotoxicity et al., 2012; Garcia-Quispes et al., 2009). Therefore,
was restricted to high doses of ZnONP in mouse neu- contradictory genotoxic effects observed between
ronal cells. Finally, Lin et al. (2009) evidenced that the wing-spot and comet assays could be explained by the
Zn ions release in the media test was very low to induce inability of the ZnONP to induce somatic mutation
cytotoxicity in human lung epithelial cells. (i.e. fixed genetic damage) in the wing cells of Droso-
Although the genotoxic effects of Zn ions was not phila. This would agree with the results of other
measured in our study with Drosophila, a previous researchers using mice and rat models, where ZnONP
work indicated that the soluble ionic compound zinc have failed to induce micronucleus formation follow-
nitrate showed inconclusive results in the Drosophila ing validated and standard methods for genotoxicity
wing spot-test (Yeşilada, 2001). This data suggest that testing (Landsiedel et al., 2010; Kwon et al., 2014).
Zn ions do not induce significant genotoxic effects in According to our results, we conclude that ZnONP
somatic cells of Drosophila and support our conclu- were unable to induce mutation and recombinational
sions that the genetic damage of ZnONP observed events in somatic cells of Drosophila (absence of
with the comet assay in Drosophila blood cells can fixed DNA damage). However, this NM can induce
be related with the NP per se. primary DNA damage in haemolymph cells of Droso-
Oxidative stress mediated by ROS production and phila under a higher dose scenario, as evidenced with
depletion of the antioxidant barrier into the cells the comet assay. As demonstrated, this genetic damage

