You are on page 1of 16

Journal of Toxicology and Environmental Health, Part A

Current Issues

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/uteh20

Toxicity of binary mixtures of Al2O3 and ZnO


nanoparticles toward fibroblast and bronchial
epithelium cells

Jéssica Schveitzer Köerich , Diego José Nogueira , Vitor Pereira Vaz , Carmen
Simioni , Marlon Luiz Neves Da Silva , Luciane Cristina Ouriques , Denice
Schulz Vicentini & William Gerson Matias

To cite this article: Jéssica Schveitzer Köerich , Diego José Nogueira , Vitor Pereira Vaz , Carmen
Simioni , Marlon Luiz Neves Da Silva , Luciane Cristina Ouriques , Denice Schulz Vicentini &
William Gerson Matias (2020): Toxicity of binary mixtures of Al2O3 and ZnO nanoparticles toward
fibroblast and bronchial epithelium cells, Journal of Toxicology and Environmental Health, Part A,
DOI: 10.1080/15287394.2020.1761496

To link to this article: https://doi.org/10.1080/15287394.2020.1761496

View supplementary material

Published online: 15 May 2020.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=uteh20
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A
https://doi.org/10.1080/15287394.2020.1761496

Toxicity of binary mixtures of Al2O3 and ZnO nanoparticles toward fibroblast


and bronchial epithelium cells
Jéssica Schveitzer Köerich a, Diego José Nogueira a, Vitor Pereira Vaz a, Carmen Simioni a
,
Marlon Luiz Neves Da Silva a, Luciane Cristina Ouriques b, Denice Schulz Vicentini a,
and William Gerson Matias a
a
Department of Sanitary and Environmental Engineering, Federal University of Santa Catarina, Florianópolis, Brazil; bDepartment Of Cell
Biology, Embryology and Genetics, Federal University of Santa Catarina, Florianópolis, Brazil

ABSTRACT KEYWORDS
The objective of this study was to examine the cytotoxic effects of binary mixtures of Al2O3 and Binary toxicity;
ZnO NPs using mouse fibroblast cells (L929) and human bronchial epithelial cells (BEAS-2B) as nanoparticles; aluminum
biological test systems. The synergistic, additive, or antagonistic behavior of the binary mixture oxide; zinc oxide; in vitro
cytotoxicity
was also investigated. In toxicity experiments, cellular morphology, mitochondrial function (MTT
assay), apoptosis, nuclear size and shape, clonogenic assays, and damage based upon oxidative
stress parameters were assessed under control and NPs exposure conditions. Although Abbott
modeling results provided no clear evidence of the binary mixture of Al2O3 and ZnO NPs
exhibiting synergistic toxicity, some specific assays such as apoptosis, nuclear size and shape,
clonogenic assay, activities of antioxidant enzymatic enzymes catalase, superoxide dismutase, and
levels of glutathione resulted in enhanced toxicity for the mixtures with 1 and 1.75 toxic units (TU)
toward both cell types. Data demonstrated that co-presence of Al2O3 and ZnO NPs in the same
environment might lead to more realistic environmental conditions. Our findings indicate cyto-
toxicity of binary mixtures of Al2O3 and ZnO NPs produced greater effects compared to toxicity of
either individual compound.

Introduction exposed to NPs during production processes or


through ingestion of contaminated water and food
The global market for nanoparticles (NPs) has
or exposure to dermal products (Wiesner et al.
experienced growth over the past two decades in
2006). Further, Donaldson et al. (2001) reported
diverse fields including aerospace, automotive,
that ultrathin metal-containing particles present in
energy, cosmetics, electronics and optics industries,
the air exerted harmful effects on health. While
biosensors, and diagnostics (Alim et al. 2018; Angell
incorporating NPs into products improves our qual-
et al. 2018; Kaweeteerawat et al. 2017). In addition,
ity of daily life, concerns regarding the consequences
these substances are used in electrochemistry to
for human health and environment are growing
obtain catalysts, coatings, paints, pigments, compo-
(Alaraby et al. 2016; Niska et al. 2018).
sites, filtration and purification systems, textiles, and
Among the most important NPs in production
many other products (Alim et al. 2018; Kengar et al.
are those comprised of zinc oxide (ZnO) and alu-
2019; Wang and Nowack 2018). Due to unique char-
minum oxide (Al2O3) (Balasubramanyam et al.
acteristics associated with reduced particle size NPs
2009; Kermanizadeh et al. 2016). Al2O3 NP pos-
exhibit a wide range of potential applications in
sesses important properties including good ther-
nanomedicine, oral care, and restorative dentistry
mal conductivity, reduced size and shape
and as antimicrobials NPs (Callister et al. 2020;
capability, and high strength and stiffness (Kar
Duran et al. 2016; Kaweeteerawat et al. 2017;
et al. 2008). Thus, Al2O3 NPs are used in the
Kermanizadeh et al. 2016; Su et al. 2018; Zhao and
production of abrasives, refractories, ceramics,
Castranova 2011). Consequently, humans may be
electrical insulators, catalysts, paper, spark plugs,

CONTACT William Gerson Matias william.g.matias@ufsc.br Laboratory of Environmental Toxicology, Department of Sanitary and Environmental
Engineering, Federal University of Santa Catarina, Florianópolis, Brazil
Supplemental data for this article can be accessed on publisher’s website.
© 2020 Taylor & Francis
2 J. S. KÖERICH ET AL.

light bulbs, artificial gems, alloys, glass, and heat et al. 2015). In addition, ZnO NPs appear to be
resistant fibers (Krewski et al. 2007). ZnO NPs more toxic than Al2O3 NPs in different cell lines
have received wide attention due to their effective (Jain et al. 2015; Jeng and Swanson 2006; Zhang
performance as antimicrobial agents, pesticides, et al. 2011; SoSong et al. 2017; Ivask et al. 2015).
fungicides, herbicides, fertilizer, soil enhancers, In this study, the mouse fibroblast (L929) cell line
nanobiosensors, and in water purification systems was selected to simulate dermal exposure as the
(Cure et al. 2018; Rajput et al. 2018). model to investigate potential effects on skin, since
Studies on the effects of Al2O3 and ZnO NPs previous studies showed n that NPs enter the skin
individually, through in vitro assessments, were barrier (Crosera et al. 2016, 2009). The human
widely reported (Chen et al. 2019; Demir, Creus, bronchial epithelial (BEAS-2B) cell line was also
and Marcos 2014; Hashimoto, Sasaki, and Imazato used in the toxicity tests since the respiratory tract
2016; Lin et al. 2008; Mancuso et al. 2018; Ng et al. is another potential route of exposure for airborne
2017; Nogueira et al. 2019). However, since the rapid NPs (Ferreira, Cemlyn-Jones, and Robalo Cordeiro
development of nanotechnology and widespread 2013). Therefore, the aim of this study was to deter-
application of various types of NPs is well established mine if the actions of binary mixtures of these NPs
it is conceivable that a complex scenario might be were additive, synergistic, or antagonistic.
due to co-exposure to these materials in the environ-
ment (Yu et al. 2016). It has been documented that
the combined effects of different types of NPs may Materials and methods
lead to additive, synergistic, or antagonistic impacts Chemicals
on aquatic organisms as compared to their indivi-
dual effects (Iswarya et al. 2016; Ye et al. 2017; Yu The powder sample of ZnO NPs was synthesized by
et al. 2016). Dávila-Grana et al. (2018) found that the polymeric precursor method according to
incubation of a lymphocytic cell line to a mixture of Vicentini, Smania, and Laranjeira (2010). The sample
metal oxide NPs such as CeO2 with ZnO resulted in of Al2O3 NPs (nanopowder, particle size 13 nm, 99.8%
a synergistic effect at low and medium concentra- purity), RPMI-1640 Medium, fetal bovine serum
tions as evidenced by greater cell mortality and acti- (FBS), penicillin, L-glutamine, phosphate buffer saline
vation of MAP kinases and increased NFκB levels. Li (PBS), 3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2 H
et al. (2015) noted that the toxicity of Cu NPs toward -tetrazolium bromide (MTT), trypan blue, parafor-
human hepatoma (HepG2) cells was higher in maldehyde, crystal violet, 4-(1,1,3,3-tetramethylbu-
a mixture containing a nontoxic ZnO concentration. tyl)phenyl-polyethylene glycol, t-octylphenoxyp
Previously Benavides et al. (2016) examined olyethoxyethanol (Triton X-100), methanol, 2′,7′-
exposure of Carassius auratus to binary metal dichlorofluorescein diacetate (DCFH-DA), acridine
mixtures and noted that Al2O3 combined with orange, and propidium iodide (AO/PI), 2-(4-amidi-
ZnO may be more toxic than compared to each nophenyl)-6-indolecarbamidine dihydrochloride, 4′,
NP individually. An increase in catalase and 6-diamidino-2-phenylindole dihydrochloride (DAPI)
superoxide dismutase activities was found in was purchased from Aldrich–Sigma (St. Louis, USA).
gills and livers of fish, especially after 14 days Trypsin–EDTA was purchased from Vitrocell
exposure, indicating synergistic response. The (Campinas, Brazil), reverse osmosis (RO) water
interaction between Al2O3 and ZnO NPs may (0.05 S/cm), and ultrapure (UP) water (18.2 mΩ)
be therefore an issue of concern and the adverse were used in the experiments.
effects of this NPs mixture need to be verified
using in vitro tests. Previous investigators
Characterization of zinc oxide and aluminum
demonstrated that individually Al2O3 and ZnO
oxide
NPs produced a reduction in cell viability,
altered cell morphology, increased number of The samples (1 g/L) were dispersed in ultrapure
apoptotic cells, and enhanced generation of reac- (UP) water and RPMI supplemented with 10% FBS
tive oxygen species (ROS) (Lin et al. 2008; by sonication (10% – 50 W power in a Q500 W
Nogueira et al. 2019; Srikanth et al. 2015; Wang Qsonica sonicator). Through transmission electron
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 3

