You are on page 1of 9

Toxicology Letters 199 (2010) 389–397

Contents lists available at ScienceDirect

Toxicology Letters
journal homepage: www.elsevier.com/locate/toxlet

Role of the dissolved zinc ion and reactive oxygen species in cytotoxicity
of ZnO nanoparticles
Wenhua Song a,b,∗ , Jinyang Zhang b , Jing Guo a , Jinhua Zhang a , Feng Ding c , Liying Li a , Zengtian Sun a
a
School of Environmental and Chemical Engineering, Tianjin Polytechnic University, Tianjin, 300160, China
b
College of Environmental Science and Engineering, Shanghai Jiaotong University, Shanghai, 200240, China
c
College of Environmental Science and Engineering, Nankai University, Tianjin 300071, China

a r t i c l e i n f o a b s t r a c t

Article history: With large-scale production and wide application of nanoscale ZnO, its health hazard has attracted exten-
Received 21 July 2010 sive worldwide attention. In this study, cytotoxicity of different sized and shaped ZnO nanoparticles in
Received in revised form mouse macrophage Ana-1 was investigated. And contribution of dissolved Zn2+ and ROS in toxicity of
28 September 2010
ZnO particles was analyzed. The results indicated that ZnO particles manifested dose-dependent toxic
Accepted 1 October 2010
effect on Ana-1 cells without size-dependence, and the particles shape may impact cytotoxicity of ZnO
Available online 8 October 2010
particles. When the concentration of dissolved Zn2+ tended to equilibrium in the complete cell medium,
the zinc ion concentration was approximately 10 ␮g/ml, inducing about 50% cell death, which was close
Keywords:
ZnO nanoparticles
to the cytotoxicity of ZnCl2 (IC50 = 13.33 ␮g Zn/ml). The Zn2+ concentration had significant correlations
Ana-1 cell with cell viability and LDH level induced by the supernatant of ZnO particle suspensions (incubation at
Cytotoxicity 37 ◦ C for 24 h). Thus, the dissolved Zn2+ played the main role in toxic effect of ZnO particles. Moreover,
ROS ROS generation assays demonstrated that ZnO particles produced intrinsically a small quantity of ROS,
Zinc ion intracellular ROS was mainly produced after ZnO particles or the dissolved Zn2+ entered into the cells.
Although intracellular ROS had significant correlations with cell viability and LDH induced by ZnO par-
ticles, intracellular ROS may not be a major factor in cytotoxicity of ZnO nanoparticles, but the cytotoxic
response.
© 2010 Elsevier Ireland Ltd. All rights reserved.

1. Introduction and 120 nm ZnO oral exposure (Wang et al., 2008). And exposures
to nanoscale or fine-sized ZnO particles by intratracheal instilla-
Nano scale ZnO has become a nano-functional material, widely tion produced potent but typical “metal fume fever”-like reversible
concerned following carbon nanotubes. It is one of the materials inflammation (Sayes et al., 2009). Moreover, many in vitro studies
with well-established production process and test method, and also also demonstrated that ZnO nanoparticles are toxic to mammalian
is one of the materials of first commercial application. Since the cells. For example, ZnO nanoparticles induced toxicity in RAW264.7
Bayer Corporation (Bayer Co., Ltd.) first provided with nano-ZnO and BEAS-2B cells, leading to the generation of reactive oxygen
products in the market, various countries started the large-scale species (ROS), oxidant injury, excitation of inflammation, and cell
industrial production of nano-ZnO. Nano-ZnO is widely used in death, and ZnO dissolution could happen in culture medium and
many fields, such as rubber manufacture, cosmetics, pigments, food endosomes (Xia et al., 2008). In addition, some studies have demon-
additives, medicine, chemical fiber and electronics industries (Ji strated that nano-ZnO is more toxic than other nanometer-scale
and Ye, 2008). However, theoretical study of nano-ZnO lags behind structured metallic oxides (Jeng and Swanson, 2006; Lai et al., 2008;
its production and application. And due to the traditional concept Horie et al., 2009).
that zinc oxide is non-toxic, the toxicology study of nano-ZnO is However, there are still a lot of controversies on the toxicity
farther behind the speed of its application development. In recent of ZnO nanoparticles. Firstly, there is not a consistent conclusion
years, with the rise and development of nanotoxicology, health on toxicity of ZnO nanoparticles and the impact of particle size on
risks of nano-ZnO are gradually concerned. The study on the acute toxicity of ZnO nanoparticles. Some research showed that differ-
toxicological impact of ZnO nanoparticles on mice showed that the ent sized ZnO nanoparticles had similar cytotoxicity on different
liver, spleen, heart, pancreas and bone are the target organs for 20 cells, 24 h IC50 was about 10–20 ␮g/ml, and particle size had no
effect on cytotoxicity (Lin et al., 2009; Deng et al., 2009; Yuan et
al., 2010). Nair et al. found 100 ␮M (8.1 ␮g/ml) ZnO nanoparticles
∗ Corresponding author at: School of Environmental and Chemical Engineering, induced osteoblast cancer cells (MG-63) viability decrease to about
Tianjin Polytechnic University, Tianjin, 300160, China. 40%, and when particle size was under 350 nm, the size did not
E-mail address: songwenhua9316@sina.com (W. Song). impact cytotoxicity, but nanoparticles were more toxic to MG-63

0378-4274/$ – see front matter © 2010 Elsevier Ireland Ltd. All rights reserved.
doi:10.1016/j.toxlet.2010.10.003
390 W. Song et al. / Toxicology Letters 199 (2010) 389–397

cell than microparticles (Nair et al., 2009). Reddy et al. reported Crystal structure of ZnO particles was characterized using X-ray diffraction (XRD,
ZnO nanoparticles (13 nm) still were not cytotoxic at 5 mmol/L Bruker D8 Discover).
(400 ␮g/ml), only at 10 mmol/L (800 ␮g/ml), ZnO nanoparticles
2.2. Cell culture and treatment with ZnO particles
caused 57% cell death. And cytotoxicity is limited to ZnO in the
nanoscale size range as no significant effect of bulk ZnO powder Mouse macrophage (Ana-1) cell line was purchased from the Cell Bank of Type
was observed (Reddy et al., 2007; Hanley et al., 2008). Secondly, Culture Collection of Chinese Academy of Sciences. Cells were maintained in the
the effect of particle shape on cytotoxicity of ZnO nanoparticles is complete cell medium, RPMI1640 supplemented with 10% fetal bovine serum (FBS),
100 U/ml penicillin, 100 ␮g/ml streptomycin, and grown at 37 ◦ C in a humidified 5%
not yet clear. Lee et al. found spherical ZnO particles induced lower
CO2 environment.
toxicity than rod-shaped ZnO particles (Lee et al., 2008). However in A stock solution up to 1 mg ZnO/ml complete cell medium was prepared to
the study of Lee et al., ZnO exposure concentration was based on the yield a more accurate weight of ZnO powder, dispersed for 20 min by a sonicator to
bottom area of culture plate, if mass-based dosage conversed, the prevent aggregation, and diluted to the specified concentrations (2.5, 5, 10, 20, 40,
exposure concentration of rod ZnO particles was much higher than 80, 100 ␮g/ml) for treatment of cells.

