You are on page 1of 12

Nanoscale

View Article Online


PAPER View Journal

Metal homeostasis disruption and mitochondrial


Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

Cite this: DOI: 10.1039/c6nr05306h


dysfunction in hepatocytes exposed to sub-toxic
doses of zinc oxide nanoparticles
M. Chevallet,*a,b,c B. Gallet,d,e,f A. Fuchs,g P. H. Jouneau,h,i K. Um,a,b,c E. Mintza,b,c
and I. Michaud-Soret*a,b,c

Increased production and use of zinc oxide nanoparticles (ZnO-NPs) in consumer products has prompted
the scientic community to investigate their potential toxicity, and understand their impact on the
environment and organisms. Molecular mechanisms involved in ZnO-NP toxicity are still under debate
and focus essentially on high dose expositions. In our study, we chose to evaluate the eect of sub-toxic
doses of ZnO-NPs on human hepatocytes (HepG2) with a focus on metal homeostasis and redox balance
disruptions. We showed massive dissolution of ZnO-NPs outside the cell, transport and accumulation of
zinc ions inside the cell but no evidence of nanoparticle entry, even when analysed by high resolution
TEM microscopy coupled with EDX. Gene expression analysis highlighted zinc homeostasis disruptions as
shown by metallothionein 1X and zinc transporter 1 and 2 (ZnT1, ZnT2) over-expression. Major oxidative
stress response genes, such as superoxide dismutase 1, 2 and catalase were not induced. Phase
2 enzymes in term of antioxidant response, such as heme oxygenase 1 (HMOX1) and the regulating
subunit of the glutamate-cysteine ligase (GCLM) were slightly upregulated, but these observations may be
Received 4th July 2016, linked solely to metal homeostasis disruptions, as these actors are involved in both metal and ROS
Accepted 19th October 2016
responses. Finally, we observed abnormal mitochondria morphologies and autophagy vesicles in response
DOI: 10.1039/c6nr05306h to ZnO-NPs, indicating a potential role of mitochondria in storing and protecting cells from zinc excess
www.rsc.org/nanoscale but ultimately causing cell death at higher doses.

1 Introduction and proliferation, dierentiation or programmed cell death.2


Maintenance of zinc homeostasis is essential and deficiency
Zinc is an essential trace element. Not only does it act as a cel- as well as excess zinc is toxic for organisms and cells. For
lular ionic signal, but in thousands of proteins, zinc partici- example, deregulation of free zinc has been implicated in the
pates in enzymatic catalysis, structural organization, and/or formation of -amyloid plaques associated with Alzheimers
regulation of function. Zn2+ is a non-redox metal ion but it is disease.3,4
involved in the redox equilibrium of the cell by binding to To complicate things further, zinc at physiological concen-
thiolate containing sites with a very high anity.1 In both its trations has also been shown to be cytoprotective as a
ionic and protein-bound forms zinc is at the crossroads of all pro-antioxidant.4
crucial decisions in the life of mammalian cells: cell growth The large number of proteins that are potentially dedicated
to Zn2+ transport (ZnT and ZIP families) and buering (metal-
lothioneins) reflect the complexity and importance of zinc
a
CNRS, Laboratoire de Chimie et Biologie des Mtaux (LCBM), UMR 5249, homeostasis.5,6 The ZnT transporters are dedicated to Zn2+
Grenoble, France. E-mail: isabelle.michaud-soret@cea.fr eux from the cytoplasm to the extracellular medium or to
b
CEA, BIG, LCBM, Grenoble, France. E-mail: mchevallet@cea.fr
c organelles. Zip transporters control Zn2+ import from the extra-
Universit Grenoble Alpes, LCBM, Grenoble, France
d
Universit Grenoble Alpes, IBS, Grenoble, France cellular medium or organelle lumens into the cytoplasm.7
e
CNRS, IBS, Grenoble, France Although the up-regulation of Zn2+ chelation and transport
f
CEA, IBS, Grenoble, France machineries appears to be directly activated by Zn2+ binding to
g
CEA, BIG, DIR, Grenoble, France the transcription factor MTF-1,8 there is still much to decipher
h
CEA, INAC, Minatec campus, Grenoble, France
i in the mechanisms governing zinc homeostasis.
Universit Grenoble Alpes, INAC-MEM-LEMMA, Grenoble, France
Electronic supplementary information (ESI) available. See DOI: 10.1039/ Due to their specific physical properties linked to surface
c6nr05306h and quantum size eects,9 nanoparticles (NPs) are finding

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

their way into a wide range of consumer products, including and focused on the early response of metal homeostatic
food packaging. This increased exposure of all living beings control and oxidative stress genes to sub-toxic doses of zinc
and their environment to NPs is driving concerns about their (i.e. conditions that do not induce cell mortality).
potential adverse impact on the environment and our well- Our results clearly demonstrate no specific nano eect of
being. Over the last decade, considerable scientific eort has the ZnO-NPs in contrast to what we previously observed with
been invested in research to better understand the risks of CuO-NPs.25 We demonstrate however that sub-toxic doses of
exposure to NPs.10,11 both ionic and nanoparticulate forms of zinc induce a metal
ZnO nanoparticles (ZnO-NPs) are among the top 5 nano- stress response with zinc homeostasis disruption and
particles produced in high tonnage (the estimated global mitochondria alterations, even after short exposure times.
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