microscopy (TEM) (JEOL JEM-1011 microscope, 530 nm, intracellular ROS content was determined
100 kV) size and morphology of particles were deter- as previously described (Keston and Brandt 1965),
mined. The images from the TEM were analyzed with the aid of a microplate reader (SpectraMax
using software of particle analyzer, ImageJ (version Paradigm, Molecular Devices). The activity of
1.46 r – http://rsb.info.nih.gov/ij/). Through CAT and SOD as well as GSH levels in cells was
NanoBrook 90Plus Zeta analyzer (Brookhaven determined. The antioxidant enzyme CAT was
Instruments Corp.) were determined the effective measured based upon the decrease of the H2O2
diameter (ED) by dynamic light scattering (DLS) absorbance at 240 nm (Aebi 1984). The SOD
and zeta potential (ZP) by phase analysis light scat- enzyme employed XOD/cytochrome c method at
tering (PALS). Using an elemental analyzer absorbance at 550 nm (McCord and Fridovich
(Quantachrome Instrument, model Nova-BET 1969). Levels of GSH were measured following
1200 E), it was possible to determine the surface reaction with DTNB at absorbance at 412 nm
area by the Brunauer–Emmett–Teller method (Beutler, Duron, and Kelly 1963). For thiobarbitu-
(BET) (Brunauer, Emmett, and Teller 1938). ric acid-reactive species (TBARS) the absorbance
at 532 nm was determined by estimating malon-
dialdehyde (MDA) according to Draper and
In vitro toxicological evaluation, cell lines,
Hadley (1990). All enzymatic/non-enzymatic
culture, and maintenance
activity data were calculated as specific activity in
The L929 and BEAS-2B cells were cultured in U/mg protein. It was used The Bradford (1976)
RPMI-1640 medium supplemented with 2% (v/v) method was employed to determine total protein
L-glutamine, 1% (v/v) penicillin, 1% (v/v) pyru- mass (mg) in order to normalize the CAT, SOD,
vate, and 10% (v/v) FBS in an incubator at 37°C GSH, and MDA results.
with 5% CO2. Cells were seeded onto 96-well
plates at a density of 5 × 104 cells/well for the AO/PI staining assay
viability assay, in 24-well plates at a density of Prior to the AO/PI staining assay, cells were trea-
1 × 105 cells/well for ROS measurements and for ted for 24 h with NPs or control. Briefly, AO/PI
nuclear morphometry (NM) and acridine orange staining was used to assess the number of viable,
and propidium iodide (AO/PI) staining. In the apoptotic, and necrotic cells as previously
clonogenic assay, 6-well plates were seeded at described (Bank 1988). Cells were then rinsed
a density of 5 × 105 cells/well and TEM culture twice with PBS and centrifuged for 4 min at
flasks (25 cm2) seeded at a density of 2.5 × 106 2000 g. Finally, cells were stained for 5 min with
cells/ml were used to evaluate cell membrane the staining mixture (100 μg/ml AO and 100 μg/ml
integrity (lipid peroxidation – LPO). PI) and then observed under a fluorescence micro-
scope (Olympus BX41) with an excitation filter of
MTT assay 480/430 nm. The viable, apoptotic, and necrotic
In assay to determine cell viability, as proposed by cells were quantified in a population of 200 cells.
Mosmann (1983), the cells were treated with dif-
ferent concentrations of Al2O3 NPs (15.62; 31.50; Nuclear morphometric assay
62.50; 125; 250 or 500 mg/L) and ZnO NPs (2.50; The cells treated with NPS for 24 h were fixed
5; 10; 25; 50 or 100 mg/L). Cell viability was using a cold paraformaldehyde solution (4%), fol-
determined by measuring the optical absorbance lowed by washing with PBS, then permeabilized
at 570 nm on BioTek ELx800 equipment through with 0.1% Triton X-100 for 10 min and incubated
a microplate reader after 24 h for treatment and with 300 nM DAPI for 30 min. The nuclear mor-
3 h for incubation with MTT. phometry was determined as previously described
(Filippi-Chiela et al. 2012), based upon photo-
Oxidative stress biomarkers assay graphs taken under the fluorescence microscope
Prior to oxidative stress biomarkers assays, cells (Olympus BX41), using the ImageJ software (ver-
were treated for 24 h. Using an excitation wave- sion 1.46 r – http://rsb.info.nih.govij/) coupled
length of 485 nm and emission wavelengths of with NII_Plugin.
4 J. S. KÖERICH ET AL.

Cell uptake of NPs concentration–response relationships of the mix-


At the end of incubation with NPs for 24 h, cells tures (Appendices 2 and 3). Using the Abbott
were removed, washed with PBS, and prefixed in model (Abbott 1925) as modified by Colby
2.5% glutaraldehyde solution in 0.1 M sodium (1967), the expected toxicity values for the binary
cacodylate buffer and post-fixed in 1% osmium nanoparticle mixtures were calculated. The mix-
tetroxide (pH 7.4), dehydrated in a graded series ture may be characterized according to the ratio
of acetone (30%, 50%, 70%, 90%, and 100%) and between observed and expected toxicity as follows:
submerged in Spurr’s resin using a graded series of less than 1 = antagonist, equal to 1 = additive, and
acetone-Spurr’s resin. Thin sections (70 nm) were greater than 1 = synergistic.
obtained by ultramicrotome (LEICA UC7) before
observation by TEM (JEOL JEM-1011, 100 kV).
Statistical analysis
Clonogenic assay Statistical analysis was performed using software
The clonogenic assay used in this study was per- GraphPad Prism® v6.0. ANOVA was carried out
formed according to Franken et al. (2006). Briefly, and assumptions of normality (Shapiro-Wilk’s
after exposure to NPs for 24 h cells were trypsi- test) and homogeneity of variances (Bartlett’s test)
nized, a specific number of cells counted (2000 were tested. Grubbs Test was applied to detect out-
cells/well) and then seeded in 6-well plates. After liers. Data were analyzed by one or two-way
10 days, the plates were fixed with methanol and ANOVA followed by the Tukey’s post hoc test for
stained with 0.1% crystal violet and rinsed with parametric data or the Kruskal–Wallis test, fol-
water, air-dried, photographed (Olympus SZX16) lowed by Dunn’s post hoc test for non-parametric
and the colonies containing more than 50 cells data. Results are presented as mean ± standard
counted. deviation (SD) and differences were considered sta-
tistically significant at p < .05.
Toxicity tests with mixtures
Binary mixtures of Al2O3 NPs with ZnO NPs were
obtained following the toxic units (TU – control, Results
0.25, 0.5, 0.75, 1, 1.5, and 1.75) design proposed by
Characterization of NPs
Sprague and Ramsay (1965). Their equitoxic pro-
portions were calculated (Cao et al. 2007) in order The TEM images illustrate the morphology of the
to acquire as much information as possible on the Al2O3 NPs (Figure 1a), ZnO (Figure 1B), and the