spherical ZnO particles. Nair et al. showed that rod ZnO nanoparti-
2.3. Cell viability assay
cles were lower toxic than spherical ZnO nanoparticles, but rod ZnO
nanoparticles were coated by starch and spherical ZnO nanopar- Cellular viability was determined using the CCK-8 assay (Beyotime). Cells were
ticles were coated by poly ethylene glycol (PEG), starch and PEG seeded with equal density in each well of 96-well plates (104 cells per well), 100 ␮l
of the cell culture medium per well, and incubated for 24 h at 37 ◦ C. Then cells
may impact on toxicity of ZnO nanoparticles (Nair et al., 2009).
were treated in 96-well plates with 2.5 ␮g/ml, 5.0 ␮g/ml, 10.0 ␮g/ml, 20.0 ␮g/ml,
In addition, now there are main two explanations on the toxicity 40.0 ␮g/ml, 80.0 ␮g/ml or 100.0 ␮g/ml of ZnO particles for 24 h at 37 ◦ C. Untreated
mechanism of ZnO nanoparticles: zinc dissolution and oxidative cells served as a control group. At the end of the treatment, CCK-8 dye was added to
damage, but the contribution of two sides is uncertain. Deng et al. each well and the plates were incubated for another 2 h at 37 ◦ C. To prevent particles
found that ZnO nanoparticles and ZnCl2 had similar toxicity, and from interfering with this assay, the solution in each well of each plate was quanti-
tatively transferred to an empty well in another plate. Subsequently, the absorbance
did not observe ZnO particles within cells with the TEM (Deng et
was measured by dual wavelength spectrophotometry at 450 nm and 630 nm using
al., 2009). Xia et al. also suggested that zinc ion played an impor- a microplate reader. Each treatment was repeated five times.
tant role in cytotoxicity of ZnO nanoparticles, and did not find ZnO In order to investigate the effect of zinc ions released from ZnO particles on
nanoparticles in cells (Xia et al., 2008). However, Jeng and Swan- cell viability, 2.5 ␮g/ml, 5.0 ␮g/ml, 10.0 ␮g/ml, 20.0 ␮g/ml, 40.0 ␮g/ml, 80.0 ␮g/ml
or 100.0 ␮g/ml of ZnO particle suspensions was incubated at 37 ◦ C in a humidi-
son compared cytotoxicity of a variety of metal oxide nanoparticles
fied 5% CO2 environment for 24 h, and then centrifuged at 20,000 rpm for 30 min.
on the Neuro-2A cell, and considered oxidative damage had an The supernatant was used to treat cells, and the cell viability was assayed by
important effect on cytotoxicity (Jeng and Swanson, 2006). Lin CCK-8. Moreover, the viability of Ana-1 exposed to ZnCl2 was also determined.
et al. showed nanoparticles and microparticles of ZnO decreased Because ZnO nanoparticles were similar toxicity to ZnCl2 in some studies (Franklin
A549 cell viability, and observed particles in cells with TEM. They et al., 2007; Deng et al., 2009), the concentrations of ZnCl2 were set accord-
ing to the concentration of ZnO particles, with 4.2 ␮g/ml, 8.4 ␮g/ml, 16.8 ␮g/ml,
still found toxicity of ZnO particles was reduced by antioxidant N-
33.6 ␮g/ml, 67.2 ␮g/ml, 134.4 ␮g/ml or 168.0 ␮g/ml (correspond to 2.5 ␮g ZnO/ml,
Acetylcysteine, this demonstrated toxicity of ZnO particles related 5.0 ␮g ZnO/ml, 10.0 ␮g ZnO/ml, 20.0 ␮g ZnO/ml, 40.0 ␮g ZnO/ml, 80.0 ␮g ZnO/ml or
with oxidative damage. Moreover, they considered that the low 100.0 ␮g ZnO/ml).
dissociation of ZnO in the cell culture medium (Ksp = 3.0 × 10−16 )
was not enough to cause cell toxicity. Thus, they inferred that 2.4. LDH measurement

oxidative damage induced by ZnO particles was due to sequential


Release of lactate dehydrogenase (LDH) to the cell culture medium indicates
oxidation–reduction reactions occurring at ZnO particles surface to cell membrane damage. LDH level in the cell culture medium was determined
produce reactive species (Lin et al., 2009). using a LDH Kit (Jiancheng Bioengineering Co. Ltd., Nanjing, China) according to the
Macrophages were multifunctional immune cells. Their phago- manufacturer’s protocols. LDH catalyzed the oxidation of lactate to pyruvate with
simultaneous reduction of NAD+ to NADH. The rate of NAD+ reduction is directly
cytosis was a main pathway to clean up particles in vivo, and
proportional to LDH level. The treatment method of cells was the same as the
macrophages had a variety of stimulating rapid response capaci- CCK-8 assay, but a cell culture medium volume of 200 ␮l was used to provide suffi-
ties, played an important part in regulating the immune response cient volume for LDH detection. After ZnO exposure, half the amount of the 200 ␮l
and inflammation. ZnO nanoparticles were widely added to the cell culture medium was collected for LDH analysis. Absorption was measured at
feed as additives. Ding et al. showed that the oral exposure to low 340 nm.

