You are on page 1of 13

Mutagenesis vol. 28 no. 3 pp. 287–299, 2013 doi:10.

1093/mutage/get014
Advance Access publication 5 March 2013

Multiple cytotoxic and genotoxic effects induced in vitro by differently shaped copper
oxide nanomaterials

Sebastiano Di Bucchianico1,*, Maria Rita Fabbrizi1, ratios have a differential impact on cellular function, including
Superb K. Misra2,3, Eugenia Valsami-Jones2,3, cell proliferation, apoptosis, cytoskeleton formation, adhesion
Deborah Berhanu2, Paul Reip4, Enrico Bergamaschi5 and and migration (1).

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Lucia Migliore1 One of the most studied types of nanomaterials is CuO and
1
Department of Translational Research and New Technologies in Medicine a number of studies have clearly demonstrated significant
and Surgery, Medical Genetics, University of Pisa, Pisa, Italy, 2Department cytogenotoxic responses both in vivo and in vitro, affecting the
of Earth Sciences, Natural History Museum, London, UK, 3School of normal functions of the nervous system (2). CuO nanoparti-
Geography, Earth and Environmental Sciences, University of Birmingham, cles (NPs) can also be cytotoxic in a concentration- and time-
Edgbaston, UK, 4Intrinsiq Materials Ltd, Cody Technology Park, Ively Road,
Farnborough, Hampshire, UK and 5Department of Clinical and Experimental
dependent manner and elicit permeability and inflammation
Medicine, University of Parma, Italy responses in human cardiac microvascular endothelial cells
*To whom correspondence should be addressed. Tel: +39 050 2211029; Fax:
in vitro (3). Moreover, CuO NPs induce genotoxic responses
+39 050 2211034; Email: sebastiano.dibucchianico@for.unipi.it through the p53 pathway in human pulmonary epithelial cells
A549 and induce oxidative stress in a dose-dependent way, as
Received on September 27, 2012; revised on December 19, 2012; accepted on indicated by depletion of glutathione and induction of lipid
January 1, 2013
peroxidation, catalase and superoxide dismutase (4). CuO
In nanotoxicology, the capacity of nanoparticles of the same NPs, like many metal oxide NPs, exhibit low mutagenic poten-
composition but different shape to induce cytotoxicity and tial toward Salmonella typhimurium. However, a dose-depend-
genotoxicity is largely unknown. A series of cytotoxic and ent inhibition of Escherichia coli WP2 was found after CuO
genotoxic responses following in vitro exposure to differently NPs exposure at concentrations ranging from 100 to 1600 μg/
shaped CuO nanoparticles (CuO NPs, mass concentrations plate (5). In vivo studies in rats have shown hepatotoxicity
from 0.1 to 100 μg/ml) were assessed in murine macrophages and nephrotoxicity and an integrated metabolomic analysis
RAW 264.7 and in peripheral whole blood from healthy suggested that mitochondrial failure, enhanced ketogenesis,
volunteers. Cytotoxicity, cytostasis and genotoxicity were fatty acid β-oxidation and glycolysis contributed to the toxic-
evaluated by the colorimetric assay of formazan reduction ity induced by CuO NPs (6). Moreover, CuO NPs were found
[3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyl tetrazolium bro- to be inflammogenic to the lungs of rats and induced a unique
mide (MTT)] and by the cytokinesis-block micronucleus inflammatory footprint after both acute and chronic treatments.
cytome (CBMN Cyt) assay. The comet assay was applied Exposure to CuO NPs was severely associated with neutro-
for detecting DNA strand breaks and information on oxi- philic/eosinophilic cytotoxic inflammation and increased lev-
dative damage to DNA (oxidised purines and pyrimidines). els of interleukin (IL)-1β, macrophage inflammatory protein-2
The MTT assay revealed a decrease in cell viability in (MIP-2), eotaxin and lactate dehydrogenase in bronchoalveo-
RAW 264.7 cells and peripheral blood lymphocytes (PBL) lar lavage fluid, 24 h after instillation (7). It was also shown
with significant dose–effect relationships for the different that CuO NPs were, by far, the most potent when compared
CuO NP shapes. The comet assay revealed a dose-depend- with several other metal oxide NPs in inducing DNA damage;
ent increase in primary DNA damage, and a significant nevertheless, little is known about the comparative responses
increase in oxidative damage to DNA was also detectable, of differently shaped CuO NPs (8).
as well as increased frequency of micronuclei in binucleated The growing need for engineered nanostructured materials,
cells, often in a dose-related manner. Proliferative activity, to meet the vast number of applications that exploit their
cytotoxicity and apoptotic markers showed a significant unique physical and chemical properties, requires the study
trend in the two cell types. Finally, we have differentiated of their cytotoxic and genotoxic potential and necessitates an
clastogenic events from aneugenic events by fluorescence in improvement of current methods. One important aspect of the
situ hybridisation with human and murine pancentromeric cytogenotoxicity assays is the use of an appropriate NP-type
probes, revealing for the first time characteristic aneugenic positive control (9). In the current OECD guidelines, there
responses related to the shape of CuO NPs and cell type. is still no specific recommendation for NPs positive control.
Independently of size and shape, all CuO NPs revealed a A positive control is used to confirm the proper performance
clear-cut cytotoxic and genotoxic potential; this suggests of the assay, but must also include a control system for
that CuO NPs are good candidates for positive controls in nanosized compounds, so that the comparisons that are made
nanotoxicology. are relevant. Indeed, the lack of a nanoparticle-type positive
control questions the suitability of the tests to identify
Introduction nanomaterials with cytotoxic and genotoxic properties. ZnO
NPs have been often used as positive controls but they have
Nanomaterials may exhibit unusual interactions with cells not always given unequivocal results, emphasising the need
and tissues, thus necessitating studies to better understand for a more careful evaluation of ZnO NPs effects across a
the different effects on biological systems relative to the spectrum of relevant cell types (10). Similar considerations
composition, size and shape of these emerging nanomaterials. can be made for other types of NPs used as a positive control,
Likewise, it has been found that particles with dissimilar aspect such as CdO, CeO2 and carbon black NPs.

© The Author 2013. Published by Oxford University Press on behalf of the UK Environmental Mutagen Society. 287
All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.
S. Di Bucchianico et al.

In this article, we report on the cytotoxic and genotoxic a Nonius PDS 120 powder diffraction system position sensitive detector.
responses induced by CuO NPs of different morphologies in an Transmission electron microscopy (TEM) was used to image the particles and
assess the size and shape of the NPs and zeta potential to study the dispersion
in vitro model of murine airway macrophages (RAW 264.7), a of the NPs used. TEM imaging was done on a HITACHI H7100, operating at
commonly used macrophage model in toxicological studies on 100 kV. Zeta potential measurements were performed on the nanoparticle sus-
nanomaterials. Airway macrophages are specialised for inter- pension as well as the particles when suspended in cell culture media, using a
acting with inhaled particles and their exposure to NPs may Malvern Zetasizer Nano ZS instrument (Malvern Instruments, UK).
lead to distinct types of macrophage activation (11), which
account for different pathologic outcomes (12). We also used Cell culture and exposure to CuO NPs
human whole blood cultures from healthy volunteers, to mimic RAW 264.7 cells were purchased from the Istituto Zooprofilattico of Brescia
(Italy). The cells were cultured in minimal essential medium (CELBIO,
possible interactions with blood cells, the major route for NPs Milan, Italy) supplemented with 10%, heat inactivated, foetal bovine serum
biodistribution (13). The aim of this study was to demonstrate (CELBIO), 1% penicillin/streptomycin (CELBIO) and 1% l-glutamine

