Abstract

Neurotoxicity of individual metals is well investigated but that of metal mixture (MM), an environmental reality, in the developing brain is relatively obscure. We investigated the combinatorial effect of arsenic (As), cadmium (Cd), and lead (Pb) on rat brain development, spanning in utero to postnatal development. MM was administered by gavage to pregnant and lactating rats, and to postweaning pups till 2 months. The pups exhibited behavioral disturbances characterized by hyperlocomotion, increased grip strength, and learning-memory deficit. Disruption of the blood-brain barrier (BBB) was associated with dose-dependent increase in deposition of the metals in developing brain. Astrocytes were affected by MM treatment as evident from their reduced density, area, perimeter, compactness, and number of processes, and increased apoptosis in cerebral cortex and cerebellum. The metals induced synergistic reduction in glial fibrillary acidic protein (GFAP) expression during brain development; however, postweaning withdrawal of MM partially restored the levels of GFAP in adults. To characterize the toxic mechanism, we treated rat primary astrocytes with MM at concentrations ranging from lethal concentration (LC)10 to LC75 of the metals. We observed synergistic downregulation in viability and increase in apoptosis of the astrocytes, which were induced by proximal activation of extra cellular signal-regulated kinase (ERK) signaling and downstream activation of Jun N-terminal kinase (JNK) pathway. Furthermore, rise in intracellular calcium ion ([Ca2+]i) and reactive oxygen species generation promoted apoptosis in the astrocytes. Taken together, these observations are the first to show that mixture of As, Cd, and Pb has the capacity to induce synergistic toxicity in astrocytes that may compromise the BBB and may cause behavioral dysfunction in developing rats.

The developing brain is vulnerable to injury from toxic metals that interfere with the critical developmental processes, i.e., cellular proliferation, migration, differentiation, synaptogenesis, myelination, and apoptosis in the central nervous system (CNS) (Rice and Barone, 2000). The limited capacity of the developing CNS to compensate for the cell loss and the disruptions in neural networking results in compromised neuronal functions (Bayer, 1989) and increased risk of neurodegeneration (Grandjean and Landrigan, 2006).

Heavy metals including arsenic (As), cadmium (Cd), and lead (Pb) have received attention as both environmental contaminants and potential neurotoxicological hazards (Brender et al., 2006; Fowler et al., 2004; Jadhav et al., 2007). Exposure to the metals in utero and in infancy is associated with risk of impaired cognitive development (Hu, 2000; Landrigan et al., 1975), subclinical brain dysfunction (Lanphear et al., 2005), and behavioral anomalies (Tsai et al., 2003; Wright et al., 2006). Studies with single metal exposure have demonstrated that As, Cd, or Pb infiltrate the immature blood-brain barrier (BBB) and accumulate in developing brain (Lidsky and Schneider, 2003; Valkonen et al., 1983; Wang et al., 2007a; Xi et al., 2010). Pb uptake through the BBB disrupts Ca2+ transport mechanism (Marchetti, 2003) and promotes activation of mitogen-activated protein (MAP) kinases in apoptotic glial cells (Posser et al., 2007). The sequestration of Pb at the level of the choroid plexus undermines brain growth and affects learning and cognitive functions of CNS (Marchetti, 2003). In vivo and in vitro studies have revealed that acute or chronic Cd exposure enhances oxidative stress in astrocytes and accumulates reactive oxygen species (ROS) that induces astrocytic death (Yang et al., 2008). Perinatal exposure to Cd induces anxiety (Minetti and Reale, 2006) and reduces learning ability of offspring (Ishitobi et al., 2007). Chronic exposure to As, even at a submicromolar concentration, promotes oxidative stress (Garcia-Chavez et al., 2006) and induces neuroglial damage in human brain (Jin et al., 2004). Intoxication with As presents deficits in spontaneous locomotor activity (SLA) and alterations in learning-memory task during postnatal development (Rodriguez et al., 2002).

Current knowledge of metal-induced neurocellular damage, however, is primarily confined to single metal exposure, and there has been increasing demand for cumulative hazard assessment of metals in mixture in the brain (Rodriguez et al., 1998; Wright and Baccarelli, 2007). The effect may be either dose additive, interactive (synergistic or antagonistic), or independent of each other. Of the very few reports on metal mixture (MM), studies with early-life low doses of Pb + Cd–associated exposure have revealed greater oxidative stress than with single Pb or Cd (Zhang et al., 2009). The mixture altered cerebellar and striatal functions that related to changes in motor activity (Antonio et al., 2002) and anxiety (Leret et al., 2003) in the long term. The synergistic toxic effect of Pb and Cd absorption in the brain cells of cerebellum, cortex, and hippocampus is accountable for the enhanced CNS damage in the mixture (Gu et al., 2009). Therefore, we hypothesized that concurrent exposure to As along with Cd and Pb in drinking water may have greater-than-additive/synergistic toxic responses to brain development.

Here, we demonstrated the toxic effect of As, Cd, and Pb on rat brain development. We investigated the behavioral impairments induced by the MM in developing rats. Because abnormal blood-brain communication is a key mechanism underlying neuronal dysfunction (Shalev et al., 2009), we examined the effect of the MM on BBB integrity and glial fibrillary acidic protein (GFAP) levels. We further focused on the toxic mechanism of action (additive, synergistic, or antagonistic) of the MM on the GFAP-expressing rat astrocytes. Collectively, our data strongly suggested that single metal risk assessment underestimates the toxic impact of the metals present in mixture and sheds new light on the harmful role of environmental metal contaminants in pediatric and long-term CNS complications.

MATERIALS AND METHODS

Reagents and Antibodies

Na-arsenite, Pb-acetate, Cd-chloride, Na-orthovanadate, NaF, Ponceau S stain, Evans blue (EB), PMSF, protease inhibitor cocktail, MTT [3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide], 1,2-Bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid tetrakis(acetoxymethyl ester) (BAPTA-AM), α-tocopherol, Hoechst 33258 stain, poly-L-lysine, dichlorofluorescein diacetate (DCF-DA), Fluo3 AM, and mammalian tissue protein extraction reagent were procured from Sigma Chemical Co. (St Louis, MO). Fluo3 AM was obtained from Molecular Probes (Carlsbad, CA). The Alexa Fluor secondary antibodies, PD98059, SB203580, LY294002, and SP600125, cell culture reagents, sample loading buffer for Western blotting, and protein markers were purchased from Invitrogen (Carlsbad, CA). The supersignal west femto maximum sensitivity substrate for Western blotting was purchased from PIERCE Biotechnology (Rockford, IL). Rabbit polyclonal antibodies to extracellular signal-regulated kinase (ERK1/2) and Jun N-terminal kinase (JNK1/2), and phospho-ERK1/2 and phospho-JNK1/2 forms were purchased from Cell Signaling Technology (Danvers, MA). Mouse monoclonal antibody to GFAP was obtained from Millipore (Temecula, CA). Mouse monoclonal antibodies to β-actin, peroxidise-conjugated secondary antibodies were from Sigma Chemical Co.. The terminal deoxynucleotidyl-transferase (TdT)-mediated dUTP nick end labeling (TUNEL) kit was purchased from Roche (Indianapolis, IN). Diaminobenzidine tetrahydrochloride (DAB) substrate kit and Vectashield medium Elite ABC kit were purchased from Vector Laboratories (Burlingame, CA). Omniscript RT Kit and SYBR-green qPCR Kit were from Qiagen (Valencia, CA).

Animals and Treatments

All animal-handling procedures were carried out following the regulations of Institutional Animal Ethics Committee and with their prior approval for using the animals. The pregnant female Wistar rats were procured from Indian Institute of Toxicology Research and were housed in a 12-h day and light cycle environment with ad libitum availability of diet and water.

The pregnant female rats were divided into nine groups and treated with the metals (Table 1) through gavage from gestation day 5. The treatment was continued in the lactating rats and postweaning pups till 2 months. The MM was treated at two different concentrations (1× and 10×) to observe the dose-dependent effect of the MM on rat brain development, 1× being the most frequently occurring concentration of the metals in groundwater sources of India (Jadhav et al., 2007). The single metals were treated at two different concentrations. One was the same as that in 10× and the other was three times of it.

TABLE 1

Metal Treatment Given to Pregnant, Lactating, and Postweaning Rats

Group 1: VehicleWater (vehicle)
Group 2: MM (1×)Pb(C2H3O2)2: 0.220 ppm + CdCl2: 0.098 ppm + NaAsO2: 0.380 ppm
Group 3: MM (10×)Pb(C2H3O2)2: 2.220 ppm + CdCl2: 0.98 ppm + NaAsO2: 3.80 ppm
Group 4: Pb individual treatmentPb(C2H3O2)2: 2.220 ppm
Group 5: Cd individual treatmentCdCl2: 0.98 ppm
Group 6: As individual treatmentNaAsO2: 3.80 ppm
Group 7: Pb individual treatment (three times group 4)Pb(C2H3O2)2: 6.660 ppm
Group 8: Cd individual treatment (three times group 5)CdCl2: 2.94 ppm
Group 9: As individual treatment (three times group 6)NaAsO2: 11.4 ppm
Group 1: VehicleWater (vehicle)
Group 2: MM (1×)Pb(C2H3O2)2: 0.220 ppm + CdCl2: 0.098 ppm + NaAsO2: 0.380 ppm
Group 3: MM (10×)Pb(C2H3O2)2: 2.220 ppm + CdCl2: 0.98 ppm + NaAsO2: 3.80 ppm
Group 4: Pb individual treatmentPb(C2H3O2)2: 2.220 ppm
Group 5: Cd individual treatmentCdCl2: 0.98 ppm
Group 6: As individual treatmentNaAsO2: 3.80 ppm
Group 7: Pb individual treatment (three times group 4)Pb(C2H3O2)2: 6.660 ppm
Group 8: Cd individual treatment (three times group 5)CdCl2: 2.94 ppm
Group 9: As individual treatment (three times group 6)NaAsO2: 11.4 ppm
TABLE 1

Metal Treatment Given to Pregnant, Lactating, and Postweaning Rats

Group 1: VehicleWater (vehicle)
Group 2: MM (1×)Pb(C2H3O2)2: 0.220 ppm + CdCl2: 0.098 ppm + NaAsO2: 0.380 ppm
Group 3: MM (10×)Pb(C2H3O2)2: 2.220 ppm + CdCl2: 0.98 ppm + NaAsO2: 3.80 ppm
Group 4: Pb individual treatmentPb(C2H3O2)2: 2.220 ppm
Group 5: Cd individual treatmentCdCl2: 0.98 ppm
Group 6: As individual treatmentNaAsO2: 3.80 ppm
Group 7: Pb individual treatment (three times group 4)Pb(C2H3O2)2: 6.660 ppm
Group 8: Cd individual treatment (three times group 5)CdCl2: 2.94 ppm
Group 9: As individual treatment (three times group 6)NaAsO2: 11.4 ppm
Group 1: VehicleWater (vehicle)
Group 2: MM (1×)Pb(C2H3O2)2: 0.220 ppm + CdCl2: 0.098 ppm + NaAsO2: 0.380 ppm
Group 3: MM (10×)Pb(C2H3O2)2: 2.220 ppm + CdCl2: 0.98 ppm + NaAsO2: 3.80 ppm
Group 4: Pb individual treatmentPb(C2H3O2)2: 2.220 ppm
Group 5: Cd individual treatmentCdCl2: 0.98 ppm
Group 6: As individual treatmentNaAsO2: 3.80 ppm
Group 7: Pb individual treatment (three times group 4)Pb(C2H3O2)2: 6.660 ppm
Group 8: Cd individual treatment (three times group 5)CdCl2: 2.94 ppm
Group 9: As individual treatment (three times group 6)NaAsO2: 11.4 ppm

The number of pregnant rats per treatment group was 30. After standardization of litters (culling), equal numbers of male and female pups were taken for each experiment, and pups from different litters were independent subjects (Holson et al., 2008).

