PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJClRC/DatoDuepGS-p. 15 RELATIONSHIP BETWEEN METHYLMERCURY-INDUCED DISRUPTION OF INTRACELLULAR CALCIUM AND NEURONAL DEATH By Tobi Leigh Limke A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Pharmacology and Toxicology and Institute for Environmental Toxicology 2001 ABSTRACT RELATIONSHIP BETWEEN METHYLMERCURY—INDUCED DISRUPTION OF INTRACELLULAR CALCIUM AND NEURONAL DEATH By Tobi Leigh Limke The mechanism by which the environmental neurotoxicant methylmercury (MeHg) causes elevations of intracellular Ca2+ ([Ca2+]i) and subsequent neuronal death was examined using single cell microfluorimetry of rat cerebellar granule neurons in primary culture. In granule cells loaded with fura-2 to monitor changes in [Ca2+]i, 0.2-1.0 uM MeHg causes early release of Ca2+ from at least one intracellular source; the possible sources examined were the smooth endoplasmic reticulum (SER) and the mitochondria. The non-specific muscarinic acetylcholine (ACh) receptor antagonist atropine delayed MeHg-induced elevations of [Ca2+]i, while down-regulation of the muscarinic receptors and the inositol-1,4,5-triphosphate (1P3) receptor with 24 hr application of the muscarinic agonist bethanechol decreased the amplitude of the MeHg-induced release of Ca2+i by approximately 30-40%. Depletion of SER Ca2+ stores with thapsigargin also reduced the amplitude of the MeHg-induced release of Ca2+i by approximately 30-40%. Inhibition of Ca2+ release through the ryanodine receptors had minimal effect. Removal of mitochondrial Ca2+ (Ca2+m) content prior to MeHg exposure using carbonyl cyanide m- chlorophenylhydrazone (CCCP) and oligomycin decreased the amplitude of the MeHg- induced release of Ca2+i by approximately 70%. Additionally, inhibition of the mitochondrial permeability transition pore (MTP) using cyclosporin A (CsA) delayed the increase in [Ca2+]i. In granule cells loaded with tetramethylrhodamine ethyl ester (TMRE) to monitor changes in mitochondrial membrane potential, CsA delayed the irreversible loss of membrane potential caused by 0.5 uM MeHg. In granule cells loaded with rhod-2 to monitor changes in [CaZ+]m, MeHg caused an early increase in [Cazjm followed several minutes later by release of dye from the mitochondria into the cytosol. The initial increase in [Ca2+]m occurred independently of extracellular Ca2+, while the release of dye from the mitochondria was delayed by prior application of thapsigargin or CsA. Use of a calcein AM-ethidium homodimer cell viability assay revealed that increasing concentrations of MeHg (0.2-1.0 uM) caused a corresponding increase in cell death at 24 hr post-exposure. Atropine and ryanodine did not protect against MeHg- induced cell death, while 24 hr BCh pretreatment significantly protected against cell killing. The BCh-mediated protection was reversed by atropine and the M3 muscarinic ACh receptor antagonist 4-diphenylacetoxyl-N-methylpiperidine methiodide (4-DAMP) but not by the M2 receptor antagonist methoctramine or the nicotinic ACh receptor antagonist dihydro-B-erythroidine hydrobromide (DHE). Thapsigargin itself was toxic, highlighting the sensitivity of granule cells to disruption of SER Ca2+. CsA also provided significant protection against cell death at 24 hr post-MeHg exposure. These results suggest that MeHg: 1) acts at M3 muscarinic ACh receptors to cause production of 1P3, which releases Ca2+ from the SER and 2) causes Ca 2+ uptake into mitochondria which is then released into the cytosol via opening of the MTP. Additionally, these disruptions of [Ca2+]; contribute to MeHg-induced neuronal death, and may underlie the specific neurotoxicity of MeHg within granule neurons of the cerebellar cortex. For the family and friends who never gave up hope, and passed that hope to me. iv ACKNOWLEDGMENTS I would like to thank my parents, grandparents and extended family for their support and encouragement through the insanity of the past 4% years. Their emotional support and love will save me many years of therapy down the road. I would like to thank Dr. William Atchison and my guidance committee for their ideas and suggestions throughout my thesis, and their willingness to share their time to serve on the committee. Their input was invaluable. I would also like to thank my adopted family in the Atchison lab who were always willing to teach me something new, help out when there was a problem, and provided the most memorable conversations imaginable. Their patience and sense of humor were appreciated every day, and will not be forgotten. This dissertation would not have been possible without the excellent technical assistance of Erica Lange Fritz, Aizhen Yao, Steve Lundback, Dr. Ravindra Hajela and Dawn Autio. In particular, the amazing efforts of several undergraduates, including Scott A. Loiselle, Ilyana C. Martinez, Bermily Maldonado, Maria Sanchez and Rebecca Hall, and one seriously talented veterinary student, Jeremy J. Bearrs, allowed for more experiments to be performed than should be humanly possible. Input and guidance from Dr. Sue Marty was greatly appreciated. PREFACE Some of the data contained in this dissertation has been published previously. Portions of Chapter 5 appeared as Limke and Atchison (2002). vi TABLE OF CONTENTS LIST OF TABLES .................................................................................. x LIST OF FIGURES ................................................................................ xi LIST OF SYMBOLS AND ABBREVIATIONS ............................................ xvi CHAPTER ONE: INTRODUCTION ........................................................... 1 A) GENERAL INTRODUCTION ....................................................... 2 B) BACKGROUND ........................................................................ 9 a. Humans and Methylmercury: Routes and Biological Consequences of Exposure ........................................................................ 9 b. Cellular and Molecular Effects of Methylmercury ....................... 15 i. MeHg-Induced Cell Death: Apoptosis or Necrosis? ........... 16 ii. Calcium and Cell Death ............................................. 18 iii. MeHg and Cholinergic Neurotransmission ...................... 20 iv. Intracellular Calcium Stores ....................................... 23 v. MeHg and the Smooth Endoplasmic Reticulum ................ 29 vi. MeHg and Mitochondria ........................................... 33 vii. MeHg-Induced Interactions Between SER and Mitochondria Ca2+ Stores ............................................................ 35 viii. Non-Ca2+ Divalent Cations ........................................ 36 C) RAT CEREBELLAR GRANULE NEURONS IN PRIMARY CULTURE AS A MODEL FOR THE SELECTIVE NEUROTOXICITY OF MeHg .......... 37 vii CHAPTER TWO: INTERACTION OF METHYLMERCURY WITH MUSCARINIC RECEPTORS CAUSES CALCIUM RELEASE FROM THE SMOOTH ENDPLASMIC RETICULUM INOSITOL-l,4,5-TRIPHOSPHATE RECEPTORS IN RAT CEREBELLAR GRANULE NEURONS ....................... 40 A) ABSTRACT ............................................................................ 41 B) INTRODUCTION ..................................................................... 42 C) MATERIALS AND METHODS ................................................... 44 D) RESULTS .............................................................................. 49 E) DISCUSSION .......................................................................... 71 CHAPTER THREE: ACUTE METHYLMERCURY EXPOSURE CAUSES NEURONAL DEATH IN RAT CEREBELLAR GRANULE CELLS THROUGH INTERACTION WITH INOSITOL-l,4,5-TRIPHOSPHATE RECEPTORS ....... 76 A) ABSTRACT ........................................................................... 77 B) INTRODUCTION .................................................................... 78 C) MATERIALS AND METHODS ................................................... 81 D) RESULTS ............................................................................... 86 E) DISCUSSION .......................................................................... 98 CHAPTER FOUR: DISRUPTION OF MITOCHONDRIAL CALCIUM SIGNIFICANTLY CONTRIBUTES TO METHYLMERCURY-INDUCED ELEVATIONS OF [Ca2+], ...................................................................... 103 A) ABSTRACT ........................................................................... 104 B) INTRODUCTION ................................................................... 105 C) MATERIALS AND METHODS .................................................. 108 D) RESULTS ............................................................................. 111 E) DISCUSSION ........................................................................ 129 viii CHAPTER FIVE: ACUTE EXPOSURE TO METHYLMERCURY OPENS THE MITOCHONDRIAL PERMEABILITY TRANSITION PORE IN RAT CEREBELAR GRANULE CELLS .......................................................... 134 A) ABSTRACT .......................................................................... 134 B) INTRODUCTION .................................................................... 136 C) MATERIALS AND METHODS .................................................. 139 D) RESULTS ............................................................................. 144 E) DISCUSSION ........................................................................ 159 CHAPTER SIX: SUMMARY AND CONCLUSIONS .................................. 165 A) SUMMARY AND CONCLUSIONS ............................................. 166 REFERENCES .................................................................................... 178 ix LIST OF TABLES Table 2.1 Effect of pharmacological inhibitors on peak fura-2 ratio induced by 1.5 min 1 mM BCh in cerebellar granule cells ................................ 51 Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 LIST OF FIGURES MeHg causes a biphasic increase in [Ca2+], in cerebellar granule cells in primary culture ...................................................................... 5 Elevations of fuIa-2 fluorescence in a cerebellar granule cell in culture exposed to 0.5 uM MeHg in CaZI-fi'ee buffer .................................. 6 Hypothesis regarding MeHg-induced elevations of {Cap} in rat cerebellar granule neurons ..................................................................... 8 Elevations of fula-2 fluorescence in a cerebellar granule cell in culture exposed to 0.5 uM MeHg in Cari-flee buffer ................................. 14 Alternating exposures to 40 mM K+ and 1 mM BCh cause increases in [Ca2+], in cerebellar granule cells ............................................... 50 Thapsigargin does not affect the time-to-onset of MeHg-induced increases in [Ca2+], ............................................................................ 52 Thapsigargin does not affect the time-to-onset of MeHg-induced [Ca2+]i increases in the absence of Ca2+e ................................................ 53 Thapsigargin decreases the amplitude of MeHg-induced fura-2 ratio increases, as measured in the absence of Caz: ................................ 54 Pretreatment with BCh hastens the MeHg-induced release of [Ca2+], without affecting the onset of MeHg-induced influx of Ca2+: ............... 56 Pretreatment with BCh hastens the MeHg-induced release of [Ca2 i, as measured in the absence of Ca2+c ................................................ 57 xi Figure 2.7 Figure 2.8 Figure 2.9 Figure 2.10 Figure 2.11 Figure 2.12 Figure 2.13 Figure 2.14 Figure 2.15 Figure 2.16 Figure 3.1 Figure 3.2 Figure 3.3 Pretreatment with BCh (1 mM, 24 hr) does not increase the [Ca2+] within mitochondria ....................................................................... 58 Pretreatment with BCh decreases the amplitude of the fura-2 ratio increase caused by MeHg, as measured in the absence of Ca”... ...................... 59 Atropine (“AP”) delays the first-phase release of Car} without affecting the second-phase influx of Ca2+e caused by MeHg ............................ 61 Atropine (“AP) delays the time-to-onset of the MeHg-induced first-phase release of Ca2+;, as measured in the absence of Ca2+¢ ........................ 62 Atropine (“AP”) decreases the amplitude of the fura-2 ratio increase caused by MeHg, as measured in the absence of Ca”c ...................... 63 Ryanodine (“Ry”) does not affect the time-to-onset of MeHg-induced increases in [Ca2+], ................................................................ 66 Ryanodine (“Ry”) does not affect the time-tO-onset of the MeHg-induced first-phase release of Ca2+i, as measured in the absence of Ca2+e ........... 67 Ryanodine does not affect the amplitude of MeHg-induced increases in [Ca2+], as measured in the absence of Ca2+e .................................... 68 Depletion of SER Ca2+ triggers SOCE in rat cerebellar granule cells. . ....69 SK&F 96365 inhibits depolarization-stimulated Ca2+ influx through VOCCS .............................................................................. 7O MeHg-induced cell death at 30 min and 24 hr post-exposure ............... 87 24 hr pretreatment with 1 mM BCh protects against MeHg-induced cell death ................................................................................. 89 Atropine plus BCh does not protect against MeHg-induced cell death. . ..90 xii Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.3 4-DAMP plus BCh does not protect against MeHg-induced cell death. . .92 The nicotinic receptor antagonist DHE does not affect BCh-mediated protection against MeHg-induced cell death .................................. 93 Atropine alone does not protect against MeHg-induced cell death. . . . . ....94 Xestospongin C does not protect against MeHg-induced cell death ....... 95 Ryanodine does not protect against MeHg-induced cell death .............. 97 CCCP and oligomycin cause a rapid decrease in TMRE fluorescence with a concomitant increase in fura—2 fluorescence ............................... 112 Effect of oligomycin on TMRE fluorescence, as well as effect of CCCP on TMRE and fura—2 fluorescence ................................................ 113 Effect of mitochondrial inhibitors on first-phase increase in [Ca2+]i caused by 0.5 11M MeHg ................................................................ 115 Effect of CCCP and oligomycin on first-phase time-tO-onset and ratio amplitude induced by 0.5 11M MeHg in the absence of Ca2+: ............. 116 Comparison of time-to—onset of Ca2+c influx caused by 0.5 11M MeHg in the presence or absence of mitochondrial inhibitors ........................ 117 Changes in fura-2 and rhod-2 fluorescence in response to increasing concentrations of extracellular K+ ............................................. 120 MeHg causes a biphasic increase in rhod-2 fluorescence in granule cells. ...................................................................................... 122 Effect of CCCP and oligomycin on MeHg-induced elevations of rhod-2 fluorescence ...................................................................... 123 xiii Figure 4.9 Figure 4.10 Figure 4.11 Figure 4.12 Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Figure 5.6 Figure 5.7 Figure 5.8 Mitochondrial membrane depolarization abolished the initial elevation of rhod-2 fluorescence without affecting the time-to-onset of the Spreading of rhOd-2 throughout the cytosol .................................................. 124 Thapsigargin delays the spreading of rhod-2 from punctate to whole-cell staining as caused by 0.5 11M MeHg .......................................... 126 CSA delays the spreading of rhod-2 from punctate to whole-cell staining as caused by 0.5 11M MeHg .................................................... 127 Effect of CsA and thapsigargin on interval between onset of first and second rhOd-2 fluorescence increases ......................................... 128 Increasing concentrations of CSA cause increasing delays in the time-to- onset of [Ca2+]: elevations caused by 0.5 11M MeHg ........................ 145 5 11M CSA significantly delays the time-to-onset of MeHg-induced [Cay], increases ........................................................................... 146 5 11M CsA delays the time-to-onset of MeHg-induced release of Ca2+i in the absence of Ca2+c ............................................................. 149 Effect of calcineurin inhibitors on MeHg-induced [Ca2+], elevations. ...150 PhAsO causes a rapid decrease in TMRE fluorescence .................... 151 Exposure to 0.5 11M MeHg causes irreversible loss of TMRE fluorescence ...................................................................... 152 CSA, but not CN inhibitors, delay the time-tO-onset of mitochondrial membrane depolarization caused by 0.5 11M MeHg ........................ 153 CSA provides limited protection against MeHg-induced cell death at 24 hr post-exposure ..................................................................... 156 xiv Figure 5.9 Figure 5.10 Figure 6.1 Comparison of protection against MeHg-induced cell death by pretreatment or Simultaneous treatment with 5 11M CSA ................... 157 Pretreatment with 20 11M W7 significantly protects against cell death caused by 0.5 11M MeHg at 24 hr post-exposure ............................ 158 Summary of findings ............................................................ 167 XV LIST OF SYMBOLS AND ABBREVIATIONS ‘I’m — inner mitochondrial membrane potential ACh — acetylcholine AMPA - alpha-aminO-3-hydroxy-5-methyl-5-isoxazole propionate AN OVA - analysis of variance Ara—C — cytosine-B—arabinofilranoside ATP -— adenosine 5'-triphosphate BAPTA - 1,2- bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid tetrakis (acetoxymethyl)ester BCh — carbamyl-B-methylcholine chloride (bethanechol) Calcein AM — calcein acetoxymethylester [Ca2+]c — extracellular Ca2+ concentration {Cay} — intracellular Ca” concentration Ca2+... — Ca2+ contained within the mitochondrial lumen CCh - carbamylcholine CCCP - carbonyl cyanide m-chlorophenylhydrazone CICR — Ca2+-induced Ca2+ release CMF-HBSS — CaZI-Mg2+ fiee Hank’s Buffered Saline Solution CN — calcineurin (protein phosphatase ZB) CNS - central nervous system CsA — cyclosporin A 4-DAMP - 4-dipheny1acetoxyl-N-methylpiperidine methiodide xvi DHE — dihydro-B—erythroidine hydrobromide DMEM — Dulbecco’s Modified Eagle Medium DNase I — deoxyribonuclease I EGTA — ethyleneglycol bis-(B-aminoethyl ether)-N,N,N',N',-tetraacetic acid EtOH -— ethanol FBS — fetal bovine serum F ura-2 AM - fura-2 acetoxymethylester GABA — gamma-aminobutyric acid GSH - glutathione HBS — HEPES-buffered saline solution HEPES — N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid hr — hour(s) 1P3 — inositOl-l,4,5-triphosphaste MeHg — methylmercury min — minute(s) MTP — mitochondrial permeability transition pore NMDA - N-methyl-D-aspartate PhAsO — phenylarsine oxide FLA; - phospholipase A2 PLC — phospholipase C PNS — peripheral nervous system ROS - reactive oxygen species SEM — standard error of the mean xvii SER — smooth endoplasmic reticulum SERCA — smooth endoplasmic reticulum Ca2+ ATPase SOCC — store—operated Ca2+ channel SOCE — store-operated Ca2+ entry TMRE — tetramethylrhodamine ethyl ester TPEN — N,N,N‘,N‘-tetrakis-(2-pyridy1methyl)ethylenediamine W7 — N-(6-aminohexyl)-5-chloro—1-napthalenesu1fonamide XeC - xestospongin C xviii CHAPTER ONE INTRODUCTION A) GENERAL INTRODUCTION Methylmercury (MeHg) is a highly lipophilic environmental contaminant which easily crosses both the blood-brain barrier and placental barrier, resulting in neurotoxicity in adults and fetuses alike (Friberg and Mottet, 1989; Aschner and Aschner, 1990; Myers and Davidson, 1998). Despite its ability to bind thiols indiscriminately with high affinity (Roberts et al., 1980), making any protein or peptide with a cysteine group a potential target, MeHg causes a surprisingly Specific pattern of neuronal damage in both the central and peripheral nervous systems (CNS and PNS, respectively). Over the past few decades, there has been considerable effort to elucidate the specific mechanism(s) by which MeHg causes neurotoxicity within discrete areas of the CNS by focusing on specific molecular targets of the compound. By identifying the molecular targets of MeHg within the CNS, particularly proximate targets in cells which exhibit heightened sensitivity, it is hoped that more reliable risk assessment can be performed in which results from in vitro assays can be correlated with in vivo effects observed at environmentally relevant exposure levels. The primary pathological result of both acute and chronic MeHg poisoning is neurotoxicity, with clinical signs ranging from ataxia and visual disturbances to paralysis and death (Rustam et al., 1975; Chang, 1980). Several areas of the brain exhibit marked pathological damage upon MeHg exposure, including the cerebral and cerebellar cortices (Hunter and Russell, 1954; Leyshon—Sorland et al. , 1994). In both humans and rats, the granule cell layer of the cerebellum is particularly sensitive to MeHg-induced cell death (Hunter and Russell, 1954) through an as yet undefined mechanism. In contrast, the neighboring Purkinje cells are less sensitive to MeHg-induced cell death despite accumulating equal or greater MeHg concentrations (Hunter and Russell, 1954). The mechanism underlying the heightened susceptibility of these neurons is not yet known, but is hypothesized to be due to interactions with specific molecular targets which have not yet been identified (Castoldi et al., 1996; Robertson and Orrenius, 2000; Castoldi et al. , 2001). Given the large number of potential molecular targets containing cysteine moieties, it is not surprising that MeHg has a number of effects which could contribute to the observed toxicity. At the cellular level, MeHg interacts with a number of Sites to cause disruption of divalent cation homeostasis, particularly Ca2+ and Zn2+ (Denny et a1. , 1993; Hare et al. , 1993), elevation of spontaneous release of neurotransmitter (Juang, 1976; Atchison and Narahashi, 1982; Atchison, 1986), generation of reactive oxygen species (Sarafian and Verity, 1991; Yee and Choi, 1994) and cell death (Kunimoto, 1994; Marty and Atchison, 1998). The disruption of intracellular Ca2+ (Ca2+i) has received much attention in recent years, with both plasma membrane Ca2+ channels and intracellular Ca2+ stores being investigated as potentially critical targets during MeHg exposure (Tan et al., 1993; Oyama et al., 1994; Sakamoto et al., 1996; Marty and Atchison, 1997; Marty and Atchison, 1998; Sirois and Atchison, 2000). Cell viability assays in which Ca2+; was buffered by the Ca2+i chelator 1,2- bis(2- aminophenoxy)ethane-N,N,N',N'—tetraacetic acid tetrakis (acetoxymethyl)ester (BAPTA) indicated that removal of Ca2+: provides significant protection against neuronal death at 3 hr post-exposure, suggesting an important role for loss of Ca2+ homeostasis in MeHg- induced neurotoxicity (Marty and Atchison, 1998). In the highly susceptible cerebellar granule cell, MeHg affects multiple subtypes of voltage-gated Ca2+ channels on the plasma membrane (Marty and Atchison, 1997; Sirois and Atchison, 2000). However, very little is known about the intracellular Ca2+ targets in this cell type. Previous studies demonstrated that MeHg disrupts intracellular Ca2+ (Ca2+i) homeostasis in cultures of rat cerebellar granule cells, resulting in a biphasic increase in [Ca2+]i: the first phase results from release of Ca2+ from intracellular pools, and the second phase results from influx of extracellular Ca2+ (C322) (Hare et al., 1993; Marty and Atchison, 1997) (Figure 1.1). The initial increase in [Caz+]t is intracellular in origin, as removal of Ca2+c abolishes the secondary fura-2 fluorescence elevation (Hare et al., 1993; Marty and Atchison, 1997) (Figure 1.2). In cerebellar granule cells, the MeHg- induced influx of Ca2+c occurs through nifedipine- and w-conotoxin—MVIIC-sensitive pathways, suggesting that L, N and/or P/Q-type voltage-gated Ca2+ channels participate in MeHg-induced Ca2+ entry; however, the intracellular Ca2+ source has not yet been identified (Marty and Atchison, 1997). Two intracellular Ca2+ stores could contribute to the early-onset elevation of [Cay], during MeHg exposure: the smooth endoplasmic reticulum (SER) and the mitochondria. The SER provides high-affinity, low-capacity Ca2+ storage while the mitochondria are generally considered to be a low-affinity, high- capacity pool (Fohrman et al., 1993; Gunter and Gunter, 1994; Simpson et al., 1995; Budd and Nicholls, 1996b; Masgrau et al. , 2000). The SER contains ryanodine- and IP3- sensitive Ca2+ pools which can release sequestered Ca2+ into the cytosol (Irving et al., 1992). In NGlO8—15 cells, part of the MeHg-induced first-phase is due to release of Ca2+ from the IP3-sensitive pool in the SER; however, emptying 0.5 1.1M MeHg 3 ’ ' : E First ; Second :Phase : Phase x . + ---- p-----—-- l l l i l 0 l l n l l. l i 0 10 20 30 40 50 60 Time (min) Figure 1.1. MeHg causes a biphasic increase in [Ca2+]i in cerebellar granule cells in primary culture. Representative tracing of the fura-2 fluorescence fiom a cerebellar granule cell exposed to 40 mM K.+ for 1 min (as a viability test), followed by HBS washout and 0.5 11M MeHg. The ratio of fluorescence emitted at excitation wavelengths of 340 nm and 380 nm gives the approximate level of [Ca2+]i, which increases in a biphasic manner. Time-tO-onset of fluorescence increases is measured from the beginning of MeHg treatment. Modified from: Marty and Atchison, Toxicol. Appl. Pharmacol. 147 :319-3 30 (1997). 0.5 uM MeHg-EGTA First Phase l-I-c- -------------- i l l l l 40 50 60 O y—a O N O DJ 0 Time (min) Figure 1.2. Elevations of fun-2 fluorescence in a cerebellar granule cell in culture exposed to 0.5 M MeHg in Ca2+-free buffer. Under these conditions, the first [Ca2+]: phase is evident while the second phase (due to influx of [Ca2+].,) is absent. Modified from: Marty and Atchison, Toxicol. Appl. Pharmacol. 147:319-330 (1997). this pool does not eliminate the first phase, suggesting that at least one additional source contributes to the first phase Ca2+ increase (Hare and Atchison, 1995). Mitochondria also store and release Ca2+, and are also affected by MeHg. Ca2+ release can occur via reversal of the uniporter, activation of the Na‘I/Ca2+ exchanger, or opening of the mitochondrial permeability transition pore (MTP). MeHg has a number of effects on mitochondria, including decreased Ca2+ uptake and increased Ca2+ efflux from preloaded rat brain mitochondria, as well as inhibition of respiration, increased permeability of the inner membrane, and depolarization of the inner mitochondrial membrane (Sone et al. , 1977; Levesque and Atchison, 1991; Hare and Atchison, 1992), any of which could contribute to the first phase Ca2+ elevation via disruption of mitochondrial Ca2+ buffering. Several lines of evidence suggest a critical relationship between intracellular Ca2+ pools and cell death, thus necessitating inquiry in this area to gain filll understanding of neuronal death induced by MeHg. AS such, the goal of this dissertation is to identify the intracellular Ca2+ pools targeted by MeHg in cerebellar granule neurons and determine their relative importance in MeHg-induced cytotoxicity of cerebellar granule neurons. This dissertation reports results of experiments designed to investigate the hypothesis that MeHg initiates Ca2+ release fi'om the SER which is buffered by mitochondria; however, MeHg impairs the ability of mitochondria to hold Ca2+ (by causing loss of mitochondrial membrane potential and inhibiting respiration), leading to opening of the mitochondrial permeability transition pore and cell death (Figure 1.3). MeHg Ryanodine Receptors Figure 1.3. Hypothesis regarding MeHg-induced elevations of [Ca2+]i in rat cerebellar granule neurons. MeHg binds M3 muscarinic ACh receptors, causing production of 1P3 and release of CaZ+ through the 1P3 receptors. This elevation of [Ca2+], triggers Ca2+-induced Ca2+ release through the ryanodine receptors. Ca2+ released from the SER is buffered by mitochondria. Excess uptake of Ca2+ into the mitochondria, combined with direct effects of MeHg on the mitochondrial respiratory chain, leads to opening of the mitochondrial permeability transition pore (MTP) and promotion of cell death. B) BACKGROUND a) Humans and MeHg: Routes and Biological Consequences of Exposure MeHg is an organic form of mercury which enters the environment through two primary routes: release of MeHg itselfas a by-product of human activities, and conversion of inorganic mercury to MeHg by biological organisms. Organic mercury has been used as a fungicide and in preservative solutions for wood, paper pulp and leather. Additionally, inorganic mercury has been used as a catalyst in several industrial processes, allowing for the release of both inorganic and organic mercury (as a by-product of the reaction) into the environment through discharge of untreated waste into the water and air (Kurland, 1973; Gerstner and Huff, 1977). Organic mercury is also produced by methylation of mercury by aquatic microorganisms in the sediment of river and lake beds (Choi and Bartha, 1994; Baldi, 1997; Guimaraes et al., 2000). MeHg is a highly lipophilic substance which, once in the environment, readily accumulates in the food chain (Rabenstein and Evans, 1978). The most common mode of MeHg intoxication for humans is through ingestion of contaminated foodstuffs (Hansen, 1990; Renzoni et al., 1998). Once ingested, MeHg readily enters the blood and is transported in the plasma throughout the body, where it easily crosses the blood-brain barrier (Steinwall and Klatzo, 1966; Chang and Hartrnann, 1972; Chang and Hartmann, 1972; Aschner and Aschner, 1990). In rats exposed to radioactively labeled MeHg, the greatest accumulation of the compound occurs in the spinal dorsal root ganglia, with the cerebral cortex and cerebellum also exhibiting high concentrations following exposure (Somjen et al. , 1973). MeHg is excreted from the body at a slow rate, with a half-life of approximately 70 days in humans (Clarkson, 1972; Neathery and Miller, 1975). While the rate of MeHg elimination via metabolism and excretion is sufficient that accumulation within the body is limited (Clarkson, 1972), there is evidence suggesting that the brain half-life is considerably longer than that of the whole body during chronic exposure (Rice, 1989; Carrier et al. , 2001). MeHg poisoning was first examined in the 19508 following the study of four patients poisoned via inhalation of MeHg from fungicidal dusts in a cereal factory. The predominant symptoms in these patients were neurological, including generalized gross ataxia, impaired speech, and severely constricted visual fields. Histological analysis of one of the four patients in autopsy revealed gross atrophy of the cerebrum as well as marked degeneration of the granule cell layer of the cerebellum and neuronal destruction in the visual area of the calcarine (visual) cortex (Hunter and Russell, 1954). The authors noted that the observed ataxia was probably due to gross alterations in the cerebellar cortex because there was no damage to the cerebellar motor pathways other than the loss of granule cells. The cerebellar damage itself was very selective, with loss of granule cells occurring in the absence of damage to nearby Purkinje cells. The two most widely known incidents of MeHg poisoning occurred in Japan in the 1950s and 19608 (Takeuchi et al., 1962; Harada, 1995) and Iraq in 1972 (Bakir et al., 1973; Clarkson er al., 1976). There were epidemics in Japan in both Minamata (from 1953 to 1960) and Niigata (from 1960 to 1965) as a result of eating contaminated seafood (Chang et al., 1977). In Minamata, both organic and inorganic mercury were part of a chemical plant’s waste effluent discharged into Minamata Bay. Once in the waterway, MeHg accumulated in the food chain, resulting in 46 deaths and 10 approximately 2200 people poisoned through consumption of contaminated seafood (Harada, 1995). In addition, prenatal exposure to MeHg resulted in many infants being born with nervous system damage. This fetal neurological syndrome is now referred to as “Minamata syndrome.” During the Iraqi poisoning incident, approximately 6500 hospitalizations and 450 deaths were reported as a result of MeHg poisoning from ingestion of contaminated seed grain in bread (Clarkson et al., 1976). In both the Minimata and Iraqi incidents, the primary sign of exposure was neurological dysfunction as characterized by cerebellar ataxia, generalized extremity weakness and sensory disturbances including impairments of speech, vision and hearing. Correlations between body burden of MeHg and fiequency of adverse symptoms (ataxia, paresthesia, etc.) in adult victims of the Iraqi poisoning indicated that the threshold for adverse effects was at a body burden level of 25 to 40 mg of mercury (Bakir et al., 1973). Additionally, Iraqi children exposed in utero, through breastfeeding, or both, exhibited marked developmental deficiencies, including delayed onset of speech and impaired motor, sensory and autonomic function (Bakir et al. , 1980). In the more severe cases, children were born blind and deaf (Bakir et al., 1980). Dose-response analysis suggests that fetal effects are observed when maternal hair mercury concentrations exceed 10 ppm (Crump et al., 2000). In adults, MeHg causes a very specific pattern of damage within the PNS and CNS that correlates with the observed symptoms of poisoning (Chang and Hartrnann, 1972; Chang et al., 1977). In the PNS, there is axonal degeneration characterized by secondary loss of the myelin sheath of the sensory branch (Hunter and Russell, 1954; Takeuchi, 1982). The disruption of motor and sensory nerve function correlates with the 11 observed extremity weakness and impairment of sensory function (Chang, 1980). Early studies focused on effects of MeHg at the neuromuscular junction, in part due to the observed increased incidence of neuromuscular weakness during the Iraqi incident (Rustam et al., 1975) and in part because the ACh release process is well characterized both biochemically and physiologically. At the neuromuscular junction, the primary effect is to decrease nerve-evoked release of acetylcholine (ACh) and alter its spontaneous quantal release by causing it to first increase, then decrease (J uang, 1976; Atchison and Narahashi, 1982). Interestingly, the elevated spontaneous release of neurotransmitter may be due to increased [Ca2+], because MeHg increases [Cay], in the absence of Ca2+,- in isolated nerve terminals (Komulainen and Bondy, 1987). Further, MeHg increases the frequency of spontaneous ACh release in the absence of Ca2+c in the neuromuscular junction (Atchison, 1986). While MeHg can damage the PNS, its effects within the CNS are of greater concern, especially in the developing organism. During chronic MeHg exposure, MeHg accumulates in the cerebellum, with the highest concentration in the Purkinje and Golgi cells of the granular layer, and a Slightly lower concentration in the granule cells, stellate cells, and basket cells (Chang et al., 1977; Moller-Madsen, 1990; Moller-Madsen, 1991) (Leyshon-Sorland et al. , 1994) (Figure 1.4). However, despite the lower accumulation of MeHg in the granule cell layer, these neurons are most affected during both acute and chronic MeHg poisoning, resulting in a characteristic loss of the granule cell layer (Hunter and Russell, 1954; Chang et al., 1977; Leyshon and Morgan, 1991). The resultant cerebellar ataxia observed in both acute and chronic MeHg poisoning is attributed to the loss of these neurons. Additionally, MeHg causes specific loss of 12 neurons from the second through fourth layers of the visual area of the calcarine cortex, resulting in constriction of the visual field (Hunter and Russell, 1954; Takeuchi, 1982). The fetal CNS exhibits greater sensitivity to MeHg than does the adult CNS, as children exposed in utero may be affected in the absence of maternal toxicity (Bakir et al., 1980; Takeuchi, 1982). Unlike the toxicity in adults, fetal poisoning produces a less specific pattern of damage that is dependent on the concentration and duration of exposure, as well as the stage of development during exposure (Rodier, 1995). Prenatal exposure interferes with neuron migration (Choi et al. , 1978) and decreases the mitotic activity of CNS neurons (Rodier et al. , 1984). Following the Iraqi incident, infants born to highly exposed mothers demonstrated permanent damage to the CNS, with symptoms including delayed achievement of developmental milestones, seizures and abnormal reflexes (Marsh et al., 1987). In several cases involving Japanese children, there was a marked decrease in the number of neurons in all lobes of the cereme cortex, as well as widespread disorganization of the cerebral cortex architecture resulting from neuronal degeneration (Choi, 1989). Similar to the adult symptoms, severely affected children exhibit severe atrophy of the cerebellar cortex such that the granule cell layer degenerates and virtually disappears (Choi, 1989). Unlike the adult cases, in which the remaining neurons of the cerebellar cortex are relatively unaffected, children exhibit a narrowing of the cortical molecular layer and a reduction in the number of Purkinje neurons (Choi, 1989). More recent studies have focused on the more subtle neurological effects resulting fiom low-concentration exposure to MeHg in the developing organism. Currently, the most common source of MeHg exposure is consumption of contaminated seafood 13 J ‘0 [ 'Efi‘lb‘ui" ‘F‘zi'. €11.75; "if‘fifil‘? \ .