ArticlePDF Available

Rat brain pro-oxidant effects of peripherally administered 5nm ceria 30 days after exposure

Authors:

Abstract

The objective of this study was to determine the residual pro-or anti-oxidant effects in rat brain 30 days after systemic administration of a 5nm citrate-stabilized ceria dispersion. A ∼4% aqueous ceria dispersion was iv-infused (0 or 85mg/kg) into rats which were terminated 30 days later. Ceria concentration, localization, and chemical speciation in the brain was assessed by inductively coupled plasma mass spectrometry (ICP-MS), light and electron microscopy (EM), and electron energy loss spectroscopy (EELS), respectively. Pro- or anti-oxidant effects were evaluated by measuring levels of protein carbonyls (PC), 3-nitrotyrosine (3NT), and protein-bound-4-hydroxy-2-trans-nonenal (HNE) in the hippocampus, cortex, and cerebellum. Glutathione reductase (GR), glutathione peroxidase (GPx), superoxide dismutase (SOD), and catalase levels and activity were measured in addition to levels of inducible nitric oxide (iNOS), and heat shock protein-70 (Hsp70). The blood brain barrier (BBB) was visibly intact and no ceria was seen in the brain cells. Ceria elevated PC and Hsp70 levels in hippocampus and cerebellum, while 3NT and iNOS levels were elevated in the cortex. Whereas glutathione peroxidase and catalase activity were decreased in the hippocampus, GR levels were decreased in the cortex, and GPx and catalase levels were decreased in the cerebellum. The GSH:GSSG ratio, an index of cellular redox status, was decreased in the hippocampus and cerebellum. The results are in accordance with the observation that this nanoscale material remains in this mammal model up to 30 days after its administration and the hypothesis that it exerts pro-oxidant effects on the brain without crossing the BBB. These results have important implications on the potential use of ceria ENM as therapeutic agents.
):5A1=>5?D;2 1:?@/7D):5A1=>5?D;2 1:?@/7D
) :;B81031) :;B81031
4195>?=D-/@8?D%@.85/-?5;:> 4195>?=D

&-?=-5:%=;$C50-:?221/?>;2%1=5<41=-88D095:5>?1=10:9&-?=-5:%=;$C50-:?221/?>;2%1=5<41=-88D095:5>?1=10:9
/1=5--D>2?1=C<;>@=1/1=5--D>2?1=C<;>@=1
'-=5?-'-=0->
):5A1=>5?D;2 1:?@/7D
>-=5?->439-58/;9
&@74>-:-'@8?-:-
):5A1=>5?D;2 1:?@/7D
=>@8?@7D10@
;A5:0*-==51=
):5A1=>5?D;2 1:?@/7D
";-:
):5A1=>5?D;2 1:?@/7D
9;0-:@7D10@
&1.1//-!8;=1:/1
):5A1=>5?D;2 1:?@/7D
'11:1C?<-312;=-005?5;:-8-@?4;=>
;88;B?45>-:0-005?5;:-8B;=7>-?4??<>@7:;B81031@7D10@/4195>?=D,2-/<@.
%-=?;2?414195/-8:35:11=5:3;99;:>4195>?=D;99;:>%4-=9-/;8;3D(;C5/;8;3D-:0
:A5=;:91:?-81-8?4;99;:>-:0?41%4-=9-/D-:0%4-=9-/1@?5/-8'/51:/1>;99;:>
&534?/85/7?;;<1:-2110.-/72;=95:-:1B?-.?;81?@>7:;B4;B?45>0;/@91:?.1:1G?>D;@&534?/85/7?;;<1:-2110.-/72;=95:-:1B?-.?;81?@>7:;B4;B?45>0;/@91:?.1:1G?>D;@
&1<;>5?;=D5?-?5;:&1<;>5?;=D5?-?5;:
-=0->'-=5?-''@8?-:-&@74>-:-*-==51=;A5:0-:";8;=1:/1&1.1//-!*@%1:3=@871
=5/(>1:3"5/4-18():=5:1->;:"=-4-9)>/45"+;718&;.1=?-:0@??1=G18088-:
&-?=-5:%=;$C50-:?221/?>;2%1=5<41=-88D095:5>?1=10:9/1=5--D>2?1=C<;>@=1
4195>?=D-/@8?D%@.85/-?5;:>

4??<>@7:;B81031@7D10@/4195>?=D,2-/<@.
(45>=?5/815>.=;@34??;D;@2;=2=11-:0;<1:-//1>>.D?414195>?=D-?) :;B81031?4->.11:-//1<?102;=
5:/8@>5;:5:4195>?=D-/@8?D%@.85/-?5;:>.D-:-@?4;=5E10-095:5>?=-?;=;2) :;B81031;=9;=15:2;=9-?5;:
<81->1/;:?-/?) :;B810318>A@7D10@
&-?=-5:%=;$C50-:?221/?>;2%1=5<41=-88D095:5>?1=10:9/1=5--D>&-?=-5:%=;$C50-:?221/?>;2%1=5<41=-88D095:5>?1=10:9/1=5--D>
2?1=C<;>@=12?1=C<;>@=1
535?-8$.61/?01:?5G1=$
4??<>0;5;=36:1@=;
#;?1>5?-?5;::2;=9-?5;:#;?1>5?-?5;::2;=9-?5;:
%@.85>4105:
#1@=;(;C5/;8;3D
A5>>@1
;<D=534?F8>1A51=:/
F(45>9-:@>/=5<?A1=>5;:5>9-01-A-58-.81@:01=?41+##85/1:>1
4??<>/=1-?5A1/;99;:>;=385/1:>1>.D:/:0
(410;/@91:?-A-58-.812;=0;B:8;-05>?41-@?4;=><;>?<11==1A51BG:-80=-2?;2?41-=?5/81
@?4;=>@?4;=>
'-=5?-'-=0->&@74>-:-'@8?-:-;A5:0*-==51=";-:&1.1//-!8;=1:/1%1:3*@=5/=@871
"5/4-18((>1:3->;:"):=5:1)>/45"=-4-9&;.1=?+;718-:088-:@??1=G180
(45>-=?5/815>-A-58-.81-?) :;B810314??<>@7:;B81031@7D10@/4195>?=D,2-/<@.
1
Rat brain pro-oxidant effects of peripherally administered 5 nm ceria 30 days after 1
exposure 2
3
Sarita S. Hardas*, Rukhsana Sultana*, Govind Warrier*, Mo Dan, Rebecca L. Florence, Peng 4
Wulll, Eric A. Grulkelll, Michael T. Tsengllll, Jason M. Unrine, Uschi M. Graham||, Robert A. 5
Yokel, § and, D. Allan Butterfield *, † 6
7
Sarita S. Hardas: *Department of Chemistry, University of Kentucky, Lexington, Kentucky 8
40506-0055. Email: Sarita.Hardas@uky.edu 9
Rukhsana Sultana: *Department of Chemistry, University of Kentucky, Lexington, Kentucky 10
40506-0055. Email: rsult2@uky.edu 11
Govind Warrier: *Department of Chemistry, University of Kentucky, Lexington, Kentucky 40506-12
0055. Email: govind.warrier@uky.edu 13
Mo Dan: ‡Department of Pharmaceutical Sciences, University of Kentucky Academic Medical 14
Center, University of Kentucky, Lexington, Kentucky 40536-0082. Email: mo.dan@uky.edu 15
Rebecca L. Florence: ‡Department of Pharmaceutical Sciences, University of Kentucky 16
Academic Medical Center, University of Kentucky, Lexington, Kentucky 40536-0082. Email: 17
rlstep2@uky.edu 18
Peng Wu: IIIChemical and Materials Engineering Department, University of Kentucky, 19
Lexington, Kentucky 40506-0503. Email: peng.wu@uky.edu 20
Eric Grulke: IIIChemical and Materials Engineering Department, University of Kentucky, 21
Lexington, Kentucky 40506-0503. Email: eric.grulke@uky.edu 22
Michael T. Tseng: IIIIDepartment of Anatomical Sciences & Neurobiology, University of 23
Louisville, Louisville, Kentucky 40202. Email: mttsen01@louisville.edu 24
Jason M. Unrine: ¶Department of Plant and Soil Sciences, University of Kentucky, Lexington, 25
Kentucky 40546-0091. Email: jason.unrine@uky.edu 26
*Manuscript
Click here to view linked References
2
Uschi M. Graham: IICenter for Applied Energy Research, University of Kentucky, Lexington, 1
Kentucky 40511. Email: uschi.graham@uky.edu 2
Robert A. Yokel: ‡Department of Pharmaceutical Sciences, University of Kentucky Academic 3
Medical Center, University of Kentucky, Lexington, Kentucky 40536-0082, § Graduate Center for 4
Toxicology, University of Kentucky Academic Medical Center, Lexington, Kentucky 40506-9983. 5
Email: ryokel@email.uky.edu 6
D. Allan Butterfield: *Department of Chemistry, University of Kentucky, Lexington, Kentucky 7
40506-0055, †Center of Membrane Sciences, University of Kentucky, Lexington, Kentucky 8
40506-0059. Email: dabcns@uky.edu 9
10
11
Corresponding author: 12
D. Allan Butterfield, Ph.D. 13
Department of Chemistry, 14
University of Kentucky, 15
Lexington, 16
Kentucky 40506-0055, 17
Phone: 859-257-3184 18
Fax: 859-323-1069 19
Email: dabcns@uky.edu 20
21
Short title: Nanoceria pro-oxidant effect on brain 22
23
24
25
26
3
Abbreviations: 1
3NT protein bound 3-nitrotyrosine 2
Ce cerium 3
Cat catalase 4
EELS electron energy loss spectroscopy 5
ENM engineered nanomaterial 6
GR glutathione reductase 7
GPx glutathione peroxidase 8
H2O2 Hydrogen peroxide 9
Hsp70 heat shock protein 70 10
HNE protein-bound 4-hydroxy-2-trans-nonenal 11
ICP-MS inductively coupled plasma mass spectrometry 12
iNOS inducible nitric oxide synthase 13
MDL method detection limit 14
PC protein carbonyl 15
ROS reactive oxygen species 16
RNS reactive nitrogen species 17
SOD superoxide dismutase 18
TEM transmission electron microscopy 19
20
21
22
4
Abstract: 1
The objective of this study was to determine the residual pro-or anti-oxidant effects in rat brain 2
30 days after systemic administration of a 5 nm citrate-stabilized ceria dispersion. A ~4% 3
aqueous ceria dispersion was iv-infused (0 or 85 mg/kg) into rats which were terminated 30 4
days later. Ceria concentration, localization, and chemical speciation in the brain was assessed 5
by inductively coupled plasma mass spectrometry (ICP-MS), light and electron microscopy 6
(EM), and electron energy loss spectroscopy (EELS), respectively. Pro- or anti-oxidant effects 7
were evaluated by measuring levels of protein carbonyls (PC), 3-nitrotyrosine (3NT), and 8
protein-bound-4-hydroxy-2-trans-nonenal (HNE) in the hippocampus, cortex, and cerebellum. 9
Glutathione reductase (GR), glutathione peroxidase (GPx), superoxide dismutase (SOD), and 10
catalase levels and activity were measured in addition to levels of inducible nitric oxide (iNOS), 11
and heat shock protein-70 (Hsp70). The blood brain barrier (BBB) was visibly intact and no ceria 12
was seen in the brain cells. Ceria elevated PC and Hsp70 levels in hippocampus and 13
cerebellum, while 3NT and iNOS levels were elevated in the cortex. Whereas glutathione 14
peroxidase and catalase activity were decreased in the hippocampus, GR levels were 15
decreased in the cortex, and GPx and catalase levels were decreased in the cerebellum. The 16
GSH: GSSG ratio, an index of cellular redox status, was decreased in the hippocampus and 17
cerebellum. The results are in accordance with the observation that this nanoscale material 18
remains in this mammal model up to 30 days after its administration and the hypothesis that it 19
exerts pro-oxidant effects on the brain without crossing the BBB. These results have important 20
implications on the potential use of ceria ENM as therapeutic agents. 21
22
Key words: oxidative stress; ceria; brain; neurotoxicity; nanomaterial, nanoparticles; rat. 23
24
25
5
Introduction: 1
Engineered nanomaterials (ENM) can be manufactured in a variety of shapes and sizes and 2
physico-chemical, surface, as well as optical and magnetic properties. ENMs have numerous 3
applications in research, medicine, electronics and other industries. Physico-chemical properties 4
of nanomaterials differ from their bulk forms mainly because of the larger surface area to mass 5
ratio, which affects reactivity, strength and electrical properties of nanomaterials. Because of 6
their comparable size with biological molecules like proteins and DNA, ENMs can gain access 7
to usually difficult to reach biological compartments in cells (Fubini et al. 2010). Increased 8
surface activity can facilitate interactions with biological molecules, which may evoke greater 9