To determine the toxic effect of metals on glial damage and behavioral aberrations during rat brain development, studies were carried out at postnatal day (P) 16 and adult P60 rats.

Behavioral Study

Spontaneous locomotor activity.

Locomotive behavior in rats was studied using the computerized Opto-Varimex (Columbus Instruments, Columbus, OH) system as previously described (Ali et al., 1990). The locomotor markers monitored were the distance traveled, number of stereotypic movements and rearings, and time moving (in minutes).

Grip strength.

Vehicle and MM-treated rats were subjected to forelimb grip strength test using a digital grip strength meter (Columbus Instruments) following the standard procedure as described previously (Terry et al., 2003). Each rat was tested five times, with a 10-s period between two successive trials. The five recordings were averaged to obtain a final reading for each individual.

Y-maze.

Learning-memory test for vehicle- and MM-treated rats was carried out as described previously using Y-maze (Wetzel and Matthies, 1982) with minor modifications. The training apparatus was a Y-maze with electrifiable grid-floored three alleys and 15-W light bulb at the end of the alleys. When the rats were being tested, only one arm with its light on (bright arm) was a safe area without footshock, whereas the other two arms with the lights off (dark arms) were unsafe areas with electric footshock (1–5 mA). The safe and unsafe areas were arbitrarily shifted during testing. The training session consisted of 30 trials per animal. Running into the dark alley of the Y-maze was counted as an error (E). Retention of the brightness discrimination was tested after 24 h and after 7 days of initial learning using a 30-trial relearning session; performance in relearning was considered as test for memory and expressed as percent (%) saving (Serota, 1971).

Analysis of EB Extravasation in Brain

Rats were anesthetized and 3% EB in saline was injected slowly through the femoral vein. The rats were then transcardially perfused and the brain isolated and incubated in formamide as previously described (Lin et al., 2010). The supernatant was collected, and the optical density (OD) at 620 nm was measured using a SPECTRA max PLUS384 spectrophotometer (Molecular devices, Sunnyvale, CA) to determine the relative amount of EB in each sample. The OD ratio was derived from the OD of the MM-treated animals over that of the vehicle-treated animals.

Fluorescence study of EB extravasation was carried out as described previously (Duran-Vilaregut et al., 2009). Briefly, EB-injected rats were perfused; cerebral cortex and cerebellum dissected, postfixed in 4% paraformaldehyde (PFA), and cryoprotected with 30% sucrose; and 20-μm-thick cryostatic sections obtained using cryomicrotome (Microm HM 520; Labcon, Germany). The cryostat sections were visualized under fluorescence microscope (Nikon Instech Co. Ltd, Kawasaki, Kanagawa, Japan) after being coverslipped on Vectashield medium (Vector Laboratories).

Evaluation of Metals in Brain

For determining the levels of As, Cd, and Pb in brain, whole-brain samples were snap frozen in liquid nitrogen and kept in a −80°C freezer till analysis. Samples for Pb, Cd, and As estimation were prepared by acid digestion as previously described (Ong et al., 2006; Singh and Rana, 2007), and analysis was carried out using atomic absorption spectrophotometer equipped with a vapor generation assembly (Varian AAS 250+ coupled with VGA 77; Varian Australia Pvt Ltd [manufacturing site], Mulgave, Australia). Detection limit of the instrument was 1 ppb (Behari and Prakash, 2006).

Immunohistochemistry and Quantitative Estimation of GFAP-Immunoreactive Astrocytes

Immunoperoxidase staining with 5-μm cryostat sections of cerebellum (transverse) and cerebral cortex (coronal) was carried out for GFAP antibody. Briefly, four pups from four different litters were taken at the developmental stages, anesthetized, and perfused and the brain was fixed and cryoprotected as described previously (Sinha et al., 2009; Zhu et al., 2001). Five-micron transverse sections were made from the cerebellum and coronal sections of the cerebral cortex using cryomicrotome (Microm HM 520; Labcon). The sections were then mounted on 3-aminopropyltriethoxysilane-coated slides. Immunoperoxidase staining for GFAP (monoclonal, 1:400) was carried out with DAB chromogen and ABC kit as described previously (Otani et al., 1999) and visualized under optical microscope (Nikon Instech Co. Ltd). Ten fields in each section were captured using ×40 objective for evaluation. Image-Pro Plus 5.1 software (Media Cybernetics Inc., Silver Spring, MD) was used for image capturing.

The images were then imported into Image-J 1.42q (http://rsb.info.nih.gov/ij/; developed by Wayne Rasband, National Institutes of Health, Bethesda, MD). The number of GFAP-immunoreactive (ir) cells and the area, perimeter, compactness, and number of processes in the GFAP-ir cells were quantified with the Shape descriptors plugin of the software.

Detection of Apoptosis (in vivo) in GFAP-ir Astrocytes

In situ detection of apoptosis was carried out by TUNEL assay in P16 and P60 rats. Briefly, four pups from four different litters were taken at the developmental stages, anesthetized, and perfused and the brain was fixed and cryoprotected as described previously (Sinha et al., 2009; Zhu et al., 2001). Five-micron transverse sections from the cerebellum and coronal sections from the cerebral cortex were made using cryomicrotome (Microm HM 520; Labcon). For the TUNEL assay, a labeling reaction was carried out with fluorescein-labeled dUTP in the presence of TdT at 37°C for 1 h. To investigate whether the apoptotic cells were astrocytes, the sections were immunostained with anti-mouse GFAP antibody (1:400 dilution in TBST [10mM Tris, pH 8.0, 150mM NaCl, 0.01% Tween 20]) according to manufacturer’s protocol. The sections were then incubated with Alexa Fluor 546–conjugated (fluorescent color: red; Abs/Em: 555/565) goat anti-mouse antibody (1:200 dilution) as previously described (Bandyopadhyay et al., 2007); counterstained with Hoechst 33258 (0.2mM) for 10 min; and visualized under a fluorescence microscope (Nikon Instech Co. Ltd) after being coverslipped on Vectashield medium (Vector Laboratories).

Protein Extraction and Western Blotting

Cerebral cortex and cerebellar tissues from five to seven postnatal rats were harvested, snap frozen in liquid nitrogen, and stored at −80° C until further investigation. SDS polyacrylamide gel electrophoresis and Western blotting were done with the tissues following an optimized protocol (Sinha et al., 2009) with GFAP and β-actin antibodies. To determine the phosphorylation of JNK1/2 and ERK1/2, previously described protocol was followed (Bandyopadhyay et al., 2006). The working dilutions for the primary antibodies were as follows: GFAP (monoclonal, 1:1000), β-actin (monoclonal, 1:1000), JNK1/2 (polyclonal, 1:1000), ERK1/2 (polyclonal, 1:1000), phospho-JNK1/2 (polyclonal, 1:1000), and phospho-ERK1/2 (polyclonal, 1:1000). The working dilutions for secondary anti-rabbit IgG (for the polyclonal primary antibodies) and anti-mouse IgG (for the monoclonal primary antibodies) conjugated to horseradish peroxidase were 1:1000 in PBS plus 0.2% Triton X-100. The samples were detected by chemiluminescence with the supersignal west femto maximum sensitivity substrate. Relative expression of each protein was determined by densitometric quantification of blots using the VersaDoc Gel Imaging System (BioRad, Hercules, CA).

Cell Culture

Astrocytes.

Rat pups (P1) were decapitated, the brain was removed, and astrocytes were cultured from the rat brain as described previously (Tanaka et al., 1998). The purity of astrocytes was 94–97% as determined by the immunofluorescence staining of GFAP.

Treatment of Cells with Metals and Cell Signaling Inhibitors

The astrocytes were grown to 80% confluence, pre-incubated in reduced serum (0.5% fetal bovine serum [FBS]) medium for 2 h, and then treated with As, Cd, Pb, or MM at concentrations ranging from 0.01 to 200μM for 18 h in a humidified tissue culture incubator at 37°C with 5% CO2-95% air.

The involvement of signaling pathways, such as PI3 kinase, MEK1/2, P38-MAPK, and JNK1/2, in MM-induced astrocyte toxicity was determined by incubating the astrocytes with the MM and cotreating with LY294002 (10μM), PD98059 (10μM), SB203580 (10μM), or SP600125 (10μM), which are inhibitors to PI3 kinase, MEK1/2, P38-MAPK, and JNK1/2 pathways, respectively.

The time course of phosphorylation of MEK1/2 (assessed by ERK1/2 phosphorylation) and JNK1/2 in MM-treated astrocytes was determined by incubating the cells with MM for 0, 5, 10, 15, 30, and 60 min.

Cytotoxicity Assay

MTT cell viability assay, a colorimetric assay, was used to determine astrocyte viability. The astrocytes were grown to 80% confluence, pre-incubated in reduced serum (0.5% FBS) medium for 2 h, and then treated with As, Cd, Pb, or MM at concentrations ranging from 0.01 to 200μM for 18 h in reduced serum medium in a humidified tissue culture incubator at 37°C with 5% CO2-95% air. The cells were then incubated with 10 μl MTT (10.4 mg/ml) and an optimized protocol was followed to determine cell viability (Sanders et al., 2000). Absorbance was measured at 595 nm, with background subtraction at 655 nm. The LC of the metals on astrocytes was analyzed by the GraphPad Prism 3.0 software.

TUNEL Assay

The in situ detection of apoptosis in astrocytes was carried out with the In Situ Cell Death Detection fluorescein kit (Roche Applied Science) for TUNEL assay, as per the manufacturer’s instruction and visualized under a fluorescence microscope (Nikon Instech Co. Ltd). Briefly, the astrocytes were grown to 80% confluence, pre-incubated in reduced serum medium for 2 h, and then treated with As, Cd, Pb, or MM in reduced serum medium at a concentration range of LC10 to LC100 of the metals for 18 h in a humidified tissue culture incubator at 37°C with 5% CO2-95% air. The cells were counterstained with Hoechst 33258. The TUNEL-positive cells were counted in five randomly selected fields. Around 10,000 cells in each cover slip were scored. The apoptotic index was expressed as the number of TUNEL-positive cells per 100 nuclei (Hoechst stained). Image-Pro Plus 5.1 software (Media Cybernetics Inc.) was used for cell counting.

Assay for [Ca2+]i

The astrocytes were grown to 80% confluence at 4000 cells/well in 96-well poly D-lysine-coated black view plates (VWR no. 62406–036; Falcon), pre-incubated in reduced serum medium for 2 h, and then treated with MM in reduced serum medium for 2 min, 5 min, 10 min, 30 min, 1 h, 2 h, and 24 h in a humidified tissue culture incubator at 37°C with 5% CO2-95% air. An increase in [Ca2+]i was measured using Fluo3 AM as an indicator dye after the addition of metals (single or in mixture) to the culture wells following an optimized protocol (Arey et al., 2005). The fluorescent signals were read by fluorescence imaging plate reader Synergy HT (BioTek, Winooski, VT).

