You are on page 1of 11

Toxicology in Vitro 29 (2015) 1492–1502

Contents lists available at ScienceDirect

Toxicology in Vitro
journal homepage: www.elsevier.com/locate/toxinvit

Magnetite nanoparticles induced adaptive mechanisms counteract


cell death in human pulmonary fibroblasts
Mihaela Radu a,b, Diana Dinu a, Cornelia Sima c, Radu Burlacu d, Anca Hermenean b,e, Aurel Ardelean e,
Anca Dinischiotu a,⇑
a
Department of Biochemistry and Molecular Biology, University of Bucharest, 91-95 Splaiul Independentei, Bucharest 050095, Romania
b
Department of Histology, Faculty of Medicine, Pharmacy and Dentistry, Vasile Goldis Western University of Arad, 1 Feleacului, Arad 310396, Romania
c
Laser Department, National Institute of Laser, Plasma and Radiation Physics, 409 Atomistilor, Bucharest-Magurele 077125, Romania
d
Department of Mathematics, University of Agriculture Sciences and Veterinary Medicine, 59 Marasti, Bucharest 011464, Romania
e
Department of Experimental and Applied Biology, Institute of Life Sciences, Vasile Goldis Western University of Arad, 86 Rebreanu, Arad 310414, Romania

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

Article history: Magnetite nanoparticles (MNP) have attracted great interest for biomedical applications due to their
Received 21 November 2014 unique chemical and physical properties, but the MNP impact on human health is not fully known.
Revised 28 May 2015 Consequently, our study proposes to highlight the biochemical mechanisms that underline the toxic
Accepted 4 June 2015
effects of MNP on a human lung fibroblast cell line (MRC-5). The cytotoxicity generated by MNP in
Available online 9 June 2015
MRC-5 cells was dose and time-dependent. MNP-treated MRC-5 cells accumulated large amount of iron
and reactive oxygen species (ROS) and exhibited elevated antioxidant scavenger enzymes. Reduced
Keywords:
glutathione (GSH) depletion and enhanced lipid peroxidation (LPO) processes were also observed. The
Magnetite nanoparticles
MRC-5 cells
cellular capacity to counteract the oxidative damage was sustained by high levels of heat shock protein
Oxidative stress 60 (Hsp60), a protein that confers resistance against ROS attack and inhibition of cell death. While signif-
Hsp60 icant augmentations in nitric oxide (NO) and prostaglandine E2 (PGE2) levels were detected after 72 h of
PGE2 MNP-exposure only, caspase-1 was activated earlier starting with 24 h post-treatment. Taken together,
Caspase-1 our results suggest that MRC-5 cells have the capacity to develop cell protection mechanisms against
MNP. Detailed knowledge of the mechanisms induced by MNP in cell culture could be essential for their
prospective use in various in vivo biochemical applications.
Ó 2015 Published by Elsevier Ltd.

1. Introduction processes. Cellular alterations induced by NPs can be associated


with diseases such as cancer and degenerative pathologies. The
Currently, living organisms are constantly exposed to naturally toxicity of NPs depends on various factors, including: size, compo-
occurring, anthropogenic or engineered nanoparticles (NPs). The sition, crystallinity, surface functionalization, surface charge,
skin, as well as the respiratory and gastrointestinal tracts are entry aggregation, and solubility, but also the individual’s genetic traits
points for these particles in organisms. NPs, which have at least that can provide or not the capacity to counteract the effects of
one dimension smaller than one micrometer, can interfere with these particles (Buzea et al., 2007).
cellular function to influence proliferation, metabolism and death Magnetite nanoparticles (MNP) are abundant constituents of
the ambient airborne particulate matter in the urban environment,
resulting especially from traffic, industry and power stations emis-
Abbreviations: MNP, magnetite nanoparticles; MRC-5, human lung fibroblast
cell line; ROS, reactive oxygen species; GSH, reduced glutathione; LPO, lipid
sions (Grobéty et al., 2010; Karlsson et al., 2008). Furthermore,
peroxidation; Hsp60, heat shock protein 60; NO, nitric oxide; PGE2, prostaglandine mixed magnetite–maghemite nanoparticles are potentially adsor-
E2; MRI, magnetic resonance imaging; SOD, superoxide dismutase; CAT, catalase; bents for arsenic and chromium which could facilitate their
GPX, glutathione peroxidase; GR, glutathione reductase; HRTEM, high resolution removal from contaminated water (Chowdhury and Yanful, 2010).
transmission electron microscopy; SAED, selected area electron diffraction; XRD, X
MNP are of special interest in biological and medical sciences
ray diffraction; MTT, 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bro-
mide; H2DCFDA, 20 ,70 -dichlorodihydrofluorescein diacetate; DCF, dichlorofluores- due to their superior biocompatibility compared to other magnetic
cein; GST, glutathione S-transferase. materials, their chemical stability in physiological media, and
⇑ Corresponding author. because they are easy and inexpensive to synthesize (Figuerola
E-mail addresses: ancadinischiotu@yahoo.com, anca.dinischiotu@bio.unibuc.ro et al., 2010). As a result, the synthesis of MNP has rapidly
(A. Dinischiotu).

http://dx.doi.org/10.1016/j.tiv.2015.06.002
0887-2333/Ó 2015 Published by Elsevier Ltd.
M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502 1493