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


that, independently from geometrical characteristics, all CuO (CELBIO). The cell cultures were maintained in a humidified atmosphere
NPs show positive responses making CuO a good positive con- (5% CO2) at 37°C. After 24 h of incubation, the cell cultures were treated
with different mass concentrations of CuO NPs (0.01–100 μg/ml) for all
trol in nano-genotoxicological studies. cytotoxicity assays and from 0.1 to 10 μg/ml for the CBMN Cyt test and
We used CuO NPs from an industrial partner (Intrinsiq the comet assay. The study was also performed on human peripheral blood
Materials Ltd and here named INT) and three CuO NPs dif- lymphocytes of two healthy male volunteers. Approximately 4–6 ml of blood
fering in shapes and surface area (rods, spheres and spindles) was drawn by venipuncture in Li-heparin vials, according to standard proce-
synthesised within the frame of the European Commission dure. The donors were chosen according to the following criteria: young age,
non-smokers, without pharmacological treatments for at least 3 weeks before
Framework Programme 7 (FP7) NanoReTox project, to address donation and without any radiological examination performed within the pre-
the environmental and human implications of increased expo- vious 3 months. Blood was drawn from the volunteers and 600 μl of whole
sure to engineered NPs and to contribute to address major con- blood was used to prepare two paired independent lymphocyte cultures for
cerns in nanotoxicity by specifically investigating a possible each subject in 4.7 ml of complete medium per culture. Standard medium was
prepared with RPMI 1640 (Gibco BRL, Italy) supplemented with 10% foetal
correlation between physicochemical properties and toxicity. bovine serum (Gibco BRL), 1.5% phytohaemagglutinin (Gibco BRL) and 1%
Cytotoxic, primary and oxidative genotoxic effects of CuO penicillin–streptomycin (Gibco BRL). CuO INT powder was weighed and
NPs were evaluated by the formazan reduction [3-[4,5-dimeth- suspended in the appropriate culture medium, then sonicated for 10 min in a
ylthiazol-2-yl]-2,5-diphenyl tetrazolium bromide (MTT)] assay sonicating bath (FALC, Treviglio, Italy). CuO spheres, spindles and rods were
and the single-cell gel-electrophoresis (comet) assay in its clas- provided in aqueous suspensions by the Natural History Museum (NHM) of
London, UK, and stored at room temperature. Aliquots of the concentrated
sical version and in the modified one with the use of restric- aqueous suspension of CuO spheres, spindles and rods were directly added
tion enzymes. A key mechanism thought to be responsible for to the culture medium (not exceeding 1% v/v) and sonicated for 10 min, to
the genotoxic effects exerted by nanomaterials, in particular obtain the final concentrations to be tested, evaluated by means of trypan blue
by CuO NPs, involves an increased production of intracellular dye exclusion (standard method). All the experiments were performed within
7 days after the CuO NPs delivery from NHM to ensure the stability of the
reactive oxygen species (ROS). ROS-induced DNA damage is suspension as indicated by the supplier.
typified by single- and double-strand DNA breaks that can be
detected by the comet assay. Thus, we applied a modification MTT assay
of the assay, which allows to determine the presence of altered Cytotoxicity was assessed by using MTT assay. Mitochondrial dehydrogenases
purines, including 8-oxoguanine and oxidised pyrimidine, of viable cells reduce MTT to water insoluble blue formazan crystals that are
using specific bacterial restriction enzymes, namely forma- then solubilised by dimethylsulphoxide (DMSO); this assay thus indicates
midopyrimidine-DNA glycosylase (Fpg) and endonuclease- cell mitochondrial activity impairment. RAW 264.7 cells were grown to sub-
III (Endo-III) (14). Chromosomal aberrations were evaluated confluence in 96-well plates before being exposed to 0.01, 0.1, 1, 10 and
100 μg/ml CuO NPs suspensions for 2 and 24 h. Hydrogen peroxide (H2O2,
using the cytokinesis-block micronucleus cytome (CBMN Cyt) 10 μM) was used as a positive control. Whole blood cultures were carried out
assays. In recent years, the micronucleus (MN) test has evolved as described above and treated with the same conditions of adherent cells.
into a comprehensive approach aimed to evaluate at the same We collected the cells (3000 rpm, 5 min) and resuspended them in 0.5 ml of
time chromosome breakage, DNA misrepair, chromosome complete medium. Erythrocyte lysis buffer (3.5 ml) [155 mM NH4Cl, 10 mM
KHCO3, 1 mM Na2 ethylenediaminetetraacetic acid (EDTA), pH 7.4] was
loss, non-disjunction, necrosis, apoptosis and cytostasis (15). added and after 1.5 min cells were collected (3000 rpm, 5 min). The cells were
Finally, we applied the fluorescence in situ hybridisation (FISH) washed in 3.5 ml of complete medium once and then pelleted (3000 rpm,
with both human and mouse pancentromeric probes to differ- 5 min). The obtained pellet was resuspended in 3.5 ml of complete medium and
entiate aneugenic and/or clastogenic mechanisms involved in 100 μl of the cell suspension was put in a 96-well plate. After exposure of all
micronuclei formation. Aneuploidy is a major cause of human cell types (nine wells for each condition), 10 μl of a 5 mg/ml MTT solution was
added. After 3 h at 37°C, medium in adherent cell culture was then replaced by
reproductive failure and an important contributor to cancer and 100 μl of DMSO and mixed thoroughly to dissolve the formazan crystals; to
it is therefore important that any increase in its frequency due cell culture in suspension, 100 μl of DMSO was added without discarding the
to chemical exposures should be recognised and controlled (16) medium. To limit potential interactions due to the possible presence of residual
especially for nanomaterials for which an adequate legislation NPs that could interfere with the assay, parallel cultures without cells (blank)
were used to take into account the absorption due to nanomaterials. Then,
is still missing. absorbance was measured at 570 nm (630 nm background) and cell viability
was determined as a percentage of the negative control (unexposed cells)
minus the percentage of concurrent condition blank absorbance. IC50 values
Materials and methods (50% inhibitory concentration) were extrapolated graphically from the plotted
absorbance data.
Characterisation of CuO NPs
The main characteristics of the CuO NPs used in this study are summarised
in Table I. CuO powder (CuO INT) was obtained from a commercial source Cytokinesis-block micronucleus cytome assay
(Intrinsiq Materials Ltd, Farnborough, UK) with the particle size specified by The CBMN Cyt on whole blood treatments was performed according to the
the manufacturer in the range 10–100 nm. CuO INT was obtained by plasma procedure described by Migliore and co-authors (13,18). After 24 h of culture,
synthesis, whereas spheres, rods and spindles were obtained by reducing Cu2+ the cells were exposed to CuO NPs at concentrations ranging from 0.1 μg/ml
by OH− in presence of acetic acid, as described previously (17). X-ray dif- to 10 μg/ml. Mitomycin C (MMC; Kyowa Hakko Kogyo Co., Tokyo) (0.17 μg/
fraction (XRD) was used for the identification of the crystal structure using ml) was used as positive control. Cytochalasin B (6 μg/ml) was added after 44 h

288
Shape-specific CuO NP cytotoxicity and genotoxicity

to block the cytokinesis process and lymphocyte cultures were harvested after cultures, 10 μM hydrogen peroxide was used as a positive control. At least two
72 h. Cells were then treated with an hypotonic solution (0.075 M KCl) to lyse independent experiments were performed for each concentration tested and
erythrocytes for 2 min, prefixed in 3:5 methanol:acetic acid, washed once with incubation was performed for 2 or 24 h. After treatment, the cell pellet was used
methanol and subsequently fixed twice with a 6:1 methanol:acetic acid fixative for the comet assay. Briefly, the slides were spread with 1% normal melting
solution. Finally, the cell solution was dropped onto cold glass slides. The stain- agarose (Sigma–Aldrich, St Louis, MO, USA) in PBS (Sigma–Aldrich) and
ing procedure was performed by immersing the air-dried slides in a 2% Giemsa left to solidify at 4°C. Cell pellets were suspended in 85 μl of 0.5% low-
solution in Sorensen’s buffer (pH 6.8). To assess the genotoxic effects of CuO melting agarose (Agarose wide range, Sigma–Aldrich) in PBS at 37°C and
NPs in RAW 264.7 cells, the CBMN Cyt assay was performed as follows. two slides per culture were prepared. The cell suspension was dropped on top
Briefly, making the appropriate dilutions, a suitable quantity of cells (2 × 104 of the first agarose layer, covered with a coverslip and allowed to solidify at
cells/cm2) was seeded into six-well plates and two paired independent cultures 4°C. A final layer of 0.5% low-melting agarose was then added to the slide.
for each concentration were prepared. After 24 h, the macrophages were treated The lysis procedure was performed in cold lysis buffer (2.5 M NaCl, 100 mM
with CuO NPs as described earlier. MMC (0.17 μg/ml) was used as a positive Na2 EDTA, 10 mM Tris, 10% DMSO and 1% Triton X-100, pH 10.0) for 1 h
control. Forty-four hours after the initial preparation of the cultures, cytocha- at 4°C. The slides were then placed for 20 min in alkaline buffer (300 mM