Assay for ROS Generation

The production of ROS, mainly H2O2, was assessed using DCF-DA, followed by semiquantitative fluorometric measurements as described previously (Izawa et al., 2009). The astrocytes were grown to 80% confluence at 4000 cells/well in 96-well poly D-lysine-coated black view plates (VWR no. 62406–036; Falcon); pre-incubated in reduced serum medium for 2 h; treated with MM for 0 min, 30 min, 60 min, 2 h, or overnight; and then incubated with DCF-DA to a working concentration of 20 μg/ml. The fluorescent level indicating intracellular ROS production was measured using a fluorescent microplate reader with excitation at 485 nm and emission at 530 nm.

Evaluation of Metal(s) Interaction in Combination

To characterize the interaction between the heavy metals for their effects on astrocyte viability, apoptotic index and GFAP level, a combination index (CI) was calculated using the software Calcusyn (Biosoft, Manchester, United Kingdom). CI values less than 1.0 indicated synergism (Zhao et al., 2004), which suggests greater than expected based on effect addition.

The U.S. Environmental Protection Agency has selected dose addition as the no-interaction definition for mixture risk assessment, so that synergism would represent toxic effects that exceed those predicted from dose addition (Hertzberg and MacDonell, 2002). Synergistic interactions have implications of gain (Berenbaum, 1985) and the combined effects are greater than the additive effect of the components (Wessinger, 1986).

Statistical Analysis

Data are presented as mean ± SE of the indicated number of experiments. Statistical analysis was carried out in SPSS 9.0 software (SPSS Inc., Chicago, IL). Data were analyzed by one-way ANOVA, followed by Student-Newman, Keuls post hoc test or Student’s t-test when appropriate.

RESULTS

Effect of MM on SLA, Grip Strength, and Learning-Memory Performance in Developing Rats

It has been previously reported that exposure to inorganic As (Rodriguez et al., 2002), Cd (Ishitobi et al., 2007), or Pb (Marchetti, 2003) caused behavioral alterations in developing rats. We, therefore, investigated whether MM treatment altered SLA, grip strength, and learning-memory performance in developing rats. The rats exhibited dose-dependent increase in the distance traveled, number of stereotypic movements, number of rearings, movement time, and grip strength (Table 2). MM treatment also demonstrated dose-dependent reduction in learning-memory performance (Table 2).

TABLE 2

MM Enhanced SLA and Grip Strength, and Reduced Learning-Memory Performance in Developing Rats

Total distance (cm/5 min)Rearings in 5 minTime moving (min) in 5 minStereotypic movements (in 5 min)Grip strength (newton)No. of errors in training phaseLearning-memory (24 h) (% saving)Learning-memory (7th day) (% saving)
Vehicle14 ± 0.86 ± 11.7 ± 0.037 ± 1154 ± 6.011.67 ± 0.674433.75
17 ± 1.3*24 ± 2*2.1 ± 0.03*9 ± 1*173 ± 2.0*13.33 ± 0.33*37.82*4.97**
10×21 ± 2.2*50 ± 4**2.8 ± .0.03**16 ± 2**200 ± 8.0**14.66 ± 0.33*34.32**0.23***
Total distance (cm/5 min)Rearings in 5 minTime moving (min) in 5 minStereotypic movements (in 5 min)Grip strength (newton)No. of errors in training phaseLearning-memory (24 h) (% saving)Learning-memory (7th day) (% saving)
Vehicle14 ± 0.86 ± 11.7 ± 0.037 ± 1154 ± 6.011.67 ± 0.674433.75
17 ± 1.3*24 ± 2*2.1 ± 0.03*9 ± 1*173 ± 2.0*13.33 ± 0.33*37.82*4.97**
10×21 ± 2.2*50 ± 4**2.8 ± .0.03**16 ± 2**200 ± 8.0**14.66 ± 0.33*34.32**0.23***

Note. The MM-treated P16 rats were assessed for SLA and grip strength. The MM-treated P60 rats were assessed for learning-memory performance (refer to the “Materials and Methods” section). Data are expressed as means of ± SE of eight postnatal rats from different litters.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with vehicle).

TABLE 2

MM Enhanced SLA and Grip Strength, and Reduced Learning-Memory Performance in Developing Rats

Total distance (cm/5 min)Rearings in 5 minTime moving (min) in 5 minStereotypic movements (in 5 min)Grip strength (newton)No. of errors in training phaseLearning-memory (24 h) (% saving)Learning-memory (7th day) (% saving)
Vehicle14 ± 0.86 ± 11.7 ± 0.037 ± 1154 ± 6.011.67 ± 0.674433.75
17 ± 1.3*24 ± 2*2.1 ± 0.03*9 ± 1*173 ± 2.0*13.33 ± 0.33*37.82*4.97**
10×21 ± 2.2*50 ± 4**2.8 ± .0.03**16 ± 2**200 ± 8.0**14.66 ± 0.33*34.32**0.23***
Total distance (cm/5 min)Rearings in 5 minTime moving (min) in 5 minStereotypic movements (in 5 min)Grip strength (newton)No. of errors in training phaseLearning-memory (24 h) (% saving)Learning-memory (7th day) (% saving)
Vehicle14 ± 0.86 ± 11.7 ± 0.037 ± 1154 ± 6.011.67 ± 0.674433.75
17 ± 1.3*24 ± 2*2.1 ± 0.03*9 ± 1*173 ± 2.0*13.33 ± 0.33*37.82*4.97**
10×21 ± 2.2*50 ± 4**2.8 ± .0.03**16 ± 2**200 ± 8.0**14.66 ± 0.33*34.32**0.23***

Note. The MM-treated P16 rats were assessed for SLA and grip strength. The MM-treated P60 rats were assessed for learning-memory performance (refer to the “Materials and Methods” section). Data are expressed as means of ± SE of eight postnatal rats from different litters.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with vehicle).

Effect of MM on BBB Permeability in Developing Rat Brain

Increase in BBB permeability is a potentially important cause of brain dysfunction including behavioral disorders (Shalev et al., 2009). Moreover, it has been reported that As and Cd disrupted the integrity of BBB (Zheng, 2001) and exposure to Pb disrupted the BBB (Lidsky and Schneider, 2003). We, therefore, determined the effect of MM on BBB permeability by measuring EB extravasation. We observed dose-dependent increase in EB extravasation in MM-treated postnatal rat brain (Figs. 1A and 1B). Transverse section of cerebellum (Figs. 1C and 1D) and coronal section of cerebral cortex (Fig. 1E) exhibited marked BBB damage, as evident from the increased red fluorescence (indicative of EB leakage) in adult rats.

MM promoted BBB permeability in developing rat brain. EB (3%) was injected through the femoral vein of the vehicle (V)– and MM (1× and 10×)–treated (A) P16 and (B) P60 rats. The whole-brain samples were incubated in formamide, and the OD of supernatant was taken at 620 nm. Average of OD ratios of the MM-treated over vehicle-treated rats was determined. Data represent means ± SE of four pups from four different litters. **p < 0.01 and ***p < 0.001 (compared with V).
FIG. 1.

MM promoted BBB permeability in developing rat brain. EB (3%) was injected through the femoral vein of the vehicle (V)– and MM (1× and 10×)–treated (A) P16 and (B) P60 rats. The whole-brain samples were incubated in formamide, and the OD of supernatant was taken at 620 nm. Average of OD ratios of the MM-treated over vehicle-treated rats was determined. Data represent means ± SE of four pups from four different litters. **p < 0.01 and ***p < 0.001 (compared with V).

EB was injected in vehicle (V)– and MM-treated (10×)-P60 rats, and 20-μm cryostat sections of cerebral cortex (coronal) and cerebellum (transverse) were observed under fluorescence microscope. (C) Representative photomicrograph (×10 magnification) showing EB fluorescence in the posterior lobe of rat cerebellum. (D) Representative photomicrograph (×10 magnification) showing EB fluorescence in white matter of rat cerebellum. (E) Representative photomicrograph (×10 magnification) of EB fluorescence in the corpus callosum. The sections are representatives of four P60 rats from four different litters.

Levels of As, Cd, or Pb in Metal-Treated Developing Rat Brain

Previous studies have demonstrated that exposure to inorganic As (Garcia-Chavez et al., 2006), Cd (Lafuente et al., 2001), or Pb (Benitez et al., 2009; Rader et al., 1981) led to deposition of the metals in rat brain. Because the metals could reach the fetus through placenta (Benitez et al., 2009; Xi et al., 2010), and lactating off-springs through mother’s milk (Counter et al., 2007; Kippler et al., 2009; Samanta et al., 2007), we investigated whether metal(s) treatment (groups 1–9) caused their deposition in developing brain. A dose-dependent increase in the levels of As, Cd, and Pb was observed in the postnatal rat brain (Table 3).

TABLE 3

Levels of As, Cs, and Pb in MM-Treated Developing Rat Brain

P16
P60
As (ppb)Cd (ppb)Pb (ppb)As (ppb)Cd (ppb)Pb (ppb)
Group 1: vehiclebdlbdlbdlbdlbdlbdl
Group 2: MM (1×)14.25 ± 3.10*17.25 ± 4.17*60.72 ± 6.01**33.7 ± 5.05*52.68 ± 6.88**87 ± 5.23**
Group 3: MM (10×)63.72 ± 7.35**52.56 ± 5.38**135.11 ± 11.20***87.23 ± 9.28**106.70 ± 9.17***128.62 ± 12.67***
Group 4: Pb individual treatmentbdlbdl143.18 ± 13.34***bdlbdl150.82 ± 13.01***
Group 5: Cd individual treatmentbdl53.67 ± 4.82**bdlbdl95 ± 8.63**bdl
Group 6: As individual treatment64.56 ± 9.16**bdlbdl90.17 ± 7.52**bdlbdl
Group 7: Pb individual treatment (three times group 4)bdlbdl288.22 ± 15.65***bdlbdl270.57 ± 21.01***
Group 8: Cd individual treatment (three times group 5)bdl139 ± 13.01***bdlbdl201 ± 29.768***bdl
Group 9: As individual treatment (three times group 6)108.32 ± 10.42***bdlbdl120.63 ± 12.39***bdlbdl
P16
P60
As (ppb)Cd (ppb)Pb (ppb)As (ppb)Cd (ppb)Pb (ppb)
Group 1: vehiclebdlbdlbdlbdlbdlbdl
Group 2: MM (1×)14.25 ± 3.10*17.25 ± 4.17*60.72 ± 6.01**33.7 ± 5.05*52.68 ± 6.88**87 ± 5.23**
Group 3: MM (10×)63.72 ± 7.35**52.56 ± 5.38**135.11 ± 11.20***87.23 ± 9.28**106.70 ± 9.17***128.62 ± 12.67***
Group 4: Pb individual treatmentbdlbdl143.18 ± 13.34***bdlbdl150.82 ± 13.01***
Group 5: Cd individual treatmentbdl53.67 ± 4.82**bdlbdl95 ± 8.63**bdl
Group 6: As individual treatment64.56 ± 9.16**bdlbdl90.17 ± 7.52**bdlbdl
Group 7: Pb individual treatment (three times group 4)bdlbdl288.22 ± 15.65***bdlbdl270.57 ± 21.01***
Group 8: Cd individual treatment (three times group 5)bdl139 ± 13.01***bdlbdl201 ± 29.768***bdl
Group 9: As individual treatment (three times group 6)108.32 ± 10.42***bdlbdl120.63 ± 12.39***bdlbdl

Note. Whole-brain samples from P16 and P60 rats were snap frozen in liquid nitrogen, acid digested and the metal content determined by atomic absorption spectrophotometer. Data expressed as means ± SE of four pups from four different litters. bdl, below detection limit.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with vehicle).