dosage of ZnO nanoparticles in mice increased the phagocytosis of


2.5. Intracellular reactive oxygen species measurement
mouse peritoneal macrophages (Ding et al., 2007). However, the
oral exposure to high dosage of ZnO nanoparticles induced the The intracellular reactive oxygen species (ROS) was determined using a well-
edema and degeneration of hepatocytes, inflammation of pancreas characterized probe, 2 ,7 -dichlorofluorescin diacetate (DCFH-DA) (Wan et al.,
and damages of stomach and spleen (Wang et al., 2008). Mean- 1993). DCFH-DA passively enters the cell, and is hydrolyzed by esterases to DCFH.
This nonfluorescent molecule is then oxidized to fluorescent compound dichlo-
while, ZnO nanoparticles at high dose still may destroy the immune rofluorescein (DCF) by cellular oxidants. A DCFH-DA stock solution (in methanol)
system. Especially, peritoneal macrophage had a close connection of 10 mM was diluted 1000-fold in the cell culture medium without serum or other
with pancreatitis (Ma et al., 2009). At present, the impact of ZnO additives to yield a 10 ␮M working solution. Cells were washed twice with PBS and
nanoparticles on the immune system is not clear. Thus, in this then incubated with DCFH-DA working solution for 20 min in the dark environ-
ment (37 ◦ C incubator) followed by treatment with ZnO particles for 24 h. Then the
study, we took mouse peritoneal macrophages Ana-1 cells as sub-
cells were washed three times with cell culture medium without serum to elimi-
ject, investigated cytotoxicity of different sized and shaped ZnO nate DCFH-DA that did not enter the cells. Cells were collected in suspension, the
particles, and analyzed the effect of the dissolved Zn2+ and ROS in fluorescence was determined at 488 nm excitation and 525 nm emission using a
cytotoxicity. fluorospectrophotometer.

2. Materials and methods 2.6. ROS activity determination of ZnO particles under the abiotic condition

2.1. ZnO particles ZnO particles suspensions (2.5, 5, 10, 20, 40, 80, 100 ␮g/ml) were prepared with
the complete cell medium, and a part of suspensions was incubated for 24 h at
Four different sized ZnO particles were supplied from commercial companies. 37 ◦ C. ROS activities of ZnO particles suspensions and supernatants after incuba-
Both fine-ZnO particles (<1 ␮m) were purchased from Hangzhou Wanjingxin Mate- tion for 24 h were determined by HRP–DCF fluorometric method (Venkatachari et
rial Co. Ltd., China. 100 ± 10 nm and 30 ± 10 nm ZnO particles were purchased from al., 2007; Jiang et al., 2008), using the complete cell medium as control. The mea-
Beijing Nachen Technology Co. Ltd., China. 10–30 nm ZnO particles were purchased sured fluorescent intensity of controls was subtracted from those of the ZnO particles
from Shenzhen Nanuo Nanomaterials Corp., China. Visualize particles size and shape suspensions or supernatants samples. Briefly, activation of DCFH-DA to DCFH was
of ZnO particles was measured by transmission electron microscopy (TEM, H7650). performed immediately prior to analysis with 1 vol of 5 mM DCFH-DA and 9 vol of
W. Song et al. / Toxicology Letters 199 (2010) 389–397 391

0.01 N sodium hydroxide incubation at room temperature in the dark for 30 min,
which yielded an unstable DCFH solution. Further dilutions were prepared in the
25 mM phosphate buffer (pH 7.2). The solution was kept on ice without exposing it
to light until use. To catalyze the reaction between DCFH and the ROS, Horseradish
peroxidase (HRP) was added to the DCFH solution in a ratio such that the working
reagent contained 2.2 units of HRP/ml of the reagent. In order to correlate the fluo-
rescent intensities and concentrations in terms of equivalent H2 O2 concentrations,
an assay of standard solutions of H2 O2 (2.5, 5, 10, 20, 30, 40 ␮M/L) was performed.
0.1 ml of H2 O2 solutions or the samples were added to 3 ml of the DCFH–HRP reagent
mixture. Blank was prepared with 0.1 ml distilled and deionized water instead of
H2 O2 . And then solutions and blank were incubated at 37 ◦ C for 15 min. The flu-
orescent intensity was measured at 488 nm excitation and 525 nm emission by a
fluorospectrophotometer.

2.7. Dissolved zinc concentrations in the cell culture medium

ZnO particle suspensions (2.5, 5, 10, 20, 40, 80, 100 ␮g/ml) were incubated for
24 h at 37 ◦ C in a humidified 5% CO2 environment. Then centrifugation at 20,000 rpm
for 30 min removed solid phase ZnO from the complete cell medium. 1.5 ml of the
supernatant was added to 1 ml of ultrahigh purity HNO3 , and digested at about
120 ◦ C until the solutions were colorless and clear. At last, the remaining solutions Fig. 1. X-ray diffraction pattern of different sized ZnO particles.
were diluted to 10 ml with 3% nitric acid. The Zn content was analyzed by ICP-AES
(Thermo IRIS Intrepid IIXSP, America).

3. Results
2.8. Statistical analysis
3.1. Characterization of ZnO particles
The data were expressed as mean ± standard deviation. For statistical analysis,
the experimental values were compared to their corresponding control values. A
statistical analysis was done by SPSS 16.0. The significant difference was judged at The X-ray diffraction analysis (Fig. 1) clearly showed that all four
p < 0.05. types of ZnO particles were the hexagonal structure.

Fig. 2. TEM image of ZnO particles.


392 W. Song et al. / Toxicology Letters 199 (2010) 389–397

Fig. 3. Cell viability of Ana-1 cell exposed to different sized ZnO particles. Control cells treated without ZnO particles are considered to have 100% activity. *p < 0.05 vs. control
cells (n = 5).

Table 1
The 24 h IC50 values and 95% confidence limits for Ana-1 cells exposed to different sized ZnO and ZnCl2 . Data were calibrated to Zn concentrations and ZnO concentrations.

ZnO ZnCl2

fine-ZnO 100 nm 30 nm 10–30 nm

Zn concentration(␮g/ml) 35.76 35.26 32.43 24.84 13.33


95% confidence limits 32.32–39.39 31.10–40.05 28.96–37.23 21.97–28.17 11.81–14.99
ZnO concentration(␮g/ml) 44.56 43.94 40.41 30.95
95% confidence limits 40.28–49.09 38.76–49.91 36.09–46.39 27.38–35.10