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


lasin B was added to each test well to a final concentration of 4 μg/ml to block NaOH, 1 mM Na2 EDTA, pH > 13.0) and electrophoresed for 20 min at 25 V
cytokinesis of dividing cells. Cell cultures were then harvested after 28 h and adjusted at 300 mA. The slides were neutralised in 0.4 M Tris buffer, pH 7.5,
treated as described earlier. Two thousand binucleated cells were examined for for 5 min, twice and in methanol absolute for 5 min, once. Coded slides were
each experimental point in a blind mode (1000 from each independent culture scored after staining with ethidium bromide (20 μg/ml) using a fluorescence
replicate) using a Nikon Eclipse 800 optical microscope (final magnification microscope (Nikon Eclipse E800) at ×200 magnification. The percentage of
of ×400). The scoring criteria adopted by Fenech (15) were followed for each DNA in the comet tail in a total of 100 randomly selected cells per sample (two
endpoint. Briefly, we evaluated the binucleate micronucleated cells frequency replicates, each with 50 cells per slide) was used as a measure of the amount
as number of binucleate cells containing one or more micronuclei per 1000 of DNA damage. Analysis was carried out by using a Comet Image Analysis
binucleated cells and the total number of micronuclei in 1000 binucleated cells. System, version 5.5 (Kinetic Imaging, Nottingham, UK). Results were reported
Moreover, 500 cells were scored to evaluate the percentage of mono-, bi-, tri- as percentage of tail DNA. To determine the presence of oxidised pyrimidine
and multinucleated cells, and the CBPI (cytokinesis-block proliferation index) and purine bases, specific bacterial enzymes, namely Endo-III and Fpg,
was calculated as an index of cytotoxicity by comparing values in the treated respectively, were used. Two independent experiments were performed at 2 and
and control cultures. The CBPI indicates the average number of cell cycles per 24 h treatment. Before electrophoresis, the slides were washed with enzyme
cell during the period of exposure to cytoB and may be used to calculate cell buffer (3.72 × 10−3 g/ml Na2EDTA (Carlo Erba, Milan, Italy), 7.45 × 10−3 g/ml
proliferation. Finally, other damage events were scored in once-divided binu- KCl (Carlo Erba), 2.38 × 10−3 g/ml HEPES (Sigma–Aldrich) and 1 × 10−3 g/
cleated cells per 1000 cells, namely: (i) nucleoplasmic bridges, a biomarker of ml BSA (Sigma–Aldrich) (only for the Fpg enzyme) three times for 15 min at
DNA misrepair and/or telomere end-fusions and (ii) nuclear buds (NBUDs), a 4°C. Next, 100 μl of either buffer alone (reference slides) or 1.5 × 10–6 g/ml of
biomarker of elimination of amplified DNA and/or DNA repair complexes. The Endo-III or Fpg (Trevigen, Gaithersburg, MD, USA) was applied to each slide,
number of apoptotic and necrotic cells as defined by Fenech (15) and mitotic which was subsequently covered with a coverslip and incubated at 37°C for
figures per 500 cells were also evaluated. 45 min (Endo-III) or 30 min (Fpg) in humidified chambers. After the treatment,
slides were placed in a horizontal electrophoresis chamber and electrophoresis
was performed as described earlier. To detect oxidised purines or pyrimidines,
Fluorescence in situ hybridisation slides were incubated without enzymes (i.e. only buffer) and compared with
Separate slides from CBMN Cyt were used to achieve FISH with human those incubated with Fpg or Endo-III enzymes, respectively. To determine
and murine pancentromeric probes (Kreatech Diagnostics, Netherlands). the number of enzyme-sensitive sites, the difference between the value of the
Slides of treatments with 1 μg/ml of NPs, positive and negative control were percentage of tail DNA obtained after digestion with each enzyme and with the
examined for the presence of centromeric spots in MN and were classified as buffer only was calculated.
centromere positive (MN C+) and centromere negative (MN C−). Slides were
pretreated in 2× standard saline citrate (SSC)/0.5% Igepal, pH 7.0, at 37°C
Statistical analysis
for 15 min and dehydrated in a graded ethanol series (70, 85 and 100%);
slides and probes were co-denatured for 10 min at 75°C and overnight at All results are expressed in terms of the mean and standard error of the mean
37°C on a humidified chamber. Post-hybridisation washing was performed for two independent duplicated experiments, except where it is differently indi-
in 1× washing buffer (0.4 × SSC/0.3% Igepal) for 2 min at 72°C. Afterwards, cated. The statistical significance of the results was calculated using Student’s
slides were washed in 2 × SSC/0.1% Igepal for 1 min at room temperature t-test. The level of significance was set at P < 0.05. A linear regression equation
and dehydrated in a graded ethanol series (70, 85 and 100%) for 1 min each. was used to calculate the effect of mass concentration on each endpoints. IC50
The slides were counterstained with DAPI (0.1 μg/ml). For FISH analysis, values are graphically interpolated from the dose–effect curves. All statistical
500 binucleated (BN) cells were scored per culture. The MN in BN were analyses were performed using STATGRAPHICS Plus (SGWIN, version 2.1).
examined using the same scoring criteria described previously for the CBMN
Cyt assay. The preparation was examined with a Nikon Eclipse 800 micro-
scope (final magnification of ×400) equipped with a 100W mercury lamp and Results
triple band-pass filters (DAPI/FITC/Texas Red or DAPI/FITC/Rhodamine)
are used to view multiple colours and single band-pass filters are used for Characterisation of CuO NPs
individual colour visualisation. Table I summarises the main characteristics of CuO NPs used
in this study. The sizes of the primary particles are measured
Comet assay and oxidative damage to DNA from the TEM images (Figure 1); INT NPs appear highly
The modified alkaline comet assay (19) was carried out on RAW 264.7 cell polydispersed and particle sizes range from 10 to 100 nm
cultures seeded at a concentration of 2 × 104 cells/cm2 in six-well plates. After
attachment, cells were treated with CuO NPs at different concentrations (see
approximately, whereas the spheres, rods and spindles were
above). In whole blood cultures, after 24 h from the beginning of culture, monodispersed with sizes of 7 ± 1 nm, 7 ± 1 × 40 ± 10 nm
the cells were exposed to the same concentrations of CuO NPs. In all cell and 1200 ± 250 × 270 ± 50 × 30 ± 10 nm, respectively. Zeta

Table I. Characterisation of the various CuO NPs used in this study

Samples Initial form Size (nm) Zeta potential (mV) Surface areaa (m2/g) Dissolutiona (%)

CuO INT Powder 10–100 (polydisperse) 42 ± 1 NA 0.66


CuO spheres Suspension 7±1 38.6 ± 0.2 60 0.7
CuO rods Suspension 7 ± 1 × 40 ± 10 47.4 ± 0.1 51 1
CuO spindles Suspension 1200 ± 250 × 270 ± 50 × 30 ± 10 26.3 ± 0.8 18 3.5

NA, not available.


a
Data taken from S. K. Misra, S. Nuseibeh, A. Dybowska, D. Berhanu, T. D. Tetley and E. Valsami-Jones, submitted for publication. Experiments were performed
in 1 mM NaNO3 medium.

289
S. Di Bucchianico et al.

isoelectric point (between 9.0 and 10.0) to ensure a relatively


good colloidal stability (12). Data on surface area and percent
dissolution presented in Table I are largely discussed by S. K.
Misra, S. Nuseibeh, A. Dybowska, D. Berhanu, T. D. Tetley and
E. Valsami-Jones, submitted for publication, who performed
dissolution studies in 1 mM NaNO3 and also in serum-free cell
culture medium.
Cytotoxicity
The MTT assay was performed with 0.01, 0.1, 1, 10 and
100 μg/ml of CuO NPs but only the data obtained in 0.1, 1 and

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


10 μg/ml exposure are shown because the 0.01 dose was not
effective and the 100 dose was toxic for all tested NPs (percent
of viable cells <50).
CuO NPs exposure for 2 h at lowest tested concentrations
(Figure 3a) led to a reduction in cell viability more evident for
nanomaterials with smaller dimensions (spheres) in both cell
systems. The CuO spindles (the largest dimension), however,
generally showed a cytotoxicity trend in a dose-dependent
manner that was statistically significant for RAW 264.7 cells.
In peripheral blood lymphocytes, dose-dependent trends were
observed for the CuO spheres. After 24 h exposure (Figure 3b),
the two cell types showed a slightly different sensitivity only
to CuO spheres except for the highest dose of CuO rods and
Fig. 1. TEM images of (a) INT CuO, (b) CuO spheres, (c) CuO rods and
CuO INT. Phagocytic cells showed a dose-dependent cytotox-
(d) CuO spindles. icity trend both in CuO spheres exposure, CuO spindles and
CuO rods, with spheres and rods treatments the most effec-
potential indicates high suspension stability in deionised tive. The highest tested concentrations of spheres, spindles
water for INT (42 ± 1), spheres (38.6 ± 0.2), rods (47.4 ± 0.1) and rods were the most effective in inducing cytotoxicity also
except for spindles (26.3 ± 0.8). XRD patterns (Figure 2) in exposed blood cultures. The IC50 values are summarised in
show that CuO spheres, spindles, rods and INT samples Table II highlighting the most effective response induced by
contain tenorite copper oxide (ICDD 48-1548); no other CuO spheres in RAW 264.7 (94.8 μg/ml) and in blood cultures
material is detected. To assess the stability of the nanoparticle after 2 h exposure (22.6 μg/ml). Following 24 h exposure, the
suspension in cell culture media, zeta potential analysis was most effective responses in RAW 264.7 were similarly induced
conducted. Particles at a concentration of 0.1 mg/ml were by CuO spheres and rods (9.9 and 10.2 μg/ml, respectively);
added to RPMI medium and Dulbecco’s modified Eagle’s CuO spheres caused the most effective cytotoxicity in blood
medium and gently vortexed and pipetted for ensuring particle cultures (12.1 μg/ml). The IC50 values bring out a slight sen-
dispersion. Zeta potential measurements showed very similar sitivity of the phagocytic cells compared with lymphocytes.
values for all the four types of samples compared with the cell Considering results obtained in RAW 264.7 after 24 h expo-
culture medium alone (e.g. ZetaRPMI = –6.8 ± 0.1; ZetaRPMI+CuO rods sures, it is possible to define the following cytotoxicity rank:
= –9.8 ± 0.4; ZetaRPMI+CuO spheres = –10 ± 1; ZetaRPMI+CuO spindles = –9 ± 1; CuO spheres > CuO rods > CuO spindles > CuO INT, which
ZetaRPMI+CuO INT = –7 ± 0.4). However, on examining the is quite the same rank of mass percent dissolution (3.5, 1, 0.66
digital images of the suspension, the particles showed high and 0.7 %, respectively) but not similar to the size rank (7,
dispersability in both media, with the exception of CuO 7 × 40, 1200 × 270 × 30, 10–100 nm, respectively) and avail-
INT samples that showed signs of particle sedimentation able surface area (60, 51, 18, CuO INT m2/g) reported by S. K.
indicating particle agglomeration. The CuO INT preparation Misra, S. Nuseibeh, A. Dybowska, D. Berhanu, T. D. Tetley and
(pH 5.5 in water and 7.5 in medium) is far enough from its E. Valsami-Jones, submitted for publication. Due to increase in