\ Q's ' 80, BC / 2 CF Seams: Figure 1.4. Schematic diagram of cerebellar cortex. The cerebellar cortex is divided into three distinct layers which lie outside the white matter (WM); these layers are the outermost molecular layer (ML), the Purkinje layer (PL), and the innermost granular layer (GL). The granular layer contains not only the cell bodies of granule cells (Gr) but also contains Golgi cell (Gol) bodies, axons and dendrites, mossy fiber (MF) terminals, and the axons of Purkinje cells with their accompanying climbing fibers (CF). The Purkinje layer contains only the cell bodies of Purkinje cells (PC). The molecular layer consists of the stellate cells (SC) and basket cells (BC), as well as the dendritic arbors of Purkinje cells and Golgi cells, and the parallel fiber (PF) branches of granule cell axons. Diagram courtesy of Dr. Yukun Yuan. l4 (Hansen, 1990; Renzoni et al., 1998); thus, MeHg exposure is of concern in regions of the world where indigenous populations heavily consume fish as part of their diet. At- risk populations include those in the Seychelles (Myers et al., 1995; Myers et al., 1997; Palumbo et al., 2000) and Faroe Islands (Weihe et al., 1996; Grandjean et al., 1999), as well as Greenland’s Inuits (Hansen, 1990) and certain tribes of North American Native Americans (Dellinger et al., 1996). While there were no adverse effects observed in a longitudinal study of children in the Seychelles Islands (Davidson et al., 2000), studies in the Faroe Islands (Grandjean et al., 1999; Steuerwald et al., 2000) and Brazil (Grandjean et al., 1999) found that long-term exposure to low concentrations of MeHg was correlated with deficiencies in motor fimction, attention, memory, and language development. The growing recognition that the greatest threat is to the developing organism is reflected in the recent lowering of the US EPA reference dose fiom 0.3 to 0.1 rig/kg per day (Rice et al. , 2000). The lower reference dose is derived from a study of children exposed in utero during the Iraqi poisoning, in which increased maternal MeHg exposure was associated with delayed neurodevelopment in the child (Marsh et al. , 1987). However, there is still considerable debate over the lowest exposure needed to produce adverse effects in the fetus. b) Cellular and Molecular Effects of Methylmercury One of the issues complicating the risk assessment of MeHg is the large number of potential molecular targets within the body. Because MeHg binds thiol groups with high affinity, any protein or peptide containing a cysteine is a potential target. Given the wide range of possible targets during MeHg exposure, the question arises whether there is 15 a Specific target, or set of targets, which make specific areas of the CNS vulnerable to MeHg-induced disruption of neuronal development and viability. This dissertation will address the question of whether MeHg targets specific receptors and ion channels involved in Ca2+ signaling in the CNS, including plasma membrane Cholinergic receptors and intracellular Ca2+ channels associated with Cay-sequestering organelles, which may be involved in the pathogenesis of MeHg. i) MeHg-Induced Cell Death: Apoptosis or necrosis? To understand whether MeHg has specific targets during neurotoxicity, it is critical to understand the mechanisms by which MeHg causes cell death. The cell death process can be divided into two morphologically and biochemically distinguishable categories: necrosis and apoptosis. Necrosis is generally characterized by cell swelling, organelle damage, loss of plasma membrane integrity and cell lysis, as well as rapid adenosine 5'-t1iphophate (ATP) depletion and severe alteration of mitochondrial ultrastructure, such that an inflammatory response is initiated. Apoptosis, on the other hand, is characterized by cell shrinkage, chromatin aggregation, plasma membrane blebbing, and DNA condensation accompanied by maintenance of ATP levels, activation of caspases, and loss of plasma membrane asymmetry (Kroemer, 1997; Mattson, 2000). However, the distinction between apoptosis and necrosis is not clear-cut, as both share several critical characteristics. For example, both modes of cell death are accompanied by elevated [Ca2+]; levels, and each can result fiom exposure to the same compound depending on the severity of the insult (Kruman and Mattson, 1999). Removal of Ca2+: with buffering agents (such as BAPTA) can protect cells from cell death in several 16 different systems, including neurons (Marty and Atchison, 1998). Both mechanisms also involve plasma membrane blebbing and mitochondrial damage, making the distinction between necrosis and apoptosis diffith in some cases. There is also evidence that the antiapoptotic gene bcl-2 can protect cells from both apoptosis and necrosis (Kane et al. , 1995; Kroemer, 1997). Indeed, although necrosis and apoptosis are considered to be different forms of cell death, there is now debate over whether they are actually part of a continuum of overlapping mechanisms of cell death (Rafliay and Cohen, 1997). In MeHg neurotoxicity, there is increasing evidence both in vivo and in vitro that MeHg causes apoptosis at lower concentrations and necrosis at higher concentrations (Nakada and Imura, 1983; Miura and Imura, 1987; Kunimoto, 1994; Nagashima et al., 1996). When studied in vivo, the cerebellar damage tends to be apoptotic, as assessed using conventional techniques such as the terminal deoxynucleotidyl transferase- mediated dUTP nick end labeling (TUNEL) assay, DNA gel electrophoresis and ultrastructural examination (Nagashima et al., 1996). When studied in vitro, cerebellar granule cells undergo apoptosis following 18 hr of exposure to lower MeHg concentrations (0.5-1.0 11M), and necrosis within 1 hr of exposure to higher concentrations (5 and 10 11M) (Castoldi et al., 2000). When granule cells in primary culture are studied after 72 hr of exposure, apoptosis occurs at 0.1-0.3 11M, with the switch to necrosis occurring at higher MeHg concentrations (Kunimoto, 1994). Hence, the mode of cell death in vitro depends on both the duration and intensity of exposure. However, there is no clear-cut threshold for which death pathway is followed, as both modes of cell death can be observed in granule cells in primary culture in response to 1 l7 11M MeHg within an hour of exposure, as determined using a combination of fluorescent microscopy and morphometric analysis, and the TUNEL assay (Castoldi er al., 2000). ii) Calcium and Cell Death The central role of Ca2+ in the cell death process is a common thread between a wide range of neurotoxic insults, including the neurotoxicity of lead and organic mercury, as well as glutamate excitotoxicity (White and Reynolds, 1996). Neurons contain a variety of mechanisms for transporting Ca2+ across both internal and external membranes. The concentration of Ca2+: is tightly regulated to maintain a relatively low [Ca2+]; (in the 100-200 nm range) compared to the [Ca2+]c (in the 1—2 mM range) (Kass and Orrenius, 1999). When the [Ca2+]; becomes elevated for sustained periods of time, the physiological signal becomes pathological as Ca2+ begins to promote cell death. Disruption of Ca2+; homeostasis results in activation of multiple enzymes which are involved in the catabolism of proteins, phospholipids and nucleic acids, thereby leading to degradation of molecules needed for cell survival (Kass and Orrenius, 1999). For example, calpains are Ca2+-activated cytosolic proteases which, when activated, are capable of promoting cell death, as indicated by the effectiveness of calpain inhibitors in preventing Ca2+-mediated cell death (Chan and Mattson, 1999). Several phosholipases, including PLC and phospholipase A2 (PLAZ) are also activated by Ca2+ either directly (PLC) or indirectly through CaZI-calmodulin (PLAz). PLC generates 1P3 which releases Ca2+; from the SER, while PLA; breaks down phospholipids to form arachidonic acid (Exton, 1990). Sustained elevations of Ca2+, can lead to sustained activation of PLA;, thereby leading to extensive membrane degradation and the generation of toxic 18 metabolites (Exton, 1990). Further, there are Ca2+-activated endonucleases which mediate DNA strand cleavage at internucelosomal linker regions to produce approximately 200 base pair long fragments (Arends et al., 1990). Although influx of Ca2+c is often associated with apoptosis, it is not necessarily an essential requirement as elevated Ca2+ originating from intracellular stores can be sufficient to induce cell death (Wei and Perry, 1996; Wei et al., 1998; Szalai et al., 1999). However, there is also evidence for a lack of correlation between changes in [Cam]: and apoptosis caused by specific insults, most notably in thymocytes and lymphocytes (Alnemri and Litwack, 1990; Beaver and Waring, 1994). MeHg causes elevated [Ca2+]; in a number of cell types and preparations, including neuroblastoma-glioma cells (Hare et al., 1993), human lymphocytes and monocytes (Shenker et al. , 1992; Shenker et al., 1993; Tan et al. , 1993), mouse thymocytes (Oyama et al., 1995), rat forebrain synaptosomes (Denny et al., 1993) and rat cerebellar neurons (Sarafian, 1993; Oyama et al. , 1994; Marty and Atchison, 1997; Mundy and Freudenrich, 2000). In cerebellar granule cells in primary culture, inhibition of MeHg-induced Ca2+c influx through voltage-operated Ca2+ channels with the Ca2+ channel blockers co-conotoxin MVIIC and nifedipine reduced the number of dead neurons at 3.5 hr post-exposure (Marty and Atchison, 1998). Additionally, chelating Ca2+i with BAPT A protected against cell death at 3.5 hr post—exposure to 0.5-1.0 11M MeHg; however, BAPT A was not protective at 24.5 hr post-exposure (Marty and Atchison, 1998). In granule cells in vitro, MeHg-induced cell death is not due to excitotoxic insult, as inhibitors of excitatory amino acid channels do not prevent MeHg- induced elevations of [Ca2+]: (Marty and Atchison, 1997) or MeHg-induced cell death 19 (Marty and Atchison, 1998; Castoldi et al., 2000). In vivo studies support the in vitro experiments, as voltage operated Ca2+ channel blockers prevented the appearance of neurological disorders and mortality in rats treated with MeHg (Sakamoto et al., 1996). Thus, MeHg neurotoxicity has a Ca2+-dependent component which, when inhibited, prevents some of the toxic effects of the compounds, including cell death. MeHg potentially causes elevations of [Ca2+]; through non-specific damage to lipid membranes on intracellular organelles and/or the plasma membrane, or through specific interactions with receptors and/or ion channels associated with Ca2+ signaling. MeHg causes non-specific membrane damage through lipoperoxidation in cerebellar granule cells in suspension (Sarafian and Verity, 1991); however, this membrane damage does not appear to contribute to MeHg-induced cell death, as prevention of lipoperoxidation is not neuroprotective (Sarafian and Verity, 1991). Alternatively, there is substantial evidence that interactions between MeHg and specific Ca2+ signaling receptors and ion channels may be important in MeHg-induced neurotoxicity, as outlined in the following sections. iii) MeHg and Cholinergic Neurotransmission MeHg alters Cholinergic transmission both pre- and post-synaptically within the CNS and PNS alike. One of the early experimental observations regarding MeHg neurotoxicity was its disruption of neurotransmitter release at the neuromuscular junction, which was studied in part because of the observed increased incidence of neuromuscular weakness during the Iraqi incident (Rustam et al., 1975) and in part because the ACh release process is well characterized both biochemically and 20 physiologically. At the neuromuscular junction, the primary presynaptic effect is to decrease nerve-evoked release of ACh and alter its spontaneous quantal release in a mechanism apparently dependent on disruption of [Ca2+], (Juang, 1976; Atchison and Narahashi, 1982). However, it is the post-synaptic effects of MeHg on Cholinergic transmission which are of interest in the study of MeHg-induced neuronal death in the CNS. MeHg binds to both nicotinic and muscarinic Cholinergic receptors in a variety of neuronal preparations. In Torpedo, MeHg inhibits binding of ACh to the muscle-type nicotinic ACh receptors (Sharnoo et al., 1976; Eldefrawi et al., 1977). In vitro, 10-30 11M MeHg inhibits the fast depolarizing response of nicotinic receptors in NIB-115 neuroblastoma cells (Quandt et al., 1982). Both muscarinic and nicotinic receptors contain a conserved pair of extracellular cysteine residues that are predicted to from a disulfide bond between two of the extracellular loops of the receptor (Sato et al., 1976; long et al., 1999). These thiol groups are critical in agonist and antagonist binding by muscarinic receptors (Hedlund and Bartfai, 1979), and agents which modify these thiol groups alter muscarinic receptor activity (Sato et al., 1976; Berstein et al., 1988). Given that MeHg binds thiol groups with high affinity, it is not surprising that MeHg interacts with several subtypes of muscarinic receptors both in vivo and in vitro. MeHg inhibits agonist binding to both M1 and M2 muscarinic receptors in rat brain cortical membranes, although its affinity for M1 receptors is much higher than for M2 receptors (IC50 values of 3.4 11M and 149 11M for M1 and M2, respectively) (Castoldi et al., 1996). In this preparation, MeHg inhibited agonist binding to the receptors in a competitive manner and decreased the affinity of the agonist carbachol for the receptor, suggesting that MeHg 21 prevents agonist binding to the receptors, thereby disrupting neurotransmission post- synaptically (Castoldi et al., 1996). MeHg also affects muscarinic receptors in vivo. In rats treated with 0.5 or 2 mg/kg/day MeHg for 16 days, there was a Significant increase in the density of muscarinic receptors in the cerebellum and hippocampus with no effect on receptor affinity (Coccini et al., 2000). This effect was region-specific, as there was no observable change in muscarinic receptor density in the cerebral cortex. Further, the increased receptor density was not observed until 2 weeks after the cessation of treatment, indicating a delayed up-regulation in the number of muscarinic receptors in these brain regions (Coccini et al. , 2000). The specific location of muscarinic receptors within the CNS may explain, in part, the specific pattern of neurotoxicity observed following MeHg exposure. For example, cerebellar granule cells express a higher density of muscarinic receptors than do any other neurons within the cerebellar cortex as assessed by [3H]]quinuclidinyl benzilate binding (Neustadt et al., 1988). The up-regulation of muscarinic receptors in the cerebellum and hippocampus is of particular interest, as MeHg damages cells and disrupts neurotransmission in these two regions (Yuan and Atchison, 1993; Yuan and Atchison, 1994; Yuan and Atchison, 1995; Yuan and Atchison, 1997; Coccini et al., 2000; Gasso et al., 2000). In cerebellar granule cells, muscarinic receptors modulate cell death induced by several toxicants, ranging from non-depolarizing growth medium to the wasp-venom peptide mastoporan (Galli et al. , 1995; Yan et al., 1995; de Luca et al. , 1996; Lin et al., 1997). Apoptosis caused by mastoparan is mediated by the M3 muscarinic receptor and not the M2 or nicotinic receptor, suggesting that activation of a 22 Specific, PLC-linked muscarinic receptor can be important in toxicant-induced death of cerebellar granule cells (Yan et al. , 1995). iv) Intracellular Calcium Stores Given the importance of maintaining low cytosolic levels of [Ca2+], and the role of Ca2+ as an intracellular signal , it is not surprising that in addition to mechanisms for Ca2+ removal fi'om the cell, there are also mechanisms for the intracellular storage of Ca2+. Within neurons, the two primary organelles which store Caz+ are the mitochondria and the SER. Mitochondria represent a low-affinity, high-capacity Ca2+ store, whereas the SER is a high-affinity, low-capacity Ca2+ pool (Somlyo et al., 1985). While the SER will move Ca2+ fi'om the cytosol into the SER lumen at low cytosolic [Ca2+]i, the mitochondria require a more powerful stimulus, with a local extra-mitochondrial Ca2+ concentration of approximately 0.5 11M needed for neuronal mitochondria (N icholls, 1978; Nicholls and Scott, 1980). Disruption of Ca2+ regulation in either of these stores can pose a serious challenge to neuronal survival. As discussed below, mitochondrial effects are crucial in Ca2+-mediated cell death, particularly apoptosis. In fact, there is substantial evidence to support the hypothesis that mitochondrial Ca2+ uptake is crucial for both apoptotic and necrotic neuronal death (lchas and Mazat, 1998; Lemasters et al. , 1998; Andreyev and Fiskum, 1999; Kruman and Mattson, 1999; Duchen, 2000). Additionally, disruption of SER Ca2+ content can be lethal to neurons (Kass and Orrenius, 1999; Mattson et al., 2000), as will also be discussed below. Mitochondria accumulate Ca2+ based on the extremely negative membrane potential within the inner mitochondrial membrane (140-180 mV, negative inside) 23 (Gunter and Gunter, 1994), making Ca2+ buffering an energetically favorable process. Uptake occurs via the ruthenium red-sensitive uniporter as well as by the so-called rapid uptake mode (Gunter and Gunter, 1994; Sparagna et al., 1995; Buntinas et al., 2001). However, Ca2+ buffering comes at a price: uptake of a large amount of Ca2+ occurs at the expense of the proton motive force, and can thus cause uncoupling of ATP synthesis from respiration. Under physiological conditions, mitochondrial Ca2+ buffering is important in regulating cytosolic [Cap] elevations which result from either influx of Ca2+, or release of Ca2+ fiom the SER, and may play a role in modulating cellular Ca2+ oscillations (Simpson and Russell, 1996; Rizzuto et al. , 1998). At low concentrations, mitochondrial matrix Ca2+ (Ca2+m) can activate several mitochondrial enzymes associated with the tricarboxylic acid (TCA) cycle, as well as stimulate the adenine nucleotide translocase to cause increased movement of ATP from the mitochondria into the cell. Further, mitochondrial buffering of physiological Ca2+ transients results in parallel depolarization of the inner membrane, regardless of whether the Ca2+ source is intracellular or extracellular (Loew et al., 1994). However, high [Ca2+]m can become hazardous to the cell via inhibition of the TCA cycle, the electron transport chain, the mitochondrial ATPase, and dissipation of the proton gradient, all of which can contribute to loss of mitochondrial ATP synthesis and cell death (Simpson and Russell, 1998b). Mitochondria are capable of releasing matrix Ca2+ by three different mechanisms (see Gunter and Pfeiffer, 1990, Gunter and Gunter, 1994, and Bemardi, 1996, for reviews of these mechanisms). The first is reversal of the uniporter, which normally extrudes cytosolic Ca2+ into the mitochondrial matrix. The second is a NaI-Ca2+ exchanger, which is the primary mechanism for mitochondrial Ca2+ efflux in mitochondria isolated fiom 24 brain (Simpson and Russell, 1998b). The third is opening of the mitochondrial permeability transition pore (MTP). The MTP is a megapore formed on the inner mitochondrial membrane in response to adverse conditions including oxidative stress, high mitochondrial matrix Ca2+ and anoxia (Dubinsky and Rothman, 1991; Petronilli et al., 1994; Chernyak et al., 1995; Byrne et al., 1999). When open, the MTP allows the passage of any molecule less than 1500 Da, including Ca2+, Zn2+ and the pro-apoptotic cytochrome c, across the usually impermeable inner mitochondrial membrane (Szabo and Zoratti, 1992; Chemyak and Bemardi, 1996; Pastorino et al., 1999; Szalai et al., 1999; He et al., 2000; Petronilli et al., 2001). Because opening of the MTP can dissipate the proton gradient within the inner membrane, a corresponding loss of 111m (depolarization) is often observed and can be used as a diagnostic tool for observing a permeability transition (White and Reynolds, 1996; Trost and Lemasters, 1997). Induction of the MTP has been correlated with conditions such as ischemia, stroke and apoptosis, although its precise physiological role is not yet understood (see Bemardi and Petronilli, 1996 for review). Sustained opening of the MTP has also been linked to cell death caused by a variety of compounds and conditions, including glutamate excitotoxicity in cerebellar granule cells (N ieminen et al., 1996; Schinder et al., 1996; White and Reynolds, 1996). Opening of the MTP can be inhibited by the immunosuppressant cyclosporin A (CSA), which acts to keep the pore closed or delay its opening in the presence of inducing agents (Broekemeier et a1. , 1989; Bemardi et al., 1993; Nienrinen et al., 1996). Because opening of the MTP can release significant amounts of Ca2+ into the cytosol, MTP induction is a possible mechanism by which MeHg causes elevated [Ca2+]i during the early stages of MeHg exposure. 25 In addition to mitochondrial Ca2+, neurons can sequester Ca2+ within the SER. The SER contains ryanodine and IP3-sensitive Ca2+ pools which are thought to be distinct in cerebellar granule cells (Irving et al., 1992). The ryanodine receptors are tetrarneric complexes formed from 550-565 kDa monomers, forming one of the largest proteins present within cells (Coronado et al. , 1994). Ryanodine receptors are characterized by the ability of the plant alkaloid ryanodine to bind with nanomolar affinity, causing Ca2+ release at low concentrations and inhibition of Ca2+ release at higher concentrations. Ca2+ release fiom the SER ryanodine-sensitive Ca2+ pool is elicited by a variety of agonists, including caffeine, low concentrations of ryanodine, Ca2+, and cyclic ADP- ribose, and can be blocked by antagonists such as high concentrations of ryanodine, dantrolene, and ruthenium red (Coronado et al., 1994). While originally identified in skeletal and cardiac muscle as a Ca2+ release mechanism during depolarization-mediated contraction, ryanodine receptors have now been identified in many tissues including neurons. Ryanodine receptors are involved in the process of Ca2+-induced Cay-release (CICR), in which cytosolic Ca2+ activates ryanodine receptors to cause Ca2+ release fiom the SER CICR has been identified in several neuronal preparations as participating in the Observed increase in Ca2+; following application of a depolarizing stimulus, and in cerebellar granule cells has been shown to be a major component of total Ca2+ signal detected following entry of Ca2+c (Simpson et al., 1993). Immunocytochemistry in rat cerebellar granule cells indicates the presence of Type II ryanodine receptors (although at very low levels), whereas neighboring Purkinje neurons contain Type I receptors (Simpson et al., 1996). Cerebellar granule cells grown in culture possess a ryanodine- sensitive SER Ca2+ pool; however, studies in this system are limited in that after 4 days in 26 culture the ryanodine receptor agonist caffeine no longer produces Ca2+ release from the SER (Irving et al., 1992). The SER IP3-releasable Ca2+ pool also plays a role in elevations of Ca2+; in a variety of cells, including neurons. The homotetrameric 1P3 receptors, while related structurally to ryanodine receptors, differ in that their subunits are smaller (305—3 13 kDa per monomer) and have a 4-10 times smaller conductance than ryanodine receptors (Marks, 1997). Of the 3 known receptor subtypes, both cerebellar granule and Purkinje neurons possess Type I 1P3 receptors, with particularly high levels in cerebellar Purkinje cells (Sharp et al., 1999). While 1P3 is the primary agonist at these receptors, their function is also antagonized by heparin and decavanadate. Activation of phosphoinositide-specific phospholipase C (PLC) causes the hydrolyis of phosphatidylinositol 4,5-bisphosphate to yield inositol 1,4,5-triphosphate (1P3) and 1,2- diacylglycerol. The IP3 then acts at 1P3 receptors on the surface of the SER membrane to cause release of Ca2+ fiom the SER These receptors are also sensitive to local cytosolic Ca2+ levels in a bell-shaped fashion, with activation at low [Ca2+], inactivation at higher [Ca2+], and peak activity at approximately 300 nM (Bezprozvanny et al., 1991). Their activity is complicated further by their interaction with nearby ryanodine receptors such that release of Ca2+ fi'om the 1P3 pool can elicit CICR from ryanodine receptors, thus magnifying the Ca2+ Signal (Irving et al., 1992). Both the 1P3 and ryanodine-sensitive Ca2+ pools are filled by SER Ca2+-ATPases (SERCAS) which allow for rapid uptake of cytoplasmic Ca2+. Although there are 5 subtypes known to date, they all possess varying degrees of sensitivity to the inhibitors thapsigargin and 2,5-Di(tert-butyl)-l ,4-benzohydroquinone (BHQ). Thapsigargin causes 27 an elevation of [Ca2+]: by inhibiting the SERCA, which leads to emptying the SER Ca2+ stores (Thastrup et al., 1990). In cerebellar granule cells, there is some distinction between the SERCAS filling the 1P3 and ryanodine pools in that BHQ will affect the IP3 but not the ryanodine pool (Simpson et al. , 1996), suggesting that a separate filling mechanism exists for the 1P3 pool. Under physiological conditions, release of SER Ca2+ does not harm the cell; however, several lines of evidence suggest that prolonged loss of SER Ca2+ levels can have deleterious effects on normal cell function. For example, depletion of SER Ca2+ can disrupt normal protein synthesis pathways, as evidenced by the inhibition of general protein synthesis following rapid release of SER Ca2+ (Brostrom and Brostrom, 1990). Depletion of SER Ca2+ can also cause the release of a Specific deoxyribonuclease involved in DNA degradation, which could contribute to apoptotic cell death (Peitsch et al. , 1993). Further, depletion of SER Ca2+ content with thapsigargin has been shown to cause cell death in a variety of cell types, including neurons (Levick et al., 1995; Wei and Perry, 1996; Wei et al., 1998). It is thus not surprising that many cells possess a mechanism for refilling the SER Ca2+ stores via store-operated (or capacitative) Ca2+ entry (SOCE) (Putney, 1986). When the SER becomes depleted of Ca2+, there is a sensing mechanism in the SER that detects the Ca2+ decrease and, through an as yet unidentified signal, sends a signal to plasma membrane Ca2+ channels to activate Ca2+ influx to replenish the stores (Putney, 1986; Ma et al., 2000). The affected Ca2+ channels are neither the classic receptor-operated nor voltage-gated channels and are thus termed store-operated Ca2+ channels (SOCCs) (Kiselyov et al., 1998). Although this system was originally identified in non-excitable cells, it is also found in neuronal preparations, including cerebellar granule neurons (Chavis et al., 1995; Simpson et al., 1995). 28 Interestingly, recent evidence suggests that opening of SOCCS may be due to physical interactions between 1P3 receptors (Kiselyov et al., 1998; Ma et al., 2000) and/or ryanodine receptors (U sachev and Thayer, 1999; Kiselyov et al. , 2000) with the SOCCS themselves. In addition to interactions between the SER and the plasma membrane, there is also increasing evidence for interaction between SER Ca2+ pools and mitochondria. Release of Ca2+ fi'om the 1P3 pool is followed closely by increased mitochondrial matrix Ca2+ in response to the microdomain of high [Ca2+] generated near the mouth of the IP3 receptor channels (Miyata et al., 1991). Use of CaZI-Sensitive photoproteins simultaneously targeted to the mitochondrial matrix and SER indicated that in living HeLa cells there is a close relationship between release of Ca2+ fi'om the SER and increased matrix Ca2+ of mitochondria which were in close apposition to 1P3 receptors (Rizzuto et al., 1998). In fact, Ca2+ released from the IF; pool caused a greater increase in mitochondrial matrix [Cay] than in the bulk cytosol, whereas intracellular perfusion of a buffered Ca2+ solution increased [Ca2+], of the SER and mitochondria to the same extent, suggesting preferential Ca2+ uptake of 1P3 Ca2+ by mitochondria. It has been hypothesized that mitochondria located near 1P3 receptors may function to buffer released Ca2+ to prevent receptor inactivation by high [Ca2+], thereby potentiating 1P3 receptor activity (Rizzuto et al., 1998; Simpson and Russell, 1998a). v) MeHg and the Smooth Endoplasmic Reticulum In disrupting [Ca2+]; homeostasis, MeHg causes Ca2+ release from intracellular stores into the cytosol in cerebellar granule cells (Sarafian, 1993; Marty and Atchison, 29 1997), synaptosomes (Kauppinen et al., 1989; Denny et al., 1993) and NG108-15 cells (Hare er al., 1993; Hare and Atchison, 1995). In T lymphocytes (Tan et al., 1993) and NG108-15 cells (Hare and Atchison, 1995), most of this Ca2+ originates in the SER. Both organic and inorganic mercurials cause rapid release of Ca2+ fi'om muscle SER- derived vesicles (Abramson et al., 1983). The release of SER Ca2+ may result fiorn the observed inhibition of the SER CaZI-ATPase by MeHg, presumably resulting from its binding to critical thiols in the Ca2+ pump (Abramson et al., 1983; Chiu et al., 1983), or from the ability of MeHg to cause non-specific damage to lipid membranes through lipid peroxidation (Sarafian and Verity, 1991; Verity et al., 1994; InSug et al., 1997). However, the Ca2+ release may also occur by more specific mechanisms mediated by intracellular Ca2+ channels located on the SER membrane, such as the 1P3 and ryanodine receptors. There are three subtypes of 1P3 receptor (Types 1, II and III) which are differentially expressed throughout the CNS. In rat brain, Type I is the primary neuronal type, expressed in large quantities in the cerebellar Purkinje neurons as well as in the neighboring granule cells (which also express Type III receptors) (Sharp et al., 1999). Type 11 receptors are found only in glia, while Type III receptors are enriched in lirnbic and basal forebrain regions, particularly in the nerve terminals (Sharp et al., 1999). Additionally, there is a developmental time-source for the expression of receptor subtypes within the CNS, such that during embryonic development, only the Type III receptor is expressed in large quantities, while the Type I receptor is expressed highly only postnatally (Sharp et al., 1999). Expression of the Type III receptor is hypothesized to play a role in the normal onset of programmed cell death in the brain during 30 development (Sharp et al., 1999; Blackshaw et al., 2000). T cells that are deficient in 1P3 Type I receptors are resistant to multiple inducers of apoptosis, suggesting that these receptors may play a role in Ca2+-mediated cell death (Jayaraman and Marks, 1997; Marks, 1997). Type III IP3 receptor expression is augmented in lymphocytes undergoing cell death, and Type HI 1P3 receptors are expressed at high levels during developmental apoptosis in postnatal cerebellar granule cells, chick dorsal root ganglia and embryonic hair follicles (Blackshaw et al., 2000). Introduction of a Type III 1P3 receptor antisense oligonucleotide prevents cell death caused by removal of nerve growth factor in chick dorsal root ganglia, implicating the Type III receptor in neuronal death processes (Blackshaw et al., 2000). Thus, 1P3 receptors are involved not only in Ca2+: signaling but also in neuronal development and death. The differential expression of these receptors becomes important when addressing MeHg neurotoxicity, as the second messenger IP3 and its receptor have been implicated as playing a role in MeHg-induced alterations of [Ca2+]i. In rat cerebellar granule cells in culture, exposure to 5 11M MeHg for 30 min resulted in a twofold increase in cellular inositol levels which potentially contributes to the observed elevation of [Cay], observed in this system (Sarafian, 1993). Presumably, this increased [1P3] contributes to the release of intracellular Ca2+ that was observed within the same time frame (Sarafian, 1993). In rat cerebellar membrane preparations, MeHg stimulates binding of [3 H] 1,4,5- triphosphate and [3H]inositol l,3,4,5-tetrakisphosphate binding to 1P3 receptors (Chetty et al., 1996). Additionally, rats treated with 2.5 mg/kg body wt/day MeHg for 7 days exhibited an increase in binding of [3H]IP3 to cerebellar membranes; this increased binding was not observed following treatment with inorganic mercury or lead (Chetty et 31 al. , 1996). MeHg-induced alterations in 1P3 production are not limited to the nervous system but are also found on the immune system. In rat T lymphocytes, MeHg causes sustained elevations in [Ca2+]: which, like those observed in cerebellar granule cells, consist of an intracellular and extracellular component (Tan et al., 1993). The release of Ca2+, was largely blocked by the 1P3 receptor antagonist heparin (Tan et al., 1993). In NG108-15 cells, depletion of the IP3-sensitive Ca2+ pool reduced, but did not abolish, the MeHg-induced increase in [Ca2+]; (Hare and Atchison, 1995). Interestingly, in these cells, MeHg did not cause an increase in [1P3], suggesting that MeHg may interact directly with the receptor to cause Ca2+ release. Finally, in mouse lymphocytes, exposure to 1 11M MeHg for 1 hr resulted in a 50% reduction in the rate of inositol metabolism, suggesting that once [1P3] becomes elevated, MeHg maintains [1P3] at higher-than-normal levels (Nakatsuru et al., 1985). Thus, MeHg causes elevated [1P3] in a variety of neurons and other cells. Ca2+ release from the SER can also occur by opening of ryanodine receptor-linked channels localized to the SER membrane. Ryanodine receptors, like 1P3 receptors, are expressed as three different isoforms (1, II and HI) which are expressed differentially in the CNS. Ryanodine receptors are involved in the process of CICR, in which cytosolic Ca2+ activates ryanodine receptors to cause Ca2+ release fiom the SER (Irving et al., 1992), and have also been implicated as mediators of neuronal death (Marks, 1997). There is also limited evidence that MeHg targets ryanodine receptors to cause Ca2+ release fi'om the SER. Inhibition of Ca2+ release through ryanodine receptors causes minor reductions in the amplitude of the MeHg-induced [Ca2+], increase in NG108-15 32 cells (Hare and Atchison, 1995). However, the role of ryanodine receptors in MeHg- induced cell death has not yet been investigated. vi) MeHg and Mitochondria A number of studies indicate that MeHg targets several parameters of mitochondrial fimction, including respiration and oxidative phosphorylation. Additionally, MeHg has serious effects on the regulation of mitochondrial Ca2+. In mitochondria derived from rat forebrain, MeHg causes release of ”Ca2+ from preloaded mitochondria while preventing further Ca2+ uptake (Levesque and Atchison, 1991). MeHg-induced alterations in neurotransmitter release appear to result primarily from perturbation of mitochondrial Ca2+ regulation, with minimal or no contribution of Ca2+ from the SER in hippocampal slices (Gasso et al. , 2000) and the rat neuromuscular junction (Levesque and Atchison, 1988). More recent evidence suggests that MeHg causes opening of the MTP, which has been implicated in apoptotic cell death caused by a wide variety of toxic insults (Ichas and Mazat, 1998; Lemasters et al., 1998). Excess accumulation of Ca“ within the mitochondrial lumen, production of reactive oxygen species, and depletion of ATP are all activators of the MTP both in cells and in isolated mitochondria (see Bemardi et al., 1998, for review). In MeHg toxicity, there is substantial evidence for the induction of all three of these conditions; thus it is not surprising that in recent studies with human monocytes (InSug et al., 1997) and lymphoid cells (Shenker et al., 2000), there is evidence of MeHg-induced Opening of the MTP. Additionally, transfection of the murine GT1-7 hypothalamic neuronal cell line with the MTP inhibitor bcl-Z reduces cell death 33 caused by MeHg, suggesting an important role for the MTP in MeHg-induced neuronal death (Sarafian et al., 1994). MeHg is also widely recognized as causing oxidative stress. MeHg-induced elevations of reactive oxygen Species (ROS), such as superoxide anion, hydroxyl radicals and hydrogen peroxide, are observed in human monocytes (InSug et al. , 1997), human T cells (Shenker et al., 1998), and murine and rat brain (Ali et al., 1992; Yee and Choi, 1994). MeHg also causes lipid peroxidation in rat cerebellar granule cells in suspension (Sarafian and Verity, 1991). The antioxidant vitamin B provides some protection against neuronal damage in vivo (Chang et al., 1978; Usuki et al., 2001), suggesting that ROS contribute to cellular damage caused by MeHg. The ability of a cell to handle ROS may be correlated with its sensitivity to MeHg. MeHg-induced increases in fi'ee radicals occur in brain regions specifically sensitive to the effects of MeHg, including the cerebellum (LeBel et al., 1990; LeBel et al., 1992). Additionally, cells expressing increased glutathione (GSH) exhibit decreased sensitivity to MeHg-induced damage (Miura et al., 1994). However, there is considerable evidence that MeHg-induced oxidative stress may not be the determinant factor in MeHg-induced neuronal death in specific brain regions. Vitamin E prevents MeHg-induced increases in ROS and lipid peroxidation but does not protect against cell death in rat cerebellar granule cells (Sarafian and Verity, 1991). Additionally, modulation of intracellular GSH levels using N-acetylcysteine (to increase GSH) or L-buthioniner(S,R)-sulfoximine (BSO) (to decrease GSH) did not alter the neurotoxic effects of MeHg in primary cultures of rat embryonic CNS cells (Du et al., 1999). Taken together, these studies suggest that while 34 MeHg causes increased levels of ROS in both adult and developing brains, the resultant damage is not a determinant factor in the selective neurotoxicity of this compound. vii) MeHg-Induced Interaction Between SER and Mitochondria Ca2+ Stores There is growing evidence that interactions between SER and mitochondria are important in determining cell survival. In permeabilized RBL-2H3 cells (Csordas et al., 1999), permeabilized hepatocytes (Hajnoczky et al., 1999) and rat pulmonary smooth muscle cells (Drummond and Tuft, 1999), Ca2+ release from the 1P3 receptor results in a strong Ca2+ signal at the mitochondria. In HeLa cells transfected with specifically targeted, Ca2+ sensitive green fluorescent proteins, Ca2+ release fiom 1P3 receptors results in a higher localized concentration of Ca2+ at the outer surface of nearby mitochondria than is observed in the cytosol (Rizzuto et al., 1998), revealing the preferential Signal received by mitochondria from the SER. This Ca2+ signal between the SER and mitochondria can be translated into a pro-apoptotic Signal; in a human tumor cell line, activation of 1P3 receptors releases Ca2+ which causes opening of the MTP in mitochondria, with subsequent release of cytochome c and promotion of apoptotic cell death (Szalai et al., 1999). Mitochondria also communicate with the SER in a negative feedback system which alters SER Ca2+ uptake and release (Landolfi et al., 1998), indicating the complex nature of this Signaling pathway. Thus, any MeHg-induced changes in the [Ca2+] of one intracellular store must be considered in the context of its effects on the other store, and vice versa, in order to account for the complicated pattern of Ca2+ signaling which occurs within neurons. 