Você está na página 1de 9

Toxicology 262 (2009) 175–183

Contents lists available at ScienceDirect

Toxicology
journal homepage: www.elsevier.com/locate/toxicol

Aroclor 1254 induced cytotoxicity and mitochondrial dysfunction in isolated


rat hepatocytes
Hamdy A.A. Aly a,b,∗ , Òscar Domènech b
a
Department of Pharmacology and Toxicology, Faculty of Pharmacy, Al-Azhar University, Nasr City, Cairo, Egypt
b
Unit of Cellular and Molecular Pharmacology, Catholic University of Louvain B-1200 Brussels, Belgium

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

Article history: Polychlorinated biphenyls (PCBs) are widespread persistent environmental contaminants that display a
Received 13 May 2009 complex spectrum of toxicological properties, including hepatotoxicity. Although Aroclor 1254 is ubiqui-
Received in revised form 21 May 2009 tous in the environment, its potential cytotoxic effect on rat hepatocytes and the mechanism underlines
Accepted 22 May 2009
its cytotoxicity are not fully investigated. Therefore, the present study was conducted to investigate: (1)
Available online 30 May 2009
the potential cytotoxicity of Aroclor 1254 in rat hepatocytes, and (2) characterization of the molecular
mechanisms involved in the Aroclor 1254-induced hepatotoxicity, particularly the role of mitochondria,
Keywords:
possibly a primary target in such event, could greatly explain the cytotoxic effect of Aroclor 1254 in rat
Hepatocytes
Mitochondria
hepatocytes. Hepatocytes were isolated from adult male albino rats and incubated for 24 h in a fresh
Aroclor 1254 media containing 0, 20, 30, 40, 50 or 60 ␮M of Aroclor 1254.
Oxidative stress At the end of incubation, hepatocytes and hepatocyte mitochondria were used for the assay. Our results
showed cytotoxicity of Aroclor 1254 in rat hepatocytes starting at a concentration of 30 ␮M as manifested
by increased lactate dehydrogenase (LDH) leakage, decreased cell viability (MTT assay) and increased lipid
peroxidation. As mitochondria are known to be one possible site of the cell damage, the effects of Aro-
clor 1254 on hepatocyte mitochondria was investigated. Aroclor 1254 induced reactive oxygen species
(ROS) generation in hepatocyte mitochondria, inhibited mitochondrial respiratory chain complexes I
and III and ␤-oxidation of free fatty acids, depletion of mitochondrial antioxidant enzymes GPx and GR
and the non-enzymatic antioxidant reduced glutathione, inhibited mitochondrial membrane potential
(␺m ), decreased mitochondrial aconitase and cardiolipin content, and elevated levels of cytochrome
P450 subfamily, CYP1A and CYP2B activities as indicated by ethoxyresorufin O-deethylase (EROD) and
pentoxyresorufin O-deethylase (PROD).
Therefore, we can conclude that Aroclor 1254 induced rat hepatocyte toxicity and our findings provide
evidence to propose that mitochondria are one of the most important and earliest cell targets in Aroclor
1254-mediated toxicity and delineate several mitochondrial processes at least, in part, by induction of
oxidative stress. These findings can be useful in future cytoprotective therapy approaches. Since mito-
chondrial events appear to be targeted in hepatocellular damage induced by Aroclor 1254, an antioxidant
therapy targeted to mitochondria may constitute an interesting strategy to ameliorate its toxicity.
© 2009 Published by Elsevier Ireland Ltd.

1. Introduction and on surfaces in homes and in factories (Sadeghi-Aliabadi et al.,


2007). PCBs mixtures contain various levels of chlorination, and
Polychlorinated biphenyls (PCBs) were widely used in a vari- were sold under several trade names, including Aroclor, Clophen,
ety of industrial and consumer products for several decades before Fenchlor, Kanechlor, Phenoclor, and Pyralene (Larsen, 2006; Simon
their production was banned in the 1970s. As a result of their et al., 2007). These compounds are highly lipophilic and accumu-
extensive use and their chemical stability, PCBs are still ubiquitous late easily in the lipid bilayer and fat deposits of the body (Safe,
environmental contaminants that are frequently found as com- 1994; Parkinson et al., 1983). For these reasons, there is a toxico-
plex mixtures of isomers and congeners in air, water, soil, dust logical concern about possible risks associated with the exposure
of humans to PCBs (Silberhorn et al., 1990). PCBs are distributed
throughout the entire ecosystem, including soil, air and water
∗ Corresponding author at: Apartment No: 62, El-Nour Tower No: 1, El-Sefarate
(Risebrough et al., 1968). They are absorbed through the skin, lungs
Quartier, Nasr City, Cairo, Egypt. Tel.: +20 173809433.
and GI tract and are transported by blood to liver and muscles (Baird,
E-mail address: hamdyaali@yahoo.com (H.A.A. Aly). 1995). Exposure to these compounds results in various harmful

0300-483X/$ – see front matter © 2009 Published by Elsevier Ireland Ltd.


doi:10.1016/j.tox.2009.05.018
176 H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183

effects including reproductive toxicity, immune suppression, birth Oxidative stress develops when there is an imbalance between
defects, cancer, developmental and behavioral changes and liver prooxidants and antioxidants ratio, and as results it leads to the
damage (Safe, 1994; Kimbrough, 1995; Adebusoye et al., 2008). accumulation of oxidative damage molecules (Sakac and Sakac,
PCBs affect two enzyme systems responsible for production and 2000). A change in AOE activities is frequently used as an indica-
elimination of reactive intermediates: the xenobiotic-metabolizing tor of the ongoing oxidative stress in a particular tissue (Andric et
enzymes (XMEs) and the antioxidant enzymes (AOEs), respec- al., 2003). Tharappel et al. (2002) have evidenced that PCB induced
tively. The effects of PCBs on the levels of XMEs, in particular toxic manifestations, which may be associated with production of
the cytochrome P450 (CYP) isozymes, and AOEs are well estab- reactive oxygen species (ROS), thereby it leads to oxidative stress. It
lished in liver (Twaroski et al., 2001a,b; Fadhel et al., 2002). A has been reported that PCBs are completely carcinogens and tumor
mechanistic link between toxicity, induction of the CYP1A sub- promoters in the liver. Induction of oxidative stress has been pro-
family, and binding to the aryl hydrocarbon (Ah) receptor has posed as a mechanism for induction of their promoting activity in
been established for PCB congeners that are structurally related to liver (Fadhel et al., 2000).
2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) (Poland and Knutson, Mitochondria are an important intracellular source of ROS as
1982). well as a sensitive target for oxidative damage under certain
Depending on their three-dimensional structure and their sub- pathological conditions such as exposure to PCB. Mitochondrial
stitution pattern, PCB congeners bind to different cellular target membranes are rich in the tetraacyl phospholipid cardiolipin.
sites and cause adverse effects by a variety of different mechanisms Its integrity is important for efficient oxidative phosphorylation.
(Zhao et al., 2009). Structure–activity studies with individual PCBs The increase in ROS generation may be related to mitochondrial
have demonstrated that congeners with chlorine atoms at the para dysfunction. Moreover, there is a lot of evidence supporting the
and meta, but not ortho, positions of both phenyl rings are able to proposal that the impairment of mitochondrial function is a key
assume a dioxin-like conformation, bind the Ah receptor with high event in ROS-signaling through apoptotic pathways (Roue et al.,
affinity, and are efficacious inducers of CYP1A (Safe, 1994). PCB 2003; Kirkland and Franklin, 2003; Shih et al., 2004; Liu et al., 2004;
congeners with multiple ortho chlorine substituents may bind to Lambert et al., 2004).
or activate the constitutive androstane receptor (CAR) (Denomme Since liver is considered as the major site for detoxification of
et al., 1983) and/or the pregnane X receptor (PXR) (Schuetz et al., toxic metabolites, this organ is considered as the potential site for
1998). Other ortho-substituted PCB congeners interact with both the PCB-induced damage. Williams et al. (1993) reported that PCBs
the aryl hydrocarbon (Ah-) receptor and CAR (Parkinson et al., caused degradative changes such as cytoplasmic loss and periph-
1983). In addition, PCB congeners with multiple ortho substituents eralization of cytoplasm and organelles in goat, rabbit, and duck
sensitize ryanodine receptors (RyR) (Pessah et al., 2006), an inter- livers. Several studies have shown that PCBs increase lipid peroxi-
action that has been shown to be enantiospecific in the case of PCB dation in rat liver (Fadhel et al., 2002). Further, PCBs cause several
136 (Pessah et al., 2009). The mono-ortho derivatives of the dioxin- toxic effects in the liver and the mechanisms of toxicity have not
like PCBs have reduced Ah receptor-mediated activities and can been determined (Tharappel et al., 2008).
also induce CYP2B enzymes. Congeners possessing chlorine atoms Studies of environmental and toxic effects of PCBs are ideally
at the ortho positions of both phenyl rings are non-dioxin-like, do performed with PCB mixtures reflecting the composition of envi-
not readily bind to the Ah receptor, and exhibit a pattern of toxic- ronmental PCB profiles to mimic actual effects and to account for
ity including carcinogenic, neurotoxic, behavioral, and endocrine complex interactions among individual PCB congeners. Congener
effects that differs from that of the coplanar PCBs (Giesy and profiles of PCBs found in environmental samples are complex and
Kannan, 1998). Nevertheless, some di-ortho chlorinated PCBs are frequently do not reflect the composition of commercial PCB formu-
extremely weak inducers of CYP1A enzymes, whereas other di- lations (Zhao et al., 2009). As it is difficult to prepare environmental
ortho chlorinated congeners do not induce CYP1A enzymes, though PCB mixture that contain a significant number of congeners, this
they induce other CYP enzymes (Ikegwuonu et al., 1996). study was conducted using Aroclor 1254.
Data from many experimental studies with mixtures of these Although the PCBs display a wide range of biological and toxico-
compounds are consistent with an additive model (Van den Berg logical properties, the potential cytotoxic effect of Aroclor 1254 on
et al., 1998). As a result of this generally accepted additivity, the rat hepatocytes and the mechanism underlines its cytotoxicity are
toxic equivalency concept was developed. It uses the relative effect not fully investigated. Therefore, the present study was conducted
potency (REP) determined for individual PCB compounds for pro- to investigate: (1) the potential cytotoxicity of Aroclor 1254 in rat
ducing toxic or biological effects relative to a reference compound, hepatocytes, and (2) characterization of the molecular mechanisms
usually 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). The total toxic involved in the Aroclor 1254-induced hepatotoxicity, particularly
equivalent (TEQ) is operationally defined by the sum of the prod- the role of mitochondria, possibly a primary target in such event,
ucts of the concentration of each compound multiplied by its TEF could greatly explain the cytotoxic effect of Aroclor 1254 in rat
value and is an estimate of the total 2,3,7,8-TCDD-like activity of the hepatocytes.
mixture (Van den Berg et al., 2006).
Commercial PCB preparations, such as Aroclor 1254, are 2. Materials and methods
described as mixed-type inducers as they contain dioxin-like,
2.1. Reagents
mono-ortho, and di-ortho congeners and thus can induce both
CYP1A and CYP2B enzymes in laboratory animals (Ngui and Aroclor 1254 (lot 6024), a commercial mixture of PCBs, and reagents for cell
Bandiera, 1999). When induced, CYP isozymes may serve as a source culture were purchased from Sigma–Aldrich Chemical Company, St. Louis, MO, USA.
Other reagents were of analytical grade.
of reactive oxygen radicals and/or may catalyze the oxidation of a
broad range of endogenous and exogenous substances, including
2.2. Isolation and culture of rat hepatocytes
PCBs themselves (Twaroski et al., 2001c). The cytochrome P450-
catalyzed oxidation of lower chlorinated biphenyls gives rise to Hepatocytes were isolated from adult male albino rats using the in situ col-
mono- and dihydroxy metabolites. The latter can autoxidize or can lagenase perfusion method (Seglen, 1976). Cell viability was determined by the
be enzymatically oxidized to semiquinones and/or quinones (Song trypan blue exclusion method, and exceeded 90%. The isolated rat hepatocytes were
cultured, at a density of 106 cells/cm2 in six-well plates, in DMEM media supple-
et al., 2008a,b). Some PCB-quinones can undergo redox cycling, with mented with 10% FBS, 2 mM l-glutamine, 0.1 U/ml insulin, 0.1 ␮M dexamethasone
the formation of ROS, thus becoming another source of oxidative and 100 ␮g/ml kanamycin sulphate for 24 h at 37 ◦ C in a humidified atmosphere of
stress (McLean et al., 2000). 95% relative humidity with 95% O2 and 5% CO2 . Then, the cells were incubated for
H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183 177