Downloaded from tih.sagepub.com at University of York on January 24, 2016


12 Toxicology and Industrial Health

could be mediated by oxidative stress. The overall Bajpayee M, Kumar A and Dhawan A (2013) The comet
data suggest that the genotoxic potential of ZnONP assay: assessment of in vitro and in vivo DNA damage.
could be considered low or weak because it promotes In: Dhawan A and Bajpayee M (eds) Genotoxicity
only primary genetic damage restricted to high-dose assessment. New York: Humana Press, pp. 325–345.
exposures, which can be easily repaired. This study Bhattacharya D, Santra CR, Ghosh AN, et al. (2014)
demonstrates the usefulness of using more than one Differential toxicity of rod and spherical zinc oxide
genotoxic test in the evaluation of the genotoxic nanoparticles on human peripheral blood mononuc-
potential of NM. Finally, it is important to remark that lear cells. Journal of Biomedical Nanotechnology
this study is the first to assess the genotoxic effects 10: 707–716.
and oxidative stress of nano and bulk ZnO particles Bolker BM, Brooks ME, Clark CJ, et al. (2009) General-
in Drosophila. ized linear mixed models: a practical guide for ecology
and evolution. Trends in Ecology & Evolution 24:
Acknowledgements 127–135.
We are grateful to MJ Rivadeneira, MM Carmona and M Borm PJ, Robbins D, Haubold S, et al. (2006) The potential
Garrison for revising the English version of the manuscript. risks of nanomaterials: a review carried out for ECE-
TOC. Particle and Fibre Toxicology 3: 11.
Declaration of Conflicting Interests Buzea C, Pacheco II and Robbie K (2007) Nanomaterials
The author(s) declared no potential conflicts of interest and nanoparticles: sources and toxicity. Biointerphases
with respect to the research, authorship and/or publication 2: MR17–MR71.
of this article. Carmona ER, Creus A and Marcos R (2011a) Genotoxicity
testing of two lead-compounds in somatic cells of
Funding Drosophila melanogaster. Mutation Research/Genetic
The author(s) disclosed receipt of the following financial Toxicology and Environmental Mutagenesis 24: 35–40.
support for the research, authorship and/or publication of Carmona ER, Guecheva TN, Creus A, et al. (2011b) Pro-
this article: The authors thank the financial support given posal of an in vivo comet assay using haemocytes of
by FONDECYT-CONICYT 11110181 project, Dirección Drosophila melanogaster. Environmental and Molecu-
General de Investigación y Postgrado, Universidad Cató- lar Mutagenesis 52: 165–169.
lica de Temuco, DGIP UCT CD 2010-01 project, and Chibber S, Ansari SA and Satar R (2013) New vision to
MECESUP UCT 0804 project. CuO, ZnO, and TiO2 nanoparticles: their outcome and
effects. Journal of Nanoparticle Research 15: 1–13.
References Cho W-S, Duffin R, Poland CA, et al. (2010) Metal oxide
Ahamed M, Posgai R, Gorey TJ, et al. (2010) Silver nano- nanoparticles induce unique inflammatory footprints in
particles induced heat shock protein 70, oxidative stress the lung: important implications for nanoparticle testing.
and apoptosis in Drosophila melanogaster. Toxicology Environmental Health Perspectives 118: 1699–1706.
and Applied Pharmacology 242: 263–269. de Andrade LR, Sandin Brito A, de Souza Melero AMG,
Alarifi S, Ali D, Alkahtani S, et al. (2013) Induction of oxi- et al. (2014) Absence of mutagenic and recombinagenic
dative stress, DNA damage, and apoptosis in a malig- activity of multi-walled carbon nanotubes in the Droso-
nant human skin melanoma cell line after exposure to phila wing-spot test and Allium cepa test. Ecotoxicology
zinc oxide nanoparticles. International Journal of Nano- and Environmental Safety 99: 92–97.
medicine 8: 983–993. Demir E (2012) In vivo genotoxicity assessment of diflu-
Alonso-Moraga A and Graf U (1989) Genotoxicity testing benzuron and spinosad in Drosophila melanogaster with
of antiparasitic nitrofurans in the Drosophila wing the comet assay using haemocytes and the smart assay.
somatic mutation and recombination test. Mutagenesis Fresenius Environmental Bulletin 21: 3894–3900.
4: 105–110. Demir E, Akça H, Kaya B, et al. (2014) Zinc oxide nano-
Arossi GA, Lehmann M, Dihl RR, et al. (2010) Induced particles: genotoxicity, interactions with UV-light and
DNA damage by dental resin monomers in somatic cell-transforming potential. Journal of Hazardous Mate-
cells. Basic & Clinical Pharmacology & Toxicology rials 264: 420–429.
106: 124–129. Demir E and Kaya B (2013) Studies on the genotoxic prop-
Azqueta A and Collins AR (2013) The essential comet erties of four benzyl derivatives in the in vivo comet
assay: a comprehensive guide to measuring DNA dam- assay using haemocytes of Drosophila melanogaster.
age and repair. Archives of Toxicology 87: 949–968. Fresenius Environmental Bulletin 22: 1590–1596.