35 viii) Non-Ca2+ Divalent Cations Elevation of intracellular concentrations of non-Ca2+ divalent cations is also observed during MeHg exposure in NG108-15 cells (Hare et al., 1993; Denny and Atchison, 1994), rat cerebellar granule cells (Marty and Atchison, 1997), and synaptosomes (Denny and Atchison, 1994). In synaptosomes, the non-Ca2+ cation was identified using 19F-NMR as Zn2+ (Denny and Atchison, 1994). The elevated [lay]; was possibly due to the displacement of endogenous Zn2+ by MeH g fiom the numerous proteins which bind Zn2+ with cysteine sulfhydryl groups (Hunt et al., 1984; Vallee and Auld, 1990). Zn2+ itself can be neurotoxic (Manev et al., 1997; Cookson and Shaw, 1999), and has recently been shown to induce opening of the MTP in cortical neurons (Sensi et al., 2000). In rats exposed to MeHg there was an elevated [ZnZI] in the brain which was not found in control or inorganic Hg-exposed animals (Muto et al., 1991). MeHg also caused elevated [Zn2+] in the kidney, another sensitive target organ during MeHg toxicity (Muto et al., 1991). In spite of the intriguing possibility that disruption of [Zn2+]i is a decisive factor in the selective toxicity of MeHg, several lines of evidence argue against this hypothesis. First, the non-Ca2+ component of the MeHg-induced increase in cerebellar granule cells is much smaller in amplitude than the Ca2+ component (Marty and Atchison, 1997). Second, chelation of non-Ca2+ divalent cations with the cell-permeant heavy metal chelator N,N,N',N'-tetrakis-(2-pyridylmethyl)ethylenediamine (TPEN) does not protect against cell death caused by low micromolar concentrations of MeHg at 3.5 hr post-exposure (Marty and Atchison, 1998). Finally, in cerebellar granule neurons in culture, Zn2+-mediated cell death was found to depend on protein synthesis (Manev et al., 1997). MeHg inhibits protein synthesis in cerebellar neurons (Sarafian et 36 al., 1984; Sarafian and Verity, 1985), cerebral cortex synaptosomes (Verity et al., 1977), and brain cortex Slices (Yoshino et al., 1966), suggesting that the mechanism by which Zn2+ exerts its toxic effects is non-functional during MeHg exposure. Thus, the contribution of elevated [Zn2+]i to MeHg-induced neurotoxicity is minimal at best. C) Rat Cerebellar Granule Neurons in Primary Culture as a Model for the Selective Neurotoxicity of MeHg As described above, MeHg causes preferential degeneration of the granule cell layer of the cerebellar cortex in humans and experimental animals alike (Hunter and Russell, 1954; Syversen et al., 1981; Leyshon and Morgan, 1991). While studies performed in vivo provide the closest parallel to human exposures, they do not provide enough information regarding the mechanisms underlying the selective neurotoxicity of MeHg. Additionally, experiments performed in vivo are complicated by the large number of interacting variables between different types of neurons within the cerebellar cortex. Thus, the experiments in this dissertation were performed in rat cerebellar granule cells in primary culture to allow for in-depth exploration of the actions of MeHg in these cells. Because cerebellar granule cells are the most populous neuronal type in the cerebellum, and are affected by several diseases and toxicants, they have been well characterized in vivo. When isolated fiom neonatal rats (approximately 6-7 days old), cerebellar granule cells grown in culture continue to develop and mature; they extend processes which bifurcate into fibers resembling the parallel fibers found in the cerebellar cortex, and make functional synapses with neighboring neurons. The presence and activity of many plasma membrane and intracellular ion channels, receptors and proteins have been well 37 characterized for the first 14 days post-isolation, with developmental profiles compiled for many cellular proteins (Gonzalez et al. , 1992; Aronica et al. , 1993; Aronica et al. , 1993; Aronica et al., 1993; Resink et al. , 1994; Xia et al. , 1995; de Luca et al., 1996). In the experiments performed for this dissertation, all cells were used at days 6-8 post- isolation to allow for the maturation of cells. At this time in culture, granule cells express L, N, P/Q and R types of voltage-gated Ca2+ channels; M2 and M3 muscarinic receptors; metabotropic glutamate receptors; excitatory amino acid (glutamate) receptors (N - methyl-D-aspartate (NMDA), alpha-amino-3 -hydroxy-5-methyl-5-isoxazole propionate (AMPA), and kainate); GABAA receptors; Type I 1P3 receptors; and Type II ryanodine receptors, among other receptors and channels (Meier et al. , 1984; Alonso et al., 1990; Pizzi et al., 1991; Aronica et al., 1993; Aronica et al., 1993; Randall and Tsien, 1995; Simpson et al., 1996; Randall and Tsien, 1997; Mhyre et al. , 2000; Mogensen and Jorgensen, 2000). Thus, the cells express many of the channels and receptors which characterize functionally mature cerebellar granule cells, making granule cells is primary culture a useful model for the study of MeHg neurotoxicity. The use of granule cells in culture is limited, however, by the lack of normal interactions with other neurons that are present in the intact animal. In particular, there is a relative lack of astrocytes in the culture, which comprise less than 10% of the cells in the culture (Marty and Atchison, 1997). Astrocytes have been intensively studied as a potential mediator of MeHg neurotoxicity (Aschner, 1996). Astrocytes concentrate MeHg and, when exposed to MeHg, release the neurotransmitter glutamate (Oyake et al. , 1966; Garman et al., 1975; Aschner et al., 1994; Aschner et al., 2000). Excess glutamate causes excitotoxic death in cerebellar granule cells by causing influx of Ca”, through 38 NMDA receptors (Ankarcrona et al. , 1995; Budd and Nicholls, 1996a; Nicholls et al., 1999). Thus, MeHg-induced release of glutamate from astrocytes may contribute to the toxicity of MeHg in vivo. In granule cells in primary culture, inhibition Of ion flux through NMDA, kainite and AMPA receptors does not alter MeHg-induced elevations of [Ca2 1 (Marty and Atchison, 1997); however, these cells are grown in the relative absence of astrocytes, which would be present in vivo. The lack of astrocytes in the culture is a disadvantage when comparing the results obtained in the culture system to those obtained in the whole animal. However, in examining the relative sensitivity of granule cells to MeHg-induced damage, the lack of astrocytes is advantageous, as it removes a potentially contributing factor to MeHg-induced neurotoxicity which is probably not unique to the cerebellar cortex. Additionally, granule cells in culture undergo cell death at much lower concentrations of MeHg than is observed in other less- sensitive cell types, suggesting that granule cells are inherently more sensitive to MeHg- induced cell death in the absence of any external contributing factors (Park et al. , 1996; Marty and Atchison, 1998). Thus, the use of granule cells in primary culture provides an opportunity to identify potential molecular targets which confer the unique sensitivity of this cell type to the effects of MeHg. 39 CHAPTER TWO INTERACTION OF METHYLMERCURY WITH MUSCARINIC RECEPTORS CAUSES CALCIUM RELEASE FROM THE SMOOTH ENDPLASMIC RETICULUM INOSITOL-l,4,5-TRIPHOSPHATE RECEPTORS IN RAT CEREBELLAR GRANULE NEURONS 40 ABSTRACT The contribution of Ca2+ release from the SER to MeHg-induced cation dysregulation was studied using Single-cell microfluorimetry of granule cells in vitro loaded with fura-2. Depletion Of SER Ca2+ with thapsigargin (10 uM) prior to MeHg (0.2-1.0 11M) exposure decreased the amplitude of MeHg-induced release of [Ca2+]; stores by approximately 30%, indicating that SER Ca2+ is released by MeHg. However, the time-to-onset of [Ca2+]; elevations by MeHg was unaffected by thapsigargin treatment. When the muscarinic receptors and inositol-1,4,5-triphosphate (1P3) receptors were down-regulated by 24 hr pretreatment with bethanechol (BCh, 1 mM), MeHg-induced release of Ca2+, was again reduced by 30-40%. BCh pretreatment significantly hastened the time-to-onset of the MeHg-induced release of Ca2+i without altering the influx of Caz} Atropine (10 11M) Significantly delayed the time-to-onset and reduced the amount of [Ca2+]; released into the cytosol by MeHg, suggesting that MeHg interacts with muscarinic receptors and produces 1P3 , which subsequently releases Ca2+ from the SER. Ryanodine (10 11M) did not affect the time-to-onset or amplitude of MeHg-induced increases in [Ca2+]i. Thus, MeHg interacts with muscarinic receptors with the end result of Ca2+ release from the SER, which contributes to the MeHg-induced loss of [Ca2+]; regulation. 41 INTRODUCTION As discussed in Chapter One, MeHg causes an early release of Ca2+ fiom at least one intracellular source in cerebellar granule cells in culture (Marty and Atchison, 1997). One possible source is the SER, which is a high-affinity, low-capacity Ca2+ pool filled by the SERCA (Irving et al. , 1992; Gunter and Gunter, 1994). Under normal conditions, Ca2+ release fiom the SER into the cytosol occurs via the [P3 receptor and/or the ryanodine receptor. In NG108-15 cells, most of the MeHg-induced first-phase is due to release of Ca2+ fiom the IP3-Sensitive pool in the SER, with some contribution from the ryanodine-sensitive pool; however, emptying this pool with the SERCA inhibitor thapsigargin does not eliminate the first phase, suggesting that at least one additional source contributes to the first phase Ca2+ increase (Hare and Atchison, 1995). In granule cells, the contribution of the SER to MeHg-induced elevations of [Ca2+]; is examined for the first time in the experiments presented in this chapter. At the SER, both organic and inorganic mercurials cause rapid release of Ca2+ from muscle SER-derived vesicles (Abramson et al., 1983). In granule cells in culture, MeHg causes an increase in [1P3] at low micromolar concentrations (Sarafian, 1993), although this increase in [1P3] is not observed in all neuronal models (Hare and Atchison, 1995) and has only been observed in granule cells after exposure to 5 uM MeHg for 30 min, which does not shed light on whether there is an early increase in [1P3] (Sarafian, 1993). The elevation of [1P3] could be due to binding of MeHg to muscarinic receptors, which has been Observed in several neuronal systems (Eldefiawi et al., 1977; Von Burg et al., 1980; Abd-Elfattah and Shamoo, 1981; Castoldi et al., 1996; Candura et al. , 1997). Granule cells express M2 and M3 muscarinic receptors, with the M3 receptors linked to 42 PLC (Fukamauchi et al., 1991; Fohrrnan et al. , 1993). MeHg is also reported to stimulate 1P3 binding to its receptor in rat cerebellar membrane preparations, suggesting that MeHg interacts with the 1P3 receptor itself (Chetty et al., 1996). The experimental approach taken to study the contribution of SER Ca2+ was similar to tint used by Hare and Atchison (1994) and Marty and Atchison (1997), in which different receptor agonists and antagonists were used to inhibit different portions of Ca2+ signaling pathways during MeHg exposure. First, the SERCA inhibitor thapsigargin was applied prior to MeHg application to deplete the SER of its Ca2+ content (Thastrup et al., 1990; Irving et al., 1992). Second, the muscarinic and 1P3 receptors were down-regulated prior to MeHg exposure using the muscarinic agonist bethanechol (BCh); this approach was taken to observe the contribution of the entire M3-PLC-IP3 receptors pathway to MeHg-induced increases in [Ca2+]; (F ohrrnan et al. , 1993; F ukamauchi et al., 1993; Wojcikiewicz et al. , 1994; Simpson et al., 1996; F ukamauchi et al., 2000). Third, the nonspecific muscarinic antagonist atropine was used to determine if MeHg interacts with these receptors to causes elevations of [Ca2+]; (Yan et al., 1995). Finally, the ryanodine receptor antagonist ryanodine was used to inhibit Ca2+ mobilization through these channels (Irving et al., 1992). The results presented here indicate that MeHg interacts with muscarinic receptors to cause Ca2+ release from the SER IP3-sensitive Ca2+ pool, and that this Ca2+ contributes to the whole-cell elevation of [Ca2+], observed during MeHg neurotoxicity. 43 MATERIALS AND METHODS Materials and Experimental Solutions. Deoxyribonuclease I (DNase I) and type II trypsin were purchased from Worthington Biochemicals (Freehold, NJ). Cell culture supplies, including Dulbecco’s Modified Eagle Medium (DMEM) and fetal bovine serum were purchased from Gibco BRL (Grand Island, NY). SK&F 96365 was purchased from Calbiochem (La J olla, CA). F ura-2 acetoxymethylester (fura-2 AM) was purchased from Molecular Probes (Eugene, OR). Thapsigargin, carbamyl-B- methylcholine chloride (bethanechol), carbamylcholine (carbachol), atropine, ryanodine, poly-D-lysine and cytosine-B-arabinofirranoside were purchased fiom Sigma (St. Louis, MO). Methyl mercuric chloride (MeHg) was purchased fi'om ICN Biochemicals Inc. (Aurora, OH) and dissolved in deionized water. Granule cell isolation. Cerebellar granule cells are isolated as described previously (Marty and Atchison, 1997). Approximately 5-10 7-day-old Sprague Dawley rats were used per isolation. Each rat was decapitated, then the skin was cut down the dorsal midline from the neck to the nose. Using the flat end of the scissors, the skull was carefully peeled back to reveal the brain. The cerebellum was cut along each side using scissors, then gently lifted away fiom the rest of the brain. Each extracted cerebellum was placed in an ice-cold beaker containing Cay-Mg2+ fi'ee Hank’s Buffered Saline Solution (CMF-HBSS). CMF-HBSS contains (mM): 5.36 KCl, 0.44 KH2P04, 136.89 NaCl, 20.02 HEPES, 0.59 EDTA, 5.5 d-glucose, and 4.17 NaI-lCO; (pH 7.3, 37°C). Once all cerebella had been dissected, the meninges were removed and the tissue finely minced using sterile scissors. The tissue was then added to a beaker containing 0.25% trypsin (w/v) and digested at 37°C in a Shaking water bath for 15 min. Cold (4°C) DMEM with 5% fetal bovine serum and 0.04% DNase I (w/v) was added to inactivate the trypsin. The cell suspension was centrifuged at 134g for 5 min (4°C), then the pellet was resuspended in DMEM with 5% F BS and 0.04% DNase I and triturated approximately 15 times with a wide-bore fire-polished Pasteur pipette. The cell suspension was allowed to sit for 4 min on ice, then the supernatant (which contained the dissociated cells) was transferred to a centrifuge tube on ice. The addition of PBS-DNase followed by trituration, settling of debris and transfer of supernatant was performed two more times. The collected supernatant was gently underlaid with 4% BSA and 3.82% MgSOa (dissolved in CMF-l-IBSS) and the tube was centrifuged at 134g at 4°C for 5 min. The resulting pellet was resuspended in the plating medium that consists of DMEM with 10% PBS, 25 mM KCl, SOuM GABA, 50uM kainate, 5 rig/ml insulin, 100 U/ml penicillin and 50 ug/ml streptomycin. Cells were plated at a density of 2.0-2.2x106 cells/35 mm dish, each containing a 25 mm glass coverslip coated with 0.1 mg/ml poly-D-lysine. After 21 hours, 10 11M cytosine-B-arabinofirranoside was added to inhibit glial cell proliferation. Penicillin-streptomycin was not added at this or subsequent medium changes because streptomycin alters Ca2+ channel function (Atchison et al. , 1988; Redman and Silinsky, 1994). Cells for all experiments were used at 6-8 days in vitro to allow for cell attachment and maturation. Measuring changes in [Ca2+]r. Unless otherwise noted, the standard buffer used for experimental solutions is HEPES Buffered Saline (HBS) which contains (mM): 150 NaCl, 5.4 KCl, 1.8 CaClz, 0.8 MgSO4, 20 d-glucose, and 20 HEPES (fi'ee acid) (pH 7.3 at room temperature of 23-25°C). The 40 mM K+ solution contains the same components 45 as HBS except with 40 mM K+ and 115.4 mM NaCl. The low-[Ca2+], EGTA-containing bufl‘er (EGTA-HBS) has the same components as HBS minus CaClz and plus 20 11M EGTA. For all experiments, pharmacological agents are dissolved in either HBS or EGTA-HBS, with appropriate controls for any additional solvents used (DMSO or ethanol, maximum final solvent of 0.01% (v/v), for all experiments). In a cell-free system, fura-2 did not interact with MeHg or any of the agents used (BCh, atropine, thapsigargin and ryanodine) at concentrations used in the present study (results not shown). To measure changes in [Ca2+]i, cells were loaded with 3-4 11M fura-2 AM in HBS for 1 hr at 37°C, followed by perfusion with HBS for 30 min. Digital fluorescent images were obtained using a Nikon Diaphot microscope (Nikon, Tokyo, Japan) or an Olympus D(70 microscope (Olympus Optical Co., Tokyo, Japan), coupled to an IonOptix system (Milton, MA) with an Open Perfusion Micro-Incubator (Medical Systems Corp.), maintained at 37°C. For each experiment, changes in emitted fluorescence at excitation wavelengths of 340 and 380 nm were monitored simultaneously in multiple soma (3-10 cells per experiment) within the same microscopic field. The fluorescence ratio (Fm/330) indicated the approximate amount of [Ca2+]i, however the data were not converted to [Ca2 i due to fura-2 interactions with other divalent cations, such as Zn”, a cation known to play a role in the MeHg response (Hare et al., 1993; Denny and Atchison, 1994). All experiments began with a l min wash with HBS (to establish baseline fluorescence), followed by l min exposure to 40 mM K+ solution; only cells which exhibited a reversible increase in [Ca2+], following the KI exposure were considered to be viable for the experiment. The K" solution was washed out for 3 min with HBS, then one 46 of the experimental protocols was followed. Pharmacological agents were used either to empty the SER of its Ca2+ content or block Ca2+ release fiom receptors on the SER membrane. To empty the SER prior to addition of MeHg, the SERCA inhibitor thapsigargin (10 11M) was applied for 5 min immediately prior to MeHg (0.2-1 .0 uM) in HBS or EGTA-HBS (Irving et al., 1992; Simpson et al., 1996). To block the muscarinic receptors, the nonspecific muscarinic receptor antagonist atropine (10 11M) was applied for 10 min prior to, as well as during, exposure to 0.2-1.0 uM MeHg in HBS or EGTA- HBS (Xu and Chuang, 1987; Yan et al., 1995). To inhibit Ca2+ release from the ryanodine receptors, 10 uM ryanodine was applied for 10 min prior to, as well as during, exposure to 0.2-1.0 11M MeHg in HBS or EGTA-HBS (Irving et al., 1992). To down- regulate the M3 muscarinic receptors and thus subsequently the IP3 receptors, 1 mM bethanechol (BCh) was added directly to the growth media 24-30 hr prior to loading cells with firra-2 (Fohrman et al., 1993; F ukarnauchi et al. , 1993; Wojcikiewicz et al., 1994; Simpson et al., 1996; Fukamauchi et al. , 2000). For BCh experiments, the fura-2 AM was added directly to the growth media, thus cells were in BCh-fiee solution for the 30 min I-IBS rinse and the 5 min viability test with 40 mM KL In these experiments, MeHg (0.2-1.0 11M) in HBS or EGTA-HBS was perfused immediately following the 3 min K+ solution wash period. Following exposure to MeHg, the time-to-onset of the first and second CaZ+i phases was determined: the first phase (from release of Ca2+: pools) was measured from the point at which the ratioed fluorescence irreversibly leaves baseline, and the second phase began at the point at which the 380 nm fluorescence sharply drops. For experiments performed in the absence of Ca2+,” the amplitude of the fura-2 ratio increase was 47 determined as the fluorescence increase above baseline, and was normalized to the peak fura-2 ratio caused by the 1 min exposure to 40 mM K" (“normalized ratio”). Ratio data was normalized to remove Slight variations in fluorescence intensity resulting from differences in cell size and fura-2 loading efficiency. For both time-to-onset and normalized ratio experiments, data from each cell observed in an experiment were averaged to provide mean time-to-onset for that dish of cells (n=1). In order to minimize differences between cell isolates, experiments using MeHg alone and MeHg with the pharmacological agent were run on the same day, and experiments using the same agents were performed in at least two separate cell isolates. Statistics. Comparisons of mean time-to-onset for “MeHg” vs the corresponding “MeHg plus inhibitor” cells are made using Student’s paired t-test, with values of p<0.05 considered to be statistically significant. 48 RESULTS The SER contributes Ca2+ to the MeHg-induced release of Car's. To determine whether the SER contributed to the first-phase [Ca2+]t elevation, the SER was depleted of its Ca2+ content with the SERCA inhibitor thapsigargin. To verify that thapsigargin abolished agonist-induced release of Ca2+ fi'om this organelle, experiments were performed using the nonspecific muscarinic agonist bethanechol (BCh, 1 mM), which activates M3 muscarinic receptors to cause production Of 1P3 and subsequent release of Ca2+ fiom the 1P3 receptor (Figure 2.1). A 5 min application of 10 11M thapsigargin abolished the BCh-activated increase in [Ca2+]i in 13 of 15 cells examined (from 3 dishes per treatment and two separate cell isolations), indicating the effectiveness of thapsigargin in reducing the SER Ca2+ pool (Table 2.1). Thapsigargin (10 11M) was then applied for 5 min immediately prior to MeHg (0.2-1.0 11M). In the presence of Ca2+“ thapsigargin did not alter the time-to-onset of the frrst-phase release of [Caz+]i or the second-phase influx of [Ca2+]; at any MeHg concentration tested (Figure 2.2). The time-to-onset for the first-phase increase in fura-2 ratio between neurons treated with MeHg only and MeHg plus thapsigargin was not different for 0.2, 0.5 or 1.0 uM MeHg. Additionally, thapsigargin did not change the time-tO-onset of the second-phase Ca2+ influx for 0.2, 0.5 or 1.0 uM MeHg. In nominally Ca2+,-fiee solution (“EGTA-HBS”), thapsigargin again did not alter the time-to-onset of the first-phase [Cay], increase caused by 0.2, 0.5 or 1.0 11M MeHg (Figure 2.3). 49 080 - g l_(_“ A § B_Ch {Eh s? 0.70 - ('1 LL. v 0 -~ 0.60 - ‘5 al ed 0.50 L L L l 0 5 10 15 20 Time (min) Figure 2.1. Alternating exposures to 40 mM K+ and 1 mM BCh cause increases in [Ca2+]. in cerebellar granule cells. Representative changes in fura-2 fluorescence (“Ratio”) caused by alternating exposure to 40 mM K+ (1 min) and 1 mM BCh (1.5 min), in a single granule cell. The amplitude of the fluorescence change in response to the BCh is normalized to the response to the preceding K+ exposure as a relative measure of the amount of Ca2+ released from the IP3-sensitive Ca2+ pool. This protocol was used to test the effect of pharmacological agents on release of Ca2+ fiom the IP3-Sensitive Ca2+ pool (see Table 2.1). 50 INHIBITOR NORMALIZED NORMALIZED PERCENT RATIO (N O RATIO (WITH DECREASE INHIBITOR) INHIBITOR) No inhibitor 0.71:1:0.11 0.56i0.1 1 21.1% (control) Thapsigargin 0.59d:0.08 0.042t0.03* 93.2% (10 11M) BCh (1 mM) 0.13d:0.01 0.03zt0.01* 76.9% Atropine 0.42:1:0.03 0.004zt0.004* 99.3% (10 11M) Ryanodine 0.51i0.09 0.242t0.04* 53.0% ( 10 11M) Table 2.1. Effect of pharmacological inhibitors on peak fura-2 ratio induced by 1.5 min 1 mM BCh in cerebellar granule cells. Cells were exposed to alternating challenges of 40 mM K+ (1 min) and 1 mM BCh (1.5 min), as shown in Figure 2.1. The pharmacological inhibitor was then applied, and the KI-BCh protocol repeated. The peak fura-2 ratio response to the BCh treatment was normalized to the preceding response to the K+ depolarization (“normalized ratio”). The peak BCh response was compared in the same cell in the absence or presence of inhibitor. Because not all cells responded to BCh, comparisons were made only in cells which responded to the initial BCh exposure. For the 24 hr 1 mM BCh experiments, comparisons were made between cells receiving 24 hr BCh pretreatment and cells receiving 24 hr vehicle pretreatment. Because the experimental protocol for the 24 hr BCh experiments prevented comparisons within the same cell, all cells were included in this data, including control cells that did not respond to the BCh treatment, to prevent any experimental bias. Thus, the normalized ratios ' under control conditions are lower than those observed for the other pharmacological agents due to the inclusion of non-responding cells in the 24 hr BCh data. For the ryanodine experiments, 1 mM carbachol was used instead of 1 mM BCh. Results are presented as meaniSEM (n = 3). The asterisk (*) indicates a value significantly different from the response prior to addition of the pharmacological agent (p<0.05). 51 A - - Thaps - + Thaps A 15 .S E: 8 G O 5': c5 .§ [-1 0.5 1.0 [MeHg] (11M) B - -Thaps - +Thaps .S E 50 8 1:: 0 5'3 ti» .§ [-1 0.5 [MeHg] (FM) Figure 2.2. Thapsigargin does not affect the time-to-onset of MeHg-induced increases in [Ca2+]i. A) Pretreatment with 10 uM thapsigargin (“thaps”) for 5 min did not alter the time-to-onset of the MeHg-induced first-phase [Ca +]i increase caused by 0.2, 0.5, or 1.0 uM MeHg. (B) In the same cells, pretreatment with 10 11M thapsigargin for 5 min also did not alter the time-to-onset of the MeHg-induced second-phase influx of Ca2+, caused by 0.2, 0.5, or 1.0 11M MeHg. Results are presented as mean i SEM (n=3- 5). 52 - - Thaps - + Thaps N O I Time-to-Onset (min) 25 O 0.5 [MeHg] (11M) Figure 2.3. Thapsigargin does not affect the time-to-onset of MeHg-induced [Ca2+]. increases in the absence of Ca2+” Pretreatment with 10 M thapsigargin (“thaps”) for 5 min did not alter the time-to-onset of the MeHg-induced first-phase 2[Caz’TJi increase caused by 0. 2, 0. 5, or 1.0 M MeHg, measured In the absence of Ca2 c. Results are presented as mean i SEM (n=3- 5). 53 - - Thaps - + Thaps N 1 Normalized Ratio O 0.2 0.5 1.0 [MeHg] (11M) Figure 2.4. Thapsigargin decreases the amplitude of MeHg-induced fura-2 ratio increases, as measured in the absence of Ca2+” Pretreatment with 10 11M thapsigargin (“thaps”) for 5 min significantly decreased the amplitude of the MeHg-induced first- phase [Ca2+]; increase caused by 0.2, 0.5, and 1.0 1.1M MeHg. The firm-2 ratio increase was normalized to the amplitude Of the 40 mM K+ response (“Normalized Ratio”). Results are presented as mean i SEM (n=3-5). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 54 However, there was a significant decrease in the amplitude of the first phase caused by 0.2 uM MeHg (from a normalized ratio of 1.11i0.2 to 0.602t0.08), 0.5 uM MeHg (from 1.38:1:024 to 0.86:1:0.08), and 1.0 uM MeHg (from 1.13i0.08 to 0.85:1:009) (Figure 2.4). This represents a decrease in the first-phase ratio amplitude of approximately 46%, 38% and 25% caused by 0.2, 0.5 and 1.0 11M MeHg, respectively. Down-regulation and desensitization of the M3-IP3 receptor pathway decreases the MeHg-induced increase in [Carin A 24 hr application of 1 mM carbachol causes down-regulation and desensitization of M2 and M3 muscarinic receptors, and IP3 receptors, in cerebellar granule cells in culture (F ohrman et al., 1993; Simpson et al., 1994). BCh, like carbachol, is a non-specific muscarinic agonist and thus would be expected to have similar effects; BCh, however, has the advantage of being more selective for muscarinic receptors with minimal activity at nicotinic ACh receptors. Pretreatment with 1 mM BCh for 24 hr significantly inhibited the Ca2+ increase in response to a 1.5 min exposure to 1 mM BCh from 0.13i0.01 to 0.032h0.01, a 77% loss of the IF; response (Table 2.1). The 24 hr BCh treatment was then tested for its effect on MeHg-induced increases of [Ca2+]. BCh pretreatment hastened the time-to-onset of the first-phase increase in fura-2 ratio for 0.2 and 0.5uM MeHg, but did not change the time- to-onset of [Cay]; increases caused by 1.0 11M MeHg (Figure 2.5A). However, the time- to-onset of the second-phase influx of Ca2+, was not changed for MeHg as compared to MeHg plus 24 hr BCh pretreatment for 0.2 uM, 0.5 GT 1.0 11M MeHg (Figure 2.53). 55 - -BCh - +BCh N O l Time-to-Onset (min) 5 0 0.2 0.5 1.0 [MeHg] (PM) B - -BCh - +BCh A 30 - .E E a 20 2 C? $ 10 0) E r— o 0.2 0.5 1.0 [MeHg] (MM) Figure 2.5. Pretreatment with BCh hastens the MeHg-induced release of [Ca2+]i without affecting the onset of MeHg-induced influx of Ca2+, A) Pretreatment with 1 mM BCh for 24 hr significantly hastened the time-to-onset of the first-phase MeHg- induced release of Ca2+, caused by 0.2 and 0.5, but not 1.0 11M, MeHg. B) In the same cells, 1 mM BCh pretreatment did not alter the time-to-onset of the second-phase influx of Ca2+, at any MeHg concentration examined. Results are presented as mean :h SEM (n=4-6). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 56 A 15 ~ .5 E r; 10 ‘8 C? $ 5 0) .§ F- o 0.2 0.5 1.0 [MeHg] (11M) Figure 2.6. Pretreatment with BCh hastens the MeHg-induced release of [Ca2+]i, as measured '- lh absence of Ca2+, Pretreatment with 1 mM BCh for 24 hr significantly hastened tin: tim-to-onset of the first-phase increase in [Ca2+]i caused by 0.2 and 0.5, but not 1.0, uM, MeHg. Results are presented as mean :1: SEM (n=4-5). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 57 CCCP + Oligomycin A 0.80 - A § § 070 - -BCh LL? v o g 0.60 a: 0.50 ' L ‘ 0 5 10 15 Time (min) CCCP + Oligomycin 0.80 — B A a3 +BCh ,5, 0.70 *- LL. v o g 0.60 ' ad 0.50 4 ‘ ‘ 0 5 10 15 Time (min) Figure 2.7. Pretreatment with BCh (1 mM, 24 hr) does not increase the [Ca2+] within mitochondria. Representative changes in fura-2 fluorescence in response to mitochondrial depolarization with carbonyl cyanide m-chlorophenylhydrazone (CCCP) (5 11M) and oligomycin (10 11M). Granule cells were treated with vehicle (“-BCh”) or 1 mM BCh (“+BCh”) for 24 hr. CCCP and oligomycin were then applied to depolarize the empty the mitochondria of their Ca2+ content (see Chapter Four) to determine the amount of Ca2+ in these organelles following BCh treatment. 58 - -BCh - +BCh Normalized Ratio 0.2 0.5 1.0 [MeHg] (11M) Figure 2.8. Pretreatment with BCh decreases the amplitude of the fura-2 ratio increase caused by MeHg, as measured in the absence of Ca2+, Granule cells were treated with 1 mM BCh for 24 hr prior to 0.2-1.0 11M MeHg in EGTA-HBS. Results are presented as mean :1: SEM (n=4-5). The asterisk (*) indicates a value significantly different fi‘om the corresponding MeHg-only data (p < 0.05). 59 In the absence of Cab}, 24 hr pretreatment with BCh again hastened the time-to-onset of the frrst-phase ratio increase for 0.2 and 0.5 11M MeHg but not 1.0 11M MeHg (Figure 2.6). The hastened onset of the first phase was not due to an increased amount of Ca2+ in the mitochondria, as mitochondrial membrane depolarization with 5 uM carbonyl cyanide m-chlorophenylhydrazone (CCCP) and 10 11M oligomycin indicated no difference in the amount of mitochondrial Ca2+ content in cells treated with BCh as compared to untreated cells (Figure 2.7). Like thapsigargin, BCh pretreatment also decreased the amplitude of the first-phase fura-2 ratio increase caused by 0.2 11M MeHg (fiom 1.11i0.13 to 0.65i0.06), 0.5 11M MeHg (from 1.14:1:0.17 to 0.70:1:0.15), and 1.0 11M MeHg (from 1.00:1:0. 12 to 0.69d:0.09) (Figure 2.8). This represents a decrease in the first-phase ratio amplitude of approximately 41%, 39% and 31% caused by 0.2, 0.5 and 1.0 uM MeHg, respectively. Blocking muscarinic receptors with atropine also decreases the MeHg- induced increase in [C32+]|. To determine whether MeHg causes Ca2+ release from the SER via binding to the M3 muscarinic receptor, experiments were performed using the nonspecific muscarinic receptor antagonist atropine. Atropine blocked the response to a brief (1 .5 min) 1 mM challenge with BCh, indicating the effectiveness of this compound in blocking muscarinic receptors (Table 2.1). Application of 10 11M atropine for 10 min prior to, as well as during, exposure to MeHg in the presence of Ca2+, indicated that atropine significantly delayed the time-to-onset of the first-phase ratio increase caused by 0.2 and 0.5 11M MeHg, but not did not change the time-to-onset of elevations of [Ca2+]; caused by 1.0 11M MeHg (Figure 2.9A). Atropine did not alter the time-to-onset of the second-phase influx of Ca2+c caused by 0.2, 0.5 or 1.0 uM MeHg (Figure 2.93). 60 - MeHg only - MeHg + AP 60 Time-TO-Onset (min) 0.2 0.5 1.0 [MeHg] (11M) - MeHg only - MeHg + AP Time-TO-Onset (min) 0.2 0.5 1.0 [MeHg] (FM) Figure 2. 9. Atropine (“AP”) delays the first-phase release of Ca2+ .without affecting the second-phase influx of Ca2+ e caused by MeHg. A) Granule cells were treated with atropine (10 11M) for 10 min prior to, as well as during, exposure to MeHg. Atropine significantly delayed the time-to-onset of the first- phase MeHg- induced release of Ca2 . caused by 0. 2 and 0. 5, but not 1.0 11M, MeHg. B) In the +same cells, atropine did not alter the time-tO-onset Of the second- phase influx of Ca2+ c at any MeHg concentration examined. Results are presented as mean :1: SEM (n=4). The asterisk (") indicates a value significantly different from the corresponding MeHg—only data (p < 0.05). 61 - MeHg only - MeHg + AP .5. o I U) Q p—i O Time-To-Onset (min) N O O 0.2 0.5 1.0 [MeHg] (HM) Figure 2.10. Atropine (“AP”) delays the time-to-onset of the MeHg-induced first- phase release of Ca2+” as measured in the absence of Ca2+, Treatment with atropine (10 11M) for 10 min prior to, as well as during, exposure to MeHg Significantly delayed the time-to-onset of the [Ca2+]: increase caused by 0.2 and 0.5, but not 1.0, 11M MeHg. Results are presented as mean i SEM (n=3-6). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 62 --AP -+AP 1.50 - .9. a ,6 1.00 O) .521. "a g 0.50 0 Z 0.00 0.2 0.5 1.0 [MeHg] (11M) Figure 2.11. Atropine (“AP”) decreases the amplitude of the fura-2 ratio increase caused by MeHg, as measured in the absence of Ca2+” Granule cells were treated with atropine (10 11M) for 10 min prior to, as well as during, exposure to MeHg. Results are presented as mean i SEM (n=3-6). The asterisk (‘) indicates a value significantly different fiom the corresponding MeHg-only data (p < 0.05). 63 In cells treated with MeHg in EGTA-HBS, atropine again delayed the time-to-onset of the first-phase ratio increase caused by 0.2 and 0.5 11M MeHg but did not change the time-to-onset Of the first phase caused by 1.0 11M MeHg (Figure 2.10). Additionally, atropine caused a Significant decrease in the first-phase amplitude caused by 0.2 uM MeHg (fi'om a normalized ratio of 1.21:1:0.06 to 0.73d:0.05), 0.5 uM MeHg (fi'om 1.042t0.10 to 0.67i0.09), and 1.0 11M MeHg (from 08810.05 to 0.54i0.06) (Figure 2.11). This represents a decrease in the first-phase ratio amplitude of approximately 40%, 36% and 39% caused by 0.2, 0.5 and 1.0 uM MeHg, respectively. Effect of ryanodine on MeHg-induced [Ca2+]. elevations. Granule cells in culture exhibit a developmental profile in the ability of their ryanodine receptors to respond to agonists, such that after 4 days in culture, they lose the ability to respond to ryanodine receptor agonists such as caffeine (Irving et al., 1992); unpublished observation). However, these receptors are still functional and participate in CICR, as evidenced by the ability of 10 11M ryanodine to inhibit CICR initiated fi'om 1P; receptors in granule cells in culture (Irving et al., 1992). Because all experiments in this dissertation were used at days 6-8 in vitro, it was impossible to test directly the ability of ryanodine, and thus experiments were performed to test the ability of 10 uM ryanodine to inhibit CICR initiated by release of Ca2+ fi'om the 1P3 receptors when activated with 1 mM carbachol for 1 min. In agreement with Irving et al., (1992) ryanodine inhibited elevations of [Ca2+], initiated by release of Ca2+ through the 1P3 receptor, as evoked by the muscarinic agonist BCh (Table 2.1). Treatment with 10 11M ryanodine for 10 min prior to, as well as during, treatment with 0.2-1.0 11M MeHg, did not change the time-to- onset of the first-phase ratio increase for 0.2, 0.5 or 1.0 11M MeHg (Figure 2.12A). Ryanodine also did not delay the time-to-onset of the second-phase Ca2+ influx caused by 0.2, 0.5 or 1.0 pM MeHg (Figure 2.123). In the nominally Ca2+-f1’ee buffer, ryanodine again did not change the time-to-onset of the first-phase ratio increase caused by 0.2, 0.5 or 1.0 11M MeHg (Figure 2.13). Ryanodine also failed to decrease amplitude of the first- phase ratio increase caused by MeHg at any concentration examined (Figure 2.14). Contribution of store-operated Ca2+ entry (SOCE). Experiments were performed to determine whether MeHg-induced emptying of the SER triggers opening of store-Operated Caz+.channels (SOCC), thereby contributing to the second-phase influx of Ca2+c. As shown in Figure 2.15, treatment with 10 11M thapsigargin for 10 min, followed by 1 mM EGTA dissolved in HBS, causes a noticeable decline in the baseline fura-2 ratio. Immediately following replacement of the EGTA solution with HBS (which contains 1.8 mM Ca2+), the baseline firra-2 ratio returns to near-pretreatment levels, which is consistent with opening of SOCCS. Experiments were then performed to determine whether the return to normal [Ca2+]; was due to opening of SOCCS. The socc inhibitor SK&F-96365 was tested for its ability to inhibit Ca2+ entry through SOCCS. SK&F-96365 is a known inhibitor of SOCCS (Lo and Thayer, 1995); however, SK&F-96365 is reported to also interact with voltage-operated Ca2+ channels (V OCCS), inhibiting Ca2+ movement through these channels (Koch et al. , 1994). For this reason, the effect of 10 uM SK&F-96365 on VOCCS was examined by adding the compound during plasma membrane depolarization-induced influx of Ca2+c caused by l min exposure to 40 mM K‘“. As seen in Figure 2.16, SK&F-96365 significantly decreased Ca2+ entry through VOCCS by approximately 50%; thus, this compound could not be used to study the contribution of SOCE to MeHg-induced influx of Ca2+e. 65 A - MeHg only - MeHg + Ry 75‘ 60 55, F 8 a o 6 1'.‘ 0 .§ i- 0.2 0.5 1.0 [MeHg] (11M) - MeHg only - MeHg + Ry E E a t: 0 c5 [1‘ 0 .§ [-1 0.2 0.5 1.0 [MeHg] (FM) Figure 2.12. Ryanodine (“Ry”) does not affect the time-to-onset of MeHg-induced increases in [Ca2+]r. A) Treatment with 10 11M ryanodine for 10 min prior to, as well as during MeHg exposure, did not alter the time-to-onset of the MeHg-induced first-phase [Ca2+]: increase caused by 0.2, 0.5, or 1.0 uM MeHg. (B) In the same cells, treatment with 10 uM ryanodine also did not alter the time-to-onset of the MeHg-induced second- phase influx of Ca2+.= caused by 0.2, 0.5, or 1.0 11M MeHg. Results are presented as mean :1: SEM (n=6-7). 66 - MeHg only - MeHg + Ry L» O 1 N O 1 p—d O I Time-To-Onset (min) O 0.5 1 .0 [MeHg] (HM) Figure 2.13. Ryanodine (“Ry”) does not affect the time-to-onset of the MeHg- induced first-phase release of Ca2+” as measured in the absence of Ca2+, Treatment with ryanodine (10 uM) for 10 min prior to, as well as during, exposure to MeHg significantly does not alter the time-to-onset of the {Cay} increase caused by 0.2, 0.5, or 1.0 11M MeHg. Results are presented as mean i SEM (n=4-7). 67 --Ry -+Ry 1.50 — .2 a ,6 1.00 - a) .u To g 0.50 0 Z 0.00 0.2 0.5 [MeHg] (11M) Figure 2.14. Ryanodine does not affect the amplitude of MeHg-induced increases in [Ca2+]i as measured in the absence of Ca2+” Ryanodine caused slight decreases in the amplitude of the MeHg-induced first-phase [Ca2+]i increase caused by 0.2, 0.5, and 1.0 uM MeHg; however, the differences were not statistically significant. Results are presented as mean i SEM (n=4-7). 68 0.45 E 13 ii BCh Th HBS BCh A ‘— a si ar in — § 0.40 - p g g EGTA ll a 3. El 035 - o .5 M) l 9‘2 0.30 - 0.25 L l l l l i l L J 0 5 10 15 20 25 30 35 40 45 Time (min) Figure 2.15. Depletion of SER Ca2+ triggers SOCE in rat cerebellar granule cells. Granule cells were exposed to 40 mM K+ (1 min) followed by 1 mM BCH (1.5 min) to release Ca2+ from the IPg-Sensitive Ca2+ pool. Thapsigargin (10 11M, 10 min) followed by a hi concentration of EGTA (1 mM, 5 min) caused a noticeable decline in the baseline [Ca +]i. Returning the cells to a Cay-containing medium (HBS) caused an immediate return to baseline [Ca2+]i, which is characteristic of SOCE. The loss of SER Ca2+ content with this treatment is confirmed by the loss of response to a second exposure to 1 mM BCh (1.5 min). 