Figure 1. TEM images for (a) Al2O3 NPs, (b) ZnO NPs, and (c) binary mixture. Size frequency distribution histograms showing
Gaussian curve (n=100 particles) for (d) Al2O3 NPs and (e) ZnO NPs.
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 5

mixture (Figure 1c). Data demonstrate that NPs stable in UP water but displayed only numerical
used were spherical and slightly agglomerated. decreases in RPMI medium over time.
Further, it can be clearly seen from Figure 1c
that by mixing the two types of NPs there was
a greater amount of Al2O3 NPs than ZnO NPs. Viability assay for NPs individually and in binary
This difference occurred because the relative sur- mixtures
face areas estimated by BET equation were Using the MTT assay to measure cell viabilityAl2O3
682.7 m2/g for Al2O3 NPs and 7.76 m2/g for and ZnO NPs individually produced a concentration-
ZnO NPs (Table 1). Based upon the histograms dependent cytotoxic effect with concentration–
exhibiting Gaussian curves for frequency distribu- response curves presented in Appendix 1. After 24 h
tion, the particle diameters were 11.19 ± 2.50 nm exposure for the two types of cells ZnO NPs were
for Al2O3 NPs and 39.34 ± 14.97 nm for ZnO NPs found to be more toxic with lower IC50 values than
(Figure 1d,e). To evaluate the dispersion stability Al2O3 NPs (Appendix 1). The L929 cells were more
of the NP suspensions, the effective diameter (ED) sensitive than BEAS-2B cells with IC50 values of
and surface charges of Al2O3 NPs and ZnO NPs approximately 280.4 and 42.9 µg/ml for Al2O3 NPs
individually and in binary mixtures were deter- and ZnO NPs, respectively. The IC50 values for BEAS-
mined in the UP water and RPMI medium con- 2B cells were approximately 378.1 and 75.2 µg/ml for
taining 10% FBS (Table 1). The ED results for the Al2O3 NPs and ZnO NPs, respectively.
Al2O3 NPs and ZnO NPs individually and in the Considering the potential simultaneous occur-
binary mixtures dispersed in ultrapure water (UP) rence of Al2O3+ ZnO NPs and based upon the
or RPMI medium confirmed the presence of par- IC50 values obtained for the individual NPs, toxi-
ticle agglomerates greater than 100 nm in dia- city tests were conducted for binary mixtures with
meter. In all cases, the ED was smaller when NPs concentrations of NPs corresponding to the toxic
were dissolved in UP water compared to RPMI units (TU) (Tables 1 and 2). Consequently, the
medium. At the same time, the zeta potential equipotential concentration of each type of NP
(ZP) values for NPs, determined individually and was obtained to prepare the binary mixtures, as
in the binary mixture (Al2O3/ZnO NPs) remained presented in Figure 2 and Table 2. The Ratio of

Table 1. Data for Al2O3 and ZnO NPs characterized in UP water and RPMI medium (1 g/L).
NPs Size±SD (nm) Surface area (m2/g) Diluent ED ± SD (nm) PZ ± SD (mV) pH
Al2O3 11.19 ± 2.50 682.7 UP water 187.05 ± 0.80 −51.13 ± 0.57 7.65
RPMI 226.32 ± 0.50 5.88 ± 1.34 8.10
ZnO −51.13 ± 0.57 7.76a UP water 276.35 ± 0.55 −34.18 ± 0.10 7.52
RPMI 887.56 ± 3.30 6.77 ± 2.51 8.09
Al2O3+ ZnO - - UP water 240.33 ± 3.05 −32.11 ± 0.36 7.43
- - RPMI 173.82 ± 0.55 7.87 ± 5.15 9.74
a
The surface area of ZnO NPs used in this study was reported previously described in the literature (Melegari et al. 2019).

Table 2. Abbott modeling representing the type of interaction between the Al2O3 and ZnO NPs in a binary mixture.
Cells TU Observed inhibition (%) Expected inhibition (%) Ratio of Inhibition (RI) Relation
L929 0.25 5.3 ± 3.80 7.25 ± 7.56 0.78 ± 0.82 Additive
0.50 12.25 ± 2.89 29.35 ± 16.05 0.46 ± 0.15 Additive
1 16.08 ± 5.12 65.40 ± 7.63 0.24 ± 0.04 Additive
1.50 25.70 ± 4.95 73.20 ± 2.54 0.35 ± 0.07* Antagonistic
1.75 34.29 ± 4.11 72.35 ± 2.89 0.47 ± 0.03* Antagonistic
2 43.50 ± 0.99 76.30 ± 3.67 0.57 ± 0.01* Antagonistic
BEAS-2B 0.25 21.9 ± 0.14 29.70 ± 3.81 0.74 ± 0.10 Additive
0.50 22.61 ± 1.69 28.15 ± 0.49 0.80 ± 0.04 Additive
1 39.46 ± 3.89 46.15 ± 1.48 0.85 ± 0.05 Additive
1.50 64.22 ± 2.85 67.50 ± 3.53 0.95 ± 0.09 Additive
1.75 64.83 ± 5.12 66.35 ± 3.74 0.98 ± 0.12 Additive
2 71.44 ± 3.91 76.60 ± 3.25 0.93 ± 0.01 Additive
*represent significative difference (p < 0.05) between toxicity induced by the binary mixture and expected toxicity.
6 J. S. KÖERICH ET AL.

Figure 2. Inhibitory effects of Al2O3 NPs, ZnO NPs, and a binary mixture on cells, measured by the MTT assay with L929 (a) and BEAS-
2B (b) cell lines.

Inhibition (RI) was calculated for the binary mix- Oxidative stress biomarkers
tures using both cell types. The RI values were less
than 1 and not significant from 1 in both cell types The ability of the mixture to induce oxidative stress
(Table 2) indicating an additive effect produced by was determined by measuring activities of CAT and
the binary mixtures at all TU conditions for BEAS- SOD as well as levels of ROS, GSH, and MDA in L929
2B cells and for 0.25 − 1 TU for L929 cells. and BEAS-2B cells incubated for 24 h (Figure 3).
However, for TU from 1.5 to 2 for L929 cells, A significant increase in ROS generation was noted
there was a significant antagonistic response. at the toxic units (TU) group 1 and 1.75 for both cells

Figure 3. Oxidative stress biomarkers (a) ROS, (b) CAT, (c) SOD, (d) GSH, and (e) MDA for L929 and BEAS-2B cells exposed to the
mixture (Al2O3+ZnO NPs) for 24 h. Values are mean ± SD (n=2). Different letters indicate significant differences between toxic units
for each cell type according to 2-way ANOVA results (p < 0.05).
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 7

compared to control and 0.25 TU (Figure 3a). Further, condensation visible as bright green patches, along
the mixture induced a significant elevation in CAT with fragments or apoptotic bodies), apoptotic cells
activity and reduction SOD activity, particularly in the (late apoptotic cells with orange to red nuclei with
case of 1.75 TU in BEAS-2B cells (Figure 3b,c). In highly fragmented chromatin), and necrotic cells (uni-
contrast, there was no marked difference in the activ- formly orange to red nuclei with organized structure,
ities of CAT and SOD in L929 cells. GSH levels were ascribable to necrotic cells).
significantly lower in both cell types from 1 TU for As shown in Figures 4a,b, in the case of the
BEAS-2B cells and 1.75 TU for L929 cells (Figure 3d). BEAS-2B cells there was a significant rise in apop-
Notably, there was a rise in lipoperoxidation (LPO) for totic cell number and diminished viable cells for 1
1 and 1.75 TU groups in BEAS-2B cells and 1.75 TU in and 1.75 TU compared to control. With respect to
L929 cells attributed to enhanced levels of MDA, the L929 cells, a significant elevation in apoptotic cell
by-product of LPO (Figure 3e). number and a significant decrease in viable cells
was found at 0.25 TU, indicating that this cell type
may be more sensitive than BEAS-2B.
Apoptosis and nuclear morphometry
Based on the ImageJ results (Figure 4c,d), the
Figure 4 illustrates the morphological changes char- cells showed a higher nuclear irregularity index
acteristic of viable cells (early apoptotic cells with and more apoptotic cells for 1 and 1.75 TU,
irregularly structured green nuclei and chromatin which is consistent with the results of the AO/IP