annual production in 2010 is over 30 000 tons)12 and are used


in commercial applications such as antibacterial coatings or as
UV absorbers in sunscreens and textiles.13,14
The literature on mammalian toxicity of ZnO-NPs published
2 Experimental
between 2009 and 2011 was recently summarized in a review.15 2.1 Nanoparticles
In vivo studies explored the main exposure means and showed ZnO-NPs with a primary particle size of less than 100 nm were
a potential risk for mammals. Through airway exposure in purchased from Sigma-Aldrich: an uncoated ZnO-NP powder
rats, ZnO-NPs elicited an increased Zn content in multiple (ref number 721077) and an aminopropyl-silane-coated
organs and in particular in the liver where it was sucient to ZnO-NP dispersion (ref number 544906). The uncoated
cause damage as shown by histopathological examination.16 In ZnO-NPs were dispersed at 50 mg mL1 in 50% FBS by soni-
skin applications, while several reviews have concluded that cation for 30 min by 1s on1s o pulses at 60% amplitude
metal oxide NPs do not penetrate the stratum corneum,17 (750 W) in a cup-horn instrument (Vibra cell, BioBlock
Adamcakova-Dodd et al.18 demonstrated that small amounts Scientific, France) under water cooling thermostated at 4 C to
of Zn from ZnO-NPs in sunscreens can pass through the skins get stable and reproducible dispersion of NPs and to avoid
protective layers and be detected in blood and urine. Oral rapid aggregation (denoted ZnO-NP-FBS). The ZnO-NP-silane
administration of ZnO-NPs also resulted in zinc accumulation were diluted directly in water at 50 mg mL1, an additional
in the liver and other organs.19 FBS coating was not necessary as these NPs were already well
In a majority of published studies, incomplete information dispersed and stable before diluting them into FBS-containing
(or characterization) of the NPs themselves and in some cases cell culture media. The actual size of the particles was deter-
the lack of information or studies on ion release and NP mined after dilution in water or in complete culture medium
uptake by cells make it dicult to compare results. by dynamic light scattering, using a Wyatt Dynapro Nanostar
In vitro studies on dierent cell lines showed consensual instrument. A Malvern instrument Zetasizer plus was used to
results: ZnO-NPs are in all cases cytotoxic via the release of determine the zeta potential. The size and shape of NPs were
Zn2+ ions, reactive oxygen species (ROS) production, and mito- verified by STEM on a Hitachi S5500.
chondria dysfunction. Some studies also revealed a genotoxic
eect of ZnO-NPs.20 Others have demonstrated that autophagy 2.2 Cell culture
is activated21,22 as well as inflammation via IL8 production.23
HepG2 cells were obtained from the American Type Culture
In hepatocytes, an alteration of albumin production was high-
collection (ATCC) (Manassas, Virginia, USA). They were grown
lighted23 and an important disruption in energy metabolism
in modified Eagles medium (MEM) supplemented with 10%
with an increase in both gluconeogenesis and glycogenolysis
v/v FBS, 20 mM L-glutamine, 10 mM sodium pyruvate, 100 g
was described.24
mL1 streptomycin and 100 U mL1 penicillin. Cells were
Moreover, most studies focused solely on NP toxicity at high
cultured at 37 C in a humidified atmosphere with 5% CO2.
toxic doses using classical tests to measure cytotoxicity, ROS
production, and DNA damages without addressing the specific
intracellular mechanisms and pathways related to metal 2.3 Cell viability
homeostasis. This fact prompted us to investigate the eects Cells were seeded at 300 000 cells per mL in 12-well plates.
of ZnO-NP exposure on metal homeostasis regulation, taking They were treated with zinc acetate or ZnO-NPs on the follow-
into account its particulate or ionic nature in situ in human ing day and harvested after a further 24 h. Cell viability was
liver-derived cells (HepG2), as the liver is the primary target for evaluated by counting Trypan Blue stained cells in a TC20
concentrating and metabolizing toxic agents. To take into Automated Cell Counter (BioRad). In order to evaluate the pro-
account toxic eects that could be linked to the surface charge tecting eect of EDTA or Ca2+, HepG2 cells were pretreated for
anionic or cationic of the nanoparticles, we chose to work 15 min with 1 mM EDTA or 2 mM CaCl2 before directly adding
with 2 types of ZnO nanoparticles, namely foetal bovine serum Zn acetate, ZnO-NP-FBS or ZnO-NP-silane at the dose of
(FBS)-coated ZnO-NPs with a negative surface charge and 250 M for 24 h. For assays including bafilomycin A (BafA),
silane-coated ZnO-NPs with a positive surface charge. We com- HepG2 cells were pretreated 1 hour with 100 nM BafA before
pared the eects of these two types of characterized ZnO-NPs directly adding Zn acetate, ZnO-NP-FBS or ZnO-NP-silane at
with the eects of the soluble form of zinc, using zinc acetate, 90 or 150 M for 24 h.

Nanoscale This journal is The Royal Society of Chemistry 2016


View Article Online

Nanoscale Paper

2.4 ICP-AES HPRT amplification signals as internal normalizers


Analysis of zinc was performed by inductively coupled plasma (to correct for total RNA content). Results are expressed as the
atomic emission spectroscopy (ICP-AES: ICPE-9000 from relative change in expression compared to the control.
Shimadzu Scientific Instruments). Results were the average (SEM) of at least 3 independents
HepG2 cells were treated with Zn acetate or ZnO-NPs for experiments.
6 h, 24 h and 48 h or kept in complete culture medium as a
control. Briefly, after removing the medium and washing with 2.6 Measuring SOD and catalase activities
PBS buer, cells were collected using trypsin, centrifuged at
SOD and catalase activities were measured as previously
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

1200 rpm for 5 min and suspended in PBS buer. Cells were
described27,28 and results were expressed as the average of
then counted using a TC20 Automated Cell Counter (BioRad)
3 independent experiments in enzymatic units per mg of total
and centrifuged again. The pellets were suspended with
protein quantified by a Bradford assay.29
100 L of pure nitric acid and mineralized overnight at 95 C
in a DigiPrep (SCP Science). The internal standard Ytterbium
was added and the samples were diluted in pure water qs 2.7 Electron microscopy
6.5 mL prior to analysis. A standard curve was performed Monolayers of HepG2 cells were fixed overnight at room temp-
using an atomic absorption standard solution of zinc (Sigma- erature in a 1 : 1 ratio mixture of 4% paraformaldehyde, 0.4%
Aldrich). glutaraldehyde in 0.2 M PHEM (60 mM PIPES, 25 mM HEPES,
To follow zinc dissolution, 90 M zinc acetate and ZnO-NPs 10 mM EGTA, 2 mM MgCl2) pH 7.2 and culture medium,
were incubated in complete culture medium in cell culture washed in 0.1 M PHEM pH 7.2, and fixed for 30 minutes in
plates. The plates were kept for 24 h in a cell culture incubator 2% paraformaldehyde, 0.2% glutaraldehyde in 0.1 M PHEM
at 37 C and 5% CO2. pH 7.2, washed in 0.1 M PHEM pH 7.2 and post-fixed in 1%
The media were retrieved at dierent times, filtered at OsO4, 1.5% potassium ferrocyanide in 0.1 M PHEM buer for
0.1 m (PVDF, Merck Millipore LTD) and centrifuged at 1 h at room temperature. After 3 washes in water, post-staining
150 000g for 45 min to sediment the ZnO-NPs. The concen- was done using 0.5% uranyl acetate in 30% ethanol for 30 min
tration of zinc ions in the supernatant was then measured by at room temperature in the dark. Cells were then dehydrated
ICP-AES. in graded ethanol series, and flat-embedded using the Epoxy
Embedding Medium kit (Sigma-Aldrich). Ultrathin sections
2.5 Real-time PCR assays & measurements (240 nm) were cut on a Leica UC7 ultra-microtome using a
DiATOME 35 diamond knife and collected on formvar carbon
HepG2 cells were harvested and RNA was isolated using
coated 100-mesh copper grids. Sections were stained in 5%
Absolutely-RNA miniprep kits (Agilent #400800). RNA concen-
uranyl acetate in water for 10 min and in 2% lead citrate for
tration was determined using a NanoDrop spectro photometer
5 min. Images were taken on a Tecnai G2 Spirit BioTwin (FEI)
(ND-1000). Reverse transcription was performed with the
at 120 kV using an ORIUS SC1000 CCD camera (Gatan).
Anity script qPCR cDNA synthesis kit (Agilent # 600559),
according to the manufacturers instructions. Gene specific
primers for MET1X were taken from.26 The other human 2.8 Study of acidic vesicular organelles
primers were designed using DNASTAR, Primer-Blast or Pick To detect and quantify acidic vesicular organelles, cells were
Primers. The designed primers are given in Table S1. stained with acridine orange (1 g mL1) for 15 min as
Quantitative PCR was performed with Brilliant II SYBR described previously.30 Green (510530 nm) fluorescence emis-
green qPCR master mix1 (Agilent #600828) using the primers sions from 30 000 cells illuminated with a blue excitation light
at 200 nM. PCR conditions ( primer concentrations, cDNA (488 nm) were measured with a MoFlo instrument (Beckman
quantity) were optimized and PCR eciency was determined Coulter).
for each target gene. PCR reaction mixtures (10 L) were
placed in the Cfx96 instrument (Bio-Rad) where they under-
went the following cycling program, optimized for a 96-well 2.9 Mitochondrial transmembrane potential assessments
block: 95 C for 15 min, immediately followed by 40 cycles of Mitochondrial potential was measured using a Rhodamine
10 s at 95 C and 30 s at 60 C. At the end, PCR products were 123 uptake assay.31 Cells were seeded in 6-well plates for 24 h
dissociated by incubating for 1 min at 95 C and then 30 s at and then treated with zinc for 24 h before adding Rhodamine
55 C, followed by a ramp up to 95 C. PCR quality and speci- 123 at a final concentration of 10 g mL1 for 15 min at 37 C.
ficity were verified by analysing the dissociation curve. For Cells were then harvested and rinsed with PBS and analysed by
each set of primers, a no-template control (NTC) and a no- flow cytometry on a MoFlo instrument (Beckman Coulter).
reverse-amplification control (NAC) were included. qRT-PCR Live cells were first selected on the basis of the size and granu-
reactions were run in triplicate, and quantification was per- larity, and their fluorescence was then measured using exci-
formed using comparative regression (Cq determination tation at 488 nm and emission at 525 nm. Cells pre-treated
mode). Quantitative PCR data were comparatively analysed with sodium azide (NaN3) at 40 mM for 30 min were used as a
using the Cfx software (Bio-Rad Cfx manager) with 36B4 and negative control.