Fig. 2. XRD pattern of all the types of CuO NPs used in this study. The patterns are shifted vertically to improve the clarity.

290
Shape-specific CuO NP cytotoxicity and genotoxicity

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 3. Effects of CuO NPs treatments on the viability of RAW 264.7 cells and whole blood by the MTT assay. Cell viability was assessed by the MTT assay
and results are presented as a percentage of negative control viability. C+: 10 µM H2O2. Cells were treated with the indicated concentrations of CuO NPs for 2 h
(a) and 24 h (b). Statistically significant differences from the control were determined by Student’s t-test (*P < 0.05, **P < 0.01, ***P < 0.001). Data represent
the mean of two separate experiments and error bars represent the standard error of the mean. In RAW 264.7 (2 h treatment) CuO spindles showed a dose–effect
relationship (r = −0.92, R2 = 84%, P = 0.03) as well as CuO spheres in PBL (2 h exposure) (r = −0.97, R2 = 94.3%, P = 0.006). In RAW 264.7 (24 h treatment),
CuO spheres showed a dose–effect relationship (r = −0.94, R2 = 89.2%, P = 0.015) as well as CuO spindles (r = −0.92, R2 = 85.3%, P = 0.025) and CuO rods
(r = −0.92, R2 = 83.8%, P = 0.03) exposures.

size, we expect the surface area of INT and spindles to be lower minimally and have the least surface area) were comparably
showing that the lower cytotoxicity can be linked to a lower cytotoxic with respect to the other CuO NPs. This complex
surface area, even if it has to be considered that the surface bioreactivity is more evident in 24 h blood responses where
area is not only dependent on the size but also on the porosity the rank of cytotoxicity is CuO spheres > CuO rods > CuO
of the NPs. spindles > CuO INT, confirming the major efficiency of CuO
However, because the IC50 values for every NP tested spheres and rods to induce cell death. This situation could be
were similar, the relationship among dissolution, potential explained by the same order of mass percent dissolution and
agglomeration and cell viability was not directly proportional size rank, taking into account the prominent phagocytic activity
to surface area. For example, CuO spindles (which dissolved of macrophages with respect to lymphocytes that could play

291
S. Di Bucchianico et al.

Table II. IC50 values of CuO NPs exposures in RAW 264.7 and PBL,
concentrations of CuO NPs suspensions. Despite the similar
graphically interpolated from linear regression equations of MTT data pattern of proliferation, the frequency of micronuclei is much
higher in phagocytic cells compared with lymphocytes except
RAW 264.7 PBL for exposure to CuO INT, where the highest tested concentra-
2 h (μg/ml) 24 h (μg/ml) 2 h (μg/ml) 24 h (μg/ml) tion induced a greater micronuclei formation in lymphocytes
and also caused a concurrent clear reduction of the CBPI
Spheres 94.8 9.9 22.6 12.1 (from 1.86 to 1.32). A linear relationship was found in the MN
Spindles 96.4 19.5 22.8 16.3 induction by CuO spheres and spindles treatments in RAW
Rods 195 10.2 29.7 15.2
INT 167 71.9 124.8 93.2 264.7 cells and by CuO spindles, rods and INT treatments in
lymphocytes.
There was a highly significant dose-dependent increase in

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


nucleoplasmic bridges frequency as a result of exposure to
an important role in NP uptake. Coherently, the IC50 values CuO spindles, rods and INT in lymphocytes cultures, whereas
obtained in lymphocytes were generally higher than those CuO spheres and rods treatments in RAW 264.7 reveal a trend
obtained in RAW 264.7. Results obtained with the MTT assay (P < 0.05 for all dose–response effects) (Figure 5). NBUDs,
were also confirmed by trypan blue dye exclusion assay (data MN-like bodies attached to the nucleus by a thin nucleoplas-
not shown). mic connection, are significantly generated only after RAW
264.7 cells exposure to NPs except CuO INT (Figure 6). Most
NBUDs in binucleate cells appear to originate from interstitial
Cytostasis and chromosomal damage or terminal acentric fragments, possibly representing nuclear
Table III summarises the results of cell proliferation and membrane entrapment of DNA that has been left in cytoplasm
cytostasis measured by scoring the CBPI and mitotic, after nuclear division, or excess DNA that is being extruded
apoptotic and necrotic indices. All doses in both cell systems from the nucleus (20). No significant NBUDs formations
reduce the proliferation index in a statistically significant were seen in lymphocytes except after CuO spheres and INT
manner. A linear relationship was found in the CBPI of exposure.
CuO INT exposed RAW 264.7 cells, whereas exposed
lymphocytes to spindles and CuO INT revealed a decrease of
cell proliferation in a dose-related manner. Exposure to CuO Clastogenic and aneuploidogenic effects of CuO NPs
NPs caused apoptotic responses measured by apoptotic figures Performing FISH analysis, we found a trend of MN induction
scored in the CBMN Cyt assay. Programmed cell death was in response to NP exposures comparable to the results obtained
increased to 2.5-fold (CuO spindles and rods) in macrophages performing the CBMN Cyt assay (Table IV). All CuO NPs
and up to 4.5-fold in lymphocytes (CuO INT), whereas severe induced an increased number of both MN C+ and MN C–
necrotic responses were observed in RAW 264.7 cells but with respect to untreated control; however, it is possible to
not in exposed whole blood. Consistent with the inhibition observe that all treatments induced a significant increase in
of proliferation, an inhibition of mitotic index was observed aneuploidogenic response (MN C+) in both cell systems, with a
in both studied cellular systems. Cytostatic responses are not consistent different behaviour of CuO NPs in RAW 264.7 cell
directly related to considered physicochemical characteristics treatments. In macrophagic cells, although the aneuploidogenic
of the NPs. events are predominant, CuO INT and spindles showed a major
Figure 4 shows the micronuclei frequency obtained by treat- degree of clastogenic events (MN C–) compared with spheres
ing RAW 264.7 cells and blood cultures with different mass and rods.

Table III. CBPI, mitotic, apoptotic and necrotic indices scored by CBMN Cyt assay