24 h in a fresh media containing 0, 20, 30, 40, 50 or 60 ␮M of Aroclor 1254. At the 2 ,7 -dichlorofluorescein (DCF) (Vanden Hoek et al., 1997; Duranteau et al., 1998).
end of incubation, hepatocytes were recovered, sonicated in 0.1 M Tris–HCl buffer, An increase in fluorescence intensity is used to quantify the generation of net mito-
pH 7.4 and centrifuged for 15 min at 240 × g (4 ◦ C). The supernatant was used for the chondrial ROS. At the end of treatment with Aroclor 1254, the mitochondria were
enzymatic and non-enzymatic assays in hepatocytes after estimating the protein preloaded with 100 ␮M DCFH-DA dissolved in methanol and incubated for 30 min
content according to the method of Smith et al. (1985). at 37 ◦ C. The medium with the DCFH-DA was then removed, and the mitochondria
The Aroclor doses selected in the present study 0, 20, 30, 40, 50 or 60 ␮M were were washed, resuspended in ice-cold PBS. The formation of the fluorescent oxi-
based on the findings presented by Thome et al. (1995) indicating that Aroclor 1254 dized derivative of DCFH, namely DCF, was measured using a spectrofluorimeter
at concentrations exceeding 25 ␮M (but not at lower concentrations) caused irre- with excitation wavelength at 485 nm and emission wavelength at 535 nm. ROS for-
versible damage to cultured rat hepatocytes. Moreover, it is important to note that all mation was quantified from a DCF-standard curve and data were expressed as pmol
of the previously published studies of the effects of Aroclor 1254 on mitochondrial DCF formed/min/mg protein.
functions utilized concentrations in the 10–200 ␮M range, which yielded a variety
of effects (Salvi and Toninello, 2001). 2.8. EROD and PROD assays

2.3. Isolation of mitochondria from hepatocytes Mitochondrial ethoxyresorufin O-deethylase (EROD) and pentoxyresorufin O-
deethylase (PROD) activities are commonly used as enzymatic markers for hepatic
Mitochondria were isolated by differential centrifugation according to the modi- CYP1A and CYP2B enzymes (Burke et al., 1994). The activity of these enzymes was
fied protocol published by Choy et al. (2008). Briefly, hepatocytes were homogenized determined spectrofluorimetrically as described previously (Burke et al., 1985),
in isolation buffer (Tris–HCl 1 mM, EGTA 1 mM, sucrose 70 mM, mannitol 220 mM with some modifications. The reaction mixture contained 0.05 M Tris–HCl (pH
and HEPES 2 mM) and centrifuged for 10 min at 460 × g. The supernatant was col- 7.4), 2.6 ␮M ethoxyresorufin (EROD) or 9.6 ␮M pentoxyresorufin (PROD), NADPH-
lected and further centrifuged for 10 min at 12,500 × g, then was decanted. The regenerative system (1 mM NADP+ , 1 U/ml glucose-6-phosphate dehydrogenase,
middle part of the pellet was collected to mix with the isolation buffer and cen- 4.4 mM glucose-6-phosphate, 5 mM MgCl2 , 1 mg BSA) and 50 ␮l of mitochondrial
trifuged for another 10 min at 12,500 × g. The pellets containing mitochondria were extract in a total volume of 1 ml. The samples were incubated for 5 min at 37 ◦ C for
suspended in ice storage buffer (sucrose 250 mM, KCl 10 mM, and Tris–HCl 10 mM). measurement of the EROD activity, and 25 min for measurement of the PROD activ-
The mitochondria were used as such for the mitochondrial assay. The protein values ity. The reaction was initiated by adding 0.5 ml of NADPH-regenerative system and
of mitochondria were measured by the method of Smith et al. (1985). terminated by adding 2 ml of methanol. After incubation, tubes were centrifuged
for 10 min at 10,000 × g to pellet the flocculated proteins. The supernatant was then
2.4. Cell viability measured for fluorescence with excitation and emission wavelengths set at 544
and 590 nm, respectively. The same standard curve was used for both assays and
2.4.1. MTT assay was constructed by measuring the fluorescence of resorufin at various concentra-
Cell viability and activity of mitochondrial electron transport chain, as indicator tions (0–70 nM). The enzymes activities for EROD and PROD were expressed as pmol
for cytotoxicity, was determined by the capacity of cells to reduce MTT (3-(4,5- resorufin/min/mg protein.
dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) to formazan (Bruggisser et
al., 2002). This assay is based on the ability of living cells to (reductively) convert 2.9. Antioxidant enzyme assays
the dissolved MTT (yellow) into the insoluble formazan (blue) in the presence of
the mitochondrial enzymes succinate dehydrogenase. The latter can be measured Glutathione reductase (GR) activity was measured according to the procedure
colorimetrically and is proportional to the metabolic activity of the cells. At the end of Hsiao et al. (2001). Briefly, 10 ␮l of mitochondrial extract was added to a reac-
of treatment of cells with Aroclor 1254, 0.5% MTT was added for 4 h at 37 ◦ C. The tion mixture (containing 0.99 ml of 100 mM potassium phosphate buffer pH 7.0,
reaction was stopped with 100 ␮l sodium dodecyl sulfate 20% and absorption was 1.1 mM MgCl2 , 5 mM GSSG and 0.1 mM NADPH) to trigger the NADPH conversion
measured at 550 nm. reaction. Changes in absorbance were monitored by a continuous-recording spec-
trophotometer at 340 nm for 5 min at 25 ◦ C. The GR activity was expressed as nM
2.4.2. Lactate dehydrogenase (LDH) leakage assay NADPH oxidized/min/mg protein. Glutathione peroxidase (GPx) activity was deter-
LDH activity in hepatocytes was measured as an index of plasma membrane mined by the method of Paglia and Valentine (1967). The reaction mixture consisted
integrity (Farber and Young, 1981). At the end of treatment with Aroclor 1254, the of 0.8 ml of 100 mM potassium phosphate buffer (pH 7.0), 1 mM EDTA, 1 mM NaN3 ,
medium was removed and measured for LDH activity by recording the absorbance 0.2 mM NADPH, 1 U/ml GR and 1 mM GSH. Fifty microliters of mitochondrial extract
change at 340 nm in an assay medium containing 1 mM pyruvate, 0.15 mM NADH were added to make a total volume of 0.9 ml. The reaction was started by the addi-
and 0.1 M sodium phosphate (pH 8.0). The percentage of LDH release was expressed tion of 100 ␮l of 2.5 mM H2 O2 , and the conversion of NADPH to NADP was monitored
as the fraction of LDH released into the culture medium relative to the total amount by a continuous recording spectrophotometer at 340 nm for 3 min. GPx activity was
of LDH (cells + medium) activity after cell lysis with 0.1% (v/v) Triton-X 100 (Ye et al., expressed as nM NADPH oxidized/min/mg protein.
2007).
2.10. Glutathione (GSH) assay
2.5. Lipid peroxidation (LPO)
GSH (gamma-l-glutamyl-l-cysteinylglycine) content was determined by a flu-
The assay for lipid peroxidation in hepatocytes was performed following the orimetric assay using o-phthalaldehyde (Hissin and Hilf, 1976). The reaction
method described previously (Iqbal and Athar, 1998), by measuring the rate of mixture containing 0.1 M sodium phosphate buffer (pH 8.0), 5.0 mM EDTA, 10 ␮l o-
production of thiobarturic acid-reactive substances (TBARS) (expressed as malon- phthaldehyde (1.0 mg/ml) and 50 ␮l of mitochondrial protein. After incubation for
dialdehyde equivalents). The amount of malonaldehyde formed was assessed by 15 min at room temperature, the fluorescence intensity was recorded with excitation
measuring the optical density at 535 nm at 37 ◦ C. The results were expressed as and emission wavelengths of 420 and 350 nm, respectively. The reduced glutathione
nmol MDA formed/mg protein. content was expressed as nmol GSH/mg protein.

