+ All Categories
Home > Documents > Cellular internalization and stress response of ingested amorphous silica nanoparticles in the...

Cellular internalization and stress response of ingested amorphous silica nanoparticles in the...

Date post: 23-Nov-2016
Category:
Upload: ashutosh-pandey
View: 212 times
Download: 0 times
Share this document with a friend
11
Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster Ashutosh Pandey a , Swati Chandra a , Lalit Kumar Singh Chauhan b , Gopeshwar Narayan c , Debapratim Kar Chowdhuri a, a Embryotoxicology Section, CSIR-Indian Institute of Toxicology Research, Lucknow 226001, Uttar Pradesh, India b Electron Microscopy Section, CSIR-Indian Institute of Toxicology Research, Lucknow 226001, Uttar Pradesh, India c Department of Molecular and Human Genetics, Banaras Hindu University, Varanasi 221005, Uttar Pradesh, India abstract article info Article history: Received 23 August 2012 Received in revised form 2 October 2012 Accepted 3 October 2012 Available online 6 October 2012 Keywords: Amorphous silica nanoparticles Heat shock proteins Oxidative stress Membrane destabilization Caspase-3 Cellular uptake Background: Amorphous silica nanoparticles (aSNPs) are used for various applications including food indus- try. However, limited in vivo studies are available on absorption/internalization of ingested aSNPs in the mid- gut cells of an organism. The study aims to examine cellular uptake of aSNPs (b 30 nm) in the midgut of Drosophila melanogaster (Oregon R + ) owing to similarities between the midgut tissue of this organism and human and subsequently cellular stress response generated by these nanoparticles. Methods: Third instar larvae of D. melanogaster were exposed orally to 1100 μg/mL of aSNPs for 1236 h and oxidative stress (OS), heat shock genes (hsgs), membrane destabilization (Acridine orange/Ethidium Bro- mide staining), cellular internalization (TEM) and apoptosis endpoints. Results: A signicant increase was observed in OS endpoints in the midgut cells of exposed Drosophila in a concentration- and time-dependent manner. Signicantly increased expression of hsp70 and hsp22 along with caspases activation, membrane destabilization and mitochondrial membrane potential loss was also ob- served. TEM analysis showed aSNPs-uptake in the midgut cells of exposed Drosophila via endocytic vesicles and by direct membrane penetration. Conclusion: aSNPs after their internalization in the midgut cells of exposed Drosophila larvae show membrane destabilization along with increased cellular stress and cell death. General signicance: Ingested aSNPs show adverse effects on the cells of GI tract of the exposed organism thus their industrial use as a food-additive may raise concern to human health. © 2012 Elsevier B.V. All rights reserved. 1. Introduction With an increase in the use of nanomaterials (NMs) in industrial and biomedical applications in recent years, concern has been raised on entry of the NMs in the environment and subsequently, their possible effects on human health. The unique physico-chemical properties of NMs such as particle size, agglomeration state, size distribution, shape, crystal structure, chemical composition, surface chemistry, surface charge and porosity supplemented by altered dimensional congura- tion and high surface area to volume ratio may account for their hazard- ous effects on biological systems [1,2]. Engineered nanoparticles (ENPs) comprise wide range of NMs which are synthesized by molecular level engineering to acquire unique properties for their vast applications. Among the ENPs, silica nanoparticles are one of the most important and most utilized NMs. Amorphous silica nanoparticles (aSNPs), characterized by their non-crystalline structure are produced in different sizes along with dif- ferent surface modications with relative ease. This has led to the ex- tensive use of aSNPs for diverse applications that include agriculture, food and animal-feed additives [3]. aSNPs can enter into an organism including human through all pos- sible routes [2]. Since aSNPs are used in food products following their approval by FDA as a food additive [4], concern has been raised on their exposure to an organism through ingestion. Earlier, different ex- posure routes such as intra-tracheal instillation [5], inhalation [6] or in- travenous injection [7] were used for in vivo studies on aSNPs in rats and mice. Limited studies on aSNPs through oral route are available [8,9], Biochimica et Biophysica Acta 1830 (2013) 22562266 Abbreviations: aSNPs, Amorphous silica nanoparticles; NMs, Nanomaterials; NPs, Nanoparticles; ENPs, Engineered nanoparticles; GI, Gastrointestinal; Hsps, Heat shock proteins; Hsgs, Heat shock genes; Sps, Stress proteins; sHsps, Small heat shock proteins; OS, Oxidative stress; ROS, Reactive oxygen species; SOD, Superoxide dismutase; CAT, Catalase; LPO, Lipid peroxidation; GSH, Glutathione; GST, Glutathione S-transferase; PC, Protein content; DNPH, Di-nitro phenyl hydrazine; PBS, Phosphate buffered saline; BSA, Bovine serum albumin; TEM, Transmission electron microscopy; DLS, Dynamic light scattering; AO/EB, Acridine orange/Ethidium Bromide; MMP, Mitochondrial membrane potential Corresponding author at: Embryotoxicology Section, CSIR-Indian Institute of Toxicology Research, Mahatma Gandhi Marg, Lucknow 226001, Uttar Pradesh, India. Tel.: +91 522 2963825 (Lab: direct), +91 522 2620107x218 (Lab: PABX); fax: +91 522 2628227. E-mail addresses: [email protected], [email protected] (D.K. Chowdhuri). 0304-4165/$ see front matter © 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.bbagen.2012.10.001 Contents lists available at SciVerse ScienceDirect Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbagen
Transcript
Page 1: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Biochimica et Biophysica Acta 1830 (2013) 2256–2266

Contents lists available at SciVerse ScienceDirect

Biochimica et Biophysica Acta

j ourna l homepage: www.e lsev ie r .com/ locate /bbagen

Cellular internalization and stress response of ingested amorphous silicananoparticles in the midgut of Drosophila melanogaster

Ashutosh Pandey a, Swati Chandra a, Lalit Kumar Singh Chauhan b,Gopeshwar Narayan c, Debapratim Kar Chowdhuri a,⁎a Embryotoxicology Section, CSIR-Indian Institute of Toxicology Research, Lucknow 226001, Uttar Pradesh, Indiab Electron Microscopy Section, CSIR-Indian Institute of Toxicology Research, Lucknow 226001, Uttar Pradesh, Indiac Department of Molecular and Human Genetics, Banaras Hindu University, Varanasi 221005, Uttar Pradesh, India

Abbreviations: aSNPs, Amorphous silica nanoparticNanoparticles; ENPs, Engineered nanoparticles; GI, Gasproteins; Hsgs, Heat shock genes; Sps, Stress proteins; sHOS, Oxidative stress; ROS, Reactive oxygen species; SOCatalase; LPO, Lipid peroxidation; GSH, Glutathione; GSTProtein content; DNPH, Di-nitro phenyl hydrazine; PBS, PBovine serum albumin; TEM, Transmission electron mscattering; AO/EB, Acridine orange/Ethidium Bromide; Mpotential⁎ Corresponding author at: Embryotoxicology Section, CS

Research, Mahatma Gandhi Marg, Lucknow 226001, Utta2963825 (Lab: direct), +91 522 2620107x218 (Lab: PABX)

E-mail addresses: [email protected], dkarchow(D.K. Chowdhuri).

0304-4165/$ – see front matter © 2012 Elsevier B.V. Alhttp://dx.doi.org/10.1016/j.bbagen.2012.10.001

a b s t r a c t

a r t i c l e i n f o

Article history:

Received 23 August 2012Received in revised form 2 October 2012Accepted 3 October 2012Available online 6 October 2012

Keywords:Amorphous silica nanoparticlesHeat shock proteinsOxidative stressMembrane destabilizationCaspase-3Cellular uptake

Background: Amorphous silica nanoparticles (aSNPs) are used for various applications including food indus-try. However, limited in vivo studies are available on absorption/internalization of ingested aSNPs in the mid-gut cells of an organism. The study aims to examine cellular uptake of aSNPs (b30 nm) in the midgut ofDrosophila melanogaster (Oregon R+) owing to similarities between the midgut tissue of this organism andhuman and subsequently cellular stress response generated by these nanoparticles.Methods: Third instar larvae of D. melanogasterwere exposed orally to 1–100 μg/mL of aSNPs for 12–36 h andoxidative stress (OS), heat shock genes (hsgs), membrane destabilization (Acridine orange/Ethidium Bro-mide staining), cellular internalization (TEM) and apoptosis endpoints.Results: A significant increase was observed in OS endpoints in the midgut cells of exposed Drosophila in aconcentration- and time-dependent manner. Significantly increased expression of hsp70 and hsp22 alongwith caspases activation, membrane destabilization and mitochondrial membrane potential loss was also ob-served. TEM analysis showed aSNPs-uptake in the midgut cells of exposed Drosophila via endocytic vesicles

and by direct membrane penetration.Conclusion: aSNPs after their internalization in the midgut cells of exposed Drosophila larvae showmembranedestabilization along with increased cellular stress and cell death.General significance: Ingested aSNPs show adverse effects on the cells of GI tract of the exposed organism thustheir industrial use as a food-additive may raise concern to human health.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction

With an increase in the use of nanomaterials (NMs) in industrial andbiomedical applications in recent years, concern has been raised onentry of the NMs in the environment and subsequently, their possibleeffects on human health. The unique physico-chemical properties ofNMs such as particle size, agglomeration state, size distribution, shape,

les; NMs, Nanomaterials; NPs,trointestinal; Hsps, Heat shocksps, Small heat shock proteins;D, Superoxide dismutase; CAT,, Glutathione S-transferase; PC,hosphate buffered saline; BSA,icroscopy; DLS, Dynamic lightMP, Mitochondrial membrane

IR-Indian Institute of Toxicologyr Pradesh, India. Tel.: +91 522; fax: +91 522 [email protected]

l rights reserved.

crystal structure, chemical composition, surface chemistry, surfacecharge and porosity supplemented by altered dimensional configura-tion and high surface area to volume ratiomay account for their hazard-ous effects on biological systems [1,2].

Engineered nanoparticles (ENPs) comprise wide range of NMswhich are synthesized by molecular level engineering to acquireunique properties for their vast applications. Among the ENPs, silicananoparticles are one of the most important and most utilized NMs.Amorphous silica nanoparticles (aSNPs), characterized by theirnon-crystalline structure are produced in different sizes along with dif-ferent surface modifications with relative ease. This has led to the ex-tensive use of aSNPs for diverse applications that include agriculture,food and animal-feed additives [3].

aSNPs can enter into an organism including human through all pos-sible routes [2]. Since aSNPs are used in food products following theirapproval by FDA as a food additive [4], concern has been raised ontheir exposure to an organism through ingestion. Earlier, different ex-posure routes such as intra-tracheal instillation [5], inhalation [6] or in-travenous injection [7]were used for in vivo studies on aSNPs in rats andmice. Limited studies on aSNPs through oral route are available [8,9],

Page 2: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

2257A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

however, with inadequate characterization of nanoparticles (NPs) anddose level [10]. Dekkers et al. [3] analyzed several food products withadded silica (E551) for the presence of silica NPs, their particle size(30–200 nm) and concentration (0.3-33.0 mg nanosilica/day) usingexperimental analytical data. Further, theoretical approach towardrisk assessment of NMs in food [included in guidelines of EuropeanFood Safety Authority (EFSA, 2011)] [11] was put forth to find whetherNPs get absorbed in gastrointestinal (GI) tract in dissolved form or inparticulate form [10]. In a recent study, Peters et al. [12] performed invitro digestion of food products (instant soup, coffee creamer, pancake)having synthetic amorphous silica as food additives and demonstratedexposure of human intestinalwall to nano-sized silica particles. Howev-er, studies on internalization/absorption of aSNPs from GI tract aftertheir entry by oral route are not available.

NPs–membrane interactions are relevant in the context of NPsmediated cellular injury and hence gaining insights on the adverse ef-fects of NPs on membrane perturbation is pertinent. aSNPs wereshown to penetrate the platelet surface membrane causing distur-bance in calcium ion channels [13] and also to cause decrease inmembrane fluidity in exposed mouse macrophage cells [14]. Pattaniet al. [15] showed membrane perturbation by chitosan NPs by usingphospholipid vesicles as a model membrane while Valant et al. [16]showed cell membrane destabilization potential of nanotized TiO2,ZnO and fullerenes (C60) in a terrestrial invertebrate, Porcellio scaberafter their ingestion. However, studies on membrane destabilizationby ingested aSNPs in GI tract are not available.

Reactive oxygen species (ROS) mediated oxidative stress (OS) hasbeen correlated with NPs mediated cellular damage [1,17]. A numberof in vitro studies are now available on the mechanism of cellular stressagainst aSNPs in the context of their size. Reduced level of glutathione(GSH) and increased malondialdehyde level indicating lipid peroxida-tion (LPO) and membrane damage were correlated with OS inducedby aSNPs (15 nm) in A549 cells [18]. GSH depletion in aSNPs (30 nm)exposed mouse keratinocytes [19] and increased ROS generation anddepletion in GSH level in exposed human hepatic cells [20] were alsoreported. Yang et al. [21] observed size dependent hazard (smaller thesize greater the hazard) of aSNPs in human hepatoma (HepG2) cellsconcordant with previous reports [19,20,22]. Summarizing these stud-ies, it is nowevident that OS induced cellular adversities of aSNPs are as-sociated with their smaller size. However, data on cellular stressresponse after oral exposure of aSNPs is lacking.

Under various stressful conditions like exposure to xenobiotics,free radicals and elevated temperature, cells set up their defense byexpressing heat shock genes (hsg) which after their translation pro-duce heat shock proteins (Hsps) [23]. While temperature shock wasused as the primary inducer to demonstrate the elevated levels ofHsps in Drosophila [24], later, diverse groups of stressors were foundto induce a number of hsgs and hence a broader term stress proteins(Sps)was coined [25]. Under adverse conditions, Hspsmaintain the un-folded proteins prior to their assembly into multi-molecular complexesin cytosol and guide the newly synthesized as well as denatured pro-teins to accomplish their appropriate conformation, thereby protectingthe cells from further damage [26]. Among the families of Hsps, hsp70 isthe most conserved across the taxa and commonly expressed. Hsp70executes functions mainly as a chaperone facilitating the refolding ofpolypeptides and dissolution of insoluble aggregates by binding tohydrophobic surfaces of denatured proteins [27]. Earlier studiesfrom this laboratory showed that elevated levels of Hsp70 againstexposure to various environmental stressors such as, pesticides andfungicides [28,29] and industrial runoffs [30], thereby establishinghsp70 expression as a first tier indicator of cellular stress [30].Along with hsp70, other Hsps, such as small heat shock proteins(sHsps) having molecular mass 12–42 kDa, also play importantroles in cellular defense [31]. sHsps have been reported to formlarge oligomers which get phosphorylated under stress conditionsand subsequently dissociate from the oligomers and bind to

non-native proteins [32]. They interact with a number of proteinsto protect cells from OS [33], and apoptosis [34].

The present study therefore is aimed to examine the internaliza-tion of aSNPs (b30 nm) in the midgut tissue of the digestive tract ofDrosophila after their oral exposure and subsequently cellular stressresponse generated by these particles. Midgut tissue of Drosophilashares similarity with that of human intestine. Moreover, Drosophilaand human have similarities in food passage, tissue, anatomy andphysiological functions of intestine and its development. Both themammalian gut and Drosophilamidgut comprise an epithelial mono-layer of enterocytes having cytoplasmic extensions and microvilli toincrease the cell surface area facing gut lumen [35].

With the environment being the ultimate sink for the disposal ofchemicals, organisms inhabiting different compartments of the envi-ronment are expected to be exposed to chemicals/materials. Duringthe past decade, alternative animal models especially the lower eukary-otemodel organisms have gained importance in the context of studyingadverse effects of environmental chemicals/materials/contaminants[36]. Drosophila, an insect model, with well-documented genetics anddevelopmental biology and high degree of homology of its genes withthat of higher mammals, is the closest invertebrate to the humans andhas been used for toxicological studies [37] and for studying human dis-eases [38]. It raises fewer ethical concerns and falls within the recom-mendations of the European Centre for the Validation of AlternativeMethods (ECVAM), aimed to prop up the scientific and regulatory ac-ceptance of alternative methods that are important in the field of bio-logical science and toward reducing, refining and replacing the use oflaboratory animals [39].

2. Materials and methods

2.1. Fly strain and culture

Wild type D. melanogaster (Oregon R+) strain was used for thestudy. Flies and larvae were reared at 24±1 °C on standard Drosophiladiet containing agar-agar, maize powder, sugar, yeast, nepagin(methyl-p-hydroxy benzoate) and propionic acid. Additional yeast sus-pension was provided for healthy growth of the organism.