physiological responses, different from the same basic material with larger particle size, the bulk 10
form equivalent of ENMs (Donaldson et al. 2004; Landsiedel et al. 2009; Xia et al. 2009). One 11
effect exhibited by ENMs is the generation of free radicals or induction of oxidative stress, which 12
is also a primary mechanism of ENM toxicity (Xia et al. 2009). Oxidative stress effects are direct 13
consequences of imbalance in the rates of reactive oxygen and / or nitrogen species (ROS or 14
RNS) production verses scavenging of ROS and / or RNS and / or antioxidant levels 15
(Butterfield et al. 2007). 16
17
Ceria ENM (a.k.a. cerium oxide; CeO2), which is one of the most used ENM employed in 18
different industrial applications (Yokel et al. 2009; Hardas et al. 2010) has been shown to have 19
both anti-inflammatory properties as well as potent toxicity. However, there is no clear 20
understanding of what exactly controls ceria’s pro-or anti-oxidant effects. A recently published 21
report summarizes findings of in vitro and in vivo studies conducted with ceria ENM under basal 22
and induced oxidative stress conditions (Celardo et al. 2011). Ceria exhibited antioxidant 23
properties evidenced by scavenging free radicals, by reducing levels of peroxides, iNOS, TNF-24
α, NF-ĸβ, and interleukin, by promoting cell viability or protecting organelles from diesel exhaust 25
and cigarette smoke-induced oxidative stress, ROS generating chemical agents, or side effects 26
6
of radiation treatment. Ceria has been suggested for potential use in the treatment of diabetic 1
cardiomyopathy, cancer, stroke, retinal degradation and Alzheimer’s disease as well as to 2
prolong life span (Chen et al. 2006; Rzigalinski et al. 2006; Das et al. 2007; Xia et al. 2008; 3
D'Angelo et al. 2009; Hirst et al. 2009; Babu et al. 2010; Colon et al. 2010; Younce et al. 2010; 4
Estevez et al. 2011; Niu et al. 2011). Antioxidant properties of ceria may be related to its SOD- 5
and catalase-mimicking activity (Korsvik et al. 2007; Pirmohamed et al. 2010) attributed to 6
Ce3+/ Ce4+ redox coupling (Celardo et al. 2011). In contrast, there are reports of ceria induced 7
pro-oxidant effects under basal conditions. In different cell culture studies, ceria ENM mediated 8
ROS injury, induced lipid peroxidation, caused membrane damage, led to elevation of the 9
cytokine, IL-8, led to depletion of GSH, and led to reduced cell viability (Brunner et al. 2006; Lin 10
et al. 2006; Park et al. 2008; Auffan et al. 2009). 11
12
To utilize ceria for therapeutic and non-therapeutic applications, it is important to know the long 13
term effects of intended and un-intended ceria exposure on mammals. Most reports on effects 14
of ceria ENM were conducted using non-mammalian organisms or in cell culture, and none of 15
these addressed long-term effects or fate of ceria. In addition to our own previous studies 16
(Yokel et al. 2009; Hardas et al. 2010), a few ceria ENM studies were conducted in intact 17
animals (Chen et al. 2006; Niu et al. 2007; Hirst et al. 2009; Amin et al. 2011; Choi et al. 2011; 18
Hirst et al. 2011; Srinivas et al. 2011; Zhou et al. 2011). One study reports that deposition and 19
retention of ceria in various vital organs and increased WBC count were seen 30 days after 20
intraperitoneal and intravenous injection to mice, but otherwise ceria was tolerated by animals 21
(Hirst et al. 2011). Ceria reduced myocardial oxidative stress in transgenic mice for ischemic 22
cardiomyopathy, rat liver from monocrotaline-induced ROS injury by induction of GSH levels 23
and intravitreal injections of ceria inhibited retinal neovascular lesions (Niu et al. 2007; Amin et 24
al. 2011; Zhou et al. 2011). However, after pulmonary inhalation of ceria ENM, granulomatous 25
pathology and GSH depletion were seen in rat lungs (Cho et al. 2010; Srinivas et al. 2011). 26
7
Animal studies have also reported that ceria ENM can accumulate in various organs, including 1
the heart and lung, irrespective of the point of entry or distance (from injection point and organ 2
specifically examined) when supplied as intravenous or intra-peritoneal injections or as a food 3
additive (Chen et al. 2006; Niu et al. 2007; Hirst et al. 2009). This accumulation may lead to 4
systemic effects involving the inflammatory response (Celardo et al. 2011) or increased ROS 5
production under normal physiological conditions (Hirst et al. 2011). 6
7
To our knowledge there is no prior information available on the long-term effects (30 d or more 8
after administration) of ceria on brain and how these effects may contrast with an immediate 9
response after the initial ENM contact. Our previous study showed moderate pro-oxidant effects 10
on rat brain, 1 and 20 h after a single acute systemic instillation of 5 nm ceria ENM (Hardas et 11
al. 2010). The current study discusses residual effects of oxidative stress parameters in brain 30 12
days after one single acute ENM peripheral administration using 5 nm ceria ENM. To address 13
the objective, the levels and activities of the antioxidant enzymes catalase, manganese 14
superoxide dismutase (Mn-SOD), glutathione peroxidase (GPx), and glutathione reductase 15
(GR), were measured along with the ratio of reduced glutathione (GSH) to its oxidized form 16
(GSSG). To understand the extent of changes in cellular redox status, the levels of oxidative 17
stress endpoints, protein carbonyl (PC), 3-nitrotyrosine (3NT), and protein bound 4- hydroxyl-2-18
trans nonenal (HNE), were measured along with heat shock protein (Hsp70) levels. 19
20
Materials and Methods: 21
All the materials, methods including the well characterized 5 nm ceria ENM are same as that 22
used in our recently published study (Hardas et al. 2010). The rats used are of the same strain, 23
sex and approximately same weight as that used in the previous study with 5 nm ceria ENM 24
(Hardas et al. 2010). Therefore, only a brief overview is presented. 25
26
8
Nanomaterial: 1
Cerium chloride heptahydrate (Sigma-Aldrich # 228931, 99.9% metal basis), ammonium 2
hydroxide (Fisher #3256, ACS, 28-30%) and citric acid monohydrate (EMD Chemicals Inc # 3
CX1725-1, GR ACS) were used without further purification. A hydrothermal method was used to 4
synthesize ~5nm ceria aqueous suspension. Briefly, a 20 ml aqueous mixture of 0.01 mol 5
cerium chloride and 0.01 M citric acid was added to 20 ml of 3 M ammonium hydroxide. After 6
stirring for 24 h at 50C, the solution was transferred into a Teflon-lined stainless steel bomb 7
and heated at 80C for 24 h to complete the reaction. The final dispersion of ceria ENM was 8
infused intravenously to the rats over 1 h without any further treatment or purification. 9
10
Ceria characterization: 11
The details of ceria ENM characterization are published in our earlier study (Hardas et al. 2010). 12
In brief, the morphology and crystallinity of the ceria was evaluated using a 200-keV field 13
emission analytical transmission electron microscope (JEOL JEM-2010F, Tokyo, Japan) 14
equipped with an Oxford energy dispersive X-ray spectrometer. The particle size distributions 15
(PSDs) were determined using dynamic light scattering (DLS: Brookhaven Instruments 16
Corporation, 90Plus NanoParticle Size Distribution Analyzer, Holtsville, NY, USA). The surface 17
area of the dried ceria powder was determined using a BETASAP 2020 surface area analyzer 18
that determines particle surface area based on nitrogen adsorption (Micromeritics Instrument 19
Corporation, Norcross, GA, USA). The ceria content of the dispersion and the potential 20
presence of contaminating elements/metals were determined by digestion of ceria dispersion 21
samples and analysis by ICP-MS. Electron energy loss spectroscopy (EELS) was performed on 22
rat liver tissue using a JEOL 2010F STEM outfitted with a URP pole piece, GATAN 2000 GIF 23
(Pleasanton, CA, USA), GATAN DigiScan II, Fischione HAADF STEM detector (Export, PA, 24
USA), and EmiSpec EsVision software (Tempe, AZ, USA). 25
9
Animals: 1
In the current study 16 male Sprague Dawley rats, weighing 328 ± 21 g (mean ± SD), were 2
obtained from Harlan, Indianapolis, IN, and were housed individually prior to study and after 3
cannulae removal (a few days after the iv infusion) in the University of Kentucky Division of 4
Laboratory Animal Resources facility under a 12:12 h light:dark cycle at 70 ± 8°F and 30 to 70% 5
humidity. The rats had ad lib access to 2018 Harlan diet and RO water. All procedures involving 6
animals were approved by the University of Kentucky Institutional Animal Care and Use 7
Committee. The research was conducted in accordance with the Guiding Principles in the Use 8
of Animals in Toxicology (http://www.toxicology.org/ai/air/air6.asp).. 9
10
Ceria administration: 11
Rats were prepared with 2 cannulae that terminated in the vena cava to administer the ceria 12
dispersion or water (controls) in one and 1.8% sodium chloride in the second, to avoid 13
agglomeration induced by sodium chloride or 10% sucrose, which are commonly added to 14
prepare iso-osmotic solutions (Yokel et al. 2009); (Hardas et al. 2010). Seven rats received 0 15
and nine rats received 85 mg ceria/kg as a single acute dose of ENM. The rats were terminated 16
30 days after the infusion. Post-mortem samples including brain cortex, hippocampus, and 17
cerebellum were harvested rapidly and were frozen in liquid nitrogen and stored at -80 0C for 18
later oxidative stress measurements. Cerium concentrations in brain cortex, blood, and liver 19
were analyzed by ICP-MS; Agilent 7500cx, (Santa Clara, CA, USA) after microwave digestion 20
as described previously (Hardas et al., 2010). The detection limits of this method were 0.089 mg 21
Ce/kg. Mean spike recovery ranged from 97 to 105%. Relative percent difference between 22
replicates was ≤ 2% for 30-120 ng Ce/ml and 18% for 1.5 ng Ce/ml. 23
Light and electron microscopic assessment: 24
After animal termination brains were removed, sliced coronally in a brain matrix (Braintree 25