The TEM micrographs demonstrated the particle shapes and of ZnCl2 to Ana-1. Cell viability obviously declined at the dose above
sizes (Fig. 2). Fine-ZnO, 100 nm ZnO and 30 nm ZnO were rod- 5 ␮g/ml. The IC50 value of ZnCl2 was 13.33 ␮g Zn/ml.
shaped, with lengths of 341.75 ± 173.34 nm, 107.59 ± 38.44 nm Cell membrane damage is reflected in the elevated LDH level in
and 70.89 ± 34.18 nm, diameters of 173.48 ± 72.73 nm, the cell culture medium. Fig. 4 shows LDH level in the cell culture
60.21 ± 21.76 nm and 40.44 ± 12.51 nm, respectively. 10–30 nm medium following 24 h exposure to ZnO particles. LDH levels fol-
ZnO was spherical with diameters of 18.76 ± 4.98 nm. Among lowing exposure to 10–30 nm ZnO particles above 10 ␮g/ml were
them, the mean size of 30 nm ZnO was very different from the increased significantly (p < 0.05). At the dose above 20 ␮g/ml, other
manufacturer’s information. three sized ZnO particles induced the steep elevation of LDH levels
(p < 0.05). However, ZnCl2 caused an obvious elevation of LDH at
5 ␮g/ml (p < 0.05).
3.2. Cytotoxicity and disruption of cell membrane induced by ZnO
particles
3.3. Cytotoxicity and disruption of cell membrane induced by
Exposure of Ana-1 cells to ZnO particles was characterized by a supernatants of ZnO particles suspensions
steep response pattern. ZnO particles initiated cytotoxicity around
5–10 ␮g/ml with a steep decline in cell viability between 10 and As shown in Figs. 5 and 6, at the dose above 10 ␮g/ml, the super-
40 ␮g/ml (Fig. 3). 10–30 nm ZnO particles at the dosage of 5 ␮g/ml natant of 10–30 nm ZnO particles suspension had a significant toxic
decreased significantly the cell viability (p < 0.05), whereas other effect on Ana-1 cell, cell viability obviously decreased (p < 0.05),
three ZnO particles decreased significantly the cell viability at the and the level of LDH evidently increased (p < 0.05). When the con-
dosage of 10 ␮g/ml. The IC50 values of four ZnO particles were 40.28, centration of ZnO particles exceeded 20 ␮g/ml, the supernatants
43.94, 40.41 and 30.95 ␮g ZnO/ml for fine-ZnO, 100 nm, 30 nm and of fine-ZnO, 100 nm and 30 nm ZnO particles suspensions reduced
10–30 nm ZnO particles, respectively (Table 1). In order to compare cell viability markedly (p < 0.05), and definitely elevated the level of
the toxicity of ZnO particles and ZnCl2 , we studied the cytotoxicity LDH. However, when the concentrations of ZnO particles exceeded

Fig. 4. LDH level in the Ana-1 cell culture medium following a 24 h exposure to ZnO particles and ZnCl2 (n = 5).
W. Song et al. / Toxicology Letters 199 (2010) 389–397 393

Fig. 5. Cytotoxicity of supernatant of ZnO particles suspensions incubated for 24 h.


*p < 0.05 vs control cells (n = 5).
Fig. 7. Zn contention of ZnO particles dissolution in the cell culture media (n = 3).

Table 2
The correlation analysis of the dissolved Zn concentration with cell viability and
LDH level induced by the supernatant (R2 ).

The dissolved Zn concentration

Fine-ZnO 100 nm 30 nm 10–30 nm


** ** **
Cell viability 0.990 0.982 0.987 0.977**
LDH level 0.991** 0.970** 0.991** 0.951**
**
p < 0.01.

for fine-ZnO, 100 nm, 30 nm and 10–30 nm, respectively, when the
dissolved Zn2+ content tended to stable. The equilibrium concentra-
tion of zinc ion was about 10 ␮g/ml. When the concentration of ZnO
Fig. 6. LDH level in the Ana-1 cell culture medium following a 24 h exposure to particles was 100 ␮g/ml, the dissolved Zn concentrations of fine-
supernatant of ZnO particles suspensions incubated for 24 h (n = 5).
ZnO, 100 nm, 30 nm and 10–30 nm ZnO particles are 9.89 ± 0.02,
9.71 ± 0.10, 9.58 ± 0.32 and 10.25 ± 0.33 ␮g/ml, respectively.
40 ␮g/ml, the supernatants of four sized ZnO particle suspensions The correlation analysis revealed there was an inverse correla-
decreased cell viability to about 50%, and cell viability had little tion between the dissolved Zn concentration of each ZnO particles
change with the increase of ZnO particles concentrations. and cell viability induced by the supernatant, and a positive correla-
tion between the dissolved Zn concentration of each ZnO particles
3.4. Dissolution of ZnO particles in cell culture medium and LDH level induced by the supernatant (Fig. 8 and Table 2).

The dissolution profiles from different sized ZnO particles in 3.5. Intracellular reactive oxygen species
cell culture medium are shown in Fig. 7. The results suggested
that the dissolution of ZnO nanoparticles was similar to fine-ZnO Intracellular ROS levels were significantly increased after expo-
particles under experiment conditions. When the concentration of sure to all sized ZnO particles at the dose above 5 ␮g/ml (p < 0.05)
ZnO particles exceeded 40 ␮g/ml, the dissolved Zn concentration (Fig. 9). But the effects of different sized ZnO particles were dif-
tended to equilibrium. The concentration of zinc ion was almost ferent. 10–30 nm ZnO particles appeared more effective, when
constant with the increases of the concentrations of ZnO parti- the dosage exceeded 10 ␮g/ml. Intracellular ROS levels induced
cles. Through calculation, we estimated that the concentrations by 10–30 nm ZnO particles were much higher than other sized
ZnO nanoparticles were 41.33, 41.33, 44.92 and 38.59 ␮g/ml and ZnO particles (p < 0.05). 30 nm, 100 nm and fine-ZnO particles
their rates of solution were 29.61%, 29.17%, 26.05% and 32.44% showed the similar characteristics. Compared to ZnO particles,

Fig. 8. The inverse correlation (R2 = 0.990) between the dissolved Zn concentration of fine-ZnO particles and cell viability induced by the supernatant (A), the positive
correlation (R2 = 0.991) between the dissolved Zn concentration of fine-ZnO particles and LDH level induced by the supernatant (B). The figures of other ZnO particles were
not showed, R2 values are shown in Table 2.
394 W. Song et al. / Toxicology Letters 199 (2010) 389–397

Fig. 9. Oxidative stress induced by exposure of Ana-1 cells to ZnO particles and ZnCl2 . *p < 0.05 vs. control cells; ␣ p < 0.05 vs. other sized ZnO particles and ZnCl2 ; ␤ p < 0.05
vs. ZnO particles.