developed over the last few years for fundamental biomedical cell death following MNP exposure may be a necessary step to
applications; their varied uses include that of contrast agents for characterize the effects of these nanoparticles on living cells.
magnetic resonance imaging (MRI) (Maity et al., 2010), heating Despite the great potential of MNP in biology and medicine, the
mediators for cancer therapy (Ito et al., 2007), as well as in vitro cell mechanisms by which they can induce toxicity in human cells has
sorting and magnetically guided drug delivery (Lin et al., 2007). In not been completely elucidated. The present study was, therefore,
addition, the ability of magnetite to combine both therapeutic and undertaken in order to investigate the MNP biological effects on
diagnostic functions into one nanoparticle has attracted great human lung fibroblasts (MRC-5). Mechanisms of induced oxidative
interest for its potential theranostic applications (Shaw et al., stress and its effects on various enzymatic and non-enzymatic
2008). Beside medical applications, ferrofluids containing MNP antioxidants, as well as the modulation of Hsp60 proteins expres-
are also used in electronic devices, mechanical engineering, analyt- sion, nitric oxide (NO), prostaglandin E2 (PGE2) and caspase-1
ical instrumentation, heat transfer and art. release were investigated.
The principal route for MNP absorption inside the human body
is respiration. Once inside, MNP translocate to the circulatory and 2. Materials and methods
lymphatic systems first, subsequently reaching tissues and organs.
Depending on particle size, inhaled nanoparticles can be deposited 2.1. Chemicals
throughout the human respiratory system including pharyngeal,
nasal, tracheobronchial and alveolar regions (Price et al., 2002). GIBCOÒ Modified Eagle’s Medium (MEM), fetal bovine serum,
As the use of NPs has increased significantly, research aimed at phosphate buffered saline (PBS), Trizol reagent and the Western
uncovering the potential risks to human health is mandatory. Breeze Chromogenic Kit-Anti-Mouse were purchased from
Understanding and controlling the interaction between MNP and Invitrogen (Carlsbad, California, USA). Nicotinamide adenine
biomolecules at the molecular and cellular level is crucial for suc- dinucleotide phosphate disodium salt (NADP+) and nicotinamide
cessfully designing MNP for potentially novel biological adenine dinucleotide phosphate reduced tetrasodium salt
applications. (NADPH) were supplied by Merck (Darmstadt, Germany).
Despite the extensive use of MNP, the studies published to date Tetramethoxypropane (TEP) and thiobarbituric acid (TBA) were
regarding their toxicity are inconsistent. Both in vitro and in vivo obtained from Fluka (Milwaukee, USA). The Detect XÒ
studies have demonstrated that the presence of MNP can exert Glutathione Colorimetric Detection Kit was purchased from Arbor
lower toxic effects than those established for molecular iron Assay (Michigan, USA). Antibodies for HSP 60 and b-actin were
(Pisanic et al., 2009), with some reports highlighting a low toxicity purchased from Santa Cruz Biotechnology (Santa Cruz, CA). All
for MNP at doses of 100 lg/mL or higher (Ankamwar et al., 2010). other chemicals used were of analytical grade and were from
By contrast, other studies have coupled the exposure to MNP to Sigma (St. Louis, Missouri, USA).
notable toxic effects such as: inflammation, the formation of
apoptotic bodies, impaired mitochondrial function, membrane
2.2. Nanoparticles
integrity alteration, generation of reactive oxygen species (ROS),
and chromosome condensation (Park et al., 2010; Ramesh et al.,
MNP were provided by NaBond Technologies Co., Ltd, China.
2012). In addition, it was demonstrated recently that magnetite
They have been characterized from the morpho-structural and size
deposition in rat lungs caused goblet cell hyper- and/or metaplasia
distribution point of view using high resolution transmission elec-
in the upper respiratory tract, increased lung and
tron microscopy (HRTEM, Philips CM120 model), selected area
lung-associated-lymph node weights and induced inflammatory
electron diffraction (SAED) and X ray diffraction (XRD,
changes in the bronchiolo-alveolar region (Pauluhn, 2012).
Bruker-AXS D8 ADVANCE).
Finally, ROS formation in the human alveolar epithelial-like
type-II cells could also contribute to MNP-induced genotoxicity
(Könczöl et al., 2011). 2.3. Cell culture conditions and treatment
Oxidative stress and ROS formation has been considered as an
important mechanism responsible for MNP-induced toxicity MRC-5 (purchased from the American Type Culture Collection)
(Singh et al., 2010). After the intracellular uptake via different is a fibroblast cell line derived from the normal lung tissue of a
mechanisms such as: receptor-mediated endocytosis, 14-week-old male and was chosen for this study due to its previ-
caveolin-mediated internalization, and macropynocitosis, MNP ously demonstrated suitability for NP toxicity studies (Li et al.,
may presumably release iron ions within lysosomes in the pres- 2010; Radu et al., 2010). The cells were cultured in minimum
ence of hydrolyzing enzymes effective at low pH values. Iron, being essential medium (MEM) supplemented with 1% antibiotic antimy-
a highly redox-active transition metal, can be involved in the cotic solution (mixture of penicillin, streptomycin and ampho-
Fenton reaction producing high reactive hydroxyl radicals that tericin B) and 10% fetal bovine serum at 37 °C in a 5% CO2
can probably attack lipids, DNA and proteins (Halliwell and humidified atmosphere. They were seeded at a density of
Gutteridge, 2007). In response, cells have developed several cellu- 7.5  105 cells/75 cm2 flask. The culture medium was changed
lar defense pathways, which under normal metabolic conditions every 3 days. All the experiments were carried out with cells
regulate the level of ROS and protect against the damage generated between passage number 11 and 20. The suspension of MNP was
by free radicals. These defense systems include both sterilized and sonicated before the cell culture treatment. The cells
low-molecular-weight free radical scavengers, such as the tripep- were incubated with nanoparticles for 24, 48 and 72 h. Controls
tide glutathione (GSH), and antioxidant enzymes, i.e. superoxide without MNP were performed for each experiment.
dismutase (SOD), catalase (CAT), glutathione peroxidase (GPX),
and glutathione reductase (GR) (Kehrer, 1993). 2.4. Cell viability/cytotoxicity
Cell survival during stress requires the induction of the heat
shock response. Production of high levels of heat shock proteins The viability of the cells was determined by the tetrazolium salt
(Hsp) can be triggered by exposure to different environmental test (Mosmann, 1983). The medium from each well was removed
stress conditions, such as an increase in temperature, the presence by aspiration, the cells were washed with 200 lL of PBS/well and
of environmental pollutants, free radicals, and so forth (Santoro, then a volume of 50 lL (1 mg/mL) of 3-(4,5-dimethylthiazo
2000). Thus, defining the interlinks between oxidative stress and l-2-yl)-2,5-diphenyltetrazolium bromide (MTT) solution was
1494 M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502

added to each well. After 2 h of incubation, the MTT solution from The glutathione S-transferase (GST) activity was assayed
each well was removed by aspiration. A volume of 50 lL iso- spectrophotometrically at 340 nm by measuring the rate of
propanol was added and the plate was shaken to dissolve the for- 1-chloro-2,4-dinitrobenzene (CDNB) conjugation with GSH,
mazan crystals. The absorbance at 595 nm was then determined according to the Habig et al. method (1974). One unit of GST activ-
using a Tecan multiplate reader (TecanGENios, Grödic, Germany), ity was expressed as one lmole of conjugated product per minute.
for each well. The absorbance for untreated cells was taken to rep- The CDNB extinction coefficient of 9.6 mM1 cm1 was used for the
resent the 100% viability. calculation.
GR activity was measured according to the method of Goldberg
2.5. Intracellular iron determination and Spooner (1983), in 0.1 M phosphate buffer, pH 7.4 with
0.66 mM GSSG and 0.1 mM NADPH by recording the decrease of
The level of intracellular iron was measured according to the absorbance at 340 nm. One unit of GR activity was defined as
Zhu assay (Zhu et al., 2012). Briefly, after treatment with MNP, cells one lmole of NADPH per minute under standard conditions.
were washed with phosphate-buffered saline (PBS), collected and All the enzymatic activities, calculated as specific activities
counted. Subsequently, the cell pellet was digested in a volume (U/mg of protein), were expressed as% from controls.
of 200 ll of 5 N HCl for 24 h at 37 °C. After centrifugation at
2000 rpm, for 10 min, a volume of 50 lL of supernatant was added 2.8. Lipid peroxidation assay
into a 96-well plate and then 50 lL of solution of 1% ammonium
persulfate were added to convert the ferrous ions to ferric ones. Malondialdehyde (MDA), a marker for lipid peroxidation, was
Finally, a volume of 100 lL of 0.5 M potassium thiocyanate was assayed via the thiobarbituric acid reaction according to
added in the solution and the mix was shaken for 5 min in order Dinischiotu et al. (2013). A volume of 200 lL of sample with a pro-
to develop the red colored iron-thiocyanate. The formed complexes tein concentration of 2 mg/mL was treated with 700 lL 0.1 N HCl
were spectrophotometrically detected at 480 nm. A standard curve and the mixture was incubated for 20 min at room temperature.
for iron was made under identical conditions using known amount Then, 900 lL of 0.025 M thiobarbituric acid (TBA) were added
of iron salts. and the total volume was incubated for 65 min at 37 °C. Finally, a
volume of 400 lL of 0.1 M Tris–HCl, 5 mM EDTA buffer, pH 7.4
was added. The fluorescence of MDA was recorded using a
2.6. Determination of intracellular ROS 520 nm/549 nm (excitation/emission) filter. A calibration curve
with 1,1,3,3-tetramethoxy propane in the range 0.05–5 lM was
A spectrofluorimetric procedure has been used to identify ROS used to calculate the MDA concentration. The results were
in MRC-5 cells by utilizing the dye 20 ,70 -dichlorodihydrofluorescein expressed as nmoles of MDA/mg protein.
diacetate (H2DCFDA) (Wan et al., 1993). After treatment with
12.5 lg/mL MNP, the cells were immediately incubated with 2.9. Reduced glutathione determination
10 lM H2DCFDA in the dark at 37 °C for 30 min. The cells were
washed twice and suspended in 2 mL of PBS solution. The fluores- The cellular lysate, deproteinized with 5% sulfosalicylic acid,
cence of dichlorofluorescein (DCF) was spectrofluorimetrically was analyzed for total glutathione and oxidized glutathione
recorded at kex 488 nm/kem 515 nm in each sample. (GSSG) using the Detect XÒ Glutathione colorimetric detection kit
and following manufacturer’s instructions. The reduced glu-
2.7. Antioxidant enzymes activities assays tathione (GSH) concentration was obtained by subtracting the
GSSG level from the total glutathione concentration. The GSH
MRC-5 cells, harvested from culture flasks, were washed with levels were calculated as nmoles/mg protein.
PBS three times, trypsinized, and centrifuged at 1500 rpm for
10 min at 4 °C. Cell pellets were resuspended in 0.3 mL of PBS 2.10. Western blotting
and then sonicated on ice three times, for 30 s each. The total
extract was centrifuged at 5000 rpm for 10 min at 4 °C. Aliquots Samples (25 lg protein) were run on 10% sodium dodecyl sul-
of the supernatant were used for the biochemical assays. The pro- fate polyacrylamide (SDS–PAGE) gels, using the Mini-Protean 3
tein concentration, expressed as mg/mL, was determinated by the system (Bio Rad Laboratories, Hercules, CA, USA), for Hsp60 detec-
Bradford method, using bovine serum albumine as standard tion. The protein bands were electro-blotted on PVDF membranes,
(Bradford, 1976). using the Mini Trans-Blot system (Bio Rad Laboratories, Hercules,
SOD activity was measured following the oxidation of NADPH at CA, USA). Membranes were maintained for 30 min in blocking
340 nm (Paoletti et al., 1986). Superoxide anions were generated solution, washed for 5 min in MiliQ water and incubated for one
from oxygen molecules in the presence of EDTA, MnCl2 and mer- hour with the primary antibodies for Hsp60. After rinsing three
captoethanol. A control was run with each set of three duplicate times with antibody wash solution the membrane was developed
samples. One unit of activity was defined as the amount of enzyme with Western Breeze Chromogenic Immunodetection Kit using
required to inhibit the rate of NADPH oxidation of the control by 5-bromo-4-chloro-3-indolyl-1-phosphate (BCIP)/nitro blue tetra-
50%. zolium (NBT) substrate for alkaline phosphatase. Beta-actin
CAT activity was assayed by monitoring the disappearance of protein (42 kDa) was used as an internal standard. The blots were
H2O2 at 240 nm, according to the Aebi method (1984). CAT activity captured and quantified using BioCapt MW and ImageJ analysis
was calculated in terms of U/mg protein, where one unit was the software. The values for each band were normalized to b-actin.
amount of enzyme that catalyzed the conversion of one lmole
H2O2 in a minute. 2.11. Quantitative RT-PCR
GPX was assayed by the Beutler (1971) method, using H2O2 and
NADPH as substrates. The conversion of NADPH to NADP+ was fol- Total RNA was extracted from MRC-5 cells using Trizol reagent
lowed by recording the changes in absorption intensity at 340 nm, according to the manufacturer’s instructions. RNA samples were
and one unit was expressed as one lmole of NADPH consumed per dissolved in diethylpyrocarbonate (DEPC)-treated water and the
minute, using a molar extinction coefficient of RNA concentration in each sample was determined by spectropho-
6.22  103 M1 cm1. tometry. One lg of each sample of total RNA was
M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502 1495