acrylic matrix BS-A 6000C). Sections containing hippocampus were fixed by immersion in 4% 26
10
buffered formalin. Cerebellum was similarly sliced and fixed. Samples were cut into 3 mm 1
cubes, dehydrated and embedded in Araldite 502. After polymerization, one micron thick 2
sections were cut and stained with toluidine blue for LM examination. Selected blocks were 3
sectioned at 80 nm, collected on 200 mesh copper grids and examined in a Philips CM 10 4
electron microscope at 60 kV. 5
6
Oxidative stress assessment, sample preparation: 7
All sample preparation and protein assays were carried out as described in our previous 8
publication (Hardas et al. 2010). Each sample was individually thawed and homogenized using 9
a 550 sonic dismembrator from Fischer Scientific for 10 to 20 s on ice. The buffer used for 10
homogenization contained 0.32 M sucrose, 0.125 M Tris, 0.6 mM MgCl2, and protease inhibitors 11
(4 µg/ul leupeptin, 4 µg/ul pepstatin A, 5 µg/ul aprotinin, and 0.2 mM phenylmethylsulfonyl 12
fluoride) at pH 8.0. Total protein concentration for each sample was measured using the 13
bicinchoninic acid assay and equal amounts of protein from control and treated samples were 14
used in each assay. 15
16
Oxidative stress markers: 17
The oxidative stress markers PC, 3NT, and HNE were assessed for each homogenized sample 18
using the slot blot technique. Specific antibodies were used to determine the levels of PC, 3NT, 19
HNE in controls and treated samples as previously described (Hardas et al. 2010). 20
21
GSH and GSSG levels: 22
The reduced (GSH) and oxidized glutathione (GSSG) levels were simultaneously measured in 23
each sample as previously described (Hissin and Hilf 1976). A small amount of brain tissue was 24
rapidly weighed, homogenized with metaphosphoric acid (25%) and sodium phosphate (0.1 M) - 25
ethylenediaminetetraacetic acid (0.005 M) buffer (pH 8) and then centrifuged. For GSH levels, 26
11
an aliquot of supernatant was further diluted with phosphate buffer and then incubated with o-1
phthaldehyde (OPT), before determination of fluorescence (λ excitation 350 nm and emission 2
420 nm). For GSSG levels equal volumes of supernatant were incubated with N-ethylmaleimide 3
(0.04 M) for 30 min and then diluted with sodium hydroxide (0.1 N), before assaying with OPT. 4
The GSH to GSSG ratio for each sample was calculated by comparing the fluorescence values 5
from each assay to their respective calibration curves. The final values are % control of mean ± 6
SEM of treated vs. control samples. 7
8
Western blot analysis: 9
The levels of the antioxidant enzymes GR, GPx, MnSOD, catalase, Hsp70 and iNOS were 10
measured using immunoblotting-Western blot techniques as described in our earlier 11
publications (Sultana et al. 2008; Hardas et al. 2010). In brief, 75 µg protein from each 12
homogenized sample was loaded and separated on SDS-PAGE alongside its respective 13
control. The separated proteins were transferred from poly-acrylamide gels to nitrocellulose 14
membranes, and then the band of specific protein identified using a specific antibody against 15
that protein. The band-intensity was quantified as previously described (Hardas et al., 2010) by 16
using the image analysis software, ImageQuant, purchased from GE Healthcare. 17
18
Enzyme activity assays: 19
The activities of GR, GPx, MnSOD and catalase enzymes were measured as described earlier 20
(Hardas et al. 2010), with commercially available kits from Cayman Chemical Company, Ann 21
Arbor, MI, USA as per the manufacturer’s instructions. Suitable protein samples from each 22
organ (i.e., 10-20 µg for brain homogenate) were mixed with assay buffer and with other specific 23
reagents for each assay on 96-well plates. Enzymatic reactions were initiated by addition of 24
reaction initiator reagents, NADPH for GR, cumene hydroperoxide for GPx, xanthine oxidase for 25
SOD, and hydrogen peroxide for the catalase assay. Progression of the each reaction was 26
12
studied separately by spectrophotometry at 340 nm (for GR and GPx), at 460 nm (for SOD) and 1
at 540 nm (for catalase). 2
Data and statistical analysis: 3
The slot blot, Western blot and enzyme assay results are presented as mean ± SEM. The 4
control mean was normalized to 100%. Grubb’s test was used to identify outliers in oxidative 5
stress parameter results. Student’s unpaired t-test was used to evaluate significant difference 6
between controls and ceria treated samples. Subsequently, two-way ANOVA was conducted to 7
determine the differential effect of ceria ENM treatment among the three brain regions studied. 8
Significance was accepted at p < 0.05. 9
10
Results: 11
Ceria composition 12
HR-TEM/HR-STEM showed the ceria ENM was polyhedral shaped (Figure 1). The XRD 13
patterns demonstrated the ceria was highly crystalline, with face centered cubic unit cells with 14
corresponding Miller indices of the most common faces of (111), (210) and (200). Evaluation of 15
a number of TEM images showed that the ceria had a number-average primary particle size of 16
~5 nm. The ceria ENM surface area was 121 m2/g, assuming the density of a ceria nanoparticle 17
7600 kg/m3 (the bulk density), the back-calculated average diameter should be 6.5nm that fits 18
our TEM observation quite well. Analysis by ICP-MS of the ceria dispersion used showed that 19
the ceria content of the as-synthesized dispersion was 4.35 ± 0.20% and the free cerium ion 20
content was 11.6 ± 0.3%. 21
Dispersion properties 22
The citrate ion was found to be a stabilizing electrolyte for ceria nanoparticles. At pH below 7.0 23
the ceria agglomerated; therefore, the ceria was maintained as an aqueous dispersion at pH 7.7 24
to 8. Zeta potential of the citrate-stabilized particles was -53 ± 7 mV at pH ~ 7.35 (Hardas et al. 25
2010). In general, dispersions with absolute values of zeta potential greater than 30 to 40 mV 26
13
are expected to be stable, and the stability of dispersion is better with higher zeta potential. The 1
cerium concentration in samples taken from the top and bottom of two ceria dispersion samples 2
in covered 15 ml centrifuge tubes that were un-disturbed for > 2 months were within 2.5% of 3
each other, demonstrating dispersion stability (data not shown). 4
5
Ceria concentration in brain and electron micrography 6
ICP-MS analysis showed that a very small amount of ceria was present in the brain compared 7
to the liver (Table 1). Electron micrographic studies suggested that ceria ENM was not present 8
in the brain, but located on the luminal side of the blood brain barrier (BBB) endothelium. The 9
hippocampus and cerebellum tissues did not show obvious ceria induced injury as no necrotic 10
neurons or elevated gliosis were observed and the BBB was visibly intact (data not shown). 11
12
EELS results 13
Electron energy loss spectroscopic (EELS) measurements on liver tissue were performed as a 14
representative organ. The 5 nm ceria agglomerates were located in the tissue 30 days after 15
infusion into rats. The ratio of Ce(3+) to Ce(4+) was evaluated using EELS measurements in 16
vivo 30 days post infusion. Further, this ratio was compared with the ratio of Ce3+/Ce4+ obtained 17
in freshly synthesized ceria. The high Ce3+/ Ce4+ ratio that was obtained in the as-synthesized, 18
fresh 5 nm ceria particles seems to have only been altered slightly in individual ceria measured 19
in liver after 30 days in vivo. However, this difference was not significant (data not shown). 20
21
Oxidative Stress Indices 22
The primary aim of these studies was to evaluate indices of oxidative stress in brain 30 days 23
following a single ceria ENM administration. 24
25
Ceria treatment affected catalase levels and activities 26
14
Previously ceria ENM was reported to have a H2O2-producing ability (Korsvik et al. 2007); 1
therefore, the effect of 5 nm ceria ENM on levels and activity of a primary H2O2-reducing 2
enzyme (catalase) was determined in the present study. Catalase activity was significantly 3
decreased in the hippocampus (~18%, *p<0.05, Figure 2b) and catalase levels were 4
significantly decreased in the cerebellum (~16%, *p<0.05, Figure 4a). To determine the 5
influence of ceria treatment on the SOD enzyme or contribution to H2O2 levels from SODs, if 6
any, the level of MnSOD and its activity were measured. There were no significant changes 7
observed in the levels or activity of MnSOD (data not shown). 8
9
Ceria treatment decreases GPx levels, activity and the GSH-GSSG ratio 10
GPx reduces H2O2 along with other peroxides using glutathione (GSH) as a source of reducing 11
equivalents. The activity of GPx was significantly decreased in the hippocampus (~69%, 12
**p<0.001, Figure 2b) and in cerebellum (~23 %, *p<0.05, Figure 4b). GPx levels showed a 13
decreasing trend in all brain regions, but was significantly decreased only in cerebellum (~27 %, 14
*p<0.05, Figure 4a). Within the three brain regions examined, hippocampal GPx activity was 15
significantly inhibited compared to that in cortex and cerebellum (**p < 0.01). GR levels were 16
significantly decreased (~24%, *p<0.05, Figure 3a) in the cortex. GR activity did not show any 17
significant change in any of the three brain regions studied. A marker of overall cellular redox 18
status was evaluated by comparing the GSH:GSSG ratio of ceria-treated samples. The GSH: 19
GSSG ratio was significantly decreased in the hippocampus (~13%, *p<0.05, Figure 2b) and 20
cerebellum (~15%, **p<0.01, Figure 4b) consistent with an increase in oxidative stress. 21
22
Ceria treatment induced protein oxidation 23
The levels of PC showed a significant increase in the hippocampus (~ 19%, *p<0.05, Figure 2c) 24
and cerebellum (~12 %, *p<0.05, Figure 4c) in treated vs. control samples. 3NT levels were 25
significantly increased in the cortex (~20%, *p<0.05, Figure 3c). There was no significant 26
15
change in protein-bound HNE levels in any of the three brain regions examined. Consistent with 1
increased 3NT levels in cortical region, iNOS levels were increased significantly (~27%, 2