69 1.00 ' g 0.80 C6 Gd 3 0.60 .52]. 7'9 0.40 E z 0.20 0.00 - + SKF-96365 (10 11M) Figure 2.16. SK&F 96365 inhibits depolarization-stimulated Ca2+ influx through VOCCs. Granule cells were exposed to 40 mM K+ for 1 min, in the presence or absence of 10 11M SK&F-96365, and the amplitude of the peak [Cay]; response to the K+ depolarization was measured. Results are presented as mean :1: SEM (n=4). The asterisk (*) indicates a value significantly different from the corresponding vehicle control data (p < 0.05). 70 DISCUSSION MeHg increases [Ca2+]; in rat cerebellar granule neurons, which contributes to MeHg-induced neuronal death (Marty and Atchison, 1997; Marty and Atchison, 1998). Because loss of Ca2+ regulation is a pivotal step in both apoptosis and necrosis, these experiments sought to identify the Ca2+ source in granule cells to understand the mechanism by which MeHg causes elevations of [Cam]; in cerebellar granule cells. MeHg interacts with muscarinic ACh receptors (Castoldi et al. , 1996; Coccini et al., 2000), causes generation of 1P3 in rat cerebellar granule cells in culture (Sarafian, 1993), and causes Ca2+ release through the 1P3 receptor in NG108-15 cells and rat T lymphocytes (Tan et al., 1993; Hare and Atchison, 1995); thus, these experiments examined whether MeHg interacts with these receptors to cause Ca2+ release fiom the SER. Data presented in this chapter reveals that Ca2+ release from the SER by MeHg contributes approximately 30-40% to the observed increase in fura-2 ratio during the initial phase of MeHg exposure. This differs fi'om results Obtained in NG108-15 cells, in which emptying the SER with thapsigargin and bradykinin prior to MeHg treatment reduced the MeHg-induced frrst-phase ratio increase by 68% (Hare and Atchison, 1995). As was found in NG108-15 cells, the ryanodine receptor Ca2+ pool in granule cells contributes minimal Ca2+ during MeHg exposure (Hare and Atchison, 1995). All four drug treatments (thapsigargin, BCh, atropine and ryanodine) decreased the first-phase ratio amplitude to a Similar extent for all MeHg concentrations examined, suggesting that MeHg activates a common signaling pathway which involves muscarinic receptors, 1P3 receptors and ryanodine receptors. 71 While pharmacological manipulation of the SER with different compounds caused similar decreases in the amount of Ca2+ released fi'om within the cell by MeHg, there was a complicated pattern of effects on the time-to-onset of the initial increase in [Ca2+]i. Atropine delayed the time-to-onset of the initial [Ca2+]; elevation, suggesting that the MeHg-induced elevation of [Ca2+]; is initiated by hiding of MeHg to muscarinic receptors. Because granule cell M3 receptors are coupled to Gq, which leads to generation of 1P3 through PLC, while M2 receptors are linked to G, which inhibits adenylate cyclase and thus is not involved in Ca2+ regulation, the effect of atropine is likely due to inhibition of M3 receptors (Whitham et al., 1991; Doble et al., 1992; Contrera et al. , 1993; F ohrman et al. , 1993). These results support the observations that MeHg binds multiple subtypes of muscarinic receptors in both the central and peripheral nervous systems (Abd-Elfattah and Shamoo, 1981; Castoldi et al., 1996; Coccini et al., 2000) and causes increased [IP3] in cerebellar granule neurons in vitro (Sarafian, 1993). The binding of MeHg to muscarinic receptors may partially explain the selective neurotoxicity of specific cell types within the central nervous system (CNS). In rats, granule cells express the highest density of muscarinic receptors in the cerebellar cortex, as assessed by [3H]quinuclidinyl benzilate binding (Neustadt et al., 1988). In vivo studies indicate that MeHg causes an up-regulation of muscarinic receptors in both the cerebellum and hippocampus, which tend to accumulate MeHg and exhibit neuronal death following MeHg exposure (Coccini et al., 2000). In granule cells, inhibition of muscarinic receptors protects against apoptosis induced by growth in non-depolarizing conditions (Y an et al., 1995), suggesting that muscarinic receptor activation can activate cell death in granule cells. Because the primary downstream target of M3 receptors is the 72 SER-localized 1P3 receptor, the functional consequence of MeHg activation of M3 receptors would be Ca2+ release from the SER, which would then contribute to the whole- cell [Cazji elevation observed during MeHg exposure. Depletion of the SER Ca2+ store with the smooth endoplasmic reticulum Ca2+ ATPase (SERCA) inhibitor thapsigargin, while decreasing the amount of Ca2+; released into the cytosol, did not alter the time-to-onset of the initial Ca+ increase. Complete emptying of the SER by thapsigargin would be expected to delay or prevent the MeHg- induced first-phase increase of [Ca2+], if it originated from the SER, as is suggested by the atropine data. However, this would only be the case if thapsigargin completely er3.311315% the SER of its CaZ+ content. In these experiments, it was found that a 5 min trea-tment with 10 11M thapsigargin abolished the response to treatment with 1 min 1 mM Bell, indicating loss of responsiveness of the M3 muscarinic receptor-1P3 receptor pat—I'lway. This was presumably due to store depletion. However, in experiments Dex-formed with rat cerebellar granule cells in primary culture, 20 min treatment with 10 MM thapsigargin only reduced the SER Ca2+ store by approximately 50% in spite of the cOmplete abolishment of response to 1 mM carbachol (Masgrau et al., 2000). The authors concluded that granule cells may contain a thapsigargin-insensitive SER Ca2+ pool; this Ca2+ pool could be the initial target of MeHg, thereby contributing to the Observed [Ca2+]; increases irrespective of thapsigargin treatment. The total SER Ca2+ pool in granule cells can be subdivided into pharrnacologically, and possibly physically, distinct subsets, making the existence of a thapsigargin-insensitive pool plausible (Simpson et al., 1996). Thus, the ability of thapsigargin treatment to decrease the amplitude of the first-phase [Ca2+]; increase without affecting the time-to-onset of the 73 initial [Ca2+]; elevation suggests that the thapsigargin-sensitive store does contribute to the overall [Ca2+]i increase without necessarily being the initial target. In order to examine the contribution of Ca2+ released fi'om the 1P3 receptor, cells were treated with BCh (1 mM, 24 hr) to down-regulate the SER 1P3 receptor. In granule cells in vitro, 24 hr treatment with 0.1-1 mM carbachol is well documented as causing down-regulation and desensitization of M2, M3 and Type I IP3 receptors (F ohrman et al., 1 993; Fukamauchi et al., 1993; Simpson et al., 1994; Simpson et al., 1996), thus a 24 hr pretreatment with 1 mM BCh should cause a similar down-regulation of these receptors. me 24 hr BCh treatment caused a 77% decrease in the peak Ca2+ response to a brief 1 mm BCh treatment (Table 2.1), which agreed with the 77i14% loss of carbachol-induced I-esponse caused by 24 hr 1 mM carbachol exposure (Simpson et al., 1996). These same authors found that while 24 hr carbachol treatment caused a 65% reduction in agonist- induced accumulation of 1P3, it did not abolish the response (Simpson et al. , 1996), suggesting that 24 hr BCh treatment might inhibit but not abolish the MeHg-induced rel ease of Caz+i if it occurs through the 1P; receptor. Accordingly, exposure to 1 mM B Ch decreased the amplitude of the first-phase Ca2+, elevation, indicating that loss of [P3 receptors corresponded with a decreased response to MeHg. Surprisingly, BCh treatment hastened the release of Ca2+; during MeHg exposure. The BCh treatment does not alter the armount of Ca2+ release by preloaded mitochondria, thus the BCh treatment does not Shi fl (3a2+ from the SER to the mitochondria. Because 24 hr carbachol treatment reduces me number of 1P3 receptors by approximately 80% (Simpson et al., 1994) and reduces but does not abolish the 1P3 accumulation following muscarinic receptor activation 74 (Simpson et al. , 1996), the remaining 1P3 receptors could exhibit increased activity to compensate for the lost receptors. Depletion Of Ca2+; stores can trigger entry of Ca2+.: via store-Operated Ca2+ entry ( SOCE) in a variety of cells, including neurons. Previous experiments indicated that MeHg causes influx of Ca2+.: which cannot be completely blocked by voltage-gated Ca2+ channel blockers (Marty and Atchison, 1997), leading to the hypothesis that at least part of this Ca2+ influx was through activation of SOCE. Rat cerebellar granule cells do undergo SOCE when the SER is depleted of Ca2+ by a combination of thapsigargin and high-EGTA medium; however, direct attempts to block SOCE with SK&F 96365 were unsuccessful, as this compound did not inhibit SOCE while causing significant inhibition of C 32+: entry through voltage-gated Ca2+ channels. MeHg-induced activation of SOCE in granule cells is doubtful, however, because of the lack of effect any SER targeting compound on the MeHg-induced influx of Cab}. Because none of the pharmacological agents used in this study altered the MeHg-induced Ca2+ influx, this suggests (but remains untested) that SOCE does not play a role in the observed loss of Ca2+, regulation. 75 CHAPTER THREE ACUTE METHYLMERCURY EXPOSURE CAUSES NEURONAL DEATH IN RAT CEREBELLAR GRAN ULE CELLS THROUGH INTERACTION WITH INOSITOL-l,4,5-TRIPHOSPHATE RECEPTORS 76 ABSTRACT Low concentrations of methylrnercury (MeHg) cause Ca2+-dependent cell death in rat cerebellar granule cells in vitro. As demonstrated in Chapter Two, MeHg interacts with muscarinic ACh receptors to cause the release of Ca2+ fiom the SER, primarily through the inositol-1,4,5-triphosphate (1P3) receptor, in cerebellar granule cells. Experiments presented in this chapter examined whether this Ca2+ release plays a role in MeHg- induced cell death in granule cells. Acute exposure to increasing concentrations of MeHg (0-2-1.0 11M) caused a corresponding increase in cell death at 24.5 hr post-exposure. Pfior down-regulation of the M3 muscarinic receptor-1P3 receptor pathway with the muscarinic receptor agonist bethanechol (BCh, 1 mM) provided significant protection against MeHg-induced cell death. This protection was reversed by the non-specific muscarinic receptor antagonist atropine (10 11M) and the M3 receptor antagonist 4- diphenylacetoxyl-N-methylpiperidine methiodide (10 uM), but not the M2 receptor antagonist methoctramine (10 uM) or neuronal nicotinic receptor antagonist dihydro-B- e I'3’11111‘oidine hydrobromide (10 11M). Atropine alone (10 uM) was not protective against MeHg-induced neuronal death. Acute or 24 hr pre-treatrnent with the [Pg-receptor antagonist xestospongin C (XeC) (1 llM) was also not neuroprotective, with 24 hr XeC itself causing significant cell death. Inhibition of the SER ryanodine receptors with 10 ”M I'D’arrodine was also not protective, indicating that MeHg-induced Ca2+ release from tIfiS store does not play a Significant role in cell death. These results indicate that ‘nhibition of MeHg-induced Ca2+ release from the SER protects against MeHg-induced Cell (1 77 INTRODUCTION Results of experiments described in Chapter Two indicate that a significant portion of the MeHg-induced increase of [Cay]; in granule cells originates from the SER, as released through the 1P3 receptor. This release of Ca2 from the IP3-sensitive Ca2+ pool appears to be due to interaction of MeHg with muscarinic receptors, although the possibility of a direct interaction between MeHg and intracellular Ca2+ channels is not ruled out. The next set of experiments was designed to test the hypothesis that MeHg- induced Ca2+ release from the SER plays a role in MeHg-induced cell death in cerebellar granule cells. To test this hypothesis, pharmacological inhibitors of specific receptors and ion channels involved in Ca2+i signaling pathways were tested for their ability to pI-event MeHg-induced cell death. The receptors and channels examined in these experiments include muscarinic ACh receptors, 1P; receptor-linked channels and ryanodine receptor-linked channels. Granule cells express a high density of M2 and M3 muscarinic receptors; aetivation of the M3 receptor leads to generation of 1P3 by activation of PLC (Alonso et 0L , l 990; Fohrman et al., 1993; Fukamauchi et al., 1993; Simpson et al., 1994). M ‘JScarinic receptors of all subtypes contain two critical thiol groups within the r<=C‘P'ID1ior’s active site, suggesting that MeHg could potentially bind to, and activate, these receli>t<>tlents of the SER-linked Ca2+ signaling pathways altered MeHg-induced cell death in rat Cerebellar granule neurons in primary culture. The first set of experiments 79 examined the effect of SER store depletion on MeHg-induced cell death using the SERCA inhibitor thapsigargin. The second set of experiments examined the effect of down-regulation of both the 1P3 receptor and the muscarinic receptors using the nonspecific muscarinic agonist bethanechol (BCh). Specific portions of this signaling pathway were also examined using specific muscarinic receptor subtype antagonists and the IF; receptor inhibitor XeC. The final set of experiments examined the role of the Iyanodine receptors in MeHg-induced cytotoxicity using the ryanodine receptor inhibitor ryanodine. 80 MATERIALS AND METHODS The materials and solutions for isolation of cerebellar granule cells are described in detail in Chapter Two. Unless specifically stated, all solutions were identical to those described in Chapter Two. The methods and solutions described below are specific to this chapter. Materials and experimental solutions. The Live/Dead Eukolight viability/cytotoxicity kit was purchased from Molecular Probes (Eugene, OR). 4-diphenylacetoxyl-N- methylpiperidine methiodide (4-DAMP), dihydro-B-erythroidine hydrobromide (DHE), and methoctramine hydrochloride (methoctramine) were purchased fiom Sigma (St. Louis, MO). Xestospongin C (XeC) was purchased fiom Calbiochem (La Jolla, CA). Cell Laocate coverslips were purchased fiom Eppendorf Scientific (Madison, WI). Experimental solutions were prepared in sterile-filtered HBS. MeHg was prepared as a 5 mM stock solution in deionized water and diluted to working concentrations. Thapsigargin was prepared as a 1 mM stock in ethanol. All other p ha-rlnacological inhibitors were prepared as stock solution in deionized water. Control solutions contained an equal volume of vehicle in sterile HBS. Rat cerebellar granule cell isolation procedure. Cerebellar granule cells were isolated as described in Chapter Two and were plated at a density of 1.8x108 cells/35 mm2 dish, ch c()ntaining a 13 mm CellLocate coverslip coated with 0.1 mg/ml poly-D-lysine. 81 C ells were maintained at 37°C in 5% C02 for 6-8 days in vitro to allow for cell maturation (Aronica et al. , 1993). Measurement of granule cell viability. The cell viability protocol was designed to parallel experimental conditions in previous experiments examining MeHg-induced e levations of [Ca2+]i in granule cells in primary culture (Marty and Atchison, 1998). To replicate these conditions, all cells were incubated for 90 min at 37°C in HBS to mimic the fura—2 AM incubation (1 hr) and HBS perfusion rinse (30 min) of prior [Ca2+]i measurement experiments. Cells were then exposed to HBS (76 min), 0.2 uM MeHg (76 min), 0.5 uM MeHg (45 min) or 1.0 uM MeHg (38 min). MeHg exposure duration and concentrations were based on previous experiments by Marty and Atchison (1997), which determined these times sufficient for approximately 97% of cells to undergo MeHg- induced release of intracellular Ca2+ stores and influx of Caz} For all experiments, appropriate solvent and drug controls were performed at the same time for 76 min each, as this was the maximum MeHg exposure interval. Following MeHg exposure, cells were returned to conditioned medium (to bind excess MeHg) for 24 hr. Previous ex periments indicated no significant loss of cell viability immediately following exposure atld significant cell death at 24 hr post-exposure, thus cells were examined at 24 hr in these experiments. At 24 hr after treatment, cells were removed from conditioned medium, rinsed twice with HBS and overlaid with Live/Dead Eukolight Viability/Cytotoxicity reagents (Molecular Probes, Eugene, OR) for 30 min at 37°C. As per tile kit instructions, the reagent buffer contains 2 uM calcein acetoxymethylester (caICQin-AM) and 4 uM ethidium homodimer which labels healthy cells green (calcein- 82 AM) and dead cells red (ethidium homodirner), both dissolved in HBS. Following the 30 min exposure, cells were examined using a Nikon Diaphot epifluorescence microscope and the number of live and dead cells in 8 randomly selected CellLocate grids were counted in order to determine percent viability for that coverslip. Because of the 30 min period in Live/Dead reagent, cells were counted at 24.5 hr after the cessation of MeHg treatment. To determine whether emptying of the SER Ca2+ pool inhibited MeHg-induced neuronal death, cells were treated with 10 pM thapsigargin for 5 min immediately prior to 0-1.0 pM MeHg. Previous work demonstrated that this protocol empties the thapsigargin-sensitive SER Ca2+ store (Simpson et al., 1996); previous chapter). Following MeHg treatment, cells were returned to conditioned medium for 24 hr, and counted at 24.5 h post-exposure. To determine if BCh treatment protected cells from MeHg-induced neuronal death, BCh was added directly to the growth medium to give a final concentration of 1 mM BCh. Similar exposure to 1 mM carbachol down-regulates the M3 receptor in granule cells as a result of continued receptor stimulation, thus 24 hr of 1 mM BCh would be expected to cause a similar down-regulation (Fohrman et al., 1993; Fukamauchi et al. , 1 993 ; Wojcikiewicz et al., 1994; Simpson et al., 1996; Fukamauchi et al., 2000). At 24 hr afier application of BCh, cells were treated with MeHg alone or MeHg plus 1 mM BCh- I" or each set of cells exposed to MeHg there were corresponding HBS, 1 mM BCh, and Vehicle control cells which were treated and counted at the same time as the MeHg- treated cells. For cells treated with BCh, all HBS washes contained 1 mM BCh; additiOnally, the conditioned medium to which they were returned after MeHg treatment 83 contained 1 mM BCh for the entire 24 hr post-treatment period. In experiments in which B Ch was added with a specific Cholinergic receptor antagonist, the antagonist was added for 10 min prior to the initial application of 1 mM BCh, as well as during BCh treatment and HBS wash. The antagonists tested were atropine (nonspecific muscarinic receptor alltagonist), methoctramine (M2 antagonist), 4-DAMP (M3 antagonist) and DHE (neuronal nicotinic ACh receptor antagonist), all at a final concentration of 10 pM. The concentrations of each antagonist were based on the results of Yan et al. (1995) and Lin et al. (1997), who used these compounds to inhibit cell death induced by non- depolarizing growth medium and mastoparan, respectively, in rat cerebellar granule neurons in primary culture. In a similar set of experiments, atropine was added directly to the conditioned medium to a final concentration of 10 pM 10 min prior to, as well as dming, exposure to 0-1.0 uM MeHg. In the next experiment, the reversible IP3-receptor inhibitor XeC (1 uM) was added for 10 min prior to, as well as during, treatment with 0-1.0 pM MeHg. This concentration of XeC was chosen based on the results of Netzeband etal., 1999, who found 1 uM XeC sufficient to prevent Ca2+ release from the 1P3 receptor in rat cerebellar gratlule neurons in primary culture. Finally, cells were treated with the ryanodine l‘e<=€=l:>t0r inhibitor ryanodine (10 pM) for 10 min prior to, as well as during, exposure to 0‘ 1 - 0 1.4M MeHg. This concentration of ryanodine has previously been shown to be e1i‘eC-‘etive at inhibiting Ca2+-induced Ca2+ release from ryanodine receptors in cerebellar granule cells (Irving et al., 1992). 84 Statistics. To account for variability between cell isolates, each set of experiments was performed in at least two separate isolates. Further, all experiments within a set (HBS only, drug only, and MeHg with or without drug) were performed on the same day (n=l), allowing the data to be paired for statistical comparison. For each exposure group, the percent viability (number of live cells/total number of cells x 100) was determined, and these percentages were normalized using an angular transformation. Values are presented as the mean i SEM. Comparisons of cell viability were made using a repeated measures analysis of variance (AN OVA) followed by Tukey’s procedure for post-hoe comparisons to compare MeHg to MeHg + inhibitor, with p < 0.05 considered to be statistically significant. 85 RESULTS These experiments were designed to examine the role of SER Ca2+ in MeHg- induced death of cerebellar granule neurons. In the first set of experiments, the effect of MeHg on granule cell viability was determined both immediately and 24 hr after exposure to 0-2 pM MeHg. Cell viability is similar to control immediately following MeHg treatment; however, at 24 hr postoexposure MeHg causes increasing cell death in a concentration-dependent manner (Figure 3.1). These results confirm earlier findings by Marty and Atchison (1998). Because MeHg does not immediately cause neuronal death, all cell viability assays were performed at 24 hr post-exposure to MeHg. The next experiment sought to determine whether emptying of the SER Caz+ pool with the SERCA inhibitor thapsigargin was protective against cell death at 24.5 h post- exposure to 0-1 uM MeHg. Cells were treated with 10 pM thapsigargin for 5 min immediately prior to MeHg; this exposure was based on earlier experiments in which this treatment depleted the SER of its Ca2+ content (previous chapter). However, this brief exposure to thapsigargin was itself toxic, causing cell death at 24.5 h post-exposure (results not shown). In the next set of experiments, the nonspecific muscarinic agonist BCh (1 mM) was applied for 24 hr prior to MeHg treatment to down-regulate the muscarinic-1P3 I'e<=el:>tor pathway. Application of 1 mM carbachol is well documented as causing down- regulation and desensitization of M2, M3 and subsequently, 1P3 receptors in granule nemons (Fohrman et al., 1993; Fukamauchi et al. , 1993; Wojcikiewicz et al., 1994; Simpson et al., 1996; Fukamauchi et al., 2000). Treatment with 1 mM BCh significantly 86 - Immediate % 24hr % Viability MeHg (PM) Figure 3.1. MeHg-induced cell death at 30 min and 24 hr post-exposure. Immediately following exposure to 0.2-2.0 uM MeHg, there is no apparent loss of g1‘ar1ule cell viability. However, at 24 hr post-exposure, there is a concentration- dependent increase in cell death. Results are presented as mean :t SEM (n = 4). ( * )significantly different from the corresponding 0 uM MeHg data (p < 0.05). 87 inhibits agonist-induced release of Ca2+ through granule cell 1P3 receptors (previous chapter), suggesting that BCh, like carbachol, also causes down—regulation of the receptors in this pathway. In the first experiment, cells were treated with 1 mM BCh for 24 hr, followed by O— 1 uM MeHg for the durations described earlier. As seen in Figure 3.2, increasing concentrations of MeHg killed an increasing percentage of cells at 24.5 h post-exposure. 'I'he loss of cell viability was consistent with that observed in previous experiments (Marty and Atchison, 1998). Treatment with 1 mM BCh significantly protected cells fiom MeHg-induced cell death at all concentrations examined. Furthermore, at 0.2 and O - 5 pM MeHg the incidence of cell death was equal to that of untreated cells. However, at 1 uM MeHg, the protection provided by 1 mM BCh was not as great as at the lower M eI-Ig concentrations. To determine whether the protection provided by BCh was due to action at muscarinic receptors, the nonspecific muscarinic antagonist atropine (10 uM) was applied for 10 min prior to BCh, as well as during the 24 hr BCh treatment for 24 hr prior to exposure to MeHg. Cell viability was again measured at 24 hr post-MeHg exposure. As seen in Figure 3.3, addition of atropine prevented the protection afforded by 1 mM BCh (Figure 3.2), resulting in no significant difference in the number of cells surviving MeHg alone or MeHg plus BCh and atropine. Granule cells express M2 and M3 receptors (Alonso et al., 1990), thus the prOtection provided by BCh could be through either or both of these receptors. In a Similar protocol to the BCh plus atropine protocol, the M2 antagonist methoctramine (10 RM) Was applied for 10 min prior to, as well as during, treatment with 1 mM BCh. 88 - MeHg - MeHg + BCl 100 — 80 .23? =1 60 7% > 40 °\° 20 o HBS o 0.2 0.5 1.( lMeHglmM) Figure 3.2. 24 hr pretreatment with 1 mM BCh protects against MeHg-i1 d ea th. Granule cells exhibit a concentration-dependent decrease in survival a post-exposure to increasing concentrations of MeHg. Cell death is attenuated pre—exposure to 1 mM BCh. Results are presented as meaniSEM (n=4). (*)5 difl‘erent fi'om MeHg-only pair (p < 0.05). 89 - MeHg - MeHg+BCh+AP % Vi bili HBS 0 0.2 0.5 1.0 [MeHg] (uM) Figure 3.3. Atropine plus BCh does not protect against MeHg-induced cell death. In contrast to the protection afforded by BCh alone (Figure 3.2), application of 10 uM atI-opine for 10 min prior to, as well as during, the 1 mM BCh pre-exposure period is not protective against cell death caused by 0.2-1.0 uM MeHg. Results are presented as meamiSEM (n = 6). 9O However, application of BCh and methoctramine alone for 24 hr resulted in a significant increase in the incidence of cell death, thus preventing examination of whether the protective effect of BCh was due to action at the M2 receptors (data not shown). Application of the M3 receptor antagonist 4-DAMP (10 uM) did prevent any protection by 1 mM BCh in a manner similar to atropine (Figure 3.4). Again, the lack of protection was complete at lower MeHg concentrations (0.2 and 0.5 uM) and partial at the higher MeHg concentration (1 uM). The actions of BCh were through muscarinic ACh receptors and not nicotinic ACh receptors, as the nicotinic receptor antagonist DHE (10 uM) did not affect the neuroprotective effects of BCh (Figure 3.5). The 24 hr BCh treatment down-regulated both the M3 receptor and the IP3- receptor, thus the next experiment sought to determine whether preventing MeHg from interacting with the M3 receptor would protect the cells fiom dying. To this end, cells were exposed to atropine alone (10 uM) for 10 min prior to, as well as during, exposure to MeHg. Cell viability was again measured at 24.5 h post-exposure. Unlike the 24 hr BCh pretreatment, atropine alone did not protect against MeHg-induced cell death (Figure 3.6). This is in contrast to its ability to decrease significantly MeHg-induced elevations in [Ca2+]; cerebellar granule neurons (previous chapter). Because atropine alone was not neuroprotective, the next step was to block the 1P 3-receptor during MeHg exposure to determine whether MeHg interacts directly with this receptor in granule cells. The reversible IP3-receptor inhibitor XeC (1 uM) was applied for 10 min prior to, as well as during, exposure to MeHg. As seen in Figure 3.7, acute application of XeC also did not prevent MeHg-induced cell death. Pretreatment for 24 hr with 1 uM XeC, which would presumably down-regulate 91 - MeHg - MeHg+BCh+4D 100 - 80 - A? a 60 23 > 40 g 20 0 HBS O 0.2 0.5 1.0 [MeHg] (HM) Figure 3.4. 4-DAMP plus BCh does not protect against MeHg-induced cell death. In contrast to the protection afforded by BCh alone (Figure 3.2), application of the M3- specific antagonist 4-DAMP for 10 min prior to, as well as during, the 1 mM BCh pre- exposure period does not protect against MeHg-induced cell death at 24 hr post-exposure. Results are presented as meaniSEM (n=4). 92 - MeHg - MeHg+BCh+DHE 100 ‘ % Viability HBS O 0.2 0.5 1.0 [MeHg] (HM) Figure 3.5. The nicotinic receptor antagonist DHE does not affect BCh-mediated protection against MeHg-induced cell death. The protective effect of 1 mM BCh was not due to action at nicotinic ACh receptors, because inclusion of the nicotinic receptor antagonist DHE (10 uM) for 10 min prior to, as well as during, MeHg exposure did not alter the effect of BCh. Results are presented as meaniSEM (n=8). (*)Significantly different from MeHg-only control (p < 0.05). 93 - MeHg - MeHg+Atropine 100 ‘ % Viability A O\ OO O O O N O O HBS O 0.2 0.5 1.0 [MeHg] (HM) Figure 3.6. Atropine alone does not protect against MeHg-induced cell death. A brief pretreatment with atropine (10 min, 10 uM), as well as treatment with atropine during MeHg exposure, did not affect MeHg-induced cell death at 24.5 h post-exposure. Results are presented as meaniSEM (n=6). 94 - MeHg - MeHg+XeC 100 ‘ 80 6O 40 20 O % Viabili HBS 0 0.2 0.5 1.0 [MeHg] (HM) Figure 3.7. Xestospongin C does not protect against MeHg-induced cell death. Pretreatment with xestosponginc C (XeC) (10 min, 1 uM), as well as inclusion of XeC in the MeHg solution, did not affect MeHg-induced cell death at 24.5 h post-exposure. Results are presented as meaniSEM (n=4). 95 the 1P3 receptors, was also not neuroprotective and was itself neurotoxic (data not shown). Finally, the participation of ryanodine receptors in MeHg-induced neuronal death was examined. Previous results indicated that inhibition of Ca2+ release through the ryanodine receptors did not affect MeHg-induced increases of [Ca2+]i in granule cells (previous chapter). To test whether ryanodine receptors are involved in MeHg-induced cell death, granule cells were exposed to 10 pM ryanodine for 10 min prior to, as well as during, exposure to MeHg, and again cell viability was determined at 24.5 h post- exposure. In results similar to those of the Ca2+ imaging experiments, ryanodine did not protect against MeHg-induced neuronal death (Figure 3 .8). 96 - MeHg - MeHg + Ry 100- 3‘ 80 :3 60 .53 > 40 O °\ 20 0 HBS O 0.2 0.5 1.0 MeHg (11M) Figure 3.8. Ryanodine does not protect against MeHg-induced cell death. Application of 10 uM ryanodine for 10 min prior to, as well as during, exposure to MeHg did not protect against cell death at 24.5 hr post-exposure to 0.2, 0.5 or 1.0 uM MeHg. Results are presented as meaniSEM (n=4). 97 DISCUSSION The purpose of these experiments was to elucidate the mechanism by which MeHg causes neuronal death in cerebellar granule cells which are highly sensitive to MeHg. The experiments focused on Ca2+ signaling pathways which converged at the SER Ca2+ pool, as this Ca2+i store has recently been implicated in MeHg-induced loss of [Ca2+]; homeostasis in granule cells (previous chapter). While there has been evidence for MeHg interacting with various components of this signaling network, including muscarinic receptors, 1P3 receptors and ryanodine receptors (Hare and Atchison, 1995; Castoldi et al., 1996; Candura et al., 1997; Coccini et al., 2000), this is the first report in which cytotoxicity in cerebellar granule cells resulting fi'om specific interactions of MeHg with these receptors is examined. In the first set of experiments, the SERCA inhibitor thapsigargin was used to examine the effect of depleting the SER of its Ca2+ content prior to MeHg exposure. Not surprisingly, thapsigargin treatment was itself neurotoxic, suggesting that disruption of SER Ca2+ regulation can induce neuronal death in granule cells. Thapsigargin-induced cell death could be due to a thapsigargin-induced increase in cytosolic [Ca2+] or induction of a SER “stress” response (Mattson et al., 2000); however, the mechanism of thapsigargin-induced cell death was not explored further. In the next set of experiments, 24 hr BCh was used to down-regulate the muscarinic and 1P3 receptors for 24 hr prior to MeHg exposure (F ohrrnan et al., 1993; F ukamauchi et al., 1993; Simpson et al., 1994). Similar treatment with 1 mM carbachol is well documented as causing down-regulation and desensitization of both the M3 and 1P3 receptors in rat cerebellar granule neurons (Fohrman et al., 1993; F ukamauchi et al. , 98 1993; Wojcikiewicz et al., 1994; Simpson et al. , 1996; Fukamauchi et al., 2000). As described in the previous chapter, 24 hr treatment with 1 mM BCh caused loss of granule cell responsiveness to subsequent muscarinic receptor stimulation by 1 mM BCh, suggesting that BCh has the same effect as carbachol. BCh was chosen based on its higher affinity for muscarinic receptors and lesser interaction with nicotinic ACh receptors than carbachol. In this study, 24 hr pretreatment with 1 mM BCh provided significant protection against MeHg-induced neuronal death at 24.5 hr post-exposure, indicating the importance of this pathway in the neurotoxicity of MeH g. Experiments using muscarinic receptor-specific antagonists suggest that the protective effect of BCh is mediated through the M3 receptor, asthe protection provided by BCh is reversed by the M3 -prefering antagonist 4-DAMP. The protection is also not mediated through an effect on nicotinic ACh receptors, as it is not reversed by DHE. However, the contribution of M2 receptors was not determined due to the toxicity of the M2 antagonist methoctramine itself and cannot be ruled out by these experiments. Experiments were then performed in an attempt to differentiate between MeHg interactions with muscarinic receptors and with the 1P3 receptor. Atropine alone did not protect against MeHg-induced neurotoxicity, which was surprising given the known interactions of MeHg with muscarinic receptors. As described in Chapter Two, atropine delayed the onset of MeHg-induced elevations of [Ca2+]i. Additionally, both in vivo and in vitro studies suggest that MeHg binds to multiple subtypes of Cholinergic receptors, inhibiting radioligand binding to both nicotinic and muscarinic receptors in a competitive fashion (Zhang and Wu, 1994; Castoldi et al., 1996; Coccini et al., 2000). At two weeks after a 16 day exposure to low concentrations of MeHg (0.5-2.0 mg/kg/day), rats exhibit 99 an increase in the number of muscarinic receptors without a change in receptor affinity in both the cerebellum and hippocampus, suggesting receptor up-regulation to compensate for loss of functional receptors via irreversible binding by MeHg (Coccini et al., 2000). Autoradiography of MeHg accumulation in mice indicates that MeHg accumulates preferentially in these two brain regions (Berlin and Ullberg, 1963), suggesting a possible correlation between the selective neurotoxicity of MeHg within specific areas of the CNS, and location of specific muscarinic receptor subtypes. However, the results presented here suggest that the binding of MeHg to muscarinic receptors does not explain the selective targeting of neurons during toxicity, as block of these receptors with atropine did not protect against cell death. Alternatively, MeHg may have higher affinity for muscarinic receptors than does atropine, resulting in displacement of atropine by MeHg and subsequent promotion of neurotoxicity. MeHg competes with M1 - and M2- prefering antagonists in rat cortical membranes, and demonstrates higher affinity for the M1 receptor (ICso = 3.4 uM) vs. the M2 receptor (IC50 = 149 uM) (Castoldi et al., 1996). MeHg also binds to cholinergic (presumably nicotinic) receptors isolated fiom Torpedo with high affinity and blocks binding of [3H]acetylcholine to the receptor (Eldefrawi et al., 1977). In guinea pig intestinal smooth muscle, MeHg inhibits contractions induced by stimulation of cholinergic nerves or by external application of ACh, suggesting a high affmity for muscarinic receptors (Fukushi and Wakui, 1985). Because MeHg cannot be removed from cholinergic receptors by mere wash with a MeHg-free solution, it is very likely that MeHg has higher affinity for muscarinic receptors than does atropine; however, the ability of MeHg to replace atropine on the muscarinic receptor has not yet been examined directly. 100 Attempts to block the 1P3 receptor with XeC also did not prevent MeHg-induced cell death. Given the protective effect of BCh and its reversibility by atropine, this result was also surprising. The 1P3 receptor has been implicated in MeHg-induced loss of [Ca2+], homeostasis in NGlO8-15 neuroblastoma cells (Hare and Atchison, 1995), T cells (Tan et al., 1993), and cerebellar granule neurons (Sarafian, 1993); previous chapter. Indeed, T cells that are deficient in 1P3 receptors are resistant to apoptosis induced by a number of agents (Jayaraman and Marks, 1997; Marks, 1997). In cerebellar granule cells in culture, neurons which have a relative deficiency in 1P3 receptors are resistant to low- [K+]- induced apoptosis (Oberdorf et a1. , 1997), suggesting that this receptor plays a critical role in granule cell death. Given that MeHg does increase [IP3], and that down- regulation of the M3 and 1P3 receptors protects against MeHg-induced neuronal death, the lack of protection afforded by xestospongin C may reflect a higher affinity for the 1P3 receptor by MeHg. Alternatively, the results suggest that MeHg-induced Ca2+ release through the 1P3 receptor does not contribute to MeHg-induced neurotoxicity, although this conclusion is not supported by the 24 hr BCh data. Inhibition of Ca2+ release through the ryanodine receptor did not protect against MeHg-induced cell death. This was not surprising, given that ryanodine receptors contribute only a fraction of the [Ca2+]i increase observed in granule cells following MeHg exposure (previous chapter). Ryanodine receptors also do not contribute significantly to MeHg-induced [Ca2+]t increases in NG108-15 cells (Hare and Atchison, l 995), suggesting that these receptors do not play a significant role in MeHg-induced neurotoxicity. Most likely, their primary contribution is through participation in Ca2+- induced Cay-release originating in IP3 receptors and/or mitochondria. 101 In summary, the experiments described in Chapters Two and Three indicate that MeHg causes Ca2+ release from the SER which contributes to MeHg-induced cell death of rat cerebellar granule cells. While the SER does not contribute the majority of the Ca2+i released during exposure, inhibition of this Ca2+ release is neuroprotective. As will be discussed in greater detail in Chapter Pour, the real significance of the release of SER Ca2+ may be in its effects on mitochondria by contributing to mitochondrial Ca2+ dysregulation and thereby promoting of cell death through mitochondrial dysfrmction. 102 CHAPTER FOUR DISRUPTION OF MITOCHONDRIAL CALCIUM SIGNIFICANTLY CONTRIBUTES TO METHYLMERCURY-INDUCED ELEVATIONS OF [Cay]: 103 ABSTRACT The contribution of Ca2+.