Table 1 2.15. Annexin-V/PI double-staining assay


Primer sets for RT-PCR analysis.

Gene name Primer sequence (50 –30 ) After exposure to MNP, MRC-5 cells were harvested, washed,
GAPDH Forward: GCCATCAGGTCACATACACG and resuspended in ice cold PBS. Apoptotic cells were determined
Reverse: GATACACGGAGCACCAGGTT with an FITC Annexin-V Apoptosis Detection Kit (Invitrogen,
hsp60 Forward: GTGGAAAAAGGAATCATTGACC California, USA) according to the manufacturer’s protocol. Briefly,
Reverse: GTAGTTAACAGAGAGGCCACACC the cells were centrifuged and subsequently incubated in the dark,
for 15 min, with 100 lL of binding buffer containing 5 lL of
Annexin V-FITC and 5 lL of propidium iodide (PI). Afterwards,
reverse-transcripted into cDNA using the complementary DNA apoptosis was analyzed by fluorescence microscopy (Olympus
synthesis kit M-MLV RT, in a total volume of 25 lL. The samples IX7 with U-RFL-T power supply unit).
were incubated for 5 min at 25 °C, 45 min at 42 °C, 10 min at
95 °C and then chilled on ice. For real-time PCR, a volume of 1 lL 2.16. DNA integrity assay
of hsp60 cDNA was amplified in 1 iQ SYBR Green Supermix con-
taining specific primer pairs (Table 1), using the iCycleriQ In order to check DNA integrity, 2  106 cells were incubated in
Real-Time PCR Detection System (Biorad), at a final volume of digestion buffer [100 mM NaCl, 10 mM Tris–HCl (pH 8), 25 mM
25 lL. PCR parameters were: an initial step for denaturation at EDTA (pH 8), 0.5% SDS, 0.1 mg/mL proteinase K] at 50 °C for 3 h
95 °C for 3 min followed by 45 cycles at 95 °C for 30 s, anneal/ex- with gentle agitation. To eliminate the RNA contamination, the
tend at 58 °C for 30 s followed by final extension at 72 °C for solution was incubated 1 h with 2 mg/mL RNase A at 37 °C. The
30 s. Following amplification, melting curves were generated to DNA from each sample was purified by phenol/chloroform/isoamyl
confirm the specificity of each primer pair with 85 cycles of alcohol (25:24:1) extraction, and then centrifuged at 10,000 rpm
increasing increments of 0.5 °C beginning with 55 °C for 10 s. The for 5 min. The supernatant containing DNA was transferred to a
samples were performed in triplicate and the clean tube and mixed with a half volume of 7.5 M ammonium
glyceraldehydes-3-phosphate dehydrogenase (GAPDH) mRNA acetate and 2 volumes of cold 100% ethanol. The precipitate was
was used for normalizing the hsp60 mRNA levels. centrifuged at 10,000 rpm for 3 min and the pellet was washed
with 70% ethanol, dried and dissolved in 50 lL of TE buffer
[10 mM Tris–HCl (pH 8), 1 mM EDTA]. The DNA samples thus
2.12. Caspase-1 activity assay
obtained were electrophoretically separated on a 1% agarose gel
at 80 V for 60 min. Subsequently, the gels were stained with ethid-
The activity of caspase-1, that recognizes the sequence YVAD,
ium bromide and visualized under UV light.
was assessed using the Caspase-1/ICE Colorimetric Protease
Assay Kit provided by BioVision. The assay was done by spec-
trophotometric detection of the chromophore p-nitroanilide 2.17. Statistical analysis
(pNA) after cleavage from the labeled substrate YVAD-pNA at
405 nm according to supplier’s instructions. Data were expressed in terms of mean ± standard deviation
(SD) of triplicate measurements for five independent experiments.
The results were compared using Student’s t-test and validated by
2.13. Nitric oxide determination
confidence intervals using the Quattro Pro X3 software. The results
were considered significant only if the P value was less than 0.05
NO concentration was measured by the Green assay (Green
and the confidence interval of control and treated cells samples
et al., 1982). In this method, nitrite is first treated with a diazotiz-
did not overlap (a = 0.05).
ing reagent, e.g., sulfanilamide (SA), in acidic media to form a tran-
sient diazonium salt. This intermediate is then allowed to react
with a coupling reagent, N-naphthyl-ethylenediamine (NED), to 3. Results
form a stable azo compound, measured spectrophotometrically at
550 nm. Thus, volumes of 80 lL of culture supernatant were added 3.1. MNP characteristics
to a microplate and mixed with a 1:1 solution of 10% NED and SA.
Then intensity of the developed color was measured spectrophoto- Nanoparticle crystallinity is demonstrated by the XRD pattern
metrically at a wavelength of 550 nm. A standard curve with differ- (Fig. 1) and the SAED image (Fig. 2C). The peaks representing mag-
ent concentrations of sodium nitrate was also performed. The NO netite in the crystalline phase can be observed in the XRD diagram
concentrations were expressed in lM. (Fig. 1).

2.14. Prostaglandin E2 immunoassay

PGE2 production was determined with a competitive


immunoassay using the Human PGE2 kit (Invitrogen, California,
US) according to the manufacturer’s instructions. Briefly,
PGE2-alkaline phosphatase tracer, PGE2-specific monoclonal anti-
body and either standard or sample (culture supernatants) were
added to each well of an 96-well microtiter plate pre-coated with
goat polyclonal anti-mouse IgG. After incubation for 2 h at room
temperature, the plate was washed and a solution of para-nitro-
phenyl phosphate (pNPP), as substrate for alkaline phosphatase
was added. The amount of yellow product formed in this enzy-
matic reaction was read at 405 nm, allowing quantification of
PGE2 in each sample. The PGE2 concentration was expressed in
pg/mL. Fig. 1. XRD of MNP.
1496 M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502

Fig. 2. HRTEM (A and B) and SAED (C) of MNP.

Fig. 4. Viability of MRC-cells following exposure to 2.5, 6.25 and 12.5 lg/mL MNP.
The values are calculated as means ± SD (n = 5) and expressed as % from controls.

P < 0.05 vs. controls.

Fig. 3. Size distribution of MNP.

The HRTEM images (Fig. 2) revealed that MNP have a rectangu-


lar shape.
The size distribution of MNP is shown in Fig. 3. This was inves-
tigated by performing a statistical assessment of the electron
microscopy images. The primary nanoparticle size is in the range
of 20–180 nm with the arithmetic mean value of approximately
50 nm. The size distribution is a lognormal function.

3.2. Cell viability


Fig. 5. Iron concentration in 12.5 lg/mL MNP treated-MRC-5 cells. Values are
The effects of MNP exposure on MRC-5 cells viability was dose calculated as means ± SD (n = 5). ⁄P < 0.05 vs. controls; ⁄⁄P < 0.01 vs. controls;
⁄⁄⁄
and time-dependent (Fig. 4). Incubation with doses of 2.5, 6.25 and P < 0.001 vs. controls.