TABLE 3

Levels of As, Cs, and Pb in MM-Treated Developing Rat Brain

P16
P60
As (ppb)Cd (ppb)Pb (ppb)As (ppb)Cd (ppb)Pb (ppb)
Group 1: vehiclebdlbdlbdlbdlbdlbdl
Group 2: MM (1×)14.25 ± 3.10*17.25 ± 4.17*60.72 ± 6.01**33.7 ± 5.05*52.68 ± 6.88**87 ± 5.23**
Group 3: MM (10×)63.72 ± 7.35**52.56 ± 5.38**135.11 ± 11.20***87.23 ± 9.28**106.70 ± 9.17***128.62 ± 12.67***
Group 4: Pb individual treatmentbdlbdl143.18 ± 13.34***bdlbdl150.82 ± 13.01***
Group 5: Cd individual treatmentbdl53.67 ± 4.82**bdlbdl95 ± 8.63**bdl
Group 6: As individual treatment64.56 ± 9.16**bdlbdl90.17 ± 7.52**bdlbdl
Group 7: Pb individual treatment (three times group 4)bdlbdl288.22 ± 15.65***bdlbdl270.57 ± 21.01***
Group 8: Cd individual treatment (three times group 5)bdl139 ± 13.01***bdlbdl201 ± 29.768***bdl
Group 9: As individual treatment (three times group 6)108.32 ± 10.42***bdlbdl120.63 ± 12.39***bdlbdl
P16
P60
As (ppb)Cd (ppb)Pb (ppb)As (ppb)Cd (ppb)Pb (ppb)
Group 1: vehiclebdlbdlbdlbdlbdlbdl
Group 2: MM (1×)14.25 ± 3.10*17.25 ± 4.17*60.72 ± 6.01**33.7 ± 5.05*52.68 ± 6.88**87 ± 5.23**
Group 3: MM (10×)63.72 ± 7.35**52.56 ± 5.38**135.11 ± 11.20***87.23 ± 9.28**106.70 ± 9.17***128.62 ± 12.67***
Group 4: Pb individual treatmentbdlbdl143.18 ± 13.34***bdlbdl150.82 ± 13.01***
Group 5: Cd individual treatmentbdl53.67 ± 4.82**bdlbdl95 ± 8.63**bdl
Group 6: As individual treatment64.56 ± 9.16**bdlbdl90.17 ± 7.52**bdlbdl
Group 7: Pb individual treatment (three times group 4)bdlbdl288.22 ± 15.65***bdlbdl270.57 ± 21.01***
Group 8: Cd individual treatment (three times group 5)bdl139 ± 13.01***bdlbdl201 ± 29.768***bdl
Group 9: As individual treatment (three times group 6)108.32 ± 10.42***bdlbdl120.63 ± 12.39***bdlbdl

Note. Whole-brain samples from P16 and P60 rats were snap frozen in liquid nitrogen, acid digested and the metal content determined by atomic absorption spectrophotometer. Data expressed as means ± SE of four pups from four different litters. bdl, below detection limit.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with vehicle).

Effect of MM on the Number, Area, Perimeter, Compactness, Processes, and Apoptosis of Astrocytes in Developing Rat Brain

Astrocytes are involved in the maintenance of BBB (Choi and Kim, 2008; Pekny et al., 1998), and loss of BBB integrity is associated with astrocyte damage under neuropathological conditions (Prior et al., 2004; Willis et al., 2004). Because we observed that MM promoted permeability of the developing rat BBB, we studied the response of the astrocytes to MM. We observed reduction in the GFAP-ir astrocyte count, area, perimeter, compactness, and number of processes in the cerebral cortex and cerebellum (Figs. 2A–D, Table 4). We further observed that the MM promoted apoptosis in the developing cerebral and cerebellar astrocytes (Figs. 2E–H).

TABLE 4

Size and Density of Astrocytes in Vehicle-Treated and MM-Treated P16 and P60 Rats

Astrocyte
Count/0.081 mm2Surface area (μm2)Perimeter (μm)CompactnessNo. of processes
Cerebral cortexV-P1635 ± 1.201762 ± 107.75149 ± 17.500.88 ± 0.028 ± 1.00
10×-P167.71 ± 1.25***722.42 ± 60.25***87.91 ± 12.20***0.722 ± 0.05**6.16 ± 0.23**
V-P6034 ± 0.752750.92 ± 15.7631184.91 ± 6.2310.90 ± 0.037.04 ± 0.80
10×-P6014.25 ± 0.75***1283.48 ± 91.561***126.83 ± 13.25**0.86 ± 0.05*5.10 ± 0.50**
CerebellumV-P1630 ± 0.801450.29 ± 52.75137.76 ± 12.500.86 ± 0.0426.97 ± 0.20
10×-P168.44 ± 0.23***768.5 ± 20.02**108.23 ± 8.92**0.84 ± 0.05*5.4366 ± 0.02**
V-P6029.75 ± 0.752058.2 ± 120.35160.18 ± 17.500.90 ± 0.0355.12 ± 0.30
10×-P6016.25 ± 0.25***1091.74 ± 52.6***115.33 ± 9.711**0.79 ± 0.025**4.20 ± 0.2**
Astrocyte
Count/0.081 mm2Surface area (μm2)Perimeter (μm)CompactnessNo. of processes
Cerebral cortexV-P1635 ± 1.201762 ± 107.75149 ± 17.500.88 ± 0.028 ± 1.00
10×-P167.71 ± 1.25***722.42 ± 60.25***87.91 ± 12.20***0.722 ± 0.05**6.16 ± 0.23**
V-P6034 ± 0.752750.92 ± 15.7631184.91 ± 6.2310.90 ± 0.037.04 ± 0.80
10×-P6014.25 ± 0.75***1283.48 ± 91.561***126.83 ± 13.25**0.86 ± 0.05*5.10 ± 0.50**
CerebellumV-P1630 ± 0.801450.29 ± 52.75137.76 ± 12.500.86 ± 0.0426.97 ± 0.20
10×-P168.44 ± 0.23***768.5 ± 20.02**108.23 ± 8.92**0.84 ± 0.05*5.4366 ± 0.02**
V-P6029.75 ± 0.752058.2 ± 120.35160.18 ± 17.500.90 ± 0.0355.12 ± 0.30
10×-P6016.25 ± 0.25***1091.74 ± 52.6***115.33 ± 9.711**0.79 ± 0.025**4.20 ± 0.2**

Note. Five-micron cryostat sections of cerebral cortex and cerebellum were made from vehicle (V)– and MM (10×)–treated P16 and P60 rats, immunostained with GFAP, and images obtained with optical microscopy. The images were then analyzed (refer to the “Materials and Methods” section). The number, surface area, perimeter, and compactness of astrocytes, and number of astrocyte processes in V- and MM-treated rats were determined. Data expressed as means ± SE of five pups from five different litters.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with V).

TABLE 4

Size and Density of Astrocytes in Vehicle-Treated and MM-Treated P16 and P60 Rats

Astrocyte
Count/0.081 mm2Surface area (μm2)Perimeter (μm)CompactnessNo. of processes
Cerebral cortexV-P1635 ± 1.201762 ± 107.75149 ± 17.500.88 ± 0.028 ± 1.00
10×-P167.71 ± 1.25***722.42 ± 60.25***87.91 ± 12.20***0.722 ± 0.05**6.16 ± 0.23**
V-P6034 ± 0.752750.92 ± 15.7631184.91 ± 6.2310.90 ± 0.037.04 ± 0.80
10×-P6014.25 ± 0.75***1283.48 ± 91.561***126.83 ± 13.25**0.86 ± 0.05*5.10 ± 0.50**
CerebellumV-P1630 ± 0.801450.29 ± 52.75137.76 ± 12.500.86 ± 0.0426.97 ± 0.20
10×-P168.44 ± 0.23***768.5 ± 20.02**108.23 ± 8.92**0.84 ± 0.05*5.4366 ± 0.02**
V-P6029.75 ± 0.752058.2 ± 120.35160.18 ± 17.500.90 ± 0.0355.12 ± 0.30
10×-P6016.25 ± 0.25***1091.74 ± 52.6***115.33 ± 9.711**0.79 ± 0.025**4.20 ± 0.2**
Astrocyte
Count/0.081 mm2Surface area (μm2)Perimeter (μm)CompactnessNo. of processes
Cerebral cortexV-P1635 ± 1.201762 ± 107.75149 ± 17.500.88 ± 0.028 ± 1.00
10×-P167.71 ± 1.25***722.42 ± 60.25***87.91 ± 12.20***0.722 ± 0.05**6.16 ± 0.23**
V-P6034 ± 0.752750.92 ± 15.7631184.91 ± 6.2310.90 ± 0.037.04 ± 0.80
10×-P6014.25 ± 0.75***1283.48 ± 91.561***126.83 ± 13.25**0.86 ± 0.05*5.10 ± 0.50**
CerebellumV-P1630 ± 0.801450.29 ± 52.75137.76 ± 12.500.86 ± 0.0426.97 ± 0.20
10×-P168.44 ± 0.23***768.5 ± 20.02**108.23 ± 8.92**0.84 ± 0.05*5.4366 ± 0.02**
V-P6029.75 ± 0.752058.2 ± 120.35160.18 ± 17.500.90 ± 0.0355.12 ± 0.30
10×-P6016.25 ± 0.25***1091.74 ± 52.6***115.33 ± 9.711**0.79 ± 0.025**4.20 ± 0.2**

Note. Five-micron cryostat sections of cerebral cortex and cerebellum were made from vehicle (V)– and MM (10×)–treated P16 and P60 rats, immunostained with GFAP, and images obtained with optical microscopy. The images were then analyzed (refer to the “Materials and Methods” section). The number, surface area, perimeter, and compactness of astrocytes, and number of astrocyte processes in V- and MM-treated rats were determined. Data expressed as means ± SE of five pups from five different litters.

*p < 0.05, **p < 0.01, and ***p < 0.001 (compared with V).

FIG. 2.

MM reduced size and number of GFAP-ir astrocytes in developing rat cortex and cerebellum. Five-micron-thick cryostat sections of cerebral cortex (coronal section) and cerebellum (transverse section) from vehicle (V)– and MM (10×)–treated P16 and P60 rats were stained for GFAP using peroxidase conjugate and DAB chromogen. (A) Representative photomicrograph (×40 magnification) of GFAP-ir astrocytes in P16 rat cerebral cortex. (B) Representative photomicrograph (×40 magnification) of GFAP-ir astrocytes in P16 rat cerebellum. (C) Representative photomicrograph (×40 magnification) of GFAP-ir astrocytes in P60 rat cerebral cortex. (D) Representative photomicrograph (×40 magnification) of rat GFAP-ir astrocytes in P60 rat cerebellum. The sections are representatives of four P16 or P60 rats from four different litters.