Figure 4. Induction of apoptosis in L929 and BEAS-2B cells exposed to the mixture (Al2O3 + ZnO NPs) for 24 h: (a) Percentage of
viable, apoptotic, and necrotic cells for L929 and (b) percentage of viable, apoptotic, and necrotic cells for BEAS-2B. Morphometric
analysis of nuclei of L929 and BEAS-2B cells exposed to the mixture for 24 h: (c) percentage of nuclei normal, apoptotic, and irregular
for L929 and (d) percentage of normal, apoptotic, and irregular nuclei for BEAS-2B. Scale bar, 200 µm. Values are mean ± SD (n=2).
Different letters indicate significant differences among toxic units of each cell according to 2-way ANOVA results (p<0.05).
8 J. S. KÖERICH ET AL.

staining assay. Once again, the L929 cells appeared (Figure 6b,c,i,j) associated with invaginations of
to be more sensitive than BEAS-2B cells, due to the cell surface.
the higher number of apoptotic nuclei and lower After 24 h treatment, the TEM images showed
number of normal nuclei when compared to the NP mixture within the endocytic vesicles in
BEAS-2B, this confirms cell death occurred at 1 different stages of maturation (Figure 6d,k) and
and 1.75 TU. large lysosomes with undigested metallic material.
The non-dissolution of NPs led to formation of
many residual bodies containing partially undi-
Clonogenic assay
gested material to be excreted (Figure 6c,e,f,h-L).

The colony-forming ability was used to determine


the long-term cytotoxic potential of the mixture
Discussion
(Al2O3+ ZnO NPs) in L929 and BEAS-2B cells
(Figure 5). A significant inhibition of colonies The study reported here is the first to assess the
was observed in L929 cells following a 24 h incu- characterization of the binary mixture of Al2O3
bation to 1 or 1.75 TU, equating to a 61.5% and and ZnO NP and their respective effect on cells
98.5%, respectively, reduction in survival. For related to human exposure. In the mixture, there
BEAS-2B cells, only exposure to 1.75 TU was was a greater agglomeration of Al2O3 than ZnO
found to significantly diminish colonies, to due to its larger estimated surface area and conse-
a 90.8% reduction in survival. quently its smaller diameter. In all cases, the EDs
were smaller when NPs were dissolved in UP
water compared to the RPMI medium since in
water there are more free ions on the NP surface
Cell uptake of NPs
and, consequently, NPs presented greater stability
TEM was used to determine the distribution and and smaller agglomerations (Melegari et al. 2019;
localization of the mixture in L929 and BEAS-2B Vicentini et al. 2017). However, some components
cells (Figure 6). The results obtained illustrate that such as serum albumin, amino acids, and ionic
cells are able to internalize the mixture since salts present in RPMI medium may influence ED
aggregates of various sizes interacted with the cell and stability of the NPs, resulting in the formation
membrane and were internalized by the cells of aggregates that allow them to disperse in larger

Figure 5. Clonogenic assay shows the long-term effects of the mixture (Al2O3 + ZnO NPs) on L929 and BEAS-2B cells. Representative
images of the colony formation (containing ?50 cells) after treatment with different TUs of the mixture for 10 d. (a–b) Graph shows
results for the percentage of colony forming units. Values are mean ± SEM (n=3). Different letters indicate significant differences in
the same groups after the exposure according to one-way ANOVA (p<0.05).
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 9

Figure 6. TEM cross-section images of randomly selected L929 and BEAS-2B cells after exposure to the mixture (Al2O3 + ZnO NPs) for
24 h, showing the location of the NPs. The ultrastructural morphology of plasma membrane invaginations/protrusions (d and k)
suggests clathrin-mediated endocytosis. (b, c, i and l) The NP aggregates can be seen inside vesicles (yellow arrow). Internalized NPs
took the same intracellular trafficking route in L929 and BEAS-2B cells: Endosomes with the mixture can be found in different stages
of maturation, (c and j) early endosome (EE), (e, h, and j) late endosome (LE) and (f and l) lysosome (l). (g) shows a more general
overview of control cell (untreated).

sizes, decreasing the exposed surface area Several investigators demonstrated that cyto-
(Maiorano et al. 2010; Nogueira et al. 2019; Wells toxicity is independent of the size of NPs (Lin
et al. 2012). At the same time, the zeta potential et al. 2008; Yuan et al. 2010). Therefore, it is
(ZP) values remained virtually unchanged indicat- essential to assess the toxicity of each specific
ing that there was no marked alteration in the material. After 24 h incubation, for the two types
physicochemical properties of the individual NPs of cells ZnO NPs were found to be induced greater
and their binary mixtures. toxicity. Previously, data reported that Al2O3 NPs
10 J. S. KÖERICH ET AL.

produced low toxicity toward mammalian cells 24 h. An increase in the ROS generation was noted
(Alshatwi et al. 2013; Nogueira et al. 2019; at TU 1 and 1.75 for both cell types. When the
Oesterling et al. 2008; Radziun et al. 2011). In generation of ROS exceeds the ability of the cell to
contrast, several in vitro studies demonstrated neutralize the effects of the radicals, the accumula-
that ZnO NPs exhibited significant cytotoxic tion of pro-oxidants occurs in the cell, leading to
effects, probably due to their solubility, leading to a state of oxidative stress (Wu et al. 2014). As the
an increase in the intracellular levels of Zn2+ oxidative stress in the cell rises, different biological
(Pandurangan and Kim 2015; Zhang et al. 2018). outcomes occur, such as a change in the activity of
With respect to the binary mixture, an antagonis- antioxidant enzymes CAT and SOD and levels of
tic effect was observed for L929 cell for the 1.5 to 2 GSH (Dayem et al. 2017). The mixture induced
TU suggesting interaction between Al2O3 and a significant elevation in CAT activity levels
ZnO NPs diminished the cytotoxic potency of accompanied by a decrease SOD activity and
the mixture. GSH levels in BEAS-2B cells. In contrast, there
Although the mechanisms underlying NPs- was no marked difference in the activity of CAT
mediated toxicity is not yet clear, one of the pos- and SOD in the L929 cells; however, the levels of
sible reasons for the high level of stress exerted by GSH were significantly reduced in L929 cells. The
ZnO NPs, rather than the interaction of the NPs oxidation of GSH indicates the inability of cells to
within the mixture, might be adsorption of Zn2+ scavenge ROS generated in response to the mix-
by the Al2O3 NPs (Sen and Sarzali 2008). Al2O3 ture, leading to the development of oxidative stress
NPs possess a relatively large surface area (88-fold in these cells, which may lead to oxidative damage
higher than of ZnO NPs), thereby providing sig- of the lipids (Li, Xia, and Nel 2008). In agreement
nificant potential for sorption of various metal our findings showed an increase in LPO in both
ions, inorganic anions, and organic ligands cells, as evidenced by elevated MDA levels, the by-
(Mahdavi, Jalali, and Afkhami 2013; Sen and product of LPO, indicative of damage the cell
Sarzali 2008). Previously Hua, Peijnenburg, and membranes (Birben et al. 2012). In several other
Vijver (2016) noted that the presence of TiO2 studies, LPO due to ROS generation was found to
NPs decreased Zn2+ toxicity toward zebrafish be associated with exposure to Al2O3 and ZnO
embryos. Dalai et al. (2014) showed that the pre- NPs, resulting in cellular toxicity (Alshatwi et al.
sence of Al2O3 NPs affected Cr(VI) toxicity, since 2013; Daoud Ali, Alkahtani, and Alarifi 2015; Król
the absorption of Cr(VI) by Al2O3 NPs was found et al. 2017; Li, Zhou, and Fan 2016; Nogueira et al.
to reduce Cr(VI) mediated toxicity toward the 2019; Roy, Das, and Dwivedi 2015; Srikanth et al.
freshwater microalgae Scenedesmus obliquus. 2015).
Therefore, the addition of Al2O3 NPs may affect Wang et al. (2015) indicated that increased ROS
the toxicity of ZnO NPs, by decreasing the toxicity levels may induce apoptosis. Our findings demon-
of ZnO NPs via a protective effect at higher con- strated a significant increase in apoptotic cell
centrations (Benavides et al. 2016; Dalai et al. numbers concomitant with a marked reduction
2014; Dávila-Grana et al. 2018). in viable cells in both cell types. Various investi-
The MTT assay was conducted to evaluate the gators noted that cellular apoptosis was initiated
cytotoxicity of the mixture (Al2O3+ ZnO NPs) by Al2O3 or ZnO NPs alone (Akhtar et al. 2012;
compared with each type of NP alone. However, Alshatwi et al. 2013; Nogueira et al. 2019; Wang
no synergistic effect was observed, probably et al. 2015). In both types of cells, there was an
because the MTT assay is a standard metabolic elevation in ROS and MDA levels accompanied by
assay based upon a simple colorimetric reaction a fall in SOD activity and GSH levels, which may
but limited by possible interference of the absor- have contributed to the observed apoptosis.
bance of particles (Almutary and Sanderson 2016). Based upon the ImageJ results, the cells exhibited
For this reason, the ability of the mixture to induce a higher nuclear irregularity index and more apop-
oxidative stress was assessed by measuring levels of totic cells after exposure to the binary mixture. The
ROS, GSH, and MDA as well as activities of CAT nuclear irregularity index, besides assessing apopto-
and SOD in L929 and BEAS-2B cells exposed for sis, is an indicator of defective activation or
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 11