“ to whole-cell [Ca2+]i elevations caused by acute exposure to low concentrations of methylrnercury (MeHg) was examined in rat cerebellar granule cells. In granule cells loaded with fura-2 to observe changes in whole-cell [Ca2+]i, prior depletion of Ca2+... using CCCP and oligomycin significantly decreased the amplitude of [Ca2+]i release from intracellular stores, as well as delayed the time-to-onset of whole-cell [Ca2+]i elevations, caused by 0.5 uM MeHg. Oligomycin delayed the MeHg-induced influx of Ca2+c, suggesting that ATP depletion triggers Ca2+ influx. In granule cells loaded with rhod-2, MeHg exposure caused a biphasic increase in fluorescence. The initial increase in fluorescence was not dependent on Ca2+¢ and was abolished by mitochondrial depolarization, indicating uptake of Ca2+ fi'om a non- mitochondrial Ca”; source. The secondary increase was associated with spreading of the dye fi'om a punctate staining to whole-cell distribution, and was significantly delayed by thapsigargin and CsA, suggesting that Ca2+ released from the SER contributes to MeHg- induced opening of the MTP. It is concluded that MeHg causes an initial release of Ca from the SER which is buffered by mitochondria, leading to elevated [Ca2+]m and opening of the MTP in rat cerebellar granule neurons. 104 INTRODUCTION As discussed in the previous chapters, acute exposure to low concentrations of MeHg causes release of Ca2+ from at least one intracellular store into the cytosol. Depletion of SER Ca2+ content reduced the amplitude of the MeHg-induced Ca2+i release by less than half; thus, at least one other intracellular store contributes to the MeHg- induced [Ca2*]i elevations. The next two chapters discuss experiments designed to examine the contribution of Ca2+.“ to MeHg-induced increases in [Ca2+]i and cell death. Mitochondria are “excitable organelles” that can accumulate a large amount of Ca2+ within the lumen and, upon excess accumulation, release Ca2+m into the cytosol via reversal of the uniporter and/or opening of the MTP (N icholls, 1978; Thayer and Miller, 1990; Budd and Nicholls, 1996b). Not only does disruption of Ca2+In contribute to opening of the MTP and release of pro-apoptotic factors (such as cytochrome c) into the cytosol, but release of sequestered Ca2+ contributes to cytosolic elevations of {Cay} which promotes cell death through activation of Cay-dependent proteases (calpains and caspases), phospholipases and endonucleases which degrade cellular protein and DNA (Cohen and Duke, 1984; Orrenius and Nicotera, 1994; Waring and Sjaarda, 1995; Moran et al., 1999). Recent studies indicate that elevation of Ca2+“. is a critical step during a variety of neurotoxic insults, including glutamate excitotoxicity, generation of reactive oxygen species, and exposure to heavy metals, including manganese and mercury (Chavez and Holguin, 1988; Chavez et al., 1989; Richter and Kass, 1991; Nicholls and Budd, 1998; Gavin et al., 1999; Kruman and Mattson, 1999; Verity, 1999) Several parameters of mitochondrial fimction are affected by MeHg. MeHg binds to mitochondria with high affinity, both in vivo and in vitro (Yoshino et al., 1966; Chang 105 and Hartmann, 1972; O'Kusky, 1983). Further, MeHg causes irreversible loss of mitochondrial membrane potential in human monocytes (InSug et al. , 1997), isolated nerve terminals (Bondy and McKee, 1991), NG108-15 neuroblastoma cells (Hare et al., 1993), and guinea pig cerebral cortex synaptosomes (Kauppinen et al., 1989). Inhibition of Ca2+... uptake and release inhibits MeHg-induced increases in the frequency of spontaneous neurotransmitter release in hemidiaphragm nerve terminals (Levesque and Atchison, 1988), and attenuates the uptake of 45Ca and release of 45Ca2+ fi'om preloaded, isolated mitochondria Ca2+ uptake (Levesque and Atchison, 1991). MeHg is also well documented as inhibiting mitochondrial respiration, decreasing cellular ATP content and causing production of reactive oxygen species, all of which can contribute to mitochondria-triggered cell death (Kauppinen et al., 1989; Sarafian et al., 1989; Levesque and Atchison, 1991; Sarafian and Verity, 1991; Yee and Choi, 1996). Because the Ca2+; chelator BAPTA protects against MeHg-induced cell death in rat granule cells in vitro, suggesting a Ca2+-dependent process (Marty and Atchison, 1998), there is a strong possibility that MeHg-induced changes in Ca2+m may be critical in MeHg-induced neuronal death. The experiments described in this chapter examined whether mitochondria contribute Ca2+m to the first-phase [Ca2+]i increase during acute exposure to low concentrations of MeHg. Because alterations in cellular and mitochondrial Ca2+ regulation are critical events in both apoptosis and necrosis, the experiments were designed to characterize the effects of MeHg on Ca2+", in the intact neuron. The first approach was to determine the relative contribution of Ca2+m to whole-cell elevations of [Caz+]t to determine whether mitochondria are a significant target in these cells. The next 106 experiments examined the effect of MeHg on regulation of Ca”... in the intact cell, including interactions between intracellular Ca2+ stores in the mitochondria and SER during MeHg exposure. The results presented here suggest that MeHg targets both the mitochondria and SER during MeHg exposure, with Ca2+ release from the SER contributing to Ca2+ release from the mitochondria via opening of the MTP. Further, these experiments indicate that Ca2+m contributes the bulk of the cytosolic [Ca2+]i observed during MeHg exposure, highlighting the importance of mitochondria during MeHg neurotoxicity. 107 on; MATERIALS AND METHODS The materials and solutions for isolation of cerebellar granule cells and measurement of changes in [Ca2+]i are described in detail in Chapter Two. Unless specifically stated, all solutions were identical to those described in Chapter Two. The methods and solutions described below are specific to this chapter. Materials and solutions. Rhod-2 acetoxymethylester (rhod-2 AM) and tetramethylrhodamine ethyl ester (TMRE) were purchased fiom Molecular Probes (Eugene, OR). Cyclosporin A (CsA), CCCP and oligomycin were purchased from Sigma (St. Louis, MO). For all experiments, pharmacological agents are dissolved in either HBS or EGTA-HBS, with appropriate controls for any additional solvents used (DMSO or ethanol, with maximum solvent of 0.01% (w/v) for all experiments). In a cell-free system, none of the pharmacological agents altered fura-2 fluorescence. Measurement of [Ca2+]i. Cells were loaded with fura-2 for 1 hr, and rinsed in HBS for 30 min, as described in Chapter Two. Changes in cellular fluorescence were measured using the IonOptix system described in Chapter Two. As before, following exposure to MeHg, the times-to-onset of the first and second phases of increased Ca2+, were determined. Additionally, in experiments performed in the nominally Ca2+-free buffer (EGTA-HBS), the amplitude of [Caz] release caused by MeHg was normalized to the peak fluorescence ratio amplitude of the 1 min exposure to 40 mM K+ to provide a “normalized ratio” relative to K+-induced depolarization response. For both time-to- onset and normalized fluorescence ratio, the data fiom each trial were averaged to 108 provide a mean time-to-onset or mean normalized ratio for that dish of cells (n=1). In order to minimize differences between cell isolates, experiments using MeHg alone and MeHg with a given pharmacological agent were run on the same day, and experiments using the same agents were replicated using at least two separate cell isolates. Measurement of [Ca2+]... To measure changes in [Ca2+]m, cells were loaded with 1 uM rhod—2 AM for 1 hr at room temperature (23-25°), followed by overnight incubation in normal culture medium (18-24 hr) (Trost and Lemasters, 1997; Peng et al., 1998; Simpson and Russell, 1998a). Because rhod-2 will fluoresce in both the cytosol and mitochondria, this protocol was used to ensure that mitochondrial loading occurred and provide time for the cytosolic dye to leak out of the cell. Cells were imaged as in the fura-2 experiments, except that the excitation wavelength was 540 nm and the emission wavelength was 590 nm. Efficiency of mitochondrial loading was confirmed using an 8 min exposure to 5 uM CCCP to uncouple oxidative phosphorylation, and 10 pM oligomycin to dissipate \ym; this combination prevents mitochondrial Ca2+ uptake during Ca2+ influx caused by K+ depolarization (40 mM), thereby abolishing changes in rhod-2 fluorescence. Like fura-2, rhod-2 data are analyzed for time-to-onset of fluorescence changes, with the time-to-onset of the initial elevation determined as the point at which fluorescence rose above baseline, and the secondary elevation (from spreading of rhod-2 fluorescence fi'om a punctate stain to whole-cell fluorescence) as the point at which there is a sharp increase in rhod-2 fluorescence. The times-to-onset of each phase of fluorescence were calculated for all cells monitored within an experiment and were averaged to provide mean time-to-onset for that dish of cells (n=1). 109 Measurement of ‘1'... The fluorescent dye tetramethylrhodamine ethyl ester (TMRE) was used to observe changes in ‘I’m. TMRE preferentially loads into mitochondria with normal membrane potentials (around —160 mV) and loses fluorescence when ‘Pm becomes depolarized (F arkas et al. , 1989). Granule cells were loaded with 100 nM TMRE in HBS for 30 min at 37°C, then rinsed in 10 nM TMRE in HBS for 30 rrrin. Cellular fluorescence was then measured continuously using the IonOptix system. In these experiments, perfusion solutions contained 10 nM TMRE to maintain dye availability for the mitochondria. Fluorescence was measured relative to an initial fluorescence of 100%. Control experiments indicated that TMRE fluorescence remained stable for at least 1 hr of observation (data not shown). Statistics Comparisons of mean time-to-onset for “MeHg” vs. the corresponding “MeHg plus inhibitor” cells were made using Student’s paired t-test, with values of p<0.05 considered to be statistically significant. 110 RESULTS Neonatal granule cells in culture exposed to a low concentration of MeHg undergo a characteristic biphasic increase in [Ca2+]i, with the first phase due to release of Ca2+ fi'om at least one intracellular store, and the second phase due to influx of Ca2+: (Marty and Atchison, 1997). To examine the contribution of the mitochondrial Ca2+ store, the mitochondria were depolarized with 5 uM CCCP and 10 uM oligomycin for 8 min immediately prior to exposure to 0.5 uM MeHg. Because mitochondria accumulate Ca2+ only when their inner membranes are sufiiciently hyperpolarized, this depolarization causes both release of Ca2+“. and prevention of further Ca2+ uptake (Budd and Nicholls, 1996b). Monitoring cells using TMRE fluorescence confirmed that this protocol was sufficient to depolarize completely the mitochondria (Figure 4.1A). The brief increase in TMRE fluorescence intensity is probably due to mitochondrial swelling, rather than mitochondrial membrane hyperpolarization, which accompanies the increased permeability of the inner mitochondrial membrane and subsequent increase in lumen volume (White and Reynolds 1996; Trost and Lemasters 1997). Further, experiments using cells loaded with fura-2 indicated that this protocol caused a large increase in [Ca2+]; (Figure 4.1B). Following 8 min exposure to CCCP plus oligomycin, the cells did not respond to another application of CCCP plus oligomycin, indicating the irreversible loss of mitochondrial membrane potential, and successful depletion of Ca2+m, with this protocol (Figure 4.1C). Treatment with 10 uM oligomycin alone caused a more gradual loss of mitochondrial membrane potential, with a 75-85% loss of TMRE fluorescence at 20 min of exposure; TMRE fluorescence could be lowered further with 5 uM CCCP, indicating that oligomycin alone did not cause complete mitochondrial membrane 111 CCCP+Oligo A a 5 e’ (U v 0 O a d) O U) Q) l— O .2 LL Time (min) B o 60 fé Ki ED 0 40 CCCP+Olrgo .9. H N at 0.20 a 1 a 4 o 5 to 15 20 25 Time (min) C 0.60 - as CCCP+OIigo CCCP+Oligo '3 0.50 — 3. K W v o E. 0.40 0.30 l l l 0 8 16 24 32 Time (min) Figure 4.1. CCCP and oligomycin cause a rapid decrease in TMRE fluorescence with a concomitant increase in fora-2 fluorescence. A) Changes in TMRE fluorescence in a single cerebellar granule neuron in culture exposed to 5 uM CCCP and 10 uM oligomycin for 8 min. CCCP and oligomycin cause a rapid decrease in TMRE fluorescence (as the mitochondrial membrane becomes depolarized). In granule cells, the rapid loss of Van caused by CCCP and oligomycin is irreversible for up to 60 min post- treatrnent. TMRE fluorescence is measured relative to the initial fluorescence intensity. B) Changes in fura-2 fluorescence (“Ratio (F 340330)”) in a single cerebellar granule neuron exposed to 5 pM CCCP and 10 uM oligomycin for 8 min. Mitochondrial membrane depolarization causes a large-amplitude increase in [Ca2+]i which recovers to baseline [Ca2+]. by the end of the 8 min exposure. C) Changes in fura-2 fluorescence (“Ratio (F 340330)”) in a single cerebellar granule neuron exposed to 5 uM CCCP and 10 pM oligomycin for 8 min, followed by 10 min HBS and another 8 min application of 5 uM CCCP and 10 uM oligomycin. The second exposure to these compounds does not result in additional increases in fura-2 ratio, indicating that a single application is sufficient to deplete Ca2+m. 112 'v. o Oligomycin é .0 m o Fluorescence (arb. units) 8 20 10 o 30 40 so 60 Time (min) A g 1.50 — CCCP C‘. :3 .23 3 1.00 v 8 5 o 0.50 t (D 23. S _ 000 r r r r ‘1‘ o 5 10 15 20 Time (min) C 0.60 r A K‘ § § 0.40 ‘ CCCP ‘1‘!“ v 0 0m 1 L L I o s 10 15 20 Time (min) Figure 4.2 Effect of oligomycin on TMRE fluorescence, as well as effect of CCCP on TMRE and fora-2 fluorescence. A) Changes in TMRE fluorescence in a single cerebellar granule neuron in culture exposed to 10 pM oligomycin for 45 min. Oligomycin causes a gradual loss of TMRE fluorescence which is further lowered by the addition of 5 uM CCCP, indicating that oligomycin alone does not completely depolarize the mitochondrial membrane. TMRE fluorescence is measured relative to the initial fluorescence intensity. B) Changes in TMRE fluorescence in a single cerebellar granule neuron in culture exposed to 5 uM CCCP for 8 min. CCCP alone causes a rapid decrease in TMRE fluorescence which does not recover during wash with CCCP-free HBS. TMRE fluorescence is measured relative to the initial fluorescence intensity. C) Changes in fora-2 fluorescence (“Ratio (F34o,3go)”) in a single cerebellar granule neuron exposed to 5 uM CCCP for 8 min. CCCP alone causes a much smaller increase in [Ca2+]; than CCCP and oligomycin together (see Figure 4.1B). 113 depolarization (Figure 4.2A). Application of 5 uM CCCP for 8 min did cause rapid, complete loss of TMRE fluorescence which did not recover after removal of CCCP and wash with HBS, indicating that CCCP alone causes rapid depolarization of the mitochondrial membrane (Figure 4.2B). However, unlike CCCP plus oligomycin, CCCP alone caused a much smaller increase in cytosolic [Ca2+] with an amplitude approximately half that caused by CCCP plus oligomycin (Figure 4.2C). Next, the effect of depleting mitochondrial Ca2+ stores on MeHg-induced elevations of [Ca2+]; was examined. Complete removal of Ca2+m using 5 uM CCCP plus 10 uM oligomycin for 8 min delayed the time-to-onset of increases in [Ca2+]; caused by 0.5 uM MeHg (from 12.6 d: 2.5 min to 37.7 i 3.0 min) (Figure 4.3). The time-to-onset of the first-phase increase in [Ca2+]; was not significantly altered by CCCP alone (11.5 i 1.4 min vs. 10.7 :t 1.7 min) or oligomycin alone (9.8 :t 1.9 rrrin vs. 17.4 :1: 5.8 min) (Figure 4.3). Similarly, emptying of the mitochondria with CCCP and oligomycin prior to exposure to 0.5 uM MeHg in EGTA-HBS revealed a similar delay in the increased [Ca2+]i, delaying the onset of the first-phase [Ca2+]; increase fiom 15 .3i3.2 to 45.4:l:3.9 min (Figure 4.4A). Additionally, CCCP and oligomycin treatment significantly reduced the amplitude of the fura-2 ratio increase from a normalized ratio of 1.252t0.26 to 0.41:1:0.13, representing a 67% decrease in the ratio amplitude (Figure 4.4B). Thus complete removal of Ca2+,“ is required to delay significantly the MeHg-induced release of Ca2+ from intracellular pools, but does not prevent an eventual increase in [Ca2+]i. Additionally, this data indicates that most of the Cay, released by MeHg originated in the mitochondria. The mitochondrial inhibitors had differing effects on the tirne-to-onset of the 114 - MeHg only - Inhibitor + MeHg E 50 " E 40 ES Cg) 3O é. 20 a5 10 .§ F-i . . CCCP Olrgo CCCP+ Olrgo Treatment Figure 4.3. Effect of mitochondrial inhibitors on first-phase increase in [01”]; caused by 0.5 pM MeHg. Granule cells were pretreated with 5 uM CCCP (8 min), 10 11M oligomycin (8 min), or 5 uM CCCP plus 10 pM oligomycin (8 min) immediately prior to 0.5 uM MeHg. Time-to-onset of the [Ca2+]; increase was compared between MeHg alone and MeHg plus inhibitor for each treatment. Results are presented as meaniSEM (n=5-8). The asterisk (*) indicates a value significantly different fi'om MeHg-only control (p<0.05). 115 - VH i; if CCCP + oligo A 50 * .S - E 3 U) c C? ,9. a') .§ H lst Phase B - MeHg only -a C+O+MeHg Normalized Ratio ~ N O Figure 4.4. Effect of CCCP and oligomycin on first-phase time-to-onset and ratio amplitude induced by 0.5 pM MeHg in the absence of Cat“, A) Measurement of time-to-onset of [Cay]; elevations caused by 0.5 uM MeHg in the absence of Ca2+, Cells were preheated with vehicle control (“VH”) or 5 uM CCCP plus 10 uM oligomycin (“CCCP+oligo”) for 8 min immediately prior to 0.5 uM MeHg. B) Measurement of the first-phase ratio amplitude in the absence of C322. Cells were exposed to 0.5 pM MeHg alone (“MeHg”) or 5 uM CCCP plus 10 uM oligomycin (8 min) immediately prior to 0.5 uM MeHg (“C+O”). The amplitude of the first-phase ratio increase was normalized to the peak fura-2 response to the 1 min 40 mM K+ test (at the beginning of each experiment). Results are presented as mearttSEM. (n=5). The asterisk (*) indicates a value significantly different from MeHg-only control (p<0.05). 116 - MeHg only - Inhibitor + MeHg Q o 1 * 4.}. (ll t-a 1J1 Time-to-Onset (min) DJ O CCCP Oligo CCCP+ Oligo Treatment Figure 4.5. Comparison of time-to-onset of Ca2+c influx caused by 0.5 uM MeHg in the presence or absence of mitochondrial inhibitors. Granule cells were pretreated with 5 uM CCCP (8 min), 10 pM oligomycin (8 min), or 5 uM CCCP plus 10 uM oligomycin (8 min) immediately prior to 0.5 uM MeHg. Comparisons are made between MeHg alone and MeHg plus inhibitor for each treatment. Results are presented as meaniSEM (n=5-8). The asterisk (*) indicates a value significantly different from MeHg-only control (p<0.05). 117 influx of Ca2+': caused by 0.5 uM MeHg. Mitochondrial membrane depolarization without maintenance of cellular ATP levels using CCCP alone hastened the influx of Caz“c caused by MeHg, from 22.0 d: 3.3 min to 15.6 i: 2.7 min (Figure 4.5). This Ca2+ influx may occur independently of any effects of MeHg, as mitochondrial membrane depolarization causes reversal of the mitochondrial ATPase, leading to conversion of ATP to ADP and depletion of ATP, thereby causing Ca2+ influx through loss of ATP for plasma membrane Ca2+ pmnps (N icholls and Scott, 1980). Oligomycin alone had the opposite effect, causing a significant delay in the time-to—onset of MeHg-induced Ca2+. influx, fi'om 17.8 i 3.7 min to 42.7 :1: 9.7 min (Figure 4.5). Because oligomycin does not completely depolarize the mitochondria by 8 min but does maintain cellular ATP levels, this delayed influx of Ca2+. suggests that MeHg-induced depletion of cellular ATP triggers onset of the second-phase increase in [Ca2+]i. Application of CCCP plus oligomycin caused a similar delay in the time-to-onset of the MeHg-induced Ca2+ influx as that caused by oligomycin alone, from 22 :1: 3.4 min to 36.3 at 8.3 min (Figure 4.5). Thus, MeHg-induced influx of Ca2+e appears to be dependent on loss of cellular ATP content as a result of MeHg-induced alteration of mitochondrial function. The next step was to examine the route by which MeHg causes uptake of Ca2+ into the mitochondria, and subsequent release of Ca2+", from the mitochondria into the cytosol. As described in Chapter Two, previous experiments suggested that MeHg causes release of Ca2+ from the SER via activation of muscarinic receptors linked to production of 1P3. Based on these results, it was hypothesized that MeHg caused Ca2+ to be released from the SER, which was then buffered by mitochondria. To test this hypothesis, the effect of MeHg on Ca2+,“ was examined using the fluorescent Ca2+,“ 118 indicator rhod—2. Cells were first loaded with fura-2 and exposed to increasing concentration of K+ (IO-100 mM) for 1 min each, to demonstrate that this protocol causes reversible increase in [Caz] which exhibit increasing fura-2 ratio amplitude as the [K+] increases (Figure 4.6A). Granule cells were then loaded with rhod-2 to monitor changes in [Ca2+]m. Control experiments using 1 min depolarizations with increasing concentrations of K+ (10- 1 00 mM) indicated that rhod-2 fluorescence increases during an increase in [Ca2+]; and returns to baseline after removal of the K+ solution (Figure 4.68). Further, this increase in fluorescence was due to a change in Ca”... and not due to cytosolic loading of the dye as the increase in rhod-2 fluorescence caused by 1 min treatment with 40 mM K+ was abolished by 8 min exposure to CCCP and oligomycin (Figure 4.6B). As seen in Figure 4.7, granule cells loaded with rhod-2 exhibit a biphasic increase in rhod-2 fluorescence during exposure to 0.5 uM MeHg. Rhod-2 fluorescence increases to a similar intensity as that caused by 1 min depolarization with 40 mM K“ (Figure 7A); this initial increase is abolished by 8 rrrin exposure to CCCP and oligomycin, indicating that the initial increase in fluorescence is due to an increase in Ca2+“, (Figure 7B). The secondary increase in fluorescence is associated with a spreading of the dye from the mitochondria to the cytosol which was apparent both in the digital image recorded by the IonOptix software, and in the fluorescence emission recording. The dye movement was associated with an increase in rhod-2 fluorescence intensity above that caused by the 1 min 40 mM K+ depolarization; thus we measured the secondary fluorescence increase as the point at which rhod-2 fluorescence became elevated above that of the 40 K+-induced depolarization. Interestingly, treatment with CCCP and oligomycin abolished the initial 119 1.40 - r-xo i 00 g 100 3‘ 50 E"; 070 . 40 l .g 10 20 l e ed 0.00 l l L J 1 0 10 20 30 40 50 Time (min) ”a? B E 6 - 100 40 D . 50 l c+o .o 40 —— 2 2° 1 <1) 4 r i ' O c: Q) 0 U) <1) H o l L l 2 2 L l 1 LL. 0 10 20 30 40 50 60 Time (min) Figure 4.6. Changes in fura-2 and rhod-2 fluorescence in response to increasing concentrations of extracellular 1C. A) Representative example of changes in fura-2 fluorescence (“Ratio (Fm/330)”) in a single cerebellar granule neuron exposed to increasing concentrations of K+ (IO-100 mM). Each K+ solution was applied for 1 min starting at the arrow, with the [K] (mM) indicated by the number above the arrow. B) Representative example of changes in rhod-2 fluorescence (“Fluorescence (arb. units)”) in a single cerebellar granule neuron exposed to increasing concentrations of K+ (IO-100 mM). Each K+ solution was applied for 1 min starting at the arrow, with the [K] (mM) indicated by the number above the arrow. At the end of the experiment, 5 uM CCCP and 10 uM oligomycin were applied for 8 min, followed by another 1 min treatment with 40 mM K". CCCP and oligomycin abolished rhod-2 fluorescence changes in response to the K4” depolarization, confirming that rhod-2 was localized within the mitochondria. 120 elevation of fluorescence but did not alter the time-to-onset of the secondary fluorescence increase associated with the dye spreading throughout the cell (Figure 4.9A). Because the initial increase of fluorescence did not occur until the spreading of dye throughout the cell occurs, the “first-phase” increase in rhod-2 fluorescence occurred simultaneously with the “second-phase” release of dye into the cytosol (Figure 4.8). When granule cells were exposed to 0.5 uM MeHg in the nominally Cay-free EGTA-HBS, the pattern of rhod-2 fluorescence changes was similar, as MeHg again caused an initial fluorescence increase which was abolished by CCCP and oligomycin, and a secondary fluorescence increase from dye movement into the cytosol (Figure 4.9B). Because MeHg caused an increase in Ca2+,“ in the absence of C322, it was hypothesized that the mitochondria were buffering Ca2+ that was released from another intracellular Ca2+ store, possibly the SER; thus, the next experiment examined the effect of SER Ca2+ depletion on MeHg-induced elevations of Ca2+m using the SERCA inhibitor thapsigargin (10 uM for 5 min). Thapsigargin treatment immediately prior to exposure to 0.5 uM MeHg did not affect the time-to-onset of the initial rhod-2 fluorescence increase, but did significantly delay the dye movement into the cytosol (Figure 4.10). Finally, experiments were performed to test whether the release of rhod-2 from the mitochondria into the cytosol was due to opening of the MTP. Related experiments indicated that the MTP inhibitor CsA delays the MeHg-induced increase in [Ca2+]; caused by 0.2 and 0.5, but not 1.0 uM MeHg (see Chapter Five). In granule cells loaded with rhod-2, application of 5 uM CsA for 10 min immediately before 0.5 uM MeHg did not alter the time-to-onset of the initial fluorescence increase but did significantly delay the 121 0.5 uM MeHg Onset of Initial Fluorescence Increase 0 10 20 30 4O 50 Relative Fluorescence (arb. units) U») Time (min) Figure 4.7. MeHg causes a biphasic increase in rhod-2 fluorescence in granule cells. Changes in rhod-2 fluorescence in a single neuron exposed to 0.5 uM MeHg. Cells were initially exposed to 1 mM K+ to verify cell viability (“K”). During MeHg exposure, there is a biphasic increase in rhod-2 fluorescence: the initial rise plateaus at a similar fluorescence intensity as that caused by the 1 min 40 mM K+; the secondary increase is associated with the dye spreading throughout the entire cytosol. 122 4 .— 3; 0.5 uM MeHg + oligomycin .6 E: 3 8 § 2 i E 1 33'. 0 m 0 iilrlrirrlrrlrlnnaalnrnrlrrrll 0 10 20 30 40 50 60 Time (min) Figure 4.8. Effect of CCCP and oligomycin on MeHg-induced elevations of rhod-2 fluorescence. Changes in rhod-2 fluorescence in a single neuron exposed to 5 pM CCCP and 10 uM oligomycin (“C+O”) for 8 min immediately before 0.5 11M MeHg. In these cells, the initial rise in rhod-2 fluorescence is absent; however, the secondary fluorescence increase from the dye spreading throughout the cytosol is still present. 123 A - MeHg only - C+O+MeHg 40 ' 30 20 10 Time-to-Onset (min) lst Phase 2nd Phase - MeHg only - C+O+MeHg 20" 10‘ Time-to-Onset (min) 1st Phase 2nd Phase Figure 4.9. Mitochondrial membrane depolarization abolished the initial elevation of rhod-2 fluorescence without affecting the time-to-onset of the spreading of rhod-2 throughout the cytosol. A) Granule cells were pretreated with 5 uM CCCP plus 10 uM oligomycin (8 min) immediately prior to 0.5 uM MeHg. Time-to—onset of the [Ca2+]i increase was compared between MeHg alone and MeHg plus CCCP and oligomycin. Results are presented as meaniSEM (n=4). B) Granule cells were pretreated with 5 uM CCCP plus 10 uM oligomycin (8 min) immediately prior to 0.5 uM MeHg in EGTA- HBS. Results are presented as mean i SEM (n=3). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 124 spreading of rhod-2 throughout the cell (Figure 4.11). A comparison of the effect of CsA and thapsigargin on the interval between onset of first and second rhod-2 fluorescence increases is shown in Figure 4.12. 125 - MeHg only - Thaps + MeHg AZOP .S. :3. _ H * 0) E’ 010- 5'2. Q5 . .§ E... 0 lst Phase 2nd Phase Figure 4.10. Thapsigargin delays the spreading of rhod-Z from punctate to whole- cell staining as caused by 0.5 uM MeHg. Granule cells were pretreated with 10 uM thapsigargin for 5 min immediately prior to 0.5 pM MeHg. The first-phase elevation is measured where rhod-2 fluorescence rises above baseline; the second-phase elevation is measured where the fluorescence rises above the fluorescence increase caused by the 1 min 40 mM 1C test. Results are presented as meaniSEM (n=5). The asterisk (’) indicates a value significantly different from MeHg-only control (p<0.05). 126 - MeHg only - CsA + MeHg 30 " N O l I Time-to-Onset (min) 8 lst Phase 2nd Phase Figure 4.11. CsA delays the spreading of rhod-2 from punctate to whole-cell staining as caused by 0.5 pM MeHg. Granule cells were pretreated with 5 uM CsA immediately prior to 0.5 uM MeHg. The first-phase elevation is measured where rhod-2 fluorescence rises above baseline; the second-phase elevation is measured where the fluorescence rises above the fluorescence increase caused by the l min 40 mM K+ test. Results are presented as meaniSEM (n=7). The asterisk (*) indicates a value significantly different fiom MeHg-only control (p<0.05). 127 - MeHg only - Drug + MeHg 20' Time (min) 8 CsA Thapsigargin Figure 4.12. Effect of CsA and thapsigargin on interval between onset of first and second rhod-2 fluorescence increases. The time between the first- and second-phase increases in rhod-2 fluorescence indicates the amount of time between the initial uptake of Ca2+ into the mitochondrial lumen and the release of rhod-2 into the cytosol. Results are presented as mean d: SEM (n=5-7). The asterisk (*) indicates a value significantly different from the corresponding MeHg-only data (p < 0.05). 128 DISCUSSION Disruption of [Ca2+]i is an early step in MeHg neurotoxicity which apparently contributes to MeHg-induced cell death in rat cerebellar granule neurons (Marty and Atchison, 1998). These experiments indicate that the bulk of this [Cay]; originates in the mitochondria. When the mitochondrial Ca2+ store is emptied following mitochondrial membrane depolarization with CCCP plus oligomycin, the amplitude of the whole-cell [Caz+]i elevation caused by MeHg is significantly reduced by approximately 65%. Removal of Ca2+". does not abolish the MeHg-induced release of {Cap}, indicating the participation of more than one intracellular Ca2+ source. The participation of more than one intracellular store agrees with results of experiments in T lymphocytes, in which MeHg causes release of Ca2+ from both mitochondria and SER (Tan et al., 1993). However, the first-phase increase in firm-2 ratio is not comprised solely of Ca2+, as there is at least one other divalent cation whose intracellular levels increase in response to MeHg (Denny et al., 1993; Hare et al., 1993; Marty and Atchison, 1997). In rat cortical synaptosomes, this cation was identified as Zn2+ by l9F-NMR (Denny and Atchison, 1994). Zn2+ has recently been identified as a toxic cation capable of opening the MTP during neurotoxic insult (W udarczyk et al. , 1999). Although non-Ca2+ divalent cations do not appear to contribute to MeHg-induced cell death in granule cells (Marty and Atchison, 1997), elevated Zn2+ could contribute to the opening of the MTP observed in this and other studies (Limke and Atchison, 2002). Thus, the non-mitochondrial component of the first-phase increase in fora-2 ratio is likely comprised of a combination of Ca2+ from another intracellular store, such as the SER, and the non-Ca2+ divalent cation, most likely Zn”. 129 The large contribution of Ca2+“. to the first-phase increase in [Ca2+]. in cerebellar granule cells is in contrast to the elevations of [Ca2+]; in NG108-15 cells, in which the primary source of Ca2+; was the SER through the 1P3 receptor, with minimal contribution fi'om the mitochondria (Hare and Atchison, 1995). The preference for Ca2+ release fi'om mitochondria as Opposed to release from another Ca2+ pool parallels experiments in the neuromuscular junction, in which MeHg-induced increases in spontaneous release of neurotransmitter were decreased to less than 10% of control by inhibition of Ca2+ release from the mitochondria and were not decreased by inhibitors of SER Ca2+ release (Levesque and Atchison, 1988). Motor nerve terminals may be one of the targets of MeHg during in vivo exposure, as a neuromuscular weakness resembling myasthenia gravis is observed (Rustam et al., 1975). The heightened susceptibility of neurons exhibiting significant changes in Ca2+". suggests a correlation between MeHg interaction with mitochondria and subsequent neuronal damage. Given the relatively small mitochondrial contribution to whole-cell {Caz} elevations in less sensitive neuronal models, such as NG108-15 cells (Hare et al., 1993; Hare and Atchison, 1995), these results suggest that disruption of mitochondrial Ca2+ regulation may be an important factor in determining cellular sensitivity to MeHg. The differential effects of mitochondrial inhibitors highlight the pivotal role of Ca2+,“ regulation in determining the extent of whole-cell increases of [Cay]; induced by MeHg. Mitochondrial membrane depolarization with CCCP alone did not cause significant release of Ca2+... into the cytosol. These results agree with previous experiments in granule cells (Budd and Nicholls, 1996b), in which mitochondrial membrane depolarization alone did not cause significant release of Ca2+m. Interestingly, 130 mitochondrial membrane depolarization alone did not alter the time-to-onset of the MeHg-induced release of Ca2+t. The smaller release of Ca2+,“ into the cytosol induced by CCCP alone suggests that enough Ca2+.“ remains to be released by MeHg. Application of oligomycin alone resulted in partial depolarization of mitochondrial membrane and, as an ATP synthase inhibitor, maintenance of cellular ATP levels (Budd and Nicholls, 1996b). Oligomycin alone did not cause an increase in cytosolic [Ca2+], indicating that partial mitochondrial membrane depolarization is insufficient to empty mitochondrial CaZ+ stores. This is supported by the lack of effect of oligomycin on the time-to—onset of the initial increase in [Ca2+]i during MeHg exposure. However, oligomycin alone did delay significantly the onset of MeHg-induced influx of Caz} [Ca2+]t is maintained, in part, by plasma membrane Ca2+ pumps which require ATP to remove Ca2+ from the cytosol (Werth et al., 1996; Morgans et al., 1998), thus depletion of cellular ATP content inhibits ATP-dependent extrusion of Ca2+ from the cell. This is supported by the hastening of the MeHg-induced, second-phase influx of Ca2+. caused by CCCP alone. When the mitochondrial membrane becomes depolarized, the mitochondrial ATP synthase reverses and begins hydrolyzing ATP into ADP in an attempt to maintain \th (N icholls and Scott, 1980). Experiments in cerebellar granule neurons in primary culture indicate that 5 min exposure to 2 uM CCCP rapidly depletes cellular ATP levels (Budd and Nicholls, 1996b), thus in our experiments, 8 nrin exposure to 5 uM CCCP would similarly be expected to deplete ATP. Under these conditions, CCCP-induced depletion of cellular ATP hastens the MeHg-induced depletion of ATP (Sarafian et al., 1989) and causes an earlier onset of Ca2+.3 influx. 131 Depletion of Ca2+m with CCCP and oligomycin delays and reduces the first-phase elevation of [Caz‘] but does not prevent an eventual rise in [Ca2+]t. The delayed movement of rhod-2 into the cytosol by thapsigargin indicates that MeHg causes Ca2+ release from the SER, with this Ca2+ subsequently buffered by mitochondria. There is substantial evidence that release of Ca2+ fiom the SER causes a rapid, reversible rise in Ca2+... which is a normal component of intracellular signaling (Landolfi et al., 1998; Pinton et al. , 1998; Rizzuto et al., 1998). In HeLa cells transfected with a Ca2+-sensitive photoprotein localized to the outer face of the inner mitochondrial membrane, Ca2+ released by activation of 1P3 receptors resulted in a high [Caz+] within the mitochondria (Rizzuto et al., 1998). This [Ca2+] was much higher than that observed in the bulk cytosol as a result of the close spatial relationship between mitochondria and SER in these cells (Pinton et al., 1998; Rizzuto et al., 1998). These experiments also provide evidence that MeHg causes opening of the MTP, because the rhod-2 movement from punctate to diffuse staining is delayed by the MTP inhibitor CsA. Because the rhod-2 fluorescence increase associated with opening of the MTP is also delayed by thapsigargin, this suggests that release of Ca2+ from the SER contributes to the MeHg-induced loss of Ca2+m regulation via opening of the MTP. This shifting of intracellular Ca2+ from the SER to the mitochondria explains the large amount of Ca2+“. released by MeHg in granule cells. Mitochondria are a low-affinity Ca2+ store which do not contain much Ca2+ at rest; high [Ca2+].m is energetically unfavorable and interferes with the respiratory chain and the tricarboxylic acid cycle (Budd and Nicholls, 1996b; Wang and Thayer, 1996). Because mitochondria normally contain little (332+, this movement of Ca2+ from the SER to the mitochondria underlies the large amplitude 132 release of Ca2+", during MeHg exposure. Additionally, it suggests in granule cells that the SER is in fact the initial Ca2+. target, with mitochondria being affected at a later time point. The sensitivity of mitochondria to MeHg-induced increases in Ca2+m may be dependent on the spatial relationship between SER Ca2+ release channels and nearby mitochondria. In NG108-15 cells, the primary source of Ca2+i during MeHg exposure originates in the IP3-sensitive Ca2+ pool in the SER (Hare and Atchison, 1995). The greater proportion of Ca2+ fi'om the SER in NG108-15 cells, as compared to granule cells, may be due to a lack of rrritochondrial buffering rather than less release of Ca2+ from this store. Alternatively, the apparent sensitivity of granule cell mitochondria may be due to the Ca2+ uptake and release characteristics of the mitochondria themselves. Mitochondria in different brain regions have differential sensitivities to induction of the MTP in different brain regions (F riberg et a1. , 1999). Because mitochondria are such a prominent target during MeHg exposure, the preferential sensitivity of the granule cells to MeHg- induced cell death may be related to the Ca2+ regulating characteristics of their mitochondria. 