12.5 lg/mL of MNP (corresponding to 1.81, 4.525 and 9.051 lg


iron/mL) for 24 and 48 h had no significant effect on viability.
Nevertheless, after 72 h exposure at all three doses, significant 11.91 ± 1.1 pg/cell for MNP incubation times of 24 h, 48 h and
and moderate decreases of cell viability were registered. Taking 72 h, respectively. Control samples of MNP untreated cells con-
into account these data, the MNP concentration used in subsequent tained close to 0.2 pg of iron per cell.
experiments was 12.5 lg/mL.

3.3. Intracellular iron concentration 3.4. ROS generation

The time-dependent changes in the iron level recorded for the A significant increase in ROS levels was observed for MRC-5
MRC-5 cells exposed to 12.5 lg/mL MNP are illustrated in Fig. 5. cells treated with for 12.5 lg/mL MNP starting with 48 h of expo-
Significant increases in iron concentration were detected for all sure, when the level increased by 52% compared to control. For
durations of MNP treatment. To be more specific, the iron accumu- longer periods of exposure, the ROS levels remained
lated in our cell line was 9.59 ± 0.8 pg/cell, 10.5 ± 0.9 pg/cell and up-regulated, but to a lesser extent (Fig. 6).
M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502 1497

by 11.7% and 38% after 48 and 72 h of treatment, respectively,


whereas CAT activity showed an increase by 9.5% and 16.9% for
the same periods of treatment. The activity of GPX was raised by
22.7% after 72 h of MNP administration. GST activity remained
unchanged after 12.5 lg/mL MNP treatment, whereas a decrease
in GR activity, by 33.2%, was recorded after 72 h of MNP
administration.

3.6. Lipid peroxidation and glutathione concentration

The levels of MDA and GSH recorded in MRC-5 cells are showed
in Table 2. Lipid peroxidation levels, expressed as MDA, were sig-
nificantly higher than in unexposed cells. A time-dependent
increase in the MDA level in MRC-5 cells was observed, with the
Fig. 6. ROS production induced by 12.5 lg/mL MNP in MRC-5 cells. Values are
most significant change of 70% being recorded after 72 h of MNP
calculated as means ± SD (n = 5) expressed as % from controls. ⁄P < 0.05 vs. controls.
exposure. By contrast, the GSH levels declined starting with 48 h
of MNP exposure, which culminated with a 60% decrease after
3.5. Antioxidant enzymes activity 72 h compared to control.

Fig. 7 shows the effects of 12.5 lg/mL MNP administration on 3.7. Protein level and gene expression of Hsp60
the activities of the antioxidant enzymes SOD (7A), CAT (7B) and
GPX (7C), as well as the activities of two enzymes involved in glu- MNP interfered with Hsp60 expression and protein level in
tathione metabolism, i.e. GST (7D) and GR (7E). An increase of SOD MRC-5 cells. A good correlation between Hsp60 mRNA expression
and CAT activities was observed starting with the second day of and Hsp60 protein level was observed. As shown in Fig. 8A and B,
MNP treatment (Fig. 7A and B), while GPX was activated after only the Hsp60 protein was up-regulated by MNP treatment. At 48 h
72 h of nanoparticles exposure (Fig. 7C). The SOD activity increased after MNP exposure, the level of this protein was 70% greater than

Fig. 7. Effects of 12.5 lg/mL MNP administration on SOD (A), CAT (B), GPX (C), GST (D) and GR (E) activities in MRC-5 cells. Values are calculated as means ± SD (n = 5) and
expressed as % from controls. ⁄P < 0.05 vs. controls.
1498 M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502

Table 2
GSH and MDA levels in MRC-5 cells after MNP treatment. Values are calculated as
means ± SD (n = 5) and expressed as % from controls.

Exposure Sample GSH MDA


time (nmoles/mg protein) (nmoles/mg protein)
24 h Control 100 ± 4.05 100 ± 6.22
MNP 106.21 ± 68.54 107.99 ± 8.6
48 h Control 100 ± 7.12 100 ± 7.3 2
MNP 52.53 ± 3.29* 115.97 ± 6.48
72 h Control 100 ± 48.29 100 ± 3.90
MNP 41.55 ± 6.46* 9.48*
*
P < 0.05 vs. controls.

Fig. 9. Activation of caspase-1 from MRC-5 cells by of 12.5 lg/mL MNP treatment.
that of the control. MNP stimulated Hsp60 mRNA expression in Values are calculated as means ± SD (n = 5) and expressed as % from controls.
MRC-5 cells in a time-dependent manner, with an increase by ⁄
p < 0.05 vs. controls; ⁄⁄p < 0.05 vs. controls.
119% and 166% after 48 and 72 h of MNP administration, respec-
tively (Fig. 8C).
(Fig. 11A). In addition, it was observed that DNA integrity was
not affected for the duration of the experiment (Fig. 11B).
3.8. Caspase-1 activity

Our studies showed that MNP induced moderate caspase-1 acti- 4. Discussion
vation in MRC-5 cells. To be more specific, the levels of this protein
were increased by 35%, 21% and 25% after 24, 48 and 72 h, respec- The tremendous potential of MNP in terms of developing new
tively (Fig. 9). therapeutic approaches make these nanoparticles of great interest
for biomedical research. As a result, it has become increasingly
important to understand the biological response to their adminis-
3.9. Inflammatory markers, NO and PGE2
tration, considering that the main limitation in MNP applications is
their toxicity. Thus, this study was initiated in order to evaluate the
We analyzed the levels of NO and PGE2, two molecules gener-
possible toxic effects of MNP on MRC-5 cells taking into account
ally associated with the inflammatory process. Significant eleva-
that human environmental, occupational and intentional expo-
tions of NO and PGE2 concentrations in the culture medium of
sures to these often occurs (i.e. due to the fact that more and more
MRC-5 cells were detected only after 72 h of MNP exposure. At this
iron oxide nanoparticles are manufactured for nanomedicine).
time point, the NO concentration and the PGE2 level were
Previous studies showed that the MNP cytotoxicity was depen-
increased by 81% (Fig. 10A) and by 82% (Fig. 10B), respectively.
dent on time and dose-exposure as well as on the characteristics of
the cell lines themselves. For example, while MNP were more toxic
3.10. Apoptotic/necrotic cells and DNA damage assessment for human umbilical vein endothelial cells (Wu et al., 2010) and
aortic endothelial cells (Ge et al., 2013), a lower toxicity was
Annexin V/PI staining revealed that magnetite nanoparticles detected in macrophage and liver cells (Brown et al., 2014;
exposure did not induce apoptosis or necrosis in MRC-5 cells Priprem et al., 2010). In addition, moderate cytotoxic effects were

Fig. 8. Hsp60 protein and gene expression in MRC-5 cells after treatment with 12.5 lg/mL MNP (at 24, 48 and 72 h). (A) Hsp60 immunoblot analysis and (B) quantitative
analyses. The control is represented as 100% and protein expression was normalized to each time control (normalized to b-actin). (C) The relative levels of Hsp60 mRNA
normalized to GAPDH mRNA level. Results are expressed as means (n = 5) ± SD. ⁄p < 0.05 vs. controls. ⁄⁄P < 0.05 vs. controls.
M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502 1499

Fig. 10. NO (A) and PGE2 (B) levels in the culture medium of MRC-5 cells after exposure to magnetite nanoparticles. Values are calculated as means ± SD (n = 5) and expressed
as % from controls. ⁄p < 0.05 vs. controls.