2.6. Mitochondrial membrane potential (␺m ) 2.11. Aconitase activity

The uptake and retention of the cationic fluorescent dye, rhodamine 123, in hep- Aconitase activity was assessed in the mitochondrial fraction by measuring the
atocytes has been used for the estimation of mitochondrial membrane potential. conversion of citrate to isocitrate, and subsequent production of ␣-ketoglutarate
This assay is based on the selective accumulation of rhodamine 123 in active mito- was determined spectrophotometrically at 340 nm. The reaction mixture contained
chondria by charge-facilitated diffusion. At the end of treatment with Aroclor 1254, 50 mM Tris–Cl buffer (pH 7.4) containing 30 mM citrate, 0.2 mM NADP+ , 0.6 mM
cells were incubated in a fresh medium containing 1.5 ␮M rhodamine 123 at 37 ◦ C for MnCl2 , 1–2 unit(s) of isocitrate dehydrogenase and 50 ␮l of mitochondrial extract
30 min. Adherent hepatocytes were washed twice with PBS, harvested by trypsiniza- in a total volume of 1 ml. The reaction was carried out at 25 ◦ C for 1 h. One unit of
tion followed by another wash with PBS. Fluorescence intensity was assessed at aconitase activity is equivalent to the production of 1 ␮mol of isocitrate per min
excitation wavelength of 490 nm and emission wavelength of 520 nm. The capac- (Chung et al., 2001).
ity of mitochondria to take up the rhodamine 123 was calculated as fluorescence
intensity of rhodamine 123 (% of control) (Shangari and O’Brien, 2004). 2.12. Mitochondrial ˇ-oxidation

2.7. Detection of mitochondrial reactive oxygen species (ROS) The ␤-oxidation of [1-14 C]palmitic acid by hepatocyte mitochondria was
assessed as described by Fréneaux et al. (1988) with some modifications. The
The production of mitochondrial ROS was measured by using the oxidation- preincubation medium (900 ␮l of 70 mmol/l sucrose, 43 mmol/l KCl, 3.6 mmol/l
sensitive fluorescent dye 2 ,7 -dichlorodihydrofluorescin diacetate (DCFH-DA). The MgCl2 , 7.2 mmol/l potassium phosphate and 36 mmol/l Tris–HCl, pH 7.4) contained
membrane-permeable diacetate form of the dye can be hydrolyzed by cellular non- 0.2 mmol/l ATP, 50 ␮mol/l l-carnitine, 15 ␮mol/l coenzyme A, 50 ␮l mitochondrial
specific esterases that cleave the acetate groups on DCFH-DA, thus trapping the protein and 5 mmol/l acetoacetate. After 5 min of preincubation at 30 ◦ C the incuba-
reduced form of the probe (DCFH) within the cell. ROS (hydrogen peroxide, as well tion mixture was brought to 1 ml by adding 100 ␮l of preincubation buffer containing
as other peroxides) in the cells oxidize DCFH, yielding the fluorescent product of [1-14 C]palmitic acid (final concentration 40 ␮mol/l; 0.05 ␮Ci/ml). The incubations
178 H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183

were carried out at 30 ◦ C for 15 min with slow shaking. It was previously shown that
the reaction is in the linear phase at this time point (Spaniol et al., 2001). The reaction
was stopped by adding 200 ␮l of 20% perchloric acid (w/v) and by placing the tubes
on ice for 20 min. After centrifugation for 5 min at 12,000 × g, the 14 C-acid-soluble
products were counted in the supernatant. Such products represent mainly ketone
bodies and citric acid cycle intermediates (Morse et al., 1988).

2.13. Cardiolipin content

Fifty microliters of mitochondrial protein was incubated at 30 ◦ C for 45 min in


the presence of 10-N-nonyl-acridine orange 5 ␮M (NAO), in a medium containing
160 mM KCl and 10 mM Hepes–KOH, pH 7.4. The excess dye was washed out by cen-
trifugation and the mitochondrial pellet appropriately diluted in the same buffer.
Fluorescence was determined in a spectrofluorimeter operating at 485 nm (excita-
tion) and 535 nm (emission). Fluorescence was converted to relative fluorescence
units using quinine (1 mg ml−1 in 0.1N H2 SO4 , ex 360 nm, em 457 nm) as reference Fig. 1. Dose–response effect of Aroclor 1254 on CYP1A (EROD) and CYP2B (PROD)
(Petit et al., 1992; Gallet et al., 1995). in rat hepatocyte mitochondria. Values are mean ± SD (n = 3). Statistical anal-
ysis (ANOVA) for differences from corresponding controls: *p < 0.05; **p < 0.01;
2.14. Respiratory chain enzyme activities ***p < 0.001.

In order to determine the effect of Aroclor 1254 on complexes I and III more
precisely, these enzyme complexes were investigated individually in the hepatocyte
mitochondria. 3.2. Lipid peroxidation, mitochondrial membrane potential
(␺m ) and ROS generation
2.14.1. Complex I
The complex I (NADH–CoQ reductase) activity was measured in mitochondrial Table 2 demonstrates that, Aroclor 1254 significantly increased
particles obtained by three times of freezing and thawing of heptocyte mitochondria
dissolved in 50 mM phosphate buffer pH 7.2 as previously described (Petrosillo et al.,
lipid peroxidation, secondary products, MDA (malondialdehyde) in
2007). The assay mixture contained 3 mM sodium azide, 1.2 ␮M antimycin A, 50 ␮M hepatocytes in a dose-related manner as compared to correspond-
decylubiquinone and 50 mM phosphate buffer pH 7.2. The mitochondrial extract ing control while ␺m was significantly decreased in the same
(50 ␮l) was added to 3 ml of the assay mixture and the reaction was started by the manner as compared to untreated cells (Table 2). Aroclor 1254 sig-
addition of 60 ␮M NADH. The reaction was measured by following the decrease in
nificantly increased ROS generation in hepatocyte mitochondria
absorbance of NADH at 340 nm in a spectrophotometer. The specific activity of the
enzyme is expressed as nmol of NADH oxidized/min/mg of mitochondrial protein. in a dose-related manner as compared to corresponding control
(Table 2).
2.14.2. Complex III
Succinate cytochrome c reductase activity was measured spectrophotometri-
cally (Tisdale, 1967) at 25 ◦ C by following the reduction of oxidized cytochrome c 3.3. EROD and PROD
(electron acceptor) by the increase in absorbance at 550 nm. The reaction was ini-
tiated by the addition of 5 mM succinate to 3 ml of the standard reaction mixture, Mitochondrial EROD and PROD were measured in the current
as previously described, supplemented with 2 mM rotenone, 1 mM KCN, 54 mM of
study to determine the effect of A1254 on this CYP subfamily.
cytochrome c (III), and 50 ␮l of mitochondrial protein.
Aroclor 1254 (60 ␮M) increased significantly EROD (fivefold) and
PROD (fourfold) activities in comparison to corresponding controls
3. Results
(Fig. 1). The increase in the activity of these two enzymes was in a
dose-related pattern. These results suggest that the Ahr (CYP1A) is
3.1. Characterization of Aroclor 1254 toxicity
critical for Aroclor-induced mitochondrial superoxide production
in hepatocytes. In addition, CYP2B may be a target for Aroclor 1254
To characterize Aroclor 1254 toxicity, we evaluated the effect of
toxicity in hepatocyte mitochondria.
increasing concentrations of Aroclor 1254 on the viability of hep-
atocytes and LDH content. Table 1 shows that hepatocytes treated
with 20 ␮M of Aroclor 1254 for 24 h did not show any significant 3.4. Antioxidant status in hepatocyte mitochondria
effect on either cell viability or LDH release into medium. Starting
at 30 ␮M for 24 h, Aroclor 1254 caused a significant decrease in cell Aroclor 1254-significantly decreased the enzymatic antioxidant
survival in a dose-related manner. On the other hand LDH released GR and GPx activities and the non-enzymatic antioxidant GSH con-
into medium was significantly increased in the same pattern in tent of hepatocyte mitochondria as compared to corresponding
comparison to related controls. controls (Table 3).