133 CHAPTERS ACUTE EXPOSURE TO METHYLMERCURY OPENS THE MITOCHONDRIAL PERMEABILITY TRANSITION PORE IN RAT CEREBELLAR GRANULE CELLS 134 ABSTRACT Following acute MeHg exposure, granule cells in culture undergo an increase in [Ca2+]; that apparently contributes to cell death. This effect consists of several temporally and kinetically distinct phases. The initial elevation arises fiom release of Ca2+; stores; the second phase results from entry of Ca2+e. These experiments tested the hypothesis that release of mitochondrial Ca2+ through the MTP contributes to the initial release of Ca2+i. Neonatal rat cerebellar granule cells in culture and single cell microfluorimetry were used to examine MeHg-induced changes in [Caz+]i and mitochondrial membrane potential (Wm). Pretreatment with the MTP inhibitor CsA (5 uM) delayed the initial phase of increased [Ca2+], induced by 0.2 and 0.5 uM MeHg, but not 1.0 uM MeHg. CsA (5 uM) also delayed the irreversible loss of ‘1’“, induced by 0.5 uM MeHg. Ca2+. was not required for either effect, because the effect of CsA on the first phase increase in [Ca2+]; and loss of ‘I’m was not altered in nominally Cay-flee buffer. Increasing concentrations of MeHg (0.2-2.0 uM) caused increasing incidence of cell death at 24 hr post-exposure. Treatment with CsA provided protection against cytotoxicity at lower MeHg concentrations (0.2-0.5 uM), but not at higher MeHg concentrations (1.0-2.0 pM). Thus the MTP appears to play a significant role in the cellular effects following acute exposure of cerebellar granule neurons to MeHg. 135 INTRODUCTION At the cellular level, MeHg interacts with a number of processes to disrupt divalent cation homeostasis (Komulainen and Bondy, 1987; Hare et al., 1993; Marty and Atchison, 1997), elevate spontaneous release of neurotransmitter (J uang, 1976; Atchison and Narahashi, 1982; Atchison et al., 1986), generate reactive oxygen species (ROS) (LeBel et al., 1990; Sarafian and Verity, 1991; Yee and Choi, 1996) and induce cell death F (Sarafian et al., 1989; Marty and Atchison, 1998). As described in previous chapters, low concentrations of MeHg (0.2-2.0 uM) disrupt the ability of rat cerebellar granule cells in primary culture to maintain [Ca2+]i, causing a biphasic increase in [Caz+]i. The I initial phase (“first phase”) involves release of Ca2+i, followed several minutes later by a secondary (“second phase”) influx of Ca2+,3 (Marty and Atchison, 1997). MeHg-induced disruption of [Ca2+]i is critical in determining neuronal survival, because agents which buffer [Ca2+]; protect against MeHg-induced granule cell cytotoxicity (Marty and Atchison, 1998), and Ca2+ channel blockers afford some protection against MeHg- induced toxicity in viva (Sakamoto et al., 1996). One possible source of the early-onset Ca2+i elevation during MeHg exposure is the mitochondria. MeHg causes a number of effects in mitochondria, including inhibition of respiration, increased permeability of the inner membrane, and depolarization of the membrane potential (w...) (Verity et al., 1975; Sone et al., 1977). Additionally, MeHg exposure disrupts Ca2+ regulation in mitochondria derived from rat brain by decreasing Ca2+ uptake and inducing Ca2+ release (Levesque and Atchison, 1991). Inhibition of mitochondrial Ca2+ release reduces MeHg-induced release of ACh from rat cortical synaptosomes, supporting the idea that mitochondrial Ca2+ release plays a key role in 136 MeHg-induced neurotoxicity (Levesque et al., 1992). In granule neurons, mitochondria can store and release large amounts of Ca2+ under both physiological and pathological conditions (Budd and Nicholls, 1996b). Given that MeHg targets mitochondria in other systems, it was hypothesized that MeHg affects mitochondrial Ca2+ regulation in rat cerebellar granule neurons, causing release of mitochondrial Ca2+ into the cytosol during MeHg exposure. One route by which mitochondria release Ca2+ is the MTP (Bemardi et al., 1992; Petronilli et al. , 1993). The MTP is a megapore formed on the inner mitochondrial membrane in response to adverse conditions including oxidative stress, elevation of mitochondrial matrix Ca2+ and anoxia (Dubinsky and Rothman, 1991; Bemardi et al. , 1993). When open, the MTP allows the passage of molecules 51.5 KDa, including Ca2+, across the usually impermeable inner mitochondrial membrane. Because MTP opening dissipates the proton gradient across the inner mitochondrial membrane, there is a corresponding loss of rum which can be used as a diagnostic tool for observing a permeability transition (White and Reynolds, 1996; Trost and Lemasters, 1997). Both in vitro and in vivo, MTP induction is linked to neuronal death caused by a variety of compounds and conditions, including ischemia, stroke and glutamate excitotoxicity (Bemardi and Petronilli, 1996; Schinder et al. , 1996; White and Reynolds, 1996). Opening of the MTP can be inhibited by the immunosuppressant CsA, even in the presence of inducing agents (Bemardi et al., 1993); (N ieminen et a1. , 1996). While there is evidence for induction of a permeability transition in MeHg- induced death of non-neuronal cells (InSug et al., 1997; Shenker et al., 1998), there is no direct experimental evidence linking MeHg to MTP induction in granule neurons. 137 Because MeHg creates conditions which favor opening of the MTP (generation of ROS, elevation of [Ca2+]i, etc.), it seemed likely that MTP induction plays a role in MeHg- induced neurotoxicity. These experiments tested the hypothesis that MeHg induces opening of the MTP (through direct or indirect action at the mitochondria), causing it to release Ca2+ into the cytoplasm, thereby contributing to the elevated [Caz+]i observed during the first, Ca2+t-dependent phase of MeHg neurotoxicity. MTP induction was measured by examining the ability of CsA to prevent changes in [Ca2+]i, rum, and cell death following acute exposure to MeHg. Because CsA also inhibits the protein phosphatase calcineurin (CN) (Bemardi and Petronilli, 1996), the effects of the CN inhibitors F K506 and W7 were examined to determine if effects on CN function contributed to the observed actions of CsA. 138 MATERIALS AND METHODS The materials and solutions for isolation of cerebellar granule cells and measurement of changes in [Ca2+]; are described in detail in Chapter Two. The procedure for the Eukolight cell viability assay is described in detail in Chapter Three. The purchase and handling of CsA, and the use of TMRE to measure changes in ‘I’m, are described in Chapter Four. Unless specifically stated, all solutions were identical to those described in Chapter Two. The methods and solutions described below are specific to this chapter. Materials and Experimental Solutions. Phenylarsine oxide (PhAsO) was purchased from Sigma (St. Louis, MO). F K506 and N-(6-aminohexyl)-5-chloro-1- napthalenesulfonamide (W7) were purchased fi'om Calbiochem (La Jolla, CA). PhAsO was prepared as a 100 mM stock solution in 100% DMSO. F K506 was dissolved in DMSO as a 1 mM stock solution. W7 was dissolved into distilled water as a 1 mM stock solution. Control solutions contained corresponding concentrations of solvent, with all final solutions containing no more than 0.1% (v/v) solvent. Measurement of [Ca2+]. Changes Using Fura-Z. To examine the role of a permeability transition in MeHg-induced changes in [Ca2+]i, the potent MTP inhibitor CsA was utilized (Bemardi et al., 1993). Because CsA also inhibits CN function, the CN inhibitors W7 [which inhibits calmodulin (Ankarcrona et al. , 1996)] and F K506 [which inhibits FK506 binding protein (Marcaida et al. , 1996)] were also used. Granule cells 139 were loaded with fora-2 for 1 hr, then rinsed with HBS as described in Chapter Two. All experiments began with a 1 min baseline recording, followed by a 1 min exposure to 40 mM KL and a 3 min washout with HBS. For a cell to be considered viable, fura-2 fluorescence must increase transiently in response to the K+ depolarization challenge and subsequently return to baseline during the washout period. In the first set of experiments, the K+ depolarization was followed by a 10 min exposure to one of the inhibitors (0.25-5 pM CsA, 20 uM W7 or 1 uM F K506) or an appropriate solvent control. Next, the perfirsion solution was replaced with either 0.5 uM MeHg (experiment) or HBS (control) for 45 min. In the second set of experiments, the K+ depolarization was followed by simultaneous application of 5 uM CsA and 0.5 pM MeHg for 45 min; corresponding buffer and vehicle controls were also performed on the same day. Finally, both the 5 pM CsA pretreatment and simultaneous treatment experiments were repeated in Ca2+-free buffer to verify the intracellular origins of the first Ca2+ phase. For the 10 min pretreatment experiments in Ca2+-free buffer, the data were also analyzed for fura—2 ratio amplitude as a measure of [Ca2+]; released. In this analysis, the fora-2 fluorescence ratio was normalized to the peak fluorescence increase caused by 1 min 40 mM K“ to account for differences in dye loading. Visualization of Changes in ‘1’... Mitochondrial depolarization is a key feature of MTP opening; thus, prevention or attenuation of mitochondrial depolarization by CsA is indicative of MTP opening (N ieminen et al., 1996; White and Reynolds, 1996; Trost and Lemasters, 1997). To measure changes in ‘I’m, cells were loaded with TMRE as described in Chapter Four. As before, all perfusion solutions contained 10 nM TMRE to 140 maintain dye availability for the mitochondria. Fluorescence was measured relative to an initial fluorescence of 100%. Control experiments to verify the ability of cerebellar granule cell mitochondria to undergo a permeability transition utilized the known MTP- inducing agent PhAsO (5-15 uM) (Kowaltowski and Castilho, 1997). Experiments with MeHg were performed in a similar manner to the [Ca2+]; experiments, with application of 40 mM K+ for 1 min (to verify cell viability), followed by 3 min rinse with HBS, a 10 min pretreatment with 5 uM CsA, and 45 min exposure to 0.5 uM MeHg. Similar experiments were performed in the absence of Ca2+... and in the absence of the l min 40 mM K+ treatment. For all MeHg experiments using TMRE, the parameter measured was time-to-onset of the mitochondrial depolarization, determined at the point at which TMRE fluorescence dropped precipitously, indicating loss of Wm. For TMRE data, fluorescence was measured relative to 100% at the beginning of the experiment. The times-to-onset fiom all cells in an experiment were averaged to provide a mean time-to- onset of mitochondrial membrane depolarization for that dish of cells (n=1). Cell Viability Assays. To determine whether inhibition of MTP activation would prevent MeHg-induced cell death, a series of experiments was performed which paralleled the fura—2 imaging experimental protocol. All cells were incubated for 90 min at 37°C in HBS to mimic the fura-2 incubation (1 hr) and HBS perfusion rinse (30 min) of the [Cay], measurement experiments. This was followed by a treatment protocol (see below), after which the cells were returned to conditioned medium (to bind excess MeHg) until the Live/Dead assay was performed (as detailed in Chapter Three). In the first set of experiments, the effect of CsA treatment on MeHg-induced cell death was examined. Cells were pre-exposed to HBS for 90 min, followed by incubation 141 in O, 1 or 5 uM CsA for 10 min prior to application of 0-2 uM MeHg at the following concentrations and durations: 0 MeHg (control) for 60 min; 0.2 uM MeHg for 76 min; 0.5 pM MeHg for 45 min; 1 pM MeHg for 38 min; or 2 uM MeHg for 30 min. These exposure times were based on the time needed for both the first and second phase [Ca2+]; increases to occur in 97.5% of the cells at each concentration (Marty and Atchison, 1998). After MeHg exposure, cells were washed with conditioned medium and returned to the incubator for 24 hr. At 24 hr post-exposure, cells were washed twice with HBS, then incubated in cytotoxicity assay reagents for 30 min and assayed for. fluorescence. Viability was also determined 24 hr after simultaneous exposure to 5 uM CsA and 0-2 uM MeHg for the given time periods. In the second set of experiments, viability was detemrined 24 hr after simultaneous exposure to 5 uM CsA and 0-2 uM MeHg for the given time periods, using the same MeHg concentrations and durations listed above. In the final set of experiments, the effect of CN inhibition on MeHg-induced cell death was examined using 20 uM W7 and 1 uM FK506. Cells were pre-exposed to HBS for 90 min, followed by 10 min treatment with either W7 or F K506 and 45 min of 0.5 uM MeHg; viability was determined at 24 hr post-exposure. For all cytotoxicity assays, appropriate solvent and buffer controls were performed simultaneously. Statistics. For fura-2 and TMRE experiments, comparisons of mean time-to-onset for “MeHg” vs the corresponding “MeHg with inhibitor” were made using Student’s paired t-test. For the cell viability experiments, the percent viability data were transformed using an angular transformation and analyzed per experimental day using a repeated 142 measures one-way analysis of variance. For all experiments, a p<0.05 was considered to be statistically significant. I43 RESULTS As noted in previous chapters, neonatal rat cerebellar granule cells in culture undergo a biphasic increase in [Ca2+]i in response to acute exposure to MeHg. Exposure to CsA for 10 min immediately prior to MeHg exposure delayed the time-to-onset of both the first and second phase of [Ca2+]i elevations in a concentration-dependent manner. That is, lower concentrations of CsA (0.25 and 0.5 uM) did not delay the time-to-onset of either phase, while higher concentrations (1.0 and 5.0 uM) significantly delayed the time- to-onset of the first-phase [Ca2+]; elevation (Figure 5.1). The higher CsA concentrations also delayed the time-to—onset of the second phase; however, these results were not quite statistically significant (p=0.05). Interestingly, both phases were delayed to a similar extent, in that the first phase was delayed from 13.6:1:1 .4 min (0.5 11M MeHg alone) to 26.9i5.3 min and 27.7:t:3 .4 min for 1.0 and 5.0 uM CsA + 0.5 uM MeHg, respectively. The second phase was also delayed to a similar extent, from 21.6:b2.5 min (0.5 uM MeHg alone) to 33.13255 and 37.1d:1.7 min for 1.0 and 5.0 uM CsA + 0.5 uM MeHg, respectively. The next set of experiments tested the effect of a single CsA concentration on the time-to-onset of [Ca2+]; elevations caused by increasing concentrations of MeHg. A 10 min pretreatment with 5 uM CsA significantly delayed the first-phase release of Ca2+; caused by 0.2 uM MeHg (fi'om 18.0i1.6 min to 345i 3.3 min) and 0.5 uM MeHg (from 12.7:l:1.2 min to 25.6:t3.7 min), but not 1.0 uM MeHg (1.8:t:0.4 min for MeHg only vs. 1.6i0.5 min for CsA pretreatment) (Figure 5.2A). The same treatment also significantly delayed the influx of Ca2+. caused by 0.5 uM MeHg (from 20.3i2.4 min to 33:89 min) and delayed, although not significantly, the influx of Ca2+c caused by 0.2 uM MeHg 144 Time to Onset (min) 0 0.25 0.5 1.0 5.0 [Cyclosporin A] (uM) Figure 5.1. Increasing concentrations of CsA cause increasing delays in the time-to- onset of [Ca2+]t elevations caused by 0.5 pM MeHg. Granule cells pretreated with increasing concentrations of CsA exhibited increasing delays in the time-to-onset of both the first and second phase increases in [Ca2+]; caused by 0.5 uM MeHg. The time-to- onset was measured from the time of initial MeHg exposure. Although the times-to-onset of the second phases for pretreatment with 1 and 5 uM CsA were not considered to be statistically different from the 0 uM CsA cells, the results were remarkably close (p = 0.05 for each). Results are presented as mean :1: SEM (n = 6-10). The asterisk (*) indicates a value significantly different from control (p < 0.05). 145 - MeHg only - CsA + MeHg 50 ' 4O 3O 20 10 0 Time-to-Onset (min) 0.2 0.5 1 .0 [MeHg] (MM) - MeHg only - CsA + MeHg 70 T 60 50 40 30 20 10 0 Time-to-Onset (min) 0.2 0.5 1.0 [MeHg] (luM) Figure 5.2. 5 pM CsA significantly delays the time-to-onset of MeHg-induced [Ca2+]: increases. (A) Granule cells pretreated with 5 pM CsA exhibited delayed the time-to-onset of the MeHg-induced first-phase [Ca2+]i increase caused by 0.2 and 0.5, but not 1.0 uM MeHg (n = 3-9). (B) 5 uM CsA pretreatment also delayed the time-to-onset of Ca2+ influx (second phase) caused by 0.2 and 0.5, but not 1.0 uM MeHg. Results are presented as mean :1: SEM (n = 3-9). The asterisk (*) indicates a value significantly different from control (p < 0.05). 146 (from 39.7:1z7.2 min to 56:1:5.5 min). CsA pretreatment did not affect the Ca2+c influx caused by 1.0 pM MeHg (Figure 5.2B). The effect of CsA is reported to be transient such that inhibition of the MTP lasts only 10-20 min (N ieminen et al., 1996). As such, continuous exposure to both CsA and MeHg might increase the delay in time-to-onset of both phases to a greater extent than was seen in the pre-exposure experiments. However, there was no difference between a m 10 min pretreatment with 5 uM CsA or 45 min simultaneous treatment with 5 uM CsA in terms of time-to-onset for the first and second phase [Ca2+]; elevations (data not shown). Simultaneous treatment with CsA delayed the effects of MeHg for both the first phase I (from 15.5i2.6 min to 27.7:t3 .4 min) and second phase (from 24.2iZ.5 min to 37.1:t:1.7 min). Attempts to potentiate the effect of CsA using trifluoperazine were unsuccessful because granule cells respond to trifluoperazine alone with elevated [Ca2+]; (data not shown). In EGTA-containing buffer, pretreatment with 5 uM CsA caused similar delays in the time-to-onset of the first-phase [Ca2+]i elevation as seen in Ca2+-containing buffer. 5 pM CsA delayed the time-to-onset of the first phase for 0.2 pM MeHg (fiom 14.7d:1.5 min to 24.2:tl.6 min) and 0.5 uM MeHg (fiom 12.3:t:2.4 min to 2233.8 min), but not for 1.0 uM MeHg (from 2.1-1:0.2 min to 2.6i0.7 min) (Figure 5.3). Measurement of the first phase ratio amplitude in these cells indicated that CsA did not alter the amount of Ca2+i released during MeHg exposure at any MeHg concentration examined (data not shown). Simultaneous exposure to 5 uM CsA and 0.5 uM MeHg in the EGTA buffer had a similar effect, delaying the first phase to the same extent (data not shown). These results are very similar to those obtained in Cay-containing medium, and implicate the 147 contribution of an intracellular Ca2+ store sensitive to CsA in the first phase elevation of [Ca2+]; caused by MeHg. Several CN inhibitors were used in an experimental protocol similar to that with CsA to test whether CN contributes to the MeHg-induced changes in [Cay]; and mm. Cells were pretreated with either 20 uM W7 or 1 pM F K506 for 10 min, followed by treatment with 0.5 uM MeHg. These concentrations were chosen based on previous E findings by Marcaida et al. (1996) and Ankarcrona et al. (1996). Neither F K506 nor W7 altered the time-to-onset of either the first or second MeHg-induced [Ca2+]Ii phases L. (Figure 5.4), indicating that CN does not play a significant role in changes in [Ca2+]i induced by MeHg. The effect of CsA on MeHg-induced changes in rpm was monitored using the potentiometric dye TMRE. Control experiments using 5-15 uM of the MTP-inducing agent PhAsO (Kowaltowski and Castilho, 1997) indicated that granule cell mitochondria can undergo a permeability transition, which is visualized as a complete loss of TMRE fluorescence due to depolarization of the inner mitochondrial membrane (Figure 5.5). As seen in Figure 5.6, MeHg causes loss of TMRE fluorescence consistent with a loss of rum. This is consistent with previous findings in intact monocytes and T cells (InSug et al. , 1997; Shenker et al. , 1998), as well as with mitochondria—containing synaptosomes (Hare and Atchison, 1992). Exposure to 0.5 uM MeHg alone caused loss of TMRE fluorescence at 25.0i2.5 min (Figure 5.7). The loss of rum was irreversible even when the MeHg was replaced with HBS, as the TMRE fluorescence did not recover during a 60 min period following MeHg exposure (figure not shown). However, this loss of fluorescence was delayed by a 10 min pretreatment with 5 uM CsA, such that neurons 148 - MeHg only - CsA + MeHg Ta“ 30' a a E a; 20- c: q 3 10 - o E O H 0.5 1.0 [MeHg](uM) Figure 5.3. 5 uM CsA delays the time-to-onset of MeHg-induced release of Ca”. in the absence of Ca2+” In the absence of C322, 5 pM CsA delayed the time-to—onset of [Ca2+]; elevations caused by 0.2 and 0.5, but not 1.0 uM MeHg. Results are presented as mean as SEM (n = 3-7). The asterisk (‘) indicates a value significantly different from control (p < 0.05). 149 Time-to-Onset (min) Time-to-Onset (min) - MeHg only - FK506 + MeHg 40 ’ 30 20 1st Phase 2nd Phase - MeHg only - W7 + MeHg 40 30 20 T‘—I_—'_—l_‘ lst Phase 2nd Phase Figure 5.4. Effect of calcineurin inhibitors on MeHg-induced [Ca2+]i elevations. A) A 10 min pretreatment with l LuM FK506 did not alter the time-to-onset of [Ca2+]; increases induced by 0.5 uM MeHg Sn = 6). B) A 10 min pretreatment with 20 11M W7 did not alter the time-to—onset of [Ca +]i increases induced by 0.5 uM MeHg (n = 6). 150 1.50 _ 15 M PhAsO (D O c: CD 0 U) g 1.00 1 a [-1-4 E 0.50 E 0) m 0.00 l l J l l O 10 20 30 40 50 60 Time (min) Figure 5.5. PhAsO causes a rapid decrease in TMRE fluorescence. Representative example of changes in TMRE fluorescence in a single granule cell. The known MTP- inducing compound PhAsO (15 [AM ) causes mitochondrial membrane depolarization, as indicated by the loss of TMRE fluorescence as measured relative to an initial fluorescence of 100%. 151 1.50 1 0.5 pM MeHg 1.00 0.50 Relative Fluorescence 0.00 A I l l ? l 1 ‘ 0 10 20 30 4O 50 60 Time (min) Figure 5.6. Exposure to 0.5 uM MeHg causes irreversible loss of TMRE fluorescence. In this representative example of a single granule cell loaded with TMRE, exposure to 0.5 11M MeHg causes rapid, irreversible loss of TMRE fluorescence, indicating mitochondrial membrane depolarization (dissipation of Wm). Time-to—onset of mitochondrial membrane depolarization is measured at the point at which TMRE fluorescence drops below baseline (dashed line). TMRE fluorescence is measured relative to an initial fluorescence of 100%. 152 - MeHg only - Treatment + MeHg A40~ * .S 530- ‘5 a 020" 52 £210- .§ [—1 0 CsA W7 FK506 Figure 5.7. CsA, but not CN inhibitors, delay the time-to-onset of mitochondrial membrane depolarization caused by 0.5 pM MeHg. Pretreatment with 5 uM CsA significantly delayed the loss of wm; however, pretreatment with 1 uM FK506 or 20 uM W7 did not. Results are presented as mean d: SEM (n = 5-9). The asterisk (*) indicates a value significantly different from the corresponding 0.5 uM MeHg data (p < 0.05). 153 receiving CsA did not lose TMRE fluorescence until 34.6:l:2.8 min. Mitochondrial membrane depolarization was not delayed by 10 min pretreatment with 1 pM FK506 or 20 nM W7 (Figure 5.7). Removal of Ca2+, with EGTA did not alter the time-to-onset of the depolarization caused by 0.5 pM MeHg (data not shown). The presence or absence of the initial K+ challenge also did not affect the MeHg-induced loss of rpm (data not shown). CsA prevents cell death caused by anoxia and glutamate excitotoxicity, presumably by inhibiting the MTP (Pastorino et al., 1995; White and Reynolds, 1996). Because CsA altered changes in [Ca2+]; and rum caused by MeHg, the ability of CsA to prevent MeHg-induced cell death was examined. Pretreatment for 10 min with 1 uM CsA protected cells against subsequent exposure to 0.2 uM MeHg, but not against higher concentrations of MeHg (Figure 5.8). Increasing the CsA concentration to 5 uM was more effective, and protected against cell death caused by 0.5 1.1M MeHg but again, not against that caused by higher MeHg concentrations. Thus CsA is cytoprotective against acute exposure to low concentrations of MeHg. Viability assays were also performed comparing the effect of 10 min pretreatment with 5 uM CsA vs. 5 uM CsA given with MeHg on the MeHg-induced cell death observed at 24 hr post-exposure (Figure 5.9). As with the comparison of the effect of pretreatment and simultaneous treatment with CsA on MeHg-induced increases in [Ca2+]i, there was no difference between these two protocols in their effectiveness in preventing MeHg-induced cell death. The ability of 20 uM W7 and 1 uM FK506 to prevent 0.5 uM MeHg-induced cell death was examined using a parallel pretreatment regimen. Treatment with 1 uM F K506 154 for 10 nrin prior to 0.5 pM MeHg did not improve cell viability at 24 hr post-exposure (data not shown). Pretreatment with 20 11M W7 for 10 min significantly protected against MeHg-induced cell death, increasing the percent viability from 41.3 i 5.6% in cells treated with 0.5 uM MeHg only, to 61.6 i 4.6% in cells treated with W7 just prior to 0.5 uM MeHg (Figure 5.10). However, W7 was unable to reverse completely the MeHg-induced cell death and did not protect cells to the extent CsA did. 155 - 5pMCsA W luMCsA l:l OpMCsA %Viability m\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\‘< MeHgGlM) Figure 5.8. CsA provides limited protection against MeHg-induced cell death at 24 hr post-exposure. Granule cell viability following exposure to 0, 1 or 5 uM CsA and increasing concentrations of MeHg at 24 hr post-MeHg exposure. Increasing concentrations of CsA resulted in increasing protection against MeHg-induced cell death. At 1 uM CsA, cells were protected from death caused by 0.2 uM MeHg, whereas 5 uM CsA also protected cells from death from MeHg at concentrations up to 0.5 11M. However, protection was lost at higher MeHg concentrations. Results are presented as mean i SEM (n = 4). The asterisk (*) indicates a value significantly different from the corresponding 0 uM MeHg data (p < 0.05). 156 - CsA pretreat - CsA + MeH [iii MeHg only 0 0.2 LL ** "f a”; '1 l ~fl< 1.0 2.0 % Viability 80 * if 60 ‘ 40 ; 20 L l O . solv 0.5 MeHg (PM) Figure 5.9. Comparison of protection against MeHg-induced cell death by pretreatment or simultaneous treatment with 5 uM CsA. Granule cell were pretreated with 5 uM CsA for 10 min (“CsA pretreat”) prior to exposure to increasing concentrations of MeHg, or with 5 uM CsA simultaneously with increasing concentrations of MeHg (“CsA+MeHg”). Results are presented as mean :t SEM (n = 4). The asterisk (*) indicates a value significantly different from the corresponding 0 uM MeHg data (p < 0.05). 157 % Viability solv 0 0.5 0.5 + W7 [MeHg] (MM) Figure 5.10. Pretreatment with 20 uM W7 significantly protects against cell death caused by 0.5 pM MeHg at 24 hr post-exposure. Results are presented as mean i SEM (n = 4). The asterisk (*) indicates a value significantly different from the corresponding 0.5 11M MeHg data (p < 0.05). 158 DISCUSSION These experiments examined the role of MTP induction in cerebellar granule cell neurotoxicity induced by acute exposure to MeHg. The MTP inhibitor CsA was used to define the role of the MTP in MeHg-induced changes in Ca2+ homeostasis. Assuming that CsA prevents mitochondrial Ca2+ release through the MTP, it was anticipated that the first phase increase of Ca2+; caused by MeHg would be affected by CsA; surprisingly, both the first and second phases were delayed, suggesting that the two phases of increased [Ca2+]i are interrelated with the first phase perhaps triggering or contributing in some way to the onset of the second phase. The CsA-mediated delay was not dependent on Ca2+¢, as indicated by the similar first phase times-to-onset in the presence and absence of Cafe. Pretreatment with CsA delayed, but did not prevent, MeHg-induced changes in [Ca2+]i, agreeing with the short-lived protection of CsA found in other systems (N ieminen et al., 1996). The CN inhibitors W7 and F K506 did not change the time-to- onset of increased Capri, indicating that CN does not contribute substantially to the delayed MeHg response induced by CsA. Thus, it appears that the MTP participates in the initial increase in [Ca2+]; by releasing Ca2+ into the cytosol upon its opening. Whether the elevated [Ca2+]i is a causative factor or a reaction to a permeability transition cannot be determined fiom these experiments. Previous studies indicate that the second phase of increased [Ca2+]; is due to entry of Ca2+e, with a possible contribution of L, N and/or P/Q-type voltage-gated Ca2+ channels (Marty and Atchison, 1997). To date, no inhibitor has been able to block completely this second phase. It was hypothesized that initiation of the second phase 159 may be dependent on the first phase, either though pathways to replenish Ca2+; stores (depleted by release of Ca2+ from the SER, for example) or through another as yet undefined mechanism. These experiments provide evidence of a link between the two phases, as delay of the first phase with CsA is accompanied by delay of the second phase. Thus, induction of the MTP may link the first and second Ca2+ phases as well as be a critical factor in determining cell survival following MeHg exposure. Using TMRE, it was found that mitochondria depolarize irreversibly in response to a low concentration (0.5 uM) of MeHg. Because an identical depolarization occurs in the absence of Ca2+c, it is concluded that this depolarization is dependent on an intracellular factor (such as elevated [Ca2+]i), on MeHg alone, or on a combination of direct damage by MeHg and other cellular factors. Neither CsA treatment nor replacement of the MeHg with normal medium reversed the loss of Wm, suggesting that MeHg binds at least one critical site with high affinity. MeHg exhibits high-affinity binding to intracellular organelles, especially mitochondria (Chang and Hartmann, 1972), thereby altering protein sulflrydryl groups and disrupting cellular functions, including respiration, ion transport and regulation, and protein synthesis (Verity et al., 1977; Kauppinen et al., 1989; Levesque and Atchison, 1991; Yee and Choi, 1996). Because MTP induction depends on critical, voltage-sensitive thiol groups (Scorrano et aI. , 1997), it is possible that direct binding of MeHg to a site on the MTP could induce a permeability transition. MeHg binding would likely be irreversible, which could explain the irreversible mitochondrial depolarization observed with MeHg exposure. This is in contrast to other MTP-inducing agents; for example, exposure to low concentrations of NMDA in cultures of rat hippocampal neurons causes a mitochondrial depolarization 160 which recovers to near normal levels when the neurons also receive CsA (Schinder et al., 1996). MeHg exposure, on the other hand, does not permit recovery of 01.... This suggests a more severe, and possibly more direct insult to the mitochondria which may be exacerbated by elevated cytosolic Ca2+, thereby overwhelming the cell and leading to neuronal death. Despite its inability to prevent mitochondrial depolarization, CsA was effective at protecting cerebellar granule cells fiom cell death at 24 hr following exposure to low concentrations of MeHg. Cytotoxicity assays performed at 24 hr post-exposure indicate that a significant fraction of the cultured cells undergo cell death; this fi'action increases with increasing MeHg concentrations. These results agree with earlier findings by Marty and Atchison, 1998. In previous experiments, treatment with the Ca2+; chelator BAPTA protected cerebellar granule cells from death 3.5 h after exposure to low concentrations of MeHg (Marty and Atchison, 1998). Because onset of a permeability transition is favored by elevated Ca2+i levels, and cell death at low MeHg concentrations is attenuated by chelating [Ca2+]i with BAPTA, these experiments suggest that elevated [Ca2+]i contributes to a CsA-sensitive permeability transition which leads to cell death. As higher MeHg concentrations are used, less protection is provided by CsA, indicating a limit to its effectiveness. The CN inhibitor W7 also reduced the percentage of cells killed by 0.5 uM MeHg; however, W7 did not significantly alter the MeHg-induced changes in [Ca2+]i or rpm. Thus, its effects on cell viability may not be due to inhibition of CN but rather to some other consequence of calmodulin inhibition. The inhibition of MeHg- induced cell death by CsA is interesting, as CsA delays but does not prevent the onset of the MTP in granule cells. The observed inhibition of cell death in these experiments may 161 simply reflect a delay in the onset of cell death in the presence of CsA, rather than actual prevention of cytotoxicity. These experiments emphasize the importance of mitochondrial Ca2+ homeostasis in MeHg-induced neurotoxicity. In recent years, the importance of mitochondrial Ca2+ regulation has become evident not only in neuronal damage caused by a variety of agents, including glutamate excitotoxicity in cerebellar granule cells (Budd and Nicholls, 1996a), but also in regulating normal cell physiology (see Duchen, 2000, for review). Loss of mitochondrial Ca2+ buffering is crucial in both apoptotic and necrotic Cay-mediated neuronal death (Kruman and Mattson, 1999). Mitochondria accumulate Ca2+ based on the extremely negative membrane potential within the inner mitochondrial membrane (140-180 mV, negative inside) (Gunter and Gunter, 1994), making Ca2+ buffering an energetically favorable process. However, Ca2+ buffering comes at a price: uptake of a large amount of Ca2+ occurs at the expense of the proton motive force, and can thus cause uncoupling of ATP synthesis from respiration. Under physiological conditions, mitochondrial Ca2+ buffering is important in regulating cytosolic [Ca2+] elevations which result fiom either influx of Ca2+: or release of Ca2+ from the SER, and may play a role in modulating cellular Ca2+ oscillations (Simpson and Russell, 1996; Rizzuto et al., 1998). However, high [Ca2+]... can become hazardous to the cell via inhibition of the electron transport chain and the mitochondrial ATPase, as well as dissipation of the proton gradient, all of which can contribute to loss of mitochondrial ATP synthesis, opening of the MTP and cell death (Gunter and Gunter, 1994; Simpson and Russell, 1998b). Loss of mitochondrial Ca2+ regulation appears to be a central feature of MeHg-induced loss of neuronal Ca2+ regulation (Levesque and Atchison, 1991). In fact, emptying of 162 mitochondrial Ca2+ stores prior to exposure to 0.5 11M MeHg decreases the amplitude of the MeHg-induced first-phase fura-2 ratio increase in granule cells by approximately 65% (see Figure 4.3B), indicating the significant contribution of mitochondrial Ca2+ stores to MeHg-induced [Ca2+]; elevations in these highly sensitive cells. By examining the changes which occur in cerebellar granule cells during MeHg exposure, several possible reasons for their heightened sensitivity to toxicity are revealed. For example, mitochondria are not all identical, and brain mitochondria exhibit differential sensitivity by region (F riberg et al. , 1999). Because these experiments indicate that MeHg induces a permeability transition in granule cells, it is possible that their mitochondria are more sensitive to conditions which favor MTP induction, such as elevated [Ca2+]i and production of ROS. Cerebellar granule neurons in culture have a relatively weak capacity to up-regulate GSH levels in response to MeHg as compared to cerebellar glia; this may contribute to a heightened sensitivity to onset of the MTP (Sarafian and Verity, 1991). Further, the difference in sensitivities may be related to different Ca2+ buffering capacities. Granule cells lack certain Ca2+ buffering proteins present in neighboring cells which are less sensitive to MeHg. For example, Purkinje cells contain high concentrations of the calcium binding proteins parvalbumin and calbindin D-zgx (Celio, 1990). Because MTP induction relies on increased extramitochondrial Ca2+ levels for Ca2+ loading, it is possible that a lack of sufficient Ca2+ binding proteins puts more of a burden on mitochondria to buffer sustained increases of [Ca2+]; in cerebellar granule cells. MeHg exposure has also been linked to elevated levels of ROS, which are believed to contribute to MeHg-induced neurotoxicity (Sarafian and Verity, 1991). 163 However, elevated levels of ROS may not be sufficient for MeHg to be lethal. In experiments using isolated rat liver mitochondria, sensitivity of the MTP to ROS was shown to be highly dependent on the mitochondrial Ca2+ load (Scorrano et al. , 1997). Moreover, mitochondrial Ca2+ uptake is an early, critical event in both apoptosis and necrosis in the neuronal model PC12 cells (Kruman and Mattson, 1999). In these cells, an initial increase in cytosolic Ca2+ was followed by a delayed increase in mitochondrial Ca2+, then accumulation of ROS and mitochondrial depolarization, caspase release and cell death. Both the mitochondrial depolarization and cell death were CsA-sensitive, implicating the permeability transition as a key event. This is remarkably similar the effect of MeHg in granule cells, in which an initial increase in [Ca2+]i (resulting fiom release of Ca2+i stores) is followed by mitochondrial depolarization and cell death, with all three components exhibiting sensitivity to CsA. Thus, it appears that the disruption of [Ca2+]; homeostasis induced in granule neurons by MeHg is the critical factor in MeHg- induced cell death, with the MTP playing a critical role. 164 CHAPTER SIX SUMMARY AND CONCLUSIONS 165 SUMMARY AND CONCLUSIONS This dissertation presents the results of experiments designed to examine the hypothesis that MeHg causes Ca2+ dysregulation in cerebellar granule cells by interacting with specific intracellular Ca2+ stores. Specifically, it was hypothesized that MeHg initiates Ca2+ release from the SER which is buffered by mitochondria; however, MeHg also impairs the ability of mitochondria to hold Ca2+ (by causing loss of mitochondrial membrane potential and inhibiting respiration), resulting in opening of the mitochondrial permeability transition pore and subsequent cell death. The results presented in this dissertation not only support this hypothesis but also provide additional information regarding the mechanisms underlying MeHg-induced elevations of [Ca2+]; and subsequent neuronal death. The summary of the results delineated in these experiments is presented in Figure 6.1. In the first experiments, the role of SER Ca2+ in MeHg-induced changes in [Ca2+]i and cell viability was examined. Overall, the results suggest that the SER plays a small role in the whole-cell changes in [Ca2+]; caused by MeHg in rat cerebellar granule cells. This is not surprising, given the small capacity of this Ca2+ store (F ohrman et al., 1993; Masgrau et al., 2000). The Ca2+ was released primarily from 1P3 receptors, with a minimal contribution of Ca2+ fiom the ryanodine receptors. Atropine delayed the time- to—onset and reduced the amplitude of MeHg-induced [Ca2+]; increases, indicating that this occurred as a result of activation of muscarinic ACh receptors. However, the potential interaction between MeHg and the 1P3 receptors themselves is not ruled out by these experiments. SER store depletion does not appear to trigger Ca2+ influx through SOCE, as none of the inhibitors of SER Ca2+ release affected 166 lCO" In: {If \ ti SERCA can Figure 6.1. Summary of findings. MeHg binds M3 muscarinic ACh receptors, causing production of 1P3 and release of Ca2+ through the 1P3 receptors. This elevation of [Ca2+], triggers Ca2+-induced Ca2+ release through the ryanodine receptors. Ca2+ released fiom the SER is buffered by mitochondria. Excess uptake of Ca2+ into the mitochondria, combined with direct effects of MeHg on the mitochondrial respiratory chain, leads to opening of the mitochondrial permeability transition pore (MTP) and promotion of cell death. However, there may be direct effects of MeHg on mitochondria which also contribute to the observed changes in cellular function (dashed arrow). 