. 3.: .1 . L535... .3. .337... .23.... w. s \s Hurt? 5.14:9: . .:l I .‘Ic‘zv’o .tov‘ds 13 II. fill-Into).- ‘Ilv . 3. mg "a”... ‘ i. .eéfia It? 5 « .ni e 3: Cat}. 2. .p 2:. It‘iuar1 7.. xIa: .3 1.2.; 13.3, «a... gift . 1'9 :31 29.2..» $2.. n. .. ( i e \ «.3.» In: a? h 11 .y n r .r:. V .1.) .44! , i 53!...” vi 71.... . 3‘ L .4: : . x. - a bun. , . . . a: «so; 14 9.4.th1: n .3 Sonic. . 2.. .. I31 , 34L :0! E .,,9£4;u .. 4.4 .3 :4! Its... )5. churn” Wu... .3. 1 J\ I ‘..o. 9.00% This is to certify that the dissertation entitled COMPARATIVE EFFECTS OF MEHG ON GABAA RECEPTOR FUNCTION IN CELLS IN CULTURE presented by CHRISTINA J. HERDEN has been accepted towards fulfillment of the requirements for the PhD. degree in Neuroscience Major Professor’s Signature 5/30/06 Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University -V-.-.-.-._. 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 c:/ClRC/DateDuelp65-p, 15 — COMPARATIVE EFFECTS OF MEHG ON GABAA RECEPTOR FUNCTION IN CELLS IN CULTURE By Christina J. Herden A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Neuroscience Program 2006 ABSTRACT COMPARATIVE EFFECTS OF MEHG ON GABAA RECEPTOR FUNCTION IN CELLS IN CULTURE By Christina Jean Herden MeHg is an environmental contaminant that still poses a significant toxicological threat. MeHg is highly neurotoxic as it readily crosses the blood brain barrier due to its lipophilic nature. In the brain, the cerebellum is one of the primary targets for MeHg. Interestingly, GABAergic synaptic transmission in the cerebellum is especially sensitive to the effects of MeHg as compared with excitatory synaptic transmission. The effects of MeHg on GABAA receptor-mediated responses are poorly understood. To characterize the effects of MeHg on IGABA in cells in culture, whole-cell IGABA were recorded in the presence of MeHg (0.1, l, and 10 M) in cerebral cortical cells in primary culture. IGABA were gradually and irreversibly suppressed by MeHg in a time- and concentration- dependent manner in cortical cells. Kinetically, MeHg prolonged [GABA slow decay rate in cortical cells in a concentration-independent manner. To determine whether or not MeHg interacts with the channel pore, IGABA were recorded at a range of voltages. The effects of MeHg on [GABA were voltage-independent, suggesting that MeHg binds to the external surface of the receptor. To determine if MeHg interacts with the GABA or diazepam binding site of the receptor, effects of MeHg on IGABA suppression and kinetics were examined in the absence or presence of GABA (10, 100, 500, 1000 M), or diazepam (0.1, 1, and 10 uM) or flumazenil (10 uM), respectively. The effects of MeHg on IGABA were not altered in the presence of GABA, diazepam, or flumazenil, indicating that MeHg does not likely interact with the GABA or diazepam site of the receptor. Cerebellar granule cells are particularly sensitive to inhibition of GABAergic synaptic transmission by MeHg as compared with neighboring Purkinje neurons. One possibility for these differential effects may be the differential expression of the a subunit, specifically the (16 subunit, in GABAA receptors in granule cells and Purkinje cells. The presence of the (16 subunit confers unique pharmacological and kinetic properties onto the GABAA receptor. To investigate this possibility, whole-cell IGABA were recorded in the presence of MeHg (O. 1, 1, and 10 M) in cerebellar granule cells, cerebral cortical cells, and in transfected human embryonic kidney (HEK-293) cells in culture. Cortical cells were used as a replacement for Purkinje cells as they are more successfully cultured and express the same ail-containing GABAA receptor phenotype as do Purkinje cells. To verify the presence of specific GABAA receptor subtypes of interest, immunocytochemistry was performed on each cell type. Findings from these studies reveal that IGABA were suppressed more rapidly by MeHg in granule cells than in cortical cells. However, IGABA recorded from (16 or Oil-containing granule cells did not differ in their responsiveness to MeHg. Similarly, the effects of MeHg did not differ in IGABA recorded from HEK-293 expressing either (16 or ail-containing receptors. These findings suggest that the differential effects of MeHg on IGABA in cerebellar cells are not likely due simply to the expression of the a6-containing GABAA receptor. DEDICATION For GABA Girl iv ACKNOWLEDGEMENTS I would like to thank my parents for their love and support, and my mother for her clerical assistance. I am also greatful to my three sisters, Corinna, Samantha, and Jamie, for their love and friendship. I would like to thank Dr. Atchison for his encouragement and support. I would like to thank Dr. Yukun Yuan for his mentorship, patience, and big heart. Thanks, also to Dr. Ravindra Hajela for his advice and expertise and help with cell transfections. I would also like to express my thanks to my guidance committee for their time, input and suggestions. I greatly appreciate the friendship, empathy and support of my dear friends James Otero and Nicole Pardo. Thanks also to Nicole Pardo and Dawn Parsell for their assistance with immunocytochemistry. TABLE OF CONTENTS LIST OF TABLES ................................................................................. viii LIST OF FIGURES .................................................................................. ix LIST OF ABBREVIATIONS ...................................................................... xi CHAPTER ONE: INTRODUCTION .............................................................. 1 A. Methylmercury .......................................................................... 2 B. GABAA receptor a. Structure .......................................................................... 8 b. Function........ ..................................................................................... 11 c. Pharmacology .................................................................. 13 C. GABAA receptors in the cerebellum ................................................. 21 D. MeHg neurotoxicity a. Cellular effects of MeHg ...................................................... 30 b. Disruption of synaptic transmission by Mehg ............................. 32 E. Disruption of GABAergic function by MeHg ....................................... 34 CHAPTER TWO: EFFECTS OF METHYLMERCURY ON DIAZEPAM-MEDIATED ENHANCEMENT OF GABAA RECEPTOR CURRENTS IN RAT CORTICAL NEURONS IN CULTURE A. Abstract .................................................................................. 42 B. Introduction ............................................................................. 43 C. Methods ................................................................................. 46 D. Results ................................................................................... 51 E. Discussion ............................................................................... 79 vi CHAPTER THREE: DIFFERENTIAL EFFECTS OF MEHG ON GABAA RECPTOR CURRENTS IN CEREBELLAR GRAN ULE AND CEREBRAL CORTICAL CELLS IN CUTLURE A. Abstract .................................................................................. 85 B. Introduction ............................................................................. 86 C. Methods ................................................................................. 88 D. Results ................................................................................... 94 E. Discussion .............................................................................. 133 CHAPTER FOUR: SUMMARY AND DISCUSSION A. Summary of research. . . . . . .. ........................................................................... 139 B. Relationship to previous work ...................................................... 143 C. Possible mechanisms underlying the differential effects of MeHg on cerebellar granule and Purkinje neurons ............................................................ 147 D. Potential mechanisms of interaction between MeHg and the GABAA Receptor .................................................................................... 158 E. Conclusion ................................................................................ 162 APPENDIX A. Suppression of [GABA by MeHg in granule cells is not reversible ............... 165 B. MeHg does not compete with GABA for its binding site on the GABAA receptor ..................................................................................... 168 BIBLIOGRAPHY ................................................................................. 17 1 vii LIST OF TABLES CHAPTER ONE 1.1. Subunit specific actions of GABAA receptor agonists and antagonists .................................................................................... 16 1.2. GABAA receptor subunit subtypes expressed by different cells in the cerebellum .............................................................................................................. 27 CHAPTER THREE 3.1. Comparative effects of MeHg on [GABA suppression in on- or GIG—containing HEK- 293 cells .................................................................................... 130 viii LIST OF FIGURES CHAPTER ONE 1.1. A schematic representation of the GABAA receptor and its subunits ............... 10 1.2. Diagram depicting the general circuitry of the cerebellum ........................... 23 CHAPTER TWO 2.1. MeHg causes a gradual suppression of bicuculline-sensitive IGABA in cortical cells .......................................................................................... 53 2.2. MeHg causes a complete and irreversible suppression of IGABA in cortical cells .......................................................................................... 55 2.3. Suppression of IGABA by MeHg in cortical cells is time- and concentration- dependent .................................................................................... 57 2.4. MeHg prolongs [GABA slow decay constant in a concentration-independent manner. ...................................................................................... 60 2.5. MeHg affects the time course of prolongation of IGABA slow decay in a concentration—dependent manner ........................................................ 62 2.6. Diazepam prolongs IGABA slow decay constant in a concentration-dependent manner ........................................................................................ 65 2.7. Co-application of MeHg and diazepam does not affect MeHg-induced prolongation of [GABA slow decay ......................................................... 68 2.8. Time course of IGABA suppression by MeHg in cortical cells is not altered by a pplication of diazepam or flumazenil .................................................... 70 2.9. Diazepam or MeHg pretreatment, following co-application, does not prevent diazepam- or MeHg-induced prolongation of IGABA slow decay .................... 73 2.10. Co-application of MeHg and flumazenil does not prevent MeHg-induced prolongation of IGABA slow decay ....................................................... 76 CHAPTER THREE 3.1. GABAA receptor (16 and a1 subunit expression in cerebellar granule and cerebral cortical cells in culture .................................................................... 96 ix 3.2. MeHg causes a time- and concentration-dependent suppression of IGABA in granule cells ................................................................................. 98 3.3. MeHg causes a time- and concentration-dependent suppression of [GABA in cortical cells ............................................................................... 101 3.4. Effects of MeHg on IGABA are voltage—independent in granule and cortical cells. .............................................................................................. 105 3.5. GABAA receptor (16 and (11 subunit expression in cerebellar granule cells at different days in culture .................................................................. 107 3.6. Effects of MeHg on granule cells at different days in culture ...................... 110 3.7. GABAA receptor (16 and a1 subunit expression in HEK-293 cells .................. 115 3.8. GABAA receptor membrane localization in (11 subunit-containing HEK-293 cells ......................................................................................... 117 3.9. GABAA receptor membrane localization in (16 subunit-containing HEK-293 cells ......................................................................................... 119 3.10. Control Secondary antibody staining in HEK-293 cells ............................ 121 3.11. Effects of diazepam on IGABA in HEK-293 cells ..................................... 123 3.12. MeHg causes a time- and concentration-dependent suppression of IGABA in HEK- 293 cells .................................................................................... 