noticed in A549 lung epithelial cells (Könczöl et al., 2011) and in the interaction between magnetite and cellular organelles, such
Vero kidney epithelial cells (Szalay et al., 2012) following MNP as mitochondria and nuclei, the activation of membrane NADPH
exposure. oxidase (Nordberg et al., 2007; Könczöl et al., 2011), or the disrup-
In our experiments, the cytotoxicity of MNP on MRC-5 cells was tion of the well-structured electronic configuration of the
found to be moderate, for all doses after 72 h (Fig. 4). Due to the nano-sized material surface which creates reactive electron donor
fact that lung cells are exposed to environmental (Buzea et al., or acceptor sites, leading to the formation of superoxide radicals
2007) and intravenously injected MNP (Tseng et al., 2012), our (Shubayev et al., 2009).
research could have important implications for human health. The low cytotoxic effect of MNP (25% decrease in cells viability
The biological relevance in the human context of the 12.5 lg/mL after 72 h) coupled to the small amounts of iron ions released (10%
dose used in this study cannot be determined readily. According of iron contents in magnetite) resulted in a relatively moderate
to World Health Organization (WHO, 2005), even if the potential amount of ROS being produced. As a multistage process,
detrimental effect of ultrafine particles (of which the majority nanoparticle-induced ROS activation of the defense antioxidant
are in the nanosize range) on human cells is accepted, the epidemi- response occurred. The cells have developed several cellular
ological data are not sufficient to conclude on the expo- defense pathways, which under normal metabolic conditions reg-
sure/response relationship of these. ulate the level of ROS and protect against the deleterious effects
The most important iron elevation was recorded at 72 h post of ROS. This defense system includes both antioxidant enzymes,
MNP exposure, when the level of iron was approximately 12 pg/- such as SOD, CAT, GPX, GST and GR, and low-molecular-weight
cell (Fig. 5). In similar studies, the intracellular iron concentrations free-radical scavengers, such as the tripeptide GSH. In order to elu-
reported vary greatly, ranging from a few pg of iron per cell (Zhang cidate the adaptation of MRC-5 to ROS production induced by
et al., 2007) up to 62.5 pg Fe/cell in the MCF-7 human breast carci- MNP, we quantified the activities of these antioxidant enzymes.
noma cell line (Kumar et al., 2012). While the amount of cellular Our results showed a similar response of SOD and CAT activities
iron is influenced by MNP concentration and exposure time, the in the cells exposed to 12.5 lg/mL MNP (Fig. 7A and B). SOD cat-
results appear to strongly depend on the type of cell line used alyzes the conversion of free-radical superoxide to hydrogen per-
(Liu and Wang, 2013). An iron content of 120 pg/cell was reported oxide. The GPX enzyme worked in tandem with CAT to scavenge
for flasks treated with a dose of 12.5 lg MNP/10 mL/75 cm2 flask; the excess of hydrogen peroxide: CAT converted the hydrogen per-
this suggests that only 10% of the iron of magnetite entered the oxide into molecular oxygen and water, while GPX used GSH as an
cells. Similarly, an uptake efficiency of 9.6% was reported in human electron donor and catalyzed the biotransformation of various
umbilical vein endothelial cell (HUVEC) treated with microparti- organic and inorganic peroxides. In our opinion, the upregulation
cles of iron oxide (van Tiel et al., 2010). The mechanisms by which in GPX activity, noticed only at 72 h post-MNP administration
MNP enters the cells are not fully understood. Some studies proved (Fig. 7C), explained the reduction in ROS level from 48 to 72 h.
that, upon their intracellular internalization via endocytosis, MNP Similarly, the inactivation of MNP induced-ROS production by
are clustered within lysosomes where, presumably, they are antioxidant enzymes has also been reported for A549 cells
degraded into iron ions due to the action of an array of hydrolyzing (Limbach et al., 2007).
enzymes at low pH in accordance with endogenous iron metabo- In our studies, the GST activities, which can detoxify the
lism pathways (Tomitaka et al., 2009; Kai et al., 2011). Once pre- endogenous compounds including peroxidized lipids, were
sent in the cell, iron is ultimately sequestrated within ferritin, unchanged (Fig. 7D); this suggests this enzymatic defense against
which limits its capacity to generate free radicals (Ghio et al., ROS-induced by MNP degradation was not activated. The decrease
2006). However, when ferritin is overloaded, free iron ions could in GR activity after 72 h of MNP exposure with about 33% (Fig. 7E)
be released in the cell. indicated an impaired reduction of GSSG to GSH, and could at least
Iron oxide nanoparticles are believed to induce redox cycling partly explained the observed depletion of about 60% in GSH con-
via the Fenton reaction, the hydroxyl radical formed being the centration (Table 2). On the other hand, ferric ions of magnetite,
most prevalent source of ROS in biological systems (Nel et al., being electrophilic, can deplete intracellular GSH by interacting
2006). The ferric ions from MNP could interact with hydrogen per- directly with sulfhydryl groups (Brodie et al., 1982) or by inhibiting
oxide physiologically formed by different types of several cellular activities including the cysteine uptake (Bannai,
enzyme-catalyzed reactions, generating another ROS type, 1984). Furthermore, the long exposure to ferric nanoparticles could
hydroperoxyl radical (Fe3+ + H2O2 ? Fe2+ + H+ + HOO) (Hurd and induce the loss of cellular GSH by glutathionylation of several pro-
Murphy, 2009). While the ROS generation leading to oxidative teins (Hare and Stamler, 2005).
stress was associated with the MNP nanotoxicity, the mechanisms Iron accumulation as a result of exposure to iron oxide nanopar-
by which these nanoparticles generate ROS are still largely ticles could potentially result in deleterious cellular consequences
unknown. Several possible mechanisms leading to the formation eventually leading to cell death (Singh et al., 2010). There is evi-
of different species of ROS have been suggested; these include dence that increased levels of ROS can induce lipid peroxidation,
1500 M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502

Fig. 11. Apoptosis detection (A): Phase-contrast (left), AnnexinV-FITC (center) and PI (right) for MRC-5 cells before and after 72 h MNP treatment. The images acquired by
fluorescence microscopy represent MRC-5 cells after 72 h treated with 10% DMSO (positive control), untreated cells and MNP-treated cells. The scale bar is the same for all
images and corresponds to 50 lm. DNA integrity (B): DNA samples were electrophoretically separated on an ethidium bromide stained-agarose gel. No DNA strand breakage
was detected.

DNA fragmentation and protein oxidation. The significant increase mouse heart (Para, 2011). The induction of Hsp60 has also been
of MDA concentration, by 116% and 170% after 48 and 72 h, respec- detected in many pulmonary cell types such as the bronchial
tively, suggests that the antioxidative system adaptation was not epithelium of patients with asthma (Fajac et al., 1997) or the
sufficient to prevent the damage of membrane lipids. human pulmonary microvascular endothelial cells following heavy
Hydroperoxyl and hydroxyl radicals are able to abstract an hydro- metal exposure (Wagner et al., 1999).
gen atom from a methylene group adjacent to double bonds of In complex with Hsp10, Hsp60 provides a suitable environment
polyunsaturated fatty acids forming carbon centered radicals that for the unfolded or misfolded protein intermediates to return to
react with molecular oxygen to form lipid peroxides (Antunes their native conformation. Unlike most Hsps with obvious
et al., 1996). An elevation in MDA level was also observed in pro-survival functions, Hsp60 appears to have both pro-survival
MRC-5 cells exposed to Fe2O3 nanoparticles (Radu et al., 2010) and pro-death functions (Chandra et al., 2007). Some studies have
and in LE rat lung epithelial cells treated with magnetite shown that Hsp60 protects cells from stress-induced death
(Ramesh et al., 2012); by contrast, a statistically insignificant through activation of extracellular signal-regulated kinase (ERK)
change in the lipid peroxidation level was observed in a A549 and inhibition of caspase-3 (Samali et al., 1999). Conversely,
human lung epithelial line treated with hematite nano-sized parti- Hsp60 has a pro-apoptotic effect consisting of the release of cyto-
cles (Guo et al., 2009). chrome c and caspase-3 activation in Jurkat cells (Samali et al.,
Cells from all organisms respond to a variety of stress condition 1999).
by a rapid synthesis of a highly conserved family of proteins, ter- Up-regulation of Hsp60 in several cancers correlates with dif-
med heat shock proteins (Hsp). Inducible forms of Hsp were orig- ferent evolution prognosis for the patients (Cappello et al., 2008).
inally described following heat stress (Garrido et al., 2001) Higher levels of Hsp60 protein help cells to be more resistant to
however, a variety of cellular stressors will lead to their induction ROS attack (Cabiscol et al., 2002), which emphasizes the potential
as part of an orchestrated stress response. Hsp can be activated by pro-survival role of this protein. By contrast, loss of Hsp60
oxidative stress (Fulda et al., 2010), conferring a transient protec- expression is associated with an increased risk of developing infil-
tion. In response to stress factors, Hsp gene expression is modu- trating recurrent bladder cancer (Lebret et al., 2003). In addition,
lated firstly by heat shock factor protein 1 (HSF-1). The an inhibition of Hsp60 under increased stress condition can sup-
mechanism of activation consists of HSF1 oligomerization and press autophagy (Kim et al., 2009). In our opinion, the
nuclear translocation, followed by enhanced DNA binding on the up-regulation of Hsp60 induced cytoprotection against MNP cell
Hsp gene promoters. stresses in the MRC-5 line, probably due to cell death
Hsp60 protein levels were significantly increased in MRC-5 cells suppression.
starting with 48 h post-MNP administration (Fig. 8A and B); by Significant elevations of NO and PGE2 concentrations in the cul-
contrast, Hsp27, Hsp70 and Hsp90 expression was not changed ture medium of MRC-5 cells were detected only after 72 h of MNP
(data not shown). The experiments showed a similar profile in both exposure. At this time, the NO concentration increased by 81%
Hsp60 mRNA and protein analysis, attesting that induced gene (Fig. 10A) and the PGE2 level was augmented by 82% (Fig. 10B).
expression accounted for the up-regulation observed in protein Hsp60 is also a ligand for Toll-like receptor 4 (TLR4) (Guo and
expression. The increase of Hsp60 gene expression in pulmonary Friedman, 2010), which is present in lung tissue (Yang et al.,
fibroblasts could explain the resistance to induced oxidative stress 2013). The activated TLR4 could trigger the canonical IKKb/NF-jB
after MNP exposure. The highest level of ROS was recorded at 48 h signaling pathway and induce the up-regulation of inducible nitric
after MNP administration, when Hsp60 gene expression began to oxide synthase (iNOS) (Newton and Dixit, 2012). Previous studies
be up-regulated. Despite the downward trend in ROS concentra- have suggested that endogenous and exogenous NO play a critical
tion recorded after this time, the MDA level continuously increased role in the release of PGE2, by direct activation of cyclooxygenase
up to 72 h (Table 2); this suggests that Hsp60 induction was not enzymes (Salvemini et al., 1993). Consequently, the increase in
sufficient to completely counteract the oxidative stress produced Hsp60 could have induced NO synthesis and indirect release of
by MNP in MRC-5 cells. PGE2, an abundant eicosanoid product with crucial role in inflam-
Few studies have been conducted on HSP expression in cells mation (Wilborn et al., 1995). Also, PGE2 can promote cell survival
treated with NPs. An in vivo study suggested that CeO2 nanoparti- by a direct inhibition of caspase-3 via S-nitrosylation (Rossig et al.,
cles inhalation caused elevation of Hsp60 and Hsp27 levels in the 1999).
M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502 1501