Effect of MM on GFAP Level in Developing Rat Brain

We studied the effect of MM on the GFAP levels in the developing cerebral cortex and cerebellum. We observed dose-dependent decrease in the level of GFAP that persisted till adulthood (Figs. 3A and 3B). Upon withdrawal of MM treatment from weaning, the GFAP level was significantly restored in adults (Figs. 3C and 3D). This suggests that the decrease in level of GFAP was primarily caused by MM treatment.

FIG. 3.

MM suppressed GFAP expression in developing rat brain. The cerebral cortex and cerebellum of vehicle (V)– and MM-treated (1× or 10×) P16 and P60 rats were immunoblotted for GFAP and β-actin. (A) Representative Western blot and densitometric analyses of relative GFAP expression normalized with β-actin in rat cerebral cortex at indicated postnatal days. Data represent means ± SE of four pups from four different litters. *p < 0.05, **p < 0.01, and ***p < 0.001 (compared with V). (B) Representative Western blot and densitometric analyses of relative GFAP expression normalized with β-actin in rat cerebellum at indicated postnatal days. Data represent means ± SE of four pups from four different litters. *p < 0.05, **p < 0.01, and ***p < 0.001 (compared with V). (C) Representative Western blot and densitometric analyses of relative GFAP expression normalized with β-actin in cerebral cortex of P60 rats after MM treatment or postweaning MM withdrawal (WD). Data represent means ± SE of four pups from four different litters. *p < 0.05 and ***p < 0.001 (compared with V). (D) Representative Western blot and densitometric analyses of relative GFAP expression normalized with β-actin in cerebellum of P60 rats after MM treatment or postweaning MM withdrawal (WD). Data represent means ± SE of five pups from five different litters. *p < 0.05 and ***p < 0.001 (compared with V).

We further investigated whether MM had synergistic or additive toxic effect on the level of GFAP in the developing rat brain. We determined the effect of the single metals (groups 4–9 in the “Materials and Methods” section) or of MM (10× concentration) on the GFAP level. We observed that the sum of the fold-reduction in GFAP level by single metals was less than the fold reduction in GFAP by MM (Figs. 3E–H), suggesting synergistic GFAP suppression by MM.

Effect of MM on Viability and Apoptosis of Rat Primary Astrocyte Culture

We investigated the toxic mechanism of action of MM in the rat primary astrocytes.

We first determined the effect of As, Cd, or Pb on the astrocyte viability. The individual metals induced cell death, and the lethal concentrations (LC10, LC25, LC50, and LC75) of the individual metals were determined (Table 5). We then treated the cells with MM and observed that MM induced greater cell death than the sum of cell death induced by individual metals (Table 6), suggesting that As, Cd, and Pb induced synergistic reduction in astrocyte viability.

TABLE 5

LC Values of As, Cd, and Pb in Rat Primary Astrocytes

LC10 (μM)LC25 (μM)LC50 (μM)LC75
As9.9924.9949.9574.93
Cd2.014.8010.0515.08
Pb1537.575.0112.5
LC10 (μM)LC25 (μM)LC50 (μM)LC75
As9.9924.9949.9574.93
Cd2.014.8010.0515.08
Pb1537.575.0112.5

Note. The 80% confluent rat primary astrocytes were pretreated in reduced serum medium for 2 h and then treated with As, Cd, and Pb at concentrations ranging from 0.01 to 200μM in reduced serum medium for 18 h, and cell viability was determined (refer to the “Materials and Methods” section).

TABLE 5

LC Values of As, Cd, and Pb in Rat Primary Astrocytes

LC10 (μM)LC25 (μM)LC50 (μM)LC75
As9.9924.9949.9574.93
Cd2.014.8010.0515.08
Pb1537.575.0112.5
LC10 (μM)LC25 (μM)LC50 (μM)LC75
As9.9924.9949.9574.93
Cd2.014.8010.0515.08
Pb1537.575.0112.5

Note. The 80% confluent rat primary astrocytes were pretreated in reduced serum medium for 2 h and then treated with As, Cd, and Pb at concentrations ranging from 0.01 to 200μM in reduced serum medium for 18 h, and cell viability was determined (refer to the “Materials and Methods” section).

TABLE 6

MM-Induced Synergistic Reduction in Viability and Increase in Apoptosis in Rat Primary Astrocytes

As, Cd, and Pb concentration in MMViability %
Apoptotic index %
AdMMAdMM
As (3.33μM) + Cd (0.67μM) + Pb (5μM)90.59 ± 0.4672.25 ± 5.39***10.31 ± 0.6222.865 ± 0.67***
As (8.33μM) + Cd (1.67μM) + Pb (12.5μM)75.25 ± 2.9753.09 ± 3.21***25.25 ± 0.7343.23 ± 1.74***
As (16.67μM) + Cd (3.33μM) + Pb (25μM)50.32 ± 1.7627.96 ± 2.73***49.87 ± 2.0367.56 ± 3.82***
As (25μM) + Cd (5μM) + Pb (37.5μM)25.16 ± 0.977.43 ± 1.52***73.98 ± 2.5698.42 ± 5.05***
As, Cd, and Pb concentration in MMViability %
Apoptotic index %
AdMMAdMM
As (3.33μM) + Cd (0.67μM) + Pb (5μM)90.59 ± 0.4672.25 ± 5.39***10.31 ± 0.6222.865 ± 0.67***
As (8.33μM) + Cd (1.67μM) + Pb (12.5μM)75.25 ± 2.9753.09 ± 3.21***25.25 ± 0.7343.23 ± 1.74***
As (16.67μM) + Cd (3.33μM) + Pb (25μM)50.32 ± 1.7627.96 ± 2.73***49.87 ± 2.0367.56 ± 3.82***
As (25μM) + Cd (5μM) + Pb (37.5μM)25.16 ± 0.977.43 ± 1.52***73.98 ± 2.5698.42 ± 5.05***

Note. The 80% confluent rat primary astrocytes were pretreated in reduced serum medium for 2 h and then treated with As, Cd, and Pb or MM at 1/3 of LC10, LC25, LC50, and LC75 of the metals in reduced serum medium for 18 h, and cell viability and apoptosis were determined (refer to the “Materials and Methods” section). The viability and apoptotic index of the astrocytes for every single metal were added (Ad). This was compared with viability and apoptotic index of MM-treated cells. Data represent means ± SE of four independent experiments in triplicates.

***p < 0.001 (compared with Ad).

TABLE 6

MM-Induced Synergistic Reduction in Viability and Increase in Apoptosis in Rat Primary Astrocytes

As, Cd, and Pb concentration in MMViability %
Apoptotic index %
AdMMAdMM
As (3.33μM) + Cd (0.67μM) + Pb (5μM)90.59 ± 0.4672.25 ± 5.39***10.31 ± 0.6222.865 ± 0.67***
As (8.33μM) + Cd (1.67μM) + Pb (12.5μM)75.25 ± 2.9753.09 ± 3.21***25.25 ± 0.7343.23 ± 1.74***
As (16.67μM) + Cd (3.33μM) + Pb (25μM)50.32 ± 1.7627.96 ± 2.73***49.87 ± 2.0367.56 ± 3.82***
As (25μM) + Cd (5μM) + Pb (37.5μM)25.16 ± 0.977.43 ± 1.52***73.98 ± 2.5698.42 ± 5.05***
As, Cd, and Pb concentration in MMViability %
Apoptotic index %
AdMMAdMM
As (3.33μM) + Cd (0.67μM) + Pb (5μM)90.59 ± 0.4672.25 ± 5.39***10.31 ± 0.6222.865 ± 0.67***
As (8.33μM) + Cd (1.67μM) + Pb (12.5μM)75.25 ± 2.9753.09 ± 3.21***25.25 ± 0.7343.23 ± 1.74***
As (16.67μM) + Cd (3.33μM) + Pb (25μM)50.32 ± 1.7627.96 ± 2.73***49.87 ± 2.0367.56 ± 3.82***
As (25μM) + Cd (5μM) + Pb (37.5μM)25.16 ± 0.977.43 ± 1.52***73.98 ± 2.5698.42 ± 5.05***

Note. The 80% confluent rat primary astrocytes were pretreated in reduced serum medium for 2 h and then treated with As, Cd, and Pb or MM at 1/3 of LC10, LC25, LC50, and LC75 of the metals in reduced serum medium for 18 h, and cell viability and apoptosis were determined (refer to the “Materials and Methods” section). The viability and apoptotic index of the astrocytes for every single metal were added (Ad). This was compared with viability and apoptotic index of MM-treated cells. Data represent means ± SE of four independent experiments in triplicates.

***p < 0.001 (compared with Ad).

We determined the effect of As, Cd, Pb, or of MM on the astrocyte apoptotic index. We observed that the MM synergistically promoted apoptosis in the astrocytes (Table 6).

Effect of MM on Activation of the MAPK Signaling Pathways in the Rat Primary Astrocytes

Possible involvement of the signaling pathways, such as PI3 kinase, MEK1/2, P38-MAPK, and JNK1/2, was determined in MM-induced toxicity in the astrocytes. We incubated the astrocytes with the MM and cotreated the cells with LY294002 (10μM), PD98059 (10μM), SB203580 (10μM), or SP600125 (10μM), which are inhibitors to PI3 kinase, MEK1/2, P38-MAPK, and JNK1/2 pathways, respectively. Inhibitors themselves were nontoxic to astrocytes. Treatment with PD98059 or SP600125 promoted viability of MM-treated astrocytes (Fig. 4A) but LY294002 or SB203580 did not, suggesting the involvement of MEK1/2 and JNK1/2 pathways in MM-induced toxicity.

MM activated MAPK signaling in MM-treated rat primary astrocytes. (A) Eighty percent confluent rat primary astrocytes were pretreated with LY294002 (LY), PD98059 (PD), SB203580 (SB), or SP600125 (SP) in reduced serum medium for 2 h. The cells were then treated with MM in reduced serum medium for additional 18 h and the MTT assay was carried out. Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with MM).
FIG. 4.

MM activated MAPK signaling in MM-treated rat primary astrocytes. (A) Eighty percent confluent rat primary astrocytes were pretreated with LY294002 (LY), PD98059 (PD), SB203580 (SB), or SP600125 (SP) in reduced serum medium for 2 h. The cells were then treated with MM in reduced serum medium for additional 18 h and the MTT assay was carried out. Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with MM).