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 13

Demir E, Turna F, Vales G, et al. (2013) In vivo genotoxi- test of Drosophila melanogaster. Toxicology and Indus-
city assessment of titanium, zirconium and aluminium trial Health 31(3): 261–267.
nanoparticles, and their microparticulated forms, in Dro- Hackenberg S, Scherzed A, Technau A, et al. (2011) Cyto-
sophila. Chemosphere 93: 2304–2310. toxic, genotoxic and pro-inflammatory effects of zinc
Demir E, Vales G, Kaya B, et al. (2011) Genotoxic analysis oxide nanoparticles in human nasal mucosa cells in
of silver nanoparticles in Drosophila. Nanotoxicology 5: vitro. Toxicology In Vitro 25: 657–663.
417–424. Halliwell B and Chirico S (1993) Lipid peroxidation: its
Dhawan A, Bajpayee M and Parmar D (2009) Comet assay: mechanism, measurement, and significance. The Amer-
a reliable tool for the assessment of DNA damage in dif- ican Journal of Clinical Nutrition 57: 715S–724S.
ferent models. Cell Biology and Toxicology 25: 5–32. Hansen SF, Baun A, Michelson E, et al. (2009) Nanomater-
Dhawan A and Sharma V (2010) Toxicity assessment of ials in consumer products. In: Linkow I and Steevens J
nanomaterials: methods and challenges. Analytical and (eds) Nanomaterials: Risks and Benefits. Netherlands:
Bioanalytical Chemistry 398: 589–605. Springer, pp. 359–367.
Espitia P, Soares ND, Coimbra JD, et al. (2012) Zinc oxide Hartmann A, Agurell E, Beevers C, et al. (2003) Recom-
nanoparticles: synthesis, antimicrobial activity and food mendations for conducting the in vivo alkaline comet
packaging applications. Food and Bioprocess Technol- assay. Mutagenesis 18: 45–51.
ogy 5: 1447–1464. Hu C, Li M, Cui Y, et al. (2010) Toxicological effects of
Fairbrother H (2010) Environmental, health, and safety TiO2 and ZnO nanoparticles in soil on earthworm Eise-
effects of engineered nanomaterials: challenges and nia fetida. Soil Biology and Biochemistry 42: 586–591.
research needs. Presented at the SPIE Defense, Security, Karadeniz A, Kaya B, Savaş B, et al. (2011) Effects of
and Sensing Conference on International Society for two plant growth regulators, indole-3-acetic acid and
Optics and Photonics, Orlando, Florida, USA, 5–9 April "-naphthoxyacetic acid, on genotoxicity in Drosophila
2010, No. 766618. SMART assay and on proliferation and viability of
Frei H and Würgler FE (1988) Statistical methods to decide HEK293 cells from the perspective of carcinogenesis.
whether mutagenicity test data from Drosophila assays Toxicology and Industrial Health 27: 840–848.
indicate a positive, negative, or inconclusive result. Karlsson HL (2010) The comet assay in nanotoxicology
Mutation Research/Environmental Mutagenesis and research. Analytical and Bioanalytical Chemistry 398:
Related Subjects 203: 297–308. 651–666.
Garcı́a Sar D, Aguado L, Montes Bayón M, et al. (2012) Kastenbaum MA and Bowman KO (1970) Tables for deter-
Relationships between cisplatin-induced adducts and mining the statistical significance of mutation frequen-
DNA strand-breaks, mutation and recombination in vivo cies. Mutation Research 9: 527–549: Medium: X.
in somatic cells of Drosophila melanogaster, under dif- Kumar A and Dhawan A (2013) Genotoxic and carcino-
ferent conditions of nucleotide excision repair. Mutation genic potential of engineered nanoparticles: an update.
Research/Genetic Toxicology and Environmental Muta- Archives of Toxicology 87: 1883–1900.
genesis 741: 81–88. Kumar A, Pandey AK, Singh SS, et al. (2011a) Cellular
Garcia-Quispes WA, Carmona ER, Creus A, et al. (2009) uptake and mutagenic potential of metal oxide nanopar-
Genotoxic evaluation of two halonitromethane disinfec- ticles in bacterial cells. Chemosphere 83: 1124–1132.
tion by-products in the Drosophila wing-spot test. Che- Kumar A, Pandey AK, Singh SS, et al. (2011b) Engineered
mosphere 75: 906–909. ZnO and TiO2 nanoparticles induce oxidative stress
Gorth DJ, Rand DM and Webster TJ (2011) Silver nanopar- and DNA damage leading to reduced viability of
ticle toxicity in Drosophila: size does matter. Interna- Escherichia coli. Free Radical Biology and Medicine
tional Journal of Nanomedicine 6: 343–350. 51: 1872–1881.
Graf U, Würgler F, Katz A, et al. (1984) Somatic mutation Kumaravel TS, Vilhar B, Faux S, et al. (2009) Comet assay
and recombination test in Drosophila melanogaster. measurements: a perspective. Cell Biology and Toxicol-
Environmental Mutagenesis 6: 153–188. ogy 25: 53–64.
Gürbüzel M, Çapoglu I, Kizilet H, et al. (2012) Genotoxic Kwon JY, Lee SY, Koedrith P, et al. (2014) Lack of geno-
evaluation of two oral antidiabetic agents in the Droso- toxic potential of ZnO nanoparticles in in vitro and in
phila wing spot test. Toxicology and Industrial Health vivo tests. Mutation Research/Genetic Toxicology and
30(4): 376–383. Environmental Mutagenesis 761: 1–9.
Gürbüzel M, Uysal H and Kizilet H (2013) Assessment of Landsiedel R, Ma-Hock L, Van Ravenzwaay B, et al.
genotoxic potential of two mycotoxins in the wing spot (2010) Gene toxicity studies on titanium dioxide and