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

3 Results and discussion higher toxicity of ZnO-NP-silane which was not statistically
significant.
3.1 Nanoparticle characterization The shape of the viability curves showed that HepG2 cells
Dynamic light scattering was used to characterize the nano- could tolerate 150 M of zinc for 24 h without loss of viability
particle populations by their hydrodynamic diameter distri- but that beyond this dose, viability decreased drastically.
bution (as % of intensity). ZnO-NP-FBS had an average dia- These dose tolerances are in agreement with the literature
meter of 237 nm (Fig. S1) and a polydispersity index of 20%, for HepG2 cells34 and other cell types such as macrophages.35
ZnO-NP-silane presented an average diameter of 79 nm and a In order to decipher the response of the cells, the
polydispersity index of 19%. The zeta potentials were expression of genes involved in metal homeostasis and oxi-
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

measured in 1 mM KCl 32,33 and were of 25 mV for the dative stress was followed.
ZnO-NP-FBS and 19.3 mV for the ZnO-NP-silane. The albumin
coating accounts for the negative surface charge at physio-
logical pH and the silane cationic coating accounts for the 3.3 Early responses of metal homeostatic control and
positive surface charge of these NPs. STEM images of coated oxidative stress genes
NPs diluted in water confirmed the size of the particles For this gene expression study we chose a sub-toxic dose of
(Fig. 1), the ZnO-NP-FBS were rod-shaped while the ZnO-NP- 90 M oering an optimum compromise between viability and
silane were more spherical. biological eect.
Indeed, at higher but still sub-toxic concentrations
3.2 HepG2 viability studies (150 M), gene expression of usual internal normalizers
ZnO-NP toxicity was first studied by measuring the viability of showed too much variability, a phenomenon which is
HepG2 cells after 24 h incubation with either zinc acetate or explained by the proliferating eect of zinc at this dose.36 In
ZnO-NP-FBS or ZnO-NP-silane in a range of concentrations order to capture early responses triggered by zinc, we also
from 0 to 300 M equivalent zinc (Fig. 2). The toxicity appeared reduced the exposition to 6 hours, a condition where no cell
to be similar among the dierent forms of zinc, with a slightly death was yet measurable, even at concentrations as high as
300 M (data not shown).
The responses of HepG2 cells to a 6 h incubation with
90 M zinc, brought either by zinc acetate, ZnO-NP-FBS or
ZnO-NP-silane were studied by quantitative RT-PCR.
For oxidative stress, we chose the classical genes CAT (cata-
lase), SOD1 (the cytoplasmic Cu, Zn superoxide dismutase),
SOD2 (the mitochondrial Mn superoxide dismutase) and
GCLM, the regulating subunit of the glutamatecysteine ligase
which is the first rate-limiting enzyme of glutathione (GSH)
synthesis. We added HMOX1, the inducible isoform of heme
Fig. 1 Characterization of ZnO-NPs. STEM images of ZnO-NP-FBS oxygenase which can act as an antioxidative protein and
(A) and ZnO-NP-silane (B) taken on a Hitachi S5500.
was already found to be induced by ZnO-NP37,38 and by
CuO-NP.25,39
For zinc homeostasis40 we chose MTF1, Met1X, Znt1 and 7
and Zip1. MTF1 is a cellular zinc sensor, which in the presence
of excess zinc, migrates to the nucleus and activates genes
involved in zinc homeostasis via binding to metal-response
elements. Metallothionein 1X (Met1X) was chosen to represent
the metallothioneins, cysteine-rich proteins which function as
cytoplasmic soft-metal chelators, bind zinc with high
anity,41 and are important in defending the cell against
metal poisoning.42 The transporters ZnT1, ZnT7 and Zip1 are
respectively a plasma membrane zinc exporter, a zinc importer
for storage in the Golgi apparatus and the main plasma mem-
brane zinc importer. Metallothioneins, ZnT1 and ZnT7 gene
expressions are known to be activated via MTF1.43
For iron homeostasis we looked at HAMP, the gene encod-
ing hepcidin, a cysteine-rich protein which controls iron
Fig. 2 Cell viability after a 24 hour incubation with zinc acetate (),
homeostasis44 and is regulated via MTF1.8,45 Finally, we added
ZnO-NP-FBS (), or ZnO-NP-silane (). The arrow shows the concen-
tration (90 M or 7.3 g mL1) chosen for the gene expression studies.
the chaperone HSPA6, from the HSP 70 family which is acti-
For sake of comparison, the ZnO-NP concentrations are expressed as vated by several soft metals such as Cd2+ and Zn2+ and
zinc equivalents. responds to CuO-NPs in HepG2 cells,25 and EGR1, an early

Nanoscale This journal is The Royal Society of Chemistry 2016


View Article Online

Nanoscale Paper

response transcription factor induced by several extracellular heme ring to form biliverdin and then bilirubin, known to
signals, including environmental stress.46 chelate metals.51 Moreover, although zinc in biology is redox
For all treatments, the highest induction was observed for inert, in contrast to copper or iron, zinc excess can indirectly
Met1X, which was up-regulated around 50-fold (Fig. 3). We influence the redox balance.1
also found a 4-fold increase in the expression of ZnT1 but no A minor but significant induction of EGR1 is in agreement
change in ZnT7 nor in Zip1. These results are in agreement with the work of Jeong et al.52 on keratinocytes and probably
with previous work of Cousins et al.47 in THP1 cells treated belongs to a global stress response signature.
with 40 M ZnSO4 showing a 3-fold increase in ZnT1 These results clearly showed an important response to zinc
expression and no significant variation for ZnT7 and Zip1. We stress after either zinc acetate or either ZnO-NP treatments,
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

found no evidence in the literature of regulation of Zip1 via possibly linked with an oxidative stress. ZnO-NP-silane elicited
MTF1. In HEK293 cells, it was shown that Zip1 is not regulated slightly higher cytotoxicity at equivalent zinc concentrations
at the transcriptional level but rather post-translationally by above 150 M and activated HSPA6 to a slightly higher extent
altering the proteins subcellular distribution.48 Finally, there than ZnO-NP-FBS or zinc acetate. However, no other dier-
was no increase in the expression of MTF1 showing that zinc ences were identified among these 3 forms of zinc. Similar
excess induces the translocation of MTF1, and does not results were found on macrophages with a slightly higher cyto-
require de novo synthesis under our conditions. Although zinc toxicity of ZnO-NP-silane versus ZnO-NP-FBS.35 This dierence
homeostasis is linked to other metal homeostasis, such as that could be explained by the higher dissolution rate of ZnO-NP-
of copper and iron, hepcidin also showed no variation in silane, as well as by the cationic coating of these nanoparticles
expression in our conditions. The chaperone HSPA6 was which can influence their interactions with the cell membrane,
slightly induced, in particular with ZnO-NP-silane. Zn2+ ion release and ion entry. Studies with polystyrene nano-
On the oxidative stress side, we found no modification of particles have also shown that amino cationic coatings are
the major enzymes dedicated to detoxification of ROS. CAT, directly responsible for toxic eects compared to anionic coat-
SOD1 and SOD2 are expressed constitutively and we can postu- ings, as NPs carrying amino groups have been shown to target
late that cells are able to cope with a moderate oxidative cell membranes through strong binding to phospholipid com-
stress without requiring additional induction of the cor- ponents.53,54 In any case, the similarity of response between
responding genes. However we found 3.54.5 fold inductions zinc ions and NPs strongly suggests that NP toxicity is essen-
for HMOX1 and GCLM which are classified as phase 2 tially due to ion release.
enzymes in term of antioxidant response, and are regulated by
the transcription factor Nrf2 via ARE, the antioxidant response 3.4 Investigation of the oxidative stress response
element.49 Interestingly, GCLM was also described to be regu- As changes observed at the transcription level are not always
lated by MTF1.50 In light of these results, it remains unclear if correlated with protein activities, we measured the activity of
90 M zinc induces an oxidative stress or solely a metal CAT and SOD enzymes after a 6 h incubation with zinc at
homeostasis disruption. As such, metal and oxidative stress 90 M (corresponding to 7.3 g mL1) and found no signifi-
responses are tightly imbricated. On the one hand, GSH which cant eect whatever the form of zinc (Fig. S2). We even
is present in millimolar concentrations in cells, binds large noticed a tendency towards decreased (P = 0.067) catalase
amounts of zinc in vitro and on the other hand metallo- activities after zinc treatments that was the opposite result to
thioneins can function as antioxidants.1,5 HMOX1 cleaves the what could have been expected. These results confirm the very