Fig. 10. The inverse correlation (R2 = 0.966) between intracellular ROS levels of fine-ZnO particles and cell viability induced by the supernatant (A), the positive correlation
(R2 = 0.981) between intracellular ROS levels of fine-ZnO particles and LDH level induced by the supernatant (B). The figures of other ZnO particles and ZnCl2 are not showed,
R2 values are shown in Table 3.

ZnCl2 induced relatively less ROS generation at the dosage above and 30 nm. After incubation at 37 ◦ C for 24 h, ROS activities of the
20 ␮g/ml (p < 0.05). supernatants were distinct lower than ZnO particles suspensions
The correlation analysis reveals that there was an inverse cor- (Fig. 12). Except that ROS activities in the supernatants of 10–30 nm
relation between intracellular ROS levels and cell viability induced ZnO particles were obviously higher than other three ZnO particles,
by ZnO particles suspensions, and a positive correlation between there was no difference between 100 nm, 30 nm and fine-ZnO par-
intracellular ROS levels and LDH level induced by ZnO particle sus- ticles. However, ROS activities of all sized ZnO particles under the
pensions (Fig. 10 and Table 3).

3.6. ROS activity determination of ZnO particles under the abiotic


condition

Fig. 11 shows acellular ROS activities of ZnO particles. ROS gen-


erated by all ZnO particles in the cell medium increased with the
increasing of the concentration. The order of ability to generate ROS
of different sized ZnO particles was 10–30 nm, fine-ZnO, 100 nm

Table 3
The correlation analysis of intracellular ROS level with cell viability and LDH level
induced by ZnO particles(R2 ).

intracellular ROS levels

Fine-ZnO 100 nm 30 nm 10–30 nm ZnCl2

Cell viability 0.966** 0.948** 0.929** 0.923** 0.911**


LDH level 0.981** 0.945** 0.922** 0.952** 0.968**
** Fig. 11. Non cellular ROS level of ZnO particles suspensions.
p < 0.01.
W. Song et al. / Toxicology Letters 199 (2010) 389–397 395

than other cells. Compared with similar cells, the toxicity of ZnO
nanoparticles toward normal mouse macrophage Ana-1 was lower
than that to cancerous macrophage RAW 264.7. After exposure
to 25 ␮g/ml ZnO particles (13 nm) for 16 h, all RAW264.7 cells
were almost dead (Xia et al., 2008). This also demonstrated that
the toxicity of ZnO particles to normal cells may be lower than
the corresponding tumor cells. Moreover, LDH leakage reflected
cell membrane damage and the degree of cell necrosis. The steep
elevation of LDH indicated that ZnO particles destroyed cell mem-
brane integrity and could cause cell necrosis. However, at low dose,
5 ␮g/ml of 10–30 nm ZnO and 10 ␮g/ml other three sized ZnO
particles obviously reduced cell viability, but failed to induce the
elevation of LDH level compared with the control, which implied
that low dose of ZnO particles did not result in cell membrane
damage and cell necrosis, but induced apoptosis.

4.2. Impact of ZnO particle size and shape on toxicity

In the present study, the toxicity of three rod-like ZnO particles


Fig. 12. Non cellular ROS level of the supernatants after incubation for 24 h. were very near, this indicated that particle size had no effect on
toxicity of ZnO particles. There was also the similar conclusion in
previous studies. For example, Lin et al. found that the toxicity of
abiotic condition were lower, not exceeding 4.0 ␮M H2 O2 , either in 70 nm ZnO particles was similar to 420 nm ZnO particles (Lin et
the supernatants or in the ZnO particles suspensions. al., 2009). Deng et al. showed 10, 30, 60 and 200 nm ZnO particles
had similar toxicity to mouse NSC (C17.2) (Deng et al., 2009). Yuan
4. Discussion et al. also found 20, 30 and 40 nm ZnO particles had comparative
toxicity in human embryonic lung fibroblasts (HELF) cells (Yuan et
4.1. Cytotoxicity of ZnO particles al., 2010). However, Nair et al. revealed that the toxicity of 40, 150
and 350 nm ZnO nanoparticles had a little difference in osteoblast
Much research has showed that ZnO nanoparticles resulted cancer cells (MG-63), but ZnO microparticles had been less toxic
in cytotoxicity to many types of cell, such as human bronchial than ZnO nanoparticles (Nair et al., 2009). Reddy et al. also showed
epithelial cells (BEAS-2B) (Huang et al., 2010), human lung epithe- that ZnO nanoparticles had a higher toxicity on human T cells than
lial cells (A549) (Lin et al., 2009; Karlsson et al., 2008; Horie et ZnO microparticles (Reddy et al., 2007; Hanley et al., 2008). Thus,
al., 2009), human keratinocyte HaCaT cells (Horie et al., 2009), we inferred there may be a critical size, when ZnO particle size
human epidermal cell (A431) (Sharma et al., 2009), L2 rat lung was below the critical size, particle size had no effect on toxicity;
epithelial cells and rat alveolar macrophages (Sayes et al., 2007, when ZnO particle size exceeded the critical size, the toxicity of ZnO
2009), the mouse macrophage cell RAW 264.7 (Xia et al., 2008), particle was relatively lower. It needs further research to prove this
mesothelioma MSTO-211H and rodent 3T3 fibroblast cells (Brunner inference.
et al., 2006), primary mouse embryo fibroblasts (PMEF) (Yang et al., At present, there are few reports about the impact of ZnO
2009), human embryonic lung fibroblasts (HELF) cells (Yuan et al., particle shape on toxicity. In this study, 10–30 nm ZnO particles
2010), mouse NSC (C17.2) (Deng et al., 2009), human astrocytoma were spherical, 30 nm, 100 nm and fine-ZnO particles were rod-
U87 (astrocytes-like) cells (Lai et al., 2008), human T lympho- shaped. 10–30 nm spherical ZnO particles were slightly highly toxic
cytes (Reddy et al., 2007; Hanley et al., 2008) and human aortic than three rod-like ZnO particles. Thus, particle shape may affect
endothelial cells (HAECs) (Gojova et al., 2007). And ZnO nanoparti- its toxicity. Yang et al. inferred that the genotoxicity of different
cles induced steep dose-cytotoxicity without cell line-dependent. nanoparticles may primarily be due to particle shape (Yang et al.,
According to previous research, ZnO nanoparticles had the highest 2009). However in the study of Deng et al., the toxicity of 10 nm
toxicity in human keratinocyte HaCaT cells and human epidermal spherical ZnO particles was similar to 30 nm, 60 nm and 200 nm
cell (A431), the concentration range of toxic effect was 5–10 ␮g/ml, rod-like ZnO particles (Deng et al., 2009). Therefore, the impact
cell activity was completely inhibited after exposure of 10 ␮g/ml of particle size on the toxicity of ZnO nanoparticles needs further
ZnO nanoparticles(Horie et al., 2009; Sharma et al., 2009). Secondly, study.
ZnO nanoparticles had comparative toxicity in most cells, for exam-
ple, lung cells (A549, BEAS-2B, MSTO-211h, RAW 264.7), fibroblast 4.3. The role of dissolved zinc ion in the toxicity of ZnO
(3T3, PMEF, HELF) and nerve cells (C17.2, U87), their IC50 value was nanoparticles
about 10–20 ␮g/ml. However, the toxicity of ZnO nanoparticles in
human T lymphocytes was much lower than other cells, exposure ZnO is slightly soluble, and can release zinc ions in solution.
of 400 ␮g/ml ZnO nanoparticles did not yet cause cytotoxicity. T Some researchers considered that dissolved zinc ions in the toxi-
lymphocytes, as a kind of immune cells, had great tolerance range city of ZnO nanoparticles played an important role. Brunner et al.
to ZnO nanoparticles. And the toxicity of ZnO nanoparticles toward inferred that toxic effects of ZnO nanoparticles on MSTO or 3T3
normal T cells was lower than that to cancerous T cells (Reddy et cells may be attributed to the dissolution of zinc ions (Brunner
al., 2007; Hanley et al., 2008). et al., 2006). Deng et al. found ZnO nanoparticles and ZnCl2 had
In this study, Ana-1 cell also was a kind of important immune the similar toxic effect on mouse NSCs (Deng et al., 2009). How-
cell. IC50 values of four ZnO particles in Ana-1 cells were 30.95, ever they did not determine the content of dissolved zinc ions.
40.41, 43.94 and 44.56 ␮g/ml for 10–30 nm, 30 nm, 100 nm and Xia et al. determined the content of dissolved zinc ions in the
fine-ZnO particles, respectively. The tolerance range of Ana-1 cell cell culture medium, and inferred when concentrations of ZnO
to ZnO nanoparticles was much lower than that of human T cells. nanoparticles were less than ZnO solubility, the cell cultures were
However, in contrast, toxicity of ZnO nanoparticles was lower mainly exposed to aqueous zinc ions (Xia et al., 2008), whereas
396 W. Song et al. / Toxicology Letters 199 (2010) 389–397