127 3.13. Effects of MeHg on IGABA are voltage-independent in HEK-293 cells ............ 132 CHAPTER FOUR 4.1. Drawing showing a cerebellar glomerulus ............................................ 150 4.2. Diagram depicting the proposed mechanism of interaction of MeHg with GABAA receptors of granule cells to alter cerebellar excitation .............................. 153 APPENDIX , A.l. Effects of MeHg on IGABA are irreversible in granule cells in culture ............. 167 A.2. MeHg does not interact with the GABA site on the GABAA receptor ............ 170 BIBLIOGRAPHY ............................................................................................................ 171 AMPA APV Ara-C ATP CNQX CNS DMEM DMSO DNase I D-pen EGTA EPSC EPSP FITC GABA GFP hr HEK-293 HEPES IGABA LIST OF ABBREVIATIONS a-amino—3-hydroxy-5-methylisoxazole-4— propionic acid DL-2-amino-5-phosphonopentanoic acid cytocine B-D-arabino-furanoside adenosine 5’- triphosphate magnesium salt 6-cyano-7-nitroquinoxaline-2,3-dione disodium salt central nervous system Dulbecco’s Modified Eagle’s Medium dimethyl sulphoxide deoxyribonuclease I D-penicillamine ethylene glycol-bis(B-aminoethyl ether)-N, N, N', N',-tetraacetic acid excitatory postsynaptic current excitatory postsynaptic potential fluorescein isothiocyanate gram y-amino-n-butyric acid green fluorescence protein hour human embryonic kidney cells 4-(2-hydroxyethyl)piperazine- 1 -ethanesulfonic acid GABAA receptor-mediated current xi 1P3 IPSC IPSP MeHg mV N MDA PBS TRITC inositol triphosphate inhibitory postsynaptic current inhibitory postsynaptic potential methylmercury minutes millivolts n-mehthyl-d-aspartate picoamperes phosphate buffered saline seconds tetrarnethylrhodamine xii CHAPTER ONE INTRODUCTION I A. Methylmercury Methylmercury (MeHg) is a ubiquitous environmental contaminant that still poses a significant toxicological threat, particularly in the Great Lakes, where levels of MeHg in fish have been found to exceed the estimated dose for adverse health effects in humans (Rice, 1995; Gerstenberger and Dellinger 2002; Gilbertson, 2004; Weis, 2004). MeHg toxicity is not limited to the United States; it occurs in populations with high fish intake all around the world including, Canada (Mckeown-Eyssen et al., 1983a,b; Forsyth et al., 2004), the Faroe Islands (Steurwald et al., 2000), the Seychelle Islands (Davidson et al., 1995, 1998; Shamlaye et al., 1995; Axtell et al., 2000), Peru (Marsh et al., 1995a, 1995b), Brazil (Kehrig et al., 1998; Grandjean et al., 1999), New Zealand (Crump et al., 1998), Madeira (Murata et al., 1999b), Greenland (Weihe et al., 2002), French Guiana (Cordier et al., 2002), France (Claisse et al., 2001), and others. In fact, MeHg was listed by the International Register for Potentially Toxic Chemicals, part of the World Health Organization (WHO), as one of the six most dangerous chemicals in the world’s environment (WHO, 1993). Environmental contamination by MeHg occurs either through MeHg itself or by conversion of elemental Hg into MeHg. Industries such as agriculture, paper, lumber, and leather have used MeHg as a preservative because of its antifungal properties. As a result, contamination has been found in agricultural runoff and waste water discharge from these industries. Additionally, elemental Hg, which is used by industries that manufacture electrical equipment, paint, and chloralkali, is vented directly into the atmosphere or is discharged in waste-water. Furthermore, Hg is released into the environment by erosion of rock containing Hg or by human activities such as combustion of fossil fuels. Importantly, aquatic microorganisms in lakes, rivers, and oceans that receive Hg waste can methylate inorganic Hg, thus converting it to MeHg (Wood et al., 1968). MeHg tends to bioaccumulate in the aquatic food chain, resulting in high concentrations in fish at the top of the food chain, such as tuna and shark (Toxicological Effects of Methylmercury, 2000). According to the Environmental Protection Agency (EPA), most fish that live in US waters have MeHg concentrations of less than 0.5 parts per million (ppm), but fish at the top of the food chain may contain more than 1 ppm (the average concentration in fish is approximately 0.25 ppm). The most common route of human exposure to MeHg is through consumption of fish and other seafood taken from contaminated waters. Frequent consumption of MeHg-contaminated food and/or water can result in significant accumulation of MeHg in the body because MeHg has a long half-life in humans (approximately 70 days) and a low rate of urinary excretion (Magos, 1975). Once consumed, MeHg is transported to most tissues in the body in plasma and red blood cells and is concentrated in the blood and the central nervous system (Chang et al., 1972). The current reference dose for MeHg has been set by the EPA at 0.1 pg/ kg/ day. Taking into consideration the average MeHg concentration of fish (0.25 ppm) (EPA, 1997), the reference dose of MeHg translates into approximately 27 g of fish per day, which is equivalent to only a few meals of fish per week. The first description of the adverse effects of MeHg demonstrated that it produced sensory and motor disturbances such as impairment of vision, hearing, numbness, and ataxia, and eventually death (Hunter et al., 1940). Histological examination of the deceased victim revealed that the brain was a primary target site for MeHg, particularly the cerebellum (Hunter and Russell, 1945). Within the cerebellum, granule cells were shown to be especially sensitive to MeHg-induced atrophy as compared to other surrounding cell types. Several years later, mass episodes of acute and chronic MeHg poisoning occurred. In the 1950’s, chronic exposure to MeHg in Japan resulted from the contamination of Minamata and Nigata Bays by local industrial mercurial waste. Over a period of years, local inhabitants ingested large amounts of fish and shell-fish taken from the Bay. As a result, at least 1500 poisonings and 46 deaths occurred in individuals of all ages (Takeuchi et al., 1962, 1968, 1970). In contaminated areas, levels of MeHg reached 50 — 85 ppm in fish and other seafoods. Over 2,000 inhabitants of the Minamata Bay area that had been exposed to MeHg were found to have neurological defects. Initially, exposure victims demonstrated slight motor disturbances such as abnormalities in gait and weakness of lower limbs, resulting in uncoordinated and unstable gait (Takeuchi et al., 1968, 1989; Takeuchi and Eto, 1999). With longer exposure time to MeHg, however, victims experienced progressive worsening of symptoms, including visual disturbances, loss of hearing, and ataxia. The gradual worsening of ataxic symptoms was associated with progressive insult to cerebellar-based motor pathways, and the onset of ataxia was closely correlated with granule cell damage in the cerebellum (Takeuchi and Eto, 1999). In 1972, a mass episode of acute MeHg contamination occurred in Iraq. Homemade bread was made from seed grain that had been contaminated with MeHg- based fungicide and eaten by inhabitants of the community. As a result, 450 individuals lost their lives and over 6500 victims were hospitalized (Bakir et al., 1973). Blood Hg concentrations in victims in 17 patients ranged from 3-4 mg/ml. At these concentrations of MeHg signs and symptoms of ataxia were observed and visual and hearing disturbances were reported. In the majority of pathological investigations of victims intoxicated with MeHg in these episodes, substantial atrophy of the cerebellum was observed, including extensive thinning of the gray matter. In the cerebellar cortex, a severe disintegration of granule cells was detected, but surprisingly, the neighboring Purkinje cells were relatively resistant to MeHg-induced cell atrophy. Poisoning with MeHg compounds can occur with even very limited acute effects. At Dartmouth University in 1997, a chemistry professor who researched the toxicity of heavy metals, accidentally spilled several drops of dimethylmercury onto one of her latex glove- covered hands (N ierenberg et al., 1998). Within days after exposure, she began to experience signs such as weight loss and diarrhea, and weeks after exposure, she experienced neurologic symptoms such as difficulty with speech and abnormal hearing, vision, and gait. Ultimately, three months after exposure, she entered into a vegetative state, followed by death. At autopsy, the cerebellum showed diffuse atrophy, including a widespread loss of granule cells, Purkinje cells, and parallel fibers in the molecular layer. It is noteworthy to mention that the developing nervous system is particularly sensitive to the effects of MeHg (Bakir et al., 1973; Grandjean etal., 1997, 1998), and that there are considerable differences in the distribution of MeHg-induced pathological changes in young brains as compared with adult brains (Takeuchi et al., 1978). In adults, the brain damage induced by MeHg is specific to loss of neurons in the visual cortex and granule cells of the cerebellum. Purkinje cells of the cerebellum are unaffected. However, in the developing brain, a more generalized and extensive distribution pattern of injury is observed, such as loss of neurons in each lobe of the brain (Choi, 1989). This is due to the fact that MeHg affects many developmental processes including formation of microtubules (Miura and Imura, 1989; Muira et al., 1999), which leads to disruption of neuronal migration, cell division, and synaptogenesis (Castoldi et al., 2001, Clarkson, 1993). In addition, MeHg can disrupt protein, DNA, and RNA synthesis and neurotransmitter production, secretion and uptake, as well as cell-signaling (Castoldi et al., 2001). Disruption of any of these processes can lead to disruption of cerebellar organization (Choi et al., 1978). Many differences exist between cerebellar granule and Purkinje neurons. Thus there are numerous possible explanations for the differences in neurotoxicity to MeHg observed in these cells. First, there is the disparity in cell soma size. Purkinje cells of the cerebellum are much larger (~ 25-30 pm) than granule cells (~ 5-8 pm), thus granule cells may simply accumulate a greater concentration of MeHg. However, Purkinje cells have been shown to accumulate as much or more MeHg as compared with granule neurons (Leyshon-Sorland and Morgan, 1994). Another difference observed between cerebellar granule cells is their differential expression of potential protective MeHg-sequestering mechanisms. Purkinje cells, but not granule cells, were shown to synthesize metallothionein, a protein that is able to sequester mercury (Leyshon-Sorland et al., 1994). In addition, Purkinje cells have a greater ability to buffer Ca”, as they contain calbindin-D29K, a Ca+2 binding protein, whereas granule cells do not (Celio, 1990). Granule and Purkinje cells also show a differential expression of certain types of ion channels including voltage-gated Ca+2 channels (Mintz et al., 1992; Randall and Tsien, 1995), N-methyl-D-aspartate (NMDA) receptors (Didier etal., 1997), AMPA/kainate receptors (Keinanen et al., 1990), and GABAA receptors (Takayama and Inoue, 2004). Most pertinent to the current investigation is the differential expression of GABAA receptor subtype in cerebellar granule and Purkinje neurons. Cerebellar Purkinje cells express an subunit—containing GABAA receptors, whereas granule cells express the (16- containing receptors in addition to a. subunit-containing ones (Wisden et al., 1996; Poltl et al., 2003). In the cerebellar slice preparation, GABAergic inhibitory synaptic transmission appears particularly sensitive to the inhibitory effects of MeHg, especially in granule cells as compared with Purkinje cells (Yuan and Atchison, 2003). Also, compared to other targets such as voltage-gated Ca+2 and K+ channels and inwardly rectifying K+ channels, GABAA receptor currents are reduced most rapidly by MeHg (Yuan et al., 2005). The exact mechanisms by which MeHg differentially interferes with GABAergic function in granule cells and Purkinje neurons are not known. One possibility for this differential sensitivity to MeHg may be the fact that granule cells express a unique type of GABAA receptor that contains the (16 subunit. This receptor subtype is especially sensitive to inhibition by heavy metals such as La+3 and Zn”, and studies have demonstrated that MeHg enhances levels of intracellular Ca+2 and Zn+2 at the concentrations used to inhibit GABAergic function (Sarafian, 1993; Denny et al., 1993; Hare et al., 1993; Hare and Atchison, 1995). Interestingly, no studies comparing the differential effects of MeHg on subunit subtype-specific populations of GABAA receptors found in the cerebellum have been conducted. This thesis was undertaken to enhance further our understanding of the differential effects of MeHg on the function of different a-subunit containing GABAA receptors found in either granule or Purkinje cells of the