Fig. 3 Quantitative PCR analysis of mRNA expression in HepG2 cells after a 6 hour incubation with Zn acetate, ZnO-NP-FBS or ZnO-NP-silane at
90 M (note the 10-fold change in scale of the right panel). Results are presented as relative expression changes after normalizing with HPRT and
36B4 mRNA. Each value represents the mean of relative expression +SEM from three independent experiments. *: p < 0.05, **: p < 0.01, ***: p <
0.001 vs. control.

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

moderate oxidative stress response generated by sub-toxic 3.6 Protection by EDTA or Ca2+ from Zn toxicity
doses of zinc seen at the transcription level. These findings are To evaluate the role of solubilised Zn2+ ion in ZnO-NP
also in agreement with the literature. Indeed, Triboulet et al.39 mediated cell death, we tested the eect of an ion chelator. We
showed, by a proteomic approach, a rather weak activation of found that pre-treatment of HepG2 cells with 1 mM EDTA,
the oxidative stress response pathway in macrophages treated known to chelate divalent metal ions, completely preserved
with ZnO-NP at 8 g mL1, the concentration corresponding to cell viability (Fig. 5) at the toxic dose of 250 M ZnO-NPs in
20% mortality (LD20). Very interestingly, they found no changes agreement with previous work.58 This result, in the light of our
in CAT and SOD protein levels. Guan et al.55 showed that SOD previous findings, is in favour of a toxicity essentially mediated
activity was significantly reduced (p < 0.05) at high concen-
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

by rapid ion release in the extracellular medium. Indeed, here


trations (above 50 g mL1) of ZnO-NPs on LO2, another liver again, similar results were found using zinc acetate. More sur-
cell line. Furthermore, Sharma et al.34 estimated the level of prisingly, we also found that pre-treatment with 1.25 mM
intracellular ROS (by the method of Wan et al.) on HepG2 cells CaCl2 showed similar protective eects.
exposed to ZnO-NPs for 6 h and found no measurable changes A possible explanation for this eect is brought by
at 8 g mL1 (close to our concentration) but a significant Johnson et al.59 who showed that dissolved Ca2+ prevents cells
increase (p < 0.05) in ROS generation at 14 and 20 g mL1. from taking up toxic amounts of zinc, suggesting that zinc
entry could be mediated by a calcium transporter. This would
3.5 Nanoparticle fate in the culture medium be a non-specific mechanism having no relation to Zip1
regulation.
As zinc ions and ZnO-NPs seemed to induce similar eects on
cells, we investigated the fate of NPs in biological media
(Fig. 4), combining filtration on 0.1 m membranes and ultra- 3.7 Cellular zinc content and subcellular localization
centrifugation to ensure the total removal of NPs. The ICP-AES Total zinc content in cells was analysed after various incu-
dosage of supernatants clearly showed a biphasic and major bation times with zinc acetate or either NPs (Fig. 6). Zinc
dissolution of NPs in complete MEM reaching 7090% of dis- increased steadily in cells up to 24 h and then remained rela-
solution after 6 h. This result is in agreement with Herrmann tively stable until 48 h, even though zinc concentrations in the
et al.s work,56 which shows a rapid dissolution of ZnO-NPs in medium remained unchanged and the amount of zinc that
a phosphate buer. Size is considered as the primary physico- entered the cells was negligible compared to the amount in
chemical property aecting the solubility of NPs and theoreti- the medium. Protective mechanisms appeared to be activated
cally, solubility increases as the particle size decreases.57 This to limit zinc level increases as already suggested by our gene
property, among many other parameters, could contribute to expression results. After 24 h with zinc followed by 24 h
the higher initial level of ZnO-NP-silane dissolution (60 vs. without zinc, we measured a dilution of zinc in cells consistent
40% for ZnO-NP-FBS), which is then followed by a phase of with cell proliferation and distribution to daughter cells. There
slower dissolution with similar rate constants for both NPs. was no evidence of active zinc export back into the medium.
The dierences in the plateaus need to be interpreted with Here again, we saw similar results with either NPs or zinc
caution as the same experiment done with zinc acetate did not acetate, an observation which corroborates a toxicity solely
reach 100%, suggesting some loss of zinc on the plastic tube linked to metal ion release.
walls and on the filter. These measurements corroborate the
extent of zinc ion release in the medium. Special care needs to
be taken to interpret DLS studies of NPs diluted in complex
culture medium as phenomena of aggregation and dissolution
may bias the correlation analysis. Critical analysis of our
results showed that both dissolution and aggregation or
precipitation occurred in MEM.

Fig. 5 Protection by EDTA or Ca2+ against Zn toxicity: HepG2 cells


were pretreated for 15 min with 1 mM EDTA or 2 mM CaCl2 before
adding Zn acetate, ZnO-NP-FBS or ZnO-NP-silane at the toxic dose of
Fig. 4 Dissolution (%) of ZnO-NP-FBS () and ZnO-NP-silane () at 250 M for 24 h. EDTA or CaCl2 treatment alone had no eect on viabi-
90 M in complete MEM culture medium. lity of HepG2 cells.