the content of zinc ions was not determined under the experimen- fore, the high level of intracellular ROS may be not a major factor
tal conditions, and the contribution of zinc ions in toxicity of ZnO in cytotoxicity of ZnO nanoparticles, but the cytotoxic response. In
nanoparticles was not clear. In this study, we measured the con- this study, intracellular ROS could produce after zinc ions or ZnO
centration of dissolved zinc ions in the cell culture medium under particles entering into the cells. On the one hand, Ana-1 cell was a
the experimental conditions, and investigated the toxic effect of kind of phagocytic cells, the NADPH oxidase was the major source
dissolved zinc ions. Our results showed that the dissolved Zn2+ of ROS production in phagocytic cells (Babior et al., 2002). On the
concentrations tended to equilibrium at 40 ␮g/ml, and the equilib- other hand, the entry of zinc ions and ZnO particles into cells led
rium concentrations of dissolved Zn2+ were about 10 ␮g/ml, which to mitochondrial damage, leading to ROS generation (Wiseman et
was close to the result of the work done by Xia et al. By equili- al., 2007; Xia et al., 2008). It has been demonstrated that nanoma-
brating excess ZnO nanopowder with the cell medium for 4 days terials of various sizes and chemical compositions preferentially
at 22 ◦ C, Xia et al. estimated that the maximum total dissolved mobilized to mitochondria and destructed mitochondrial function
zinc concentrations could be 190 ␮M in bronchial epithelial growth (Foley et al., 2002; Li et al., 2003; Savic et al., 2003). And the ZnO par-
medium (corresponding to 12.35 ␮g/ml) and 225 ␮M in complete ticles taken up by cells may cause rapid dissolution, releasing Zn2+
DMEM (corresponding to 14.63 ␮g/ml) (Xia et al., 2008). Never- ions (Xia et al., 2008), elevated intracellular Zn2+ induced the dis-
theless, Horie et al. found 1 mg/ml ultrafine ZnO particles released ruption of mitochondrial function (Bishop et al., 2007; Wiseman et
about 60 ␮g/ml Zn2+ into the DMEM-FBS (Horie et al., 2009). More- al., 2007) and triggered ROS generation (Xia et al., 2008). Moreover,
over, ZnO nanoparticles in simulated uterine fluid containing HAS metabolisation of particles by P450 cytochrome can also be a source
released 7.6 ␮g/ml zinc ions (Yang and Xie, 2006). Rapid dissolu- of ROS, as demonstrated in the case of other particles (Takano et al.,
tion of the bulk ZnO and nanoparticulate ZnO in Milli-Q water (pH 2002; Lanone and Boczkowski, 2006).
7.6) was also observed, a dialyzed zinc concentration was approx-
imately 8 mg/l by 6 h, an equilibrium concentration was 16 mg/l 5. Conclusions
after 72 h (Franklin et al., 2007). And the solubility of ZnO nanopar-
ticles in seawater was 3.7 ␮g/ml (Wong et al., 2010). Thus, the In summary, different sized ZnO particles induced slightly dif-
solubility of ZnO nanoparticles in different medium was differ- ferent cytotoxicity. The particles shape may impact on cytotoxicity
ent. However, zinc ions in different media all had an important of ZnO particles. And the dissolved Zn2+ equilibrium concentra-
effect on the toxicity of ZnO nanoparticles. Toxicity test results of tions of ZnO particles in the complete cell medium were about
supernatants showed that dissolved zinc ions severely inhibited 10 ␮g/ml, meanwhile, the supernatants induce 50% cell death. Thus,
cell viability. When the concentration of zinc ions reached equilib- the dissolved Zn2+ played a main role in toxic effect of ZnO parti-
rium (10 ␮g/ml), cell viability reduced by 50%, the toxic effect was cles. Furthermore, ROS generation of ZnO particles under abiotic
close to ZnCl2 (IC50 = 13.33 ␮g/ml). Furthermore, dissolved zinc ions conditions is not enough to cause cytotoxicity. Although intracel-
still induced LDH leakage. And correlation analysis showed that the lular ROS has the significant correlation with cell viability and LDH
zinc ion content had a significant correlation with cell viability and induced by ZnO particles, ROS was not necessarily the main factor
LDH level. Thus, zinc ions played a main role in the toxicity of ZnO of cytotoxicity of ZnO particles.
particles.
Comparing the toxicity of supernatants with ZnO particle sus- Conflict of interest statement
pensions, we found that the toxic effect of supernatants was similar
to ZnO particles suspensions at dosage under 10 ␮g/ml, which indi- None.
cated that the toxicity was mainly due to dissolved zinc ions at
the low dosage. And when the concentration of ZnO particles was
Acknowledgement
10–40 ␮g/ml, the sharp decline of cell viability may be related to
rapid increase of dissolved Zn2+ content. Bozym et al. found free
The authors are grateful to the financial support from
zinc ions outside a narrow concentration range were toxic to a vari-
the National Natural Science Foundation of China (30771771,
ety of cells in vitro (Bozym et al., 2010). When the concentration of
30800934 and 30901221) and the Natural Science Foudation of
ZnO particles exceeded 40 ␮g/ml, the effect of zinc ions on cytotox-
Tianjin (10JCZDJC17100).
icity was almost the same, the enhancement of cytotoxicity with
the increase of ZnO particles concentration was mainly caused by
ZnO particles. References