CBPI Mitotic index Apoptotic index Necrotic index


RAW 264.7 PBL RAW 264.7 PBL RAW 264.7 PBL RAW 264.7 PBL

C− 1.83 ± 0.02 1.86 ± 0.06 4.3 ± 0.2 4.9 ± 0.4 1.7 ± 0.2 0.7 ± 0.1 0.9 ± 0.1 0.6 ± 0.2
C+ 1.38 ± 0.07*** 1.40 ± 0.05*** 1.9 ± 0.3*** 2.3 ± 0.3*** 4.6 ± 0.5*** 2.2 ± 0.2*** 2.1 ± 0.2 1.9 ± 0.4
Spheres
0.1 1.64 ± 0.03** 1.59 ± 0.08*** 1.8 ± 0.2*** 3.9 ± 0.4* 3.0 ± 0.4*** 1.5 ± 0.2*** 1.9 ± 0.3** 0.8 ± 0.3
1 1.60 ± 0.02** 1.55 ± 0.07*** 1.7 ± 0.1*** 3.4 ± 0.3** 4.9 ± 0.7*** 1.6 ± 0.5** 2.9 ± 0.3*** 1.5 ± 0.1**
10 1.60 ± 0.02** 1.50 ± 0.06*** 1.6 ± 0.2*** 3.0 ± 0.2*** 2.6 ± 0.2** 1.7 ± 0.1*** 3.9 ± 0.3*** 1.3 ± 0.2*
Spindles
0.1 1.60 ± 0.03** 1.65 ± 0.04*** 1.8 ± 0.1*** 4.5 ± 0.5 3.4 ± 0.2*** 1.6 ± 0.5** 1.9 ± 0.1** 1.2 ± 0.4
1 1.53 ± 0.03*** 1.65 ± 0.03*** 1.5 ± 0.2*** 3.7 ± 0.7 3.7 ± 0.3*** 1.7 ± 0.3*** 4.7 ± 0.7*** 1.2 ± 0.2*
10 1.62 ± 0.06** 1.59 ± 0.04*** 1.6 ± 0.2*** 3.3 ± 0.3*** 4.3 ± 0.5*** 2.5 ± 0.4*** 5.3 ± 0.4*** 1.6 ± 0.3**
Rods
0.1 1.70 ± 0.03** 1.72 ± 0.03** 1.9 ± 0.1*** 4.1 ± 0.6 2.5 ± 0.4* 1.4 ± 0.2** 1.2 ± 0.3 1.1 ± 0.2*
1 1.70 ± 0.03** 1.75 ± 0.02* 1.8 ± 0.2*** 3.3 ± 0.4** 2.8 ± 0.7* 1.6 ± 0.4** 1.8 ± 0.4* 1.3 ± 0.2*
10 1.69 ± 0.08* 1.45 ± 0.08*** 1.9 ± 0.2*** 2.7 ± 0.2*** 4.3 ± 0.5*** 2.2 ± 0.2*** 2.3 ± 0.2** 2.2 ± 0.4***
INT
0.1 1.73 ± 0.05* 1.60 ± 0.03*** 2.2 ± 0.4*** 3.5 ± 0.5* 1.9 ± 0.5 1.8 ± 0.2** 0.7 ± 0.3 1.2 ± 0.2*
1 1.49 ± 0.05*** 1.67 ± 0.02*** 2.0 ± 0.2*** 2.3 ± 0.5*** 2.7 ± 0.2** 2.6 ± 0.5*** 1.7 ± 0.2* 1.4 ± 0.1**
10 1.35 ± 0.05*** 1.32 ± 0.01*** 1.1 ± 0.4*** 1.8 ± 0.2*** 4.2 ± 0.6*** 3.2 ± 0.4*** 2.5 ± 0.5*** 1.9 ± 0.2***

Data represent the mean ± SEM of two separate experiments. Statistically significant differences from the control were determined by Student’s t-test (*P < 0.05;
**P < 0.01; ***P < 0.001).

292
Shape-specific CuO NP cytotoxicity and genotoxicity

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 4. Micronuclei frequency determined by CBMN Cyt assay in RAW 264.7 and PBL cells from whole blood after CuO NPs exposure at indicated mass
concentrations (µg/ml). C+, 0.17 µg/ml mitomycin C. Statistically significant differences from the control were determined by Student’s t-test (*P < 0.05,
**P < 0.01, ***P < 0.001). Raw 264.7 CuO spheres treatment showed a dose–effect relationship (r = 0.96, R2 = 92.5%, P = 0.009) as well as the CuO spindles
exposure (r = 0.89, R2 = 79.6%, P = 0.042). PBL CuO spindles treatment showed a dose–effect relationship (r = 0.85, R2= 72.8%, P = 0.05) as well as both CuO
rods (r = 0.95, R2 = 90.4%, P = 0.045) and CuO INT (r = 0.91, R2 = 84.5%, P = 0.027) exposures.

Fig. 5. Nucleoplasmic bridges frequency determined by CBMN Cyt assay after CuO NPs exposure at indicated mass concentrations (microgram per millilitre).
C+, 0.17 µg/ml mitomycin C. Statistically significant differences from the control were determined by Student’s t-test (**P < 0.01, ***P < 0.001). RAW 264.7
CuO spheres treatment showed a dose–effect relationship (r = 0.87, R2 = 76.4%, P = 0.05) and CuO rods exposure (r = 0.99, R2 = 99.6%, P = 0.002). PBL CuO
spindles treatment showed a dose–effect relationship (r = 0.95, R2 = 90.1%, P = 0.012) as well as both CuO rods (r = 0.96, R2 = 91.3%, P = 0.044) and CuO INT
(r = 0.96, R2 = 92.2%, P = 0.04) exposures.

Primary and oxidative DNA damage and 6-fold higher than in untreated control, respectively), whereas
Figure 7 shows data obtained in evaluating the CuO NP-induced CuO INT and rods were the most effective after 24 h exposure.
DNA damage by the comet assay. Cells were exposed for 2 and The 2 and 24 h treatments were also used to investigate
24 h, with the same range of CuO NPs concentrations used for oxidative damage to DNA, by using the repair enzymes Endo-III
the CBMN Cyt assay. A significant increase in DNA migration (Figure 8) and Fpg (Figure 9). Both 2 and 24 h NPs treatments
(i.e. primary DNA damage) was detected after 2 and 24 h demonstrated the induction of Endo-III sensitive DNA sites
treatments. (oxidised pyrimidine) in both cell systems. Specifically, after
Comparing the two cell systems, it is evident that they react 2 h of treatment, RAW 264.7 and PBL showed linear dose–
differently to NP treatments. Significant trends were found in effect curves with phagocytic cells particularly susceptible to
PBL (2 and 24 h), whereas the RAW 264.7 cells showed a dose- spindles and rods treatments, whereas the lymphocytes appear
dependent effect only after the short treatment. CuO INT expo- to be generally more sensitive to 24 h CuO spheres exposures.
sures were the most effective in inducing DNA fragmentation Fpg, instead, induced significant increases in strand breaks due
after 2 h exposure both in RAW 264.7 and PBL (about 4.5-fold to oxidised purines at all concentrations tested at 2 h for RAW

293
S. Di Bucchianico et al.

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 6. Nuclear bud frequency determined by CBMN Cyt assay after CuO NPs exposure at indicated mass concentrations (microgram per millilitre). C+, 0.17 µg/
ml mitomycin C. Statistically significant differences from the control were determined by Student’s t-test (*P < 0.05, **P < 0.01, ***P < 0.001). RAW 264.7
CuO spindles treatment showed a dose–effect relationship (r = 0.91, R2 = 83%, P = 0.032). PBL CuO rods treatment showed a dose–effect relationship (r = 0.99,
R2 = 98.5%, P = 0.007).

Table IV. Percentage of centromere negative (MN C−) and centromere


attributable to any of the individual NPs characteristics. For
positive (MN C+) micronuclei in binucleated RAW 264.7 and PBL cells instance, despite the significant differences in solubility
(BNMN) incubated with 1 µg/ml of CuO NPs as evaluated with the CBMN between the four sets of CuO NPs (S. K. Misra, S. Nuseibeh,
Cyt assay in combination with fluorescence in situ hybridisation A. Dybowska, D. Berhanu, T. D. Tetley and E. Valsami-Jones,
Cell types Treatment % BNMN % MN C− % MN C+
submitted for publication), there is only a slight variation in
cytotoxic responses. Our studies were performed in 1 mM
RAW 264.7 Negative control 0.8 ± 0.1 61.4 ± 3.2 38.6 ± 3.6 NaNO3, but even in serum-free cell culture medium (DCCM-
MMC 3.8 ± 0.4 74.3 ± 2.6 25.4 ± 1.8 1) a significantly higher percentage dissolution was found for
Spheres 2.1 ± 0.1 23.7 ± 3.7 76.4 ± 2.8 all the nanoparticles, whereas the relative properties within
Spindles 1.9 ± 0.1 36.7 ± 3.4 63.3 ± 3.3
Rods 2.3 ± 0.3 30.4 ± 1.2 69.6 ± 1.6 the different shapes still persist (S. K. Misra, S. Nuseibeh,
INT 1.6 ± 0.2 46.2 ± 2.3 53.8 ± 1.8 A. Dybowska, D. Berhanu, T. D. Tetley and E. Valsami-Jones,
PBL Negative control 0.5 ± 0.1 68.6 ± 2.2 31.4 ± 1.4 submitted for publication).
MMC 4.6 ± 0.4 72.4 ± 4.2 27.6 ± 2.2 In terms of links with physicochemical properties, there
Spheres 0.8 ± 0.1 26.8 ± 2.2 73.2 ± 3.2
Spindles 0.7 ± 0.1 32.3 ± 1.8 67.7 ± 3.6 seems to be a better correlation between NPs available surface
Rods 1.2 ± 0.3 25.8 ± 3.6 74.2 ± 2.6 area and cytotoxicity (data not shown) as would be expected due
INT 1.1 ± 0.2 24.7 ± 2.3 75.3 ± 3.4 to the lower surface area of the larger particles. However, this
correlation is not significant and it is not discussed further here.
MMC, 0.17 µg/ml mitomycin C. Statistically significant differences from
the negative control were determined by Student’s t-test and all data result Moreover, we have also to take into account the relative differ-
P < 0.05 respect to relative control. ences in zeta potential that can lead to a higher agglomeration that
could retard the dissolution (21) and the diffusion of metal ions
(S. K. Misra, S. Nuseibeh, A. Dybowska, D. Berhanu, T. D. Tetley
264.7 and PBL. Moreover, 24 h of PBL treatment (CuO spheres and E. Valsami-Jones, submitted for publication). However, all
and rods) showed dose-dependent effects more than RAW264.7 CuO NPs used in this study showed strong cytotoxic responses,
cells (CuO spindles). CuO spheres treatments were the most independently from their differences in zeta potential.
effective in macrophages, whereas rods exposures were the Cytostatic responses, evaluated by scoring apoptotic,
most effective in lymphocytes. necrotic and mitotic figures as well as the proliferation index
in CBMN Cyt assay, could not be linked to a specific physico-
chemical property.
Discussion CuO NPs also induced a significant increase in MN fre-
Cytotoxic and genotoxic responses following in vitro quency in PBL and in RAW 264.7, and FISH analysis revealed
exposure to CuO INT and differently shaped CuO NPs (mass that all CuO NPs are able to induce a significant increase in
concentrations from 0.1 to 100 μg/ml) were assessed in murine malsegregation events. It is known that many nanomateri-
macrophages and in PBL. The MTT assay revealed a decrease als such as multi-walled carbon nanotubes, amorphous SiO2,
of cell viability in RAW 264.7 cells and PBL with significant TiO2 and CoCr can induce aneuploidy by mechanical interfer-
dose–effect relationships over the tested concentrations and ence with the cytoskeleton, causing chromosome missegrega-
with distinctive patterns for different CuO NPs; the comparison tion (22). Overall, this study provides the first evidence of
between cytotoxicity and physicochemical characteristics of the CuO NPs potential to induce clastogenic and aneugenic
NPs showed a differentiated bioreactivity, however, not directly events.