*p<0.05, Figure 3d), and there was a positive correlation between 3NT and iNOS levels (r= 3
0.67, p < 0.05, Figure 3d). 4
5
Hsp-70 levels increased after ceria treatment 6
Hsp-70 is a member of the heat shock protein family and inducible by oxidative stress. Hsp-70 7
levels were significantly increased in the hippocampus (~49%, *p<0.05, Figure 2a), as well as 8
compared to that of cortical and cerebellar Hsp70 levels (*p<0.05). In the cerebellum Hsp70 9
levels were increased (~40% *p<0.05, Figure 4a). These results are consistent with diminution 10
of GSH levels indicated by the decreased GSH: GSSG ratio in these brain regions. 11
12
Discussion: 13
The present work was conducted to evaluate the oxidative stress effects of ceria ENM in brain 14
30 d after a single acute peripheral administration of 5 nm ceria ENM. ROS and RNS are 15
inevitable byproducts of all major metabolic cellular processes and conspicuous by their high 16
reactivity and high cytotoxicity. Apart from being a cause for many pathological conditions or 17
cellular disturbances, ROS/ RNS also play important roles in cell signaling pathways and 18
cellular homeostasis (Celardo et al. 2011; Butterfield et al., 2001). Therefore, balancing 19
excessive and insufficient ROS/RNS is critical, and endogenous enzymes like GPx, GR, 20
catalase and SOD as well as antioxidants like glutathione maintain this redox balance efficiently. 21
Any perturbation to this redox balance causes oxidative stress. One of the mechanisms by 22
which nanomaterials induce toxicity is by inducing ROS production, elevating oxidative stress, 23
which damages proteins, lipids or DNA (Butterfield and Stadtman, 1997; Nel et al. 2006; 24
Sharma Sharma 2007; Mocan et al. 2010). 25
16
Owing to its redox switching between two oxidation states, 3+ and 4+, ceria exhibits catalytic 1
activity, which has made it useful in industrial applications (Celardo et al. 2011). Ceria’s surface 2
redox capability can affect the immediate surroundings and has been strongly linked to particle 3
size (Gilliss et al. 2005). The characteristic oxidation and reduction in ceria ENM is linked to the 4
continued possibility to absorb and release oxygen by inducing oxygen vacancies close to the 5
particle surface. The Ce3+/ Ce4+ valance switch resembles redox behavior of some biological 6
antioxidant enzymes like SOD and catalase. The high Ce3+/ Ce4+ ratio is responsible for the 7
ceria ENM SOD-mimetic activity (Korsvik et al. 2007). 8
In our previous study, the same 5 nm ceria ENM at the same dose as used in the present study 9
showed moderate effects, on brain redox status, while catalase levels and activities were 10
increased in hippocampus after 1 and 20 h, respectively, and catalase activity was decreased in 11
cerebellum after 1 h. No change was seen in PC, 3NT and HNE levels (Hardas et al. 2010). 12
Although, ceria ENM was not found in the brain, but located on the luminal side of the BBB 13
endothelial cells and the BBB was intact, the current study demonstrated that 5 nm ceria ENM 14
produced significant pro-oxidant effects in the brain 30 days following administration and their 15
retention in peripheral organs (Figure 5). Based on the EELS analysis, it would appear that even 16
after a long-term exposure time of 30 days ceria continued to show a significant +3 valence on 17
the surface. Similar high +3 valence was observed on surface of ceria ENM after short-term 18
retention inside the rat, as seen in our previous study (Hardas et al. 2010), which was not 19
significantly different compared to freshly prepared ceria. This finding clearly defines an 20
enhanced oxygen storage capacity (Nesic et al.) of the ceria surfaces that does not diminish 21
greatly throughout the 30 day exposure period. Yet, ceria with enhanced OSC was shown in the 22
current study to have a significant pro-oxidative effect in the brain after the long-term exposure 23
following a single administration. 24
17
As depicted in Figure 5, a decline in the GSH: GSSG ratio (in hippocampus and cerebellum) in 1
the current study indicates elevated oxidative stress in the cellular environment. A similar 2
observation was reported in Park et. al. in which 30 nm ceria depleted GSH levels in human 3
lung epithelial cells in a dose-dependent manner (Park et al. 2008). Hydroxylated derivatives of 4
fullerenes also decreased the GSH: GSSG ratio and induced lipid peroxidation (Nakagawa et al. 5
2011). At the cellular level the GSH: GSSG ratio is dependent on GPx and GR enzymes, and in 6
the current study we observed decreased GPx activity (in hippocampus and cerebellum), 7
decreased GPx levels (in cerebellum) and decreased GR levels (cortex). Similarly, 15 nm silver 8
and 90 nm copper ENM down-regulated GPx gene expression (Wang et al. 2009), 25 nm silver 9
ENM (Rahman et al. 2009) and SWCNT (Wang et al. 2011) inhibited GPx activity, and silica 10
ENM reduced GR activity (Akhtar et al. 2010). Thus it may be possible that after 30 days 11
following administration, ENM have deleterious effects on enzymes needed for maintenance of 12
the reduced thiol status of the cells (Figure 5). 13
Figure 5 also shows that inhibition of catalase and GPx activities may lead to accumulation of 14
H2O2 and ultimately increased production of hydroxyl radicals (OH•). Activity of catalase can be 15
inhibited by hydroxyl radicals (OH•) and that of GPx by H2O2 and hydroperoxides (Pigeolet et al. 16
1990). Ceria ENM can produce H2O2 under abiotic conditions (Korsvik et al. 2007; Xia et al. 17
2008), and H2O2 has high membrane permeability (Halliwell 1992). Further H2O2 can undergo a 18
Fenton-type reaction to produce highly potent OH• radicals as noted above (Figure 5). A lack of 19
change in SOD activity and level in the current study may imply that endogenous SOD does not 20
account for ceria-induced elevated oxidative stress in brain. Therefore, ceria ENM treatment 21
may have caused induction in ROS that led to oxidative inhibition of antioxidant enzyme 22
activities, decreased the GSH: GSSG ratio and increased in PC and 3NT levels observed in the 23
present study. Similar consequences of oxidative stress were seen after exposure to various 24
18
other ENM, such as TiO2 (Hao et al. 2009; Liang et al. 2009; Xiong et al. 2011), SWCNT(Wang 1
et al. 2011), MWCNT(Guo et al. 2011), hematite (Radu et al. 2010), and ZnO (Xia et al. 2008). 2
3
The elevated 3NT levels, a marker for increased nitrosive stress, are consistent with elevated 4
levels of inducible nitric oxide synthase (iNOS). iNOS exists at extremely low levels under 5
normal physiological conditions, but it is inducible by endotoxin and inflammatory cytokines 6
among other stresses (Calabrese et al. 2000). We found a correlation between 3NT levels and 7
iNOS levels in the cortical region of rats examined 30 d after ceria ENM treatment. It may 8
indicate that increased 3NT levels are due to increased NO production in the cortex and we 9
speculate that iNOS levels concurrently may be induced following ceria treatment via increased 10
cytokine production. 11
12
Electron micrograph analysis showed an intact BBB and an absence of any significant amount 13
of ceria in the brain. However, cobalt-chromium ENMs have damaged DNAs without crossing 14
cellular membranes (Bhabra et al. 2009), and the chemotherapeutic drug doxorubicin led to 15
neurotoxic effects without ever crossing the BBB (Tangpong et al. 2006). Similarly, it is 16
conceivable that the observed oxidative stress response in brain regions could be due to the 17
accumulation of 5 nm ceria in peripheral organs and subsequent elevation of BBB-permeable 18
inflammatory cytokines. Studies to test this notion are in progress. 19
20
Induction of heat shock protein Hsp-70 levels as seen in the present study is in agreement with 21
other literature reports. Similar to effects of silver ENM (Ahamed et al. 2010) and fullerene C60 22
(Usenko et al. 2008) treatments, Hsp-70 levels and other oxidative stress markers were induced 23
with concomitant decrease in the GSH: GSSG ratio. In a transgenic mouse model for 24
cardiomyopathy, Hsp-70 levels were increased as an oxidative stress marker of ER stress. 25
Ceria ENM treatment rescued these cells from ER stress and as a result Hsp-70 expression 26
19
was down regulated (Niu et al. 2007). In the present study, 5 nm ceria ENM indirectly induced 1
ROS production in brain causing depletion of GSH, which initially induces antioxidant levels but 2
over 30 days eventually inhibits H2O2-scavenging catalase and GPx enzyme activities. This may 3
increase H2O2 levels and therefore OH• production via the Fenton reaction, which may cause 4
protein oxidation and induction of Hsp 70 levels (Figure 5). 5
6
The pro-oxidant effects observed in the brain 30 days after a single intravenous administration 7
of 5 nm ceria ENM are similar to those observed in age-related or Alzheimer disease-related 8
oxidative stress effects previously reported by our laboratory (Butterfield et al., 2001; Butterfield 9
and Stadtman, 1997). Although all three brain regions showed an effect of 5 nm ceria ENM 10
treatment, the extent of changes in oxidative stress indices was not same for all three brain 11
regions. Therefore, depending upon which brain region is affected, the function of that brain 12
region will be compromised. The EELS data suggests that mechanisms other than the valence 13
switching between Ce4+ and Ce3+ oxidation states and the possibility to absorb and release 14
oxygen by inducing oxygen vacancies must play a critically important role in pro-oxidant effects 15
of ceria ENM. At present it is difficult to speculate why 5 nm ceria ENM did not cross the BBB. 16
However, oxidative stress effects induced in brain may have been caused by some peripheral 17
inflammatory cytokines that cross the BBB or by ROS generated as a result of long-term 18
accumulation of 5 nm ceria ENM in peripheral organs. 19
20
Conclusions: 21