Downloaded from tih.sagepub.com at University of York on January 24, 2016


14 Toxicology and Industrial Health

zinc oxide nanomaterials used for UV-protection in cos- Pasupuleti S, Alapati S, Ganapathy S, et al. (2012) Toxicity
metic formulations. Nanotoxicology 4: 364–381. of zinc oxide nanoparticles through oral route. Toxicol-
Leeuw TK, Reith RM, Simonette RA, et al. (2007) Single- ogy and Industrial Health 28: 675–686.
walled carbon nanotubes in the intact organism: near-IR Philbrook NA, Winn LM, Afrooz ARMN, et al. (2011)
imaging and biocompatibility studies in Drosophila. The effect of TiO2 and Ag nanoparticles on reproduc-
Nano Letters 7: 2650–2654. tion and development of Drosophila melanogaster and
Lin W, Xu Y, Huang C-C, et al. (2009) Toxicity of nano- CD-1 mice. Toxicology and Applied Pharmacology
and micro-sized ZnO particles in human lung epithelial 257: 429–436.
cells. Journal of Nanoparticle Research 11: 25–39. Pompa PP, Vecchio G, Galeone A, et al. (2011) In vivo
Lindsley DL and Zimm GG (1992) The Genome of toxicity assessment of gold nanoparticles in Drosophila
Drosophila Melanogaster. San Diego: Academic Press, melanogaster. Nano Research 4: 405–413.
p. 1133. Posgai R, Cipolla-McCulloch CB, Murphy KR, et al.
Liu X, Vinson D, Abt D, et al. (2009) Differential toxicity (2011) Differential toxicity of silver and titanium diox-
of carbon nanomaterials in Drosophila: larval dietary ide nanoparticles on Drosophila melanogaster develop-
uptake is benign, but adult exposure causes locomotor ment, reproductive effort, and viability: size, coatings
impairment and mortality. Environmental Science & and antioxidants matter. Chemosphere 85: 34–42.
Technology 43: 6357–6363. Raj S, Jose S, Sumod U, et al. (2012) Nanotechnology in
Luo G, Santoro IM, McDaniel LD, et al. (2000) Cancer pre- cosmetics: opportunities and challenges. Journal of
disposition caused by elevated mitotic recombination in Pharmacy & Bioallied Sciences 4: 186–193.
Bloom mice. Nature Genetics 26: 424–429. Sabella S, Brunetti V, Vecchio G, et al. (2011) Toxicity of
Machado N, Lopes J, Saturnino R, et al. (2013) Lack of citrate-capped AuNPs: an in vitro and in vivo assess-
mutagenic effect by multi-walled functionalized carbon ment. Journal of Nanoparticle Research 13: 6821–6835.
nanotubes in the somatic cells of Drosophila melanoga- Savolainen K, Alenius H, Norppa H, et al. (2010) Risk
ster. Food and Chemical Toxicology 62: 355–360. assessment of engineered nanomaterials and nano-
Magdolenova Z, Collins A, Kumar A, et al. (2014) technologies–a review. Toxicology 269: 92–104.
Mechanisms of genotoxicity. A review of in vitro and Sengstag C (1994) The role of mitotic recombination in
in vivo studies with engineered nanoparticles. Nanotox- carcinogenesis. CRC Critical Reviews in Toxicology
icology 8: 233–278. 24: 323–353.
Marcos R and Carmona ER (2013) The wing-spot and Sharma V, Anderson D and Dhawan A (2012a) Zinc oxide
the comet tests as useful assays detecting genotoxi- nanoparticles induce oxidative DNA damage and ROS-
city in Drosophila. Methods in Molecular Biology triggered mitochondria mediated apoptosis in human
1044: 417–427. liver cells (HepG2). Apoptosis 17: 852–870.
Mukhopadhyay I, Chowdhuri DK, Bajpayee M, et al. Sharma V, Shukla RK, Saxena N, et al. (2009) DNA dama-
(2004) Evaluation of in vivo genotoxicity of cyperme- ging potential of zinc oxide nanoparticles in human epi-
thrin in Drosophila melanogaster using the alkaline dermal cells. Toxicology Letters 185: 211–218.
comet assay. Mutagenesis 19: 85–90. Sharma V, Singh P, Pandey AK, et al. (2012b) Induction of
Nam SH, Kim SW and An YJ (2013) No evidence of the oxidative stress, DNA damage and apoptosis in mouse
genotoxic potential of gold, silver, zinc oxide and tita- liver after sub-acute oral exposure to zinc oxide nano-
nium dioxide nanoparticles in the SOS chromotest. particles. Mutation Research/Genetic Toxicology and
Journal of Applied Toxicology 33: 1061–1069. Environmental Mutagenesis 745: 84–91.
Nounou H, Attia H, Shalaby M, et al. (2013) Oral exposure Shrivastava R, Raza S, Yadav A, et al. (2014) Effects of
to zinc oxide nanoparticles induced oxidative damage, sub-acute exposure to TiO2, ZnO and Al2O3 nanoparti-
inflammation and genotoxicity in rat’s lung. Life Sci- cles on oxidative stress and histological changes in
ence Journal 10: 1969–1979. mouse liver and brain. Drug and Chemical Toxicology
Osmond MJ and McCall MJ (2010) Zinc oxide nanoparti- 37(3): 336–347.
cles in modern sunscreens: an analysis of potential expo- Singh NP, McCoy MT, Tice RR, et al. (1988) A simple
sure and hazard. Nanotoxicology 4: 15–41. technique for quantitation of low levels of DNA damage
Pan X, Redding JE, Wiley PA, et al. (2010) Mutageni- in individual cells. Experimental Cell Research 175:
city evaluation of metal oxide nanoparticles by the 184–191.
bacterial reverse mutation assay. Chemosphere 79: Tice R, Agurell E, Anderson D, et al. (2000) Single cell gel/
113–116. comet assay: guidelines for in vitro and in vivo genetic