Table 1
Dose–response effect of Aroclor 1254 on cell viability and LDH release.

Control 20 ␮M 30 ␮M 40 ␮M 50 ␮M 60 ␮M

Cell viability 93.67 ± 1.53 90.33 ± 2.52 85.67 ± 1.53* 83 ± 2** 79 ± 2.65*** 75.67 ± 4.93***
% of LDH released 27.33 ± 2.52 30.67 ± 3.06 36 ± 1* 38.67 ± 3.21** 42.33 ± 4.16*** 51.67 ± 2.52***

Values are mean ± SD (n = 3). Statistical analysis (ANOVA) for differences from corresponding controls: * p < 0.05; ** p < 0.01; *** p < 0.001.

Table 2
Dose–response effect of Aroclor 1254 on lipid peroxidation, reactive oxygen species production and mitochondrial membrane potential (␺m ) in rat hepatocytes.

Control 20 ␮M 30 ␮M 40 ␮M 50 ␮M 60 ␮M

Lipid peroxidation 1.63 ± 0.23 2.37 ± 0.15* 2.6 ± 0.1** 2.8 ± 0.2*** 3.57 ± 0.21*** 4.37 ± 0.32***
␺m (% of control) 100 ± 9.61 78.45 ± 4.61* 70.67 ± 5.33** 59.99 ± 4.61*** 49.22 ± 7.05*** 38.46 ± 7.05***
ROS generation 3.2 ± 0.2 3.97 ± 0.21* 4.23 ± 0.21** 6.23 ± 0.21*** 8.37 ± 0.25*** 10.37 ± 0.32***

Lipid peroxidation: nmol of MDA equivalent formed/mg protein; ␺m : fluorescence intensity of rhodamine 123 (% of control); ROS: pmol DCF formed/min/mg protein. Values
are mean ± SD (n = 3). Statistical analysis (ANOVA) for differences from corresponding controls: * p < 0.05; ** p < 0.01; *** p < 0.001.
H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183 179

Table 3
Dose–response effect of Aroclor 1254 on ROS generation and antioxidant status of rat hepatocyte mitochondrial.

Control 20 ␮M 30 ␮M 40 ␮M 50 ␮M 60 ␮M

GPx 30.67 ± 2.08 24 ± 1*


22 ± 2.65 **
20.53 ± 2.84 **
17 ± 2***
14.67 ± 2.52***
GR 49.67 ± 2.08 42 ± 1.73* 40 ± 3* 36.93 ± 3** 33.83 ± 3.75*** 32.33 ± 2.52***
GSH 42 ± 2 34.67 ± 1.53* 32.33 ± 2.08** 28.67 ± 3.06*** 24.67 ± 2.52*** 21.17 ± 2.93***

GPx and GR: nmol NADPH oxidized/min/mg protein; GSH: nmol GSH/mg protein. Values are mean ± SD (n = 3). Statistical analysis (ANOVA) for differences from corresponding
controls: * p < 0.05; ** p < 0.01; *** p < 0.001.

Table 4
Dose–response effect of Aroclor 1254 on aconitase, ␤-oxidation and cardiolipin content in mitochondria of rat hepatocytes.

Control 20 ␮M 30 ␮M 40 ␮M 50 ␮M 60 ␮M

Aconitase 7.93 ± 0.4 6.7 ± 0.3*


5.97 ± 0.55 **
5 ± 0.03 ***
4.4 ± 0.46 ***
4 ± 0.5***
␤-Oxidation 0.68 ± 0.03 0.57 ± 0.07* 0.53 ± 0.04** 0.50 ± 0.03** 0.46 ± 0.02*** 0.41 ± 0.03***
Cardiolipin 34.33 ± 4.04 27 ± 1* 25 ± 2** 18.33 ± 1.53*** 15.33 ± 3.06*** 9.67 ± 1.53***

Aconitase: mU/mg protein; ␤-oxidation: [14 C] ␤-oxidation products (nmol/min/mg protein); cardiolipin: nmol/mg protein. Values are mean ± SD (n = 3). Statistical analysis
(ANOVA) for differences from corresponding controls: * p < 0.05; ** p < 0.01; *** p < 0.001.

Table 5
Dose–response effect of Aroclor 1254 on mitochondrial respiratory chain complexes I and III of rat hepatocytes.

Control 20 ␮M 30 ␮M 40 ␮M 50 ␮M 60 ␮M

Complex I 69 ± 3.61 57.33 ± 2.52** 48 ± 2*** 36.67 ± 4.16*** 26.67 ± 3.51*** 20.67 ± 3.06***
Complex III 248.38 ± 12.58 211.67 ± 7.64* 193 ± 12.12** 180.67 ± 11.02*** 165.67 ± 14.36*** 134 ± 14.42***

Complex I: nmol of NADH oxidized/min/mg protein; complex III: nmol of cytochrome c reduced/min/mg protein. Values are mean ± SD (n = 3). Statistical analysis (ANOVA)
for differences from corresponding controls: * p < 0.05; ** p < 0.01; *** p < 0.001.

3.5. Aconitase, ˇ-oxidation and cardiolipin content of hepatocyte was indeed due to the impairment of mitochondrial function by
mitochondria Aroclor 1254, the mitochondrial membrane potential was mea-
sured using the fluorescent dye rhodamine 123. The increase in
In order to evaluate the role of Aroclor 1254 in generating a supernatant LDH activity directly correlates to the reduced amount
mitochondrial oxidative stress, we examined mitochondrial aconi- of formazan formed in MTT assay. These findings are in line with
tase activity, an enzyme that is inactivated by superoxide. Aconitase Thome et al. (1995) who reported that Aroclor 1254 at concentra-
inactivation has been suggested as a reliable marker for mitochon- tions exceeding 25 ␮M causes irreversible damage to cultured rat
drial ROS production (Hausladen and Fridovich, 1994, 1996). Aroclor hepatocytes.
1254-treated hepatocytes significantly decreased mitochondrial Several studies had suggested that increased cellular oxidative
aconitase activity in a dose-related manner as compared to cor- stress be a critical underlying mechanism of PCB-mediated cell acti-
responding control (Table 4). Aroclor 1254 impaired mitochondrial vation and dysfunction (Mariussen et al., 2002; Twaroski et al.,
formation of acid-soluble products, which reflects ␤-oxidation. The 2001c). Therefore, this toxic effect of Aroclor 1254 on hepatocytes
content of the mitochondrial lipid cardiolipin was significantly may be due to the increased lipid peroxidation, as shown in this
decreased in a dose-related pattern, probably due to oxidation pro- study, which was based on the critical role played by lipids in main-
cess (Table 4). taining membrane structure and function, and ultimately cellular
viability, and ROS production, as also shown in this study, which in
3.6. Mitochondrial respiratory chain complexes I and III turn induces oxidative stress (Sridhar et al., 2004; Banudevi et al.,
2005). The increased lipid peroxidation may be due to increased
Complex I transfers electrons to complexes III and IV. Aroclor ROS or decreased redox status. These findings come in agreement
1254 significantly inhibited state 3 oxidation, complexes I and III, with Aly et al. (2009).
reflecting impairment of the function of the electron transport In addition to the cytotoxicity study on hepatocytes, the toxic
chain and uncouple oxidative phosphorylation (Table 5). effect of Aroclor 1254 was investigated on isolated rat hepatocyte
mitochondria in order to characterize the molecular mechanisms
4. Discussion involved in the hepatotoxicity, focusing on the role played by the
mitochondria, a possible early target in this process.
In the present study, Aroclor 1254, starting at 30 ␮M, seri- Mitochondria are the main cellular source and target of ROS.
ously damaged cell membrane integrity, decreased cell viability On the other hand, mitochondria possess an antioxidant defense
and significantly increased LDH release. These morphological and system that, under physiological conditions, decreases ROS pro-
biochemical changes in hepatocytes manifested clear evidence duction or removes generated ROS (Czarna and Jarmuszkiewicz,
of the toxicity of Aroclor 1254 on hepatocytes. As mitochon- 2006). Among, GPx, GR and reduced glutathione GSH are key com-
dria are often a target of toxicity, an MTT test was performed ponents of this mitochondrial antioxidant defense system (Droge,
to assess cytotoxicity and therefore cell viability, which mainly 2002).
reflects the mitochondrial reductive activity (Bruggisser et al., ROS have been implicated in the pathogenesis of several liver
2002). diseases (Petrosillo et al., 2007). ROS can alter vital cell compo-
Since reduction of MTT is dependent on the activity of the mito- nents like polyunsaturated fatty acids, proteins and nucleic acids.
chondrial electron transport chain, impaired formazan production To a lesser extent, carbohydrates are also the targets of ROS. These
from MTT is usually interpreted as indicative for mitochondrial reactions can alter intrinsic membrane properties like fluidity, ion
damage. To determine whether the reduced production of formazan transport, loss of enzyme activity, inhibition of protein synthe-
180 H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183