Nanoscale This journal is The Royal Society of Chemistry 2016


View Article Online

Nanoscale Paper
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

Fig. 6 Cellular zinc content measured by ICP-AES after incubation with


90 M zinc acetate, ZnO-NP-FBS or ZnO-NP-silane for 6 h, 24 h and
48 h or for 24 h followed by a 24 h recovery.

In the literature, three dierent scenarii have been


described to explain ZnO-NP toxicity. In some studies, toxicity
appears to be solely linked to ions released in the biological
media and to the deleterious eects of ion entry in
cells.57,58,60,61 Others support a Trojan horse scenario with NP
dissolution in the acidic lysosomal compartment occurring
after cell entry.35,62 Yet in others there appears to be a specific
NP eect.63,64 In the latter case, NP did not enter the cell,
however dissolution and release of soluble zinc could only par-
tially account for the toxicity that was also linked to a direct
Fig. 7 TEM observation of HepG2 sections, A, B: control; C, E, G:
contact of NPs with the cell walls.
ZnO-NP-FBS treated cells (90 M, 6 h); D, F, H: ZnO-NP-silane treated
Our work is in agreement with the first scenario which is cells (90 M, 6 h). M: mitochondria, N: nucleus, G: Golgi apparatus, RE:
described in the majority of studies. This controversy on the endoplasmic reticulum, AM: abnormal mitochondria, MLB: multilamellar
cytotoxic potency of the ionic versus the NP form of zinc is body, A: autophagosome.
probably linked to experimental variables such as the NP
characteristics: their chemical nature, size, coating, mode of
dispersion, presence of corona65 and concentration as well as
cell types15,66 and culture media (DMEM or MEM versus
zinc. In contrast to our observations pertaining to CuO-NP
RPMI).67 All these parameters could influence the aggregation
uptake at similar concentrations in HepG2 cells,25 we conclude
and the kinetics of dissolution of NPs and thereby, their
that very little, if any, zinc enters and subsists in cells in the
toxicity.
nanoparticulate form. In the literature, a rare number of
studies showed ZnO-NPs inside hepatocytes only at high toxic
3.8 ZnO-NPs fate in HepG2 cells and eects on organelle doses, compatible with a saturation plateau of NP dissolution.
structure The maximum total dissolved zinc concentration that could be
A debate surrounds the eective entry of the nanoparticulate reached was estimated to be 225 M in complete DMEM by
form of zinc in cells via endocytosis or phagocytosis. ZnO-NPs Xia et al.53 and around 10 g mL1 (125 M) in complete
have been shown to be internalized by macrophages, cham- RPMI.70 Sharma et al.34 assessed cellular uptake of ZnO-NPs in
pions of rapid phagocytosis.35,68,69 Mechanisms of entry are HepG2 cells by an indirect measure using flow cytometry (light
however less clear for other cell types. In the present study, we scattering) no entry was found at 98 M but entry was detected
investigated the intracellular localization of ZnO-NPs using at 170 and 245 M (12 h). Using TEM, Wang et al.71 observed
transmission electronic microscopy (TEM) on 240 nm sections possible ZnO-NPs both inside and outside the cells, as well as
of embedded cells. Hepatocytes are energetically active cells, cell membrane damages, at the concentration of 2.45 mM
presenting an unusually high number of mitochondria and after 48 h, but both these studies did not confirm their nature
ribosomes as well as many dark lipid rich vesicles revealed by unequivocally by a chemical analysis such as EDX. In con-
osmium staining (Fig. 7). After 6 or 24 h of ZnO-NP treatment, clusion, it appears that only a few publications show convin-
we were unable to find any nanoparticles inside the cells. cing proof of ZnO-NP entry by TEM coupled to EDX35 or
Dense elements were analysed by EDX but were all negative for synchrotron61,72 observations and then, only in macrophages.

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

Macrophages are cultured in RPMI and as ZnO-NPs seem to 3.9 Zinc entry in mitochondria
dissolve less and re-precipitate more under phosphate and Proteins that transport zinc into mitochondria and distribute
carbonate forms66 in this medium, this coupled with the pha- it within them, as well as the control of these processes, are
gocytic nature of macrophages could explain this particularity. largely unknown. However accumulation of zinc has been visu-
Further TEM analysis revealed significant cell ultrastructure alized on (individual) isolated mitochondria via the uniporter-
alterations induced by sub-toxic doses of zinc. While gene dependent (calcium specific) and independent transport
expression analysis can only give an averaged result over the mechanisms.78 Interestingly, co-treatments with calcium
cell population, TEM analysis oers information on individual reduced mitochondrial zinc uptake, and this also explains the
cell responses. The first striking observation was that a subset
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

protective eect of calcium on entire cells that we report here


of cells showed no morphological alterations while others pre- (see section 3.6). Metallothionein has been also described to
sented a significant number of mitochondria with abnormal function as a metallochaperone for zinc delivery into prostate
morphologies and branching and in some cases doughnut and liver mitochondria,79 and zinc-loaded metallothioneins
shapes (Fig. 7C and D), suggesting disruptions in the fission imported into liver were proposed to modulate respiration.80
and fusion dynamics of the mitochondrial network at these Thus, increasing the amount of zinc-loaded metallothionein
sub-toxic doses. These cells also presented an increased in cells could lead to mitochondria disruptions. Furthermore,
number of autophagosomes (Fig. 7G and H) and the appear- ZnT2 is the first zinc transporter to be found directly associ-
ance of multilamellar bodies (MLBs) (Fig. 7E and F). Increased ated with mitochondria and possibly modulating their func-
autophagy in response to zinc was confirmed by observation of tion in secretory cells.81 MTF1 was shown to regulate the
more acidic vesicular organelles by flow cytometry analysis expression of ZnT2 in pancreatic acinar cells.82 Interestingly,
after treatment by acridine orange (Fig. S3). Autophagy is ZnT2 mRNA, while not normally detected in rat liver, shows
stimulated in response to a variety of environmental stresses to high expression in hepatocytes after an acute oral dose of
enable cellular survival;73 it plays a housekeeping role in zinc.43 In the light of these observations, we decided to investi-
removing misfolded or aggregated proteins, and clearing gate the expression of ZnT2 in HepG2 cells after zinc treatment
damaged organelles. ZnO-NPs have also been described to (Fig. 8). As the ZnT2 basal expression was very low in hepato-
enhance autophagy in macrophages and to simultaneously cytes, we worked with high concentrations of cDNA. We
induce apoptosis;21 indeed, the biogenesis of MLBs has been showed a 3040 fold over-expression of ZnT2 after all zinc
shown to result from autophagy in lung type II alveolar cells.74 treatments, corroborating a possible active sequestration of
In our case, we found no changes in the morphology of the zinc in mitochondria via ZnT2.
endoplasmic reticulum (ER) in contrast to what was observed
in human umbilical vein endothelial cells (HUVECs) where
doses of 120 M ZnO-NPs or ZnCl2 led to aggregation and 3.10 Investigation of mitochondria dysfunction
swelling of the ER, to activation of ER stress-responsive path- We measured the relative gene expression of Mfn1, Mfn2 and
ways and to a significant decrease in the number of mitochon- Drp1. Mfn1 and 2 are essential proteins for the mitochondrial
dria.58 Pathways involved in the response to ZnO-NP stress fusion process while Drp1 is an essential protein for the
thus appear to be cell type dependent. In agreement with our fission process. No significant dierences of expression were
results, zinc accumulation in mitochondria was demonstrated seen in our conditions of treatment (Fig. S4). Minor activation
using a selective fluorescent dye after ZnO-NP treatment at of these genes had been shown in rat primary astrocytes
similar doses, on leukemia Jurkat cells and human lung carci- exposed to titanium dioxide nanoparticles but only at high
noma H1355. Interestingly, higher doses induced also toxic doses.83
depolarization of mitochondrial membranes, caspase
activation and cell apoptosis.75
Excess zinc ions is known to be sequestered by metallothio-
nein, excreted via ZnT1 and also stored in intracellular com-
partments, but the nature of these compartments stirs contro-
versy and probably depends on the cell type. Evidence was
found that zinc can be stored in mitochondria, in the ER, in
the Golgi apparatus, in lysosomes40 as well as in the contro-
versial zincosomes.2
Our working hypothesis is that zinc accumulates primarily
in mitochondria where it aects mitochondrial network
dynamics. Once autophagy has started, MLBs are formed to
sequester the damaged mitochondria. This hypothesis is also Fig. 8 Quantitative PCR analysis of ZnT2 mRNA expression in HepG2
cells after a 6 hour incubation with Zn acetate, ZnO-NP-FBS or
backed by the observations that orally administered zinc salts
ZnO-NP-silane at 90 M. Results are presented as relative expression
increase zinc content in rat liver mitochondria,76 reduces mito- changes after normalizing with HPRT and 36B4 mRNA. Each value rep-
chondrial potential, alters mitochondrial metabolism and resents the mean of relative expression +SEM of three independent
leads to decreased ATP production in hepatocytes.77 experiments.