294
Shape-specific CuO NP cytotoxicity and genotoxicity

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 7. Primary DNA damage. Comet assay revealed an increase in primary DNA damage, with different sensitivity shown by cell types at 2 (above) and 24 h
(bottom) treatments. C+: 10 µM H2O2. CuO spheres treatment for 2 h in Raw 264.7 cells showed a dose–effect relationship (r = 0.95, R2 = 90.7%, P = 0.04) as
well as in CuO spindles (r = 0.99, R2 = 97.6%, P = 0.01), rods (r = 0.99, R2 = 97.4%, P = 0.01) and CuO INT (r = 0.99, R2 = 98.6%, P = 0.007) exposures. CuO
spindles (2 h treatment) showed in PBL a dose–effect relationship (r = 0.97, R2 = 93.6%, P = 0.007) as well as both CuO rods (r = 0.99, R2 = 99.3%, P = 0.0002)
and INT (r = 0.95, R2 = 89.9%, P = 0.05). Only CuO spindles RAW 264.7 (24 h treatment) continue to show a trend (r = 0.95, R2 = 89.9%, P = 0.05), whereas
in PBL (24 h treatment) a dose-dependent relationship was found for spheres (r = 0.98, R2 = 96%, P = 0.004), spindles (r = 0.93, R2 = 86.8%, P = 0.02) and rods
(r = 0.97, R2 = 94%, P = 0.007). Statistically significant differences from the control were determined by Student’s t-test (**P < 0.01, ***P < 0.001).

Moreover, there was a highly significant increase in nucleo- by the comet assay, but a more modest effect in terms of MN
plasmic bridges frequency indicating chromosome breakage frequency. The two assays have different sensitivity especially
and fusions in both cell systems, whereas NBUDs were sig- with regard to the potential to induce missegregation as already
nificantly generated only in RAW 264.7 cells. The induction demonstrated by Hartmann and co-authors (23). In fact, geno-
of apoptosis and chromosomal rearrangements often in a dose- toxic insults induced by aneugens are not readily detectable by
dependent manner and specifically at low doses of NPs, accom- the comet assay and CuO INT NPs, the least aneugenic NPs
panied by proliferation arrest at high concentrations, suggests among those tested by us (as shown here by FISH analysis),
differential sensitivity of NPs concentrations independently are the most effective in inducing DNA damage assessed by
from cell types. A possible mechanism could be interpreted as the comet assay.
a status where cells tolerate DNA damage and gain resistance This study shows, therefore, that two sets of CuO NPs, syn-
to cell death at low dose exposures. thesised using very different protocols (lab aqueous synthesis
The comet assay also showed an increase in primary DNA and commercial plasma synthesis) both can induce cytotoxic,
damage after 2 and 24 h treatments, whereas oxidative damage cytostatic and genotoxic effects. These findings are consist-
to DNA was found with NPs shape-related trends. ent with a wide range of data available in the literature and
Interestingly, for the RAW 264.7 cells the CuO INT NPs demonstrate that all CuO NPs, effectively independently from
caused the highest response in primary DNA damage scored their precise size and shape, clearly cause positive responses.

295
S. Di Bucchianico et al.

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 8. Oxidative damage to DNA pyrimidines after 2 h (above) and 24 h (bottom) CuO NPs treatment in cell cultures. Endo III sensitive sites in treated,
untreated (C−) and positive control (C+: 10 µM H2O2) cultures are shown. Statistically significant differences from the control were determined by Student’s t-test
(*P < 0.05, **P < 0.01, ***P < 0.001). CuO spheres treatment for 2 h in RAW 264.7 showed a dose–effect relationship (r = 0.97, R2 = 94%, P = 0.03) as well
as CuO INT exposure (r = 0.98, R2 = 96.7%, P = 0.016). In 2 h whole blood treatment, a dose–effect relationship was found in spheres (r = 0.99, R2 = 98.4%,
P = 0.008), rods (r = 0.99, R2 = 98.6%, P = 0.007) and CuO INT (r = 0.98, R2 = 95.2%, P = 0.02). After 24 h treatment, CuO spheres and spindles showed a dose–
effect relationship in RAW 264.7 (r = 0.99, R2 = 98.8%, P = 0.006 and r = 0.96, R2 = 91.4%, P = 0.04, respectively) and also in PBL 24 h exposure a dose–effect
relationship was found for CuO spheres (r = 0.99, R2 = 98%, P = 0.01), rods (r = 0.99, R2 = 99.4%, P = 0.003) and CuO INT (r = 0.98, R2 = 96.6%, P = 0.02).

Among different metal oxide NPs for which a certain amount RAW 264.7 cells from the cytotoxicity induced by submaximal
of studies are available, CuO appears to trigger the highest doses of this nanomaterial (12).
responses for inducing cytotoxicity, intracellular ROS genera- Moreover, by evaluating their ability to cause cell death,
tion and DNA damage (8). Karlsson and coworkers (8) showed mitochondrial damage, DNA damage and oxidative DNA
that the A549 cell viability determined after 18 h of exposure lesions, NPs of CuO were found highly toxic when compared
to CuO NPs at doses of 40 μg/ml and 80 μg/ml was reduced with other metal oxide NPs and much more toxic compared
by 90 and 96%, respectively. These toxicity data were consist- with CuO micrometric-sized particles (24). In human laryn-
ent with our results and the results they obtained by the comet geal epithelial cells (HEp-2), CuO induced the greatest amount
assay using Fpg, demonstrating oxidative lesions induced in of cytotoxicity in a dose-dependent manner and was able to
DNA. We found effectively a clear induction of oxidative dam- overwhelm antioxidant defences (e.g. catalase and glutathione
age to DNA detected not only by using FPg, an enzyme that reductase) compared with other oxides such as silicon oxide
recognises and excises oxidised purines (including 8-hydroxy- and ferric oxide NPs. Moreover, a significant increase in the
20-deoxyguanosine), but also by using Endo-III that recognises level of 8-isoprostanes and in the ratio of GSSG to total glu-
and excises oxidised pyrimidines. Oxidative damage to DNA tathione in cells exposed to CuO, together with a partial reduc-
seems to be the central element in the induction of the series of tion of the cytotoxic effect when HEp-2 cells were co-treated
multiple effects shown in this study. Consistent with the sup- with resveratrol was observed (25).
posed oxidative CuO NPs effects, the addition of the antioxi- Even in human A549 cells, CuO NPs were found to induce
dant N-acetyl-cysteine has been shown to significantly protect oxidative stress in a dose-dependent manner indicated by

296
Shape-specific CuO NP cytotoxicity and genotoxicity

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


Fig. 9. Oxidative damage to DNA purines after 2 h (above) and 24 h (bottom) CuO NPs treatment in cell cultures. Fpg sensitive sites in treated, untreated (C−) and
positive control (C+: 10 µM H2O2) cultures are shown. Statistically significant differences from the control were determined by the Student’s t-test (***P < 0.001).
CuO spheres treatment for 2 h in RAW 264.7 showed a dose–effect relationship (r = 0.95, R2 = 91%, P = 0.04) as well as spindles (r = 0.99, R2 = 97.4%,
P = 0.013), rods (r = 0.95, R2 = 90%, P = 0.04) and CuO INT (r = 0.95, R2 = 89.7%, P = 0.05) exposure. For 2 h whole blood treatments, a dose–effect
relationship was found in spheres (r = 0.98, R2 = 96%, P = 0.02), rods (r = 0.97, R2 = 93.5%, P = 0.03) and CuO INT (r = 0.97, R2 = 93.5%, P = 0.03). After 24 h,
only CuO spindles showed a dose–effect relationship in RAW 264.7 (r = 0.99, R2 = 97.3%, P = 0.01), whereas in PBL 24 h exposures, CuO spheres (r = 0.97,
R2 = 94%, P = 0.03) and rods (r = 0.99, R2 = 98.3%, P = 0.009) showed a dose–effect relationship.