sis, DNA damage, ultimately resulting in cell death (Halliwell and and CYP1A2 genes are controlled by the Ahr; consequently, follow-
Gutteridge, 1990). Oxidative stress develops when there is an imbal- ing dioxin exposure there is a robust accumulation of CYP1A1 and
ance between prooxidants and antioxidants ratio, and as results it CYP1A2 in mitochondria of mouse liver (Senft et al., 2002). Such
leads to the accumulation of oxidative damaged molecules (Sakac findings corroborate with our results as we had seen the elevated
and Sakac, 2000). levels of total CYP1A and CYP2B activities as indicated by EROD and
The involvement of ROS in Aroclor 1254-induced toxicity was PROD, respectively. Moreover, Chu et al. (2008) demonstrated that
assayed using the fluorescence probe DCFH-DA. This is a sensitive Aroclor 1254 induced microsomal enzyme activities in rat liver.
and rapid method, which is well suited for detecting overall oxida- The cytochrome P450 monooxygenases sequentially transfer
tive stress after stimulation with ROS-forming compounds in low two electrons to oxygen from mitochondrial adrenodoxin, with sub-
concentrations (Myhre et al., 2000). sequent formation of oxygenated substrate and water (Poulos and
In the present study, ROS was shown to be induced by Aro- Raag, 1992). Although electron transfer is normally a well-coupled
clor 1254. These results come in accordance with Schilderman et process, superoxide and H2 O2 may be released, particularly in the
al. (2000), Banudevi et al. (2006) and Venkataraman et al. (2008) presence of CYP1A1 ligands that are poorly metabolized. Burqin
who reported that Aroclor 1254 induced ROS production which can et al. (2001) reported that Aroclor 1254 induced hepatic ethoxy-,
damage the cellular elements in its target organs such as rat liver, methoxy-, and pentoxyresorufin O-deethylase (EROD, MROD, and
spleen, brain, seminal vesicle, prostate and epididymis. Moreover, PROD, respectively) activities in rat liver. Recent study demon-
Aly et al. (2009) demonstrated that Aroclor 1254 induced ROS in rat strated that PCB-induced oxidative stress in rat kidney might be
testicular mitochondria and impaired spermatogenesis. Although attributable to induction of CYP1A1 (Lu et al., 2009). Further,
mitochondria are likely the major site of ROS generation (James et Bonfanti et al. (2009) reported that PCB induced hepatic CYP450
al., 2005; Talbot et al., 2004) in most cells, little is known about ROS (total CYP450, CYP1A, CYP1B). Therefore, it is reasonable to sup-
production within hepatocyte mitochondria. Here we report new pose that CYP-system may be implicated in Aroclor 1254 triggered
information concerning the mechanism leading to hepatocyte ROS ROS production in hepatocyte mitochondria. The increase in ROS
production at the mitochondrial level. generation would lead to irreversible damage to mitochondrial
Respiratory chain complex I transfers the electrons to com- membrane lipids and proteins resulting in mitochondrial dys-
plexes III and IV. Our study shows that, Aroclor 1254 inhibited function and ultimately in cell death (Kowaltowski and Vercesi,
mitochondrial respiratory chain complexes I and III. This inhi- 1999).
bition of activity reflects mitochondrial impairment. Damaged The decreased levels of cardiolipin in the Aroclor 1254-treated
mitochondrial respiratory chain, as shown in the present study, hepatocytes, as shown in this study, might also have accounted for
is considered an important source of ROS (Perez-Carreras et al., the alteration of the mitochondrial membrane lipid order, as car-
2003). It is conceivable that mitochondrial-mediated ROS gener- diolipin is a mitochondrial anionic phospholipid known to confer
ation leads to primary reactions and damages in the immediate stability and fluidity to the mitochondrial membrane (Fariss et al.,
area surrounding where these ROS are produced, given that they 2005). Moreover, cardiolipin also functions as a support for the elec-
are a highly reactive and short-lived species. Therefore, as a major tron transport chain, since it anchors complexes III and IV, forming
source of ROS production, mitochondria also could be the major a supercomplex (Hauff and Hatch, 2006). Furthermore, cardiolipin
target of ROS attack. This effect of ROS should be greatest at the has been shown to be specifically required for the optimal function-
level of mitochondrial membrane constituents including the com- ing of mitochondrial complex I (Sharpley et al., 2006). Therefore, our
plexes of the respiratory chain and phospholipid constituents rich findings are in accordance with this proposition, since the oxida-
in unsaturated fatty acids, particularly cardiolipin (Martins et al., tion of cardiolipin could have caused disruption of the respiratory
2008). chain, observed in this study. Further our result concerning cardi-
Beside the effect on the respiratory chain, ␤-oxidation of fatty olipin corroborate with Zhou and Zhang (2005) who reported that
acids represents another target of mitochondrial toxicity. In the A1254 interfered with the integrity of cell membrane and caused
present study, Aroclor 1254 inhibited mitochondrial ␤-oxidation. hepatocyte damage. Thus, it is reasonable that ROS-induced oxida-
Since the activity of the electron transport chain can be rate-limiting tive damage to mitochondrial cardiolipin may be responsible, at
for mitochondrial fatty acid metabolism (Latipää et al., 1986), the least in part, for the observed defect in complex I activity in Aroclor
inhibition of ␤-oxidation may be secondary to impaired function of 1254-treated hepatocytes.
the electron transport chain by Aroclor 1254. Since inhibition of the respiratory chain and cytochrome P450
The findings of the current study so far have shown that can be associated with the generation of ROS (Kaufmann et al.,
Aroclor 1254 inhibited the respiratory chain and mitochondrial 2005), the redox status of the mitochondria of Aroclor 1254-treated
␤-oxidation, can lead to a decrease in the ␺m . Our data clearly hepatocyte was assessed by measuring GR and GPx activities and
show that Aroclor 1254 caused mitochondrial dysfunction, char- the reduced glutathione content.
acterized by decreased ␺m which may subsequently depress Our results showed that in Aroclor 1254-treated hepatocytes, GR
cellular viability and leading to cell death. Conditions favoring and GPx activities in the hepatocyte mitochondria were decreased.
increased mitochondrial generation of reactive oxygen species (e.g., The decreased activities of GR and GPx may be due to excessive ROS
the presence of high Ca2+ concentrations or when the mitochondrial production in the mitochondria as shown in this study. Meanwhile,
antioxidant defense mechanisms are compromised) may induce the the decreased activity in GR and GPx may be responsible for the ele-
mitochondrial pore transition (MPT) and death of the cell (Hoek et vated lipid peroxidation. This study also showed that Aroclor 1254
al., 1995). The consequence of ␺m disruption is the uncoupling (≥30 ␮M)-treated hepatocytes resulted in a significant decrease of
of the oxidative phosphorylation and the generation of superoxide GSH content in mitochondria, which indicated that damage of cellu-
anion via the uncoupled respiratory chain (Wilson and Prochaska, lar antioxidant systems was induced. Castell et al. (1997) suggested
1990). that GSH depletion might ultimately lead to cell death by impairing
Another proposed mechanism for ROS generation and induction the cell defense against oxidative damage.
of lipid peroxidation by Aroclor 1254 is the induction of cytochrome GSH is an important cellular antioxidant that plays a major role
P450 enzymes, with the subsequent release of ROS (Schlezinger et in protecting cells against oxidative stress and involves in antiox-
al., 1999). We focused in this study to determine whether the CYP1A idant defense of hepatocytes. GSH may also serve as a cofactor
and CYP2B are determinants in the pathway leading to Aroclor for (a) several drug-metabolizing enzymes (i.e. GSTs) where it is
1254-induced toxicity in the hepatocyte mitochondria. The CYP1A1 consumed, or (b) for antioxidant enzymes (i.e. GPX) where it func-
H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183 181