Babior, B., Lambeth, J., Nauseef, W., 2002. The neutrophil NADPH oxidase. Arch.
4.4. The role of ROS in the toxicity of ZnO nanoparticles Biochem. Biophys. 397, 342–344.
Bishop, G.M., Dringen, R., Robinson, S.R., 2007. Zinc stimulates the production of toxic
Evidently, ROS plays an important role in toxicity of nanopar- reactive oxygen species (ROS) and inhibits glutathione reductase in astrocytes.
Free Radic. Biol. Med. 42, 1222–1230.
ticles (Nel et al., 2006). In order to study the effect of ROS in Bozym, R.A., Chimienti, F., Giblin, L.J., Gross, G.W., Korichneva, I., Li, Y., Libert, S.,
cytotoxicity of ZnO nanoparticles, we measured ROS generation Maret, W., Parviz, M., Frederickson, C.J., Thompson, R.B., 2010. Free zinc ions
under abiotic conditions and intracellular ROS after exposure to outside a narrow concentration range are toxic to a variety of cells in vitro. Exp.
Biol. Med. 235, 741–750.
ZnO particles. The determination results of intracellular ROS levels Brunner, T.J., Wick, P., Manser, P., Spohn, P., Grass, P.N., Limbach, L., Bruinink, A., Stark,
showed that intracellular ROS levels were significantly increased W.J., 2006. In vitro cytotoxicity of oxide nanoparticles: comparison to asbestos,
after exposure to all ZnO particles at the dose above 5 ␮g/ml. And silica, and the effect of particle solubility. Environ. Sci. Technol. 40, 4374–4381.
Deng, X., Luan, Q., Chen, W., Wang, Y., Wu, M., Zhang, H., Jiao, Z., 2009. Nanosized zinc
there was an inverse correlation between intracellular ROS levels
oxide particles induce neural stem cell apoptosis. Nanotechnology 20, 115101
and cell viability, and a positive correlation between intracellular (7 pp.).
ROS levels and LDH level after a 24 h exposure to ZnO particles and Ding, X.B., Wen, L.X., Niu, T.L., Wang, G.Q., Long, X.M., 2007. The impact of nano-ZnO
on mice immune function. Feed Res. 9, 1–4.
ZnCl2 , which is similar to the result of the work done by Deng et al.
Foley, S., Crowley, C., Smaihi, M., Bonfils, C., Erlanger, F.B., Seta, P., Larroque, C., 2002.
(2009). However, test results of ROS generation under abiotic con- Cellular localization of a water-soluble fullerene derivatives. Biochem. Biophys.
ditions showed that although ROS generations evidently enhanced Res. Commun. 294, 116–119.
with the increase of ZnO particles concentrate, they were lower Franklin, N.M., Rogers, N.J., Apte, S.C., Batley, G.E., Gadd, G.E., Casey, P.S., 2007. Com-
parative toxicity of nanoparticulate ZnO, bulk ZnO, and ZnCl2 to a freshwater
in both ZnO particles suspensions and supernatants, which could microalga (Pseudokirchneriella subcapitata): the importance of particle solubil-
not be enough to cause cytotoxicity (Wiseman et al., 2007). There- ity. Environ. Sci. Technol. 41, 8484–8490.
W. Song et al. / Toxicology Letters 199 (2010) 389–397 397