Although short term time exposure to ceria in vivo leads to a relatively small initial biochemical 22
response (Hardas et al. 2010), ceria administration may prove to be more harmful in the long 23
term. As reported here a single acute dose of 5 nm ceria ENM can adversely affect brain redox 24
status after 30 d. The important point to be noted is that, ceria may induce pro-oxidant effects 25
without crossing or disturbing BBB. The elevated oxidative stress in particular brain regions may 26
20
compromise brain functions and conceivably could even lead to neurodegeneration. Thus, 1
implications of our study are profound for the proposed use of ceria ENM in therapeutic/non-2
therapeutic applications, which may lead to human exposure. Therefore, caution is suggested 3
until and unless such effects as we demonstrated here can be mitigated. 4
5
Acknowledgements: This work was supported by United States Environmental Protection 6
Agency Science to Achieve Results [grant number RD-833772]. Although the research 7
described in this article has been funded wholly or in part by the United States Environmental 8
Protection Agency through STAR Grant RD-833772, it has not been subjected to the Agency’s 9
required peer and policy review and therefore does not necessarily reflect the views of the 10
Agency and no official endorsement should be inferred. 11
12
21
References: 1
Ahamed, M., R. Posgai, et al. (2010). "Silver nanoparticles induced heat shock protein 70, 2 oxidative stress and apoptosis in Drosophila melanogaster." Toxicol. Appl. Pharmacol. 3 242(3): 263-269. 4 Akhtar, M. J., M. Ahamed, et al. (2010). "Nanotoxicity of pure silica mediated through oxidant 5 generation rather than glutathione depletion in human lung epithelial cells." Toxicology 6 276(2): 95-102. 7 Amin, K. A., M. S. Hassan, et al. (2011). "The protective effects of cerium oxide nanoparticles 8 against hepatic oxidative damage induced by monocrotaline." International journal of 9 nanomedicine 6: 143-149. 10 Auffan, M., J. Rose, et al. (2009). "CeO2 nanoparticles induce DNA damage towards human 11 dermal fibroblasts in vitro." Nanotoxicology 3(2): 161-171. 12 Babu, S., J.-H. Cho, et al. (2010). "Multicolored redox active upconverter cerium oxide 13 nanoparticle for bio-imaging and therapeutics." Chem. Commun. 46(37): 6915-6917. 14 Bhabra, G., A. Sood, et al. (2009). "Nanoparticles can cause DNA damage across a cellular 15 barrier." Nat Nano 4(12): 876-883. 16 Brunner, T. J., P. Wick, et al. (2006). "In vitro cytotoxicity of oxide nanoparticles: comparison to 17 asbestos, silica, and the effect of particle solubility." Environ. Sci. Technol. 40(14): 4374-18 4381. 19 Butterfield, D.A., Drake, J., Pocernich, C., Castegna, A. (2001) "Evidence of oxidative damage 20 in Alzheimer's disease brain: Central role of amyloid -peptide." Trends Molec. Med. 21 7:548-554. 22 Butterfield, D.A., Stadtman, E.R. (1997) "Protein oxidation processes in aging brain." Adv. Cell 23 Aging Gerontol. 2:161-191. 24 Butterfield, D. A., T. Reed, et al. (2007). "Roles of amyloid [beta]-peptide-associated oxidative 25 stress and brain protein modifications in the pathogenesis of Alzheimer's disease and 26 mild cognitive impairment." Free Radic. Biol. Med. 43(5): 658-677. 27 Calabrese, V., T. E. Bates, et al. (2000). "NO synthase and NO-dependent signal pathways in 28 brain aging and neurodegenerative disorders: The role of oxidant/antioxidant balance." 29 Neurochemical Research 25(9-10): 1315-1341. 30 Celardo, I., J. Z. Pedersen, et al. (2011). "Pharmacological potential of cerium oxide 31 nanoparticles." Nanoscale 3(4): 1411-1420. 32 Chen, J. P., S. Patil, et al. (2006). "Rare earth nanoparticles prevent retinal degeneration induced 33 by intracellular peroxides." Nat. Nanotechnol. 1(2): 142-150. 34 Cho, W.-S., R. Duffin, et al. (2010). "Metal oxide nanoparticles induce unique inflammatory 35 footprints in the lung: important implications for nanoparticle testing." Environmental 36 Health Perspectives 118(12): 1699-1706. 37 Choi, J., V. Reipa, et al. (2011). "Physicochemical characterization and in vitro hemolysis 38 evaluation of silver nanoparticles." Toxicological Sciences 123(1): 133-143. 39 Colon, J., N. Hsieh, et al. (2010). "Cerium oxide nanoparticles protect gastrointestinal epithelium 40 from radiation-induced damage by reduction of reactive oxygen species and upregulation 41 of superoxide dismutase 2." Nanomedicine : nanotechnology, biology, and medicine 6(5): 42 698-705. 43 D'Angelo, B., S. Santucci, et al. (2009). "Cerium oxide nanoparticles trigger neuronal survival in 44 a human Alzheimer disease model by modulating BDNF pathway." Current Nanoscience 45 5(2): 167-176. 46
22
Das, M., S. Patil, et al. (2007). "Auto-catalytic ceria nanoparticles offer neuroprotection to adult 1 rat spinal cord neurons." Biomaterials 28(10): 1918-1925. 2 Donaldson, K., V. Stone, et al. (2004). "Nanotoxicology." Occupational and Environmental 3 Medicine 61(9): 727-728. 4 Estevez, A. Y., S. Pritchard, et al. (2011). "Neuroprotective mechanisms of cerium oxide 5 nanoparticles in a mouse hippocampal brain slice model of ischemia." Free Radic. Biol. 6 Med. 51(6): 1155-1163. 7 Fubini, B., M. Ghiazza, et al. (2010). "Physico-chemical features of engineered nanoparticles 8 relevant to their toxicity." Nanotoxicology 4(4): 347-363. 9 Gilliss, S. R., J. Bentley, et al. (2005). "Electron energy-loss spectroscopic study of the surface of 10 ceria abrasives." Applied Surface Science 241(1-2): 61-67. 11 Guo, Y.-Y., J. Zhang, et al. (2011). "Cytotoxic and genotoxic effects of multi-wall carbon 12 nanotubes on human umbilical vein endothelial cells in vitro." Mutation 13 Research/Genetic Toxicology and Environmental Mutagenesis 721(2): 184-191. 14 Halliwell, B. (1992). "Reactive oxygen species and the central-nervous-system." J. Neurochem. 15 59(5): 1609-1623. 16 Hao, L., Z. Wang, et al. (2009). "Effect of sub-acute exposure to TiO2 nanoparticles on oxidative 17 stress and histopathological changes in Juvenile Carp (Cyprinus carpio)." Journal of 18 Environmental Sciences 21(10): 1459-1466. 19 Hardas, S. S., D. A. Butterfield, et al. (2010). "Brain distribution and toxicological evaluation of 20 a systemically delivered engineered nanoscale ceria." Toxicological Sciences 116(2): 21 562-576. 22 Hirst, S. M., A. Karakoti, et al. (2011). "Bio-distribution and in vivo antioxidant effects of 23 cerium oxide nanoparticles in mice." Environmental Toxicology: n/a-n/a. 24 Hirst, S. M., A. S. Karakoti, et al. (2009). "Anti-inflammatory Properties of Cerium Oxide 25 Nanoparticles." Small 5(24): 2848-2856. 26 Hissin, P. J. and R. Hilf (1976). "A fluorometric method for determination of oxidized and 27 reduced glutathione in tissues." Analytical Biochemistry 74(1): 214-226. 28 Korsvik, C., S. Patil, et al. (2007). "Superoxide dismutase mimetic properties exhibited by 29 vacancy engineered ceria nanoparticles." Chem. Commun.(10): 1056-1058. 30 Landsiedel, R., M. D. Kapp, et al. (2009). "Genotoxicity investigations on nanomaterials: 31 Methods, preparation and characterization of test material, potential artifacts and 32 limitationsMany questions, some answers." Mutation Research/Reviews in Mutation 33 Research 681(2-3): 241-258. 34 Liang, G., Y. Pu, et al. (2009). "Influence of Different Sizes of Titanium Dioxide Nanoparticles 35 on Hepatic and Renal Functions in Rats with Correlation to Oxidative Stress." J Toxicol 36 Environ Health A 72(11-12): 740-745. 37 Lin, W. S., Y. W. Huang, et al. (2006). "Toxicity of cerium oxide nanoparticles in human lung 38 cancer cells." Int. J. Toxicol. 25(6): 451-457. 39 Mocan, T., S. Clichici, et al. (2010). "Implications of oxidative stress mechanisms in toxicity of 40 nanoparticles (review)." Acta Physiologica Hungarica 97(3): 247-255. 41 Nakagawa, Y., T. Suzuki, et al. (2011). "Cytotoxic effects of hydroxylated fullerenes on isolated 42 rat hepatocytes via mitochondrial dysfunction." Arch. Toxicol. 85(11): 1429-1440. 43 Nel, A., T. Xia, et al. (2006). "Toxic potential of materials at the nanolevel." Science 311(5761): 44 622-627. 45
23
Nesic, O., G.-Y. Xu, et al. (2004). "IL-1 Receptor Antagonist Prevents Apoptosis and Caspase-3 1 Activation after Spinal Cord Injury " Journal of Neurotrauma 18(9): 947-956. 2 Niu, J., A. Azfer, et al. (2007). "Cardioprotective effects of cerium oxide nanoparticles in a 3 transgenic murine model of cardiomyopathy." Cardiovascular Research 73(3): 549-559. 4 Niu, J., K. Wang, et al. (2011). "Cerium Oxide Nanoparticles Inhibits Oxidative Stress and 5 Nuclear Factor-κB Activation in H9c2 Cardiomyocytes Exposed to Cigarette Smoke 6 Extract." J. Pharmacol. Exp. Ther. 338(1): 53-61. 7 Park, E. J., J. Choi, et al. (2008). "Oxidative stress induced by cerium oxide nanoparticles in 8 cultured BEAS-2B cells." Toxicology 245(1-2): 90-100. 9 Pigeolet, E., P. Corbisier, et al. (1990). "Glutathione-peroxidase, superoxide-dismutase, and 10 catalase inactivation by peroxides and oxygen derived free-radicals." Mechanisms of 11 Ageing and Development 51(3): 283-297. 12 Pirmohamed, T., J. M. Dowding, et al. (2010). "Nanoceria exhibit redox state-dependent catalase 13 mimetic activity." Chem. Commun. 46(16): 2736-2738. 14 Radu, M., M. C. Munteanu, et al. (2010). "Depletion of intracellular glutathione and increased 15 lipid peroxidation mediate cytotoxicity of hematite nanoparticles in MRC-5 cells." Acta 16 Biochemica Polonica 57(3): 355-360. 17 Rahman, M. F., J. Wang, et al. (2009). "Expression of genes related to oxidative stress in the 18 mouse brain after exposure to silver-25 nanoparticles." Toxicology Letters 187(1): 15-21. 19 Rzigalinski, B. A., K. Meehan, et al. (2006). "Radical nanomedicine." Nanomedicine (Lond) 20 1(4): 399-412. 21 Sharma, H. S. and A. Sharma (2007). Nanoparticles aggravate heat stress induced cognitive 22 deficits, bloodbrain barrier disruption, edema formation and brain pathology. Progress 23 in Brain Research, Elsevier. 162: 245-273. 24 Srinivas, A., P. J. Rao, et al. (2011). "Acute inhalation toxicity of cerium oxide nanoparticles in 25 rats." Toxicology Letters 205(2): 105-115. 