inactivation of the cell cycle checkpoint signaling propagation of inflammation or induction of oxidative
processes (Filippi-Chiela et al. 2012). The L929 stress in the cells. As shown by Il et al. (2015), who
cells appeared to be more sensitive than BEAS-2B examined different particle shapes and sizes of Al2O3
cells, due to the higher number of apoptotic nuclei, NPs, cytotoxicity occurred via a mechanism involving
lower number of normal nuclei and decrease in lysosomal destabilization and generation of ROS, which
viable cells for a low UT compared to BEAS-2B. leads to mitochondrial perturbation resulting in cellular
This confirms the presence of cell death for 1 and death by apoptosis (Fink and Cookson 2005).
1.75 TU, as apoptotic cells showed nuclear conden-
sation and fragmentation (Filippi-Chiela et al. 2012).
Conclusions
Colony formation was inhibited in L929 cells
exposed at 1 and 1.75 TU and for BEAS-2B calls at The mixture at 1 and 1.75 toxic units (TU) was
1.75 TU possibly due to enhanced intracellular uptake more toxic toward both types of cells, even though
of the NPs through the endocytosis process. Taccola it presented antagonism at 1.75 TU toward the
et al. (2011) reported that endocytosis may occur L929 cell line. Antagonism may be attributed to
when ZnO NPs came into contact with the growth the protective effect of Al2O3, due to its large sur-
medium proteins that adhere to their surfaces. These face area and ion sorption potential, reducing the
proteins are then identified by cell surface receptors toxicity of ZnO. In addition, the L929 cells were
(Boucrot and Kirchhausen 2007). The interaction of found to be more sensitive than BEAS-2B cells due
NPs with cellular proteins is dependent upon cell to the higher number of apoptotic cells, larger
cycle phases. Patel et al. (2016) found higher ZnO apoptotic nuclei, and greater colony growth inhi-
uptake occurs in the G2/M cell phase and NPs induce bition predominantly at 1 and 1.75 TU. In sum-
cell cycle arrest by activating the ROS. For Al2O3, the mary, the results reported herein highlight the
reduction of cells in the G2/M cell phase might induce need to better understand how mixtures of differ-
the cell cycle to stop in the G0-G1 cell phases ent NPs affect specific cells in distinct ways.
(Periasamy, Athinarayanan, and Alshatwi 2016),
where cell growth occurs. According to the prolifera-
Acknowledgments
tion results, the cells were no longer able to undergo
normal cell division, which indicates that the cell cycle We would like to thank the staff of the Central Laboratory of
was affected following treatment with the NPs. Electron Microscopy (LCME) and Multiuser Laboratory of Biology
One of the most probable causes for the observed Studies (LAMEB) at the Federal University of Santa Catarina,
Florianópolis, SC, Brazil for assistance with microscopy analysis.
decrease in proliferation of L929 and BEAS-2B cells is
the high uptake of aggregates of NPs of various sizes
that was evident in the TEM results. The dynamics of Declaration of interest
uptake of Al2O3 NPs and ZnO NPs by cells was
The authors report no conflicts of interest. The authors are
comparable to that previously observed for various solely responsible for the content and writing of this article.
cell lines (Jeng and Swanson 2006; Kim, Baek, and
Choi 2010; Nogueira et al. 2019; Roy, Das, and
Dwivedi 2015; Zhang et al. 2013), suggesting that the Funding
uptake process may be induced by different mechan- This work was funded by the National Council for Scientific
isms such as phagocytosis, pinocytosis, macropinocyto- and Technological Development (CNPq) and Coordination
sis, or receptor-mediated endocytosis. In addition, most of Improvement of Higher Education Personnel (CAPES).
uptake processes involve lysosomes, where a variety of
macromolecules are degraded (Böhme et al. 2014).
ORCID
However, the mixture that accumulates in endosomes
and lysosomes may undergo dissolution in lysosomes Jéssica Schveitzer Köerich http://orcid.org/0000-0002-
and initiate lysosomal destabilization, which enables 8862-1105
Diego José Nogueira http://orcid.org/0000-0001-5831-
NPs to enter the cytosol (Stern, Adiseshaiah, and 4043
Crist 2012; Zhang et al. 2013). Finally, the TEM images Vitor Pereira Vaz http://orcid.org/0000-0003-3689-2656
displayed that the cytotoxic effect may be due to Carmen Simioni http://orcid.org/0000-0001-5629-9361
12 J. S. KÖERICH ET AL.