cerebellum. Specifically, it was performed to 1) characterize the effects of MeHg on GABAA receptor-mediated function 2) to determine if MeHg differentially affects the "1 F functional properties of GABAA receptors in (16 subunit-containing and non-a6 subunit- containing cells. . GABAA Receptor a. Structure GABAA receptors are members of the Cys-loop family of receptors which include the nicotinic acetylcholine, glycine, and serotonin 5-HT3 receptors (Karlin and Akabas, 1995, Le Novere et al., 2002). All subunits in this receptor family have a disulfide-linked loop in the extracellular domain consisting of 15 residues. GABAA receptors are ligand- gated Cl' ion channels that mediate the majority of fast inhibitory synaptic transmission in the mammalian central nervous system. Approximately 20-50% of all neuronal synapses use GABA as a neurotransmitter (Bloom and Iversen, 1971). GABAA receptors are 275 kDa heteropentameric glycoproteins composed of different subunit families and their subtypes (cl-(16, [Bl-B3, 71-73, 5, e, p, 1:) (Olsen and Tobin, 1990, Wisden and Seeburg, 1992; Sieghart, 1995; Hevers and Luddens, 1998) (Fig. 1.1). The exact stoichiometry of GABAA receptors in their native environment is unknown, but based on studies in recombinant receptors, the GABAA receptor stoichiometry appears to consist of two a, two [3, and one y subunit. One phenotype of GABAA receptor predominates in most regions of the brain (on, 02/3, 72) (Chang et al., 1996). Each subunit has a similar overall topology consisting of four transmembrane domains (M1, M2, M3, and M4), which are responsible for forming the ion-conducting pore (Schofield et al., 1987; Karlin and Akabas, 1995). Furthermore, each subunit has a large extracellular N-terminus and a smaller C-terminus, a small loop between domain M2 and M3, and an intracellular loop Figure 1.1. A schematic representation of the GABAA receptor and its subunits. The (11, [32/3, and y; phenotype of receptor predominates in most regions of the brain. The binding site for GABA is located at the interface between the o. and 0 subunits. Each subunit is comprised of four transmembrane domains (M l-M4), a large extracellular N-terminus and a smaller C-terminus. Intracellular Figure 1.1 10 between M3 and M4 (Schofield et al., 1987; Karlin and Akabas, 1995). The extracellular N-terminus is believed to form the agonist binding sites and the loop between M3 and M4 has been shown to be a site for regulation by phosphorylation and for localization at synapses (Brandon et al., 2002; Moss and Smart, 1996). The loop between domain M2 and M3 has been implicated in the signal transduction process, specifically in coupling conformational changes between the two domains and regulation of channel gating (Grosman et at., 2000; Beta et al., 2002). b. Function GABA is released from presynaptic terminals into the synaptic cleft and reaches concentrations between 500 - 1000 M in less than 1 ms (Jones and Westbrook, 1995). GABA concentrations in the cleft decay rapidly (within ms) as a result of diffusion and reuptake by the presynaptic terminal. Postsynaptically, GABA binds to its receptor at the interface between the extracellular B and C loops of the [3 subunit and the D loop of the a subunit (Galzi at al., 1994). Two GABA molecules must bind to trigger a rotation of the extracellular domains of these receptor subunits, resulting in opening of the channel pore (Sakmann et al., 1983). However, GABA binding can also result in the entrance of the receptor into a desensitized state, during which two GABA molecules are bound but the channel is closed (MacDonald et al., 1989a; Twyman et al., 1990; Twyman and MacDonald, 1992; Jones and Westbrook; 1995). From the desensitized state, unbinding of the ligand from the receptor, or deactivation, is slowed because it must backtrack through each of the preceding states. The number and order of occurrence of open, closed, and desensitized states of the GABAA receptor channel can vary depending upon 11 which binding model is considered. One model of single channel GABAA receptor activity at the main conductance level, for instance, incorporates a reaction scheme that involves two sequential binding sites, three open states, ten closed states, and one desensitized state (MacDonald et al., 1989a; Twyman et al., 1990; Twyman and MacDonald, 1992). GABAA receptor chloride channels open to at least three conductance states (27- 30 pS, 17-19 pS, and 11-12 pS) (MacDonald et al., 1989a). However, most (95%) of the Cl' current though the channel flows through the highest conductance state (27-30 pS). At the highest conductance state, two types of GABAA receptor channel openings occur: brief openings (~ 2.5 ms) and bursts of openings (longer openings interrupted by brief periods of closure). Bursts of openings are thought to represent fluctuations of the receptor between the double-1i ganded open and double-liganded closed state. Silent periods between single brief openings and bursts are believed to correspond to a double- liganded closed, or desensitized, state of the receptor (Hammond, 2001). Desensitization of the GABAA receptor increases with increased GABA concentrations and is voltage- dependent, increasing its rate at more positive potentials (Mellor and Randall, 1998). The typical GABAA receptor-mediated whole-cell response is characterized by rapid activation followed by a fast and a slow decay component (Hamill et al., 1983; MacDonald et al., 1989a). The fast decay component corresponds to the desensitization of receptors during prolonged GABA exposure, resulting in a decrease in the frequency of channel opening. The later slow decay phase corresponds to the deactivation, or GABA unbinding, of receptors. Functional properties of GABAA receptor binding, desensitization, and deactivation are dependent upon subunit subtype composition of the 12 receptor, in addition to other factors such as protein phosphorylation, voltage, and ligand concentration (MacDonald and Olson, 1994). Both the rate and extent of GABAA receptor desensitization are dependent on subunit composition. Studies using transfected cells expressing different recombinant GABAA receptors found that receptors containing the (16 subunit (along with 0272 subunits) deactivated more slowly as compared to those receptors containing the at subunit (Tia et al., 1996). Furthermore, prolonged GABA application produced desensitization of (110272 receptors that was characterized by a fast exponential decay and a slow decay component. In contrast, GABAA receptors expressing the (:6 subunit did not readily desensitize. Cells cotransfected with on and (:6 subunits (along with [5272 subunits) showed unique deactivation and desensitization kinetics, distinct from those of receptors containing only one subtype of a. subunit (Tia et al., 1996). Furthermore, studies using GABAA receptor a] subunit deficient mice revealed the presence of slower current decay rates in these "(11 knockout" animals, suggesting that the a; subunit is responsible for fast decay rates of inhibitory synaptic currents (Vicini et al., 2001). Thus, it appears that the GABAA receptor a subunit plays a critical role in receptor desensitization and deactivation. c. Pharmacology GABAA receptors have drug-binding sites for GABA as well as agonists such as benzodiazepines, barbiturates, and general anesthetics (Olsen, 1981, 1987; Olsen et al., 1986). GABAergic function can also be modulated by neurosteroids (Callachan et al., 1987; Puia et al., 1990, 1993, 2003; Zhu et al., 1996) and La+3 (Ma and Narahashi, 13 1993a) and antagonized by agents such as picrotoxin, bicuculline, furosemide, penicillin, and Zn+2 (MacDonald and Barker, 1978; Westbrook and Mayer, 1987; Smart and Constanti, 1990; Twyman et al., 1992; Korpi and Luddens, 1993; Barberis et al., 2000; Kolbaev et al., 2002). The affinity of GABAA receptors for these agents is highly dependent upon the subunit composition of the receptor (see Table 1.1). As numerous drugs act at the GABAAreceptor, only those compounds that are relevant to the present thesis will be discussed. GABA binds to the GABAA receptor to regulate the opening or closing, also known as gating, of the chloride ion channel (Sakmann et al., 1983). Functional GABAA receptors are formed following expression of only the a and [3 subunits (Baumann et al., 2001). The nature of the a subunit appears to affect the affinity of the receptor to GABA. Receptors expressing a6 subunits have a much higher affinity for GABA than do receptors containing (1; subunits (Fisher et al., 1997). Benzodiazepines, such as diazepam, are compounds that produce a range of sedative, anti-convulsant, and/or anxiolytic actions (Bellantuono et al., 1980). Benzodiazepines potentiate GABAergic responses at submaximal concentrations of GABA (Study and Barker, 1981). Single-channel analyses of the actions of benzodiazepines have revealed that they enhance GABAergic responses by increasing open and burst frequencies without altering channel conductance or open or burst duration (Study and Barker, 1981). Enhanced open frequency results from an increase in the probability of Cl' channel opening in response to GABA by increasing the binding rate of the GABA molecule to the receptor (Costa and Guidotti, 1979; Study and Barker, 1981; Vicini et al., 1987; Sieghart, 1992). Furthermore, a predominant effect of 14 Table 1.1. Subunit specific actions of GABAA receptor agonists and antagonists. Excitatory actions of agents on specific receptor subunits are indicated by a plus (+) symbol. Enhanced sensitivity to excitation is indicated by a double plus (++) symbol. Inhibitory responses are indicated by a minus (-) symbol. Enhanced sensitivity to inhibition is indicated by a double minus (--) symbol. Insensitivity of the subunit to the agent is depicted by a (0) symbol. 15 Drug or Drug GABAA Receptor Subunit Group 0‘1 Cl<5 7 + 0 Benzodiazepines + ++ + 0 Barbiturates + ++ + 0 Furosemide - + + 0 Lanthanum Zinc Table 1.1 16 benzodiazepines on GABAergic responses is a prolongation of the time course by slowing of decay (Segal and Barker, 1984; Vicini et al., 1986; Zhang et al., 1993). The slowing of decay by benzodiazepines is believed to be due to slowing of current deactivation and speeding up of current desensitization (Mellor and Randall, 1997). The actions of benzodiazepines at the GABAA receptor are primarily dependent on a and 7 subunit combinations (Luddens and Wisden, 1991). The presence of the 72 subunit confers benzodiazepine sensitivity onto the receptor, and the a subunit, when expressed with the B and 72 subunits, affects the efficacy and affinity of the GABAA receptor for benzodiazepines (Pritchett et al., 1989). For example, GABAA receptors containing the 0.6 subunit bind the inverse-agonist ligand [3H]Ro 15-4513, but do not bind benzodiazepines such as diazepam (W ieland et al., 1992). In contrast, the function of GABAA receptors containing the on subunit is potentiated by benzodiazepines (Makela et al., 1997). Barbiturates, such as pentobarbital, are usually anesthetic compounds. Their effects on GABAergic responses vary depending on the concentration at which they are applied. At low concentrations, barbiturates enhance GABA-mediated chloride conductance by increasing the mean channel open time (Nicoll and Wojtowicz, 1980; Schulz and MacDonald, 1981). At higher concentrations, barbiturates open the GABAA receptor directly, independent of the presence of GABA (Nicoll and Wojtowicz, 1980; Schulz and MacDonald, 1981; Thompson et al., 1996). At very high (millimolar) concentrations, they inhibit GABAA receptor function, most likely by blocking the open channel (Akk and Steinbach, 2000). Barbiturate sensitivity is dependent on the GABAA receptor [3 subunit. Receptor mutation studies have shown that [33 subunit-containing 17 GABAA receptors are insensitive to pentobarbital, whereas [31 subunit—containing receptors are sensitive (Cestari et al., 2000). Moreover, the a subunit affects the sensitivity of the GABAA receptor to barbiturates. Receptors containing the (16 subunit are about three times more responsive to pentobarbital activation than are non (1‘ subunit- containing receptors (Fisher et al., 1997). Steroids, particularly neurosteroids such as pregnanalone, enhance the actions of GABA at the GABAA receptor by prolonging the open time and burst duration of the chloride channel and by enhancing the binding affinity of receptor agonists such as muscimol or benzodiazepines (Barker et al., 1987; Callachan et al., 1997a). The steroids androsterone and pregnanolone have been shown to increase channel open and burst durations (Twyman and MacDonald, 1992). The prolonged open duration is due to an increase in open frequency and an increase in frequency of occurrence of longer open states. Increased burst duration is concentration-dependent and results in an increase in proportion of bursts with longer durations. The mechanism for