MNP treated MRC-5 cells, probably subsequent to caspase-3 Acknowledgements


inhibition, did not present DNA fragmentation. Similar results were
observed in the A549 human lung epithelial cells exposed to This work was supported by the strategic grant
20–60 nm MNP at a concentration of 10 lg/mL (Könczöl et al., POSDRU/159/1.5/S/133391, Project ‘‘Doctoral and Post-doctoral
2011) but the treatment with higher doses generated DNA oxida- programs of excellence for highly qualified human resources train-
tive damage in the same cells (Karlsson et al., 2009). ing for research in the field of Life sciences, Environment and Earth
Activated TLR 4 generates caspase-1 activation, a Science’’ co-financed by the European Social Found within the
pro-inflammatory cysteine protease, which is responsible for pro- Sectorial Operational Program Human Resources Development
teolytic processing and activation of the proform of IL-1b (Staal 2007–2013 and by the PNII-IDEI 340/2007 project financed by
et al., 2011). National Council of Scientific Research (CNCSIS), Romania.
In our study, MNP exposure caused the up-regulation and activa-
tion of caspase-1 starting with 24 h of exposure, suggesting an
inflammatory response. The activity of caspase-1 in pulmonary References
fibroblasts remained elevated up to 72 h of MNP exposure.
Recently, a new type of programmed cell death named pyroptosis, Aebi, H., 1984. Catalase in vitro. In: Bergmayer, H.U. (Ed.), Methods of enzymatic
or caspase-1-dependent cell death, was discovered (Labbé and analysis. FRG, Weinheim, pp. 673–684.
Ankamwar, B., Lai, T.C., Huang, J.H., Liu, R.S., Hsiao, M., Chen, C.H., Hwu, Y.K.,
Saleh, 2011). In this pathway, the autocatalysis of inactive 2010. Biocompatibility of Fe3O4 nanoparticles evaluated by in vitro
pro-caspase-1 to active caspase-1 is triggered by the formation of cytotoxicity assays using normal, glia and breast cancer cells.
a cytosolic complex named the ‘‘inflammasome’’ (Lamkanfi et al., Nanotechnology 21, 75102.
Antunes, F., Salvador, A., Marinho, H.S., Alves, R., Pinto, R.E., 1996. Lipid peroxidation
2007). Although in our experiment caspase-1 activation occurred, in mitochondrial inner membranes. I. An integrative kinetic model. Free Rad.
no form of cell death was observed during the analyzed time inter- Biol. Med. 21, 917–943.
vals (Fig. 9). It is possible that while caspase-1 activation in immune Bannai, S., 1984. Transport of cystine and cysteine in mammalian cells. Biochim.
Biophys. Acta 779, 289–306.
cells triggers pyroptosis and influences the development of adaptive Bergsbaken, T., Fink, S.L., Cookson, B.T., 2009. Pyroptosis: host cell death and
immune responses (Bergsbaken et al., 2009), epithelial cells activate inflammation. Nat. Rev. Microbiol. 7, 99–109.
caspase-1 to prevent cell death (Gurcel et al., 2006). This enzyme Beutler, E., 1971. Red cell metabolism. In: Grune, Stratton (Eds.), A manual of
biochemical methods. New York, pp. 71–73.
supports survival and confers resistance to mammalian cells against Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of
pathogenic bacteria that express a pore-forming toxin that, in turn, microgram quantities of protein utilizing the principle of proteindye binding.
can create holes in the plasma membrane of the host cell (Sollberger Anal. Biochem. 72, 248–254.
Brodie, A.E., Potter, J., Reed, D.J., 1982. Unique characteristics of rat spleen
et al., 2014). The repair of toxin-induced damage to the plasma
lymphocyte, L1210 lymphoma and HeLa cells in glutathione biosynthesis
membrane requires the activation of the lipid biogenesis pathways, from sulfur-containing amino acids. Eur. J. Biochem. 123, 159–164.
controlled by two transcription factors known as sterol regulatory Brown, D.M., Johnston, H., Gubbins, E., Stone, V.-J., 2014. Serum enhanced cytokine
element-binding protein 1 (SREBP1) and SREBP2 (Lamkanfi, 2011). responses of macrophages to silica and iron oxide particles and nanomaterials:
a comparison of serum to lung lining fluid and albumin dispersions. J. Appl.
The molecular mechanism by which caspase-1 activates SREBPs is Toxicol. 34, 1177–1187.
completely unknown (Sollberger et al., 2014). In our case, MNP could Buzea, C., Pacheco, I.I., Robbie, K., 2007. Nanomaterials and nanoparticles: sources
generate the same events due to the increased lipid peroxidation of and toxicity. Biointerphases 2, MR17–MR172.
Cabiscol, E., Bell, G., Tamarit, J., Echave, P., Herrero, E., Ros, J., 2002. Mitochondrial
polyunsaturated acids from the plasma membrane. Hsp60, resistance to oxidative stress, and the labile iron pool are closely
connected in Saccharomyces cerevisiae. J. Biol. Chem. 277, 44531–44538.
Cappello, F., Conway de Macario, E., Marasà, L., Zummo, G., Macario, A.J.L., 2008.
5. Conclusions Hsp60 expression, new locations, functions and perspectives for cancer
diagnosis and therapy. Cancer Biol. Ther. 7, 801–809.
The MNP treatment of MRC-5 cells resulted in a time-dependent Chandra, D., Choy, G., Tang, D.G., 2007. Cytosolic accumulation of HSP60 during
apoptosis with or without apparent mitochondrial release: evidence that its
accumulation of Fe ions inside the cells that led to ROS generation. pro-apoptotic or pro-survival functions involve differential interactions with
The increased CAT and SOD activity, GSH depletion and MDA gener- caspase-3. J. Biol. Chem. 282, 1289–1301.
ation reflected the oxidative stress induced in MRC-5 cells. Based on Chowdhury, S.R., Yanful, E.K., 2010. Arsenic and chromium removal by mixed
magnetite–maghemite nanoparticles and the effect of phosphate on removal. J.
these findings, the increase in NO and PGE2 production coupled with Environ. Manage. 91, 2238–2247.
the surge of caspase-1 activity and Hsp60 expression could suggest Dinischiotu, A., Stanca, L., Gradinaru, D., Petrache, S.N., Radu, M., Serban, A.I., 2013.
that MRC-5 cells successfully developed cell protection mechanisms Lipid peroxidation due to in vitro and in vivo exposure of biological samples to
nanoparticles. Oxid. Stress Nanotechnol. 1028, 155–164.
and suppressed cell death. When cells exposure to MNP doses higher Fajac, I., Roisman, G.L., Lacronique, J., Polla, B.S., Dusser, D.J., 1997. Bronchial gamma
that 12.5 lg/mL occurs, an additional amount of ROS could be pro- delta T-lymphocytes and expression of heat shock proteins in mild asthma. Eur.
duced, and consequently, the processes mentioned above would Respir. J. 10, 633–638.
Figuerola, A., Di Corato, R., Manna, L., Pellegrino, T., 2010. From iron oxide
reach saturation. If these processes are saturated, most cellular com-
nanoparticles towards advanced iron-based inorganic materials designed for
ponents are likely to be targets of oxidative damage, such as lipid biomedical applications. Pharmacol. Res. 62, 126–143.
peroxidation, protein oxidation, GSH depletion and DNA single Fulda, S., Gorman, A.M., Hori, O., Samali, A., 2010. Cellular stress responses: cell
strand breaks, that are all initiated by ROS excess. Taken together, survival and cell death. Int. J. Cell Biol., 1–23, ID 21407
Garrido, C., Gurbuxani, S., Ravagnan, L., Kroemer, G., 2001. Heat shock proteins:
these events could lead to cellular dysfunction and ultimately cell endogenous modulators of apoptotic cell death. Biochem. Biophys. Res.
death. As a result, the need for a better understanding of this mech- Commun. 286, 433–442.
anism warrants further investigations. Ge, G., Wu, H., Xiong, F., Zhang, Y., Guo, Z., Bian, Z., Xu, J., Gu, C., Gu, N., Chen, X.,
Yang, D., 2013. The cytotoxicity evaluation of magnetic iron oxide nanoparticles
on human aortic endothelial cells. Nanoscale Res. Lett. 8, 215–224.
Conflict of Interest Ghio, J.A., Turi, J.L., Yang, F., Garrick, L.M., Garrick, M.D., 2006. Iron homeostasis in
the lung. Biol. Res. 39, 67–77.
Goldberg, D.M., Spooner, R.J., 1983. Glutathione reductase. In: Bergmayer, H.U. (Ed.),
The authors declare that there are no conflicts of interest. Methods of Enzymatic Analysis. Verlag Chemie, Dearfield Beach, pp. 258–265.
Green, L.C., Wagner, D.A., Glogowski, J., Skipper, P.L., Wishnok, J.S., Tannenbaum,
S.R., 1982. Analysis of nitrate, nitrite, and [15N] nitrate in biological samples.
Transparency Document Anal. Biochem. 126, 131–138.
Grobéty, B., Gieré, R., Dietze, V., Stille, P., 2010. Airborne particles in the urban
environment. Elements 6, 229–234.
The Transparency document associated with this article can be Guo, J.F., Friedman, S.L., 2010. Toll-like receptor 4 signalling in liver injury and
found in the online version. hepaticfibrogenesis. Fibrogen. Tissue Repair 3, 21.
1502 M. Radu et al. / Toxicology in Vitro 29 (2015) 1492–1502