26 Sultana, R., M. Piroddi, et al. (2008). "Protein Levels and Activity of Some Antioxidant 27 Enzymes in Hippocampus of Subjects with Amnestic Mild Cognitive Impairment." 28 Neurochem Res 33(12): 2540-2546. 29 Tangpong, J., M. P. Cole, et al. (2006). "Adriamycin-induced, TNF-α-mediated central nervous 30 system toxicity." Neurobiology of Disease 23(1): 127-139. 31 Usenko, C. Y., S. L. Harper, et al. (2008). "Fullerene C60 exposure elicits an oxidative stress 32 response in embryonic zebrafish." Toxicol. Appl. Pharmacol. 229(1): 44-55. 33 Wang, J., M. F. Rahman, et al. (2009). "Expression changes of dopaminergic system-related 34 genes in PC12 cells induced by manganese, silver, or copper nanoparticles." 35 NeuroToxicology 30(6): 926-933. 36 Wang, J., P. Sun, et al. (2011). "Cytotoxicity of single-walled carbon nanotubes on PC12 cells." 37 Toxicol. Vitro 25(1): 242-250. 38 Xia, T., M. Kovochich, et al. (2008). "Comparison of the mechanism of toxicity of zinc oxide 39 and cerium oxide nanoparticles based on dissolution and oxidative stress properties." Acs 40 Nano 2(10): 2121-2134. 41 Xia, T., N. Li, et al. (2009). "Potential health impact of nanoparticles." Annual Review of Public 42 Health 30(1): 137-150. 43 Xiong, D., T. Fang, et al. (2011). "Effects of nano-scale TiO2, ZnO and their bulk counterparts 44 on zebrafish: Acute toxicity, oxidative stress and oxidative damage." Science of The 45 Total Environment 409(8): 1444-1452. 46
24
Yokel, R. A., R. L. Florence, et al. (2009). "Biodistribution and oxidative stress effects of a 1 systemically-introduced commercial ceria engineered nanomaterial." Nanotoxicology 2 3(3): 234-248. 3 Yokel, R. A., R. L. Florence, et al. (2009). "Biodistribution and oxidative stress effects of a 4 systemically-introduced commercial ceria engineered nanomaterial." Nanotoxicology 5 3(4): 234-248. 6 Younce, C. W., K. Wang, et al. (2010). "Hyperglycaemia-induced cardiomyocyte death is 7 mediated via MCP-1 production and induction of a novel zinc-finger protein MCPIP." 8 Cardiovascular Research 87(4): 665-674. 9 Zhou, X., L. L. Wong, et al. (2011). "Nanoceria inhibit the development and promote the 10 regression of pathologic retinal neovascularization in the vldlr knockout mouse." PLoS 11 ONE 6(2): e16733. 12 13 14
15
25
Table legend: 1
Table 1: Cerium concentrations in blood, brain, and liver, expressed as concentration and as a 2
percentage of the ceria ENM dose. [Ce] in mg/kg wet weight of the blood or tissue (mean ± SD). 3
4
5
Table 1: Cerium concentrations in blood, brain, and liver 6
Cerium concentration [Ce] (mg/kg) wet weight and as a % of the ceria
ENM dosea
Blood (mg/kg)
Brain (mg/kg)
Liver (mg/kg)
[Ce]
0.11 ± 0.16
0.38 ± 0.52
505 ± 238
% dose
0.01 ± 0.02
0.008 ± 0.009
44 ± 27
7
a Based on reference volume of blood in the rat (7% of body weight) or weight of the brain or 8
liver times the ceria concentration. 9
10
11
12
26
Figure legends: 1
Figure 1: Ceria ENM imaged using HRTEM. The ceria were dispersed on a carbon film. 2
Visually they have a narrow size distribution ranging from 4 to 6 nm, the majority 5 nm. 3
4
Figure 2: In hippocampus 30 d after a single acute dose of 5 nm ceria ENM: a) histograms 5
showing GR, GPx, catalase, antioxidant enzymes and Hsp70 heat shock protein levels and 6
corresponding Western blot experiments showing protein levels in control [C] and treated [T] 7
samples. The intensity of each band was normalized with intensity of corresponding band of β-8
actin-loading control (not shown); b) GR, GPx catalase antioxidant activities, and GSH: GSSG 9
ratio measured in control and ceria treated samples; and c) oxidative stress markers PC, 3NT 10
and HNE levels. The values are calculated as % control for each measurement expressed as 11
mean ± SEM, control n = 7, treated n = 9, *p <0.05, **p < 0.01, compared to control. 12
13
Figure 3: In cortex 30 d after a single acute dose of 5 nm ceria ENM: a) histograms showing 14
GR, GPx, catalase, antioxidant enzymes and Hsp70 heat shock protein levels and 15
corresponding Western blot experiments showing protein levels in control [C] and treated [T] 16
samples. The intensity of each band was normalized with intensity of corresponding band of β-17
actin-loading control (not shown); b) GR, GPx, catalase, antioxidants activities and GSH: GSSG 18
ratio measured in control and ceria treated samples; c) oxidative stress markers PC, 3NT and 19
HNE levels, and d) iNOS levels and correlation between 3NT levels and iNOS levels in ceria 20
treated samples, n=9, r=0.67, p < 0.05. The values are calculated as % control for each 21
measurement expressed as mean ± SEM, control n = 7, treated n = 9, *p <0.05, **p < 0.01, 22
compared to control. 23
24
Figure 4: In cerebellum 30 d after a single acute dose of 5 nm ceria ENM: a) histograms 25
showing GR, GPx, catalase, antioxidant enzymes and Hsp70 heat shock protein levels and 26
27
corresponding Western blot experiments showing protein levels in control [C] and treated [T] 1
samples. The intensity of each band was normalized with intensity of corresponding band of β-2
actin-loading control (not shown); b) GR, GPx, catalase, antioxidants activities, and GSH: 3
GSSG ratio measured in control and ceria treated samples, and c) oxidative stress markers PC, 4
3NT and HNE levels. The values are calculated as % control for each measurement expressed 5
as mean ± SEM, control n = 7, treated n = 8, *p <0.05, **p < 0.01, compared to control. 6
7
Figure 5: In this proposed pathway ceria ENM induces pro-oxidant effects on rat brain without 8
crossing BBB. Ceria ENM indirectly induces ROS production leading to GSH depletion and 9
inhibition of catalase and GPx enzymes. This inhibition of H2O2 reducing enzyme activity can 10
induce H2O2 levels and therefore hydroxyl radicals (OH•) production mediated by Fenton 11
reaction. Increased OH• can further oxidize the proteins and may hamper their regular function. 12
Hydroxyl radicals may also inhibit H2O2 reducing catalase and GPx activity by way of oxidative 13
modification. These biochemical reactions could make cellular environment more oxidizing, 14
triggering a cellular stress response to induce Hsp-70 levels. Hsp-70 is a chaperone protein that 15
can shepherd oxidized proteins to the 20S proteosome for degradation for further cellular 16
clearance (shown as dotted arrow). If timely clearance of oxidized protein takes place then there 17
may not be any change in cellular PC levels. As there was no evidence of the presence of ceria 18
ENM inside the brain, it is further proposed that ceria ENM exert their pro-oxidant effect in the 19
brain secondary to its peripheral effects. 20
21
22
23
24
25
26
28
1
2
3
4
5
6
7
8
9
10
11
29
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
30
1
2
3
4
5
6
7
8
9
10
31
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
32
1
2
... The ability to scavenge free radicals of the CeNP was observed in studies with J774A.1 murine macrophage cells (Hirst et al., 2009); and with hippocampus and cerebellum (Hardas et al., 2012). ...
... This hydrothermal reaction led to smaller particles~4 nm. This particle size is similar of the other studies than used similar rote of synthesis (Hardas et al., 2012;Goharshadi et al., 2011). However, CeNP with upper crystalline size has been synthesized by autoclave-mediated hydrothermal (Shajahan et al., 2020) as well as microwave-mediated hydrothermal synthesis (Soren et al., 2015). ...
... The CeNPs surface area observed in this study were twofold higher than of previous nanoparticle studies (Yokel et al., 2013;Hardas et al., 2012;Celardo et al., 2011), but lower than another study (Zeyons et al., 2009). Thus, the small size of the CeNPs of this study is within previous reported values and it could be confirmed by TEM images and also confirmed by BET surface area. ...
Article
Full-text available
Fungal infection is a public health problem. Antifungal agents resistance is often seen in common Candida albicans in the hospital environment. Nanoparticles have been reported in the literature as a promising development of health products. The toxicity, antioxidant and antifungal activity of cerium oxide nanoparticles (CeNP) were evaluated. After synthesis and characterization of the physicochemical properties, the new CeNP was evaluated by biological tests of antifungal activity. The antioxidant activity of CeNP was evaluated by scavenging free radicals of 1,1-diphenyl-2-picrylhydrazyl hydrate (DPPH). The DPPH scavenging activity was monitored by % color inhibition the absorbance. In vivo acute toxicity studies of CeNP were carried out by oral administration in mice (300 and 2000 mg/kg, n = 3/ group), and by brine shrimp (Artemia salina) model (0.001-25 mg/ mL, n = 10/ group). CeNP was able to interfere in the fungal growth, depending on the strain and dosage used. Acute lethal dose 50% is greater than 2000 mg/kg and CeNP did not induce toxicity for A. salina. Antioxidant activity was not significant. The current antifungal and toxicity features results support the use of CeNP as antifungal agent against Candida albicans strains, which may find applications in biotechnology and biomedical area in the development of a new nano-biomaterial for clinical applications.
... L'absorbance du DNPH est mesurée à une longueur d'onde de 370 nm (Ɛ = 22000 mol.l-1.cm-1 ; Reznick et Packer, 1994). Bien que le dosage des protéines carbonylées soit largement utilisé dans la littérature pour la quantification du stress oxydant et en présence de NPC (Arya et al., 2016;Hardas et al., 2012;Niu et al., 2007;, nous n'avons pas réussi à démontrer une augmentation de la production de protéines carbonylées par le glutamate à 100mM ou par l'H2O2 à 2mM. La mise en évidence des atteintes des protéines pourrait être évaluée par la quantification de la 3-nitrotyrosine (3-NT) comme cela a déjà été réalisé et a permis de montrer l'effet protecteur des NPC Estevez et al., 2011;Hardas et al., 2012;Niu et al., 2007). ...