prolongation of open and burst durations of IGABA is similar to that described for barbiturates, suggesting that both may regulate the GABAA receptor channel through at least one common mechanism (MacDonald et al., 1989b). There is no absolute subunit specificity for neurosteroid modulation of GABAA receptors, however the actions of neurosteroids seem to be particularly evident in receptors containing a, y, or 6 subunits (Puia et al., 1990, 1993; Korpi and Luddens, 1993; Zhu et al., 1996). GABAA receptors that contain the 6 subunit, for instance, have been shown to have reduced sensitivity to neurosteroids (Zhu et al., 1996; Vicini et al., 2002) 18 La”, a rare earth metal, potentiates GABAA receptor-mediated currents (Ma and Narahashi, 1993a). It does so by increasing the sensitivity of the receptor for GABA (Narahashi et al., 1994) and by increasing the channel open duration (Zhu et al., 1998). The exact molecular mechanism(s) by which La"3 affects GABAA receptor-mediated currents in not known. The modulation of GABAergic responses by La+3 is dependent greatly on the presence of the a subunit. La“3 potentiates a103yzL receptor currents (Saxena et al., 1997), whereas 06B3Y2L currents are only weakly potentiated. In addition, (161336 receptor currents are strongly inhibited by La+3 (Saxena et al., 1997). Furthermore, in autoradiographic binding assays of a6 subunit knockout granule cells, GABA- antagonistic stimulation of the binding by La+3 was abolished (Makela et al., 1997). Furosemide is a loop diuretic which acts to reduce the amount of water in the body by acting on the kidneys to increase the flow of urine. Furosemide non- competitively antagonizes GABA currents (Makela et al., 1999). Competing evidence exists for location of the binding site for furosemide on the GABAA receptor. Kolbaev et al., (2002) demonstrated that furosemide interacted with a site thought to be inside the channel pore of the receptor, resulting in an open-channel block. However, other evidence suggests that the main structural domain for furosemide antagonism resides on a specific 0.6 subunit domain (J ackel etal., 1998). Interestingly, the effects of furosemide are unique to GABAA receptors containing (16 or (:4 subunits, in combination with [32/3 and 72 subunits (Korpi et al., 1995). Furosemide rapidly and reversibly antagonizes GABAergic responses of recombinant 05B2/3Y2 receptors expressed in Xenopus oocytes, however GABAA receptors containing (110272 subunits are insensitive to the inhibitory effects of furosemide (Korpi et al., 1995). 19 Zn“2 slows the onset of GABAA responses by decreasing the rate of GABA binding and the rates of gating transitions from open to closed states (Barberis et al., 2000). Additionally, Zn+2 enhances the rate of deactivation of GABAA receptor currents by increasing the unbinding rate of GABA (Barberis et al., 2000). The effects of Zn+2 are independent of voltage or GABA concentration, indicating that Zn+2 binds to an external site of the receptor distinct from the GABA binding site (Narahashi et al., 1994). Specifically, Nagaya and MacDonald (2001) revealed that the extracellular N-terminus of the 72L subunit is a major determinant of Zn+2 sensitivity by using 72/6 subunit chimeras. Sensitivity of GABAA receptor currents to Zn+2 is highly dependent on subunit composition. The (16036 GABAA receptor currents are the most sensitive to inhibition by Zn”, whereas replacement of the 6 subunit by a 7 subunit reduces the sensitivity to inhibition by Zn+2 (Knoflach et al., 1996; Saxena and MacDonald, 1996; Makela etal., 1997). Furthermore, sensitivity to Zn+2 inhibition is reduced when the 06 subunit of the GABAA receptor is replaced by the al subunit (Draguhn etal., 1990; Saxena and MacDonald, 1994). Along the same lines, studies using granule cells and Purkinje neurons in primary culture revealed that GABAA receptors of Purkinje cells, which contain the on subunit receptor, were significantly less sensitive to the inhibitory effects of Zn”, as compared to granule cells, which have (16 subunit containing receptors (Zempel and Steinbach, 1995). Finally, picrotoxin, a plant-derived convulsant, inhibits GABAA receptor currents by a reduction in the channel's average open and burst durations (Olsen, 1981; Olsen and Tobin, 1990). It does this via binding to a site inside the channel pore, preventing either the Cl' channel from opening, or producing a physical block of the channel. Similarly, 20 penicillin, which can also be a convulsant, inhibits GABAA responses by blocking the open-channel, resulting in a reduction in channel open duration and increase in average burst duration (MacDonald and Barker, 1978; Twyman et al., 1992). In sum, GABAA receptors containing the (16 subunit are particularly sensitive to GABA and pentobarbital activation, insensitive to diazepam potentiation, and especially sensitive to inhibition by Zn+2 and Ln“3 (Draguhn et al., 1990; Luddens et al., 1990; Knoflach et al., 1996; Saxena and MacDonald, 1996; Makela et al., 1997; Zhu et al., 1998; Makela et al., 1999). In contrast, GABAA receptors containing the (:1 subunit are potentiated by benzodiazepines and La”, and are relatively insensitive to inhibition by Zn+2 and furosemide. Additionally, the on subunit containing receptor has a lower efficacy and affinity for pentobarbital (Fisher et al., 1997). C. GABAA Receptors in the Cerebellum Cerebellar cortex consists of three layers (molecular, Purkinje, and granule) that contain only five types of neurons - the granule, Purkinje, Golgi, stellate, and basket cells. The synaptic circuitry between cell types is well-defined (Wisden et al., 1996) (Fig. 1.2). In the granular layer, granule cells receive excitatory glutamatergic input from the mossy fibers, which originate from a variety of brain stem nuclei that receive inputs from the cerebral cortex and spinal cord. Granule cells receive inhibitory GABAergic inputs from Golgi cells. In the molecular layer, granule cell axons, or parallel fibers, form excitatory connections with Purkinje cells and local GABAergic intemeurons (i.e. basket and stellate cells). These intemeurons form inhibitory synapses with Purkinje 21 Figure 1.2. Diagram depicting the general circuitry of the cerebellum. Five types of neurons are organized into three layers in the cerebellar cortex: granule cells (GC), Purkinje cells (PC), Golgi cells (GOL), basket cells (BC), and stellate cells (SC). Mossy fibers (MF), which originate from brainstem nuclei, form excitatory synapses (+) with granule cells and Golgi cells. Granule cells axons, or parallel fibers (PF) form excitatory synapses (+) with Purkinje cells. Golgi cells receive excitatory input (+) from granule cell parallel fibers, and in turn, form inhibitory synapses (-) with granule cells. Climbing fibers (CF), originating from the inferior olivary nucleus of the medulla, form excitatory synapses (+) with Purkinje cells. Stellate and basket cells from inhibitory synapses (-) with Purkinje cells. Purkinje cells form inhibitory synapses (-) with deep nuclei of the cerebellum. 22 CF 1] fl L:_l Deep cerebellar nuclei Inferior olivary nucleus Figure 1.2 23 Spinal cord via brain stem nuclei cells. Purkinje cells also receive excitatory glutamatergic input from the climbing fibers originating from the inferior olive. Purkinje cell axons, which project onto nuclei in the white matter of the cerebellar cortex, serve as the only output of the system. The mossy fiber-granule cell-Purkinje cell pathway through the cerebellum, defined by its high frequency discharge, correlates with and controls behavior (Bastian and Thach, 1995). The climbing fiber-Purkinje cell pathway does not directly control moment-to—moment behavior, as its low frequency discharge increases only when errors in motor performance occur and during adaptation or learning of movement (Bastian and Thach, 1995). Combined electrical stimulation of mossy and climbing fibers depresses parallel fiber Purkinje cell synapses that are concurrently active and spares those that are inactive. The circuitry of the cerebellum is believed to allow the system to receive sensory and motor input and adjust the motor output by comparing intended movements with performed movements, recognizing errors in behavior, and computing corrections (Kandel et al., 2000). Nearly every motor center in the CNS receives excitatory input from nuclei located centrally, in the white matter of the cerebellum. One of these centers is the spinal cord, which receives input from the cerebellum, indirectly, via brain stem nuclei such as the vestibular and reticular nucleus. Other centers include the primary, prefrontal, and premotor cerebral cortices, which receive cerebellar input indirectly from the thalamus (Kandel et al., 2000). Activity in deep cerebellar nuclei correlates with behavior as these nuclei differentially control motor systems and their respective functions. The vestibular and fastigial nuclei control equilibrium, upright stance, and gait; the interpositus controls stretch, contact, placing and other reflexes; and the dentate controls voluntary movements of the extremities, such as those involved in reaching and 24 grasping for objects (Bastian and Thach, 1995). Thus, disruption of the cerebellar circuitry and its output can result in many dysfunctions including abnormalities of limb and eye movements, decrease in muscle tone, and production of incoordinated movement, such as ataxia. A limited number of GABAA receptor isoforms are expressed in the cerebellum (Wisden et al., 1996; Poltl et al., 2003) (see Table 1.2). Studies using a mixture of two subunit-specific antibodies to immunoprecipitate GABAA receptors from the cerebellum In“ of rats have found that 28% of the total cerebellar GABAA receptors were the (110,72 I. isoform, 36% were the (160272 isoform, and 23% were the 0:605 isoform (Quirk etal., 1994). In addition to having strikingly different pharmacological and biophysical properties, receptors containing the (11 and/or 0.6 subunits have markedly different cellular distributions in the cerebellum. Studies using in situ hybridization and immunocytochemistry of rat brain have shown that cerebellar Purkinje cells express GABAA receptors containing (11, [32, [33, and 7 subunits (Wisden et al., 1996). Purkinje cells contain greater levels of B2 mRNA than that of B3. In the adult rat, two splice variants of the 7 subunit exist, 72L (long) and 725 (short), and the mRNA expression levels of 721. are stronger than those for 723- Both 'th and 723 subunit-containing receptors are most likely localized to Purkinje cell dendrites and spines. The predominant GABAA receptor isoform in Purkinje cells is thought to be on, [32, y. In contrast, the composition of GABAA receptors in cerebellar granule cells is more diverse and complex. The major subunit genes expressed in cerebellar granule cells are al, 0.6, B2, [33, 72, and 6 (Laurie et al., 1992). Interestingly, the 0.6 subunit is only expressed in GABAA receptors of cerebellar granule cells and its expression is developmentally regulated both in vivo and 25 Table 1.2. GABAA receptor subunit subtypes expressed by different cells in the cerebellum. The - symbol indicates no expression of the subunit subtype, and the ? symbol indicates the subtype is unknown or not tested. Note that only cerebellar granule cells express (:6 and 6 subunit—containing GABAA receptors. Note also that for certain subtypes expression is limited to the prenatal or the postnatal stage of development. Contributions to this table include studies by Laurie et al., 1992; Pollard etal., 1993; Khan et al., 1994; Thompson et al., 1994; Gao et al., 1995; Wisden et al., 1996; Nusser et al., 1996, 1998; and Poltl et al., 2003. 