2.2. NPs and their characterization by transmission electron microscopy(TEM) and dynamic light scattering (DLS)

aSNPs (20–30 nm, purity>99%, specific surface area>600 m2/g)(NANOSHEL LLC, Wilmington DE, USA) were used. The morphologyand average size of aSNPs were analyzed by TEM after preparing thesample by drop coating aSNPs suspension (50 μg/mL in MilliQ water)on carbon-coated copper TEM grids. Films on the TEM grids wereallowed to dry. Measurement of aSNPs was performed on a FEI TecnaiG2 sprit Twin Transmission Electron Microscope equipped with a CCDCamera (Netherlands) and at an accelerating voltage of 80 kV. Averagesize of aSNPs (n=30) was measured from three independently pre-pared grids. Hydrodynamic size of the aSNPs was determined usingDLS method in a Zetasizer (Nano-ZS, Model ZEN3600) equipped withthe 4 mW, 633 nm laser (Malvern instruments Ltd., UK). aSNPs weresuspended in Milli-Q water and 5% sucrose solution (used as mediumfor aSNPs exposure) at a concentration of 200 μg/mL and probe sonicat-ed at 30 W for 10 min before the DLSmeasurement. DLSmeasurementswere carried out thrice with three replicates prepared from indepen-dent pools.

2.3. Treatment schedule

Three different concentrations (1, 10, 100 μg/mL) of aSNPs wereused.We essentially followed the exposure regimen as described previ-ously [40]. Briefly, third instar larvae (68–92 h) of Oregon R+ were ex-posed to different concentrations of aSNPs dispersed in 5% sucrose

Page 3: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

2258 A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

solution (20 μL sucrose solution/10 larvae) for 12–36 h. Control groupreceived 5% sucrose solution without aSNPs.

2.4. Temperature shock treatment

Healthy third instar larvae of Oregon R+ were given temperatureshock (37±1 °C) for 1 h as described previously [41]. Temperatureshock was used as a positive control for the induction of hsgs.

2.5. Reverse transcriptase-PCR (RT-PCR) analysis of hsgs in the midguttissue of third instar Oregon R+ larvae

Total RNA from control and treated third instar larvae of Oregon R+

was extracted using TRI reagent (Ambion, Austin, TX, USA) and reversetranscribed into cDNA using cDNA synthesis kit (Fermentas, MD, USA)essentially following the manufacturer's instructions. Forward and re-verse primers for hsp83, hsp70, hsp60, hsp27, hsp26, hsp23, hsp22 andgapdh (Table 1) were synthesized as previously described [42]. Follow-ing PCR, the amplicons were separated on an 1.5% agarose gelcontaining ethidium bromide at 5 V cm−1 and visualized under aVERSA DOC Imaging System Model 1000 (Bio-Rad, Hercules, CA, USA).The intensity of the bands was quantified by Quantity One software ofBio-Rad, CA, USA. For comparison, data were presented in terms of rel-ative intensity, calculated by normalization using gapdh as the internalcontrol. Each experiment was carried out thrice with three replicatesprepared from independent pools.

2.6. Assay of OS markers

To evaluate the oxidative damage caused by the aSNPs, ROS genera-tion, superoxide dismutase (SOD), catalase (CAT) and glutathioneS-transferase (GST) activities, glutathione (GSH) content, protein car-bonyl (PC) content and lipid peroxidation (LPO) product were assayedin control and exposed third instar larvae of Oregon R+. Except for theROS measurement for which single cell suspension was used, rest ofthe abovementioned assayswere carried out in 10% tissue homogenate.Each experiment was carried out thrice with three replicates preparedfrom independent pools.

2.6.1. Preparation of tissue homogenateLarval midgut tissue homogenate was prepared essentially follow-

ing a method as described previously [42]. The homogenate wascentrifuged at 10,000 ×g and the supernatant was used for differentassays and protein estimation.

2.6.2. Preparation of single cell suspensionMidgut tissues of 15 larvae from control and treated groupswere in-

cubated in collagenase (0.5 mg/ml) for 15 min at 24±1 °C. The cellswere passed through a nylon mesh (85 μm) and were washed with

Table 1Genes and their primer sequences used in RT-PCR amplification.

hsp22 Forward (F) 5′CGAGCCGCCCGTTTGGAGT3′Reverse (R) 5′GACGAAGCGGCGGAGGAAGTG3′

hsp23 Forward (F) 5′GAGCCTTGCCGACGATTTG3′Reverse (R) 5′GGCGCCCACCTGTTTCTC3′

hsp26 Forward (F) 5′CAAGCAGCTGAACAAGCTAACAATCTG3′Reverse (R) 5′GCATGATGTGACCATGGTCGTCCTGG3′

hsp27 Forward (F) 5′CATCCGCGTCGCCTGCTACTG3′Reverse (R) 5′CTCGCGCTCCTCGTGCTTCC3′

hsp60 Forward (F) 5′CCTCCGGCGGCATTGTCTTC3′Reverse (R) 5′AGCGCATCGTAGCCGTAGTCACC3′

hsp70 Forward (F) 5′GAACGGGCCAAGCGCACACTCTC3′Reverse (R) 5′TCCTGGATCTTGCCGCTCTGGTCTC3′

hsp83 Forward (F) 5′CCCGTGGCTTCGAGGTGGTCT3′Reverse (R) 5′TCTGGGCATCGTCGGTAGTCATAGG3′

gapdh Forward (F) 5′AATTCCGATCTTCGACATGG 3′Reverse (R) 5′GAAAAAGCGGCAGTCGTAAT3′

phosphate buffered saline (PBS) to remove collagenase. Finally, thecells were processed for measurement of ROS generation.

2.6.3. Measurement of ROS

Intracellular ROS generation in the midgut cells of control andtreated larvae was estimated by flowcytometry using a dye, 2′,7′-dichlorodihydrofluorescein diacetate (DCF-DA; Sigma Chemicals,St. Louise, MO, USA) [42,43]. Data were analyzed using cell quest soft-ware (Mac OS 8.6) and 10,000 events were counted per sample.

2.6.4. Superoxide dismutase (SOD) (superoxide: superoxideoxidoreductase EC 1.15.1.1)

SOD activity was estimated as described by Nishikimi et al. [44]with minor modifications as reported by Singh et al. [42]. One unitof enzyme activity is defined as enzyme concentration required forinhibiting chromogen production (optical density 560 nm) by 50%in 1 min under assay condition and the data were expressed as thespecific activity in units/min/mg larval protein.

2.6.5. Catalase (CAT) (H2O2:H2O2 oxidoreductase EC 1.11.1.6)CAT activity in the control and treated larvaewasmeasured as de-

scribed previously [45]. Enzyme activity was expressed as μmol H2O2

decomposed/min/mg/larval protein.

2.6.6. Glutathione S-transferase (GST, EC 2.5.1.18)Glutathione S-transferase (GST) activity was determined by the

method of Habig et al. [46]. The enzyme activity was calculated asnmol CDNB reduced/min/mg larval protein.

2.6.7. Glutathione (GSH) contentGSH content in the control and exposed larvae was quantified using

Elmann's reagent [47]. The assay was performed essentially following apreviously describedmethod [42]. GSH content was expressed in termsof nmol/mg larval protein.

2.6.8. Assay for lipid peroxidation (LPO)Malonyl dialdehyde (MDA) content as a measure of LPO was

assayed using tetraethoxypropane as an external standard [48]. Lipidperoxide level was expressed in terms of nmoles MDA formed/h/mglarval protein.

2.6.9. Determination of protein carbonyl (PC) contentWemeasured PC content in control and treated groups as described

previously [49]. In brief, absorption of di-nitro phenyl hydrazine(DNPH) was determined at 370 nm using 2 M HCl as a blank. The re-sults were expressed as nmoles of DNPH incorporated mg−1 protein.

2.6.10. Protein estimationProtein concentrationwas determined by themethod of Lowry et al.

[50] using Folin reagent and bovine serum albumin (BSA) as thestandard.

2.7. Determination of mitochondrial membrane potential (MMP)

Depolarization of mitochondrial membranes was analyzed by fluo-rochrome 5, 5′, 6, 6′-Tetrachloro-1, 1′, 3, 3′-tetraethyl benzimidazolylcarbocyanine iodide (JC-1) (Invitrogen, USA) following a previouslypublished procedure [51]. Results were expressed as the percentage ofdepolarized cells. Each experiment was carried out thrice with threereplicates prepared from independent pools.