Gojova, A., Guo, B., Kota, R.S., Rutledge, J.C., Kennedy, I.M., Barakat, A., 2007. Induction Reddy, K.M., Feris, K., Bell, J., Wingett, D.G., Hanley, C., Punnoose, A., 2007. Selective
of inflammation in vascular endothelial cells by metal oxide nanoparticles: effect toxicity of zinc oxide nanoparticles to prokaryotic and eukaryotic systems. Appl.
of particle composition. Environ. Health Perspect. 115, 403–409. Phys. Lett. 90, 213902.
Hanley, C., Layne, J., Punnoose, A., Reddy, K.M., Coombs, I., Coombs, A., Feris, K., Savic, R., Luo, L., Eisenberg, A., Maysinger, D., 2003. Micellar nanocontainers dis-
Wingett, D., 2008. Preferential killing of cancer cells and activated human t cells tribute to defined cytoplasmic organelles. Science 300, 615–618.
using ZnO nanoparticles. Nanotechnology 19, 295203. Sayes, C.M., Reed, K.L., Subramoney, S., Abrams, L., Warheit, D.B., 2009. Can in vitro
Horie, M., Nishio, K., Fujita, K., Endoh, S., Miyauchi, A., Saito, Y., Iwahashi, H., assays substitute for in vivo studies in assessing the pulmonary hazards of fine
Yamamoto, K., Murayama, H., Nakano, H., Nanashima, N., Niki, E., Yoshida, Y., and nanoscale materials? J. Nanopart Res. 11, 421–431.
2009. Protein adsorption of ultrafine metal oxide and its influence on cytotoxi- Sayes, C.M., Reed, K.L., Warheit, D.B., 2007. Assessing toxicity of fine and nanopar-
city toward cultured cells. Chem. Res. Toxicol. 22, 543–553. ticles: comparing in vitro measurements to in vivo pulmonary toxicity profiles.
Huang, C.-C., Aronstam, R.S., Chen, D.-R., Huang, Y.-W., 2010. Oxidative stress, cal- Toxicol. Sci. 97 (1), 163–180.
cium homeostasis, and altered gene expression in human lung epithelial cells Sharma, V., Shukla, R.V., Saxena, N., Parmar, D., Das, M., Dhawan, A., 2009. DNA dam-
exposed to ZnO nanoparticles. Toxicol. in Vitro 24, 45–55. aging potential of zinc oxide nanoparticles in human epidermal cells. Toxicol.
Jeng, H.A., Swanson, J., 2006. Toxicity of metal oxide nanoparticles in mammalian Lett. 185, 211–218.
cells. J. Environ. Sci. Health A: Toxic Hazard. Subst. Environ. Eng. 41 (12), Takano, H., Yanagisawa, R., Ichinose, T., Sadakane, K., Inoue, K., Yoshida, S., Takeda,
2699–2711. K., Yoshino, S., Yoshikawa, T., Morita, M., 2002. Lung expression of cytochrome
Ji, S.L., Ye, C.H., 2008. Synthesis, growth mechanism, and applications of zinc oxide P450 1A1 as a possible biomarker of exposure to diesel exhaust particles. Arch.
nanomaterials. J. Mater. Sci. Technol. 24, 457–472. Toxicol. 76, 146–151.
Jiang, J., Oberdörster, G., Elder, A., Gelein, R., Mercer, P., Biswas, P., 2008. Does Venkatachari, P., Hopke, P.K., Brune, W.H., Ren, X., Lesher, R., Mao, J., Mitchell, M.,
nanoparticle activity depend upon size and crystal phase? Nanotoxicology 2 2007. Characterization of wintertime reactive oxygen species concentrations in
(1), 33–42. flushing, New York. Aerosol Science and Technology 41, 97–111.
Karlsson, H.L., Cronholm, P., Gustafsson, J., Möller, L., 2008. Copper oxide nanopar- Wan, C.P., Myung, E., Lau, B.H., 1993. An automated micro-fluorometric assay for
ticles are highly toxic: a comparison between metal oxide nanoparticles and monitoring oxidative burst activity of phagocytes. J. Immunol. Methods 159,
carbon nanotubes. Chem. Res. Toxicol. 21, 1726–1732. 131–138.
Lai, J.C.K., Lai, M.B., Jandhyam, S., Dukhande, V.V., Bhushan, A., Daniels, C.K., Leung, Wang, B., Feng, W.Y., Wang, M., Wang, T., Gu, Y., Zhu, M., Ouyang, H., Shi, J., Zhang, F.,
S.W., 2008. Exposure to titanium dioxide and other metallic oxide nanoparticles Zhao, Y., Chai, Z., Wang, H., Wang, J., 2008. Acute toxicological impact of nano-
induces cytotoxicity on human neural cells and fibroblasts. Int. J. Nanomed. 3 and submicro-scaled zinc oxide powder on healthy adult mice. J. Nanopart. Res.
(4), 533–545. 10 (2), 263–276.
Lanone, S., Boczkowski, J., 2006. Biomedical applications and potential health risks Wiseman, D.A., Wells, S.M., Hubbard, M., Welker, J.E., Black, S.M., 2007. Alterations
of nanomaterials: molecular mechanisms. Curr. Mol. Med. 6, 651–663. in zinc homeostasis underlie endothelial cell death induced by oxidative stress
Lee, J., Kang, B.S., Hicks, B., Chancellor Jr., T.F., Chu, B.H., Wang, H.T., Keselowsky, B.G., from acute exposure to hydrogen peroxide. Am. J. Physiol. Lung Cell. Mol. Physiol.
Ren, F., Lele, T.P., 2008. The control of cell adhesion and viability by zinc oxide 292, L165–L177.
nanorods. Biomaterials 29, 3743–3749. Wong, S.W.Y., Leung, P.T.Y., Djurišić, A.B., Leung, K.M.Y., 2010. Toxicities of nano zinc
Li, N., Sioutas, C., Cho, A., Schmitz, D., Misra, C., Sempf, J., Wang, M., Oberley, T., oxide to five marine organisms: influences of aggregate size and ion solubility.
Froines, J., Nel, A., 2003. Ultrafine particulate pollutants induce oxidative stress Anal. Bioanal. Chem. 396, 609–618.
and mitochondrial damage. Environ. Health Perspect. 111, 455–460. Xia, T., Kovochich, M., Liong, M., Madler, L., Gilbert, B., Shi, H., Yeh, J.I., Zink, J.I., Nel,
Lin, W., Xu, Y., Huang, C.-C., Ma, Y., Shannon, K.B., Chen, D.-R., Huang, Y.-W., 2009. A.E., 2008. Comparison of the mechanism of toxicity of zinc oxide and cerium
Toxicity of nano- and micro-sized ZnO particles in human lung epithelial cells. oxide nanoparticles based on dissolution and oxidative stress properties. ACS
J. Nanopart. Res. 11, 25–39. Nano 2 (10), 2121–2134.
Ma, Z., Ma, Q., Sha, H., Wang, L., Zhang, M., 2009. Relationship between peritoneal Yang, H., Liu, C., Yang, D., Zhang, H., Xi, Z., 2009. Comparative study of cytotoxicity,
macrophages and inflammatory reaction in a rat model of severe acute pancre- oxidative stress and genotoxicity induced by four typical nanomaterials: the role
atitis. Acad. J. Xi’an Jiaotong Univ. 21, 238–241. of particle size, shape and composition. J. Appl. Toxicol. 29, 69–78.
Nair, S., Sasidharan, A., Divya Rani, V.V., Menon, D., Nair, S., Manzoor, K., Raina, Yang, Z., Xie, C., 2006. Zn2+ release from zinc and zinc oxide particles in simulated
S., 2009. Role of size scale of ZnO nanoparticles and microparticles on toxic- uterine solution. Colloids Surf. B 47, 140–145.
ity toward bacteria and osteoblast cancer cells. J. Mater. Sci.: Mater. Med. 20, Yuan, J.H., Chen, Y., Zha, H.X., Song, L.J., Li, C.Y., Li, J.Q., Xia, X.H., 2010. Determination,
S235–S241. characterization and cytotoxicity on HELF cells of ZnO nanoparticles. Colloids
Nel, A., Xia, T., Mädler, L., Li, N., 2006. Toxic potential of materials at the nanolevel. Surf. B: Biointerfaces 76, 145–150.
Science 311, 622–627.

You might also like