The 80% confluent rat primary astrocytes were pre-incubated in reduced serum medium for 2h, treated with MM in reduced serum medium, and the cell lysates collected at the indicated time points. Western blotting on the cell lysates was carried out sequentially with antibodies to polyclonal phospho-ERK1/2 and polyclonal ERK1/2 or polyclonal phospho-JNK1/2 (54 and 46 kDa) and polyclonal JNK1/2 (54 and 46 kDa) antibodies. (B) Representative Western blot and densitometric analyses of the phospho-ERK1/2 (44 and 42 kDa) relative to ERK1/2 (44 and 42 kDa). Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with 0 min). (C) Representative Western blot and densitometric analyses of the phospho-JNK1/2 (54 and 46 kDa) relative to JNK1/2 (54 and 46 kDa). Data represent means ± SE of four independent experiments in triplicates. *p < 0.05 and ***p < 0.001 (compared with 0 min). Eighty percent confluent rat primary astrocytes were pre-incubated in reduced serum medium with PD98059 or SP600125 for 2 h and then treated with MM (at individual LC10 values). The cell lysates were collected at the indicated time points. Western blotting on the cell lysates was carried out sequentially with polyclonal phospho-JNK and polyclonal JNK antibodies (for PD98059) or phospho-ERK1/2 and polyclonal ERK1/2 antibodies (for SP600125). (D) Representative Western blot showing phospho-JNK (54 and 46 kDa) and JNK bands (54 and 46 kDa). (E) Representative Western blot and densitometric analyses of phospho-ERK1/2 relative to ERK1/2 bands (44 and 42 kDa). Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with 0 min).

We then studied the time course of phosphorylation of MEK1/2 (assessed by ERK1/2 phosphorylation) and JNK1/2. We observed that phosphorylation of ERK1/2 started before 5 min, reached its peak at 10 min, and declined to less than the basal level in 30 min (Fig. 4B). The phosphorylation of JNK1/2 started at 10 min, reached its peak at 15 min, and declined thereafter to the basal level in 30 min (Fig. 4C). Because MM-induced phosphorylation of ERK1/2 preceded phosphorylations of JNK1/2, we determined whether activation of JNK1/2 was ERK1/2 dependent. We simultaneously incubated the primary astrocytes with PD98059 and MM, and observed that PD98059 blocked MM-stimulated phosphorylation of JNK1/2 (Fig. 4D). But upon incubation of the cells with SP600125 (10μM) and MM, the former failed to block MM-stimulated phosphorylation of ERK1/2 (Fig. 4E). These data suggested that MM toxicity in rat primary astrocytes involves an upstream MEK1/2 activation followed by JNK1/2.

Effect of MM on [Ca2+]i and ROS Generation in the Rat Primary Astrocytes

It has been previously reported that induction of apoptosis by metals correlates with [Ca2+]i release and ROS generation (Yang et al., 2008) in astrocytes, and therefore, we investigated the effect of MM on them.

We treated the astrocytes with MM and observed that the MM triggered [Ca2+]i release. The [Ca2+]i release reached its peak after 30 min of MM treatment (Fig. 5A). Similarly, MM triggered ROS generation, and the ROS generation reached its peak after 1 h of MM treatment (Fig. 5B).

MM-induced apoptosis involved activation of ERK1/2 and JNK1/2, increased [Ca2+]i release, and ROS generation in rat primary astrocytes. (A) Eighty percent confluent rat primary astrocytes were pre-incubated in reduced serum medium for 2 h and then treated with MM (at LC10 values of metals) for 30 min, followed by [Ca2+]i assay (refer to the “Materials and Methods” section). Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with V). (B) Eighty percent confluent rat primary astrocytes were pre-incubated in reduced serum medium for 2 h and then treated with MM (at LC10 values of metals) for 1 h, followed by ROS assay (refer to the “Materials and Methods” section). Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with V). (C) Eighty percent confluent rat primary astrocytes were pretreated with α-tocopherol, SP600125, or PD98059 in reduced serum medium and then treated with MM for additional 30 min, followed by [Ca2+]i assay. Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with MM). (D) Eighty percent confluent rat primary astrocytes were pretreated with BAPTA-AM, SP600125, or PD98059 in reduced serum medium and then treated with MM for additional 1 h, followed by ROS assay. Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with MM). (E) Eighty percent confluent rat primary astrocytes were pretreated with α-tocopherol, BAPTA-AM, SP600125, or PD98059 in reduced serum medium, and then treated with MM for additional 18 h. Apoptotic index was then determined through TUNEL assay. Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with MM).
FIG. 5.

MM-induced apoptosis involved activation of ERK1/2 and JNK1/2, increased [Ca2+]i release, and ROS generation in rat primary astrocytes. (A) Eighty percent confluent rat primary astrocytes were pre-incubated in reduced serum medium for 2 h and then treated with MM (at LC10 values of metals) for 30 min, followed by [Ca2+]i assay (refer to the “Materials and Methods” section). Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with V). (B) Eighty percent confluent rat primary astrocytes were pre-incubated in reduced serum medium for 2 h and then treated with MM (at LC10 values of metals) for 1 h, followed by ROS assay (refer to the “Materials and Methods” section). Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with V). (C) Eighty percent confluent rat primary astrocytes were pretreated with α-tocopherol, SP600125, or PD98059 in reduced serum medium and then treated with MM for additional 30 min, followed by [Ca2+]i assay. Data represent means ± SE of four independent experiments in triplicates. ***p < 0.001 (compared with MM). (D) Eighty percent confluent rat primary astrocytes were pretreated with BAPTA-AM, SP600125, or PD98059 in reduced serum medium and then treated with MM for additional 1 h, followed by ROS assay. Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with MM). (E) Eighty percent confluent rat primary astrocytes were pretreated with α-tocopherol, BAPTA-AM, SP600125, or PD98059 in reduced serum medium, and then treated with MM for additional 18 h. Apoptotic index was then determined through TUNEL assay. Data represent means ± SE of four independent experiments in triplicates. **p < 0.01 and ***p < 0.001 (compared with MM).

To investigate whether the [Ca2+]i release was ROS, ERK1/2, or JNK1/2 –dependent, we incubated the MM-treated astrocytes with an antioxidant (α-tocopherol, 200 μg/ml), PD98059 (10μM), or SP600125 (10μM). α-Tocopherol itself was nontoxic. We observed that PD98059 (10μM) or SP600125 (10μM) suppressed [Ca2+]i release, but α-tocopherol (200 μg/ml) did not (Fig. 5C). This suggested that [Ca2+]i release in MM-treated astrocytes was ERK1/2 and JNK1/2 dependent.

To investigate whether the ROS generation was [Ca2+]i, ERK1/2, or JNK1/2 dependent, we incubated the MM-treated astrocytes with [Ca2+]i chelator (BAPTA-AM, 5μM), PD98059 (10μM), or SP600125 (10μM). BAPTA-AM itself was nontoxic to astrocytes. We observed that BAPTA-AM, PD98059, or SP600125 suppressed ROS generation in MM-treated rat primary astrocytes (Fig. 5D), suggesting that ROS generation was [Ca2+]i, ERK1/2, and JNK1/2 dependent.

We next investigated whether ERK1/2, JNK1/2, [Ca2+]i and ROS signaling resulted in apoptosis. We treated the MM-treated astrocytes with α-tocopherol (200 μg/ml), PD98059 (10μM), BAPTA-AM (5μM), or SP600125 (10μM) and observed that they all suppressed apoptosis (Fig. 5E). This suggested that activation of ERK1/2 and JNK1/2, followed by increased [Ca2+]i and ROS generation, resulted in apoptosis in the MM-treated astrocytes.

DISCUSSION

Understanding the cellular mechanisms leading to pathological alterations in developing rat brain by a mixture of heavy metals containing As, Cd, and Pb prompted us to undertake this study. Major findings of our study are that MM affects neurobehavioral parameters in developing rats that may be triggered by a compromised blood-brain communication. The metals act in synergy to reduce the expression of GFAP, an important protein component of BBB (McCall et al., 1996), accompanied by increase in apoptosis and morphological alterations in GFAP-expressing astrocytes. The synergistic apoptotic effect on astrocytes involves initial activation of ERK, which in turn activates the JNK pathway, followed by an increase in Ca2+i release and ROS generation.

We investigated neurobehavioral impact of MM during rat brain development. Our data show that MM treatment induced hyperlocomotion, enhanced grip strength, and reduced learning-memory functions in rats. The abnormalities in neurobehavior can be explained by altered glia-neuron interactions (Sykova, 2001) resulting from reduced GFAP level that plays a vital role in modulating synaptic efficacy in the CNS (Garcia et al., 2003; McCall et al., 1996). Moreover, the synapse being tightly ensheathed by astrocyte processes (Sykova, 2001, 2005), change in the number of processes may modulate synaptic activity (McCall et al., 1996) leading to behavioral impairments.

The synthesis of γ-amino butyric acid (GABA; the chief inhibitory neurotransmitter in the mammalian CNS) in neurons is closely associated with astrocyte metabolism (Sonnewald et al., 1993). Thus, it can be suggested that the damaged astrocytes induced by MM may play an important role in the pathophysiology of GABA hypofunction (Kondziella et al., 2006)–induced hyperactivity (hyperlocomotion and enhanced grip strength). Moreover, marked leakage of BBB in the white matter, corpus callosum, and posterior lobe of cerebral cortex and cerebellum may cause profound modulation in coordinated stimulation and integrated information from different neural circuits (Habas, 2001; Poliak and Peles, 2003) leading to the observed behavioral alterations.

Astrocytes play an important role in long-term potentiation (LTP) (McCall et al., 1996) that is widely considered as one of the major mechanisms that underlies learning and memory (Bliss and Collingridge, 1993; Cooke and Bliss, 2006). The astrocytes are a source of neurotrophic factors (Muller et al., 1995; Rudge et al., 1992), and reduced GFAP level interferes with their production (McCall et al., 1996). Moreover, GFAP directly participates in the change in arborization of astrocytic processes during LTP (Wenzel et al., 1991). Thus, reduced GFAP expression by MM may interfere in the necessary signaling between astrocytes and neurons or synapses, and impair learning-memory performance in developing rats.

An immature or disrupted BBB appeared to be vulnerable to metal permeation (Lidsky and Schneider, 2003; Wang et al., 2007b) that led to significant dose-dependent rise in levels of As, Cd, and Pb. The rise in metal content in developing rat brain confirmed metal passage to the postnatal rats from lactating mothers. Prolonged metal treatment sustained BBB damage in the P60 rats indicating reduced resistance to the effect of the metals even in adults.

GFAP is an established marker of astrocyte maturation and reactivity (Gomes et al., 1999). We observed that MM induced reduction in cortical and cerebellar GFAP levels, and increased apoptosis in GFAP-ir astrocytes. MM exposure reduced GFAP-expressing astrocyte count, size, and number of projections, thus suggesting that it adversely impacts astrocyte function. Our data extend previous observation of reduced glial differentiation by either Pb or Cd in developing rat brain (Chow et al., 2008; Zawia and Harry, 1996). In addition, our data show that the reduction in GFAP levels by MM exposure is more than the sum of effects of individual metal, thus suggesting higher risk of the metals in mixture than expected or theoretically predicted. Therefore, addition of As enhances previously reported synergistic toxic effect of Pb and Cd in the brain cells of cerebellum and cortex (Gu et al., 2009).

Inhibition of GFAP immunoreactivity by MM in developing brain appears to be caused by astrocyte apoptosis. In primary cultures of astrocytes, our data show that MM synergistically induced apoptosis. This was manifested by the activation of MEK/ERK, followed by the activation of JNK pathways, which then enhanced intracellular Ca2+ levels and subsequently ROS generation. There are reports on the activation of JNK1/2 downstream to ERK1/2 in pluripotent HT29 intestinal cell line (Salh et al., 2000) and human melanoma (Lopez-Bergami et al., 2007). Cd is reported to induce apoptosis via the activation of JNK-mediated signaling where Ca2+ and oxidative stress act as the pivotal mediators of apoptotic signaling in skin epidermal cell line (Son et al., 2010). But we are not aware of any report describing our observed sequences of events leading to astrocyte apoptosis by metals.