2.8. Colorimetric assay for IETD- and DEVD-ase activities

Caspase activities (as apoptotic endpoint marker) weremeasuredin control and treated larvae using synthetic tetrapeptide substrates

Page 4: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 1. Characterization of aSNPs by TEM. Transmission electron microscopic images ofaSNPs showing (A) dispersion inwaterwith smaller aggregates and (B)morphology. Rep-resentative image is from the grid prepared from one of the three independent pools. Datarepresent are mean±SD of aSNPs size (n=30) from three independently prepared grids.

Table 2Dynamic light scattering (DLS) measurements of aSNPs showing hydrodynamic sizeand polydispersity index (PDI) in water and 5% sucrose solution.

Dispersion medium Average hydrodynamicsize (nm)

Polydispersity Index (PDI)

MilliQ water 138.63±7.81 0.2615% sucrose solution 148.14±9.96 0.267

2259A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

against each enzyme essentially following a previously describedmethod [52]. Each experiment was carried out thrice with three rep-licates prepared from independent pools.

2.9. Electron microscopy in larval gut tissue

Midgut tissues dissected from third instar larvae of control and NPstreated groups were fixed in a solution of 2% glutaraldehyde and 4%paraformaldehyde in 0.1 M cacodylate buffer (pH 7.2), post-fixed in2% osmium tetroxide, dehydrated and embedded in Araldite 502 (TedPella, Inc., Redding, CA, USA). Ultrathin sections (70–90 nm thick,using an Ultra Cut UCT ultra-microtome, Leica) of the blocks (preparedfrom three independent pools) were picked up on copper grids, stainedwith uranyl acetate and lead citrate, and analyzed on a FEI Tecnai G2sprit Twin Transmission Electron Microscope equipped with CCD Cam-era (Netherlands) at 80 kV.

2.10. Acridine orange (AO)/Ethidium Bromide (EB) staining for mem-brane stability in midgut cells of Oregon R+ larvae

Cell membrane destabilization was measured following a recentlydescribed method [53]. Isolated gut tissues from control and aSNPs ex-posed groups were stained with a mixture of AO and EB for 5 min andobserved on a Leica TCS SPE confocal microscope (Nussloch, Germany).Cell membrane integrity was assessed visually by counting the EBstained cells from three independent pools with three replicates(30 larvae/replicate). Scale of membrane destabilization was set asnormal (b5%), medium (b95%) and severe (>95%).

2.11. Statistical analysis

Statistical significance of the mean values for different parame-ters was monitored in control and exposed larvae using ANOVAfollowed by post hoc tests after ascertaining the homogeneity of var-iance and normality of data. For multiple comparisons One-way andtwo-way ANOVA were followed by Dunnett and Bonferroni's test re-spectively. Each endpoint was considered as dependent variablewhile concentration and time of exposure as independent variables.Statistical significance level was ascribed as *Pb0.05 or **Pb0.001.Pearson's correlations were calculated and then linear regressionanalysis was carried out. Prism computer program (GraphPad ver-sion 5.0, San Diego, CA, USA) was used for the statistical analysis.

3. Results

During the course of study, no overt signs of toxicity were ob-served in the exposed organisms.

3.1. Characterization of NPs

TEM analysis showed spherical aSNPs (average particle size of27.28±4.1 nm) dispersed with small aggregates (Fig. 1). The averagehydrodynamic size of aSNPs by DLS measurement was found to be138.63±7.81 nm inwater and 148.14±9.96 nm in 5% sucrose solution(Table 2). Size distribution by intensity of the aSNPs in both themediumis presented in Fig. 2.

3.2. aSNPs induced hsp70 and hsp22 expression in exposed third instarlarvae of D. melanogaster (Oregon R+)

To examine aSNPs-induced expression of selected stress genes in theexposed Oregon R+ larvae, semi-quantitative RT-PCR assay for hsp83,hsp70, hsp60, hsp27, hsp26, hsp23 and hsp22 was carried out. We ob-served a non-significant (P>0.05) change in the expression of hsp83,hsp60, hsp27 and hsp26 in the larvae exposed to 1–100 μg/mL aSNPsfor 12–36 h as compared to their respective control. A significant

(Pb0.05; Pb0.001) concentration-dependent increase in the expressionof hsp70 and hsp22was observed in the exposed larvae after 36 h (72.5and 78% increase in the expression of hsp70 and hsp22 at 100 μg/mLaSNPs after 36 h respectively) (Fig. 3).

3.3. Elevated ROS generation in aSNPs exposed third instar larvae ofOregon R+

Fig. 4 shows flowcytometricmeasurement of ROS generation in con-trol and aSNPs-exposed larvae. A concentration- and time-dependentsignificant (Pb0.05; Pb0.001) increase in ROS generationwas observedin the exposed larvae (3.5-fold increase in ROS generation in 100 μg/mLaSNP treated larvae after 36 h).

3.4. Effect of aSNPs on anti-oxidant markers in exposed third instarOregon R+ larvae

Fig. 5A shows SOD activity in control and aSNPs-exposed larvae. Weobserved a non-significant (P>0.05) change in SOD activity in 1 μg/mLaSNPs-treated larvae throughout the exposure periods while atime-dependent significant (Pb0.05; Pb0.001) increase in the enzymeactivity was observed in 100 μg/mL aSNPs-treated group (44.5% in-crease in the enzyme activity at 100 μg/mL aSNPs concentration after36 h). A similar trend was observed for CAT activity and PC content inthe exposed larvae (41.4 and 42.8% increase in CAT activity and PC

Page 5: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 2. Characterization of aSNPs by DLS method. DLS measurements showing size intensity curves in (C) water and (D) 5% sucrose solution. Data represent are mean±SD of hy-drodynamic size from three independent measurements.

2260 A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

content at 100 μg/mL aSNPs after 36 h respectively) (Fig. 5B–C). A sig-nificant (Pb0.05; Pb0.001) concentration- and time-dependent reduc-tion in GSH content was observed in the exposed larvaewithmaximumreduction in GSH content (53%) at 100 μg/mL aSNPs after 36 h (Fig. 5D).We observed a significantly increased MDA content and GST activity inthe exposed larvae in a concentration- and time-dependent manner(Fig. 5E–F).

3.5. Uptake and distribution of aSNPs in gut epithelial cells of exposedOregon R+ larvae

Ultrastructural morphology in control cells show distinct mito-chondria, Golgi bodies, smooth and rough endoplasmic reticulum,fewer vacuoles along with intact microvilli. Deviation from the con-trol cells was evident in 100 μg/mL aSNPs treated organism as re-vealed by damaged mitochondria (membrane and cristae damage),increased vacuolization and shrink and damaged microvilli. aSNPsand their aggregates were found to be entrapped in endocytic vesi-cles, accumulated in mitochondria and internalized after their pene-tration through cell membrane. More distinctly, aSNPs were found

Fig. 3. RT-PCR analysis for mRNA expression of hsp70 and hsp22. (A) Agarose gel images andaSNP treated third instar larvae of D. melanogaster for 12–36 h. Data represent are mean±Splate] of three independent experiments with three replicates. Significance is ascribed as *P

to be dispersed in cytoplasm in the form of small aggregates(Fig. 6A–H).

3.6. Membrane stability in midgut cells of exposed third instar Oregon R+

larvae

To confirm our observation of cell membrane andmicrovilli damagein the exposed larvae by TEM, AO/EB staining was performed in controland treated group to examine membrane destabilization. We observedsignificant (Pb0.05; Pb0.001) membrane destabilization of gut cells(>5% EB stained nuclei) in a concentration dependent manner (severedestabilization in 17% of the larvae exposed to 100 μg/mL aSNPs for36 h) (Fig. 7).

3.7. Loss of mitochondrial membrane potential (MMP) in midgut cells ofexposed third instar Oregon R+ larvae

Disintegrated mitochondrial structure along with aSNPs depositionin mitochondria observed by TEM prompted us to examine the effectof aSNPs on MMP in the exposed larvae. A significant (Pb0.05;

(B) densitometric representation of mRNA expression of hsp70 and hsp22 in control andD of relative intensity [relative to internal control (gapdh), originating from same tem-b0.05 or **Pb0.001 as compared to control. PC, positive control.

Page 6: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 4. Estimation of intracellular ROS. ROS generation in control and aSNPs treatedthird instar larvae of D. melanogaster for 12–36 h. Data represent are mean±SD ofthree independent experiments made in three replicates. Significance is ascribed as*Pb0.05 or **Pb0.001 as compared to control.

2261A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

Pb0.001) concentration-dependent loss of MMP (expressed as percentof depolarized cells) was observed in the cells of exposed larvae after36 h in comparison to control (Fig. 8).