26 Cell GABAA Receptor Subunit Subtypes Type a I3 y 5 Purkinje 0‘1 [32/ B3 YZL’YZS ' Granule 0.1 , 0‘6 , [32,133,030 YZS’YZL’CY3) 8 (a4) Basket/ a1 ,(a3) I32 729th ‘ Stellate Golgi (12(7), ? Y 1(?) - (13(7) Table 1.2 27 in vitro (Laurie et al., 1992; Thompson et al., 1994; Gao et al., 1995). Studies using in situ hybridization, immunocytochemistry, and radioligand binding have shown that there is a lack of expression of the 0.6 subunit during early cerebellar development and an increased expression of a6 subunits in rats during maturation (Laurie et al., 1992, Thompson et al., 1994). In contrast, GABAA receptors containing the al subunit are expressed throughout cerebellar development and maturation (Laurie et al., 1992, Thompson et al., 1994). There is a great degree of segregation of distinct GABAA receptor subtypes on the surface of cerebellar granule cells (Nusser et al., 1996; Nusser et al., 1998). This segregation suggests different functional roles of the GABAA receptor depending upon subunit composition and subcellular location. Receptors containing the (16, [32/3, and 6 subunits are present exclusively in the extrasynaptic membrane. These (166 subunit- containing receptors have a high affinity for GABA spilled over from Golgi cells (Rossi et al., 1998), and do not desensitize to the prolonged application of agonist (Saxena et al., 1994). It is therefore believed that this GABAA receptor subtype plays a role in mediating tonic inhibition, which originates from the persistent opening of GABAA receptor channels (Brickley et al., 1996; Nusser et al., 1998; Semyanov et al., 2004; Hanchar et al., 2005). Receptors containing the al, (16, (32,3, and 72 subunits (01102572 and (16021372, or (110602372) are primarily present in Golgi cell to granule cell synapses. These receptors are thought to play a role in mediation of phasic inhibition, which originates from synchronous opening of GABAA receptors. Furthermore, granule cell axon terminals that form synapses with Purkinje cells possess 06 subunit containing receptors which are thought to function in increasing glutamate release onto Purkinje cells (Nusser 28 et al., 1998). A subtype of GABAA receptor containing the (16 subunit ((1602372) is also present in some granule cells that form synapses with glutamatergic mossy fibers, suggesting that it may play a functional role in regulation of these excitatory synapses. GABAA receptors containing combinations of a1, a6, and (ll/0.6 also coexist in cerebellar granule cells, suggesting that a variety of different GABAA receptor subtypes coexist in these cells including (110,72, (160,72, (110160.72, a60x6 and a1a60x6 (Pollard et al., 1993; Khan et al., 1994; Nusser et al., 1996, 1998). Almost half (42- 45 %) of all GABAA receptors in the cerebellum contain (16 subunits (Khan et al., 1996; Jones et al., 1997; Jechlinger et al., 1998). The pharmacological properties unique to 06 receptors disappeared in mice lacking the (16 gene; "a6 knockout" animals (u6-/-) were insensitive to the benzodiazepine receptor ligand [3H]Ro 15-4513, insensitive to furosemide and Zn+2 inhibition, and showed a reduction in high affinity binding of the GABA agonist [3H]muscimol (Makela et al., 1997). In these "cu, knockout" mice there was a dramatic loss in the 6 subunit protein, most likely contributing to the reduction of high affinity GABA agonist binding and to the decreased Zn+2 inhibition observed in these animals. Along the same lines, rats containing a point mutation in the (16 gene, also termed alcohol-non-tolerant (ANT) rats, showed diazepam-mediated potentiation of Ion“. This increased affinity for benzodiazepines may contribute to the increased susceptibility to benzodiazepine induced ataxia and impairment of postural reflexes observed in these animals (Korpi et al., 1993). Thus, the (:6 subunit containing GABAA receptors, which are especially sensitive to inhibition by heavy metals such as Zn+2 and La”, may play a critical role in cerebellar- 29 based motor coordination, the disruption of which can result from MeHg exposure (Chang, 1977). D. MeHg Neurotoxicity (1. Cellular eflects of MeHg The mechanisms by which MeHg produces neurotoxicity are not clear. MeHg has a high affinity for sulfhydryl and disulfide groups on cysteines that are numerous in proteins and enzymes. Thus, MeHg has the potential to interfere with many cell processes including disruption of protein synthesis (Syversen, 1981), disruption of neurotransmitter release (Atchison and Narahashi, 1982; Atchison, 1986; Traxinger and Atchison, 1987), and disturbance of divalent cation homeostasis (Ca"2 and Zn”) (Levesque and Atchison, 1991; Hewett and Atchison, 1992; Hare et al., 1993; Denny et al., 1993; Sirois and Atchison, 1996; Marty and Atchison 1997, 1998; Limke and Atchison, 2002; Edwards etal., 2005). This is important, as GABAA receptor-mediated activity can be modulated by numerous divalent cations. For example, Zn+2 inhibits [GABA (Barberis et al., 2000) and Hg+2 potentiates GABAergic currents (Arakawa et al., 1991). Furthermore, Ca+2 can activate intracellular signaling cascades that can alter GABAA receptor function (Moss and Smart, 1996; Brandon et al., 2000). Alteration of Ca“2 concentrations by MeHg can also have effects on synaptic transmission, such as increased presynaptic neurotransmitter release and altered postsynaptic response of the receptor (Sirois and Atchison, 1996; Yuan and Atchison, 1993, 1995, 1999). Studies using single cell Ca+2 imaging with the fluorescent dye fura-2 have shown that MeHg induces changes in divalent cation homeostasis in a neuroblastoma- glioma 30 hybrid cell line NG108-15, synaptosomes, and in cerebellar granule and Purkinje cells (Denny et al., 1993; Hare et al., 1993; Hare et al., 1995, Marty and Atchison 1997, 1998; Limke and Atchison, 2002; Edwards et al., 2005). Using the ratio of fura—2 fluorescence at the wavelengths of 340/380 nm as an indicator of relative intracellular Ca“2 concentrations, our lab has demonstrated that MeHg increases levels of intracellular Ca+2 in a multiphasic, kinetically-distinct manner. The initial increase in cytosolic Ca+2 concentration by MeHg is related to Ca+2 release from internal stores (i.e. endoplasmic reticulum and mitochondria) (Denny et al., 1993; Hare et al., 1993; Marty and Atchison, 1997; Limke and Atchison, 2002), and the final phase results from Ca+2 entry into the cell. Furthermore, a gradual increase in excitation ratio following the initial phase was also observed, which was inhibited or reversed by the divalent heavy metal chelator, TPEN, and was shown to be Can-independent. Studies determined that these Ca”- independent alterations in fura-2 fluorescence resulted from elevations of intracellular Zn”, resulting from Zn+2 displacement from cytoplasmic proteins (Denny and Atchison, 1994). The response of cerebellar granule cells and NG108-15 cells to MeHg differed only in their sensitivity to MeHg; granule cells were about ten times more sensitive to MeHg than were NG108-15 cells. Likewise, the response of granule cells to MeHg was similar to those of Purkinje cells, however the effects of MeHg in granule cells occured at earlier time points than those in Purkinje cells (Edwards et al., 2005). In addition, Purkinje cells were less sensitive to the MeHg-induced increase in [Ca+2]i and subsequent cell death. Thus, granule cells appear to be more sensitive to the effects of MeHg as compared to NG108-15 cells and Purkinje cells. 31 1!, E g E é b. Disruption of Synaptic Transmission by MeHg Any of the adverse cellular effects of MeHg described above could influence GABAergic synaptic transmission and play an important role in its neurotoxicity. This is important as GABAA receptors are vital in maintaining the excitability levels of cells. They accomplish this by mediating inhibitory synaptic transmission and regulating glutamatergic excitatory synaptic transmission (DeLorey and Olsen, 1992; Olsen and Avoli, 1997; Kardos, 1999). As cells do not operate in isolation and every cell is part of a vast network of cells, changes in excitability of one cell can affect the excitability of many others. Excess excitation can result in exitotoxic cell death (Kardos et al., 1999). Disruption of synaptic transmission and alteration of channel-mediated synaptic events by mercurials was first demonstrated in the peripheral nervous system in isolated nerve muscle preparations from cat cervical ganglion, frog sartorius muscle, and rat diaphram (Kostial and Landeka, 1975; Manalis and Cooper, 1975; Atchison and Narahashi, 1982). Intracellular recordings from these isolated nerve-muscle preparations demonstrated that both HgCl2 and MeHg completely reduced the nerve-evoked release of acetylcholine (ACh). In addition, MeHg affected excitatory endplate potentials (EPPs) by first producing a transient increase in EPP amplitude, followed by complete suppression of response (Atchison and Narahashi, 1982; Traxinger and Atchison, 1987). MeHg has been shown to inhibit mammalian neuromuscular synaptic transmission by decreasing the mean quantal content of neurotransmitter released. This reduction in mean quantal content is a result of a decrease in the available store of transmitter and an increase in the probability of its release (by increasing [Ca+2]i) (Atchison and Narahashi, 1982). This 32 suggests the involvement of MeHg in presynaptic mechanisms of synaptic transmission in the peripheral nervous system. Acute application of MeHg also disrupts synaptic transmission in the central nervous system. Similar to the effects of MeHg observed on EPPs, MeHg first increases and later inhibits population spike amplitudes in hippocampal neurons (Yuan and Atchison, 1993; 1994). These effects of MeHg are also observed on excitatory post- synaptic potential (EPSP) amplitudes in hippocampal slice (Yuan and Atchison, 1995). In this preparation, MeHg causes an initial increase followed by a complete reduction of both extracellular population spikes and of action potentials evoked by stimulation of Schaffer collaterals (Yuan and Atchison, 1993; 1994; 1995). These effects are time- and concentration-dependent. Increasing the stimulation intensity following the inhibition of population spikes by MeHg, causes a complete recovery of responses. This suggests that MeHg changes the excitability of the membrane, effectively increasing the threshold needed for firing of an action potential. Furthermore, in cell bodies of CA1 pyramidal neurons, MeHg inhibits both orthodromically- and antidromically-activated population spikes over a similar period of time. Because antidromic activation of population spikes does not involve synaptic connections (because current is carried to the cell body following axonal stimulation) it should not be as sensitive to the presynaptic effects of MeHg as would orthodromically-activated population spikes. Orthodromically-activated responses, which are induced by stimulation of Schaffer collaterals in region CA1, involve a synaptic mechanism that incorporates activation of CA1 neuron dendrites by axons of stimulated fibers. These responses should be especially sensitive to any presynaptic effects of MeHg. Because findings revealed that both orthodromically- and 33 antidromically-activated population spikes are reduced by MeHg over a similar time- scale, it suggests that MeHg predominately affects postsynaptic sites in hippocampus. Similar to the effects observed in hippocampus, MeHg also affects synaptic transmission in the cerebellum. Using both extracellular and intracellular recording techniques, acute bath application of MeHg causes an initial stimulation followed by a depression of cerebellar synaptic transmission between parallel fibers and Purkinje cells or between climbing fibers and Purkinje neurons (Yuan and Atchison, 1999). These effects of MeHg are both concentration- and time-dependent because higher concentrations of MeHg (100 pM) inhibit responses more rapidly than do lower concentrations (20 M). Furthermore, MeHg also hyperpolarizes and then depolarizes Purkinje cell membranes and suppresses the spontaneous activity of these cells. Recently, studies using whole-cell voltage-clamp of granule cells in cerebellar slice have revealed that spontaneous GABAergic inhibitory currents (sIPSCs) are more sensitive to MeHg inhibition than are spontaneous excitatory synaptic currents (sEPSCs) because sIPSCs are reduced more rapidly than are sEPSCs (Yuan and Atchison, 2003). Thus, GABAA receptors appear to be particularly sensitive to the inhibitory effects of MeHg. E. Disruption of GABAergic Function by MeHg The effects of mercury (Hg‘z), the divalent inorganic ion which is methylated to MeHg, on GABAergic function have been extensively investigated. Studies using the whole-cell patch clamp technique on dorsal root ganglia neurons of rats have found that HgCl2 potentiates GABAA receptor-mediated currents in the presence of low concentrations of GABA (30 11M) (Arakawa et al., 1991). The enhancement by HgCl2 34 was prevented by cysteine and iodoacetamide, suggesting the involvement of sulfliydryl groups in these actions (Huang and N arahashi, 1996). Hg+2 potentiation of GABA- mediated currents was shown to be concentration-dependent, in that larger concentrations of Hg+2 caused greater potentiation of responses. The effects of HgCl2 (0.1 - 100 M) on GABAA receptor mediated currents were voltage-independent, suggesting that the binding site for Hg2 was located on the external side of the receptor. Contradictory to these findings, the response produced by Hg+2 was also shown to be use-dependent, in that potentiation of GABAergic currents was enhanced with increases in frequency of channel opening. In addition, Hg+2 accelerated the rate of desensitization of GABAA receptor currents by decreasing the time constants for the fast and the slow phase of desensitization (Huang and Narahashi, 1996). Hg+2 was also shown to induce a slow inward current, insensitive to antagonists of voltage gated Na“, K+, or Ca+2 channels (Arakawa et al. 1991). Non-specific cation channels were thought to carry this current. Examination of HgCl2 potentiation of GABA-induced currents in the presence and absence of ZnCl2 revealed that the binding site of HgCl2 on the receptor differed from that of Zn+2 (Huang and Narahashi, 1996). Evidence has shown that stimulation of G5 protein by choleratoxin and inhibition of GilGo proteins by pertussis toxin decreased GABAA receptor-mediated currents (Huang and Narahashi, 1997a). In a follow-up study it was found that protein kinase A (PKA) negatively regulated GABAA receptor activity and abolished the Hg“2 potentiation of GABAergic responses (Huang and Narahashi, 1997 b). It was suggested that G-proteins, specifically - Gi/G0 acting to inhibit PKA - are involved in the Hg”- induced potentiation of IGABA. 