1P, Receptors RYanodtne Receptors ism 167 the MeHg-induced influx of Cafe. Despite the small amount of Ca2+ release fi'om the SER store, inhibition of this Ca2+ release was a determinant factor in later cell death caused by MeHg. Cell death was prevented by down-regulation of muscarinic and IF; receptors using a 24 hr treatment with BCh. This protection was likely mediated by the M3 receptor and not the nicotinic ACh receptor, as protection was not afforded by BCh when 4-DAMP or atropine were included, but was not affected by the addition of DHE. The M3 receptor is coupled to P3 production in granule cells (Whitham et al., 1991; F ohrman et al., 1993), thus inhibition of M3 receptors presumably prevents a MeHg-induced increase in [1P3]. The potential contribution of the M2 receptor to MeHg-induced neurotoxicity was not examined as a result of the toxicity of methoctramine alone. However, typically activation of M2 muscarinic receptors is coupled to activation of potassium conductance (Pan and Williams, 1994), which would not be expected to facilitate Ca2+ release or entry. Indiscriminate release of Ca2+ from the SER is insufiicient to contribute to MeHg- induced cell death, as indicated by the lack of protection afforded by ryanodine. Surprisingly, atropine alone and XeC were not neuroprotective, in contrast to the ability of atropine to delay MeHg-induced elevations of [Ca2+]i. In sum, these experiments provide evidence that the SER participates in MeHg-induced elevations of [Ca2+]; and subsequent neuronal death. In the second set of experiments, the role of mitochondrial Ca2+ in MeHg-induced elevations of [Ca2+]; and cell death was examined. Unlike the small contribution of SER Ca2+, mitochondria contribute a large portion of the [Ca2+]; observed in cerebellar granule cells during exposure to MeHg. Removal of mitochondrial Ca2+ prior to MeHg exposure 168 decreased the amplitude of the Ca2+ release from intracellular sources by more than 50%. Observation of cells loaded with rhod-2 to monitor changes in [Ca2+]... indicated that MeHg causes an early increase in [Ca2+]... fiom Ca2+ originating within the cell. This increase in [Ca2+],fl is followed several minutes later by spreading of the dye from a punctate stain, presumably mitochondrial in origin, to more generalized whole-cell localization. The spreading of the dye beyond the mitochondria was delayed by thapsigargin, suggesting that SER Ca2+ contributes to the elevated [Ca2+]m. The inhibitor of MTP opening CsA also delayed the spreading of rhod-2 beyond the punctate pattern, suggesting that the dye movement results from opening of the MTP to allow the dye to cross rapidly through the usually impermeable inner mitochondrial membrane. The participation of the MTP in MeHg neurotoxicity is implicated further by the ability of CsA to attenuate MeHg-induced elevations of [Ca2+]i, loss of mitochondrial membrane potential, and cell death. Calcineurin inhibitors did not affect the MeHg-induced increase in [Ca2+]; or loss of mitochondrial membrane potential. However, the calmodan inhibitor W7 did protect against MeHg-induced cell death at 24 hr post-exposure. This is not surprising, because calmodulin is known to play a role in Calf-mediated cell death (Orrenius and Nicotera, 1994; Urushitani et al., 2001). The impact of MeHg-induced alterations on [Ca2+]... has widespread implications for the viability of affected neurons, leading to a proposed scenario in which MeHg- induced disruption of mitochondrial firnction is a focal point for many of the observed changes which occur in cerebellar granule neurons. As demonstrated in this dissertation, MeHg causes an early increase in [Ca2+]In as a result of Ca2+ uptake into the mitochondrial lumen. Initially, this uptake of Ca2+ into the mitochondria by the Ca2+ 169 uniporter would be driven by the mitochondrial membrane potential (140-180 mV, negative inside) (N icholls, 1974); however, as the [Ca2+]m increases, the inner membrane potential becomes more positive, thereby making uptake of Ca2+ less favorable energetically (N icholls and Scott, 1980). As the transmembrane potential dissipates, mitochondria buffer Ca2+ using ATP-dependent mechanisms (N icholls and Scott, 1980), which diverts a portion of the neuron’s ATP away fiem other cellular processes. The inner membrane depolarization reduces the proton motive force which drives ATP mitochondrial ATP synthesis, thus elevated [Ca2+]... can effectively uncouple ATP synthesis from respiration (Hillered et al., 1983). Additionally, high [03*]... can inhibit the electron transport chain and the mitochondrial ATP synthase, which results in inhibition of mitochondrial ATP synthesis (see Gunter and Gunter, 1994, and Simpson and Russell, 1998b, for review). In an effort to maintain trim, the ATP synthase can reverse, and begin hydrolyzing ATP to ADP to produce the protons necessary to maintain the negative inner membrane potential (Leyssens et al., 1996). The end result is depletion of cellular ATP, which has a wide range of effects within the cell. Loss of ATP deprives plasma membrane Ca2+ extrusion proteins of the energy necessary to maintain the steep [Cay] gradient across the plasma membrane, and reduces the amount of ATP available for Ca2+ pumps on storage organelles, such as the SERCA pump of the SER. Additionally, loss of ATP inhibits the Nai’-K+ ATPase, which maintains the membrane potential across the plasma membrane. Inhibition of the Na+-K+ ATPase would lead to plasma membrane depolarization, thereby leading to activation of plasma membrane voltage-gated Ca2+ channels and influx of Ca2+c into a cell now compromised in its ability to extrude Ca2+ fiom the cytoplasm. ATP depletion, loss of rpm and elevated [Ca2+]... all 170 contribute to induction of the MTP and promotion of cell death (Bemardi and Petronilli, 1996), and each of these actions has been shown to occur with MeHg (Levesque and Atchison, 1991; Hare and Atchison, 1992; present thesis). The far-reaching consequences of mitochondrial effects are supported by experiments presented in this dissertation and elsewhere. For example, the ability of oligomycin to delay the time-to- onset of MeHg-induced influx of Ca2+c (Figure 4.5) suggests that MeHg-induced depletion of ATP triggers influx of Ca2+, in cerebellar granule cells. This influx occurs primarily through voltage-gated Ca2+ channels (Marty and Atchison, 1997) and occurs at approximately the same time as the onset of plasma membrane depolarization (Stringfellow et al. , 2000). Thus, the mitochondria are not only affected directly by MeHg, but are also potentially responsible for a large number of the observed cellular responses which lead to neuronal damage and death. The results presented in this dissertation support the idea that MeHg neurotoxicity results fiom disruption of an intricate signaling network in which disruption of one portion of the network affects the functioning of another portion such that studying each component separately does not provide an adequate picture of the mechanism underlying MeHg neurotoxicity. As suggested by results presented here, the importance of the SER as a target in MeHg neurotoxicity may lie in the effect of Ca2+ released from the SER on mitochondria. Under normal conditions, granule cell mitochondria do not store Ca2+; however, when they are exposed to a high extra-mitochondrial [Caz+], they usually respond by buffering this Ca2+ via several uptake mechanisms (the uniporter, the rapid active mode transporter, etc.) (Budd and Nicholls, 1996a; Budd and Nicholls, 1996b). Experiments using specifically-targeted, Cay-sensitive proteins demonstrate that opening 171 of the IP3 receptor-activated channels results in higher [Ca2+] at the mitochondrial surface than in the general cytosol (Rizzuto et al. , 1998), suggesting a preferential signaling pathway in which Ca2+ stored within the SER lumen is released to communicate a pro- apoptotic signal to mitochondria. Excess mitochondrial matrix Ca2+ contributes to cell death by promoting release of cytochrome C, enhancing production of reactive oxygen species, inhibiting mitochondrial ATP synthesis, and opening the MTP, all of which can contribute to cellular damage and death (Bemardi and Petronilli, 1996; Szalai et al. , 1999). In HeLa cells, overexpression of calreticulirr, an SER luminal protein which binds Ca2+, causes an increased sensitivity to induction of apoptosis and increase in release of cytochrome C fiom the mitochondria, highlighting the role of SER-mitochondria communication in apoptosis (N akamura et al. , 2000). By causing pathological release of Ca2+ from the SER which contributes to the mitochondrial dysfunction, MeHg causes a complicated pattern of Ca2+; changes which contribute to cell death. In interpreting the data presented in this dissertation, there are several inconsistencies which complicate the process of delineating a precise, mechanistic model for the actions of MeHg in cerebellar granule cells. The model (as shown in Figure 6.1) focuses only on effects of MeHg on Ca2+ signaling systems within the cell. In this model, there is an assumption that any effects on Ca2+ signaling are due to a direct interaction between MeHg and the receptor signaling system being examined. In reality, MeHg- induced disruption of [Ca2+]i may reflect disruption of a non-Ca2+ system which in turn affects [Ca2+]i. For example, MeHg increases free [Zn2+] within the cytosol in rat cortical synaptosomes (Denny et al., 1993; Denny and Atchison, 1994) and possibly within NG108-15 neuroblastoma cells (Hare et al., 1993) and rat cerebellar granule cells (Marty 172 and Atchison, 1997). MeHg also causes an elevation of ROS within rat cerebellar granule cells and human T lymphocytes (Sarafian and Verity, 1991; Oyama et al., 1994; Yee and Choi, 1996; Shenker et al., 1999). Because both Zn2+ and ROS are known inducers of the MTP (Wudarczyk et al. , 1999), the observed MeHg-induced elevation of [Ca2+]; that is CsA-sensitive (see Chapter Five) may be an indirect result of MeHg- induced elevations of Zn2+ and/or ROS. Similarly, MeHg depletes cellular concentrations of ATP (Kuznetsov et al., 1987; Sarafian et al. , 1989), which are critical for maintenance of [Ca2+]; by providing energy to both intracellular Ca2+ pumps (such as the SERCA) and plasma membrane extrusion proteins (such as the plasma membrane Ca2+ pump) (Orrenius and Nicotera, 1994; Guerini et al., 1998). Like Zn2+ and ROS, mitochondrial ATP depletion is a contributing factor to opening of the MTP (Simbula et al., 1997), which could then contribute to elevation of [Cay]; through opening of the MTP. Thus, the observed elevation of [Cay]; in cerebellar granule cells may be symptomatic of other, non-Ca2+ effects of MeHg within the cell. Additionally, the non-Ca2+ effects may contribute to the direct effects of MeHg on Ca2+ signaling pathways that are shown in Figure 6.1. For example, MeHg causes mitochondrial membrane depolarization associated with opening of the MTP (see Chapters Four and Five). However, mitochondrial depolarization may not be sufficient to cause opening of the MTP without contributing factors, such as elevations of Ca2+... or elevation of ROS. In hepatocytes, mitochondrial membrane depolarization with CCCP is not suflicient to open the MTP; however, CCCP plus the pro-oxidant t- butylhydroperoxide causes rapid opening of the MTP which is accompanied by ATP depletion and subsequent cell death (N ieminen et al., 1995). This phenomenon of 173 mitochondrial membrane depolarization without opening of the MTP may underlie the effect of CCCP and oligomycin on MeHg-induced elevations of rhod-2 fluorescence (Figure 4.8). The combination of CCCP and oligomycin causes rapid loss of rum but does not prevent the movement of rhod-2 fiom localized to diffuse staining when cells are exposed to 0.5 uM MeHg. This suggests that there is another factor contributing to opening of the MTP, such as production of ROS or elevation of cytosolic [Zn2+], which occurs during MeHg exposure and which was not considered in the experiments undertaken for this dissertation. Another inconsistency in this model is the assumption that MeHg has specific interactions which occur in the same temporal order during every MeHg exposure in cerebellar granule cells. The model presented in Figure 6.1 indicates an initial interaction with muscarinic receptors, followed by release of Ca2+ fi'om the SER and subsequent disruption of Ca2+... regulation. The problem with this assumption is apparent in the inconsistencies observed in the time-to—onset data for the first-phase increase of [Ca2+]; caused by MeHg. For example, both atropine, which indirectly affects release of Ca2+ fi-om the SER, and CsA, which inhibits opening of the MTP, delay the time-to-onset of MeHg-induced release of [Ca2+]t. These data contradict the proposed model because they suggest that MeHg is targeting both the SER and mitochondria simultaneously rather than sequentially, as suggested by the thapsigargin-sensitive nature of MeHg-induced increases of C322... Another discrepancy is apparent in the lack of effect of thapsigargin on the time-to-onset of MeHg-induced elevations of [Ca2+]i. Both BCh and atropine changed the time-to-onset, but thapsigargin had no effect on the timing of the [Ca2+]; increase. While this can be explained by the presence of a thapsigargin-insensitive store 174 in cerebellar granule cells, as has been suggested by others (Masgrau et al., 2000), this hypothesis was not tested. Thus, while the data presented in this dissertation provide strong evidence for participation of M3 muscarinic receptors, 1P3 receptors and the MTP in MeHg-induced elevations of [Ca2+]; and subsequent neuronal death, they do not yet provide a clear picture of the temporal order of the events which take place during the initial exposure period in cerebellar granule cells. Regardless of the source(s) of the initial increase of [Ca2+]i, in all experiments there is a latent period between exposure to MeHg and the increase in [Ca2+]; which has been observed in this dissertation and elsewhere (Hare et al., 1993; Marty and Atchison, 1997). This latent period decreases as the concentration of MeHg increases, suggesting a threshold for the Ca2+ response. The physiological basis for this threshold is unknown. Hypothetically, the threshold could depend on saturation of binding of MeHg to proteins not directly involved in the observed elevation of [Ca2+]i. For example, MeHg binds thiols contained in cytosolic GSH (Sarafian and Verity, 1991; Sanfeliu et al., 2001). Saturation of intracellular GSH would not only increase the concentration of non-bound MeHg in the cytosol but would also deplete the GSH reserve, thereby interfering with the cell’s ability to manage oxidative stress (Sanfeliu et al., 2001). As discussed above, oxidative stress contributes to opening of the MTP, thus MeHg-induced elevations of [Ca2+]i could potentially depend on depletion of GSH. The latent period could also result from saturation of extracellular, non-specific binding sites. Many plasma membrane proteins, including muscarinic and nicotinic ACh receptors, as well as NMDA receptors, contain cysteine residues which regulate receptor activity and which could potentially participate in non-specific binding of MeHg (Sato et al., 1976; Janaky et al., 1993; Zeng 175 et al. , 1999). Whereas none of these possibilities was tested directly in experiments presented in this dissertation, MeHg has been shown to impair function of neuronal nicotinic and muscarinic receptor-mediated responses in NIB-115 neuroblastoma cells (Quandt et al., 1982) and of glutamate receptors (Yuan and Atchison, 1997). The model presented in this dissertation is based on pharmacological manipulations which did not provide sufficient evidence to answer unequivocally the question of whether MeHg interacts with specific, potentially unique molecular targets in cerebellar granule cells. An alternative approach which could potentially provide this answer would be based on genetic manipulation of biological systems to add and/or remove specific proteins fiom cells and, when possible, experimental animals. Theoretically, a better approach would use transgenic methodology to create, for example, 1P3 knockout mice to remove the 1P3 receptor fi'om mouse cells and observe whether its removal alters the effects of MeHg in the whole animal and/or cells derived from these animals. However, experiments of this nature are limited by cost, the time needed to engineer genetically-altered animals, and the limited viability of some strains of knockout mice. Additionally, animals often compensate by substituting a different isoform of the receptor, which can complicate interpretation of data gathered fi'om this type of experiment. Thus, genetic manipulation provides a potentially enlightening experimental approach to determining the specific molecular target(s) of MeHg, but as observed with the pharmacological approaches, is also constrained by potential problems. In conclusion, this dissertation provides evidence that MeHg alters [Ca2+]; regulation by causing early, irreversible release of Ca2+ from both the SER and mitochondria to cause cell death in cerebellar granule cells. In cerebellar granule cells, 176 MeHg-induced neuronal death is prevented by inhibition of opening of the MTP or by inhibition of Ca2+ release through the 1P3 receptor, with protection falling off as the MeHg concentration exceeds 1 uM. It is interesting to note that, while MeHg has many effects which can be harmful (i. e., generation of reactive oxygen species, disruption of protein and ATP synthesis, etc.), the highly sensitive granule cells can be protected by interfering with the MeHg—induced release of Ca2+ fi'om intracellular stores (Sarafian et al., 1984; Sarafian, 1999). Because there is great diversity in the location and abundance of intracellular Ca2+-release channels, particularly the IP3 receptor, and the location of their upstream receptors, such as the muscarinic receptors, within different areas of the nervous system, this may contribute to the selective neurotoxicity of MeHg within the cerebellum. 177 REFERENCES ABD-ELFATTAH, A. S. and SHAMOO, A. E. (1981). Regeneration of a functionally active rat brain muscarinic receptor by D-penicillarnine after inhibition with methylmercury and mercuric chloride. Mol. Pharmacol. 20(3): 492-497. ABRAMSON, J. J., TRIMM, J. L., WEDEN, L. and SALAMA, G. (1983). Heavy metals induce rapid calcium release from sarcoplasmic reticulum vesicles isolated from skeletal muscle. Proc. Natl. Acad Sci. U. S. A. 80(6): 1526-1530. ALI, S. F., LEBEL, C. P. and BONDY, S. C. (1992). Reactive oxygen species formation as a biomarker of methylmercury and trimethyltin neurotoxicity. NeuroToxicoIogy 13(3): 637-648. ALNEMRI, E. S. and LITWACK, G. (1990). Activation of intemucleosomal DNA cleavage in human CEM lymphocytes by glucocorticoid and novobiocin. Evidence for a non-Ca2m-requiring mechanism(s). J. Biol. Chem. 265(28): 17323-17333. ALONSO, R., DIDIER, M. and SOUBRIE, P. (1990). [3H]N-methylscopolamine binding studies reveal M2 and M3 muscarinic receptor subtypes on cerebellar granule cells in primary culture. J. Neurochem. 55(1): 334-337. ANDREYEV, A. and F ISKUM, G. (1999). Calcium induced release of mitochondrial cytochrome c by different mechanisms selective for brain versus liver. Cell Death Differ. 6(9): 825-832. - ANKARCRONA, M, DYPBUKT, J. M, BONFOCO, E., ZHIVOTOVSKY, B, ORRENIUS, s, LIPTON, s. A. and NICOTERA, P. (1995). Glutamate-induced neuronal death: a succession of necrosis or apoptosis depending on mitochondrial frmction. Neuron 15(4): 961-973. ANKARCRONA, M., DYPBUKT, J. M, ORRENIUS, S. and NICOTERA, P. (1996). Calcineurin and mitochondrial function in glutamate-induced neuronal cell death. FEBS Lett. 394(3): 321-324. ARENDS, M. J ., MORRIS, R. G. and WYLLIE, A. H. (1990). Apoptosis. The role of the endonuclease. Am. J. Pathol. 136(3): 593-608. ARONICA, E., CONDORELLI, D. F., NICOLETTI, F., DELL'ALBANI, P., AMICO, C. and BALAZS, R. (1993). Metabotropic glutamate receptors in cultured cerebellar granule cells: developmental profile. J. Neurochem. 60(2): 559-565. 178 ARONICA, E., DELL'ALBANI, P, CONDORELLI, D. F., NICOLETI‘I, F., HACK, N. and BALAZS, R. (1993). Mechanisms underlying developmental changes in the expression of metabotropic glutamate receptors in cultured cerebellar granule cells: homologous desensitization and interactive effects involving N-methyl-D- aspartate receptors. Mol. Pharmacol. 44(5): 981-989. ARONICA, E., NICOLETII, F., CONDORELLI, D. F. and BALAZS, R. (1993). Pharmacological characterization of metabotropic glutamate receptors in cultured cerebellar granule cells. Neurochem. Res. 18(5): 605-612. ASCHNER, M. (1996). Astrocytes as modulators of mercury-induced neurotoxicity. NeuroToxicology 17(3-4): 663-669. ASCHNER, M. and ASCHNER, J. L. (1990). Mercury neurotoxicity: mechanisms of blood-brain barrier transport. Neurosci. Biobehav. Rev. 14(2): 169-176. ASCHNER, M, MULLANEY, K. J ., WAGONER, D, LASH, L. H. and KIMELBERG, H. K. (1994). Intracellular glutathione (GSH) levels modulate mercuric chloride (MC)- and methylmercuric chloride (MeHgCl)-induced amino acid release from neonatal rat primary astrocytes cultures. Brain Res. 664(1-2): 133-140. ASCHNER, M, YAO, C. P, ALLEN, J. W. and TAN, K. H. (2000). Methylmercury alters glutamate transport in astrocytes. Neurochem. Int. 37(2-3): 199-206. ATCHISON, W. D. (1986). Extracellular calcium-dependent and -independent effects of methylmercury on spontaneous and potassium-evoked release of acetylcholine at the neuromuscular junction. J. Pharmacol. Exp. Ther. 237(2): 672-680. ATCHISON, W. D, ADGATE, L. and BEAMAN, C. M. (1988). Effects of antibiotics on uptake of calcium into isolated nerve terminals. J. Pharmacol. Exp. Ther. 245(2): 394-401. ATCHISON, W. D, JOSHI, U. and THORNBURG, J. E. (1986). Irreversible suppression of calcium entry into nerve terminals by methylmercury. J. Pharmacol. Exp. Ther. 238(2): 618-624. ATCHISON, W. D. and NARAHASHI, T. (1982). Methylmercury-induced depression of neuromuscular transmission in the rat. Neuro Toxicology 3(3): 37-50. BAKIR, F., DAMLUJI, S. F., AMIN-ZAKI, L, MURTADHA, M, KHALIDI, A, AL- RAWI, N. Y, TIKRITI, S, DAHAHIR, H. I., CLARKSON, T. W, SMITH, J. C. and DOHERTY, R. A. (1973). Methylmercury poisoning in Iraq. Science 181(96): 230-241. 179 BAKIR, F., RUSTAM, H, TIKRITI, S, AL-DAMLUJI, S. F. and SHIHRISTANI, H. (1980). Clinical and epidemiological aspects of methylmercury poisoning. Postgrad Med. J. 56(651): 1-10. BALDI, F. (1997). Microbial transformation of mercury species and their importance in the biogeochemical cycle of mercury. Met. Ions Biol. syst. 34: 213-257. BEAVER, J. P. and WARIN G, P. (1994). Lack of correlation between early intracellular calcium ion rises and the onset of apoptosis in thymocytes. Immunol. Cell Biol. 72(6): 489-499. BERLIN, M. and ULLBERG, S. (1963). Accumulation and retention of mercury in the mouse. III: An autoradiographic comparison of methyl mercuric dicyandiamide with inorganic mercury. Arch Environ. Health 6: 610-616. BERNARDI, P. (1996). The permeability transition pore. Control points of a cyclosporin A- sensitive mitochondrial channel involved in cell death. Biochim. Biophys. Acta 1275(1-2): 5-9. BERNARDI, P, COLONNA, R., COSTANTINI, P, ERIKSSON, 0., FONTAINE, E., ICHAS, F ., MASSARI, S, NICOLLI, A, PETRONILLI, V. and SCORRANO, L. (1998). The mitochondrial permeability transition. Biofactors 8(3-4): 273-281. BERNARDI, P. and PETRONILLI, V. (1996). The permeability transition pore as a mitochondrial calcium release channel: a critical appraisal. J. Bioenerg. Biomembr. 28(2): 13 1 -l 38. BERNARDI, P, VASSANELLI, S, VERONESE, P, COLONNA, R., SZABO, I. and ZORATTI, M. (1992). Modulation of the mitochondrial permeability transition pore. Effect of protons and divalent cations. J. Biol. Chem. 267(5): 2934-2939. BERNARDI, P, VERONESE, P. and PETRONILLI, V. (1993). Modulation of the mitochondrial cyclosporin A-sensitive permeability transition pore. 1. Evidence for two separate Me2+ binding sites with opposing effects on the pore open probability. J. Biol. Chem. 268(2): 1005—1010. BERSTEIN, G, HAGA, K, HAGA, T. and ICHIYAMA, A. (1988). Agonist and antagonist binding of muscarinic acetylcholine receptors purified from porcine brain: interconversion of high- and low-affinity sites by sulfhydryl reagents. J. Neurochem. 50(6): 1687-1694. BEZPROZVANN Y, I, WATRAS, J. and EHRLICH, B. E. (1991). Bell-shaped calcium- response curves of Ins(l,4,5)P3- and calcium-gated channels from endoplasmic reticulum of cerebellum. Nature 351(6329): 751-754. 180 BLACKSHAW, S, SAWA, A, SHARP, A. H, ROSS, C. A, SNYDER, S. H. and KHAN, A. A. (2000). Type 3 inositol 1,4,5-trisphosphate receptor modulates cell death. FASEBJ. 14(10): 1375-1379. BONDY, S. C. and MCKBE, M. (1991). Disruption of the potential across the synaptosomal plasma membrane and mitochondria by neurotoxic agents. Toxicol. Lett. 58(1): 13-21. BROEKEMEIER, K. M, DEMPSEY, M. E. and PFEIFFER, D. R. (1989). Cyclosporin A is a potent inhibitor of the inner membrane permeability transition in liver mitochondria. J. Biol. Chem. 264(14): 7826-7830. BROSTROM, C. O. and BROSTROM, M. A. (1990). Calcium-dependent regulation of protein synthesis in intact mammalian cells. Annu. Rev. Physiol. 52: 577-590. BUDD, S. L. and NICHOLLS, D. G. (1996a). Mitochondria, calcium regulation, and acute glutamate excitotoxicity in cultured cerebellar granule cells. J. Neurochem. 67(6): 2282-2291. BUDD, S. L. and NICHOLLS, D. G. (1996b). A reevaluation of the role of mitochondria in neuronal Ca2+ homeostasis. J. Neurochem. 66(1): 403-411. BUNTINAS, L, GUNTER, K. K, SPARAGNA, G. C. and GUNTER, T. E. (2001). The rapid mode of calcium uptake into heart mitochondria (RaM): comparison to RaM in liver mitochondria. Biochim. Biophys. Acta 1504(2-3): 248-261. BYRNE, A. M, LEMASTERS, J. J. and NIEMINEN, A. L. (1999). Contribution of increased mitochondrial free Ca2+ to the mitochondrial permeability transition induced by tert-butylhydroperoxide in rat hepatocytes. Hepatology 29(5): 1523- 1531. CANDURA, S. M, D'AGOSTINO, G, CASTOLDI, A. F., MESSORI, E., LIUZZI, M, MANZO, L. and TONINI, M. (1997). Effects of mercuric chloride and methyl mercury on cholinergic neuromuscular transmission in the guinea-pig ileum. Pharmacol. T oxicol. 80(5): 218-224. CARRIER, G, BRUNET, R. C, CAZA, M. and BOUCHARD, M. (2001). A toxicokinetic model for predicting the tissue distribution and elimination of organic and inorganic mercury following exposure to methyl mercury in animals and humans. 1. Development and validation of the model using experimental data in rats. T oxicol. Appl. Pharmacol. 171(1): 38-49. CASTOLDI, A. F., BARNI, S, TURIN, I, GANDINI, C. and MANZO, L. (2000). Early acute necrosis, delayed apoptosis and cytoskeletal breakdown in cultured cerebellar granule neurons exposed to methylmercury. J. Neurosci. Res. 59(6): ; 775-7 87 . ‘ 181 CASTOLDI, A. F., CANDURA, S. M, COSTA, P, MANZO, L. and COSTA, L. G. (1996). Interaction of mercury compounds with muscarinic receptor subtypes in the rat brain. NeuroToxicology 17(3-4): 735-741. CASTOLDI, A. F., COCCINI, T, CECCATELLI, S. and MANZO, L. (2001). Neurotoxicity and molecular effects of methylmercury. Brain Res. Bull. 55(2): 197-203. CELIO, M. R. (1990). Calbindin D-28k and parvalbumin in the rat nervous system. Neuroscience 35(2): 375—475. CHAN, S. L. and MATTSON, M. P. (1999). Caspase and calpain substrates: roles in synaptic plasticity and cell death. J. Neurosci. Res. 58(1): 167-190. CHANG, L. W. (1980). Mercury. In: Experimental and Clinical Neurotoxicology (Spencer, P. S. and Schaumburg, H. H, eds.). The Williams and Wilkins Co., Baltimore, pp. 508-526. CHANG, L. W, GILBERT, M. and SPRECHER, J. (1978). Modification of methylmercury neurotoxicity by vitamin E. Environ. Res. 17(3): 356-366. CHANG, L. W. and HARTMANN, H. A. (1972). Blood-brain barrier dysfunction in experimental mercury intoxication. Acta Neuropathol. 21(3): 179-184. CHANG, L. W. and HARTMANN, H. A. (1972). Electron microscopic histochemical study on the localization and distribution of mercury in the nervous system after mercury intoxication. Exp. Neurol. 35(1): 122-137. CHANG, L. W, REUHL, K. R. and SPYKER, J. M. (1977). Ultrastructural study of the latent effects of methyl mercury on the nervous system after prenatal exposure. Environ. Res. 13(2): 171-185. CHAVEZ, E. and HOLGUIN, J. A. (1988). Mitochondrial calcium release as induced by Hg2+. J Biol Chem 263(8): 3582-3587. CHAVEZ, E., ZAZUETA, C, DIAZ, E. and HOLQUIN, J. A. 1989). Characterization by Hg2+ of two different pathways for mitochondrial Ca + release. Biochim. Biophys. Acta 986(1): 27-32. CHAVIS, P, NOONEY, J. M, BOCKAERT, J ., FAGNI, L, FELTZ, A. and BOSSU, J. L. (1995). Facilitatory coupling between a glutamate metabotropic receptor and dihydropyridine-sensitive calcium channels in cultured cerebellar granule cells. J. Neurosci. 15(1 Pt 1): 135-143. 182 CHERNYAK, B. V. and BERNARDI, P. (1996). The mitochondrial permeability transition pore is modulated by oxidative agents through both pyridine nucleotides and glutathione at two separate sites. Eur. J. Biochem. 238(3): 623-630. CHERNYAK, B. V, DEDOV, V. N. and CHERNYAK, V. (1995). Ca(2+)-tr'iggered membrane permeability transition in deenergized mitochondria fiom rat liver. FEBS Lett. 365(1): 75-78. CHE'ITY, C. S, RAJANNA, S, HALL, E., YALLAPRAGADA, P. R. and RAJANNA, B. (1996). In vitro and in vivo effects of lead, methyl mercury and mercury on inositol 1,4,5-trisphosphate and l,3,4,5-tetrakisphosphate receptor bindings in rat brain. Toxicol. Lett. 87(1): 11-17. CHIU, V. C, MOURING, D. and HAYNES, D. H. (1983). Action of mercurials on the active and passive transport properties of sarcoplasmic reticulum. J. Bioenerg. Biomembr. 15(1): 13-25. CHOI, B. H. (1989). The effects of methylmercury on the developing brain. Prog. Neurobiol. 32(6): 447-470. CHOI, B. H, LAPHAM, L. W, AMIN-ZAKI, L. and SALEEM, T. (1978). Abnormal neuronal migration, deranged cerebral cortical organization, and diffuse white matter astrocytosis of human fetal brain: a major effect of methylmercury poisoning in utero. J. Neuropathol. Exp. Neurol. 37(6): 719-733. CHOI, S. C. and BARTHA, R. (1994). Environmental factors affecting mercury methylation in estuarine sediments. Bull. Environ. Contam. T oxicol. 53(6): 805- 812. CLARKSON, T. W. (1972). The biological properties and distribution of mercury. Biochem. J. 130(2): 61P-63P. CLARKSON, T. W, AMIN-ZAKI, L. and AL-TIKRITI, S. K. (1976). An outbreak of methylmercury poisoning due to consumption of contaminated grain. Fed. Proc. 35(12): 2395-2399. COCCINI, T, RANDINE, G, CANDURA, S. M, NAPPI, R. E., PROCKOP, L. D. and MANZO, L. (2000). Low-level exposure to methylmercury modifies muscarinic cholinergic receptor binding characteristics in rat brain and lymphocytes: physiologic implications and new opportunities in biologic monitoring. Environ. Health Perspect. 108(1): 29-33. COHEN, J. J. and DUKE, R. C. (1984). Glucocorticoid activation of a calcium- dependent endonuclease in thymocyte nuclei leads to cell death. J Immunol. 132(1): 38-42. 183 CONTRERA, J. G, MCLESKEY, S. W, HOLOPAINEN, I, XU, J. and WOJCIK, W. J. (1993). Muscarinic m2 receptors in cerebellar granule cell cultures from rat: mechanism of short-term desensitization. J. Pharmacol. Exp. Ther. 265(1): 433- 440. COOKSON, M. R and SHAW, P. J. (1999). Oxidative stress and motor neurone disease. Brain Pathol. 9(1): 165-186. CORONADO, R., MORRISSETTE, J ., SUKHAREVA, M. and VAUGHAN, D. M. (1994). Structure and function of ryanodine receptors. Am. J. Physiol. 266(6 Pt 1): C1485-1504. CRUMP, K. S, VAN LANDINGHAM, C, SHAMLAYE, C, COX, C, DAVIDSON, P. W, MYERS, G. J. and CLARKSON, T. W. (2000). Benchmark concentrations for methylmercury obtained from the Seychelles Child Development Study. Environ. Health Perspect. 108(3): 257-263. CSORDAS, G, THOMAS, A. P. and HAJNOCZKY, G. (1999). Quasi-synaptic calcium signal transmission between endoplasmic reticulum and mitochondria. EMBO J. 18(1): 96-108. DAVIDSON, P. W, PALUMBO, D, MYERS, G. J ., COX, C, SHAMLAYE, C. F., SLOANE-REEVES, J ., CERNICHIARI, E., WILDING, G. E. and CLARKSON, T. W. (2000). Neurodevelopmental outcomes of Seychellois children from the pilot cohort at 108 months following prenatal exposure to methylmercury fi'om a maternal fish diet. Environ. Res. 84(1): 1-11. DE LUCA, A, WELLER, M, FREI, K. and FONTANA, A. (1996). Maturation- dependent modulation of apoptosis in cultured cerebellar granule neurons by cytokines and neurotrophins. Eur. J. Neurosci. 8(9): 1994-2005. DELLINGER, J. A, MEYERS, R. M, GEBHARDT, K. J. and HANSEN, L. K. (1996). The Ojibwa Health Study: fish residue comparisons for Lakes Superior, Michigan, and Huron. Toxicol. Ind Health 12(3-4): 393-402. DENNY, M. F. and ATCHISON, W. D. (1994). Methylmercury-induced elevations in intrasynaptosomal zinc concentrations: an 19F-NMR study. J. Neurochem. 63(1): 383-386. DENNY, M. F., HARE, M. F. and ATCHISON, W. D. (1993). Methylmercury alters intrasynaptosomal concentrations of endogenous polyvalent cations. T oxicol. Appl. Pharmacol. 122(2): 222-232. 184 DOBLE, A, HUBERT, J. P. and BLANCHARD, J. C. (1992). Pertussis toxin pretreatment abolishes the inhibitory effect of riluzole and carbachol on D- [3H]aspartate release from cultured cerebellar granule cells. Neurosci. Lett. 140(2): 251-254. DRUMMOND, R. M. and TUFT, R. A. (1999). Release of Ca2+ fi'om the sarcoplasmic reticulum increases mitochondrial [Ca2+] in rat pulmonary artery smooth muscle cells. J. Physiol. (Land) 516(Pt 1): 139-147. DUBINSKY, J. M. and ROTHMAN, S. M. (1991). Intracellular calcium concentrations during "chemical hypoxia" and excitotoxic neuronal injury. J. Neurosci. 11(8): 2545-2551. DUCHEN, M. R. (2000). Mitochondria and calcium: from cell signalling to cell death. J. Physiol. 529 Pt 1: 57-68. ELDEFRAWI, M. E., MANSOUR, N. A. and ELDEFRAWI, A. T. (1977). Interactions of acetylcholine receptors with organic mercury compounds. Adv. Exp. Med Biol. 84: 449-463. EXTON, J. H. (1990). Signaling through phosphatidylcholine breakdown. J. Biol. Chem. 265(1): 14. FARKAS, D. L, WEI, M. D, FEBBRORIELLO, P, CARSON, J. H. and LOEW, L. M. (1989). Simultaneous imaging of cell and mitochondrial membrane potentials. Biophys. J. 56(6): 1053-1069. FOHRMAN, E. B, DE ERAUSQUIN, G, COSTA, E. and WOJCIK, W. J. (1993). Muscarinic m3 receptors and dynamics of intracellular Ca2+ cerebellar granule neurons. Eur. J. Pharmacol. 245(3): 263-271. FRIBERG, H, CONNERN, C, HALESTRAP, A. P. and WIELOCH, T. (1999). Differences in the activation of the mitochondrial permeability transition among brain regions in the rat correlate with selective vulnerability. J. Neurochem. 72(6): 2488-2497. F RIBERG, L. and MOTTET, N. K. (1989). Accumulation of methylmercury and inorganic mercury in the brain. Biol. Trace Elem. Res. 21: 201-206. FUKAMAUCHI, F., AIHARA, O. and KUSAKABE, M. (2000). Intemalization and down-regulation of muscarinic acetylcholine receptors in cerebellar granule cells of tenascin-gene deficient mice. Neurochem. Int. 36(2): 153-158. FUKAMAUCHI, F., HOUGH, C. and CHUANG, D. M. (1991). Expression and agonist- induced down-regulation of mRNAs of m2- and m3- muscarinic acetylcholine receptors in cultured cerebellar granule cells. J. Neurochem. 56(2): 716-719. 185 FUKAMAUCHI, F., SAUNDERS, P. A, HOUGH, C. and CHUANG, D. M. (1993). Agonist-induced down-regulation and antagonist-induced up-regulation of m2- and m3-muscarinic acetylcholine receptor mRN A and protein in cultured cerebellar granule cells. Mol. Pharmacol. 44(5): 940-949. GALLI, C, MEUCCI, O, SCORZIELLO, A, WERGB, T. M, CALISSANO, P. and SCHETTINI, G. (1995). Apoptosis in cerebellar granule cells is blocked by high KCl, forskolin, and IGF-1 through distinct mechanisms of action: the involvement of intracellular calcium and RNA synthesis. J. Neurosci. 15(2): 1172-1179. GARMAN, R. H, WEISS, B. and EVANS, H. L. (1975). Alkylrnercurial encephalopathy in the monkey (Saimiri sciureus and Macaca arctoides): a histopathologic and autoradiographic study. Acta Neuropathol. 32(1): 61-74. GASSO, S, SUNOL, C, SANFELIU, C, RODRIGUEZ-PARKE, E. and CRISTOFOL, R M. (2000). Pharmacological characterization of the effects of methylmercury and mercuric chloride on spontaneous noradrenaline release from rat hippocampal slices. Life Sci. 67(10): 1219-1231. GAVIN, C. E., GUNTER, K. K. and GUNTER, T. E. (1999). Manganese and calcium transport in mitochondria: implications for manganese toxicity. NeuroToxicology 20(2-3): 445-453. GERSTNER, H. B. and HUFF, J. E. (1977). Selected case histories and epidemiologic examples of human mercury poisoning. Clin. Toxicol. 11(2): 131-150. GONZALEZ, B, LEROUX, P, LAMACZ, M, BODENANT, C, BALAZS, R. and VAUDRY, H. (1992). Somatostatin receptors are expressed by immature cerebellar granule cells: evidence for a direct inhibitory effect of somatostatin on neuroblast activity. Proc. Natl. Acad. Sci. U. S. A. 89(20): 9627-9631. GRANDJEAN, P, BUDTZ-JORGENSEN, E., WHITE, IL F., JORGENSEN, P. J., WEIHE, P, DEBES, F. and KEIDING, N. (1999). Methylmercury exposure biomarkers as indicators of neurotoxicity in children aged 7 years. Am. J. Epidemiol. 150(3): 301-305. GRANDJEAN, P, WHITE, R. F., NIELSEN, A, CLEARY, D. and DE OLIVEIRA SANTOS, E. C. (1999). Methylmercury neurotoxicity in Amazonian children downstream from gold mining. Environ. Health Perspect. 107(7): 587-591. GUERINI, D, GARCIA-MARTIN, E., ZECCA, A, GUIDI, F. and CARAFOLI, E. (1998). The calcium pump of the plasma membrane: membrane targeting, calcium binding sites, tissue-specific isoform expression. Acta Physiol. Scand. Suppl. 643: 265-273. 186 GUIMARAES, J. R., MEILI, M, HYLANDER, L. D, DE CASTRO E SILVA, E., ROULET, M, MAURO, J. B. and DE LEMOS, R. (2000). Mercury net methylation in five tropical flood plain regions of Brazil. Sci. Total Environ. 261(1-3): 99-107. GUNTER, K. K. and GUNTER, T. E. (1994). Transport of calcium by mitochondria. J. Bioenerg. Biomembr. 26(5): 471-485. GUNTER, T. E. and PFEIFFER, D. R. (1990). Mechanisms by which mitochondria transport calcium. Am. J. Physiol. 258(5 Pt 1): C755-786. HAJNOCZKY, G, HAGER, R. and THOMAS, A. P. (1999). Mitochondria suppress local feedback activation of inositol 1,4, 5- trisphosphate receptors by Ca”. J. Biol. Chem. 274(20): 14157-14162. HANSEN, J. C. (1990). Human exposure to heavy metals through consumption of marine foods: a case study of exceptionally high intake among Greenlanders. In: Heavy Metals in the Marine Environment (Furnow, A. W. and Rainbow, P. S, eds). CRC Press, New York, pp. 227-243. HARADA, M. (1995). Minarnata disease: methylmercury poisoning in Japan caused by environmental pollution Crit. Rev. Toxicol. 