... Bien que le dosage des protéines carbonylées soit largement utilisé dans la littérature pour la quantification du stress oxydant et en présence de NPC (Arya et al., 2016;Hardas et al., 2012;Niu et al., 2007;, nous n'avons pas réussi à démontrer une augmentation de la production de protéines carbonylées par le glutamate à 100mM ou par l'H2O2 à 2mM. La mise en évidence des atteintes des protéines pourrait être évaluée par la quantification de la 3-nitrotyrosine (3-NT) comme cela a déjà été réalisé et a permis de montrer l'effet protecteur des NPC Estevez et al., 2011;Hardas et al., 2012;Niu et al., 2007). ...
Thesis
Les accidents vasculaires cérébraux (AVC) constituent la 2ème cause de mortalité dans le monde et la 1ère chez les femmes en France. Pour les AVC ischémiques (AVCi), seules des stratégies de recanalisation pharmacologique ou mécanique ont été approuvées mais aucune stratégie protectrice n'est aujourd'hui disponible. Bien que le rôle délétère du stress oxydant ait été clairement établi dans les lésions neuronales et vasculaires à la suite d'une ischémie cérébrale (IC) dans les études précliniques, aucune stratégie anti-oxydante n'a démontré d'efficacité clinique à ce jour. Or, les nanoparticules d'oxyde de cérium (NPC) possèdent de multiples capacités antioxydantes (enzymatique et non enzymatique). Afin d'améliorer la biocompatibilité des NPC, la société Specific Polymers® a développé des copolymères de polyéthylène glycol (PEG)/ polyméthacrylate de méthyle/ phosphonate pour recouvrir leur surface. De plus, ces polymères peuvent être fonctionnalisés avec un peptide ciblant l'endothélium ce qui permettrait d'y concentrer l'effet antioxydant des NPC, afin de réduire la survenue d'hémorragies cérébrales, complications graves chez les patients victimes d'AVCi. L'objectif de cette thèse est d'évaluer l'impact du recouvrement des NPC sur leur potentiel thérapeutique dans l'IC. Les études ont été menées in vitro pour établir la toxicité, l'effet antioxydant et l'internalisation cellulaire des NPC et in vivo, pour examiner leur biodistribution et leur toxicité, ainsi que leur potentiel thérapeutique dans un modèle d'IC. Les études in vitro ont été effectuées sur des cellules endothéliales cérébrales murines de la lignée b.End3. Nous avons démontré que les NPC n'induisaient ni mortalité, ni perturbation de l'activité métabolique jusqu'à 100µg/ml. A 1000µg/mL, les NPC nues augmentent la mortalité, contrairement aux NPC PEGylées. Nous avons modélisé l'excitotoxicité survenant lors d'une IC et qui contribue au stress oxydant, grâce à un traitement des cellules par le glutamate. L'augmentation de la production d'espèces réactives de l'oxygène par les cellules b.End3 et l'oxydation des acides nucléiques dans ces conditions ont été réduites par les NPC, démontrant que leur recouvrement n'interfère pas avec leurs propriétés anti-oxydantes. La fonctionnalisation des NPC a permis le greffage d'un fluorophore pour suivre leur internalisation par cytométrie en flux et microscopie confocale. Nous avons ainsi mis en évidence que les NPC étaient rapidement internalisées dans les cellules b.End3. Des études de microscopie électronique ont ensuite montré que les NPC sont principalement localisées dans des endosomes périnucléaires. Enfin, nous avons réalisé le greffage sur les NPC d'un peptide ciblant une protéine d'adhésion vasculaire surexprimée lors de l'IC. La suite de ces études consistera à vérifier l'interaction spécifique de ces NPC avec la molécule d'adhésion. Les études in vivo ont permis d'établir la biodistribution des NPC chez des souris Swiss : des NPC sont retrouvées durant les premières heures suivant leur injection, avant leur élimination par voie rénale. L'histopathologie n'a révélé aucune toxicité des NPC sur le foie, les reins, la rate, les poumons et le cerveau de ces souris et aucune modification de leur numération sanguine n'a été observée. Les NPC ont ensuite été administrées dans un modèle murin d'IC, mais n'ont pas réduit le volume de la lésion dans nos conditions. En conclusion, le recouvrement des NPC par des polymères innovants a réduit leur toxicité sans altérer leurs capacités antioxydantes et leur internalisation dans des cellules endothéliales cérébrales. L'absence d'accumulation à long terme et de toxicité in vivo sont encourageantes quant à leur biocompatibilité. Bien que les NPC n'aient pas montré d'effet protecteur in vivo, celles ciblant l'endothélium pourraient réduire les lésions vasculaires et le risque hémorragique consécutif à une IC.
... For example, rat brain pro-oxidant effects were reported after peripheral administration of 85 mg/kg nanoceria. Without permeating the BBB, 5-nm nanoceria indirectly decreased the ratio of reduced to oxidized glutathione in the hippocampus and cerebellum, which was an indicator of oxidative stress [165]. To date, there is little evidence of toxicity in response to CNS injury treatment with nanoceria in vivo, as the administration doses are generally very low, mostly lower than 1 mg/kg (Table 1). ...
Article
Full-text available
Central nervous system (CNS) injury, induced by ischemic/hemorrhagic or traumatic damage, is one of the most common causes of death and long-term disability worldwide. Reactive oxygen and nitrogen species (RONS) resulting in oxidative/nitrosative stress play a critical role in the pathological cascade of molecular events after CNS injury. Therefore, by targeting RONS, antioxidant therapies have been intensively explored in previous studies. However, traditional antioxidants have achieved limited success thus far, and the development of new antioxidants to achieve highly effective RONS modulation in CNS injury still remains a great challenge. With the rapid development of nanotechnology, novel nanomaterials provided promising opportunities to address this challenge. Within these, nanoceria has gained much attention due to its regenerative and excellent RONS elimination capability. To promote its practical application, it is important to know what has been done and what has yet to be done. This review aims to present the opportunities and challenges of nanoceria in treating CNS injury. The physicochemical properties of nanoceria and its interaction with RONS are described. The applications of nanoceria for stroke and neurotrauma treatment are summarized. The possible directions for future application of nanoceria in CNS injury treatment are proposed.
... Compared to previous studies, the particle size of the nanoparticles produced in this research is similar to the other syntheses [24]. However, an autoclave-assisted hydrothermal process and microwave-mediated hydrothermal synthesis created the new-generation CeNP with the upper-range crystalline size [25,26]. ...
Preprint
Full-text available
Background: Nanotechnology plays a significant role in medicine, especially in diagnosis and treatment as a carrier to drugs and vaccinology. Several studies were conducted using nanoparticles as an adjuvant. The main aim of this study was in vivo and in vitro toxicity evaluation of synthesized Cerium Nanoparticles (CeNPs). Methods: In the present study, cerium nanoparticles were prepared using the wet chemical method. The formation of cerium nanoparticles was confirmed by scanning electron microscopy, transmission electron microscopes, x-ray diffraction analysis, dynamic light scattering. In vivo and in vitro toxicity of synthesized nanoparticles was evaluated in three different amounts of cerium nanoparticles (30 µg, 50 µg, & 100 µg) in mice and human fibroblast cell lines, respectively. Results: Cerium nanoparticles were successfully synthesized, and the identity was confirmed by x-ray diffraction analysis. The shape and size of nanoparticles were spherical and <100 nm, respectively. The prepared nanoparticles had a charge of-26.6 mV and a hydrodynamic radius of 446 nm. MTT assay indicated that none of the concentration of cerium was toxic, and in vivo toxicity also clarified the safety of cerium nanoparticles in mice; no significant un-normal behavioral and physical symptoms were observed in mice after CeNP administration Conclusion: Cerium nanoparticles have special properties, especially low toxicity, unique capabilities in stimulating the immune system. Cerium nanoparticles can be considered an effective and safe candidate in vaccines.
... While the dispersed nanoparticles haven't shown any such effects at all. Some of the cerium based nanoparticles along with its host and their effect on biological function is depicted in table 01 [44]. ...
Article
Full-text available
The advancement in the production and usage of the cerium oxide nanoparticles have diverted the attention of scientists towards their usage in medical field and therapeutic usage. The clinical usage of these Nano ceria is based on their ability to moderate the oxidative stress and this is only because of their ability to change their valent state from +3 to +4 which makes them ideal for scavenging radicals for use in a number of systemic and neurodegenerative disorders. This review aims to synthesize the basic methods used for the synthesis of nanoparticles along with the use of ligand, stabilizing agent and other components. This review also concludes that how various physical and chemical properties of nanoparticles effect the basic biological activities such as antimicrobial activity, cytotoxicity and many others. However, during the standardization, some of the physiochemical properties, methods used for preparation and catalytic abilities must be taken into account.
... Further, studies showed that nanoceria can protect cells against reactive oxygen species (ROS) such as superoxide radical anion and hydrogen peroxide, thereby suggesting it might have SOD-and catalase-mimicking activity [3,9,12,13]. In contrast, there are reports of nanoceriainduced pro-oxidant effects including lipid peroxidation, elevation of cytokines, and GSH depletion [10,[14][15][16]. ...
Article
Nanoceria (CeO 2 , cerium oxide nanoparticles) is proposed as a therapeutic for multiple disorders. In blood, nanoceria becomes protein-coated, changing its surface properties to yield a different presentation to cells. There is little information on the interaction of nanoceria with blood proteins. The current study is the first to report the proteomics identification of plasma and serum proteins adsorbed to nanoceria. The results identify a number of plasma and serum proteins interacting with nanoceria, proteins whose normal activities regulate numerous cell functions: antioxidant/detoxification, energy regulation, lipoproteins, signaling, complement, immune function, coagulation, iron homeostasis, proteolysis, inflammation, protein folding, protease inhibition, adhesion, protein/RNA degradation, and hormonal. The principal implications of this study are: 1) The protein corona may positively or negatively affect nanoceria cellular uptake, subsequent organ bioprocessing, and effects; and 2) Nanoceria adsorption may alter protein structure and function, including pro-and inflammatory effects. Consequently, prior to their use as therapeutic agents, better understanding of the effects of nanoceria protein coating is warranted.