35 In contrast to the extensively investigated effects of Hg+2 on GABAA receptor currents, little is known about the effects of MeHg on GABAergic function. Whole-cell patch clamp recordings in dorsal root ganglia neurons have shown that high concentrations of MeHg (100 pM) suppress the peak currents induced by low concentrations of GABA (30 M). Additionally, a slow inward current, presumed to be mediated by non-specific cation channels, was observed similar to that seen in response to Hg+2 (Arakawa et al., 1991). In hippocampus, GABAergic inhibitory synaptic transmission is more sensitive to inhibition by MeHg than is excitatory transmission (Yuan and Atchison, 1995, 1997). This preferential effect of MeHg on GABAA receptor- mediated inhibitory synaptic transmission was shown to underlie the early enhancement of excitatory synaptic transmission (Yuan and Atchison, 1997). Importantly, GABAergic synaptic transmission was shown to be more sensitive to MeHg inhibition in granule cells than in Purkinje cells because suppression of granule cell IPSCs occurs much earlier than does suppression of Purkinje cell currents (Yuan and Atchison, 2003). A spontaneous, transient, slow inward Cl'-mediated current was also seen in the slice preparation that was determined not to be GABAA receptor-mediated (Yukun and Atchison, 2005). The role of this current in MeHg-induced neurotoxicity is unknown. In primary cultures of cerebellar granule cells, MeHg (0.1 pM - 10 M) significantly suppressed whole cell GABAA receptor-mediated currents in a time- and concentration- dependent manner. MeHg also induced a slow inward current in these cells that was insensitive to voltage-activated Na", K”, and Ca+2 channel blockers. The Zn“2 chelator, TPEN, was able to delay the onset of this non-specific inward current as well as delay the time to decrease of GABAA currents (Xu and Atchison, 1997). 36 Several studies have suggested that MeHg interaction with the GABAA receptor enhances benzodiazepine binding in rodent cerebellum and cerebral cortex (Corda et al., 1981; Concas et al., 1983; Fonfria et al., 2001). Corda et a1. (1981) observed that a single administration of MeHg given by gavage produced a long-lasting increase in 3H- diazepam binding in rat cortex. Similarly, Concas et a1. (1983) showed that long-term administration of MeHg enhanced 3H-diazepam binding in rat cerebellum. In primary cultures of mice cerebellar granule cells, Fonfria et a1. (2001) observed that 30 min exposure to MeHg induced an increase in [3H]flunitrazepam binding at the GABAA receptor. However, in direct contradiction to these results, Komulainen and Tuomisto (1985) found no effect of either acute or chronic exposure to MeHg on the number of benzodiazepine receptors labeled with [3H]flunitrazepam in rat cerebellum. 37 SPECIFIC AIMS 1. What effects does MeHg have on GABAA receptor function in cells in culture? Does MeHg interact with a known modulatory site(s) on the GABAA receptor? Or does it bind to a different site(s) on the receptor? 2. Does MeHg differentially affect the functional properties of GABAA receptors in (16 subunit-containing and non-a6 subunit-containing cells? Are these actions subunit specific? Are dig-containing GABAA receptors of granule cells especially sensitive to the inhibitory effects of MeHg? Specific Aim 1 is addressed in Chapter 2. Time-course-of-[GABA-suppression experiments were performed with three different concentrations of MeHg (0.1, 1.0, and 10 uM) in cerebral cortical cells in culture. In addition, the effect of each concentration of MeHg on GABAA receptor mediated current kinetics was determined. Furthermore, the effects of diazepam (0.1, 1.0, and 10 uM) on [GABA current kinetics were investigated. These concentrations of MeHg are clinically relevant, as they are within the range of concentrations found in the blood of individuals poisoned with MeHg in the mass poisoning which occurred in Iraq in the 1970’s (Bakir et al., 1973). Following acute exposure to MeHg, signs of toxicity such as muscle weakness and ataxia occurred in poison-victims at concentrations ranging between 0.99 -1.56 moles of Hg. The concentration of MeHg associated with toxicity (1.0 moles of Hg) can be converted to approximately 12.5 uM, when the average adult male weight (80 kg) and tissue density (1.0 kg/liter) is considered. The lower concentrations of MeHg (0.1 and 1 uM) were used 38 in the present study to determine the concentrations at which the effects of MeHg become evident. To determine whether or not MeHg interacts with the diazepam binding site of the receptor, effects of MeHg on long». suppression and kinetics were examined in the presence of diazepam, either with or without pretreatment with MeHg or diazepam. Experiments which examined the effects of MeHg on IGABA in the presence of a benzodiazepine antagonist, flumazenil (10 uM), were also performed. The experiments from Specific Aim 2 are described in Chapter 3. Sensitivity to MeHg of cells containing (16 or al phenotyes of GABAA receptor was investigated using time-course-of-IGABA-suppression experiments performed with three different concentrations of MeHg (0.1, 1.0, and 10 uM) in both cerebellar granule (an-containing) and cerebral cortical (on-containing) cells in culture. To verify the presence of specific GABAA receptor subtypes of interest, immunocytochemistry was also performed in both cell types. In addition, the effect of each concentration of MeHg on the current-voltage relationship in granule and cortical cells was determined. Furthermore, to verify the findings obtained in native cells, experiments which examined the effects of MeHg on IGABA in human embryonic kidney (HEK-293) cells expressing either (16 or (xi-containing GABAA receptors were performed. To investigate these aims, several techniques were employed. To examine the effects of MeHg on functional properties of native GABAA receptors, whole-cell voltage- patch clamp recordings were performed. The patch-clamp technique is ideal in that it allows for measurements of whole-cell or single-channel currents at a millisecond time resolution (Hamill et al., 1981). GABAA receptor-mediated responses typically occur 39 within hundreds of milliseconds (Maconochie etal., 1994). Whole-cell recordings were performed on rat cerebellar granule and cortical neurons in primary culture. Cortical cells were used as a substitute for cerebellar Purkinje cells, as they contain the same type of GABAA receptor phenotype as do Purkinje cells (Hansen et al., 2001) and they are more successfully cultured. Purkinje cells are difficult to culture because they have an early mortality rate and often become contaminated with other cell types, particularly glial cells (Edwards, J ., personal communication). Furthermore, the distribution of Purkinje cells in the cerebellum is sparse (as compared to granule cells), they do not produce a large yield (Edwards, J., personal communication), and Purkinje cells have a very arborized dendritic tree, which is clipped in culture. To verify results obtained in granule and cortical neurons, experiments were also conducted in HEK-293 cells transfected with rat cDNA for either (11 or (16 subunit-containing GABAA receptors. To identify the target site(s) for MeHg, pharmacological characterization of GABAergic responses was examined using subunit-subtype specific agonists and antagonists. In addition, the presence of either on or (16 subunit-containing GABAA receptors in cells was determined using immunoctyochemistry. 40 CHAPTER TWO EFFECTS OF METHYLMERCURY ON DIAZEPAM-MEDIATED ENHANCEMENT OF GABAA RECEPTOR CURRENTS IN RAT CORTICAL NEURONS IN CULTURE 41 ABSTRACT GABAA receptor-mediated synaptic transmission is extremely sensitive to the effects of MeHg, as it is inhibited more quickly by MeHg than is excitatory synaptic transmission. The GABAA receptor has many drug binding sites, including those for barbiturates and benzodiazepines. However, the site at which MeHg interacts with the GABAA receptor has not yet been determined. Previous work in other labs suggested an effect of MeHg on benzodiazepine receptors. To characterize the interaction of MeHg with the GABAA receptor and to test if MeHg interacts with the benzodiazepine binding site of the receptor, diazepam (0.1, 1.0, and 10 M) or flumazenil (10 M), a benzodiazepine antagonist, was applied to cerebral cortical cells in primary culture in the absence and presence of MeHg (1 uM). [GABA were obtained by means of the whole-cell voltage patch-clamp technique using a symmetrical Cl' concentration. MeHg (0.1, 1.0, and 10 M) caused a progressive suppression of 15AM. Kinetically, MeHg prolonged IGABA slow decay time at each concentration tested. Similarly, IGABA slow decay time was prolonged at each [diazepam] tested. When MeHg was applied in conjunction with diazepam, the decay time of IGABA was prolonged further as compared to either MeHg or diazepam alone. Additionally, co-application of MeHg plus diazepam still prolonged IGABA slow decay following pretreatment with either MeHg or diazepam alone. Finally, flumazenil did not prevent or reduce the effects of MeHg on IGABA slow decay time. These results suggest that MeHg and diazepam have independent effects on the GABAA receptor and that MeHg does not appear to compete with the same binding site on the receptor as does diazepam, although MeHg does alter decay kinetics. 42 INTRODUCTION MeHg is a pervasive environmental contaminant that continues to pose a significant present-day toxicological threat, particularly in the Great Lakes, where levels of MeHg in fish have been found to exceed the estimated dose for adverse health effects in humans (Rice, 1995; Gerstenberger and Dellinger, 2002; Gilbertson, 2004; Weis, 2004). MeHg is highly neurotoxic and, in humans, can cause severe neurological signs including constriction of the visual field, speech impairment, mental disturbances, and ataxia (Chang, 1977). MeHg has a high affinity for sulfliydryl groups on cysteines which are numerous in cell membranes, proteins and enzymes. Thus, MeHg has the potential to bind to cell membrane proteins, change their conformation, and interfere with many cellular processes (Atchison and Hare, 1994; Sirois and Atchison, 1996; Denny and Atchison, 1996; Limke et al., 2004); Among these are disruption of excitatory and inhibitory synaptic transmission (Yuan and Atchison, 1993, 1997, 1999, 2003). GABAergic synaptic transmission is inhibited by MeHg. Whole-cell patch clamp recordings in dorsal root ganglion neurons have shown that high concentrations of MeHg (100 uM) suppress the peak currents induced by low concentrations of GABA (Arakawa etal., 1991). In hippocampal slice, MeHg was shown to decrease gradually inhibitory postsynaptic potential (IPSP) amplitudes to complete suppression (Yuan and Atchison, 1995). Suppression of inhibitory synaptic transmission by low concentrations of MeHg occurs earlier than does suppression of glutamatergic synaptic transmission (Yuan and Atchison, 1995, 1997), suggesting that inhibitory synaptic transmission is more sensitive to inhibition by MeHg than is excitatory transmission. Similarly, studies conducted in rats treated chronically with MeHg have suggested that MeHg impairs 43 inhibitory GABAergic neurons in cerebral cortex, neostriatum, and caudate putamen (O'Kusky and McGeer, 1985, 1989). Importantly, some cell types appear to be particularly sensitive to the effects of MeHg such as granule cells of the cerebellum. Suppression of inhibitory post synaptic currents (IPSCs) in cerebellar granule cells, for instance, occurs much earlier than does suppression in neighboring Purkinje cells (Yuan and Atchison, 2003). Whether or not MeHg differentially interferes with GABAergic function in different cell types is not known, nor has a specific mechanism for suppression of GABAA receptor function by MeHg been identified. One possibility for this difference in sensitivity to MeHg may be a differential expression of GABAA receptor subunits in various cell types, as different GABAA receptor subunit composition confers GABAA receptors with different pharmacological and electrophysiological properties. Current carried by GABAA receptors containing the (:1 subunit, for instance, is potentiated by benzodiazepines