depletion of glutathione and induction of lipid peroxidation, the leached copper–peptide complex induces a multiple-fold
catalase and superoxide dismutase (6). Moreover, that study increase in intracellular ROS generation and reduces the frac-
showed that CuO NPs induce significant effects on cell viabil- tion of viable cells, resulting in the overall inhibition of bio-
ity, even at the lowest dose tested (10 μg/ml) and that A549 cell mass growth (28).
viability is reduced to 48% at the highest concentration tested NP cytotoxicity also depends on intracellular solubility as
(50 μg/ml) after 24 h of exposure. demonstrated by a comparison of stabilised copper metal NPs
The ability of the CuO NPs to cause cell death, DNA dam- and degradable CuO NPs (29). Using copper as a representa-
age and oxidative DNA lesions was evaluated by Wang and co- tive example, Studer and collaborators (29) compared the cyto-
authors (26) and one key mechanism proposed was the ability toxicity of copper metal NPs stabilised by a carbon layer with
of CuO NPs to damage mitochondria. CuO NPs showed signifi- copper oxide NPs using two different cell lines. Measuring the
cant toxicity mediated by oxidative stress in both HepG2 cells intra- and extracellular solubility of the two materials in stand-
and catfish primary hepatocytes (26). The toxicity observed not ard buffers, the authors explained the cytotoxicity difference
only highlights ROS-induced cell death but also was linked to on the basis of altered copper release. In fact, the well-known
cell and mitochondrial membranes. Furthermore, high content mechanism of intracellular dissolution (Trojan horse-type
bright-field imaging demonstrated potent and dose-dependent mechanism) illustrates how NPs can bypass the protection of
hatching interference in the zebrafish embryos compromising mammalian cells against heavy metal ions and the dominating
the activity of the hatching enzyme, ZHE1 (27). This hypoth- influence of physicochemical parameters on the cytotoxicity of
esis is based on the presence of metal-sensitive histidine in the a given metal (30). However, the electronic properties of the
catalytic centre of this enzyme. Furthermore, complexation- oxide NPs, determined by their chemical composition, can be
mediated leaching of CuO NPs by amino acids was identified modulated by crystalline structure, defect, size, shape and mor-
as the source of toxicity towards Escherichia coli showing that phology, and may play an important role in their cytotoxicity.

297
S. Di Bucchianico et al.

For instance, although no significant differences were observed Positive controls are needed to demonstrate the ability of
in the cytoxicity of spherical and sheet-shaped ZnO NPs on the cells used and the test protocols to detect compounds able
BEAS-2B and RAW 264.7 cells, there were differences in their to induce adverse effects (39). Moreover, the appropriateness
cellular association and inflammatory potential (31). Even so, of available testing systems for nanomaterials requires con-
ZnO NPs with different sizes and shapes can induce different firmation and standardisation (40). The availability of refer-
cytotoxic effects in A549 human lung epithelial cells as well as ence materials is important to function as a benchmark for
different influences on the mitochondria activity and chemokine the adverse effects of materials with complex structures and
production (32). Likely, the differential effects of NPs size and to compare toxicological data from various studies appropri-
shape on their cytotoxicity can also be closely linked to the cell ately (41).
type. NPs with different specific surface area, different sizes and
different raw material but the same hydrodynamic diameter, dif-
Funding

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


ferentially disturbed the Caco-2 cell monolayer integrity, were
cytotoxic and triggered an increase of several transcripts cod- European Community’s Seventh Framework Programme
ing for pro-inflammatory cytokines and chemokines; this dif- (FP7/2007–2013) under grant agreement No CP-FP 214478-2
ferential toxicity was only partly correlated with the release (project name NanoRetox; www.nanoretox.eu).
of copper and with the shape of the NPs (33). For a number
of biomarkers, the biological responses are significantly more Conflict of interest statement: Paul Reip is an employee of Intrinsiq Materials
Ltd, which sourced (and also manufactures) the CuO used in this study.
important in presence of CuO NPs than in presence of soluble
Cu suggesting a specific nanoparticle effect (34). CuO NPs were
clearly observed in both the cell nucleus and the mitochondria
References
after being taken up by human pulmonary epithelial A549 cells
mainly through endocytosis, although part of the internalised 1. Huang, X., Teng, X., Chen, D., Tang, F. and He, J. (2010) The effect of the
CuO NPs were dynamically excreted to the extracellular envi- shape of mesoporous silica nanoparticles on cellular uptake and cell func-
tion. Biomaterials, 31, 438–448.
ronment (35). Intracellular localisation was also observed from 2. Xu, L. J., Zhao, J. X., Zhang, T., Ren, G. G. and Yang, Z. (2009) In vitro
2 to 24 h of incubation with both rod-shaped and spherical CuO study on influence of nano particles of CuO on CA1 pyramidal neurons of
NPs in HepG2 cells (36). In addition, numerous CuO NPs were rat hippocampus potassium. Curr. Environ. Toxicol., 24, 211–217.
observed in dead A549 cells compared with viable cells, sug- 3. Sun, J., Wang, S., Zhao, D., Hun, F. H., Weng, L. and Liu, H. (2011)
gesting that the difference in the amount of CuO NPs that are Cytotoxicity, permeability, and inflammation of metal oxide nanoparticles
in human cardiac microvascular endothelial cells. Cell Biol. Toxicol., 25,
absorbed into individual cells may determine their fate (37). 333–342.
This interesting study by Hanagata and co-authors (37) also 4. Ahamed, M., Siddiqui, M. A., Akhtar, M. J., Ahmad, I., Pant, A. B. and
showed that up-regulation of DNA damage-inducible 45 β and Alhadlaq, H. A. (2010) Genotoxic potential of copper oxide nanoparticles
γ genes expression was not observed with Cu ions released into in human lung epithelial cells. Biochem. Biophys. Res. Commun., 396,
578–583.
medium but was observed in cells exposed to CuO NPs, high- 5. Pan, X., Redding, J. E., Wiley, P. A., Wen, L., McConnell, J. S. and Zhang,
lighting that CuO NPs themselves cause genotoxicity. B. (2010) Mutagenicity evaluation of metal oxide nanoparticles by the bac-
Other than size or shape, the response of NPs in the cells may terial reverse mutation assay. Chemosphere, 79, 113–116.
be directed by complications regarding dosage. For chemicals, 6. Lei, R., Wu, C., Yang, B., Ma, H., Shi, C., Wang, Q., Wang, Q., Yuan, Y. and
dosage simply implies molar concentration, whereas for NPs Liao, M. (2008) Integrated metabolomic analysis of the nano-sized cop-
per particle-induced hepatotoxicity and nephrotoxicity in rats: a rapid in
the concentration can be viewed as a 2D entity, one compo- vivo screening method for nanotoxicity. Toxicol. Appl. Pharmacol., 232,
nent being the atomic or molar density, the other being the size, 292–301.
shape or other surface properties (38). Thus, a complex inter- 7. Cho, W. S., Duffin, R., Poland, C. A., Howie, S. E., MacNee, W., Bradley,
play prevails between size, shape, dissolution, zeta potential, M., Megson, I. L. and Donaldson, K. (2010) Metal oxide nanopar-
ticles induce unique inflammatory footprints in the lung: important
surface area and nature of surface functionalisation, as well as implications for nanoparticle testing. Environ. Health Perspect., 118,
the nature of protein corona that defines the biological identity 1699–1706.
of NPs. Nevertheless, numerous controversies persist about the 8. Karlsson, H. L., Cronholm, P., Gustafsson, J. and Möller, L. (2008)
biological impact of NPs physicochemical properties and fur- Copper oxide nanoparticles are highly toxic: a comparison between
ther studies are needed due to relatively few reports describing metal oxide nanoparticles and carbon nanotubes. Chem. Res. Toxicol., 21,
1726–1732.
the extents to which the NP properties influence their toxicity. 9. Fubini, B., Ghiazza, M. and Fenoglio, I. (2010) Physico-chemical features
Overall, the data obtained here together with the available lit- of engineered nanoparticles relevant to their toxicity. Nanotoxicology, 4,
erature suggest that CuO NPs should be considered as a valuable 347–363.
positive control for cytotoxicity and genotoxicity studies of nano- 10. Hanley, C., Thurber, A., Hanna, C., Punnoose, A., Zhang, J. and Wingett,
D. G. (2009) The Influences of Cell Type and ZnO Nanoparticle Size on
structured materials, at least in in vitro studies with murine and Immune Cell Cytotoxicity and Cytokine Induction. Nanoscale Res. Lett., 4,
mammalian cells, due to the fact that they exert a clear cytotoxic 1409–1420.
and genotoxic action, independent of their size and shape. For 11. Di Giorgio, M. L., Di Bucchianico, S., Ragnelli, A. M., Aimola, P., Santucci,
instance, Rotoli and co-authors (12) observed that CuO NPs (the S. and Poma, A. (2011) Effects of single and multi walled carbon nanotubes
same CuO INT used for our work) had very marked effects on on macrophages: cyto and genotoxicity and electron microscopy. Mutat.
Res., 722, 20–31.
cell viability, comparable to that observed upon treatment with 12. Rotoli, B. M., Bussolati, O., Costa, A. L. et al. (2012) Comparative effects
Ni NPs, used as a positive control in comparative studies of metal of metal oxide nanoparticles on human airway epithelial cells and mac-
oxide NPs on human airway epithelial cells and macrophages. rophages. J. Nanopart. Res., 14, 1069.
However, although the primary role of oxidative stress mecha- 13. Migliore, L., Frenzilli, G., Nesti, C., Fortaner, S. and Sabbioni, E. (2002)
Cytogenetic and oxidative damage induced in human lymphocytes by plati-
nisms and the contribution of aneugenic responses in cytotoxic- num, rhodium and palladium compounds. Mutagenesis, 17, 411–417.
ity and genotoxicity induced by CuO NPs is clear, it would need 14. Collins, A. R., Oscoz, A. A., Brunborg, G., Gaivão, I., Giovannelli, L.,
to be combined with a better understanding of cellular uptake Kruszewski, M., Smith, C. C. and Stetina, R. (2008) The comet assay: topi-
levels, in particular in the function of different cell types. cal issues. Mutagenesis, 23, 143–151.