Marlon Luiz Neves Da Silva http://orcid.org/0000-0002- aluminum (Al2O3) and zinc (ZnO) oxide nanoparticles in
6602-2910 a freshwater fish. Carassius Auratus. Environ. Sci. Pollut. Res.
Luciane Cristina Ouriques http://orcid.org/0000-0003- 23:24578–91. doi:doi.10.1007/s11356-016-7915-3.
2964-6772 Beutler, E., O. Duron, and B. M. Kelly. 1963. Improved
Denice Schulz Vicentini http://orcid.org/0000-0001-7080- method for the determination of blood glutathione.
9911 J. Lab. Clin. Med. 61:882–88.
William Gerson Matias http://orcid.org/0000-0002-2386- Birben, E., U. M. Sahiner, C. Sackesen, S. Erzurum, and
0578 O. Kalayci. 2012. Oxidative stress and antioxidant
defense. World Allergy Organ. J. 5:9–19. doi:doi.10.1097/
WOX.0b013e3182439613.
References Böhme, S., H.-J. Stärk, T. Meißner, A. Springer, T. Reemtsma,
D. Kühnel, and W. Busch. 2014. Quantification of Al2O3
Abbott, W. S. 1925. A method of computing the effectiveness nanoparticles in human cell lines applying inductively
of an insecticide. J. Econ. Entomol. 18 (2):265–67. coupled plasma mass spectrometry (neb-ICP-MS, LA-ICP-
doi:10.1093/jee/18.2.265a. MS) and flow cytometry-based methods. J. Nanopart. Res.
Aebi, H. 1984. Catalase in vitro. Meth. Enzymol. 105:121–26. 16:2592. doi:doi.10.1007/s11051-014-2592-y.
doi:doi.10.1016/S0076-6879(84)05016-3. Boucrot, E., and T. Kirchhausen. 2007. Endosomal recycling
Akhtar, M. J., M. Ahamed, S. Kumar, M. A. Majeed Khan, controls plasma membrane area during mitosis. Proc. Natl.
J. Ahmad, and S. A. Alrokayan. 2012. Zinc oxide nanopar- Acad. Sci. U.S.A. 104:7939–44. doi:doi.10.1073/
ticles selectively induce apoptosis in human cancer cells pnas.0702511104.
through reactive oxygen species. Int. J. Nanomedicine Bradford, M. M. 1976. A rapid and sensitive method for the
7:845–57. doi:10.2147/IJN.S29129. quantitation of microgram quantities of protein utilizing
Alaraby, M., B. Annangi, R. Marcos, and A. Hernández. 2016. the principle of protein-dye binding. Anal. Biochem.
Drosophila melanogaster as a suitable in vivo model to 72:248–54. doi:doi.10.1016/0003-2697(76)90527-3.
determine potential side effects of nanomaterials: A Brunauer, S., P. H. Emmett, and E. Teller. 1938. Adsorption
review. J. Toxicol. Environ. Health B 19:65–104. doi: of gases in multimolecular layers. J. Am. Chem. Soc.
doi.10.1080/10937404.2016.1166466. 60:309–19. doi:doi.10.1021/ja01269a023.
Alim, S., J. Vejayan, M. M. Yusoff, and A. K. M. Kafi. 2018. Callister, C., M. Callister, M. Nolan, and E. Nolan. 2020.
Recent uses of carbon nanotubes & gold nanoparticles in Preventative agents: The multiple uses of silver nanoparti-
electrochemistry with application in biosensing: A review. cles in dentistry. Compend. Continuing Educ. Dent.
Biosens. Bioelectron. 121:125–36. doi:doi.10.1016/j. 41:143–47.
bios.2018.08.051. Cao, Q., Q. H. Hu, S. Khan, Z. J. Wang, A. J. Lin, X. Du, and
Almutary, A., and B. J. S. Sanderson. 2016. The MTT and Y. G. Zhu. 2007. Wheat phytotoxicity from arsenic and
crystal violet assays. Int. J. Toxicol. 35:454–62. doi: cadmium separately and together in solution culture and
doi.10.1177/1091581816648906. in a calcareous soil. J. Hazard. Mater. 148:377–82. doi:
Alshatwi, A. A., P. V. Subbarayan, E. Ramesh, A. A. Al- doi.10.1016/j.jhazmat.2007.02.050.
Hazzani, M. A. Alsaif, and A. A. Alwarthan. 2013. Chen, P., H. Wang, M. He, B. Chen, B. Yang, and B. Hu.
Aluminium oxide nanoparticles induce 2019. Size-dependent cytotoxicity study of ZnO nanopar-
mitochondrial-mediated oxidative stress and alter the ticles in HepG2 cells. Ecotoxicol. Environ. Saf. 171:337–46.
expression of antioxidant enzymes in human mesenchymal doi:10.1016/j.ecoenv.2018.12.096.
stem cells. Food Addit. Contam. A 30:1–10. doi: Colby, S. R. 1967. Calculating synergistic and antagonistic
doi.10.1080/19440049.2012.729160. responses of herbicide combinations. Weeds 15:20.
Angell, C., M. Kai, S. Xie, X. Dong, and Y. Chen. 2018. doi:10.2307/4041058.
Bioderived DNA nanomachines for potential uses in biosen- Crosera, M., G. Adami, M. Mauro, M. Bovenzi, E. Baracchini,
sing, diagnostics, and therapeutic applications. Adv. Healthc. and F. Larese Filon. 2016. In vitro dermal penetration of
Mater. 7:1701189. doi:doi.10.1002/adhm.201701189. nickel nanoparticles. Chemosphere 145:301–06. doi:
Balasubramanyam, A., N. Sailaja, M. Mahboob, doi.10.1016/j.chemosphere.2015.11.076.
M. F. Rahman, S. M. Hussain, and P. Grover. 2009. In Crosera, M., M. Bovenzi, G. Maina, G. Adami, C. Zanette,
vivo genotoxicity assessment of aluminium oxide nanoma- C. Florio, and F. Filon Larese. 2009. Nanoparticle dermal
terials in rat peripheral blood cells using the comet assay absorption and toxicity: A review of the literature. Int.
and micronucleus test. Mutagenesis 24:245–51. doi: Arch. Occup. Environ. Health 82:1043–55. doi:
doi.10.1093/mutage/gep003. doi.10.1007/s00420-009-0458-x.
Bank, H. L. 1988. Rapid assessment of islet viability with Cure, J., H. Assi, K. Cocq, L. Marìn, K. Fajerwerg, P. Fau,
acridine orange and propidium iodide. In Vitro Cell. Dev. E. Bêche, Y. J. Chabal, A. Estève, and C. Rossi. 2018.
Biol. 24:266–73. doi:doi.10.1007/BF02628826. Controlled growth and grafting of high-density Au nanopar-
Benavides, M., J. Fernández-Lodeiro, P. Coelho, C. Lodeiro, ticles on zinc oxide thin films by photo-deposition. Langmuir
and M. S. Diniz. 2016. Single and combined effects of 34:1932–40. doi:doi.10.1021/acs.langmuir.7b04105.
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 13

Dalai, S., S. Pakrashi, M. Bhuvaneshwari, V. Iswarya, nanoparticles and Zn ions on zebrafish embryos (Danio
N. Chandrasekaran, and A. Mukherjee. 2014. Toxic effect rerio). NanoImpact 2:45–53. doi:doi.10.1016/j.
of Cr(VI) in presence of n-TiO2 and n-Al2O3 particles impact.2016.06.005.
towards freshwater microalgae. Aquat. Toxicol. 146:28–37. Il, K. B., Y. H. Joo, P. J. Pak, J.-S. Kim, and N. Chung. 2015.
doi:10.1016/j.aquatox.2013.10.029. Different shapes of Al2O3 particles induce differential
Daoud Ali, D., S. Alkahtani, and S. Alarifi. 2015. cytotoxicity via a mechanism involving lysosomal destabi-
Nanoalumina induces apoptosis by impairing antioxidant lization and reactive oxygen species generation. J. Korean
enzyme systems in human hepatocarcinoma cells. Int. Soc. Appl. Biol. Chem. 58:433–42. doi:doi.10.1007/s13765-
J. Nanomedicine 10:3751. doi:10.2147/IJN.S82050. 015-0038-6.
Dávila-Grana, Á., L. Diego-González, Á. González- Iswarya, V., M. Bhuvaneshwari, N. Chandrasekaran, and
Fernández, and R. Simón-Vázquez. 2018. Synergistic effect A. Mukherjee. 2016. Individual and binary toxicity of
of metal oxide nanoparticles on cell viability and activation anatase and rutile nanoparticles towards Ceriodaphnia
of MAP kinases and NFκB. Int. J. Mol. Sci. 19:246. doi: dubia. Aquat. Toxicol. 178:209–21. doi:10.1016/j.
doi.10.3390/ijms19010246. aquatox.2016.08.007.
Dayem, A. A., M. K. Hossain, S. Bin Lee, K. Kim, S. K. Saha, Ivask, A., T. Titma, M. Visnapuu, H. Vija, A. Kakinen,
G. M. Yang, H. Y. Choi, and S. G. Cho. 2017. The role of M. Sihtmae, S. Pokhrel, L. Madler, M. Heinlaan, V. Kisand,
reactive oxygen species (ROS) in the biological activities of et al. 2015. Toxicity of 11 metal oxide nanoparticles to three
metallic nanoparticles. Int. J. Mol. Sci. 18. doi:doi.10.3390/ mammalian cell types in vitro. Curr. Top Med. Chem.
ijms18010120. 15:1914–29. doi:10.2174/1568026615666150506150109.
Demir, E., A. Creus, and R. Marcos. 2014. Genotoxicity and Jain, K., E. Kohli, D. Prasad, K. Kamal, S. M. Hussain, and
DNA repair processes of zinc oxide nanoparticles. S. B. Singh. 2015. In vitro cytotoxicity assessment of metal
J. Toxicol. Environ. Health A 77:1292–303. doi:10.1080/ oxide nanoparticles. Nanomed. Nanobiol. 1:10–19.
15287394.2014.935540. doi:10.1166/nmb.2014.1003.
Donaldson, K., V. Stone, A. Clouter, L. Renwick, and Jeng, H. A., and J. Swanson. 2006. Toxicity of metal oxide
W. MacNee. 2001. Ultrafine particles. Occup. Environ. nanoparticles in mammalian cells. J. Environ. Sci. Health A
Med. 58:211–16. doi:doi.10.1136/oem.58.3.211. 41:2699–2611. doi:doi.10.1080/10934520600966177.
Draper, H. H., and M. Hadley. 1990. Malondialdehyde deter- Kar, K. K., S. Srivastava, A. Rahaman, and S. K. Nayak. 2008.
mination as index of lipid peroxidation. Meth. Enzymol. Acrylonitrile-butadiene-styrene nanocomposites filled with
186:421–31. doi:doi.10.1016/0076-6879(90)86135-I. nanosized alumina. Polym. Compos. 29:489–99. doi:
Duran, N., M. Duran, M. B. DeJesus, A. B. Seabra, W. J. Favaro, doi.10.1002/pc.20387.
and G. Nakazato. 2016. Silver nanoparticles: A new view on Kaweeteerawat, C., P. N. Ubol, S. Sangmuang,
mechanistic aspects of antimicrobial activity. Nanomedicine A. Auerviriyavit, and R. Maniratanachote. 2017.
12:789–99. doi:10.1016/j.nano.2015.11.016. Mechanisms of antibiotic resistance in bacteria mediated
Ferreira, A. J., J. Cemlyn-Jones, and C. Robalo Cordeiro. by silver nanoparticles. J. Toxicol. Environ. Health A
2013. Nanoparticles, nanotechnology and pulmonary 80:1276–89. doi:10.1080/15287394.2017.1376727.
nanotoxicology. Rev. Port. Pneumol. 19:28–37. doi: Kengar, M. D., A. A. Jadhav, S. B. Kumbhar, and
doi.10.1016/j.rppneu.2012.09.003. R. P. Jadhav. 2019. A review on nanoparticles and its
Filippi-Chiela, E. C., M. M. Oliveira, B. Jurkovski, application. Asian J. Pharm. Technol. 9:115. doi:
S. M. Callegari-Jacques, V. D. da Silva, and G. Lenz. doi.10.5958/2231-5713.2019.00020.5.
2012. Nuclear morphometric analysis (NMA): Screening Kermanizadeh, A., I. Gosens, L. MacCalman, H. Johnston,
of senescence, apoptosis and nuclear irregularities. PLoS P. H. Danielsen, N. R. Jacobsen, A. G. Lenz, T. Fernandes,
ONE 7. doi:doi.10.1371/journal.pone.0042522. R. P. F. Schins, F. R Cassee et al. 2016. A multilaboratory
Fink, S. L., and B. T. Cookson. 2005. Apoptosis, pyroptosis, toxicological assessment of a panel of 10 engineered nano-
and necrosis: Mechanistic description of dead and dying materials to human health - ENPRA project - the high-
eukaryotic cells. Infect. Immun. 73:1907–16. doi: lights, limitations, and current and future challenges.
doi.10.1128/IAI.73.4.1907-1916.2005. J. Toxicol. Environ. Health B 19:1–28. doi:doi.10.1080/
Franken, N. A. P., H. M. Rodermond, J. Stap, J. Haveman, 10937404.2015.1126210.
and C. van Bree. 2006. Clonogenic assay of cells in vitro. Keston, A. S., and R. Brandt. 1965. The fluorometric analysis
Nat. Protoc. 1:2315–19. doi:doi.10.1038/nprot.2006.339. of ultramicro quantities of hydrogen peroxide. Anal.
Hashimoto, M., J. I. Sasaki, and S. Imazato. 2016. Biochem. 11:1–5. doi:doi.10.1016/0003-2697(65)90034-5.
Investigation of the cytotoxicity of aluminum oxide nano- Kim, I.-S., M. Baek, and S.-J. Choi. 2010. Comparative cyto-
particles and nanowires and their localization in L929 toxicity of Al2O3, CeO2, TiO2 and ZnO nanoparticles to
fibroblasts and RAW264 macrophages. J. Biomed. Mater. human lung cells. J. Nanosci. Nanotechnol. 10:3453–58.
Res. B: Appl. Biomater. 104:241–52. doi:doi.10.1002/jbm. doi:doi.10.1166/jnn.2010.2340.
b.33377. Krewski, D., R. A. Yokel, E. Nieboer, D. Borchelt, J. Cohen,
Hua, J., W. J. G. M. Peijnenburg, and M. G. Vijver. 2016. J. Harry, S. Kacew, J. Lindsay, A. M. Mahfouz, and
TiO2 nanoparticles reduce the effects of ZnO V. Rondeau. 2007. Human health risk assessment for
14 J. S. KÖERICH ET AL.