Fig. 5. Measurement of OS endpoints. (A–B) Superoxide dismutase (SOD) and catalase (C(MDA) contents and (F) glutathione S-transferase (GST) activity in control and aSNP exposthree independent experiments made in three replicates. Significance is ascribed as *Pb0.0

3.8. IETD- andDEVD-ase activities in exposed third instar Oregon R+ larvae

To investigate aSNPs-induced apoptosis in the exposed organism,we measured IETD- and DEVD-ase activities as markers of apoptosisin control and treated group after 36 h. Both these enzyme activitieswere found to be increased in the exposed larvae in a concentration de-pendent manner (3- and 2.5-folds increase in IETD- and DEVD-ase ac-tivities at 100 μg/mL aSNPs concentration) (Fig. 9).

3.9. Correlation among different stress parameters

A correlation was drawn among different endpoints relevant toaSNPs-induced adverse effects in the exposed organism (Table 3). A sig-nificant negative correlation (r=−0.987; Pb0.001) was observed be-tween MDA content (lipid peroxidation) and GSH content. Further,GSH content was found to be negatively correlated (r=0.969;Pb0.001) with ROS generation. A significant positive correlation wasdrawn between ROS generation and hsp70 (r=0.996; Pb0.001), MMP(r=0.995; Pb0.001), ROS vs. hsp22 (r=0.996; Pb0.001) and ROS vs.DEVDase activity (r=0.969; Pb0.001).

AT) activities, (C–E) protein carbonyl (PC), glutathione (GSH) and malondialdehydeed third instar larvae of D. melanogaster for 12–36 h. Data represent are mean±SD of5 or **Pb0.001 as compared to control.

Page 7: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 6. TEM analysis of midgut tissue. TEM images of midgut tissue of control and aSNP exposed third instar larvae of D. melanogaster for 36 h showing intracellular organization in(A) control and (B) treated, morphology of microvilli in (C) control and (D) treated (E) aSNPs aggregates in cytoplasm, (F) mitochondrial disintegration, (G) endocytic uptake and(H) direct invasion of aSNPs aggregates inside the cell. M, mitochondria; V, vesicles; N, nucleus; MV, microvilli. Arrows indicate the presence of aSNPs.

2262 A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

Page 8: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 7. Membrane destabilization in midgut cells. (A) Graphical representation and(B) confocal images showing AO/EB staining in the midgut cells of control and aSNPsexposed third instar larvae of D. melanogaster after 36 h. Red and white arrows denoteAO and EB stained nuclei respectively. Data represent are mean±SD of three indepen-dent experiments made in three replicates. Significance is ascribed as *Pb0.05 or**Pb0.001 as compared to control.

2263A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

4. Discussion

The present in vivo study was carried out to examine absorption/internalization of ingested aSNPs in cells lining the midgut and alsocellular stress response in exposed D. melanogaster larvae.

Physiochemical properties of NPs are essential to be characterizedbefore they are evaluated for the adverse effects [19,54] and this ismore so due to alteration in their properties viz. dispersity, agglomera-tion and aggregation in different mediums. We used aSNPs of b30 nm,average particle size of which was further confirmed by TEM analysis.During the state of suspension, aSNPs exhibitedminimum precipitationand aggregation which was apparent from the correlograms in waterand 5% sucrose solution as the end of curves were almost smooth(please see S1). According to the emerging picture about NPs adversi-ties, several in vitro studies had documented that smaller NPs, with a di-ameter of ≤10 nm, are consistently more detrimental to the biologicalsystems than the larger analogues, i.e., >100 nm [20–22]. Taking thisinto consideration, aSNPs of b30 nm were chosen for the study.

Excessive ROS generation and decreased antioxidant capacity inan organism have been suggested to play role in the toxicity of NPs[55,56]. In agreement with the earlier studies, we also observed sig-nificant generation of ROS in the aSNPs exposed organism along with

Fig. 8. Measurement of MMP in midgut cells. Alteration in MMP in control and aSNPstreated third instar larvae of D. melanogaster for 36 h. Data represent are mean±SDof three independent experiments made in three replicates. Significance is ascribedas *Pb0.05 or **Pb0.001 as compared to control.

induction of OS such as significant alteration in LPO, GSH content,anti-oxidant enzyme activities (SOD, CAT) and PC content. We deter-mined the generation ROS in control and in aSNPs exposed organismby flowcytometry using a dye DCF-DA. Its use is recommended as asuitable probe for the measurement of total ROS generation withproper methodological considerations to avoid technical pitfalls[57]. However, direct reaction of this dye with cytochrome c and cy-toplasmic oxidases or peroxidases and self ROS generation by thisdye are also reported [58]. Thus, possibility of overestimating thegeneration of ROS in the present experimental set up cannot beruled out. Significantly enhanced SOD and CAT activities observed inthe exposed organism may be correlated with the activation of cellulardefense to abate the adverse effect of free radicals [59]. Utilization ofGSH for scavenging free peroxide radicals in the aSNPs exposed organ-ism was reflected by a significant negative correlation drawn betweenMDA content and GSH content (YGSH=−7.34XMDA+56.42; r=0.987; Pb0.001). Similarly, a decline in anti-oxidant level in the exposedorganismmay be correlated with an increased free radical/ROS state asevident by a significant negative correlation drawn between GSH con-tent and ROS (YROS=−0.861XGSH+41.89; r=0.969; Pb0.001).

Parallel to the activation of anti-oxidant defense system in a xenobi-otic exposed organism, another mode of cellular defense operates, thatis induction of hsgs. Increased expression of hsgs occurs as a defensivemean through non-specific mechanisms of toxicity that involves thegeneration of abnormal proteins and alteration of cellular functions.We, therefore, tested this possibility by measuring the induction of se-lected hsgs viz. hsp83, hsp70, hsp60, hsp27, hsp26, hsp23 and hsp22.Among these hsgs, only hsp70 and hsp22were found to be significantlyinduced in the aSNPs exposed organism. Hsp70 has been reported as anearly indicator of cellular stress induced by several environmentalchemicals [31] including NPs [55]. We also observed a significant in-crease in the induction of hsp70 in the exposed organism reflecting cel-lular stress caused by the aSNPs. ROS levels have been documentedearlier to play an important role in the regulation of gene expressionby activating transcription factors that, in turn, canmodulate the induc-tion of proteins involved in cellular response to environmental condi-tions [60]. Therefore, it is likely that ROS generated in the exposedorganism would activate transcription factors that can trigger the in-duction of hsps, viz., hsp70. This was further corroborated when a posi-tive correlation was drawn between hsp70 expression and ROS levels(YHSP70=1.548XROS+25.48; r=0.995; Pb0.001).

Screening of only one particular class of stress protein may not beadequate for identifying a sensitive bioindicator against a broad rangeof pollutants due to different levels of expression of different familiesof stress proteins against different agents [61,62]. Studies from this labas well by others reported induction of other hsgs either parallel tothe induction of hsp70 or without the significant induction of hsp70 bydifferent environmental chemicals [37,42,63]. In this context, we ob-served a significant induction of hsp22 along with hsp70. Parallel tothe positive correlation drawn between hsp70 and ROS levels, we alsoobserved a positive correlation drawn between hsp22 and ROS levels(YHSP22=2.051XROS+23.44; r=0.995, Pb0.001) in the aSNPs exposedgroups. Morrow et al. [64] demonstrated an increased expression ofhsp22 and its association with mitochondrial damage in Drosophila. Inthis study too, significant loss of MMP as demonstrated byflowcytometry and mitochondrial membrane disintegration by TEMfurther indicates that hsp22 induction might have a role in such cellularresponses in the exposed organism.

Excess OS generated in an organism after xenobiotics exposure canlead to cell death [65,66]. Parallel to the increased OS generated inaSNPs exposed Drosophila, we observed a significant increase in IETD-and DEVD-ase activities (corresponding to Caspase −9 and 3) concur-rent with a significant change in MMP indicating increased cell deathin the organism. This was further corroborated when we drew a posi-tive correlation between OS endpoints and IETD- and DEVD-ase activi-ties measured in this study (YDEVDase=0.002XROS+0.017; please see

Page 9: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 9. Colorimetric estimation of caspases activity. (A) IETDase (caspase 9) and (B) DEVDase (caspase 3) activities in control and aSNPs treated third instar larvae of D. melanogasterfor 36 h. Data represent are mean±SD of three independent experiments made in three replicates. Significance is ascribed as *Pb0.05 or **Pb0.001 as compared to control.