tions as a redox partner (Uhlig and Wendel, 1992). Mitochondrial mixture on cytochrome P450 and P-glycoprotein expressions in fetuses and
GSH detoxifies free oxygen radicals physiologically generated by pregnant rats. Chemosphere 75, 572–579.
Bray, T.M., Taylor, C.G., 1993. Tissue glutathione, nutrition, and oxidative stress. Can.
the respiratory chain (Droge, 2002). During oxidation, GSH forms a J. Physiol. Pharmacol. 71, 746–751.
dimer, glutathione disulfide (GSSG), which, in turn, can be reduced Bruggisser, R., von Daeniken, K., Jundt, G., Schaffner, W., Tullberg-Reinert, H., 2002.
by the enzyme GR at the expense of NADPH (Bray and Taylor, 1993). Interference of plant extracts, phytoestrogens and antioxidants with the MTT
tetrazolium assay. Planta Med. 68, 445–448.
The decrease in GR and GPx activities and GSH levels in mitochon- Burke, M.D., Thompson, S., Elcombe, C.R., Halpert, J., Haaparanta, T., Mayer, R.T., 1985.
dria may possibly contribute to the Aroclor 1254-induced toxicity. Ethoxy-, pentoxy- and benzyloxyphenoxazones and homologues: a series of sub-
Enhanced generation of ROS and/or impaired antioxidant detoxifi- strates to distinguish between different induced cytochromes P-450. Biochem.
Pharmacol. 34, 3337–3345.
cation functions contribute to an imbalance between oxidative and Burke, M.D., Thompson, S., Weaver, R.J., Wolf, C.R., Mayer, R.T., 1994. Cytochrome
reductive reactions, which alters thiol/disulfide redox (Moriarty- P450 specificities of alkoxyresorufin O-dealkylation in human and rat liver.
Craige and Jones, 2004). Biochem. Pharmacol. 48, 923–936.
Burqin, D.E., Diliberto, J.J., Derr-Yellin, E.C., Kannan, N., Kodavanti, P.R., Birnbaum, L.S.,
Mitochondria are the organelles of eukaryotic cells for carbohy-
2001. Differential effects of two lots of aroclor 1254 on enzyme induction, thyroid
drate metabolism, and mitochondrial oxidative phosphorylation is hormones, and oxidative stress. Environ. Health Perspect. 109, 1163–1168.
the major energy transduction pathway. Aconitase activity, which Castell, J.V., Gomez-Lechon, M.J., Ponsoda, X., Bort, R., 1997. In vitro investigation of
represents mitochondrial function, was measured by the turnover the molecular mechanisms of hepatotoxicity. In: Castell, J.V., Gomez-Lechon, M.J.
(Eds.), In Vitro Methods in Pharmaceutical Research. Academic Press, London, pp.
of citrate to isocitrate as an index of cellular energy utilization. The 375–410.
activity of aconitase enzyme was monitored in mitochondria. This Choy, C.S., Cheah, K.P., Chiou, H.Y., Li, J.S., Liu, Y.H., Yong, S.F., Chiu, W.T., Liao, J.W.,
enzyme can be used to estimate indirectly the superoxide concen- Hu, C.M., 2008. Induction of hepatotoxicity by sanguinarine is associated with
oxidation of protein thiols and disturbance of mitochondrial respiration. J. Appl.
tration, since its activity is very sensitive to superoxide exposure Toxicol. 28, 945–956.
and its inactivation, as shown in this study, is a reliable marker for Chu, I., Bowers, W.J., Caldwell, D., Nakai, J., Wade, M.G., Yagminas, A., Li, N., Moir, D., El
mitochondrial superoxide production (Gardner, 2002; James et al., Abbas, L., Håkansson, H., Gill, S., Mueller, R., Pulido, O., 2008. Toxicological effects
of in utero and lactational exposure of rats to a mixture of environmental con-
2005). taminants detected in Canadian Arctic human populations. J. Toxicol. Environ.
In summary, our results showed that Aroclor 1254 caused Health A 71, 93–108.
hepatotoxicity by induction of oxidative stress in hepatocyte mito- Chung, H.C., Kim, S.H., Lee, M.G., Cho, C.K., Kim, T.H., Lee, D.H., Kim, S.G., 2001.
Mitochondrial dysfunction by gamma-irradiation accompanies the induction of
chondria, by promoting mitochondrial structural and functional cytochrome P450 2E1 (CYP2E1) in rat liver. Toxicology 161, 79–91.
damage, inhibition of mitochondrial respiratory complexes I and Czarna, M., Jarmuszkiewicz, W., 2006. Role of mitochondria in reactive oxygen
III and ␤-oxidation of fatty acids, reduction in ␺m , depletion of species generation and removal, relevance to signaling and programmed cell
death. Postepy. Biochem. 52, 145–156.
antioxidant defense system with consequent redox status alter-
Denomme, M.A., Bandiera, S., Lambert, I., Copp, L., Safe, L., Safe, S., 1983. Polychlo-
ation, decrease in mitochondrial cardiolipin and aconitase content, rinated biphenyls as phenobarbitone-type inducers of microsomal enzymes.
increase in mitochondrial EROD and PROD activity and cellular Structure–activity relationships for a series of 2,4-dichloro-substituted con-
death. Therefore, we can conclude that Aroclor 1254 induced rat geners. Biochem. Pharmacol. 32, 2955–2963.
Droge, W., 2002. Free radicals in the physiological control of cell function. Physiol.
hepatocyte toxicity and our findings provide evidence to propose Rev. 82, 47–95.
that mitochondria are one of the most important and earliest cell Duranteau, J., Chandel, N.S., Kulisz, A., Shao, Z., Schumacker, P.T., 1998. Intracellular
targets in Aroclor 1254-mediated toxicity and delineate several signaling by reactive oxygen species during hypoxia in cardiomyocytes. J. Biol.
Chem. 273, 11619–11624.
mitochondrial processes at least, in part, by induction of oxidative Fadhel, Z.A., Lu, Z., Robertson, L.W., Glauert, H.P., 2000. Role of cytochrome P450
stress. These findings can be useful in future cytoprotective ther- induction and lipid peroxidation in 3,3 ,4,4 -tetrachlorobiphenyl (PCB-77) and
apy approaches. Since mitochondrial events appear to be targeted 2,2 ,4,4 ,5,5 -hexachlorobiphenyl (PCB-153)-induced toxicity in liver of male
rats. Toxicol. Sci. 54, 78.
in hepatocellular damage induced by Aroclor 1254, an antioxidant Fadhel, Z., Lu, Z., Robertson, L.W., Glauert, H.P., 2002. Effect of 3,3 ,4,4 -
therapy targeted to mitochondria may constitute an interesting tetrachlorobiphenyl and 2,2 ,4,4 ,5,5 -hexachlorobiphenyl on the induction of
strategy to ameliorate its toxicity. hepatic lipid peroxidation and cytochrome P-450 associated enzyme activities
in rats. Toxicology 175, 15–25.
Farber, J.L., Young, E.E., 1981. Accelerated phospholipid degradation in anoxic rat
Conflict of interest hepatocytes. Arch. Biochem. Biophys. 211, 312–320.
Fariss, M.W., Chan, C.B., Patel, M., Van Houten, B., 2005. Role of mitochondria in toxic
oxidative stress. Mol. Interv. 5, 94–111.
None declared. Fréneaux, E., Labbe, G., Letteron, P., The Le Dinh, Degott, C., Genève, J., Larrey, D.,
Pessayre, D., 1988. Inhibition of the mitochondrial oxidation of fatty acids by
tetracycline in mice and in man: possible role in microvesicular steatosis induced
References by this antibiotic. Hepatology 8, 1056–1062.
Gallet, P.F., Maftah, A., Petit, J.M., Denis-Gay, M., Julien, R., 1995. Direct cardiolipin
Adebusoye, S.A., Ilori, M.O., Picardal, F.W., Amund, O.O., 2008. Extensive biodegra- assay in yeast using the red fluorescence emission of 10-N-nonyl acridine orange.
dation of polychlorinated biphenyls in Aroclor 1242 and electrical transformer Eur. J. Biochem. 15, 113–119.
fluid (Askarel) by natural strains of microorganisms indigenous to contaminated Gardner, P.R., 2002. Aconitase: sensitive target and measure of superoxide. Methods
African systems. Chemosphere 73, 126–132. Enzymol. 349, 9–23.
Aly, H.A.A., Domenech, O., Abdel-Naim, A.B., 2009. Aroclor 1254 impairs spermato- Giesy, J.P., Kannan, K., 1998. Dioxin-like and non-dioxin-like toxic effects of polychlo-
genesis and induces oxidative stress in rat testicular mitochondria. Food Chem. rinated biphenyls (PCBs): implications for risk assessment. Crit. Rev. Toxicol. 28,
Toxicol., Mar 21. 511–569.
Andric, N.L., Andric, S.A., Zoric, S.N., Kostic, T.S., Stojilkovic, S.S., Kovacevic, R.Z., Halliwell, B., Gutteridge, J.M.C., 1990. Role of free radicals and catalytic metal ions in
2003. Parallelism and dissociation in the actions of an Aroclor 1260-based trans- human disease: an overview. Meth. Enzymol. 186, 1–85.
former fluid on testicular androgenesis and antioxidant enzymes. Toxicology Hauff, K.D., Hatch, G.M., 2006. Cardiolipin metabolism and Barth syndrome. Prog.
194, 65–75. Lipid Res. 45, 91–101.
Baird, L., 1995. Toxic organic chemicals. In: Environmental Chemistry. WH Freeman Hausladen, A., Fridovich, I., 1994. Superoxide and peroxynitrite inactivate aconitases,
and Company, New York, pp. 252–274. but nitric oxide does not. J. Biol. Chem. 269, 29405–29408.
Banudevi, S., Arunkumar, A., Sharmila, M., Senthilkumar, J., Balasubramanian, K., Hausladen, A., Fridovich, I., 1996. Measuring nitric oxide and superoxide: rate con-
Srinivasan, N., Aruldhas, M.M., Arunakaran, J., 2005. Diallyl disulfide-induced stants for aconitase reactivity. Methods Enzymol. 269, 37–41.
modulation of a few phase I and II drug metabolizing enzymes on Aroclor 1254 Hissin, P.J., Hilf, R., 1976. A fluorometric method for determination of oxidized and
toxicity in Rattus norvegicus liver and ventral prostate. J. Clin. Biochem. Nutr. 36, reduced glutathione in tissues. Anal. Biochem. 74, 214–226.
59–65. Hoek, J.B., Farber, J.L., Thomas, A.P., Wang, X., 1995. Calcium ion-dependent signalling
Banudevi, S., Krishnamoorthy, G., Venkataraman, P., Vignesh, C., Aruldhas, M.M., and mitochondrial dysfunction: mitochondrial calcium uptake during hormonal
Arunakaran, J., 2006. Role of alpha-tocopherol on antioxidant status in liver, stimulation in intact liver cells and its implication for the mitochondrial perme-
lung and kidney of PCB exposed male albino rats. Food Chem. Toxicol. 44, ability transition. Biochim. Biophys. Acta 1272, 93–102.
2040–2046. Hsiao, G., Lin, Y.H., Lin, C.H., Chou, D.S., Lin, W.C., Sheu, J.R., 2001. The protective
Bonfanti, P., Colombo, A., Villa, S., Comelli, F., Costa, B., Santagostino, A., 2009. The effects of PMC against chronic carbon tetrachloride-induced hepatotoxicity in
effects of accumulation of an environmentally relevant polychlorinated biphenyl vivo. Biol. Pharm. Bull. 4, 1271–1276.
182 H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183