Astrocytes are networked together by a series of gap junctions permitting to propagate Ca2+ waves through the linked network (Lobsiger and Cleveland, 2007), and Ca2+-mediated intercellular communication is a mechanism by which astrocytes communicate with each other and modulate the activity of adjacent cells (Verderio and Matteoli, 2001). MM-induced alteration in astrocyte morphology may influence [Ca2+]i (Barres et al., 1989); in contrast, an increase in [Ca2+]i may also play a key role in altering astrocyte cytoskeleton, affecting the glia-neuron interaction (Shelton and McCarthy, 2000).

Previous reports show that As (Jin et al., 2004) and Pb (Zhang et al., 2008) induced apoptosis in astrocyte culture, and Ca2+ chelation and increase in antioxidant defense system protected astrocytes from Cd insults (Yang et al., 2008). Our data, for the first time, shed new light on the toxic mechanism of the three metals in combination involving synergistic loss in viability and enhanced apoptosis in rat primary astrocytes.

Overall, this study emphasizes that MM exposure during brain development has greater than additive response on astrocyte toxicity with adverse impacts on BBB function that culminate into neurological deficits in developing rats. This study underscores the importance of synergistic action of metals to that compared individually and hence would warrant revision of the threshold of metal toxicity and the current cutoff values of such environmental contaminants in assessing relative neurodevelopmental risks.

FUNDING

SERC Fast Track Scheme For Young Scientists of the Department of Science and Technology, Government of India; Supra Institutional Project of the Council of Scientific and Industrial Research, India.

We thank Dr Naibedya Chattopadhyay (scientist, Central Drug Research Institute), Dr Rohit A. Sinha (researcher, Duke-Nus Graduate Medical School), Dr A. K. Agarwal (emeritus scientist, Indian Institute of Toxicology Research), and Dr Preeti Srivastava (scientist, Indian Institute of Toxicology Research) for helpful suggestions and insightful comments.

References

Ali
MM
Mathur
N
Chandra
SV

Effect of chronic cadmium exposure on locomotor behaviour of rats
Indian J. Exp. Biol.
1990
28
653
656

Antonio
MT
Lopez
N
Leret
ML

Pb and Cd poisoning during development alters cerebellar and striatal function in rats
Toxicology
2002
176
1–2
59
66

Arey
BJ
Seethala
R
Ma
Z
Fura
A
Morin
J
Swartz
J
Vyas
V
Yang
W
Dickson
JK
Jr
Feyen
JH

A novel calcium-sensing receptor antagonist transiently stimulates parathyroid hormone secretion in vivo
Endocrinology
2005
146
2015
2022

Bandyopadhyay
S
Hartley
DM
Cahill
CM
Lahiri
DK
Chattopadhyay
N
Rogers
JT

Interleukin-1alpha stimulates non-amyloidogenic pathway by alpha-secretase (ADAM-10 and ADAM-17) cleavage of APP in human astrocytic cells involving p38 MAP kinase
J. Neurosci. Res.
2006
84
106
118

Bandyopadhyay
S
Jeong
KH
Hansen
JT
Vassilev
PM
Brown
EM
Chattopadhyay
N

Calcium-sensing receptor stimulates secretion of an interferon-gamma-induced monokine (CXCL10) and monocyte chemoattractant protein-3 in immortalized GnRH neurons
J. Neurosci. Res.
2007
85
882
895

Barres
BA
Chun
LL
Corey
DP

Calcium current in cortical astrocytes: induction by cAMP and neurotransmitters and permissive effect of serum factors
J. Neurosci.
1989
9
3169
3175

Bayer
SA

Cellular aspects of brain development
Neurotoxicology
1989
10
307
320

Behari
JR
Prakash
R

Determination of total arsenic content in water by atomic absorption spectroscopy (AAS) using vapour generation assembly (VGA)
Chemosphere
2006
63
17
21

Benitez
MA
Mendez-Armenta
M
Montes
S
Rembao
D
Sanin
LH
Rios
C

Mother-fetus transference of lead and cadmium in rats: involvement of metallothionein
Histol. Histopathol.
2009
24
1523
1530

Berenbaum
MC

The expected effect of a combination of agents: the general solution
J. Theor. Biol.
1985
114
413
431

Bliss
TV
Collingridge
GL

A synaptic model of memory: long-term potentiation in the hippocampus
Nature
1993
361
6407
31
39

Brender
JD
Suarez
L
Felkner
M
Gilani
Z
Stinchcomb
D
Moody
K
Henry
J
Hendricks
K

Maternal exposure to arsenic, cadmium, lead, and mercury and neural tube defects in offspring
Environ. Res.
2006
101
132
139

Choi
YK
Kim
KW

Blood-neural barrier: its diversity and coordinated cell-to-cell communication
BMB Rep.
2008
41
345
352

Chow
ES
Hui
MN
Lin
CC
Cheng
SH

Cadmium inhibits neurogenesis in zebrafish embryonic brain development
Aquat. Toxicol.
2008
87
157
169

Cooke
SF
Bliss
TV

Plasticity in the human central nervous system
Brain
2006
129
Pt 7
1659
1673

Counter
SA
Buchanan
LH
Ortega
F

Lead concentrations in maternal blood and breast milk and pediatric blood of Andean villagers: 2006 follow-up investigation
J. Occup. Environ. Med.
2007
49
302
309

Duran-Vilaregut
J
del Valle
J
Camins
A
Pallas
M
Pelegri
C
Vilaplana
J

Blood-brain barrier disruption in the striatum of rats treated with 3-nitropropionic acid
Neurotoxicology
2009
30
136
143

Fowler
BA
Whittaker
MH
Lipsky
M
Wang
G
Chen
XQ

Oxidative stress induced by lead, cadmium and arsenic mixtures: 30-day, 90-day, and 180-day drinking water studies in rats: an overview
Biometals
2004
17
567
568

Garcia-Chavez
E
Jimenez
I
Segura
B
Del Razo
LM

Lipid oxidative damage and distribution of inorganic arsenic and its metabolites in the rat nervous system after arsenite exposure: influence of alpha tocopherol supplementation
Neurotoxicology
2006
27
1024
1031

Garcia
SJ
Seidler
FJ
Slotkin
TA

Developmental neurotoxicity elicited by prenatal or postnatal chlorpyrifos exposure: effects on neurospecific proteins indicate changing vulnerabilities
Environ. Health Perspect.
2003
111
297
303

Gomes
FC
Paulin
D
Moura Neto
V

Glial fibrillary acidic protein (GFAP): modulation by growth factors and its implication in astrocyte differentiation
Braz. J. Med. Biol. Res.
1999
32
619
631

Grandjean
P
Landrigan
PJ

Developmental neurotoxicity of industrial chemicals
Lancet
2006
368
2167
2178

Gu
C
Chen
S
Xu
X
Zheng
L
Li
Y
Wu
K
Liu
J
Qi
Z
Han
D
Chen
G
et al. 

Lead and cadmium synergistically enhance the expression of divalent metal transporter 1 protein in central nervous system of developing rats
Neurochem. Res.
2009
34
1150
1150

Habas
C

[The cerebellum: from motor coordination to cognitive function]
Rev. Neurol. (Paris)
2001
157
1471
1497

Hertzberg
RC
MacDonell
MM

Synergy and other ineffective mixture risk definitions
Sci. Total Environ.
2002
288
1–2
31
42

Holson
RR
Freshwater
L
Maurissen
JP
Moser
VC
Phang
W

Statistical issues and techniques appropriate for developmental neurotoxicity testing: a report from the ILSI Research Foundation/Risk Science Institute expert working group on neurodevelopmental endpoints
Neurotoxicol. Teratol.
2008
30
326
348

Hu
H

Exposure to metals
Prim. Care
2000
27
983
996

Ishitobi
H
Mori
K
Yoshida
K
Watanabe
C

Effects of perinatal exposure to low-dose cadmium on thyroid hormone-related and sex hormone receptor gene expressions in brain of offspring
Neurotoxicology
2007
28
790
797

Izawa
Y
Takahashi
S
Suzuki
N

Pioglitazone enhances pyruvate and lactate oxidation in cultured neurons but not in cultured astroglia
Brain Res.
2009
1305
64
73

Jadhav
SH
Sarkar
SN
Patil
RD
Tripathi
HC

Effects of subchronic exposure via drinking water to a mixture of eight water-contaminating metals: a biochemical and histopathological study in male rats
Arch. Environ. Contam. Toxicol.
2007
53
667
677

Jin
Y
Sun
G
Li
X
Li
G
Lu
C
Qu
L

Study on the toxic effects induced by different arsenicals in primary cultured rat astroglia
Toxicol. Appl. Pharmacol.
2004
196
396
403

Kippler
M
Lonnerdal
B
Goessler
W
Ekstrom
EC
Arifeen
SE
Vahter
M

Cadmium interacts with the transport of essential micronutrients in the mammary gland—a study in rural Bangladeshi women
Toxicology
2009
257
1–2
64
69

Kondziella
D
Brenner
E
Eyjolfsson
EM
Markinhuhta
KR
Carlsson
ML
Sonnewald
U

Glial-neuronal interactions are impaired in the schizophrenia model of repeated MK801 exposure
Neuropsychopharmacology
2006
31
1880
1887

Lafuente
A
Marquez
N
Pazo
D
Esquifino
AI

Cadmium effects on dopamine turnover and plasma levels of prolactin, GH and ACTH
J. Physiol. Biochem.
2001
57
231
236

Landrigan
PJ
Whitworth
RH
Baloh
RW
Staehling
NW
Barthel
WF
Rosenblum
BF

Neuropsychological dysfunction in children with chronic low-level lead absorption
Lancet
1975
1
708
712

Lanphear
BP
Hornung
R
Khoury
J
Yolton
K
Baghurst
P
Bellinger
DC
Canfield
RL
Dietrich
KN
Bornschein
R
Greene
T
et al. 

Low-level environmental lead exposure and children’s intellectual function: an international pooled analysis
Environ. Health Perspect.
2005
113
894
899

Leret
ML
Millan
JA
Antonio
MT

Perinatal exposure to lead and cadmium affects anxiety-like behaviour
Toxicology
2003
186
1–2
125
130

Lidsky
TI
Schneider
JS

Lead neurotoxicity in children: basic mechanisms and clinical correlates
Brain
2003
126
Pt 1
5
19

Lin
JL
Huang
YH
Shen
YC
Huang
HC
Liu
PH

Ascorbic acid prevents blood-brain barrier disruption and sensory deficit caused by sustained compression of primary somatosensory cortex
J. Cereb. Blood Flow Metab
2010
30
1121
1136

Lobsiger
CS
Cleveland
DW

Glial cells as intrinsic components of non-cell-autonomous neurodegenerative disease
Nat. Neurosci.
2007
10
1355
1360

Lopez-Bergami
P
Huang
C
Goydos
JS
Yip
D
Bar-Eli
M
Herlyn
M
Smalley
KS
Mahale
A
Eroshkin
A
Aaronson
S
et al. 