2264 A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

S2). Hsps have been reported to have both pro- and anti-apoptoticproperties [67]. While Hsp70 has been reported to be largelyanti-apoptotic, Hsp22 has been suggested to possess both pro- andanti-apoptotic properties depending on the cell types [68]. Significantincrease in hsp70 and hsp22 expression observed in the aSNPs exposedorganism in the context of increased cell death might be a consequenceof subtle balance in the expression of these two hsgs under the presentexperimental condition. At present, we are unable to critically evaluatethe role of each hsg on apoptotic cell death in the exposed organism.

Membranes as the biological boundary of cells facilitate the uptake/internalization of different molecules inside cells while endocytosis is aphysiological process for engulfing themolecules/substances and isme-diated by the formation of endocytic vesicles (endosomes). Theentrapped molecules are eventually localized in intracellular environ-ment. NPs of varying size and compositionmay be taken up by differentendocytic mechanisms ending up in different intracellular traffickingpathways [69]. Earlier reports have shown internalization of NPsthrough the process of endocytosis [70,71]. Thus, release of these NPsfrom the endosomes into cytoplasm [72] and subsequently, their expo-sure to intracellular moieties and organelles remains a possibility. Weobserved entrapped aSNPs inside the endocytic vesicles in aggregatedform in the midgut cells of exposed Drosophila. In agreement with ear-lier in vitro reports [70,73], we also observed the dispersion of aSNPs inthe cytoplasm of cells of the exposed organism by TEM analysis. Thus,internalization of aSNPs inside the cells taking place through direct pen-etration of cell membrane along with endocytic uptake cannot be ruledout.

To gain insight on the mechanism of NPs-induced cellular stress,earlier studies using either model membrane systems (lipid bilayer)or cell lines or in vivomodel systems revealed NPsmediatedmembranedamage through particle-membrane interaction [14,16,18]. Membranedisintegration due to exposure of NPs and their aggregates can lead to

Table 3Correlation among different stress endpoints.

hsp70 hsp22 ROS LPO GSH MMP DEVDase(caspase-3)

hsp70 1hsp22 0.998 1ROS 0.996 0.996 1LPO 0.987 0.988 0.995 1GSH −0.952 −0.953 −0.969 −0.987 1MMP 0.993 0.994 0.995 0.981 −0.942 1DEVDase(caspase-3

0.988 0.989 0.996 0.997 −0.986 0.984 1

Pb0.001.

the damage of proteins and lipid molecules. We hypothesized thataSNPs induced cellular stress in the exposed organism could be associ-ated with membrane damage caused by these particles. Our observa-tion of abnormal morphology (thinning and shortening in size) ofmicrovilli of themidgut of exposedDrosophila by TEMalongwithmem-brane destabilization as evident by positive AO/EB staining in the mid-gut cells of exposed organism at the highest tested concentration ofaSNPs further confirms the above possibility. It is tempting to speculatethat physical obstruction to the membrane by aggregates may happenas the concentration of aSNPs increase (1–100 μg/mL) in the exposedorganism leading to severe membrane destabilization. Among differentcellular organelles, mitochondria are responsible for intracellular ROSmetabolism [74]. Safeguarding of ROS homeostasis in mitochondria isreported to be dependent on respiratory chain andmembrane potential[75]. Our observation of aSNPs accumulation in mitochondria alongwith a significant loss ofMMP in the exposed organism raises a possibil-ity of mitochondria mediated ROS generation leading to cellular stressin the organism. Since we have used collagenase to make a suspensionof dissociated cells for assaying MMP and ROS, a possibility that the en-zyme influencing stability of the cellular membranes of the aSNPs ex-posed organism cannot be ruled out. To the best of our knowledge,this is the first report on aSNPs induced cellular stress response vis avis their cellular internalization using a well accepted in vivo modelDrosophila which has minimum ethical concern.

5. Conclusion

Taken together, the study suggests a negative impact of ingestedaSNPs (b30 nm) on the midgut tissues of exposed Drosophila alongwith elucidation of possible mechanism of cellular stress response(Fig. 10). The study also recommends the use of Drosophila as an invivo model for studying adverse effects of NPs which might havewider applications in the context of similarities existing between themidgut tissue of this organism and that in higher organisms.

Supplementary data to this article can be found online at http://dx.doi.org/10.1016/j.bbagen.2012.10.001.

Conflict of interest

The authors declare that they have no conflict of interest.

Acknowledgments

The authors are grateful to Dr. K.C. Gupta, Director, CSIR-Indian Insti-tute of Toxicology Research, Lucknow for support and for Oregon R+

stock by Drosophila Stock Centre, Bloomington, USA. Financial

Page 10: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

Fig. 10. Schematic illustration of aSNPs-uptake and cellular response. Schematic model showing the mechanism of internalization of aSNPs and cellular stress evoked by these NPsin midgut cells of exposed Drosophila larvae. aSNPs, amorphous silica nanoparticles; PM, plasma membrane; EV, endocytic vesicle; LPO, lipid peroxidation GSH, glutathione; ROS,reactive oxygen species; MT, mitochondria; MMP, mitochondrial membrane potential; HSPs, heat shock proteins; SOD, superoxide dismutase; CAT, catalase.

2265A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

assistance to AP as JRF and SRF [20-6/2008(ii)EU-IV], SC as JRF [20-12/2009(ii)EU-IV] from University Grants Commission (UGC), New Delhiand to DKC from Council of Scientific of Industrial Research (CSIR),NewDelhi (SIP-08 and NWP-34) is thankfully acknowledged. IITR com-munication number 3061.

References

[1] A. Nel, T. Xia, L. Madler, N. Li, Toxic potential of materials at the nanolevel, Science(New York, N.Y.) 311 (2006) 622–627.

[2] G. Oberdorster, E. Oberdorster, J. Oberdorster, Nanotoxicology: an emerging dis-cipline evolving from studies of ultrafine particles, Environ. Health Perspect.113 (2005) 823–839.

[3] S. Dekkers, P. Krystek, R.J. Peters, D.P. Lankveld, B.G. Bokkers, P.H. vanHoeven-Arentzen, H. Bouwmeester, A.G. Oomen, Presence and risks of nanosilicain food products, Nanotoxicology 5 (2011) 393–405.

[4] C.A. Barnes, A. Elsaesser, J. Arkusz, A. Smok, J. Palus, A. Lesniak, A. Salvati, J.P.Hanrahan, W.H. Jong, E. Dziubaltowska, M. Stepnik, K. Rydzynski, G. McKerr, I.Lynch, K.A. Dawson, C.V. Howard, Reproducible comet assay of amorphous silicananoparticles detects no genotoxicity, Nano Lett. 8 (2008) 3069–3074.

[5] W.S. Cho, M. Choi, B.S. Han, M. Cho, J. Oh, K. Park, S.J. Kim, S.H. Kim, J. Jeong, In-flammatory mediators induced by intratracheal instillation of ultrafine amor-phous silica particles, Toxicol. Lett. 175 (2007) 24–33.

[6] C.M. Sayes, K.L. Reed, K.P. Glover, K.A. Swain, M.L. Ostraat, E.M. Donner, D.B.Warheit, Changing the dose metric for inhalation toxicity studies: short-termstudy in rats with engineered aerosolized amorphous silica nanoparticles, Inhal.Toxicol. 22 (2010) 348–354.

[7] R. Kumar, I. Roy, T.Y. Ohulchanskky, L.A. Vathy, E.J. Bergey, M. Sajjad, P.N. Prasad,In vivo biodistribution and clearance studies using multimodal organically modi-fied silica nanoparticles, ACS Nano 4 (2010) 699–708.

[8] K.E. Cha, H. Myung, Cytotoxic effects of nanoparticles assessed in vitro and in vivo,J. Microbiol. Biotechnol. 17 (2007) 1573–1578.

[9] S.J. So, I.S. Jang, C.S. Han, Effect of micro/nano silica particle feeding for mice,J. Nanosci. Nanotechnol. 8 (2008) 5367–5371.

[10] S. Dekkers, H. Bouwmeester, P. Bos, R. Peters, A. Rietveld, A. Oomen, Knowledge gapsin risk assessment of nanosilica in food: evaluation of the dissolution and toxicityof different forms of silica, Nanotoxicology (2012), http://dx.doi.org/10.3109/17435390.2012.662250 (Epub).

[11] E.F.S.A. (EFSA), Guidance on the risk assessment of the application of nanoscienceand nanotechnologies in the food and feed chain, EFSA J. 9 (2011).