Conference Paper
The emergence of nanotechnology has been an interesting innovation in this current age which is permeating rapidly into the various fields such as medicine, materials science, pharmacy, environmental protection, agriculture. Consequently, the high reactivity of these compounds is of great concern as there is more interaction with humans, biological and environmental elements of the ecosystem. However, there is limited knowledge on the risk and toxicity assessment of these nanomaterials. This review explores studies on the modes of contact and the possible damage to the human body system. It also suggests feasible procedures that can be deployed to assess the toxicity and the synthesis of nanoparticles/nanomaterials from biological origin (green synthesis) frequently.
Article
The acquisition of new nanomaterials with unique properties for biomedical application is an essential requirement and many nanoparticles are being studied to test their possible application in therapy. The present study aimed to evaluate the effect of nano-cerium/hydroxyapatite (nCe/HAp) as a regenerative agent for the juvenile hormones and the improvement of dyslipidemia in adult and aged rats. This study included two main groups of albino male rats (adult and aged), each of sixteen rats. Each group was divided equally into two subgroups, control and treated. Treated rats were injected intravenously with 75 mg/kg b. w. nCe/HAp weekly for four weeks. At the end of the experiment, the blood samples were collected between 12:00 and 1:00 (mid-dark) for biochemical analysis. The measurements of dehydroepianrsterone, dehydroepianrsterone sulphate, melatonin, and testosterone hormones were performed as well as lipid profile was analyzed. The obtained results revealed that nCe/HAp has a pronounced positive effect on juvenile hormones value and lipid profile status in adults more than in aged rats. Based on the experimental results, it can be established that nCe/HAp has a potential effect on improving the level of juvenile hormones in adult and aged rats. Therefore, it may candidate nCe/HAp composite as a regenerative drug in the event of hormonal imbalance resulting from dysfunction or aging in the future.
Article
Full-text available
Cerium oxide nanoparticles, so-called nanoceria, are engineered nanomaterials prepared by many methods that result in products with varying physicochemical properties and applications. Those used industrially are often calcined, an example is NM-212. Other nanoceria have beneficial pharmaceutical properties and are often prepared by solvothermal synthesis. Solvothermally synthesized nanoceria dissolve in acidic environments, accelerated by carboxylic acids. NM-212 dissolution has been reported to be minimal. To gain insight into the role of high-temperature exposure on nanoceria dissolution, product susceptibility to carboxylic acid-accelerated dissolution, and its effect on biological and catalytic properties of nanoceria, the dissolution of NM-212, a solvothermally synthesized nanoceria material, and a calcined form of the solvothermally synthesized nanoceria material (ca. 40, 4, and 40 nm diameter, respectively) was investigated. Two dissolution methods were employed. Dissolution of NM-212 and the calcined nanoceria was much slower than that of the non-calcined form. The decreased solubility was attributed to an increased amount of surface Ce ⁴⁺ species induced by the high temperature. Carboxylic acids doubled the very low dissolution rate of NM-212. Nanoceria dissolution releases Ce ³⁺ ions, which, with phosphate, form insoluble cerium phosphate in vivo. The addition of immobilized phosphates did not accelerate nanoceria dissolution, suggesting that the Ce ³⁺ ion release during nanoceria dissolution was phosphate-independent. Smaller particles resulting from partial nanoceria dissolution led to less cellular protein carbonyl formation, attributed to an increased amount of surface Ce ³⁺ species. Surface reactivity was greater for the solvothermally synthesized nanoceria, which had more Ce ³⁺ species at the surface. The results show that temperature treatment of nanoceria can produce significant differences in solubility and surface cerium valence, which affect the biological and catalytic properties of nanoceria.
Article
The resorptive effects of 0.01 M cerium nanodioxide sol upon single intraperitoneal administration to rats have been studied. The acute exposure to nanoparticles was found to have a dose-dependent general toxic effect on the body (weight loss, inflammatory changes in the abdominal organs, modification of individual behavior, hematological changes, metabolic imbalance), which develops on the background of POL activation. The prooxidant effect of cerium dioxide nanoparticles is demonstratively manifested at relatively high exposure levels (80–8 mg / kg). The threshold dose for the general toxic effect (Limch integr) is equal to 0.8 mg / kg.
Article
Full-text available
Cerium dioxide nanoparticles have been proposed for an increasing number of applications in biomedicine, cosmetic, as polishing materials and also as byproducts from automotive fuel additives. The aim of this study was to examine the potential in vitro cyto- and genotoxicity of nano-sized CeO2 (7 nm) on human dermal fibroblasts. By combining a physicochemical and a (geno)toxicological approach, we defined the causal mechanisms linking the physico-chemical properties of nano-CeO2 with their biological effects. Using X-ray absorption spectroscopy, we observed a reduction of 21 +/- 4% of the Ce4+ atoms localized at the surface of CeO2 nanoparticles due to the interactions with organic molecules present in biological media. These particles induced strong DNA lesions and chromosome damage related to an oxidative stress. These genotoxic effects occurred at very low doses, which highlighted the importance of a genotoxicological approach during the assessment of the toxicity of nanoparticles.
Article
Surfaces of ceria (CeO2) particles have been studied by electron energy-loss spectroscopy in a field-emission gun scanning transmission electron microscope. All the ceria particles analyzed contained Ce3+ at the surface. Rare-earth impurities such as La were enriched at the surface and were observed for particles ranging from tens to hundreds of nanometers in size. Fluorine in the abrasives corresponded to a lower average cerium valence. Time series investigations indicate that fluorine substitutes on the oxygen sub-lattice and is charge-balanced by some cerium changing from Ce4+ to Ce3+.
Article
One of the consequences of cytokine-orchestrated inflammation after CNS trauma is apoptosis. Our hypothesis is that cell death in the spinal cord after injury results in part from increased synthesis and release of IL-1beta. Using a ribonuclease protection assay, we demonstrated that there is increased transient expression of IL-1beta mRNA and, by using IL-1beta protein ELISA assay, that there are increased IL-1beta protein levels in the contused rat spinal cord, initially localized to the impact region of the spinal cord (segment T8). Using an ELISA cell death assay, we showed that there is apoptosis in the spinal cord 72 h after injury, a finding that was confirmed by measuring caspase-3 activity, which also significantly increased at the site of injury 72 h after trauma. Treatment of the contused spinal cord at the site of injury with the IL-1 receptor antagonist (rmIL-lra, 750 ng/mL) for 72 h using an osmotic minipump completely abolished the increases in contusion-induced apoptosis and caspase-3 activity.
Article
Publisher Summary Aging and age-related neurological disorders, especially Alzheimer's disease (AD) and stroke, affect millions of people worldwide. Free radical-associated protein oxidation in these brain disorders appears fundamental to the pathogenesis and etiology, and, hence, treatment of each. Other neurological disorders of the brain are associated with free radical oxidative stress, for example, Parkinson's disease, amyotrophic lateral sclerosis, Wilson's disease, and traumatic brain injury. Hence, greater understanding of free radical processes and their treatment and prevention in AD and stroke likely will provide insight into the basis of and treatment for other neurological disorders of oxidative stress. Membrane and cytosolic proteins, along with bilayer lipids, are primary targets for free radical oxidation in brain cells. This chapter summarizes some of the studies on protein oxidation and its involvement in aging, AD, and stroke. Several methods are widely used to assess the role of protein oxidation in oxidative stress in various diseases and in aging. The formation of cross-linked protein aggregates is of particular significance in the accumulation of reactive oxygen species (ROS)-mediated protein damage during aging and oxidative stress because such aggregates are resistant to degradation by proteases that preferentially degrade the oxidized forms of proteins.
Article
The objective was to characterize the biodistribution of nanoscale ceria from blood and its effects on oxidative stress endpoints. A commercial 5% crystalline ceria dispersion in water (average particle size ~31±4 nm) was infused intravenously into rats (0, 50, 250 and 750 mg/kg), which were terminated 1 or 20 h later. Biodistribution in rat tissues was assessed by microscopy and ICP-AES/MS. Oxidative stress effects were assessed by protein-bound 4-hydroxy 2-trans-nonenal (HNE), protein-bound 3-nitrotyrosine (3-NT), and protein carbonyls. Evans blue (EB)-albumin and Na fluorescein (Na2F) were given intravenously as blood-brain barrier integrity markers. The initial ceria t½ in blood was ~7 min. Brain EB and Na2F increased some at 20 h. Microscopy revealed peripheral organ ceria agglomerations but little in the brain. Spleen Ce concentration was >liver >blood >brain. Reticuloendothelial tissues cleared ceria. HNE was significantly increased in the hippocampus at 20 h. Protein carbonyl and 3-NT changes were small. The nanoparticle characterizations before and after biodistribution, linked with the physiological responses, provide a foundation for evaluating the effects of engineered nanomaterial physico-chemical properties on peripheral organ distribution, brain entry and resultant toxicity.
Article
Cerium oxide nanoparticles have oxygen defects in their lattice structure that enables them to act as a regenerative free radical scavenger in a physiological environment. We performed a comprehensive in vivo analysis of the biological distribution and antioxidant capabilities of nanoceria administered to mice perorally (PO), intravenously (IV), or intraperitoneally (IP) by dosing animals weekly for 2 or 5 weeks with 0.5 mg kg−1 nanoceria. Next, we examined if nanoceria administration would decrease ROS production in mice treated with carbon tetrachloride (CCl4). Our results showed that the most extensive and cumulative nano-deposition was via IV and IP administered while PO administration showed mice excreted greater than 95% of their nanoceria within 24 h. Organ deposition for IV and IP mice was greatest in the spleen followed by the liver, lungs, and kidneys. Elimination for all administration routes was through feces. Nanoceria administration showed no overt toxicity, however, WBC counts were elevated with IV and IP administration. Our in vivo studies show that nanoceria administration to mice with induced liver toxicity (by CCl4) showed similar findings to mice treated with N-acetyl cystine (NAC), a common therapeutic to reduce oxidative stress. Taken together, our studies show that nanoceria remains deposited in tissues and may decrease ROS, thereby suggesting that cerium oxide nanoparticles may be a useful antioxidant treatment for oxidative stress. © 2011 Wiley Periodicals, Inc. Environ Toxicol 2011.
Article
Although mankind stands to obtain great benefit from nanotechnology, it is important to consider the potential health impacts of nanomaterials (NMs). This consideration has launched the field of nanotoxicology, which is charged with assessing toxicological potential as well as promoting safe design and use of NMs. Although no human ailments have been ascribed to NMs thus far, early experimental studies indicate that NMs could initiate adverse biological responses that can lead to toxicological outcomes. One of the principal mechanisms is the generation of reactive oxygen species and oxidant injury. Because oxidant injury is also a major mechanism by which ambient ultrafine particles can induce adverse health effects, it is useful to consider the lessons learned from studying ambient particles. This review discusses the toxicological potential of NMs by comparing the possible injury mechanisms and adverse health effects of engineered and ambient ultrafine particles.