Nanoscale This journal is The Royal Society of Chemistry 2016


View Article Online

Nanoscale Paper
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

Fig. 9 Measurement of mitochondrial transmembrane potential after


incubation with 90 M zinc acetate, ZnO-NP-FBS or ZnO-NP-silane for
24 h. This gure shows the amount of rhodamine 123 internalized in the
cells expressed as a percentage of the mean uorescence of control
Fig. 10 Eect of balomycin A on Zn toxicity: HepG2 cells were pre-
cells (three independent experiments). ***: p < 0.001.
treated 1 hour with 100 nM Baf A before adding Zn acetate,
ZnO-NP-FBS or ZnO-NP-silane at subtoxic doses of 90 and 150 M for
24 h. Baf A treatment alone had no eect on viability of HepG2 cells.
In parallel, we explored mitochondrial integrity by measur-
ing mitochondrial transmembrane potential. In agreement
with literature on ZnO-NPs and others nanoparticles,24,34,39 we liberation from IONPs.88 Therefore, when NPs are internalized
observed a decrease in the mitochondrial potential using a and further dissolved inside the acidic lysosomes (known as
Rhodamine 123 uptake assay (Fig. 9) confirming mitochon- the Trojan horse mechanism), blocking lysosomal acidifica-
drial damage after zinc treatments. In neurones, it has also tion has a protective eect. In our conditions, Baf A did not
been shown that consequences of cellular zinc overload have this protective eect, suggesting again that zinc enters
include increased ROS production, loss of mitochondrial cells in its ionic form using another mechanism. We showed
membrane potential and reduced cellular ATP levels.84 ROS an increase in acidic vesicles after treatment with sub-toxic
generation could be a consequence of zinc stress and not the zinc concentrations, suggesting a defence mechanism by auto-
first actor leading to toxicity. Interestingly it has been shown phagy, limiting in these conditions zinc release from aected
that fusion of mitochondria in mammalian cells is dependent mitochondria. At higher doses, in contrast, autophagy could
on the mitochondrial inner membrane potential.85 Therefore, trigger apoptosis. Consequently, inhibiting autophagy by Baf A
the morphological modifications of mitochondria that we have at low doses of zinc may prevent cells from protecting them-
observed in TEM could be a direct consequences of the selves against it and therefore potentiates zinc-induced
decrease of mitochondrial potential. toxicity. In parallel to this eect, recently published results
In a study on ZnO-NP treated macrophages, Aude-Garcia suggest that lysosomes play a role as zinc sinks, temporarily
et al.86 showed as well convincing arguments in favour of res- storing zinc. Indeed, Kukic et al.89 found that the inhibition of
piratory chain defects rather than ROS generation as direct the lysosomal zinc sink function by Baf A increased cyto-
causes of mitochondrial potential alteration. We can propose plasmic zinc levels and favoured apoptosis. They found that
the same scenario in hepatocytes based on our results. lysosomes actively absorb zinc and secrete it after fusion with
the plasma membrane. Such phenomena could also explain
3.11 Role of autophagy in cell survival in response to zinc our results on zinc accumulation in cells which reaches a
In order to test the role of autophagy in cell response to zinc, plateau in 24 to 48 h. Blocking this process could lead to more
we blocked the lysosomal proton pump in HepG2 cells using accumulation of zinc in cells. ZnT2 transporters are generally
bafilomycin A (Baf A) prior to exposing the cells to 90 or 150 M present only in mitochondria, however some authors90 have
of ZnO-NPs or zinc acetate (Fig. 10). We checked the eciency found them also in lysosomes. So we can hypothesise an
of the Baf A treatment with an acridine orange assay and con- accumulation of zinc in lysosomes via autophagy of poisoned
firmed the total disappearance of acidic vesicles in these con- mitochondria or directly via ZnT2 or both. Maintenance of
ditions. Interestingly we found that the pre-treatment by Baf zinc homeostasis is a complex process and redundancy is to be
A potentiated the toxicity of zinc in all cases, ions or NPs, in expected, which makes it unlikely that mitochondria are the
contrast to what is usually reported in NP toxicity studies. only organelles acting as zinc sinks.
Indeed, Baf A has been shown to inhibit copper ion release
in acidic lysosomes and to significantly rescue A549 cells from
cell death induced by copper NP internalization and dis- 4 Conclusion
solution.87 In a similar study, the toxicity of iron oxide nano-
particles (IONPs) in microglial cultures was prevented by In this work on HepG2 cells, at a sub-toxic dose of 90 M, we
neutralizing lysosomal pH by Baf A, thus avoiding rapid iron found no evidence of zinc entry in its particulate form and on

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

the contrary, many arguments in favour of major dissolution 4 Q. Hao and W. Maret, JAD, J. Alzheimers Dis., 2005, 8, 161
of nanoparticles outside the cell and zinc entry in its ionic 170; discussion 209115.
form. 5 W. Maret and Y. Li, Chem. Rev., 2009, 109, 46824707.
We demonstrate that the entry of zinc induced a zinc stress 6 I. Sekler, S. L. Sensi, M. Hershfinkel and W. F. Silverman,
response and the up-regulation of MET, ZnT1, ZnT2. No Mol. Med., 2007, 13, 337343.
change in CAT, SOD1 and SOD2 expression, nor in the enzy- 7 L. A. Lichten and R. J. Cousins, Annu. Rev. Nutr., 2009, 29,
matic activities of catalase and SOD, were highlighted, imply- 153176.
ing a minimal activation of oxidative stress. TEM analysis 8 V. Gunther, U. Lindert and W. Schaner, Biochim. Biophys.
however put to light early mitochondria injury under sub-toxic Acta, 2012, 1823, 14161425.
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