[12] R. Peters, E. Kramer, A.G. Oomen, Z.E.H. Rivera, G. Oegema, P.C. Tromp, R. Fokkink, A.Rietveld, H.J.P. Marvin, S. Weigel, A.A.C.M. Peijnenburg, H. Bouwmeester, Presenceof nano-sized silica during in vitro digestion of food containing silica as a food addi-tive, ACS Nano 6 (2012) 2441–2451.

[13] J.J. Corbalan, C. Medina, A. Jacoby, T. Malinski, M.W. Radomski, Amorphous silicananoparticles aggregate human platelets: potential implications for vascular ho-meostasis, Int. J. Nanomed. 7 (2012) 631–639.

[14] H. Yang, Q. Wu, M. Tang, L. Kong, Z. Lu, Cell membrane injury induced by silicananoparticles in mouse macrophage, J. Biomed. Nanotechnol. 5 (2009)528–535.

[15] A. Pattani, V.B. Patravale, L. Panicker, P.D. Potdar, Immunological effects andmembrane interactions of chitosan nanoparticles, Mol. Pharm. 6 (2009) 345–352.

[16] J. Valant, D. Drobne, K. Sepcic, A. Jemec, K. Kogej, R. Kostanjsek, Hazardous potentialof manufactured nanoparticles identified by in vivo assay, J. Hazard. Mater. 171(2009) 160–165.

[17] H.R.S. Meng, T. Xia, S. George, A.E. Nel, A predictive toxicological paradigm for thesafety assessment of nanomaterials, ACS Nano 3 (2009) 1620–1627.

[18] W. Lin, Y.W. Huang, X.D. Zhou, Y. Ma, In vitro toxicity of silica nanoparticles inhuman lung cancer cells, Toxicol. Appl. Pharmacol. 217 (2006) 252–259.

[19] K.O. Yu, C.M. Grabinski, A.M. Schrand, R.C. Murdock, W. Wang, B. Gu, J.J. Schlager,S.M. Hussain, Toxicity of amorphous silica nanoparticles in mouse keratinocytes,J. Nanopart. Res. 11 (2009) 15–24.

[20] Y. Ye, J. Liu, J. Xu, L. Sun, M. Chen, M. Lan, Nano-SiO2 induces apoptosis via activa-tion of p53 and Bax mediated by oxidative stress in human hepatic cell line,Toxicol. In Vitro 24 (2010) 751–758.

[21] L. Yang, S. Lei, J. Minghua, D. Zhongjun, L. Xiaomei, G. Caixia, L. Yanbo, H. Peili, S.Zhiwei, Size-dependent cytotoxiciy of amorphous silica nanoparticles in humanhepatoma HepG2 cells, Toxicol. In Vitro 25 (2011) 1343–1352.

[22] H. Nabeshi, T. Yoshikawa, K. Matsuyama, Y. Nakazato, A. Arimori, M. Isobe, S.Tochigi, S. Kondoh, T. Hirai, T. Akase, T. Yamashita, K. Yamashita, T. Yoshida, K.Nagano, Y. Abe, Y. Yoshioka, H. Kamada, T. Imazawa, N. Itoh, S. Tsunoda, Y.Tsutsumi, Size-dependent cytotoxic effects of amorphous silica nanoparticles onLangerhans cells, Pharmazie 65 (2010) 199–201.

[23] B.G. Atkinson, Synthesis of heat-shock proteins by cells undergoing myogenesis,J. Cell Biol. 89 (1981) 666–673.

[24] F. Ritossa, A new puffing pattern induced by temperature shock and DNP inDrosophila, Experientia 18 (1962) 571–573.

[25] D.R. Boreham, R.E. Mitchel, Regulation of heat and radiation stress responses inyeast by hsp-104, Radiat. Res. 137 (1994) 190–195.

[26] R.I. Morimoto, Cells in stress: transcriptional activation of heat shock genes, Science259 (1993) 1409–1410.

[27] N. Yokoyama, M. Hirata, K. Ohtsuka, Y. Nishiyama, K. Fujii, M. Fujita, K. Kuzushima,T. Kiyono, T. Tsurumi, Co-expression of human chaperone Hsp70 and Hsdj or Hsp40

Page 11: Cellular internalization and stress response of ingested amorphous silica nanoparticles in the midgut of Drosophila melanogaster

2266 A. Pandey et al. / Biochimica et Biophysica Acta 1830 (2013) 2256–2266

co-factor increases solubility of overexpressed target proteins in insect cells,Biochim. Biophys. Acta 1493 (2000) 119–124.

[28] A. Nazir, I. Mukhopadhyay, D.K. Saxena, D.K. Chowdhuri, Evaluation of the no ob-served adverse effect level of solvent dimethyl sulfoxide in Drosophilamelanogaster, Toxicol. Mech. Methods 13 (2003) 147–152.

[29] A. Nazir, I. Mukhopadhyay, D.K. Saxena, D. Kar Chowdhuri, Chlorpyrifos-inducedhsp70 expression and effect on reproductive performance in transgenic Drosoph-ila melanogaster (hsp70-lacZ) Bg9, Arch. Environ. Contam. Toxicol. 41 (2001)443–449.

[30] I. Mukhopadhyay, D.K. Saxena, D.K. Chowdhuri, Hazardous effects of effluentfrom the chrome plating industry: 70 kDa heat shock protein expression as amarker of cellular damage in transgenic Drosophila melanogaster (hsp70-lacZ),Environ. Health Perspect. 111 (2003) 1926–1932.

[31] S.C. Gupta, A. Sharma, M. Mishra, R.K. Mishra, D.K. Chowdhuri, Heat shock pro-teins in toxicology: how close and how far? Life Sci. 86 (2010) 377–384.

[32] T. Stromer, M. Ehrnsperger, M. Gaestel, J. Buchner, Analysis of the interaction ofsmall heat shock proteins with unfolding proteins, J. Biol. Chem. 278 (2003)18015–18021.

[33] B.K. Derham, J.C. Ellory, A.J. Bron, J.J. Harding, The molecular chaperonealpha-crystallin incorporated into red cell ghosts protects membrane Na/K-ATPaseagainst glycation and oxidative stress, Eur. J. Biochem. 270 (2003) 2605–2611.

[34] A.P. Arrigo, In search of the molecular mechanism by which small stress proteinscounteract apoptosis during cellular differentiation, J. Cell. Biochem. 94 (2005)241–246.

[35] Y. Apidianakis, L.G. Rahme, Drosophila melanogaster as a model for human intes-tinal infection and pathology, Dis. Model. Mech. 4 (2011) 21–30.

[36] E.G. Stringham, E.P.M. Candido, Transgenic hsp16-lacZ strains of the soil nema-tode Caenorhabditis elegans as biological monitors of environmental stress, Envi-ron. Toxicol. Chem. 13 (1994) 1211–1220.

[37] A. Sharma, M. Mishra, K.R. Ram, R. Kumar, M.Z. Abdin, D.K. Chowdhuri,Transcriptome analysis provides insights for understanding the adverse effectsof endosulfan in Drosophila melanogaster, Chemosphere 82 (2011) 370–376.

[38] A. Jeibmann, W. Paulus, Drosophila melanogaster as a model organism of brain dis-eases, Int. J. Mol. Sci. 10 (2009) 407–440.

[39] D.J. Benford, A.B. Hanley, K. Bottrill, S. Oehlschlager, M. Balls, F. Branca, J.J.Castegnaro, J. Descotes, K. Hemminiki, D. Lindsay, B. Schilter, Biomarkers as pre-dictive tools in toxicity testing, Altern. Lab. Anim. 28 (2000) 119–131.

[40] R. Goyal, S.K. Tripathi, S. Tyagi, A. Sharma, P. Kumar, K.R. Ram, D.K. Chowdhuri, Y.Shukla, K.C. Gupta, In vitro and in vivo evaluation of linear polyethyleniminenanoparticles, J. Biomed. Nanotechnol. 7 (2011) 52–53.

[41] S.C. Lakhotia, T. Mukherjee, Specific activation of puff 93D of Drosophilamelanogaster by benzamide and the effect of benzamide treatment on the heatshock induced puffing activity, Chromosoma 81 (1980) 125–136.

[42] M.P. Singh, M.M. Reddy, N. Mathur, D.K. Saxena, D.K. Chowdhuri, Induction ofhsp70, hsp60, hsp83 and hsp26 and oxidative stress markers in benzene, tolueneand xylene exposed Drosophila melanogaster: role of ROS generation, Toxicol.Appl. Pharmacol. 235 (2009) 226–243.