Guo, B., Zebda, R., Drake, S.J., Sayes, C.M., 2009. Synergistic effect of co-exposure of Park, E.-J., Kim, H., Kim, Y., Yi, J., Choi, K., Park, K., 2010. Inflammatory responses may
carbon black and Fe2O3 nanoparticles on oxidative stress in cultured lung be induced by a single intratracheal instillation of iron nanoparticles in mice.
epithelial cells. Part. Fibre. Toxicol. 6, 4. Toxicology 275, 65–71.
Gurcel, L., Abrami, L., Girardin, S., Tscopp, J., van der Goot, F.G., 2006. Caspase-1 Pauluhn, J., 2012. Subchronic inhalation toxicity of iron oxide (magnetite, Fe3O4) in
activation of lipid metabolic pathways in response to bacterial pore-forming rats: pulmonary toxicity is determined by the particle kinetics typical of poorly
toxins promotes cell survival. Cell 126, 1135–1145. soluble particles. J. Appl. Toxicol. 32, 488–504.
Habig, W.H., Pabst, M.J., Jakoby, W.B., 1974. Glutathione-S-transferase. The first Pisanic, T.R., Jin, S., Shubayev, V.I., 2009. Iron oxide magnetic nanoparticle
enzymatic step in mercapturic acid formation. J. Biol. Chem. 249, 7130–7139. nanotoxicity: incidence and mechanisms. In: Sahu, S., Casciano, D. (Eds.),
Halliwell, B., Gutteridge, J.M.C., 2007. Free Radicals in Biology and Medicine. Oxford Nanotoxicity: From in vivo and in vitro Models to Health Risks. John Wiley &
University Press, New York. Sons Ltd, pp. 397–425.
Hare, J.M., Stamler, J.S., 2005. NO/redox disequilibrium in the failing heart and Price, O.T., Asgharian, B., Miller, F.J., Cassee, F.R., Winter-Sorkina, R., 2002. Multiple
cardiovascular system. J. Clin. Invest. 115, 509–517. Path Particle Dosimetry model: A Model for Human and Rat Airway Particle
Hurd, T.R., Murphy, M.P., 2009. Biological systems relevant for redox signalling and Dosimetry, (MPPD V1.0), RIVM Report, 650010030.
control. In: Jacob, C., Winyard, P.G. (Eds.), Redox Signalling and Regulation in Priprem, A., Mahakunakorn, P., Thomas, C., Thomas, I., 2010. Cytotoxicity studies of
Biology and Medicine. Wiley-VCH VerlagGmbH&Co. KGaA, Weinheim, pp. 13– superparamagnetic iron oxide nanoparticles in macrophage and liver cells. Am.
45. J. Nanotechnol. 1, 78–85.
Ito, A., Fujioka, M., Yoshida, T., Wakamatsu, K., Ito, S., Yamashita, T., Jimbow, K., Radu, M., Munteanu, M.C., Petrache, S., Serban, A.I., Dinu, D., Hermenean, A., Sima,
Honda, H., 2007. 4-S-Cysteaminylphenol-ioded magnetite cationic liposomes C., Dinischiotu, A., 2010. Depletion of intracellular glutathione and increased
for combination therapy of hypertermia with chemotherapy against malignant lipid peroxidation mediate cytotoxicity of hematite nanoparticles in MRC-5
melanoma. Cancer Sci. 98 (3), 424–430. cells. Acta. Biochim. Polon. 57, 355–360.
Kai, W., Xiaojun, X., Ximing, P., Zhenqing, H., Qiqing, Z., 2011. Cytotoxic effects and Ramesh, V., Ravichandran, P., Copeland, C.L., Gopikrishnan, R., Biradar, S., Goornavar,
the mechanism of three types of magnetic nanoparticles on human hepatoma V., Ramesh, G.T., Hall, J.C., 2012. Magnetite induces oxidative stress and
BEL-7402 cells. Nanoscale Res. Lett. 6, 480. apoptosis in lung epithelial cells. Mol. Cell. Biochem. 363, 225–234.
Karlsson, H.L., Holgersson, A., Moller, L., 2008. Mechanisms related to the Rossig, L., Fichtlscherer, B., Breitschopf, K., Haendeler, J., Zeiher, A.M., Mulsch, A.,
genotoxicity of particles in the subway and from other sources. Chem. Res. Dimmeler, S., 1999. Nitric oxide inhibits caspase-3 by S-nitrosylation in vivo. J.
Toxicol. 21, 726–731. Biol. Chem. 274, 6823–6826.
Karlsson, H.L., Gustafsson, J., Cronholm, P., Moller, L., 2009. Size-dependent toxicity Salvemini, D., Misko, T.P., Masferrer, J.L., Seibert, K., Currie, M.G., Needleman, P.,
of metal oxide particles—a comparison between nano- and micrometer size. 1993. Nitric oxide activates cyclooxygenase enzymes. Proc. Natl. Acad. Sci. 90,
Toxicol. Lett. 188, 112–118. 7240–7244.
Kehrer, J.P., 1993. Free radicals as mediators in tissue injury and disease. Crit. Rev. Samali, A., Cai, J., Zhivotovsky, B., Jones, D.P., Orrenius, S., 1999. Presence of a pre-
Toxicol. 23, 21–48. apoptotic complex of pro-caspase-3, Hsp60 and Hsp10 in the mitochondrial
Kim, H., Choi, J., Ryu, J., Park, S.G., Cho, S., Park, B.C., Lee do, H., 2009. Activation of fraction of Jurkat cells. The EMBO J. 18, 2040–2048.
autophagy during glutamate-induced HT22 cell death. Biochem. Biophys. Res. Santoro, M.G., 2000. Heat shock factors and the control of the stress response.
Commun. 388, 339–344. Biochem. Pharmacol. 59, 55–63.
Könczöl, M., Ebeling, S., Goldenberg, E., Treude, F., Gminski, R., Gieré, R., Grobérty, B., Shaw, S.Y., Westly, E.C., Pittet, M.J., Subramanian, A., Schreiber, S.L., Weissleder, R.,
Rothen-Rutishauser, B., Merfort, I., Mersch-Sunderman, V., 2011. Cytotoxicity 2008. Perturbational profiling of nanomaterial biologic activity. Proc. Natl. Acad.
and genotoxicity of size-fractionated iron oxide (magnetite) in A549 human Sci. 105, 7387–7392.
lung epithelial cells: role of ROS, JNK, and NF-B. Chem. Res. Toxicol. 24, 1460– Shubayev, V., Pisanic, T.R., Jin, S., 2009. Magnetic nanoparticles for theragnostics.
1475. Adv. Drug Deliv. Rev. 61, 467–477.
Kumar, M., Singh, G., Arora, V., Mewar, S., Sharma, U., Jagannathan, N.R., Sapra, S., Singh, N., Jenkins, G.J.S., Asadi, R., Doak, S.H., 2010. Potential toxicity of
Dinda, A.K., Kharbanda, S., Singh, H., 2012. Cellular interaction of folic acid superparamagnetic iron oxide nanoparticles (SPION). Nano Rev. 1, 5358.
conjugated superparamagnetic iron oxide nanoparticles and its use as contrast Sollberger, G., Strittmatter, G.E., Garstkiewicz, M., Sand, J., Beer, H.-D., 2014.