exposure to zinc, suggesting that excess zinc is mainly trans- 9 E. Roduner, Chem. Soc. Rev., 2006, 35, 583592.
ported and stored in mitochondria. ZnT2 up-regulation 10 O. Bondarenko, K. Juganson, A. Ivask, K. Kasemets,
strengthens this hypothesis as well as the induction of auto- M. Mortimer and A. Kahru, Arch. Toxicol., 2013, 87, 1181
phagy. Zinc entry in mitochondria could be driven by MET 1200.
and/or ZnT2 and lead to mitochondria disruption. Under sub- 11 G. Oberdorster, A. Maynard, K. Donaldson, V. Castranova,
toxic conditions, cells are able to deal with excess zinc and J. Fitzpatrick, K. Ausman, J. Carter, B. Karn, W. Kreyling,
eliminate deleterious mitochondria by increasing autophagy D. Lai, S. Olin, N. Monteiro-Riviere, D. Warheit and
but at higher doses, mitochondria injury could lead to apopto- H. Yang, Part. Fibre Toxicol., 2005, 2, 8.
sis. Working at low doses allowed us to study the early 12 A. A. Keller, S. McFerran, A. Lazareva and S. Suh,
response of hepatocytes and to highlight the first element of J. Nanopart. Res., 2013, 15, 1692.
the response cascade. Future perspectives concern the study of 13 M. J. Osmond and M. J. McCall, Nanotoxicology, 2010, 4,
the subcellular fate of zinc in cells after exposure to ZnO-NPs 1541.
using synchrotron approaches in order to confirm zinc 14 J. T. Seil and T. J. Webster, Int. J. Nanomed., 2012, 7, 2767
accumulation in mitochondria. 2781.
15 R. J. Vandebriel and W. H. De Jong, Nanotechnol., Sci. Appl.,
2012, 5, 6171.
Acknowledgements 16 L. Wang, W. Ding and F. Zhang, J. Nanosci. Nanotechnol.,
2010, 10, 86178624.
The authors thank Aurlien Deniaud for useful discussions, 17 G. J. Nohynek, J. Lademann, C. Ribaud and M. S. Roberts,
Vronique Collin-Faure for FACS experiments and Josiane Crit. Rev. Toxicol., 2007, 37, 251277.
Arnaud (CHU Grenoble) for metal quantification. This work 18 A. Adamcakova-Dodd, L. V. Stebounova, J. S. Kim,
was funded by the CEA-Toxicology Transversal Program S. U. Vorrink, A. P. Ault, P. T. OShaughnessy, V. H. Grassian
through the NanoStress grant. This research is part of the and P. S. Thorne, Part. Fibre Toxicol., 2014, 11, 15.
LabEx SERENADE (grant ANR-11-LABX-0064) and the LabEx 19 C. H. Li, C. C. Shen, Y. W. Cheng, S. H. Huang, C. C. Wu,
ARCANE (grant ANR-11-LABX-0003-01). The platforms of the C. C. Kao, J. W. Liao and J. J. Kang, Nanotoxicology, 2012, 6,
Grenoble Instruct Centre (ISBG; UMS 3518 746756.
CNRS-CEA-UJF-EMBL) were used, with support from FRISBI 20 J. Heim, E. Felder, M. N. Tahir, A. Kaltbeitzel,
(ANR-10-INSB-05-02) and GRAL (ANR-10-LABX-49-01), within U. R. Heinrich, C. Brochhausen, V. Mailander, W. Tremel
the Grenoble Partnership for Structural Biology (PSB). The and J. Brieger, Nanoscale, 2015, 7, 89318938.
electron microscope facility is supported by the Rhne-Alpes 21 R. Roy, S. K. Singh, L. K. Chauhan, M. Das, A. Tripathi and
Region, the Fondation pour la Recherche Mdicale, the Fonds P. D. Dwivedi, Toxicol. Lett., 2014, 227, 2940.
Europen de Dveloppement conomique et Rgional, the 22 K. N. Yu, T. J. Yoon, A. Minai-Tehrani, J. E. Kim, S. J. Park,
Centre National de la Recherche Scientifique, the M. S. Jeong, S. W. Ha, J. K. Lee, J. S. Kim and M. H. Cho,
Commissariat lnergie Atomique, the Universit de Toxicol. In Vitro, 2013, 27, 11871195.
Grenoble Alpes, EMBL and the GIS-Infrastructures en Biologie 23 A. Kermanizadeh, G. Pojana, B. K. Gaiser, R. Birkedal,
Sant et Agronomie (IBISA). D. Bilanicova, H. Wallin, K. A. Jensen, B. Sellergren,
G. R. Hutchison, A. Marcomini and V. Stone,
Nanotoxicology, 2013, 7, 301313.
References 24 C. Filippi, A. Pryde, P. Cowan, T. Lee, P. Hayes,
K. Donaldson, J. Plevris and V. Stone, Nanotoxicology, 2015,
1 W. Maret, Biochemistry, 2004, 43, 33013309. 9, 126134.
2 D. Beyersmann and H. Haase, Biometals, 2001, 14, 331341. 25 M. Cuillel, M. Chevallet, P. Charbonnier, C. Fauquant,
3 R. A. Cherny, C. S. Atwood, M. E. Xilinas, D. N. Gray, I. Pignot-Paintrand, J. Arnaud, D. Cassio, I. Michaud-Soret
W. D. Jones, C. A. McLean, K. J. Barnham, I. Volitakis, and E. Mintz, Nanoscale, 2014, 6, 17071715.
F. W. Fraser, Y. Kim, X. Huang, L. E. Goldstein, R. D. Moir, 26 P. Muller, H. van Bakel, B. van de Sluis, F. Holstege,
J. T. Lim, K. Beyreuther, H. Zheng, R. E. Tanzi, C. Wijmenga and L. W. Klomp, JBIC, J. Biol. Inorg. Chem.,
C. L. Masters and A. I. Bush, Neuron, 2001, 30, 665676. 2007, 12, 495507.

Nanoscale This journal is The Royal Society of Chemistry 2016


View Article Online

Nanoscale Paper

27 R. F. Beers and I. W. Sizer, J. Biol. Chem., 1952, 195, 133 50 G. K. Andrews, Biometals, 2001, 14, 223237.
140. 51 J. D. Van Norman and R. Szentirmay, Bioinorg. Chem.,
28 S. Marklund and G. Marklund, Eur. J. Biochem., 1974, 47, 1974, 4, 3743.
469474. 52 S. H. Jeong, H. J. Kim, H. J. Ryu, W. I. Ryu, Y. H. Park,
29 M. M. Bradford, Anal. Biochem., 1976, 72, 248254. H. C. Bae, Y. S. Jang and S. W. Son, J. Dermatol. Sci., 2013,
30 S. Paglin, T. Hollister, T. Delohery, N. Hackett, 72, 263273.
M. McMahill, E. Sphicas, D. Domingo and J. Yahalom, 53 T. Xia, M. Kovochich, M. Liong, J. I. Zink and A. E. Nel, ACS
Cancer Res., 2001, 61, 439444. Nano, 2008, 2, 8596.
31 L. V. Johnson, M. L. Walsh and L. B. Chen, Proc. Natl. Acad. 54 C. Loos, T. Syrovets, A. Musyanovych, V. Mailnder,
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