Rewired ERK-JNK signaling pathways in melanoma
Cancer Cell
2007
11
447
460

Marchetti
C

Molecular targets of lead in brain neurotoxicity
Neurotox. Res.
2003
5
221
236

McCall
MA
Gregg
RG
Behringer
RR
Brenner
M
Delaney
CL
Galbreath
EJ
Zhang
CL
Pearce
RA
Chiu
SY
Messing
A

Targeted deletion in astrocyte intermediate filament (Gfap) alters neuronal physiology
Proc. Natl. Acad. Sci. U.S.A.
1996
93
6361
6366

Minetti
A
Reale
CA

Sensorimotor developmental delays and lower anxiety in rats prenatally exposed to cadmium
J. Appl. Toxicol.
2006
26
35
41

Muller
HW
Junghans
U
Kappler
J

Astroglial neurotrophic and neurite-promoting factors
Pharmacol. Ther.
1995
65
1
18

Ong
WY
He
X
Chua
LH
Ong
CN

Increased uptake of divalent metals lead and cadmium into the brain after kainite-induced neuronal injury
Exp. Brain Res.
2006
173
468
474

Otani
A
Takagi
H
Oh
H
Koyama
S
Matsumura
M
Honda
Y

Expressions of angiopoietins and Tie2 in human choroidal neovascular membranes
Invest. Ophthalmol. Vis. Sci.
1999
40
1912
1920

Pekny
M
Stanness
KA
Eliasson
C
Betsholtz
C
Janigro
D

Impaired induction of blood-brain barrier properties in aortic endothelial cells by astrocytes from GFAP-deficient mice
Glia
1998
22
390
400

Poliak
S
Peles
E

The local differentiation of myelinated axons at nodes of Ranvier
Nat. Rev. Neurosci.
2003
4
968
980

Posser
T
de Aguiar
CB
Garcez
RC
Rossi
FM
Oliveira
CS
Trentin
AG
Neto
VM
Leal
RB

Exposure of C6 glioma cells to Pb(II) increases the phosphorylation of p38(MAPK) and JNK1/2 but not of ERK1/2
Arch. Toxicol.
2007
81
407
414

Prior
MJ
Brown
AM
Mavroudis
G
Lister
T
Ray
DE

MRI characterisation of a novel rat model of focal astrocyte loss
MAGMA
2004
17
125
132

Rader
JI
Peeler
JT
Mahaffey
KR

Comparative toxicity and tissue distribution of lead acetate in weanling and adult rats
Environ. Health Perspect.
1981
42
187
195

Rice
D
Barone
S
Jr

Critical periods of vulnerability for the developing nervous system: evidence from humans and animal models
Environ. Health Perspect.
2000
108
Suppl. 3
511
533

Rodriguez
VM
Carrizales
L
Mendoza
MS
Fajardo
OR
Giordano
M

Effects of sodium arsenite exposure on development and behavior in the rat
Neurotoxicol. Teratol.
2002
24
743
750

Rodriguez
VM
Dufour
L
Carrizales
L
Diaz-Barriga
F
Jimenez-Capdeville
ME

Effects of oral exposure to mining waste on in vivo dopamine release from rat striatum
Environ. Health Perspect.
1998
106
487
491

Rudge
JS
Alderson
RF
Pasnikowski
E
McClain
J
Ip
NY
Lindsay
RM

Expression of ciliary neurotrophic factor and the neurotrophins-nerve growth factor, brain-derived neurotrophic factor and neurotrophin 3-in cultured rat hippocampal astrocytes
Eur. J. Neurosci.
1992
4
459
471

Salh
BS
Martens
J
Hundal
RS
Yoganathan
N
Charest
D
Mui
A
Gomez-Munoz
A

PD98059 attenuates hydrogen peroxide-induced cell death through inhibition of Jun N-Terminal Kinase in HT29 cells
Mol. Cell Biol. Res. Commun.
2000
4
158
165

Samanta
G
Das
D
Mandal
BK
Chowdhury
TR
Chakraborti
D
Pal
A
Ahamed
S

Arsenic in the breast milk of lactating women in arsenic-affected areas of West Bengal, India and its effect on infants
J. Environ. Sci. Health A Tox. Hazard. Subst. Environ. Eng.
2007
42
1815
1825

Sanders
JL
Chattopadhyay
N
Kifor
O
Yamaguchi
T
Butters
RR
Brown
EM

Extracellular calcium-sensing receptor expression and its potential role in regulating parathyroid hormone-related peptide secretion in human breast cancer cell lines
Endocrinology
2000
141
4357
4364

Serota
RG

Acetoxycycloheximide and transient amnesia in the rat
Proc. Natl. Acad. Sci. U.S.A.
1971
68
1249
1250

Shalev
H
Serlin
Y
Friedman
A

Breaching the blood-brain barrier as a gate to psychiatric disorder
Cardiovasc. Psychiatry Neurol.
2009
2009
278531

Shelton
MK
McCarthy
KD

Hippocampal astrocytes exhibit Ca2+-elevating muscarinic cholinergic and histaminergic receptors in situ
J. Neurochem.
2000
74
555
563

Singh
S
Rana
SV

Amelioration of arsenic toxicity by L-Ascorbic acid in laboratory rat
J. Environ. Biol.
2007
28
2 Suppl.
377
384

Sinha
RA
Khare
P
Rai
A
Maurya
SK
Pathak
A
Mohan
V
Nagar
GK
Mudiam
MK
Godbole
MM
Bandyopadhyay
S

Anti-apoptotic role of omega-3-fatty acids in developing brain: perinatal hypothyroid rat cerebellum as apoptotic model
Int. J. Dev. Neurosci.
2009
27
377
383

Son
YO
Lee
JC
Hitron
JA
Pan
J
Zhang
Z
Shi
X

Cadmium induces intracellular Ca2+- and H2O2-dependent apoptosis through JNK- and p53-mediated pathways in skin epidermal cell line
Toxicol. Sci.
2010
113
127
137

Sonnewald
U
Westergaard
N
Schousboe
A
Svendsen
JS
Unsgard
G
Petersen
SB

Direct demonstration by [13C]NMR spectroscopy that glutamine from astrocytes is a precursor for GABA synthesis in neurons
Neurochem. Int.
1993
22
19
29

Sykova
E

Glial diffusion barriers during aging and pathological states
Prog. Brain Res.
2001
132
339
363

Sykova
E

Glia and volume transmission during physiological and pathological states
J. Neural Transm.
2005
112
137
147

Tanaka
J
Toku
K
Matsuda
S
Sudo
S
Fujita
H
Sakanaka
M
Maeda
N

Induction of resting microglia in culture medium devoid of glycine and serine
Glia
1998
24
198
215

Terry
AV
Jr
Stone
JD
Buccafusco
JJ
Sickles
DW
Sood
A
Prendergast
MA

Repeated exposures to subthreshold doses of chlorpyrifos in rats: hippocampal damage, impaired axonal transport, and deficits in spatial learning
J. Pharmacol. Exp. Ther.
2003
305
375
384

Tsai
SY
Chou
HY
The
HW
Chen
CM
Chen
CJ

The effects of chronic arsenic exposure from drinking water on the neurobehavioral development in adolescence
Neurotoxicology
2003
24
747
753

Valkonen
S
Savolainen
H
Jarvisalo
J

Arsenic distribution and neurochemical effects in peroral sodium arsenite exposure of rats
Bull. Environ. Contam. Toxicol.
1983
30
303
308

Verderio
C
Matteoli
M

ATP mediates calcium signaling between astrocytes and microglial cells: modulation by IFN-gamma
J. Immunol.
2001
166
6383
6391

Wang
KY
Jin
TY
Li
H
Shi
XQ

[Effects of cadmium on zinc metabolism and its functions in rats]
Zhonghua Lao Dong Wei Sheng Zhi Ye Bing Za Zhi
2007a
25
2
77
79

Wang
Q
Luo
W
Zheng
W
Liu
Y
Xu
H
Zheng
G
Dai
Z
Zhang
W
Chen
Y
Chen
J

Iron supplement prevents lead-induced disruption of the blood-brain barrier during rat development
Toxicol. Appl. Pharmacol.
2007b
219
33
41

Wenzel
J
Lammert
G
Meyer
U
Krug
M

The influence of long-term potentiation on the spatial relationship between astrocyte processes and potentiated synapses in the dentate gyrus neuropil of rat brain
Brain Res.
1991
560
122
131

Wessinger
WD

Approaches to the study of drug interactions in behavioral pharmacology
Neurosci. Biobehav. Rev.
1986
10
103
113

Wetzel
W
Matthies
H

Differences in Y-maze retention of rats: selection according to open field activity
Physiol. Behav.
1982
28
619
622

Willis
CL
Leach
L
Clarke
GJ
Nolan
CC
Ray
DE

Reversible disruption of tight junction complexes in the rat blood-brain barrier, following transitory focal astrocyte loss
Glia
2004
48
1
13

Wright
RO
Amarasiriwardena
C
Woolf
AD
Jim
R
Bellinger
DC

Neuropsychological correlates of hair arsenic, manganese, and cadmium levels in school-age children residing near a hazardous waste site
Neurotoxicology
2006
27
210
216

Wright
RO
Baccarelli
A

Metals and neurotoxicology
J. Nutr.
2007
137
2809
2813

Xi
S
Jin
Y
Lv
X
Sun
G

Distribution and speciation of arsenic by transplacental and early life exposure to inorganic arsenic in offspring rats
Biol. Trace Elem. Res.
2010
134
84
97

Yang
CS
Tzou
BC
Liu
YP
Tsai
MJ
Shyue
SK
Tzeng
SF

Inhibition of cadmium-induced oxidative injury in rat primary astrocytes by the addition of antioxidants and the reduction of intracellular calcium
J. Cell. Biochem.
2008
103
825
834

Zawia
NH
Harry
GJ

Developmental exposure to lead interferes with glial and neuronal differential gene expression in the rat cerebellum
Toxicol. Appl. Pharmacol.
1996
138
43
47

Zhang
Y
Sun
LG
Ye
LP
Wang
B
Li
Y

Lead-induced stress response in endoplasmic reticulum of astrocytes in CNS
Toxicol. Mech. Methods
2008
18
751
757

Zhang
YM
Liu
XZ
Lu
H
Mei
L
Liu
ZP

Lipid peroxidation and ultrastructural modifications in brain after perinatal exposure to lead and/or cadmium in rat pups
Biomed. Environ. Sci.
2009
22
423
429

Zhao
L
Wientjes
MG
Au
JL

Evaluation of combination chemotherapy: integration of nonlinear regression, curve shift, isobologram, and combination index analyses
Clin. Cancer Res.
2004
10
7994
8004

Zheng
W

Toxicology of choroid plexus: special reference to metal-induced neurotoxicities
Microsc. Res. Tech.
2001
52
89
103

Zhu
Y
Romero
MI
Ghosh
P
Ye
Z
Charnay
P
Rushing
EJ
Marth
JD
Parada
LF

Ablation of NF1 function in neurons induces abnormal development of cerebral cortex and reactive gliosis in the brain
Genes Dev.
2001
15
859
876

Comments

0 Comments
Submit a comment
You have entered an invalid code
Thank you for submitting a comment on this article. Your comment will be reviewed and published at the journal's discretion. Please check for further notifications by email.