aluminium, aluminium oxide, and aluminium hydroxide. Ng, C. T., L. Q. Yong, M. P. Hande, C. N. Ong, L. E. Yu,
J. Toxicol. Environ. Health B 10:1–269. doi:doi.10.1080/ B. H. Bay, and G. H. Baeg. 2017. Zinc oxide nanoparticles
10937400701597766. exhibit cytotoxicity and genotoxicity through oxidative
Król, A., P. Pomastowski, K. Rafińska, V. Railean-Plugaru, stress responses in human lung fibroblasts and
and B. Buszewski. 2017. Zinc oxide nanoparticles: Drosophila melanogaster. Int. J. Nanomedicine
Synthesis, antiseptic activity and toxicity mechanism. 12:1621–37. doi:doi.10.2147/IJN.S124403.
Adv. Colloid Interface Sci. 249:37–52. doi:10.1016/j. Niska, K., E. Zielinska, M. W. Radomski, and I. Inkielewicz-
cis.2017.07.033. Stepniak. 2018. Metal nanoparticles in dermatology and
Li, L., M. L. Fernández-Cruz, M. Connolly, E. Conde, cosmetology: Interactions with human skin cells. Chem.
M. Fernández, M. Schuster, and J. M. Navas. 2015. The Biol. Interact. 295:38–51. doi:10.1016/j.cbi.2017.06.018.
potentiation effect makes the difference: Non-toxic con- Nogueira, D. J., M. Arl, J. S. Köerich, C. Simioni,
centrations of ZnO nanoparticles enhance Cu nanoparticle L. C. Ouriques, D. S. Vicentini, and W. G. Matias. 2019.
toxicity in vitro. Sci. Total Environ. 505:253–60. Comparison of cytotoxicity of α-Al2O3 and η-Al2O3
doi:10.1016/j.scitotenv.2014.10.020. nanoparticles toward neuronal and bronchial cells.
Li, N., T. Xia, and A. E. Nel. 2008. The role of oxidative stress Toxicol. In Vitro 61. doi:doi.10.1016/j.tiv.2019.104596.
in ambient particulate matter-induced lung diseases and its Oesterling, E., N. Chopra, V. Gavalas, X. Arzuaga, E. J. Lim,
implications in the toxicity of engineered nanoparticles. R. Sultana, D. A. Butterfield, L. Bachas, and B. Hennig.
Free Radic. Biol. Med. 44:1689–99. doi:doi.10.1016/j. 2008. Alumina nanoparticles induce expression of
freeradbiomed.2008.01.028. endothelial cell adhesion molecules. Toxicol. Lett.
Li, X., S. Zhou, and W. Fan. 2016. Effect of nano-Al2O3 on 178:160–66. doi:10.1016/j.toxlet.2008.03.011.
the toxicity and oxidative stress of copper towards Pandurangan, M., and D. H. Kim. 2015. In vitro toxicity of
Scenedesmus obliquus. Int. J. Environ. Res. Public Health zinc oxide nanoparticles: A review. J. Nanopart. Res. 17.
13:575. doi:doi.10.3390/ijerph13060575. doi:doi.10.1007/s11051-015-2958-9.
Lin, W., I. Stayton, Y. W. Huang, X. D. Zhou, and Y. Ma. Patel, P., K. Kansara, V. A. Senapati, R. Shanker, A. Dhawan,
2008. Cytotoxicity and cell membrane depolarization and A. Kumar. 2016. Cell cycle dependent cellular uptake
induced by aluminum oxide nanoparticles in human lung of zinc oxide nanoparticles in human epidermal cells.
epithelial cells A549. Toxicol. Environ. Chem. 90:983–96. Mutagenesis 31:481–90. doi:doi.10.1093/mutage/gew014.
doi:doi.10.1080/02772240701802559. Periasamy, V. S., J. Athinarayanan, and A. A. Alshatwi. 2016.
Mahdavi, S., M. Jalali, and A. Afkhami. 2013. Heavy metals Aluminum oxide nanoparticles alter cell cycle progression
removal from aqueous solutions using TiO2, MgO, and through CCND1 and EGR1 gene expression in human
Al2O3 nanoparticles. Chem. Eng. Commun. 200:448–70. mesenchymal stem cells. Biotechnol. Appl. Biochem.
doi:doi.10.1080/00986445.2012.686939. 63:320–27. doi:doi.10.1002/bab.1368.
Maiorano, G., S. Sabella, B. Sorce, V. Brunetti, Radziun, E., J. Dudkiewicz Wilczyńska, I. Ksiazek, K. Nowak,
M. A. Malvindi, R. Cingolani, and P. P. Pompa. 2010. E. L. Anuszewska, A. Kunicki, A. Olszyna, and
Effects of cell culture media on the dynamic formation of T. Zabkowski. 2011. Assessment of the cytotoxicity of
protein−nanoparticle complexes and influence on the cel- aluminium oxide nanoparticles on selected mammalian
lular response. ACS Nano 4:7481–91. doi:doi.10.1021/ cells. Toxicol. In Vitro 25:1694–700. doi:doi.10.1016/j.
nn101557e. tiv.2011.07.010.
Mancuso, L., C. Manis, A. Murgia, M. Isola, A. Salis, F. Piras, Rajput, V. D., T. M. Minkina, A. Behal, S. N. Sushkova,
P. Caboni, and G. Cao. 2018. Effect of ZnO nanoparticles S. Mandzhieva, R. Singh, A. Gorovtsov,
on human bone marrow mesenchymal stem cells: V. S. Tsitsuashvili, W. O. Purvis, K. A. Ghazaryan, et al.
Viability, morphology, particles uptake, cell cycle and 2018. Effects of zinc-oxide nanoparticles on soil, plants,
metabolites. Biosci. Biotechnol. Res. Asia 15:751–65. doi: animals and soil organisms: A review. Environ.
doi.10.13005/bbra/2684. Nanotechnol. Monit. Manage. 9:76–84. doi:10.1016/j.
McCord, J. M., and I. Fridovich. 1969. Superoxide dismutase enmm.2017.12.006.
an enzymic function for erythrocuprein (hemocuprein). Roy, R., M. Das, and P. D. Dwivedi. 2015. Toxicological
J. Biol. Chem. 244:6049–55. mode of action of ZnO nanoparticles: Impact on immune
Melegari, S. P., C. F. Fuzinatto, R. A. Gonçalves, B. V. Oscar, cells. Mol. Immunol. 63:184–92. doi:doi.10.1016/j.
D. S. Vicentini, and W. G. Matias. 2019. Can the surface molimm.2014.08.001.
modification and/or morphology affect the ecotoxicity of Sen, T. K., and M. V. Sarzali. 2008. Removal of cadmium
zinc oxide nanomaterials? Chemosphere 224:237–46. metal ion (Cd2+) from its aqueous solution by aluminium
doi:10.1016/j.chemosphere.2019.02.093. oxide (Al2O3): A kinetic and equilibrium study. Chem.
Mosmann, T. 1983. Rapid colorimetric assay for cellular Eng. J. 142:256–62. doi:10.1016/j.cej.2007.12.001.
growth and survival: Application to proliferation and cyto- Song, Z. M., H. Tang, X. Deng, K. Xiang, A. Cao, Y. Liu, and
toxicity assays. J. Immunol. Methods 65:55–63. doi:10.1016/ H. Wang. 2017. Comparing toxicity of alumina and zinc
0022-1759(83)90303-4. oxide nanoparticles on the human intestinal epithelium
JOURNAL OF TOXICOLOGY AND ENVIRONMENTAL HEALTH, PART A 15