Ikegwuonu, F.I., Ganen, L.G., Larsen, M.C., Shen, X., Jefcoate, C.R., 1996. The regu- Poland, A., Knutson, J.C., 1982. 2,3,7,8-Tetrachlorodibenzo-p-dioxin and related halo-
lation by gender, strain, dose, and feeding status of the induction of multiple genated aromatic hydrocarbons: examination of the mechanism of toxicity.
forms of cytochrome P450 isozymes in rat hepatic microsomes by 2,4,5,29,49,59- Annu. Rev. Pharmacol. Toxicol. 22, 517–554.
hexachlorobiphenyl. Toxicol. Appl. Pharmacol. 139, 33–41. Poulos, T.L., Raag, R., 1992. Cytochrome P450cam: crystallography, oxygen activation,
Iqbal, M., Athar, M., 1998. Attenuation of iron-nitrilotriacetate mediated renal and electron transfer. FASEB J. 6, 674–679.
oxidative stress, toxicity and hyperproliferative response by the prophylactic Risebrough, R.W., Rieche, P., Peakall, D.B., Herman, S.G., Kirven, M.N., 1968. Polychlo-
treatment of rats with garlic oil. Food Chem. Toxicol. 36, 485–495. rinated biphenyls in the global ecosystem. Nature 220, 1098–1102.
James, A.M., Cochemé, H.M., Smith, R.A., Murphy, M.P., 2005. Interactions of Roue, G., Bitton, N., Yuste, V.J., Montange, T., Rubio, M., Dessauge, F., Delettre, C.,
mitochondria-targeted and untargeted ubiquinones with the mitochondrial Merle-Beral, H., Sarfati, M., Susin, S.A., 2003. Mitochondrial dysfunction in CD47-
respiratory chain and reactive oxygen species. Implications for the use of mediated caspase-independent cell death: ROS production in the absence of
exogenous ubiquinines as therapies and experimental tools. J. Biol. Chem. 280, cytochrome c and AIF release. Biochimie 85, 741–746.
21295–212312. Sadeghi-Aliabadi, H., Chan, K., Lehmler, H.J., Robertson, L.W., O’Brien, P.J.,
Kaufmann, P., Török, M., Hänni, A., Roberts, P., Gasser, R., Krähenbühl, S., 2005. Mech- 2007. Molecular cytotoxic mechanisms of catecholic polychlorinated biphenyl
anisms of benzarone and benzbromarone-induced hepatic toxicity. Hepatology metabolites in isolated rat hepatocytes. Chem. Biol. Interact. 167, 184–192.
41, 925–935. Safe, S.H., 1994. Polychlorinated biphenyls (PCBs): environmental impact, biochem-
Kimbrough, R.D., 1995. Polychlorinated biphenyls (PCBs) and human health: an ical and toxic responses, and implications for risk assessment. Crit. Rev. Toxicol.
update. Crit. Rev. Toxicol. 25, 133–163. 24, 87–149.
Kirkland, R.A., Franklin, J.L., 2003. Bax, reactive oxygen, and cytochrome c release in Sakac, V., Sakac, M., 2000. Free oxygen radicals and kidney diseases—part I. Med.
neuronal apoptosis. Antioxid. Redox. Signal. 5, 589–596. Pregl. 53, 463–474.
Kowaltowski, A.J., Vercesi, A.E., 1999. Mitochondrial damage induced by conditions Salvi, M., Toninello, A., 2001. Aroclor 1254 inhibits the mitochondrial permeability
of oxidative stress. Free Radical Bio. med. 26, 463–471. transition and release of cytochrome c: a possible mechanism for its in vivo
Lambert, C., Apel, K., Biesalski, H.K., Frank, J., 2004. 2-Methoxyestradiol induces toxicity. Toxicol. Appl. Pharmacol. 176, 92–100.
caspase-independent, mitochondria-centered apoptosis in DS-sarcoma cells. Schilderman, P.A., Mass, L.M., Pachen, D.M., Dekok, T.M., Kleinjans, J.C., Vanschooten,
Int. J. Cancer 108, 493–501. F.S., 2000. Induction of DNA adducts by several polychlorinated biphenyls. Env-
Larsen, J.C., 2006. Risk assessments of polychlorinated dibenzo-p-dioxins, polychlo- iron. Mol. Mutagen. 36, 79–86.
rinated dibenzofurans, and dioxin-like polychlorinated biphenyls in food. Mol. Schlezinger, J.J., White, R.D., Stegeman, J.J., 1999. Oxidative inactivation of
Nutr. Food Res. 50, 885–896. cytochrome p-4501A (CYP1A) stimulated by 3,3 ,4,4 -tetrachlorobiphenyl: pro-
Latipää, P.M., Kärki, T.T., Hiltunen, J.K., Hassinen, I.E., 1986. Regulation of palmitoyl- duction of reactive oxygen by vertebrate CYP1As. Mol. Pharmacol. 56, 588–597.
carnitine oxidation in isolated rat liver mitochondria. Role of the redox state of Schuetz, E.G., Brimer, C., Schuetz, J.D., 1998. Environmental xenobiotics and the anti-
NAD(H). Biochim. Biophys. Acta 875, 293–300. hormones cyproterone acetate and spironolactone use the nuclear hormone
Liu, M.J., Wang, Z., Li, H.X., Wu, R.C., Liu, Y.Z., Wu, Q.Y., 2004. Mitochon- pregnenolone X receptor to activate the CYP3A23 hormone response element.
drial dysfunction as an early event in the process of apoptosis induced by Mol. Pharmacol. 54, 1113–1117.
woodfordin I in human leukemia K562 cells. Toxicol. Appl. Pharmacol. 194, Seglen, P.O., 1976. Preparation of isolated rat liver cells. Methods Cell Biol. 13, 29–83.
141–155. Senft, A.P., Dalton, T.P., Nebert, D.W., Genter, M.B., Puga, A., Hutchinson, R.J., Kerzee,
Lu, C.F., Wang, Y.M., Peng, S.Q., Zou, L.B., Tan, D.H., Liu, G., Fu, Z., Wang, J.K., Uno, S., Shertzer, H.G., 2002. Mitochondrial reactive oxygen production is
Q.X., Zhao, J., 2009. Combined Effects of Repeated Administration of 2,3,7,8- dependent on the aromatic hydrocarbon receptor. Free Radic. Biol. Med. 33,
Tetrachlorodibenzo-p-dioxin and Polychlorinated Biphenyls on Kidneys of Male 1268–1278.
Rats. Arch Environ Contam Toxicol., Apr 17. Shangari, N., O’Brien, P.J., 2004. The cytotoxic mechanism of glyoxal involves oxida-
Mariussen, E., Myhre, O., Reistad, T., Fonnum, F., 2002. The polychlorinated biphenyl tive stress. Biochem. Pharmacol. 68, 1433–1442.
mixture Aroclor 1254 induces death of rat cerebellar granule cells: the involve- Sharpley, M.S., Shannon, R.J., Draghi, F., Hirst, J., 2006. Interactions between
ment of the N-methyl-d-aspartate receptor and reactive oxygen species. Toxicol. phospholipids and NADH:ubiquinone oxidoreductase (complex I) from bovine
Appl. Pharmacol. 179, 137–144. mitochondria. Biochemistry 45, 241–248.
Martins, N.M., Santos, N.A.G., Curti, C., Bianchi, M.L.P., Santos, A.C., 2008. Cisplatin Shih, C.M., Ko, W.C., Wu, J.S., Wei, Y.H., Wang, L.F., Chang, E.E., Lo, T.Y., Cheng,
induces mitochondrial oxidative stress with resultant energetic metabolism H.H., Chen, C.T., 2004. Mediating of caspaseindependent apoptosis by cadmium
impairment, membrane rigidification and apoptosis in rat liver. J. Appl. Toxicol. through the mitochondria-ROS pathway in MRC-5 fibroblasts. J. Cell Biochem.
28, 337–344. 91, 384–397.
McLean, M.R., Twaroski, T.P., Robertson, L.W., 2000. Redox cycling of 2-(x’-mono,- Silberhorn, E.M., Glauert, H.P., Robertson, L.W., 1990. Carcinogenicity of polyhalo-
di,-trichlorophenyl)-1, 4-benzoquinones, oxidation products of polychlorinated genated biphenyls: PCBs and PBBs. Crit. Rev. Toxicol. 20, 440–496.
biphenyls. Arch. Biochem. Biophys. 376, 449–455. Simon, T., Britt, J.K., James, R.C., 2007. Development of a neurotoxic equivalence
Moriarty-Craige, S.E., Jones, D.P., 2004. Extracellular thiols and thiol/disulfide redox scheme of relative potency for assessing the risk of PCB mixtures. Regul. Toxicol.
in metabolism. Annu. Rev. Nutr. 24, 481–509. Pharm. 48, 148–170.
Morse, R.M., Valenzuela, G.A., Greenwald, T.P., Eulie, P.J., Wesley, R.C., McCallum, R.W., Smith, P.K., Krohn, R.I., Hermanson, G.T., Mallia, A.K., Gartner, F.H., Provenzano, M.D.,
1988. Amiodarone-induced liver toxicity. Ann. Intern. Med. 109, 838–840. Fujimoto, E.K., Goeke, N.M., Olson, B.J., Klenk, D.C., 1985. Measurement of protein
Myhre, O., Vestad, T.A., Sagstuen, E., Aarnes, H., Fonnum, F., 2000. The effects using bicinchoninic acid. Anal. Biochem. 150, 76–85.
of aliphatic (n-nonane), naphtenic (1,2,4-trimethylcyclohexane) and aromatic Song, Y., Buettner, G.R., Parkin, S., Wagner, B.A., Robertson, L.W., Lehmler, H.J., 2008b.
(1,2,4-trimethylbenzene) hydrocarbons on respiratory burst in human neu- Chlorination increases the persistence of semiquinone free radicals derived
trophil granulocytes. Toxicol. Appl. Pharmacol. 167, 222–230. from polychlorinated biphenyl hydroquinones and quinones. J. Org. Chem. 73,
Ngui, J.S., Bandiera, S.M., 1999. Induction of hepatic CYP2B is a more sensitive indica- 8296–8304.
tor of exposure to aroclor 1260 than CYP1A in male rats. Toxicol. Appl. Pharmacol. Song, Y., Wagner, B.A., Lehmler, H.J., Buettner, G.R., 2008a. Semiquinone radicals
161, 160–167. from oxygenated polychlorinated biphenyls: electron paramagnetic resonance
Paglia, D.E., Valentine, W.N., 1967. Studies on quantitative and qualitative char- studies. Chem. Res. Toxicol. 21, 1359–1367.
acterization of erythrocyte glutathione peroxidase. J. Lab. Clin. Med. 70, Spaniol, M., Bracher, R., Ha, H.R., Follath, F., Krähenbühl, S., 2001. Toxicity of amio-
158–169. darone and amiodarone analogues on isolated rat liver mitochondria. J. Hepatol.
Parkinson, A., Safe, S.H., Robertson, L.W., Thomas, P.E., Ryan, D.E., Reik, L.M., Levin, 35, 628–636.
W., 1983. Immunochemical quantitation of cytochrome P-450 isozymes and Sridhar, M., Venkataraman, P., Siva, D., Arunkumar, A., Aruldhas, M.M., Srinivasan,
epoxide hydrolase in liver microsomes from polychlorinated or polybrominated N., Arunakaran, J., 2004. Impact of polychlorinated biphenyl (Aroclor 1254) and
biphenyl-treated rats. A study of structure–activity relationships. J. Biol. Chem. vitamin C on antioxidant system of rat ventral prostate. Asian J. Androl. 6, 19–22.
258, 5967–5976. Talbot, D.A., Lambert, A.J., Brand, M.D., 2004. Production of endogenous matrix
Perez-Carreras, M., Del Hoyo, P., Martin, M.A., Rubio, J.C., Martin, A., Castellano, G., superoxide from mitochondrial complex I leads to activation of uncoupling pro-
Colina, F., Arenas, J., Solis-Herruzo, J.A., 2003. Defective hepatic mitochondrial tein 3. FEBS Lett. 556, 111–115.
respiratory chain in patients with nonalcoholic steatohepatitis. Hepatology 38, Tharappel, J.C., Lee, E.Y., Robertson, L.W., Spear, B.T., Glauert, H.P., 2002. Regula-
999–1007. tion of cell proliferation, apoptosis, and transcription factor activities during the
Pessah, I.N., Hansen, L.G., Albertson, T.E., Garner, C.E., Ta, T.A., Do, Z., et al., 2006. promotion of liver carcinogenesis by polychlorinated biphenyls. Toxicol. Appl.
Structure–activity relationship for noncoplanar polychlorinated biphenyl con- Pharmacol. 179, 172–184.
geners toward the ryanodine receptor–Ca2+ channel complex type 1 (RyR1). Tharappel, J.C., Lehmler, H.J., Srinivasan, C., Robertson, L.W., Spear, B.T., Glauert,
Chem. Res. Toxicol. 19, 92–101. H.P., 2008. Effect of antioxidant phytochemicals on the hepatic tumor promot-
Pessah, I.N., Lehmler, H.-J., Robertson, L.W., Perez, C.F., Cabrales, E., Bose, D.D., et al., ing activity of 3,3 ,4,4 -tetrachlorobiphenyl (PCB-77). Food Chem. Toxicol. 46,
2009. Enantiomeric specificity of (−)-2,2 ,3,3 ,6,6 -hexachlorobiphenyl toward 3467–3474.
ryanodine receptor types 1 and 2. Chem. Res. Toxicol. 22, 201–207. Thome, J.P., Roelandt, L., Goffinet, G., Stouvenakers, N., Kremers, P., 1995. Cytotoxic
Petit, J.M., Maftah, A., Ratinaud, M.H., Julien, R., 1992. 10-N-nonyl acridine orange effects of Aroclor 1254 on ultrastructure and biochemical parameters in cultured
interacts with cardiolipin and allows the quantification of this phospholipid in foetal rat hepatocytes. Toxicology 98, 83–94.
isolated mitochondria. Eur. J. Biochem. 209, 267–273. Tisdale, H.D., 1967. Preparation and properties of succinic-cytochrome c reductase
Petrosillo, G., Portincasa, P., Grattagliano, I., Casanova, G., Matera, M., Ruggiero, F.M., (complex II and III). Methods Enzymol. 10, 213–215.
Ferri, D., Paradies, G., 2007. Mitochondrial dysfunction in rat with nonalcoholic Twaroski, T.P., O’Brien, M.L., Larmonier, N., Glauert, H.P., Robertson, L.W., 2001a.
fatty liver involvement of complex I, reactive oxygen species and cardiolipin. Polychlorinated biphenyl-induced effects on metabolic enzymes AP-1 binding,
Biochim. Biophys. Acta 1767, 1260–1267. vitamin E, and oxidative stress in rat liver. Toxicol. Appl. Pharmacol. 171, 85–93.
H.A.A. Aly, Ò. Domènech / Toxicology 262 (2009) 175–183 183