Sci. U. S. A., 1980, 77, 990994. K. Landfester, G. U. Nienhaus and T. Simmet, Beilstein
32 K. W. Powers, S. C. Brown, V. B. Krishna, S. C. Wasdo, J. Nanotechnol., 2014, 5, 24032412.
B. M. Moudgil and S. M. Roberts, Toxicol. Sci., 2006, 90, 55 R. Guan, T. Kang, F. Lu, Z. Zhang, H. Shen and M. Liu,
296303. Nanoscale Res. Lett., 2012, 7, 602.
33 C. M. Sayes and D. B. Warheit, Wiley Interdiscip. Rev.: 56 R. Herrmann, F. J. Garcia-Garcia and A. Reller, Beilstein
Nanomed. Nanobiotechnol., 2009, 1, 660670. J. Nanotechnol., 2014, 5, 20072015.
34 V. Sharma, D. Anderson and A. Dhawan, Apoptosis, 2012, 57 S. K. Misra, A. Dybowska, D. Berhanu, S. N. Luoma and
17, 852870. E. Valsami-Jones, Sci. Total Environ., 2012, 438, 225232.
35 S. Triboulet, C. Aude-Garcia, L. Armand, A. Gerdil, 58 T. Buerki-Thurnherr, L. Xiao, L. Diener, O. Arslan,
H. Diemer, F. Proamer, V. Collin-Faure, A. Habert, C. Hirsch, X. Maeder-Althaus, K. Grieder, B. Wampfler,
J. M. Strub, D. Hanau, N. Herlin, M. Carriere, A. Van S. Mathur, P. Wick and H. F. Krug, Nanotoxicology, 2013, 7,
Dorsselaer and T. Rabilloud, Nanoscale, 2014, 6, 61026114. 402416.
36 M. O. Parat, M. J. Richard, C. Meplan, A. Favier and 59 B. M. Johnson, J. A. Fraietta, D. T. Gracias, J. L. Hope, C.
J. C. Beani, Biol. Trace Elem. Res., 1999, 70, 5168. J. Stairiker, P. R. Patel, Y. M. Mueller, M. D. McHugh, L.
37 E. Emri, E. Miko, P. Bai, G. Boros, G. Nagy, D. Rozsa, J. Jablonowski, M. A. Wheatley and P. D. Katsikis,
T. Juhasz, C. Hegedus, I. Horkay, E. Remenyik and G. Emri, Nanotoxicology, 2015, 9, 737748.
Metallomics, 2015, 7, 499507. 60 R. Chen, L. Huo, X. Shi, R. Bai, Z. Zhang, Y. Zhao, Y. Chang
38 P. J. Moos, K. Olszewski, M. Honeggar, P. Cassidy, and C. Chen, ACS Nano, 2014, 8, 25622574.
S. Leachman, D. Woessner, N. S. Cutler and J. M. Veranth, 61 V. Wilhelmi, U. Fischer, H. Weighardt, K. Schulze-Ostho,
Metallomics, 2011, 3, 11991211. C. Nickel, B. Stahlmecke, T. A. Kuhlbusch, A. M. Scherbart,
39 S. Triboulet, C. Aude-Garcia, M. Carriere, H. Diemer, C. Esser, R. P. Schins and C. Albrecht, PLoS One, 2013, 8,
F. Proamer, A. Habert, M. Chevallet, V. Collin-Faure, e65704.
J. M. Strub, D. Hanau, A. Van Dorsselaer, N. Herlin-Boime 62 S. A. James, B. N. Feltis, M. D. de Jonge, M. Sridhar,
and T. Rabilloud, Mol. Cell. Proteomics, 2013, 12, 3108 J. A. Kimpton, M. Altissimo, S. Mayo, C. Zheng,
3122. A. Hastings, D. L. Howard, D. J. Paterson, P. F. Wright,
40 R. A. Colvin, W. R. Holmes, C. P. Fontaine and W. Maret, G. F. Moorhead, T. W. Turney and J. Fu, ACS Nano, 2013, 7,
Metallomics, 2010, 2, 306317. 1062110635.
41 M. Vasak and G. Meloni, JBIC, J. Biol. Inorg. Chem., 2011, 63 P. J. Moos, K. Chung, D. Woessner, M. Honeggar,
16, 10671078. N. S. Cutler and J. M. Veranth, Chem. Res. Toxicol., 2010,
42 P. Babula, M. Masarik, V. Adam, T. Eckschlager, 23, 733739.
M. Stiborova, L. Trnkova, H. Skutkova, I. Provaznik, 64 V. Valdiglesias, C. Costa, G. Kilic, S. Costa, E. Pasaro,
J. Hubalek and R. Kizek, Metallomics, 2012, 4, 739750. B. Laon and J. P. Teixeira, Environ. Int., 2013, 55, 92100.
43 J. P. Liuzzi and R. J. Cousins, Annu. Rev. Nutr., 2004, 24, 65 I. L. Hsiao and Y. J. Huang, J. Nanopart. Res., 2013, 15,
151172. 1829.
44 T. Ganz and E. Nemeth, Biochim. Biophys. Acta, 2012, 1823, 66 T. Xia, M. Kovochich, M. Liong, L. Madler, B. Gilbert,
14341443. H. Shi, J. I. Yeh, J. I. Zink and A. E. Nel, ACS Nano, 2008, 2,
45 S. Balesaria, B. Ramesh, H. McArdle, H. K. Bayele and 21212134.
S. K. Srai, FEBS Lett., 2010, 584, 719725. 67 R. B. Reed, D. A. Ladner, C. P. Higgins, P. Westerho and
46 G. Thiel and G. Cibelli, J. Cell. Physiol., 2002, 193, 287292. J. F. Ranville, Environ. Toxicol. Chem., 2012, 31, 9399.
47 R. J. Cousins, R. K. Blanchard, M. P. Popp, L. Liu, J. Cao, 68 R. Roy, V. Parashar, L. K. Chauhan, R. Shanker, M. Das,
J. B. Moore and C. L. Green, Proc. Natl. Acad. Sci. U. S. A., A. Tripathi and P. D. Dwivedi, Toxicol. In Vitro, 2014, 28,
2003, 100, 69526957. 457467.
48 F. Wang, J. Dufner-Beattie, B.-E. Kim, M. J. Petris, 69 C. Shen, S. A. James, M. D. de Jonge, T. W. Turney,
G. Andrews and D. J. Eide, J. Biol. Chem., 2004, 279, 24631 P. F. Wright and B. N. Feltis, Toxicol. Sci., 2013, 136, 120
24639. 130.
49 S. K. Niture, R. Khatri and A. K. Jaiswal, Free Radical Biol. 70 W. Song, J. Zhang, J. Guo, F. Ding, L. Li and Z. Sun, Toxicol.
Med., 2014, 66, 3644. Lett., 2010, 199, 389397.

This journal is The Royal Society of Chemistry 2016 Nanoscale


View Article Online

Paper Nanoscale

71 Y. Wang, W. G. Aker, H. M. Hwang, C. G. Yedjou, H. Yu 82 L. Guo, L. A. Lichten, M. S. Ryu, J. P. Liuzzi, F. Wang and
and P. B. Tchounwou, Sci. Total Environ., 2011, 409, R. J. Cousins, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2818
47534762. 2823.
72 B. Gilbert, S. C. Fakra, T. Xia, S. Pokhrel, L. Madler and 83 C. L. Wilson, V. Natarajan, S. L. Hayward,
A. E. Nel, ACS Nano, 2012, 6, 49214930. O. Khalimonchuk and S. Kidambi, Nanoscale, 2015, 7,
73 D. Glick, S. Barth and K. F. Macleod, J. Pathol., 2010, 221, 1847718488.
312. 84 K. E. Dineley, T. V. Votyakova and I. J. Reynolds,
74 M. Hariri, G. Millane, M. P. Guimond, G. Guay, J. W. Dennis J. Neurochem., 2003, 85, 563570.
and I. R. Nabi, Mol. Biol. Cell, 2000, 11, 255268. 85 Y. Mattenberger, D. I. James and J.-C. Martinou, FEBS Lett.,
Published on 20 October 2016. Downloaded by Cornell University Library on 27/10/2016 06:54:32.

75 Y. Y. Kao, Y. C. Chen, T. J. Cheng, Y. M. Chiung and 2003, 538, 5359.


P. S. Liu, Toxicol. Sci., 2012, 125, 462472. 86 C. Aude-Garcia, B. Dalzon, J.-L. Ravanat, V. Collin-Faure,
76 M. Yamaguchi, M. Kura and S. Okada, Chem. Pharm. Bull., H. Diemer, J. M. Strub, S. Cianferani, A. Van Dorsselaer,
1981, 29, 23702374. M. Carrire and T. Rabilloud, J. Proteomics, 2016, 134, 174
77 J. Lemire, R. Mailloux and V. D. Appanna, J. Appl. Toxicol., 185.
2008, 28, 175182. 87 E. Moschini, M. Gualtieri, M. Colombo, U. Fascio,
78 L. M. Malaiyandi, O. Vergun, K. E. Dineley and M. Camatini and P. Mantecca, Toxicol. Lett., 2013, 222,
I. J. Reynolds, J. Neurochem., 2005, 93, 12421250. 102116.
79 L. C. Costello, Z. Guan, R. B. Franklin and P. Feng, J. Inorg. 88 C. Petters, K. Thiel and R. Dringen, Nanotoxicology, 2015,
Biochem., 2004, 98, 664666. 111, DOI: 10.3109/17435390.2015.1071445.
80 B. Ye, W. Maret and B. L. Vallee, Proc. Natl. Acad. 89 I. Kukic, S. L. Kelleher and K. Kiselyov, J. Cell Sci., 2014,
Sci. U. S. A., 2001, 98, 23172322. 127, 30943103.
81 Y. A. Seo, V. Lopez and S. L. Kelleher, Am. J. Physiol.: Cell 90 R. D. Palmiter, T. B. Cole and S. D. Findley, EMBO J., 1996,
Physiol., 2011, 300, C1479C1489. 15, 17841791.

Nanoscale This journal is The Royal Society of Chemistry 2016

You might also like