and La”, and is relatively insensitive to inhibition by Zn+2 and furosemide, as compared with that carried by (16 subunit-containing receptors. In addition, the on subunit-containing receptor has a low efficacy and affinity for pentobarbital. Furthermore, receptors containing the (1; subunit desensitize more slowly than do other subtypes of GABA receptors, such as those containing the (1., subunit (Tia et aL,1996) Several studies have suggested that MeHg interaction with the GABAA receptor involves the benzodiazepine binding site. MeHg enhances benzodiazepine binding to GABAA receptors in rodent cerebellum and cerebral cortex (Corda et al., 1981; Concas et al., 1983; Fonfria et al., 2001). Corda et a1. (1981) observed that a single administration of MeHg produced a long-lasting increase in 3H-diazepam binding in rat cortex. Similarly, Concas et a1. (1983) showed that long-term administration of MeHg enhanced 3H-diazepam binding in rat cerebellum. In primary cultures of mice cerebellar granule cells, Fonfria et a1. (2001) observed that short-term exposure to MeHg induced an increase in [3H]flunitrazepam binding at the GABAA receptor. However, in direct contradiction to these results, Komulainen and Tuomisto (1985) found no effect of either acute or chronic exposure to MeHg on the number of benzodiazepine receptors labeled with [3H]flunitrazepam in rat cerebellum. Thus, whether or not MeHg interacts with the benzodiazepine binding site of the GABAA receptor to affect GABAA receptor function remains to be determined. The present study was designed specifically to characterize the interaction of MeHg with the (11 subunit-containing GABAA receptor, as the diazepam binding site has been implicated in the interactions of MeHg with the GABAA receptor. Using the whole-cell voltage patch-clamp technique, results from the present study show that the effects of MeHg are not changed in the presence of diazepam or a benzodiazepine antagonist. These findings suggest that MeHg interacts at a site on the GABAA receptor different from that used by diazepam. 45 METHODS Solutions and Chemicals Methylmercuric chloride (MeHg) (ICN Biomedical lnc., Costa Mesa, CA, USA) was dissolved in deionized water to a final concentration of 10 mM, which served as a stock solution. On the day of experiments, MeHg working solutions (0.1, 1, or 10 uM) were diluted in extracellular solution consisting of (in mM): NaCl, 125; CaCl2, 2.0; KCl, 2.5; MgCl2, 1.0; KH2PO4, 1.25; NaHC03, 26.0, and D-glucose, 20.0, pH-adjusted to 7.4 using 95% 02/5% CO2. Diazepam, 7-amino-n-butyric acid (GABA), flumazenil, bicuculline, D-penicillamine, 6-cyano-7-nitroquinoxaline-2,3-dione disodium salt (CNQX), DL-2-amino-5-phosphonopentanoic acid (APV), glutamine, 4-(2- hydroxyethyl)piperazine-l-ethanesulfonic acid (HEPES), gentamycin, deoxyribonuclease I (DNase I), cytosine B-D-arabino-furanoside (Ara-C), adenosine 5'- triphosphate magnesium salt (ATP), ethylene glycol-bis(B-aminoethyl ether)-N, N, N', N',-tetraacetic acid (EGTA), trypsin, and poly-D-lysine were all purchased from Sigma Chemical Co. (St. Louis, MO). Dulbecco’s Modified Eagle’s Medium (DMEM) was purchased from Mediatech, Inc. (Hemdon, VA), and heat inactivated horse serum was purchased from GIBCO-Invitrogen Inc. (Carlsbad, CA). Diazepam and flumazenil were dissolved in dimethyl sulphoxide (DMSO; < 0.01% (v/v) final concentration) and diluted in extracellular solution prior to use. Other drugs were dissolved in deionized water and diluted in extracellular solution prior to use. 46 Preparation of Primary Cortical Cell Culture Primary cultures of cortical neurons were obtained from 3-4 newborn Sprague- Dawley rat pups (Charles River Laboratories, Wilmington, MA) following methods described by Inglefield et al. (2001). Briefly, following removal of the brain and isolation of cortex, cells were digested for 4.5 min at 37° C with trypsin 0.025 % (w/v) in buffer that contained (in mM): KCI (5.0), KH2PO4 (0.205), NaCl (137.0), Na2HPO4 (0.17), D-glucose (5.0), sucrose (59.0), and gentamycin (0.1 mg/ml), pH 7.3. Trypsin was inactivated by addition of the above-described buffer containing 0.016 % (w/v) DNase I for 4.5 min at 37° C. To this mixture, warmed DMEM supplemented with 10 % (w/v) horse serum, 10 mM HEPES, 2 mM glutamine, and gentamycin (0.1 mg/ml) was added and the cells were centrifuged (SOOxg, 5 min). The resulting pellet was resuspended in DMEM containing DNase I and was centrifuged again. The cell suspension was then resuspended in DMEM and triturated with a Pasteur pipette. Cells were plated at a density of l x 106 cells/ dish onto 35-mm Petri dishes (Corning 1nc., Corning, NY) coated with poly-D-lysine. Cells were maintained in a 37° C incubator with 95% O2 and 5% CO2. DMEM was replaced every two days. Cytosine [3- arabinoside (5 M) was added to cell cultures 72 hr after plating to inhibit glial cell proliferation. Cells were used for recordings 10-14 days after plating. All experiments were done in compliance with NIH and university standards and were approved by the Michigan State University Committee on Animal Use and Care. Whole-cell Recordings in Cortical Cells Prior to making whole-cell patch voltage-clamp recordings, the Petri dish containing the cells was rinsed and then covered with approximately 1 ml of recording 47 solution. CNQX (10 M) and APV (50 M) were added to the extracellular solution to inhibit ionotropic glutamate receptor-mediated responses. Recording patch electrodes were fabricated from borosilicate capillary glass (0D: 1.5 mm, 1D: 1.0 mm; Garner Glass Co., Claremont, CA) and had an impedance of 3 - 7 MO when filled with a pipette solution consisting of (in mM): CsCl, 140; NaCl, 4.0; CaCl2, 0.5; HEPES, 10.0; EGTA- CsOH, 5.0; Mg-ATP, 2.0, pH adjusted to 7.3. Following acquisition of a cell-attached patch and cancellation of capacitative currents, the whole-cell configuration was obtained by negative pressure applied to the syringe. Cells were voltage clamped at -60 mV so that GABA application produced inward currents under our approximately symmetric [Cl'] conditions. The zero current potential was near 0 mV, which is close to the expected Cl' equilibrium potential. Capacitance transient neutralization and series resistance compensation were optimized. All experiments were carried out at room temperature of 22 - 25° C. GABA was pulsed onto neurons using a picospritzer (Picospritzer II, General Valve Corporation, Fairfield, NJ) delivered through a glass pipette (impedance of ~ 900 k0) placed close to the cell. GABA (50 M) was applied onto the target cells by a 10 ms pulse given at intervals of 30 s. This was long enough for GABAA receptors to recover completely from desensitization (results not shown). Physiological saline alone or containing MeHg, diazepam and/or flumazenil was delivered into the extracellular bathing solution at a rate of approximately 0.45 ml/min via a gravity—powered perfusion system through a series of tubes placed close to the cell. To observe the effects of MeHg on IGABA, recordings were made in the presence or absence of MeHg. In some experiments, diazepam (0.1- 10 M) or flumazenil (10 M), a benzodiazepine antagonist, 48 was applied simultaneously with MeHg to test for any competitive interactions of the drugs. In other experiments, cells were pre-treated with either MeHg (1 M) or diazepam (1 uM) prior to co-application of both agents to allow sufficient time for MeHg or diazepam to interact with the receptor and affect IGABA. Pretreatment time with MeHg was approximately 15 — 20 min, the time needed for MeHg to produce prolongation of IGABA slow decay rate. Pretreatment with diazepam was approximately 1- 2 min, or until maximum effects on prolongation of IGABA decay rate were evident. In the present experiments, three MeHg concentrations were used (0.1, 1.0, and 10 pM). Data Acquisition and Analysis Data were acquired every 30 s using an IBM compatible computer equipped with a DigiData 1200 interface and pClamp 8.1 software (Axon Instruments, Union City, CA). IGABA were recorded using an Axopatch l-D amplifier and low-pass filtered at 1 kHz with an eight pole filter before digitization at 10 kHz, storage and display. Off-line analysis was performed using Clampfit 8.1 software (Axon Instruments) and the MiniAnalysis program 5.6.10 (Synaptosoft Inc., Decatur, GA). Using MiniAnalysis software, effects of MeHg on kinetics of Iona». were examined. Each GABAA receptor-mediated event was selected manually and marked. The maximum amplitude for each event was measured at the peak of the inward current. The decay phase was best fitted using the Levenberg- Marquardt least-squares method (Levenberg, 1944; Marquardt, 1963) with a double- exponential function of the form 2 ant", where n is the number of exponential components, a is the relative amplitude of the component, and r is the time constant. Statistical analyses of data collected before and during application of MeHg were 49 conducted using a Student's paired t-test or a one-way analysis of variance, depending upon the number of groups being compared. The Tukey-Kramer test was used for post hoc comparisons. Values were considered statistically significant at p<0.05. The data are presented as mean :SEM and the number of replications is given. The data were obtained from approximately 46 separate cell cultures, and the number of replications refers to individual cells obtained from at least 2 separate cell cultures. 50 o i t '— 4. ‘V T L 5. ‘. ‘ -i i RESULTS MeHg Suppresses [GABA and Prolongs [GABA Slow Decay Under symmetrical [Cl'], application of a GABA pulse produced an inward current, which was completely abolished by bath application the GABAA receptor antagonist bicuculline (10 M) (Fig. 2.1A), indicating that this inward current was a GABAA receptor-mediated response. Unlike bicuculline, which decreases IGABA rapidly and reversibly, continuous perfusion of 1 uM MeHg into the bathing solution had little immediate effect on IGABA (Fig. 2.1B). However, as shown in Figure 2.2, with additional time of exposure, MeHg caused a progressive reduction of IGABA. The effects of MeHg never reached a steady-state level less than complete suppression of IGABA (Fig. 2.2A). This effect was not reversible by washing cells for 5 min with MeHg-free solution containing D-penicillamine (20 M), a MeHg chelator (Fig. 2.2B). The time course of [GABA reduction by MeHg is shown in Figure 3A. Onset of reduction typically occurred between 2 - 8 min, depending upon the MeHg concentration used. Increasing the concentration of MeHg shortened the time course of suppression. For each concentration of MeHg tested, the time-course of MeHg suppression was approximately linear. Time to total reduction of [GABA by MeHg differed significantly between each concentration of MeHg tested (p<0.05) (Fig. 2.3B). At 10 M MeHg, the highest concentration tested, complete reduction of IGABA occurred within approximately 30 min. IGABA in cortical cells typically decayed in a bi-exponential manner consisting of a fast (In...) and slow (15.0w) component as shown in Figure 2.4A. The fast component decayed at an average of 650 :I: 102 ms and the slow component at 51 Figure 2.1. MeHg causes a gradual suppression of bicuculline-sensitive IGABA in cortical cells. A, Effects of 10 pM bicuculline on GABAA receptor-mediated currents. Whole-cell IGABA were evoked from a holding potential of -— 60 mV by a 10 ms pulse of GABA (50 M, black dot), at intervals of 30 s, in the presence of CNQX (10 M) and APV (50 pM) in the external solution to block glutamatergic currents. Data were collected before (control), after 150 s (hash bar) exposure to bicuculline, and after 180 s (hash bar) wash with bicuculline-free solution. Bicuculline caused a complete and reversible suppression of IGABA. B, Comparison of early time course of effects of 10 M bicuculline (top) and 1 uM MeHg (bottom) on consecutive GABAA receptor-mediated currents. IGABA were obtained under identical conditions to those described above. Data were collected before (control) and after application of bicuculline (top) or MeHg (bottom) (as indicated by an arrow). Bicuculline caused immediate reduction of [GABA whereas over the same time course MeHg did not affect IGABA. 52 A Control Bicuculline Wash ' /o 1[/0 2000 ms V B Control Bicuculline 30 s < D. O 8 2000 ms Control MeHg 30 S r i <2: G. O O 1.0 2000 ms Figure 2.1 53 Figure 2.2. MeHg causes a complete and irreversible suppression of loan in cortical cells. A, Time course of effects of 1 uM MeHg on GABAA receptor-mediated currents. IGABA were recorded under identical conditions to those described in Figure 2.1. Data were collected before (control) and at various time points after exposure to MeHg (6, 8, 11, 16, and 37 min). MeHg caused a progressive and complete reduction of IGABA. B, Effects of MeHg on IGABA are irreversible, Currents were collected before (control), after 35 min (hash bar) exposure to MeHg (1 M), and following 10 min (hash bar) wash with D-penicillamine (D-pen)-containing (20 uM), MeHg-free solution. 