agent for targeted magnetic imaging of tumor cells. Int. J. Nanomed. 7, 3503– Caspase-1: the inflammasome and beyond. Innate Immun. 20, 115–125.
3516. Staal, J., Bekaert, T., Beyaert, R., 2011. Regulation of NF-jB signaling by caspases and
Labbé, K., Saleh, M., 2011. Pyroptosis: a caspase-1-dependent programmed cell MALT1 paracaspase. Cell Res. 21, 40–54.
death and a barrier to infection. In: Couillin, I., Pétrilli, V., Martinon, F. (Eds.), Szalay, B., Tátrai, E., Nyíro} , G., Vezér, T., Dura, D., 2012. Potential toxic effects of iron
The Inflammasomes, Progress in Inflammation Research. Springer, Bassel AG, oxide nanoparticles in in vivo and in vitro experiments. J. Appl. Toxicol. 32,
pp. 17–37. 446–453.
Lamkanfi, M., 2011. Emerging inflammasome effector mechanisms. Nat. Rev./ Tomitaka, A., Hirukawa, A., Yamada, T., Morishita, S., Takemura, Y., 2009.
Immunol. 11, 213–220. Biocompatibility of various ferrite nanoparticles evaluated by in vitro
Lamkanfi, M., Kanneganti, T.D., Franchi, L., Nunez, G., 2007. Caspase-1 cytotoxicity assays using HeLa cells. J. Magn. Magn. Mater. 321, 482–1484.
inflammasomes in infection and inflammation. J. Leukoc. Biol. 82, 220–225. Tseng, W.-K., Chieh, J.-J., Yang, C.-K., Chiang, C.-K., Chero, Y.-L., Shieh, Y.Y., Horng, H.-
Lebret, T., Watson, R.W., Molinie, V., O’Neill, A., Gabriel, C., Fitzpatrick, J.M., Botto, H., E., Yang, H.-C., Wu, C.-C., 2012. A noninvasive method to determine the fate of
2003. Heat shock proteins HSP27, HSP60, HSP70, and HSP90: expression in Fe3O4 nanoparticles following intravenous injection using scanning SQUID
bladder carcinoma. Cancer 98, 970–977. biosusceptometry. PLoS ONE 7 (11), e48510.
Li, J.J., Hartono, D., Ong, C.N., Bay, B.H., Yung, L.Y., 2010. Autophagy and oxidative van Tiel, S.T., Wielopolski, P.A., Houston, G.C., Krestin, G.P., Bernsen, M.R., 2010.
stress associated with gold nanoparticles. Biomaterials 31, 5996–6003. Variations in labeling protocol influence incorporation, distribution and
Limbach, L.K., Wick, P., Manser, P., Grass, R.N., Bruinink, A., Stark, W.J., 2007. retention of iron oxide nanoparticles into human umbilical vein endothelial
Exposure of engineered nanoparticles to human lung epithelial cells: influence cells. Contrast Media Mol. Imaging 5, 247–257.
of chemical composition and catalytic activity on oxidative stress. Environ. Sci. Wagner, M., Hermanns, I., Bittinger, F., Kirkpatrick, C.J., 1999. Induction of stress
Technol. 41, 4158–4163. proteins in human endothelial cells by heavy metal ions and heat shock. Am. J.
Lin, B.L., Shen, X.D., Cui, S., 2007. Application of nanosized Fe3O4 in anticancer drug Physiol. 277, L1026–L1033.
carriers with target-orientation and sustained-release properties. Biomed. Wan, C.P., Myung, E., Lau, B.H.S., 1993. An automated microfluorometric assay for
Mater. 2, 132–134. monitoring oxidative burst activity of phagocytes. J. Immunol. Meth. 159, 131–
Liu, Y., Wang, J., 2013. Effects of DMSA-coated Fe3O4 nanoparticles on the 138.
transcription of genes related to iron and osmosis homeostasis. Toxicol. Sci. WHO. Word Health Organisation, 2005. Guidelines for air quality.
131, 521–536. Wilborn, J., Crofford, L.J., Burdick, M.D., Kunkel, S.L., Strieter, R.M., Peters-Golden, M.,
Maity, D., Chandrasekharan, P., Yang, C.T., Chuang, K.H., Shuter, B., Xue, J.M., Ding, J., 1995. Cultured lung fibroblasts isolated from patients with idiopathic
Feng, S.S., 2010. Facile synthesis of water-stable magnetite nanoparticles for pulmonary fibrosis have a diminished capacity to synthesize prostaglandin E2
clinical MRI and magnetic hyperthermia applications. Nanomedicine UK.5, and to express cyclooxygenase-2. J. Clin. Invest. 95, 1861–1868.
1571–1584. Wu, X., Tan, Y., Mao, H., Zhang, M., 2010. Toxic effects of iron oxide nanoparticles on
Mosmann, J., 1983. Rapid colorimetric assay for cellular growth and survived. J. human umbilical vein endothelial cells. Int. J. Nanomed. 5, 385–399.
Immunol. Meth. 65, 55–63. Yang, Z., Deng, Y., Su, D., Tian, J., Gao, Y., He, Z., Wang, X., 2013. TLR 4 as receptor for
Nel, A., Xia, T., Mädler, L., Li, N., 2006. Toxic potential of materials at the nanolevel. HMGB1-mediated acute lung injury after liver ischemia/reperfusion injury. Lab.
Science 311, 622–627. Invest. 93, 792–800.
Newton, K., Dixit, V.M., 2012. Signalling in innate immunity and inflammation. Cold Zhang, C., Wangler, B., Morgenstern, B., Zentgraf, H., Eisenhut, M., Untenecker, H.,
Spring Harb. Perspect. Biol. 4, pii: a006049. Kruger, R., Huss, R., Seliger, C., Semmler, W., Kiessling, F., 2007. Silica- and
Nordberg, G.F., Fowler, B.A., Nordberg, M., Friberg, L.T., 2007. Handbook on the alkoxysilane-coated ultra small superparamagnetic iron oxide particles: a
Toxicology of Metals, third ed. Academic press, New York. promising tool to label cells for magnetic resonance imaging. Langmuir 23,
Paoletti, F., Aldinucci, D., Mocali, A., Caparrini, A., 1986. A sensitive 1427–1431.
spectrophotometric method for the determination of superoxide dismutase Zhu, X.M., Wang, Y.X., Leung, K.C., Lee, S.F., Zhao, F., Wang, D.W., Lai, J.M., Wan, C.,
activity in tissue extracts. Anal. Biochem. 154, 538–541. Cheng, C.H., Ahuja, A.T., 2012. Enhanced cellular uptake of aminosilane-coated
Para, R., 2011. Evaluation of Toxicological effects on Intra Tracheal Instilled CeO2 superparamagnetic iron oxide nanoparticles in mammalian cell lines. Int. J.
Nanoparticles on the Heart of Male Spargue–Dawley Rats. Theses, Dissertations Nanomed. 7, 953–964.
and Capstones. Paper 48.

You might also like