[43] J.A. Royall,H. Ischiropoulos, Evaluationof 2′,7′-dichlorofluorescin anddihydrorhodamine123 as fluorescent probes for intracellular H2O2 in cultured endothelial cells, Arch.Biochem. Biophys. 302 (1993) 348–355.

[44] M. Nishikimi, N. Appaji, K. Yagi, The occurrence of superoxide anion in the reac-tion of reduced phenazine methosulfate and molecular oxygen, Biochem.Biophys. Res. Commun. 46 (1972) 849–854.

[45] A.K. Sinha, Colorimetric assay of catalase, Anal. Biochem. 47 (1972) 389–394.[46] W.H. Habig, M.J. Pabst, W.B. Jakoby, Glutathione S-transferases. The first en-

zymatic step in mercapturic acid formation, J. Biol. Chem. 249 (1974)7130–7139.

[47] G.L. Elmann, Tissue sulfhydryl groups, Arch. Bichem. Biophys. 82 (1959) 70–77.[48] H. Ohkawa, N. Ohishi, K. Yagi, Assay for lipid peroxides in animal tissues by

thiobarbituric acid reaction, Anal. Biochem. 95 (1979) 351–358.[49] R.L. Levine, D. Garland, C.N. Oliver, A. Amici, I. Climent, A.G. Lenz, B.W. Ahn, S.

Shaltiel, E.R. Stadtman, Determination of carbonyl content in oxidatively modi-fied proteins, Meth. Enzymol. 186 (1990) 464–478.

[50] O.H. Lowry, N.J. Rosebrough, A.L. Farr, R.J. Randall, Protein measurement with theFolin phenol reagent, J. Biol. Chem. 193 (1951) 265–275.

[51] M.V. Taussat, T. Camilli, Y. Aron, C. Meplan, P. Hainaut, B.S. Polla, B. Wekslar, Effectsof tobacco smoke and benzo (a) pyrene on human endothelial cell and monocytestress responses, Am. J. Physiol. Heart Circ. Physiol. 280 (2001) 1293–1300.

[52] S.C. Gupta, H.R. Siddique, N. Mathur, A.L. Vishwakarma, R.K. Mishra, D.K.Saxena, D.K. Chowdhuri, Induction of hsp70, alterations in oxidative stressmarkers and apoptosis against dichlorvos exposure in transgenic Drosophila

melanogaster: modulation by reactive oxygen species, Biochim. Biophys. Acta1770 (2007) 1382–1394.

[53] S. Novak, D. Drobne, A. Menard, Prolonged feeding of terrestrial isopod (Porcellioscaber, Isopoda, Crustacea) on TiO (2) nanoparticles. Absence of toxic effect,ZooKeys (2012) 261–273.

[54] Y. Li, L. Sun, M. Jin, Z. Du, X. Liu, C. Guo, Y. Li, P. Huang, Z. Sun, Size-dependentcytotoxicity of amorphous silica nanoparticles in human hepatoma HepG2cells, Toxicol. In Vitro 25 (2011) 1343–1352.

[55] M. Ahamed, R. Posgai, T.J. Gorey, M. Nielsen, S.M. Hussain, J.J. Rowe, Silvernanoparticles induced heat shock protein 70, oxidative stress and apoptosis inDrosophila melanogaster, Toxicol. Appl. Pharmacol. 242 (2010) 263–269.

[56] J.P. Wise Sr., B.C. Goodale, S.S. Wise, G.A. Craig, A.F. Pongan, R.B. Walter, W.D.Thompson, A.K. Ng, A.M. Aboueissa, H. Mitani, M.J. Spalding, M.D. Mason, Silvernanospheres are cytotoxic and genotoxic to fish cells, Aquat. Toxicol. 97 (2010) 34–41.

[57] X. Chen, Z. Zhong, Z. Xu, L. Chen, Y. Wang, 2′,7′-Dichlorodihydrofluorescein as afluorescent probe for reactive oxygen species measurement: forty years of appli-cation and controversy, Free Radic. Res. 44 (2010) 587–604.

[58] B. Kalyanaraman, V.D. Usmar, K.J.A. Davies, P.A. Dennery, H.J. Forman, M.B.Grisham, G.E. Mann, K. Moore, L.J. Roberts, H. Ischiropoulos, Measuring reactiveoxygen and nitrogen species with fluorescent probes: challenges and limitations,Free Radic. Biol. Med. 52 (2012) 1–6.

[59] C. Lin, J. Kang, R. Zheng, Vitamin K3 triggers human leukemia cell death throughhydrogen peroxide generation and histone hyperacetylation, Pharmazie 60(2005) 765–771.

[60] V. Calabrese, G. Scapagnini, A. Ravagna, R.G. Fariello, A.M. Giuffrida Stella, N.G.Abraham, Regional distribution of heme oxygenase, HSP70, and glutathione inbrain: relevance for endogenous oxidant/antioxidant balance and stress toler-ance, J. Neurosci. Res. 68 (2002) 65–75.

[61] D.I. De Pomerai, Transgenesis in animal systems: a view from within, Sci. Christ.Belief 8 (1996) 39–60.

[62] H.R. Kohler, G. Alberti, A. Seniezak, Lead induced hsp70 and hsp60 pattern trans-formation and leg malformation during postembryonic development in orbatidmite, Archegozetes longisetosus Aoki, Comp. Biochem. Physiol. C Toxicol.Pharmacol. 141 (2005) 398–405.

[63] Q. Wan, I. Whang, J. Lee, Molecular and functional characterization of HdHSP20: abiomarker of environmental stresses in disk abalone Haliotis discus discus, FishShellfish Immunol. 33 (2012) 48–59.

[64] G. Morrow, M. Samson, S. Michaud, R.M. Tanguay, Overexpression of the smallmitochondrial Hsp22 extends Drosophila life span and increases resistance to ox-idative stress, FASEB J. 18 (2004) 598–599.

[65] D.S. Kim, B.K. Jeon, Y.E. Lee, W.H. Woo, Y.J. Mun, Diosgenin induces apoptosis inHepG2 cells through generation of reactive oxygen species and mitochondrialpathway, Evid. Based Complement. Alternat. Med. (2012) (Epub).

[66] F. Zhao, Q. Wang, The protective effect of peroxiredoxin II on oxidative stressinduced apoptosis in pancreatic beta-cells, Cell Biosci. 22 (2012), http://dx.doi.org/10.1155/2012/981675 (Epub).

[67] H.M. Beere, “The stress of dying”: the role of heat shock proteins in the regulationof apoptosis, J. Cell Sci. 117 (2004) 2641–2651.

[68] A.A. Shemetov, A.S. Seit-Nebi, N.B. Gusev, Structure, properties, and functions ofthe human small heat-shock protein HSP22 (HspB8, H11, E2IG1): a critical re-view, J. Neurosci. Res. 86 (2008) 264–269.

[69] T.G. Iversen, T. Skotland, K. Sandvig, Endocytosis and intracellular transport ofnanoparticles: present knowledge and need for future studies, Nano Today 6(2011) 176–185.

[70] E. Besic Gyenge, X. Darphin, A. Wirth, U. Pieles, H. Walt, M. Bredell, C. Maake, Up-take and fate of surface modified silica nanoparticles in head and neck squamouscell carcinoma, J. Nanobiotech. 9 (2011) 32–45.

[71] M.S. Cartiera, K.M. Johnson, V. Rajendran, M.J. Caplan, W.M. Saltzman, The uptakeand intracellular fate of PLGA nanoparticles in epithelial cells, Biomaterials 30(2009) 2790–2798.

[72] O. Faklaris, V. Joshi, T. Irinopoulou, P. Tauc, M. Sennour, H. Girard, C. Gesset, J.C.Arnault, A. Thorel, J.P. Boudou, P.A. Curmi, F. Treussart, Photoluminescent dia-mond nanoparticles for cell labeling: study of the uptake mechanism in mamma-lian cells, ACS Nano 3 (2009) 3955–3962.

[73] L. Sun, Y. Li, X. Liu, M. Jin, L. Zhang, Z. Du, C. Guo, P. Huang, Z. Sun, Cytotoxicity andmitochondrial damage caused by silica nanoparticles, Toxicol. In Vitro 25 (2011)1619–1629.

[74] T.P. Cash, Y. Pan, M.C. Simon, Reactive oxygen species and cellular oxygen sens-ing, Free Radic. Biol. Med. 43 (2007) 1219–1225.

[75] A.J. Kowaltowski, N.C. de Souza-Pinto, R.F. Castilho, A.E. Vercesi, Mitochondriaand reactive oxygen species, Free Radic. Biol. Med. 47 (2009) 333–343.


Recommended