298
Shape-specific CuO NP cytotoxicity and genotoxicity

15. Fenech, M. (2007) Cytokinesis-block micronucleus cytome assay. Nat. cytotoxicity depends on intracellular solubility: comparison of stabilized
Protoc., 2, 1084–1104. copper metal and degradable copper oxide nanoparticles. Toxicol. Lett.,
16. Parry, E. M., Parry, J. M., Corso, C. et al. (2002) Detection and characteri- 197, 169–174.
zation of mechanisms of action of aneugenic chemicals. Mutagenesis, 17, 30. Xia, T., Kovochich, M., Liong, M., Mädler, L., Gilbert, B., Shi, H., Yeh, J.
509–521. I., Zink, J. I. and Nel, A. E. (2008) Comparison of the mechanism of toxic-
17. Zhu, J., Li, D., Chen, H., Yang, X., Lu, L. and Wang, X. (2004) Highly dis- ity of zinc oxide and cerium oxide nanoparticles based on dissolution and
persed CuO nanoparticles prepared by a novel quick precipitation method. oxidative stress properties. ACS Nano, 2, 2121–2134.
Mater. Lett. 58, 3324–3327. 31. Heng, B. C., Zhao, X., Tan, E. C., Khamis, N., Assodani, A., Xiong, S.,
18. Migliore, L., Saracino, D., Bonelli, A., Colognato, R., D’Errico, M. R., Ruedl, C., Ng, K. W. and Loo, J. S. (2011) Evaluation of the cytotoxic and
Magrini, A., Bergamaschi, A. and Bergamaschi, E. (2010) Carbon nano- inflammatory potential of differentially shaped zinc oxide nanoparticles.
tubes induce oxidative DNA damage in RAW 264.7 cells. Environ. Mol. Arch. Toxicol., 85, 1517–1528.
Mutagen., 51, 294–303. 32. Hsiao, I. L. and Huang, Y. J. (2011) Effects of various physicochemical
19. Collins, A. R., Dusinská, M., Gedik, C. M. and Stĕtina, R. (1996) Oxidative characteristics on the toxicities of ZnO and TiO nanoparticles toward

Downloaded from https://academic.oup.com/mutage/article/28/3/287/1295389 by Universidade de Bras�lia user on 03 December 2021


damage to DNA: do we have a reliable biomarker? Environ. Health human lung epithelial cells. Sci. Total Environ., 409, 1219–1228.
Perspect., 104(Suppl. 3), 465–469.20. 33. Piret, J. P., Vankoningsloo, S., Mejia, J. et al. (2012) Differential toxicity
20. Lindberg, H. K., Wang, X., Järventaus, H., Falck, G. C., Norppa, H. and of copper (II) oxide nanoparticles of similar hydrodynamic diameter on
Fenech, M. (2007) Origin of nuclear buds and micronuclei in normal and human differentiated intestinal Caco-2 cell monolayers is correlated in part
folate-deprived human lymphocytes. Mutat. Res., 617, 33–45. to copper release and shape. Nanotoxicology, 6, 789–803.
21. Peng, X., Palma, S., Fisher, N. S. and Wong, S. S. (2011) Effect of mor- 34. Buffet, P. E., Tankoua, O. F., Pan, J. F. et al. (2011) Behavioural and bio-
phology of ZnO nanostructures on their toxicity to marine algae. Aquat. chemical responses of two marine invertebrates Scrobicularia plana and
Toxicol., 102, 186–196. Hediste diversicolor to copper oxide nanoparticles. Chemosphere, 84,
22. Gonzalez, L., Decordier, I. and Kirsch-Volders, M. (2010) Induction of 166–174.
chromosome malsegregation by nanomaterials. Biochem. Soc. Trans., 38, 35. Wang, Z., Li, N., Zhao, J., White, J. C., Qu, P. and Xing, B. (2012) CuO
1691–1697. nanoparticle interaction with human epithelial cells: cellular uptake, loca-
23. Hartmann, A., Schumacher, M., Plappert-Helbig, U., Lowe, P., Suter, W. tion, export, and genotoxicity. Chem. Res. Toxicol., 25, 1512–1521.
and Mueller, L. (2004) Use of the alkaline in vivo Comet assay for mecha- 36. Piret, J. P., Jacques, D., Audinot, J. N. et al. (2012) Copper(ii) oxide
nistic genotoxicity investigations. Mutagenesis, 19, 51–59. nanoparticles penetrate into HepG2 cells, exert cytotoxicity via oxi-
24. Karlsson, H. L., Gustafsson, J., Cronholm, P. and Möller, L. (2009) Size- dative stress and induce pro-inflammatory response. Nanoscale, 4,
dependent toxicity of metal oxide particles–a comparison between nano- 7168–7184.
and micrometer size. Toxicol. Lett., 188, 112–118. 37. Hanagata, N., Zhuang, F., Connolly, S., Li, J., Ogawa, N. and Xu, M. (2011)
25. Fahmy, B. and Cormier, S. A. (2009) Copper oxide nanoparticles induce Molecular responses of human lung epithelial cells to the toxicity of copper
oxidative stress and cytotoxicity in airway epithelial cells. Toxicol. In Vitro, oxide nanoparticles inferred from whole genome expression analysis. ACS
23, 1365–1371. Nano, 5, 9326–9338.
26. Wang, Y., Aker, W. G., Hwang, H. M., Yedjou, C. G., Yu, H. and Tchounwou, 38. Patra, H. K. and Dasgupta, A. K. (2011) Cancer cell response to nanopar-
P. B. (2011) A study of the mechanism of in vitro cytotoxicity of metal ticles: criticality and optimality. Nanomedicine 2011 November 15 [Epub
oxide nanoparticles using catfish primary hepatocytes and human HepG2 ahead of print].
cells. Sci. Total Environ., 409, 4753–4762. 39. Donaldson, K., Poland, C. A. and Schins, R. P. (2010) Possible geno-
27. Lin, S., Zhao, Y., Xia, T. et al. (2011) High content screening in zebrafish toxic mechanisms of nanoparticles: criteria for improved test strategies.
speeds up hazard ranking of transition metal oxide nanoparticles. ACS Nanotoxicology, 4, 414–420.
Nano, 5, 7284–7295. 40. Warheit, D. B. and Donner, E. M. (2010) Rationale of genotoxicity testing
28. Gunawan, C., Teoh, W. Y., Marquis, C. P. and Amal, R. (2011) Cytotoxic of nanomaterials: regulatory requirements and appropriateness of available
origin of copper(II) oxide nanoparticles: comparative studies with micron- OECD test guidelines. Nanotoxicology, 4, 409–413.
sized particles, leachate, and metal salts. ACS Nano, 5, 7214–7225. 41. Lai, D. Y. (2012) Toward toxicity testing of nanomaterials in the 21st cen-
29. Studer, A. M., Limbach, L. K., Van Duc, L., Krumeich, F., Athanassiou, tury: a paradigm for moving forward. Wiley Interdiscip. Rev. Nanomed.
E. K., Gerber, L. C., Moch, H. and Stark, W. J. (2010) Nanoparticle Nanobiotechnol., 4, 1–15.

299

You might also like