25(1): 1-24. HARE, M. F. and ATCHISON, W. D. (1992). Comparative action of methylmercury and divalent inorganic mercury on nerve terminal and intraterrninal mitochondrial membrane potentials. J. Pharmacol. Exp. Ther. 261(1): 166-172. HARE, M. F. and ATCHISON, w. D. (1995). Methylmercury mobilizes Ca” from intracellular stores sensitive to inositol 1,4,5-trisphosphate in NGlO8-15 cells. J. Pharmacol. Exp. Ther. 272(3): 1016-1023. HARE, M. F., MCGINNIS, K. M. and ATCHISON, W. D. (1993). Methylmercury increases intracellular concentrations of CaH and heavy metals in NGlO8-15 cells. J. Pharmacol. Exp. Ther. 266(3): 1626-1635. HE, L, POBLENZ, A. T, MEDRANO, C. J. and FOX, D. A. (2000). Lead and calcium produce rod photoreceptor cell apoptosis by opening the mitochondrial permeability transition pore. J. Biol. Chem. 275(16): 12175-12184. HEDLUND, B. and BARTFAI, T. (1979). The importance of thiol- and disulfide groups in agonist and antagonist binding to the muscarinic receptor. Mol. Pharmacol. 15(3): 531-544. HILLERED, L, MUCHIRI, P. M, NORDENBRAND, K. and ERNSTER, L. (1983). Mn2+ prevents the Ca2+-induced inhibition of ATP synthesis in brain mitochondria. F EBS Lett. 154(2): 247-250. 187 HUNT, J. B, NEECE, S. H, SCHACHMAN, H. K. and GINSBURG, A. (1984). Mercurial-promoted Zn2+ release from Escherichia coli aspartate transcarbamoylase. J. Biol. Chem. 259(23): 14793—14803. HUNTER, D. and RUSSELL, D. S. (1954). Focal cerebral and cerebellar atrophy in a human subject due to organic mercury compounds. Neurol. Neurosurg. Psychiat. 17: 169-235. ICHAS, F. and MAZAT, J. P. (1998). From calcium signaling to cell death: two conformations for the mitochondrial permeability transition pore. Switching from low- to high- conductance state. Biochim. Biophys. Acta 1366(1-2): 33-50. INSUG, O, DATAR, S, KOCH, C. J., SHAPIRO, I. M. and SHENKER, B. J. (1997). Mercuric compounds inhibit human monocyte function by inducing apoptosis: evidence for formation of reactive oxygen species, development of mitochondrial membrane permeability transition and loss of reductive reserve. Toxicology 124(3): 211-224. IRVING, A. J., COLLINGRIDGE, G. L. and SCHOFIELD, J. o. (1992). Interactions between Ca2+ mobilizing mechanisms in cultured rat cerebellar granule cells. J. Physiol. (Lond) 456: 667-680. IRVING, A. J ., COLLINGRIDGE, G. L. and SCHOFIELD, J. G. (1992). L-glutarnate and acetylcholine mobilise Ca2+ fi'om the same intracellular pool in cerebellar granule cells using transduction mechanisms with different Ca2+ sensitivities. Cell Calcium 13(5): 293-301. JANAKY, R., VARGA, V, SARANSAARI, P. and OJA, S. S. (1993). Glutathione modulates the N-methyl-D-aspartate receptor-activated calcium influx into cultured rat cerebellar granule cells. Neurosci. Lett. 156(1-2): 153-157. JAYARAMAN, T. and MARKS, A. R. (1997). T cells deficient in inositol 1,4,5- trisphosphate receptor are resistant to apoptosis. Mol. Cell Biol. 17(6): 3005-3012. JU AN G, M. S. (1976). Depression of frog muscle contraction by methylmercuric chloride and mercuric chloride. T oxicol. Appl. Pharmacol. 35(1): 183-185. JUAN G, M. S. (1976). An electrophysiological study of the action of methylmercuric chloride and mercru'ic chloride on the sciatic nerve-sartorius muscle preparation of the frog. Toxicol. Appl. Pharmacol. 37(2): 339-348. KANE, D. J ., ORD, T, ANTON, R. and BREDESEN, D. E. (1995). Expression of bcl-2 inhibits necrotic neural cell death. J. Neurosci. Res. 40(2): 269-275. KASS, G. E. and ORRENIUS, S. (1999). Calcium signaling and cytotoxicity. Environ. Health Perspect. 107 Suppl 1: 25-35. 188 KAUPPINEN, R. A, KOMULAINEN, H. and TAIPALE, H. (1989). Cellular mechanisms underlying the increase in cytosolic free calcium concentration induced by methylmercury in cerebrocortical synaptosomes from guinea pig. J. Pharmacol. Exp. Ther. 248(3): 1248-1254. KISELYOV, K, XU, X, MOZHAYEVA, G, KUO, T, PESSAH, I, MIGNERY, G, ZHU, X, BIRNBAUMER, L. and MUALLEM, S. (1998). Functional interaction between InsP3 receptors and store-operated Htrp3 channels. Nature 396(6710): 478-482. KISELYOV, K. I, SHIN, D. M, WANG, Y, PESSAH, I. N, ALLEN, P. D. and MUALLEM, S. (2000). Gating of store-operated channels by conformational coupling to ryanodine receptors. Mol. Cell 6(2): 421-431. KOCH, B. D, FAUROT, G. F., KOPANITSA, M. V. and SWINNEY, D. C. (1994). Pharmacology of a Ca(2+)-influx pathway activated by emptying the intracellular Ca2+ stores in HL-60 cells: evidence that a cytochrome P- 450 is not involved. Biochem. J. 302(Pt 1): 187-190. KOMULAINEN, H. and BONDY, S. C. (1987). Increased fiee intrasynaptosomal Ca2+ by neurotoxic organometals: distinctive mechanisms. Toxicol. Appl. Pharmacol. 88(1): 77-86. KOWALTOWSKI, A. J. and CASTILHO, R. F. (1997). Ca2+ acting at the external side of the inner mitochondrial membrane can stimulate mitochondrial permeability transition induced by phenylarsine oxide. Biochim. Biophys. Acta 1322(2-3): 221- 229. KROEMER, G. (1997). The proto-oncogene Bel-2 and its role in regulating apoptosis. Nat. Med 3(6): 614-620. KRUMAN, I. I. and MATTSON, M. P. (1999). Pivotal role of mitochondrial calcium uptake in neural cell apoptosis and necrosis. J. Neurochem. 72(2): 529-540. KUNIMOTO, M. (1994). Methylmercury induces apoptosis of rat cerebellar neurons in primary culture. Biochem. Biophys. Res. Commun. 204(1): 310-317. KURLAND, L. T. (1973). Methyl-mercury sources, mode of action and clinical and pathological effects on the developing nervous system. Res. Publ. Assoc. Res. Nerv. Ment. Dis. 51(10): 283-297. 189 KUZNETSOV, D. A, ZAVIJALOV, N. V, GOVORKOV, A. V. and SIBILEVA, T. M. (1987). Methyl mercury-induced nonselective blocking of phosphorylation processes as a possible cause of protein synthesis inhibition in vitro and in vivo. Toxicol. Lett. 36(2): 1 53-160. LANDOLFI, B, CURCI, S, DEBELLIS, L, POZZAN, T. and HOFER, A. M. (1998). Ca2+ homeostasis in the agonist-sensitive internal store: functional interactions between mitochondria and the ER measured In situ in intact cells. J. Cell Biol. 142(5): 1235-1243. LEBEL, C. P, ALI, S. F. and BONDY, S. C. (1992). Deferoxarnine inhibits methyl mercury-induced increases in reactive oxygen species formation in rat brain. T oxicol. Appl. Pharmacol. 112(1): 161-165. LEBEL, C. P, ALI, S. F., MCKEE, M. and BONDY, S. C. (1990). Organometal-induced increases in oxygen reactive species: the potential of 2',7'-dichlorofluorescin diacetate as an index of neurotoxic damage. Toxicol. Appl. Pharmacol. 104(1): 17-24. LEMASTERS, J. J ., NIEMINEN, A. L, QIAN, T, TROST, L. C, ELMORE, S. P, NISHIMURA, Y, CROWE, R. A, CASCIO, W. E., BRADHAM, C. A, BRENNER, D. A. and HERMAN, B. (1998). The mitochondrial permeability transition in cell death: a common mechanism in necrosis, apoptosis and autophagy. Biochim. Biophys. Acta 1366(1-2): 177-196. LEVESQUE, P. C. and ATCHISON, W. D. (1988). Effect of alteration of nerve terminal Ca2+ regulation on increased spontaneous quantal release of acetylcholine by methyl mercury. Toxicol. Appl. Pharmacol. 94(1): 55-65. LEVESQUE, P. C. and ATCHISON, W. D. (1991). Disruption of brain mitochondrial calcium sequestration by methylmercury. J. Pharmacol. Exp. Ther. 256(1): 236- 242. LEVESQUE, P. C, HARE, M. F. and ATCHISON, W. D. (1992). Inhibition of mitochondrial Ca2+ release diminishes the effectiveness of methyl mercury to release acetylcholine fiom synaptosomes. Toxicol. Appl. Pharmacol. 115(1): 11- 20. LEVICK, V, COFFEY, H. and D'MELLO, S. R. (1995). Opposing effects of thapsigargin on the survival of developing cerebellar granule neurons in culture. Brain Res. 676(2): 325-335. LEYSHON, K. and MORGAN, A. J. (1991). An integrated study of the morphological and gross-elemental consequences of methyl mercury intoxication in rats, with particular attention on the cerebellum. Scanning Microsc. 5(3): 895-904. 190 LEYSHON-SORLAND, K, JASANI, B. and MORGAN, A. J. (1994). The localization of mercury and metallothionein in the cerebellum of rats experimentally exposed to methylmercury. Histochem. J. 26(2): 161-169. LEYSSENS, A, NOWICKY, A. V, PATTERSON, L, CROMPTON, M. and DUCHEN, M. R. (1996). The relationship between mitochondrial state, ATP hydrolysis, [Mgzji and [Ca2+], studied in isolated rat cardiomyocytes. J. Physiol. 496 (Pt 1): 111-128. LIMKE, T. L. and ATCHISON, W. D. (2002). Acute exposure to methylmercury opens the mitochondrial permeability transition pore in rat cerebellar granule cells. Toxicol. Appl. Pharmacol. (in press). LIN, S. Z, YAN, G. M, KOCH, K. E., PAUL, S. M. and IRWIN, R. P. (1997). Mastoparan-induced apoptosis of cultured cerebellar granule neurons is initiated by calcium release from intracellular stores. Brain Res. 771(2): 184-195. LOEW, L. M, CARRINGTON, W, TUFT, R. A. and FAY, F. S. (1994). Physiological cytosolic Ca” transients evoke concurrent mitochondrial depolarizations. Proc. Natl. Acad. Sci. U. S. A. 91(26): 12579-12583. MA, H. T, PATTERSON, R. L, VAN ROSSUM, D. B, BIRNBAUMER, L, MIKOSHIBA, K. and GILL, D. L. (2000). Requirement of the inositol trisphosphate receptor for activation of store-operated Ca2+ channels. Science 287(5458): 1647-1651. MANEV, H, KHARLAMOV, E., UZ, T, MASON, R. P. and CAGNOLI, C. M. (1997). Characterization of zinc-induced neuronal death in primary cultures of rat cerebellar granule cells. Exp. Neurol. 146(1): 171-178. MARCAIDA, G, KOSENKO, E., MINANA, M. D, GRISOLIA, s. and FELIPO, v. (1996). Glutamate induces a calcineurin-mediated dephosphorylation of Na+,K(+’- ATPase that results in its activation in cerebellar neurons in culture. J. Neurochem. 66(1): 99-104. MARKS, A. R. (1997). Intracellular calcium-release channels: regulators of cell life and death. Am. J. Physiol. 272(2 Pt 2): H597-605. MARSH, D. O, CLARKSON, T. W, COX, C, MYERS, G. J ., AMIN-ZAKI, L. and AL-TIKRITI, S. (1987). Fetal methylmercury poisoning. Relationship between concentration in single strands of maternal hair and child effects. Arch. Neurol. 44(10): 1017-1022. MARTY, M. S. and ATCHISON, W. D. (1997). Pathways mediating Ca2+ entry in rat cerebellar granule cells following in vitro exposure to methyl mercury. T oxicoI. Appl. Pharmacol. 147(2): 319-330. 191 MARTY, M. S. and ATCHISON, W. D. (1998). Elevations of intracellular Ca2+ as a probable contributor to decreased viability in cerebellar granule cells following acute exposure to methylmercury. Toxicol. Appl. Pharmacol. 150(1): 98-105. MASGRAU, R., SERVITJA, J. M, SARRI, E., YOUNG, K. W, NAHORSKI, S. R. and PICATOSTE, F. (2000). Intracellular Ca2+ stores regulate muscarinic receptor stimulation of phospholipase C in cerebellar granule cells. J. Neurochem. 74(2): 818-826. MATTSON, M. P. (2000). Apoptosis in neurodegenerative disorders. Nat. Rev. Mol. Cell. Biol. 1(2): 120-129. MATTSON, M. P, LAFERLA, F. M, CHAN, S. L, LEISSRING, M. A, SHEPEL, P. N. and GEIGER, J. D. (2000). Calcium signaling in the ER: its role in neuronal plasticity and neurodegenerative disorders. Trends Neurosci. 23(5): 222-229. MEIER, E., DREJER, J. and SCHOUSBOE, A. (1984). GABA induces functionally active low-affinity GABA receptors on cultured cerebellar granule cells. J. Neurochem. 43(6): 1737-1744. MHYRE, T. R., MAINE, D. N. and HOLLIDAY, J. (2000). Calcium-induced calcium release from intracellular stores is developmentally regulated in primary cultures of cerebellar granule neurons. J. Neurobiol. 42(1): 134-147. MIURA, K, CLARKSON, T. W, IKEDA, K, NAGANUMA, A. and IMURA, N. (1994). Establishment and characterization of methylmercury-resistant PC12 cell line. Environ. Health Perspect. 102 Suppl 3: 313-315. MIURA, K. and IMURA, N. (1987). Mechanism of methylmercury cytotoxicity. Crit. Rev. Toxicol. 18(3): 161-188. MIYATA, H, SILVERMAN, H. s, SOLLOTI‘, s. J., LAKATTA, E. G, STERN, M. D. and HANSFORD, R. G. (1991). Measurement of mitochondrial free Ca2+ concentration in living single rat cardiac myocytes. Am J. Physiol. 261(4 Pt 2): H1123-1134. MOGENSEN, H. S. and JORGENSEN, O. S. (2000). AMPA receptor subunit mRNAs and intracellular [Ca(2+)] in cultured mouse and rat cerebellar granule cells. Int. J. Dev. Neurosci. 18(1): 61-68. MOLLER-MADSEN, B. (1990). Localization of mercury in CNS of the rat. II. Intraperitoneal injection of methyhnercuric chloride (CH3HgCl) and mercuric chloride (I-IgClz). Toxicol. Appl. Pharmacol. 103(2): 303-323. 192 MOLLER-MADSEN, B. (1991). Localization of mercury in CNS of the rat. III. Oral administration of methylmercuric chloride (CH3HgCl). F undam. Appl. T oxicol. 16(1): 172-187. MORAN, J , ITOH, T, REDDY, U. R., CHEN, M, ALNEMRI, E. S. and PLEASURE, D. (1999). Caspase-3 expression by cerebellar granule neurons is regulated by calcium and cyclic AMP. J. Neurochem. 73(2): 568-577. MORGANS, C. W, EL FAR, O, BERNTSON, A, WASSLE, H. and TAYLOR, W. R. (1998). Calcium extrusion from mammalian photoreceptor terminals. J. Neurosci. 18(7): 2467-2474. MUNDY, W. R. and FREUDENRICH, T. M. (2000). Sensitivity of immature neurons in culture to metal-induced changes in reactive oxygen species and intracellular free calcium. NeuroToxicology 21(6): 1 135-1 144. MUTO, H, SHINADA, M, TOKUTA, K. and TAKIZAWA, Y. (1991). Rapid changes in concentrations of essential elements in organs of rats exposed to methylmercury chloride and mercuric chloride as shown by simultaneous multielemental analysis. Br. J. Ind. Med. 48(6): 382-388. MYERS, G. J. and DAVIDSON, P. W. (1998). Prenatal methylmercury exposure and children: neurologic, developmental, and behavioral research. Environ Health Perspect. 106 Suppl 3: 841-847. MYERS, G. J ., DAVIDSON, P. W, SHAMLAYE, C. F., AXTELL, C. D, CERNICHIARI, E., CHOISY, O, CHOI, A, COX, C. and CLARKSON, T. W. (1997). Effects of prenatal methylmercury exposure from a high fish diet on developmental milestones in the Seychelles Child Development Study. NeuroToxicology 18(3): 819-829. MYERS, G. J., MARSH, D. O, COX, C, DAVIDSON, P. W, SHAMLAYE, C. F., TANNER, M. A, CHOI, A, CERNICHIARI, E., CHOISY, O. and CLARKSON, T. W. (1995). A pilot neurodevelopmental study of Seychellois children following in utero exposure to methylmercury from a maternal fish diet. Neuro Toxicology 16(4): 629-638. NAGASHIMA, K, FUJII, Y, TSUKAMOTO, T, NUKUZUMA, S, SATOH, M, FUJITA, M, FUJIOKA, Y. and AKAGI, H. (1996). Apoptotic process of cerebellar degeneration in experimental methylmercury intoxication of rats. Acta Neuropathol. 91(1): 72-77. NAKADA, S. and IMURA, N. (1983). Susceptibility of lipids to mercurials. J. Appl. T oxicol. 3(3): 131-134. 193 NAKAMURA, K, BOSSY-WETZEL, E., BURNS, K, FADEL, M. P, LOZYK, M, GOPING, I. S, OPAS, M, BLEACKLEY, R. C, GREEN, D. R. and MICHALAK, M. (2000). Changes in endoplasmic reticulum luminal environment affect cell sensitivity to apoptosis. J. Cell Biol. 150(4): 731-740. NAKATSURU, S, OOHASHI, J ., NOZAKI, H, NAKADA, S. and IMURA, N. (1985). Effect of mercurials on lymphocyte functions in vitro. Toxicology 36(4): 297-305. NEATHERY, M. W. and MILLER, W. J. (1975). Metabolism and toxicity of cadmium, mercury, and lead in animals: a review. J. Dairy Sci. 58(12): 1767-1781. NEUSTADT, A, FROSTHOLM, A. and ROTTER, A. (1988). Topographical distribution of muscarinic cholinergic receptors in the cerebellar cortex of the mouse, rat, guinea pig, and rabbit: a species comparison. .1. Comp. Neurol. 272(3): 317-330. NICHOLLS, D. G. (1974).The influence of respiration and ATP hydrolysis on the protonelectrochemical gradient across the inner membrane of rat liver mitochondria as determined by ion distribution. Eur. J. Biochem. 50: 305-315. NICHOLLS, D. G. (1978). The regulation of extramitochondrial free calcium ion concentration by rat liver mitochondria. Biochem. J. 176(2): 463 -474. NICHOLLS, D. G. and BUDD, S. L. (1998). Mitochondria and neuronal glutamate excitotoxicity. Biochim. Biophys. Acta 1366(1-2): 97-112. NICHOLLS, D. G, BUDD, S. L, CASTILHO, R. F. and WARD, M. W. (1999). Glutamate excitotoxicity and neuronal energy metabolism. Ann. N. Y. Acad Sci. 893: 1-12. NICHOLLS, D. G. and SCO'I'I‘, I. D. (1980). The regulation of brain mitochondrial calcium-ion transport. The role of ATP in the discrimination between kinetic and membrane-potential- dependent calcium-ion efflux mechanisms. Biochem. J. 186(3): 833-839. NIEMINEN, A. L, PETRIE, T. G, LEMASTERS, J. J. and SELMAN, W. R. (1996). Cyclosporin A delays mitochondrial depolarization induced by N-methyl-D- aspartate in cortical neurons: evidence of the mitochondrial permeability transition. Neuroscience 75(4): 993-997. NIEMINEN, A. L, SAYLOR, A. K, TESFAI, S. A, HERMAN, B. and LEMASTERS, J. J. (1995). Contribution of the mitochondrial permeability transition to lethal injury after exposure of hepatocytes to t-butylhydroperoxide. Biochem. J. 307(Pt 1 ): 99-1 06. 194 OBERDORF, J ., VALLANO, M. L. and WOJCIKIEWICZ, R J. (1997). Expression and regulation of types I and H inositol 1,4,5- trisphosphate receptors in rat cerebellar granule cell preparations. J. Neurochem. 69(5): 1897-1903. O'KUSKY, J. (1983). Methylmercury poisoning of the developing nervous system: morphological changes in neuronal mitochondria. Acta Neuropathol. 61(2): 116- 122. ORRENIUS, S. and NICOTERA, P. (1994). The calcium ion and cell death. J. Neural T ransm. Suppl. 43: 1-11. OU, Y. C, WHITE, C. C, KREJSA, C. M, PONCE, R. A, KAVANAGH, T. J. and FAUSTMAN, E. M. (1999). The role of intracellular glutathione in methyhnercury-induced toxicity in embryonic neuronal cells. Neuro Toxicology 20(5): 793-804. OYAKE, Y, TANAKA, M, KUBO, H. and CHICHIBU, M. (1966). [Neuropathological studies on organic mercury poisoning with special reference to the staining and distribution of mercury granules]. Shinlcei Kenkyu No Shimpo 10(4): 744-750. OYAMA, Y, CARPENTER, D. O, UENO, S, HAYASHI, H. and TOMIYOSHI, F. (1995). Methylmercury induces Ca(2+)'dependent hyperpolarization of mouse thymocytes: a flow cytometric study using fluorescent dyes. Eur. J. Pharmacol. 293(2): 101-107. OYAMA, Y, TOMIYOSHI, F ., UENO, S, FURUKAWA, K. and CHIKAHISA, L. (1994). Methylmercury-induced augmentation of oxidative metabolism in cerebellar neurons dissociated from the rats: its dependence on intracellular Ca2+. Brain Res. 660(1): 154-157. PALUMBO, D. R., COX, C, DAVIDSON, P. W, MYERS, G. J ., CHOI, A, SHAMLAYE, C, SLOANE-REEVES, J ., CERNICHIARI, E. and CLARKSON, T. W. (2000). Association between prenatal exposure to methylmercury and cognitive functioning in Seychellois children: a reanalysis of the McCarthy Scales of Children's Ability fiom the main cohort study. Environ. Res. 84(2): 81-88. PAN, Z. Z. and WILLIAMS, J. T. (1994) Muscarine hyperpolarizes a subpopulation of neurons by activating an M2 muscarinic receptor in rat nucleus raphe magnus in vitro. J. Neurosci. 14: 1332-1338. PARK, S. T, LIM, K. T, CHUNG, Y. T. and KIM, S. U. (1996). Methylmercury- induced neurotoxicity in cerebral neuron culture is blocked by antioxidants and NMDA receptor antagonists. NeuroToxicology 17(1): 37-45. 195 PASTORINO, J. G, SNYDER, J. W, HOEK, J. B. and FARBER, J. L. (1995). Ca2+ depletion prevents anoxic death of hepatocytes by inhibiting mitochondrial permeability transition. Am. J. Physiol. 268(3 Pt 1): C676-685. PASTORINO, J. G, TAFANI, M, ROTHMAN, R. J ., MARCINKEVICIUTE, A, HOEK, J. B, FARBER, J. L. and MARCINEVICIUTE, A. (1999). Functional consequences of the sustained or transient activation by Bax of the mitochondrial permeability transition pore. J. Biol. Chem. 274(44): 31734-31739. PEITSCH, M. C, POLZAR, B, STEPHAN, H, CROMPTON, T, MACDONALD, H. R., MANNHERZ, H. G. and TSCHOPP, J. (1993). Characterization of the endogenous deoxyribonuclease involved in nuclear DNA degradation during apoptosis (programmed cell death). EMBO J. 12(1): 371-377. PENG, T. 1., JOU, M. J., SHEU, S. S. and GREENAMYRE, J. T. (1998). Visualization of NMDA receptor-induced mitochondrial calcium accumulation in striatal neurons. Exp. Neurol. 149(1): 1-12. PETRONILLI, V, COLA, C. and BERNARDI, P. (1993). Modulation of the mitochondrial cyclosporin A-sensitive permeability transition pore. II. The minimal requirements for pore induction underscore a key role for transmembrane electrical potential, matrix pH, and matrix Ca2+. J. Biol. Chem. 268(2): 1011- 101 6. PETRONILLI, V, COSTANTINI, P, SCORRANO, L, COLONNA, R., PASSAMONTI, S. and BERNARDI, P. (1994). The voltage sensor of the mitochondrial permeability transition pore is tuned by the oxidation-reduction state of vicinal thiols. Increase of the gating potential by oxidants and its reversal by reducing agents. J. Biol. Chem. 269(24): 16638-16642. PETRONILLI, V, PENZO, D, SCORRANO, L, BERNARDI, P. and DI LISA, F. (2001). The mitochondrial permeability transition, release of cytochrome c and cell death. Correlation with the duration of pore openings in situ. J. Biol. Chem. 276(15): 12030-12034. PINTON, P, BRINI, M, BASTIANU'ITO, C, TUFT, R. A, POZZAN, T. and RIZZUTO, R. (1998). New light on mitochondrial calcium. Biofactors 8(3-4): 243-253. PIZZI, M, RIBOLA, M, VALERIO, A, MEMO, M. and SPANO, P. (1991). Various Ca2+ entry blockers prevent glutamate-induced neurotoxicity. Eur. J. Pharmacol. 209(3): 169-173. PUTNEY, J. W, JR. (1986). A model for receptor-regulated calcium entry. Cell Calcium 7(1): 1-12. 196 QUANDT, F. N, KATO, E. and NARAHASHI, T. (1982). Effects of methylmercury on electrical responses of neuroblastoma cells. NeuroToxicology 3(4): 205-220. RABENSTEIN, D. L. and EVANS, C. A. (1978). The mobility of methylmercury in biological systems. Bioinorg. Chem. 8(2): 107-101. RAFFRAY, M. and COHEN, G. M. (1997). Apoptosis and necrosis in toxicology: a continuum or distinct modes of cell death? Pharmacol. Ther. 75(3): 153-177. RANDALL, A. and TSIEN, R. W. (1995). Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J. Neurosci. 15(4): 2995-3012. RANDALL, A. D. and TSIEN, R. W. (1997). Contrasting biophysical and pharmacological properties of T-type and R- type calcium channels. Neuropharmacology 36(7): 879-893. REDMAN, R. S. and SILINSKY, E. M. (1994). Decrease in calcium currents induced by aminoglycoside antibiotics in frog motor nerve endings. Br. J. Pharmacol. 113(2): 375-378. RENZONI, A, ZINO, F. and FRANCHI, E. (1998). Mercury levels along the food chain and risk for exposed populations. Environ. Res. 77(2): 68-72. RESINK, A, HACK, N, BOER, G. J. and BALAZS, R. (1994). Growth conditions differentially modulate the vulnerability of developing cerebellar granule cells to excitatory amino acids. Brain Res. 655(1-2): 222-232. RICE, D. C. (1989). Brain and tissue levels of mercury after chronic methylmercury exposure in the monkey. J. T oxicol. Environ. Health 27(2): 189-198. RICE, G, SWARTOUT, J ., MAHAFFEY, K. and SCHOENY, R. (2000). Derivation of US. EPA's oral Reference Dose (RfD) for methylmercury. Drug Chem. T oxicol. 23(1): 41-54. RICHTER, C. and KASS, G. E. (1991). Oxidative stress in mitochondria: its relationship to cellular Ca2+ homeostasis, cell death, proliferation, and differentiation. Chem. Biol. Interact. 77(1): 1-23. RIZZUTO, R., PINTON, P, CARRINGTON, W, FAY, F. S, FOGARTY, K. E., LIFSHITZ, L. M, I UP I, R. A. and POZZAN, T. (1998). Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses. Science 280(5370): 1763-1766. ROBERTS, L. R, STEINRAUF, L. K. and BLICKENSTAF F, R. T. (1980). New evaluation of potential methylmercury scavengers. J. Pharm. Sci. 69(8): 964-967. 197 ROBERTSON, J. D. and ORRENIUS, S. (2000). Molecular mechanisms of apoptosis induced by cytotoxic chemicals. Crit. Rev. T oxicol. 30(5): 609-627. RODIER, P. M. (1995). Developing brain as a target of toxicity. Environ. Health Perspect. 103 Suppl 6: 73-76. RODIER, P. M, ASCHNER, M. and SAGER, P. R. (1984). Mitotic arrest in the developing CNS after prenatal exposure to methylmercury. Neurobehav. T oxicol. T eratol. 6(5): 379-3 85. RUSTAM, H, VON BURG, R., AMIN-ZAKI, L. and EL HASSANI, S. (1975). Evidence for a neuromuscular disorder in methylmercury poisoning. Arch. Environ. Health 30(4): 190-195. SAKAMOTO, M, IKEGAMI, N. and NAKANO, A. (1996). Protective effects of Ca2+ channel blockers against methyl mercury toxicity. Pharmacol. Toxicol. 78(3): 193-199. SANFELIU, C, SEBASTIA, J. and K1, S. U. (2001). Methylmercury neurotoxicity in cultures of human neurons, astrocytes, neuroblastoma cells. NeuroToxicology 22(3): 317-327. SARAFIAN, T, HAGLER, J ., VARTAVARIAN, L. and VERITY, M. A. (1989). Rapid cell death induced by methyl mercury in suspension of cerebellar granule neurons. J. Neuropathol. Exp. Neurol. 48(1): 1-10. SARAFIAN, T. and VERITY, M. A. (1985). Inhibition of RNA and protein synthesis in isolated cerebellar cells by in vitro and in vivo methyl mercury. Neurochem. Pathol. 3(1): 27-39. SARAF IAN, T. and VERITY, M. A. (1991). Oxidative mechanisms underlying methyl mercury neurotoxicity. Int. J. Dev. Neurosci. 9(2): 147-153. SARAF IAN , T. A. (1993). Methyl mercury increases intracellular Ca2+ and inositol phosphate levels in cultured cerebellar granule neurons. J. Neurochem. 61(2): 648-657. SARAF IAN, T. A. (1999). Methylmercury-induced generation of free radicals: biological implications. Met. Ions Biol. Syst. 36: 415-444. SARAFIAN, T. A, CHEUNG, M. K. and VERITY, M. A. (1984). In vitro methyl mercury inhibition of protein synthesis in neonatal cerebellar perikarya. Neuropathol. Appl. Neurobiol. 10(2): 85-100. 198 SARAFIAN, T. A, VARTAVARIAN, L, KANE, D. J., BREDESEN, D. E. and VERITY, M. A. (1994). bcl-2 expression decreases methyl mercury-induced free- radical generation and cell killing in a neural cell line. Toxicol. Lett. 74(2): 149- 155. SATO, T, SATO, M. and SAWADA, M. (1976). Effects of disulfide bond reduction on the excitatory and inhibitory postsynaptic responses of Aplysia ganglion cells. Jpn. J. Physiol. 26(5): 471-485. SCHINDER, A. F., OLSON, E. C, SPITZER, N. C. and MONTAL, M. (1996). Mitochondrial dysfunction is a primary event in glutamate neurotoxicity. J. Neurosci. 16(19): 6125-6133. SCORRANO, L, PETRONILLI, V. and BERNARDI, P. (1997). On the voltage dependence of the mitochondrial permeability transition pore. A critical appraisal. J. Biol. Chem. 272(19): 12295-12299. SENSI, S. L, YIN, H. Z. and WEISS, J. H. (2000). AMPA/kainate receptor-triggered Zn” entry into cortical neurons induces mitochondrial Zn2+ uptake and persistent mitochondrial dysfunction. Eur. J Neurosci. 12(10): 3813-3818. SHAMOO, A. E., MACLENNAN, D. H. and ELDERFRAWI, M. E. (1976). Differential effects of mercurial compounds on excitable tissues. Chem. Biol. Interact. 12(1): 41-52. SHARP, A. H, NUCIFORA, F. C, JR, BLONDEL, O, SHEPPARD, C. A, ZHANG, C, SNYDER, S. H, RUSSELL, J. T, RYUGO, D. K. and ROSS, C. A. (1999). Differential cellular expression of isoforms of inositol 1,4,5- triphosphate receptors in neurons and glia in brain. J. Comp. Neurol. 406(2): 207-220. SHENKER, B. J., BERTHOLD, P, DECKER, S, MAYRO, J., ROONEY, C, VITALE, L. and SHAPIRO, I. M. (1992). Immunotoxic effects of mercuric compounds on human lymphocytes and monocytes. II. Alterations in cell viability. Immunopharmacol. Immunotoxicol. 14(3): 555-577. SHENKER, B. J ., BERTHOLD, P, ROONEY, C, VITALE, L, DEBOLT, K. and SHAPIRO, I. M. (1993). Immunotoxic effects of mercuric compounds on human lymphocytes and monocytes. III. Alterations in B-cell function and viability. Immunopharmacol. Immunotoxicol. 15(1): 87-112. SHENKER, B. J ., GUO, T. L, O, I. and SHAPIRO, I. M. (1999). Induction of apoptosis in human T-cells by methyl mercury: temporal relationship between mitochondrial dysfunction and loss of reductive reserve. Toxicol. Appl. Pharmacol. 157(1): 23-35. 199 SHENKER, B. J ., GUO, T. L. and SHAPIRO, I. M. (1998). Low-level methylmercury exposure causes human T-cells to undergo apoptosis: evidence of mitochondrial dysfunction. Environ. Res. 77(2): 149-159. SHENKER, B. J ., GUO, T. L. and SHAPIRO, I. M. (2000). Mercury-induced apoptosis in human lymphoid cells: evidence that the apoptotic pathway is mercurial species dependent. Environ. Res. 84(2): 89-99. SIMBULA, G, GLASCOTT, P. A, JR, AKIT A, S, HOEK, J. B. and FARBER, J. L. (1997). Two mechanisms by which ATP depletion potentiates induction of the mitochondrial permeability transition. Am. J. Physiol. 273(2 Pt 1): C479-488. SIMPSON, P. B, CHALLISS, R A. and NAHORSKI, S. R. (1993). Involvement of intracellular stores in the Ca2+ responses to N-Methyl-D- aspartate and depolarization in cerebellar granule cells. J. Neurochem. 61(2): 760-763. SIMPSON, P. B, CHALLISS, R. A. and NAHORSKI, S. R. (1994). Chronic activation of muscarinic and metabotropic glutamate receptors down-regulates type I inositol 1,4,5-trisphosphate receptor expression in cerebellar granule cells. J. Neurochem. 63(6): 2369-2372. SIMPSON, P. B, CHALLISS, R A. and NAHORSKI, S. R. (1995). Divalent cation entry in cultured rat cerebellar granule cells measured using Mn2+ quench of fura 2 fluorescence. Eur. J Neurosci. 7(5): 831-840. SIMPSON, P. B, CHALLISS, R. A. and NAHORSKI, S. R. (1995). Neuronal Ca2+ stores: activation and function. Trends Neurosci. 18(7): 299-306. SIMPSON, P. B, NAHORSKI, S. R. and CHALLISS, R. A. (1996). Agonist-evoked Ca2+ mobilization from stores expressing inositol 1,4,5- trisphosphate receptors and ryanodine receptors in cerebellar granule neurones. J. Neurochem. 67(1): 364-373. SIMPSON, P. B. and RUSSELL, J. T. (1996). Mitochondria support inositol 1,4,5- trisphosphate-mediated Ca2+ waves in cultured oligodendrocytes. J. Biol. Chem. 271(52): 33493-33501. SIMPSON, P. B. and RUSSELL, J. T. (1998a). Mitochondrial Ca2+ uptake and release influence metabotropic and ionotropic cytosolic Ca2+ responses in rat oligodendrocyte progenitors. J. Physiol. (Land) 508(Pt 2): 413-426. SIMPSON, P. B. and RUSSELL, J. T. (1998b). Role of mitochondrial Ca2+ regulation in neuronal and glial cell signalling. Brain Res. Brain Res. Rev. 26(1): 72-81. 200 SIROIS, J. E. and ATCHISON, W. D. (2000). Methylmercury affects multiple subtypes of calcium channels in rat cerebellar granule cells. T oxicol. Appl. Pharmacol. 167(1): 1-11. SOMJEN, G. G, HERMAN, S. P, KLEIN, R, BRUBAKER, P. E., BRINER, W. H, GOODRICH, J. K, KRIGMAN, M. R. and HASEMAN, J. K. (1973). The uptake of methyl mercury (203 Hg) in different tissues related to its neurotoxic effects. J. Pharmacol. Exp. Ther. 187(3): 602-611. SOMLYO, A. P, BOND, M. and SOMLYO, A. V. (1985). Calcium content of mitochondria and endoplasmic reticulum in liver frozen rapidly in viva. Nature 314(6012): 622-625. SONE, N, LARSS'I'U VOLD, M. K. and KAGAWA, Y. (1977). Effect of methyl mercury on phosphorylation, transport, and oxidation in mammalian mitochondria. J. Biochem. (Tokyo) 82(3): 859-868. SPARAGNA, G. C, GUNTER, K. K, SHEU, S. S. and GUNTER, T. E. (1995). Mitochondrial calcium uptake fi'om physiological-type pulses of calcium. A description of the rapid uptake mode. J. Biol. Chem. 270(46): 27510-27515. STEINWALL, O. and KLATZO, I. (1966). Selective vulnerability of the blood-brain barrier in chemically induced lesions. J. Neuropathol. Exp. Neurol. 25(4): 542- 559. STEUERWALD, U, WEIHE, P, JORGENSEN, P. J , BJERVE, K, BROCK, J ., HEINZOW, B, BUDTZ-JORGEN SEN, E. and GRANDJEAN, P. (2000). Maternal seafood diet, methylmercury exposure, and neonatal neurologic function. J. Pediatr. 136(5): 599-605. STRINGFELLOW, T. L, MARTY, M. S, FRITZ, E. L. AND ATCHISON, W. D. (2000). K+ depolarization delays methylmercury-induced elevations of cerebellar granule cell intracellular calcium independently of changes in plasma membrane and mitochondrial membrane potentials. Soc. for Neurosci. (Abstracts) 26: 598. SYVERSEN, T. L, TOTLAND, G. and FLOOD, P. R. (1981). Early morphological changes in rat cerebellum caused by a single dose of methyhnercury. Arch. Toxicol. 47(2): 101-111. SZABO, I. and ZORATTI, M. (1992). The mitochondrial megachannel is the permeability transition pore. J. Bioenerg. Biomembr. 24(1): 111-117. SZALAI, G, KRISHNAMURTHY, R. and HAJNOCZKY, G. (1999). Apoptosis driven by IP(3)-linked mitochondrial calcium signals. EMBO J. 18(22): 6349-6361. 201 TAKEUCHI, T. (1982). Pathology of Minamata disease. With special reference to its pathogenesis. Acta Pathol. Jpn. 32(Suppl. 1): 73-99. TAKEUCHI, T, MORIKAWA, N, MATSUMOTO, H. and SHIRAISHI, Y. (1962). A pathological study of Minimata disease in Japan. Acta Neuropathol. 2: 40-57. TAN, X. X, TANG, C, CASTOLDI, A. F., MANZO, L. and COSTA, L. G. (1993). Effects of inorganic and organic mercury on intracellular calcium levels in rat T lymphocytes. J. T oxicol. Environ. Health 38(2): 159-170. THASTRUP, O, CULLEN, P. J ., DROBAK, B. K, HANLEY, M. R. and DAWSON, A. P. (1990). Thapsigargin, a tumor promoter, discharges intracellular Ca2+ stores by specific inhibition of the endoplasmic reticulum Ca w-ATPase. Proc. Natl. Acad Sci. U. S. A. 87(7): 2466-2470. THAYER, S. A. and MILLER, R. J. (1990). Regulation of the intracellular free calcium concentration in single rat dorsal root ganglion neurones in vitro. J. Physiol. (Land) 425: 85-115. TROST, L. C. and LEMASTERS, J. J. (1997). Role of the mitochondrial permeability transition in salicylate toxicity to cultured rat hepatocytes: implications for the pathogenesis of Reye's syndrome. T oxicol. Appl. Pharmacol. 147(2): 431-441. URUSHITANI, M, NAKAMIZO, T, INOUE, R., SAWADA, H, KIHARA, T, HONDA, K, AKAIKE, A. and SHIMOHAMA, S. (2001). N-methyl-D-aspartate receptor-mediated mitochondrial Cam) overload in acute excitotoxic motor neuron death: a mechanism distinct from chronic neurotoxicity after Ca0+) influx. J. Neurosci. Res. 63(5): 377-387. USACHEV, Y. M. and THAYER, s. A. (1999). Ca2+ influx in resting rat sensory neurones that regulates and is regulated by ryanodine-sensitive Ca2+ stores. J. Physiol. (Land) 519 Pt 1: 115-130. USUKI, F., YASUTAKE, A, UMEHARA, F., TOKUNAGA, H, MATSUMOTO, M, ETO, K, ISHIURA, S. and HIGUCHI, I. (2001). In vivo protection of a water- soluble derivative of vitamin E, Trolox, against methylmercury-intoxication in the rat. Neurosci. Lett. 304(3): 199-203. VALLEE, B. L. and AULD, D. S. (1990). Active-site zinc ligands and activated H20 of zinc enzymes. Proc. Natl. Acad Sci. U. S. A. 87(1): 220-224. VERITY, M. A. (1999). Manganese neurotoxicity: a mechanistic hypothesis. NeuroToxicalogy 20(2-3): 489-497. 202 VERITY, M. A, BROWN, W. J. and CHEUNG, M. (1975). Organic mercurial encephalopathy: in vivo and in vitro effects of methyl mercury on synaptosomal respiration. J. Neurochem. 25(6): 759-766. VERITY, M. A, BROWN, W. J., CHEUNG, M. and CZER, G. (1977). Methyl mercury inhibition of synaptosome and brain slice protein synthesis: in vivo and in vitro studies. J. Neurochem. 29(4): 673-679. VERITY, M. A, SARAF IAN , T, PACIFICI, E. H. and SEVANIAN, A. (1994). Phospholipase A2 stimulation by methyl mercury in neuron culture. J. Neurochem. 62(2): 705-714. VON BURG, R., NORTHINGTON, F. K. and SHAMOO, A. (1980). Methylmercury inhibition of rat brain muscarinic receptors. T oxicol. Appl. Pharmacol. 53(2): 285-292. WANG, G. J. and THAYER, S. A. (1996). Sequestration of glutamate-induced Ca2+ loads by mitochondria in cultured rat hippocampal neurons. J. Neurophysiol. 76(3): 1 61 l -1 62 1 . WARING, P. and SJAARDA, A. (1995). Extracellular calcium is not required for gliotoxin or dexamethasone- induced DNA fragmentation: a reappraisal of the use of EGTA. Int. J. Immunopharmacol. 17(5): 403-410. WEI, H, LEEDS, P, CHEN, R. W, WEI, W, LENG, Y, BREDESEN, D. E. and CHUAN G, D. M. (2000). Neuronal apoptosis induced by pharmacological concentrations of 3- hydroxykynurenine: characterization and protection by dantrolene and Bel-2 overexpression. J. Neurochem. 75(1): 81-90. WEI, H. and PERRY, D. C. (1996). Dantrolene is cytoprotective in two models of neuronal cell death. J. Neurochem. 67(6): 2390-2398. WEI, H, WEI, W, BREDESEN, D. E. and PERRY, D. C. (1998). Bel-2 protects against apoptosis in neuronal cell line caused by thapsigargin-induced depletion of intracellular calcium stores. J. Neurochem. 70(6): 2305-2314. WEIHE, P, GRANDJEAN, P, DEBES, F. and WHITE, R. (1996). Health implications for Faroe islanders of heavy metals and PCBs from pilot whales. Sci. Total Environ. 186(1-2): 141-148. WERTH, J. L, USACHEV, Y. M. and THAYER, S. A. (1996). Modulation of calcium efflux fi'om cultured rat dorsal root ganglion neurons. J. Neurosci. 16(3): 1008- 1015. 203 WHITE, R. J. and REYNOLDS, I. J. (1996). Mitochondrial depolarization in glutamate- stirnulated neurons: an early signal specific to excitotoxin exposure. J. Neurosci. 16(18): 5688-5697. WHITHAM, E. M, CHALLISS, R. A. and NAHORSKI, S. R (1991). M3 muscarinic cholinoceptors are linked to phosphoinositide metabolism in rat cerebellar granule cells. Eur. J. Pharmacol. 206(3): 181-189. WOJCIKIEWICZ, R. J ., FURUICHI, T, NAKADE, S, MIKOSHIBA, K. and NAHORSKI, S. R. (1994). Muscarinic receptor activation down-regulates the type I inositol 1,4,5- trisphosphate receptor by accelerating its degradation. J. Biol. Chem. 269(11): 7963-7969. WUDARCZYK, J ., DEBSKA, G. and LENARTOWICZ, E. (1999). Zinc as an inducer of the membrane permeability transition in rat liver mitochondria. Arch Biochem. Biophys. 363(1): 1-8. XIA, Y, RAGAN, R. E., SEAH, E. E., MICHAELIS, M. L. and MICHAELIS, E. K. (1995). Developmental expression of N-methyl-D-aspartate (NMDA)-induced neurotoxicity, NMDA receptor function, and the NMDARI and glutarnate- binding protein subtmits in cerebellar granule cells in primary cultures. Neurochem. Res. 20(5): 617-629. XU, J. and CHUANG, D. M. (1987). Muscarinic acetylcholine receptor-mediated phosphoinositide turnover in cultured cerebellar granule cells: desensitization by receptor agonists. J. Pharmacol. Exp. Ther. 242(1): 238-244. YAN, G. M, LIN, S. Z, IRWIN, R. P. and PAUL, S. M. (1995). Activation of G proteins bidirectionally affects apoptosis of cultured cerebellar granule neurons. J Neurochem. 65(6): 2425-2431. YAN, G. M, LIN, S. Z, IRWIN, R. P. and PAUL, S. M. (1995). Activation of muscarinic cholinergic receptors blocks apoptosis of cultured cerebellar granule neurons. Mol. Pharmacol. 47(2): 248-257. YEE, S. and CHOI, B. H. (1994). Methylmercury poisoning induces oxidative stress in the mouse brain. Exp. Mol. Pathol. 60(3): 188-196. YEE, S. and CHOI, B. H. (1996). Oxidative stress in neurotoxic effects of methylmercury poisoning. NeuroToxicology 17(1): 17-26. YOSHINO, Y, MOZAI, T. and NAKAO, K. (1966). Distribution of mercury in the brain and its subcellular units in experimental organic mercury poisoning. J. Neurochem. 13: 397-406. 204 YUAN, Y. and ATCHISON, W. D. (1993). Disruption by methylmercury of membrane excitability and synaptic transmission of CA1 neurons in hippocampal slices of the rat. Toxicol. Appl. Pharmacol. 120(2): 203-215. YUAN, Y. and ATCHISON, W. D. (1994). Comparative effects of inorganic divalent mercury, methylmercury and phenylrnercury on membrane excitability and synaptic transmission of CA1 neurons in hippocampal slices of the rat. NeuroToxicology 15(2): 403-41 1. YUAN, Y. and ATCHISON, W. D. (1995). Methylmercury acts at multiple sites to block hippocampal synaptic transmission. J. Pharmacol. Exp. Ther. 275(3): 1308-1316. YUAN, Y. and ATCHISON, W. D. (1997). Action of methylmercury on GABA(A) receptor-mediated inhibitory synaptic transmission is primarily responsible for its early stimulatory effects on hippocampal CA1 excitatory synaptic transmission. J. Pharmacol. Exp. Ther. 282(1): 64-73. ZENG, F. Y, SOLDNER, A, SCHONEBERG, T. and WESS, J. (1999). Conserved extracellular cysteine pair in the M3 muscarinic acetylcholine receptor is essential for proper receptor cell surface localization but not for G protein coupling. J. Neurochem. 72(6): 2404-2414. ZHAN G, Z. and WU, D. (1994). [Effect of methylmercury chloride on brain muscarinic receptor in rat]. Hua Xi Yi Ke Da Xue Xue Bao 25(4): 388-392. 205 lullliltiflll‘rfll’llfillll