Downloaded from tih.sagepub.com at University of York on January 24, 2016


Carmona et al. 15

toxicology testing. Environmental and Molecular Muta- contaminants. Mutation Research/Genetic Toxicology
genesis 35: 206–221. and Environmental Mutagenesis 605: 78–86.
Tironi VA, Tomás MC and Añón MC (2007) Lipid and Wang H, Wu F, Meng W, et al. (2013) Engineered nano-
protein deterioration during the chilled storage of particles may induce genotoxicity. Environmental Sci-
minced sea salmon (Pseudopercis semifasciata). Jour- ence & Technology 47: 13212–13214.
nal of the Science of Food and Agriculture 87: Wang B, Zhang Y, Mao Z, et al. (2014) Toxicity of ZnO
2239–2246. nanoparticles to macrophages due to cell uptake and
Valdiglesias V, Costa C, Kiliç G, et al. (2013) Neuronal intracellular release of zinc ions. Journal of Nanoscience
cytotoxicity and genotoxicity induced by zinc oxide and Nanotechnology 14(8): 5688–5696.
nanoparticles. Environment International 55: 92–100. Xia T, Kovochich M, Liong M, et al. (2008) Comparison of
Vales G, Demir E, Kaya B, et al. (2012) Genotoxicity of the mechanism of toxicity of zinc oxide and cerium
cobalt nanoparticles and ions in Drosophila. Nanotoxi- oxide nanoparticles based on dissolution and oxidative
cology 7: 462–468. stress properties. ACS Nano 2(10): 2121–2134.
Vecchio G, Galeone A, Brunetti V, et al. (2012) Mutagenic Xia T, Li N and Nel AE (2009) Potential health impact of
effects of gold nanoparticles induce aberrant phenotypes nanoparticles. Annual Review of Public Health 30:
in Drosophila melanogaster. Nanomedicine: Nanotech- 137–150.
nology, Biology and Medicine 8: 1–7. Yeşilada E (2001) Genotoxicity testing of some metals in
Villela IV, de Oliveira IM, da Silva J, et al. (2006) DNA the Drosophila wing somatic mutation and recombina-
damage and repair in haemolymph cells of golden mus- tion test. Bulletin of Environmental Contamination and
sel (Limnoperna fortunei) exposed to environmental Toxicology 66(4): 464–469.

Downloaded from tih.sagepub.com at University of York on January 24, 2016

You might also like