in vitro model. J. Nanosci. Nanotechnol. 17:2881–91. doi: Wells, M. A., A. Abid, I. M. Kennedy, and A. I. Barakat. 2012.
doi.10.1166/jnn.2017.13056. Serum proteins prevent aggregation of Fe2O3 and ZnO
Sprague, J. B., and B. A. Ramsay. 1965. Lethal levels of mixed nanoparticles. Nanotoxicology 6:837–46. doi:10.3109/
copper–zinc solutions for juvenile salmon. J. Fish. Res. 17435390.2011.625131.
Board Can. 22:425–32. doi:doi.10.1139/f65-042. Wiesner, M. R., G. V. Lowry, P. Alvarez, D. Dionysiou, and
Srikanth, K., A. Mahajan, E. Pereira, A. C. Duarte, and P. Biswas. 2006. Assessing the risks of manufactured
J. Venkateswara Rao. 2015. Aluminium oxide nanoparti- nanomaterials. Environ. Sci. Technol. 40:4336–45. doi:
cles induced morphological changes, cytotoxicity and oxi- doi.10.1021/es062726m.
dative stress in chinook salmon (CHSE-214) cells. J. Appl. Wu, H., J. J. Yin, W. G. Wamer, M. Zeng, and Y. M. Lo. 2014.
Toxicol. 35:1133–40. doi:doi.10.1002/jat.3142. Reactive oxygen species-related activities of nano-iron
Stern, S. T., P. P. Adiseshaiah, and R. M. Crist. 2012. metal and nano-iron oxides. J. Food Drug Anal. 22:86–94.
Autophagy and lysosomal dysfunction as emerging doi:doi.10.1016/j.jfda.2014.01.007.
mechanisms of nanomaterial toxicity. Part. Fibre Toxicol. Ye, N., Z. Wang, H. Fang, S. Wang, and F. Zhang. 2017.
9:1. doi:doi.10.1186/1743-8977-9-20. Combined ecotoxicity of binary zinc oxide and copper oxide
Su, H., Y. Wang, Y. Gu, L. Bowman, J. Zhao, and M. Ding. nanoparticles to Scenedesmus obliquus. J. Environ. Sci. Health A
2018. Potential applications and human biosafety of nano- 52:555–5. doi:doi.10.1080/10934529.2017.1284434.
materials used in nanomedicine. J. Appl. Toxicol. 38:3–24. Yu, R., J. Wu, M. Liu, G. Zhu, L. Chen, Y. Chang, and H. Lu.
doi:doi.10.1002/jat.3476. 2016. Toxicity of binary mixtures of metal oxide nanopar-
Taccola, L., V. Raffa, C. Riggio, O. Vittorio, M. C. Iorio, ticles to Nitrosomonas europaea. Chemosphere 153:187–97.
R. Vanacore, A. Pietrabissa, and A. Cuschieri. 2011. Zinc doi:10.1016/j.chemosphere.2016.03.065.
oxide nanoparticles as selective killers of proliferating cells. Yuan, J. H., Y. Chen, H. X. Zha, L. J. Song, C. Y. Li, J. Q. Li, and
Int. J. Nanomedicine 6:1129–40. doi:doi.10.2147/IJN. X. H. Xia. 2010. Determination, characterization and cytotoxi-
S16581. city on HELF cells of ZnO nanoparticles. Colloids Surf.
Vicentini, D. S., R. C. Puerari, K. G. Oliveira, M. Arl, B Biointerfaces 76:145–50. doi:10.1016/j.colsurfb.2009.10.028.
S. P. Melegari, and W. G. Matias. 2017. Toxicological Zhang, Q., L. Xu, J. Wang, E. Sabbioni, L. Piao, M. Di
impact of morphology and surface functionalization of Gioacchino, and Q. Niu. 2013. Lysosomes involved in the
amorphous SiO2 nanomaterials. NanoImpact 5:6–12. cellular toxicity of nano-alumina: Combined effects of
doi:10.1016/j.impact.2016.11.003. particle size and chemical composition. J. Biol. Regul.
Vicentini, D. S., A. Smania, and M. C. M. Laranjeira. 2010. Homeost. Agents 27:365–75.
Chitosan/poly (vinyl alcohol) films containing ZnO nano- Zhang, X. Q., L. H. Yin, M. Tang, and Y. P. Pu. 2011. ZnO,
particles and plasticizers. Mater. Sci. Eng. C 30:503–08. TiO2, SiO2, and Al2O3 nanoparticles-induced toxic effects
doi:10.1016/j.msec.2009.01.026. on human fetal lung fibroblasts. Biomed. Environ. Sci.
Wang, C., X. Hu, Y. Gao, and Y. Ji. 2015. ZnO nanoparticles 24:661–69. doi:doi.10.3967/0895-3988.2011.06.011.
treatment induces apoptosis by increasing intracellular Zhang, Z. J. I., Z. J. Tang, Z. Y. Zhu, Z. M. Cao, H. J. Chen,
ROS levels in LTEP-a-2 cells. Biomed. Res. Int. 2015. doi: W. J. Zheng, X. Hu, H. Z. Lian, and L. Mao. 2018. The
doi.10.1155/2015/423287. time-dependent cellular response mechanism upon expo-
Wang, Y., and B. Nowack. 2018. Dynamic probabilistic material sure to zinc oxide nanoparticles. J. Nanopart. Res. 20. doi:
flow analysis of nano-SiO2, nano iron oxides, nano-CeO2, doi.10.1007/s11051-018-4333-0.
nano-Al2O3, and quantum dots in seven European regions. Zhao, J., and V. Castranova. 2011. Toxicology of nanomater-
Environ. Pollut. 235:589–601. doi:10.1016/j.envpol.201 ials used in nanomedicine. J. Toxicol. Environ. Health B
8.01.004. 14:593–532. doi:doi.10.1080/10937404.2011.615113.

You might also like