Twaroski, T.P., O’Brien, M.L., Robertson, L.W., 2001b. In: Robertson, L.W., Hansen, Venkataraman, P., Krishnamoorthy, G., Vengatesh, G., Srinivasan, N., Aruldhas, M.M.,
L.G. (Eds.), Recent Advances in the Environmental Toxicology and Health Effects. Arunakaran, J., 2008. Protective role of melatonin on PCB (Aroclor 1254) induced
Effects of PCB 77 (3,3 ,4,4 -Tetrachlorobiphenyl) on Glutathione Peroxidase oxidative stress and changes in acetylcholine esterases and membrane bound
Activity and Regulatory Mechanisms. The University Press of Kentucky, pp. ATPases in cerebellum, cerebral cortex and hippocampus of adult rat brain. Int.
317–319. J. Dev. Neurosci. 26, 585–591.
Twaroski, T.P., O’Brien, M.L., Robertson, L.W., 2001c. Effects of selected polychlo- Williams, C.S., Jones, B.C., Nchege, O., Chung, R.A., 1993. Hepatocyte ultrastructure
rinated biphenyl (PCB) congener on hepatic glutathione, glutathione-related following exposure to aroclors and pure polychlorinated biphenyls. J. Environ.
enzymes, and selenium status: implications for oxidative stress. Biochem. Phar- Pathol. Toxicol. Oncol. 12, 17–33.
macol. 62, 273–281. Wilson, K.S., Prochaska, L.J., 1990. Phospholipid vesicles containing bovine heart
Uhlig, S., Wendel, A., 1992. The physiological consequences of glutathione variations. mitochondrial cytochrome c oxidase and subunit III-deficient enzyme: anal-
Life Sci. 51, 1083–1094. ysis of respiratory control and proton translocating activities. Arch. Biochem.
Van den Berg, M., Birnbaum, L., Bosveld, A.T., Brunstrom, B., Cook, P., Feeley, M., Biophys. 282, 413–420.
Giesy, J.P., Hanberg, A., Hasegawa, R., Kennedy, S.W., et al., 1998. Toxic equivalency Ye, S.F., Hou, Z.Q., Zhang, Q.Q., 2007. Protective effects of Phellinus linteus extract
factors (TEFs) for PCBs, PCDDs, PCDFs for humans and wildlife. Environ. Health against iron overload-mediated oxidative stress in cultured rat hepatocytes.
Perspect. 106, 775–792. Phytother. Res. 21, 948–953.
Van den Berg, M., Birnbaum, L.S., Denison, M., De Vito, M., Farland, W., Feeley, M., Zhao, H.X., Adamcakova-Dodd, A., Hu, D., Hornbuckle, K.C., Just, C.L., Robertson, L.W.,
Fiedler, H., Hakansson, H., Hanberg, A., Haws, L., Rose, M., Safe, S., Schrenk, D., Thorne, P.S., Lehmler, H.J., 2009. Development of a synthetic PCB mixture resem-
Tohyama, C., Tritscher, A., Tuomisto, J., Tysklind, M., Walker, N., Peterso, R.E., bling the average polychlorinated biphenyl profile in Chicago air. Environ. Int.,
2006. The 2005 world health organization reevaluation of human and mam- Apr 16.
malian toxic equivalency factors for dioxins and dioxin-like compounds. Toxicol. Zhou, C., Zhang, C., 2005. Protective effects of antioxidant vitamins on Aroclor 1254-
Sci. 93, 223–241. induced toxicity in cultured chicken embryo hepatocytes. Toxicol. In Vitro 19,
Vanden Hoek, T.L., Li, C., Shao, Z., Schumacker, P.T., Becker, L.B., 1997. Significant levels 665–673.
of oxidants are generated by isolated cardiomyocytes during ischemia prior to
reperfusion. J. Mol. Cell. Cardiol. 29, 2571–2583.

Você também pode gostar