54 GABA O 37 min 16 min 11 min 8 min 6 min Control 1000 ms (0 min) B Control MeHg D-pen Figure 2.2 55 Figure 2.3. Suppression of IGABA by MeHg in cortical cells is time- and concentration-dependent. A, IGABA were recorded in rat cortical cells in primary culture at [MeHg] of 0.1 uM (open squares), 1.0 7.1M (filled triangles), or 10 M (open circles). Recording conditions were as described in Figure 2.1. Data were collected continually before and after MeHg exposure. Current amplitude in response to MeHg was normalized to that of control. Each datum point represents the mean value recorded from 3-5 cells. B, Comparative effects of different concentrations of MeHg on time to complete reduction of IGABA. Each bar represents the mean values obtained from 3-5 cells. The times to total reduction of IGABA by MeHg differed significantly between each concentration of MeHg tested, as is indicated by the asterisks (*, p<0.05). 56 0| ’7; 25 E 8 °\° 50 < 2 .6 75 100 III 60r Time to Total Reduction (min) 0.1 1.0 10 [MeHg] (HM) Figure 2.3 57 approximately 1291 i 161 ms. In the presence of MeHg, the fast component appeared to be unaffected (data not shown), but 15.0,, was prolonged significantly (p<0.05) (Fig. 2.4A). However, the percent increase over control in 15.0“, produced by MeHg was concentration-independent (Fig. 2.4B). The time courses of prolongation of rslow of [GABA by MeHg are shown in Figure 2.5A. Onset of slow decay prolongation typically occurred between 1.5 — 18 min, depending upon the MeHg concentration used (Fig. 2.5B). Increasing the concentration of MeHg shortened the time to onset of slow decay prolongation. Similarly, time to peak prolongation of slow decay by MeHg was concentration-dependent (Fig. 2.5C). The higher the concentration of MeHg, the less time was required to reach the peak prolongation of IGABA slow decay. Effects of MeHg on [GABA are Not Altered in the Presence of Diazepam For comparison, effects of diazepam on IGABA in cortical cells were also examined. Diazepam (0.1, 1.0, or 10 M) was continuously bath-perfused onto cortical cells during 30 s pulses of GABA. Consistent with previous reports on the effects of benzodiazepines on IGABA (Vicini et a1, 1986; Sieghart, 1992), diazepam prolonged the decay of IGABA. In particular, diazepam prolonged the slower component of decay (Tsrow) (Fig. 2.6A), usually within 30 — 60 s after application. The percent increase over control in rslow produced by diazepam was concentration-dependent, as it differed significantly between each concentration of diazepam tested (Fig. 2.6B) (p<0.05). Thus, diazepam and MeHg may act at a similar site via a similar mechanism to prolong IGABA slow decay. If this is the case, then diazepam and MeHg could compete for the same binding site on the GABAAreceptor. 58 Figure 2.4. MeHg prolongs [GABA slow decay constant in a concentration- independent manner. A, Representative IGABA (control, dark line; 1510“,: 981.4 ms) showing the effect of 1 uM MeHg (dotted line; rsiow: 1451.5 ms) on slow decay rate (15.0w). Current amplitude in response to MeHg was scaled to that of control. IGABA were recorded under conditions identical to those described in Figure 2.1. B, Comparative effects of MeHg on prolongation of Tsjow before (control, black bar) and after exposure to MeHg at the concentrations indicated (0.1 uM - 10 M, grey bars). Each bar represents the mean values of rsiow prolongation recorded from 15-16 cells. MeHg prolonged mm in a concentration-independent manner that differed significantly from control, as indicated by the asterisk (*, p< 0.05). 59 GABA O '1 :981.4 ms slow <2 500 1000 ms Control 0.1 1,0 10 [MeHg] (MM) Figure 2.4 60 Figure 2.5. MeHg affects the time course of prolongation of [GABA slow decay in a concentration-dependent manner. A, IGABA were recorded at [MeHg] of 10 M (dashed line, open circles), 1.0 7.1M (filled line, open triangles), or 0.1 uM (dashed and dotted line, filled crosses). Recording conditions were as described in Figure 2.1. Data were collected immediately before and approximately every 60 -90 s after MeHg exposure. Each trace represents the mean values recorded from 3-5 cells. B, Comparative effects of MeHg on time to onset of prolongation of tslow, Data were collected after exposure to MeHg at the concentrations indicated (0.1 uM — 10 M, grey bars). Each bar represents the mean values obtained from 3-5 cells. Times to onset of rsiow prolongation differed significantly between each concentration of MeHg tested, such that times were lessened by higher concentrations of MeHg, as indicated by an asterisk (*, p< 0.05). C, Comparative effects of MeHg on time to peak of prolongation of Tsjow. Data were collected after exposure to MeHg at the concentrations indicated (0.1 uM — 10 M, grey bars). Each bar represents the mean values obtained from 3-5 cells. Times to peak of Tsjow prolongation differed significantly between each concentration of MeHg tested, such that times were reduced at higher concentrations of MeHg, as indicated by the asterisk (*, p< 0.05). 61 A —+- 0.1 uM 200 ' 50 — "9" 10uM + 1pM coscoo i .3? 40 30 20 Time (min) 10 B 30” 2E5 580 2 25 [MeHg] (MM) Figure 2.5 62 O E E x (U G) o. .9. (D .E I— 10 1.0 0.1 [MeHg] (71M) Figure 2.5 63 Figure 2.6. Diazepam prolongs chnn slow decay constant in a concentration- dependent manner. A, Representative IGABA (control, dark line, rslow: 413 ms) showing the effects of 1 uM diazepam (dashed line; Tsjowi 722.6 ms) on slow decay rate. Current amplitude in response to MeHg was scaled to that of control. B, Comparative effects of diazepam on concentration-dependent prolongation of Tsjow (control, black bar; diazepam, 0.1 uM - 10 M, hashed bars). Each bar represents the averaged values obtained from 14 -15 cells. Prolongation of Tsjow (% control) differed significantly among concentrations of diazepam tested as indicated by the asterisk (*, p<0.05). A GABA O U) 200 150 Tsiow (°/o Control) Figure 2.6 Control I I 1.10.; 413.0 ms ’ 1' -722.6 ms I 4— slow' 0pA O 10 500 ms 0.1 1.0 10 [Diazepam] (11M) 65 To test this, the combined effects of MeHg and diazepam on IGABA were examined. The intermediate [MeHg], 1 7.1M , was chosen for all these studies, inasmuch as the percent reduction of [GABA and the percent prolongation of Tsjow by MeHg appeared to be concentration-independent. Co-application of MeHg with diazepam (1 and 10 pM) prolonged significantly Tsjow as compared to the effect of MeHg alone (p<0.05) (Fig. 2.7). However, prolongation of [GABA Tsjow by MeHg with diazepam (0.1 uM) did not reach significance (p>0.05). Furthermore, co-application of diazepam with MeHg did not alter the time course of IGABA reduction by MeHg (Fig. 2.8). Next, cells were pretreated with diazepam prior to co—application of MeHg plus diazepam to allow sufficient time for diazepam to interact with the receptor. Pretreatment with diazepam did not interfere with the effects of MeHg on 10AM. As shown in Figure 2.9A-B, under these conditions MeHg plus diazepam prolonged rsiow to a significantly greater extent as compared to the effect of diazepam pretreatment alone (p<0.05). At a lower concentration of diazepam (0.1 uM), however, a significant prolongation of rslow following diazepam pretreatment was not observed (p>0.05). In addition, the time course of [GABA reduction by MeHg was not affected by co-application, following pretreatment with diazepam (Fig. 2.8) (p>0.05). Effects of Diazepam on IGABA Are Not Altered in the Presence of MeHg To allow sufficient time for MeHg to interact with the receptor, cells were then pretreated with MeHg prior to a combined application of diazepam and MeHg. Pretreatment with MeHg did not interfere with the effects of diazepam on long... As 66 Figure 2.7. Co-application of MeHg and diazepam does not affect MeHg-induced prolongation of IGABA slow decay. A, Representative traces depict effects of co- application of 1 pM MeHg plus 1 uM diazepam (dashed line; rsiow: 817.5 ms) on slow decay time constant of [GABA (control, dark line, Tslow: 665.4 ms). Current amplitude in response to MeHg was scaled to that of control. B, Comparison of effects of 1 uM MeHg (dark grey bar) and co-application of 1 uM MeHg plus diazepam (0.1 uM - 10 M, hashed bars) on prolongation of Tsiow (control, black bar). Each bar represents the averaged values obtained from 6 - 8 cells. Prolongation of 1510,. produced by MeHg plus diazepam differed significantly from control and the effects produced by 1 pM MeHg alone at 1 M and 10 M diazepam as indicated by the asterisk (*, p<0.05). 67 ' rslow: 665.4 ms ‘— Tslow: 817.5 ms B 1000 ms g 150:: T %/_ / (C3: 100 — -. % / / IBIZebnjrll (11:4; +1 ill; Merl; 68 Figure 2.8. Time course of [GABA suppression by MeHg in cortical cells is not altered by application of diazepam or flumazenil. A, Comparison of time courses of IGABA reduction by 1 pM MeHg (filled circles, dotted line), co-application of MeHg plus 1 uM diazepam, following MeHg pretreatment (open circles), co-application of MeHg plus diazepam following diazepam pretreatment (filled triangles, hashed line), co-application of MeHg plus 10 pM flumazenil (open diamonds), and co-application of MeHg plus diazepam (open squares). Each datum point represents the averaged values obtained from 3-5 cells. B, Comparative effects on time to reduction of [GABA by MeHg (filled circles), MeHg plus flumazenil (open diamonds), and MeHg plus diazepam, either with MeHg (open circles) or diazepam pretreatment (filled triangles) or without (open squares). Each bar represents the averaged values obtained from 4 - 5 cells. Time to [GABA reduction by MeHg did not differ significantly between any group tested (p>0.05). 69 o I -+- MeHg —9— MeHg Pre-treat C: —i'- DZ Pre-treat 9 25 7 —°— MeHg+Flu E MeHg+DZ O O r- °\o 50 < 3 _0 75 - 1oo ' ' 7“. o‘. . . 0 10 20 30 40 50 Time (min) E 607 E C l .9 ‘6 40 r 3 '0 (D 0: E 20 ‘ O l— 9 (D 0 g MeHg MeHg DZ MeHg MeHg I- Pre- Pre- +Flu +DZ Figure 2.8 70 Figure 2.9. Diazepam or Mel-lg pretreatment, following co-application, does not prevent diazepam- or MeHg-induced prolongation of IGABA slow decay. A, Representative traces depict effects of 1 uM diazepam (dashed line; 1510“,: 722.6 ms) and 1 uM MeHg plus diazepam (dotted line; Tslowt 897.2 ms), following diazepam pretreatment (control, dark line; Tslowi 413.0 ms) on IGABA 15.0w. Current amplitudes in response to diazepam or diazepam plus MeHg were scaled to those of control. B, Comparative effects of 1 uM MeHg (dark grey bar) and co-application of 1 7.1M MeHg plus diazepam, following diazepam pretreatment (0.1 uM — 10 71M, hashed bars) on prolongation of Tsjow (control, black bar). Each bar represents the averaged values obtained from 4-5 cells. Prolongation of IGABA slow decay produced by diazepam (0.1, 1.0, 10 M) alone was significantly different from control as indicated by the asterisk (*, p<0.05). Prolongation of IGABA slow decay produced by the combination of MeHg plus 1 or 10 M diazepam, following diazepam pretreatment, differed significantly from that produced by diazepam alone as indicated by the cross (T, p<0.05). C, Representative traces depict effects of application of 1 uM MeHg alone (hashed line; rslow: 1451.5 ms; and co-application of 1 11M MeHg plus 1 uM diazepam, following MeHg pretreatment (dotted line, Tsjow: 1731.0 ms), on slow decay time constant of IGABA (control, dark line, 15.0w: 981.4 ms). Current amplitudes in response to MeHg or MeHg plus diazepam were scaled to those of control. D, Comparative effects of 1 uM MeHg (dark grey bar) and co-application of 1 pM MeHg plus diazepam, following diazepam pretreatment (0.1 uM - 10 pM, hashed bars) on prolongation of Tsjow (control, black bar). Each bar represents the averaged values obtained from 6 - 8 cells. Prolongation of Tsjow produced by MeHg plus 1 or 10 M 71 diazepam, following MeHg pretreatment, differed significantly from control and that produced by MeHg alone as indicated by the asterisk (*, p<0.05). 72 '- I'I’rsmw: 722.6 ms ism: 897.2 ms rm: 413.0 ms ........ .,_ .....,...... .. in ..w . .. .. . . .. .1 . fl ...i.. .. . . . . _ 000000 55555 coacoo as as 10 1.0 [Diazepam] (uM) Figure 2.9 73 GABA O ... I'.‘ "' "I‘I‘q r at . 'r.___;,.i_ Tslow' 1451.5 ms ' a‘” . ”mm“ *— Tsiow- 1731.0 ms , 3,1!“ rslow: 981.4 ms 500 pA D s: W 0 Control MeHg 0.1 1.0 10 Figure 2.9 [Diazepam] (uM) + 1 11M MeHg 74 Figure 2.10. Co-application of MeHg and flumazenil does not prevent MeHg- induced prolongation of [GABA slow decay. A, The effects of diazepam on IGABA were inhibited by 10 uM flumazenil. IGABA were obtained as described in Figure 2.1. Current traces were collected before (control), after exposure to diazepam (1 M), and after exposure to diazepam plus flumazenil. B, Flumazenil itself produced no effects on IGABA. Currents were recorded before (control), after flumazenil exposure, and after flumazenil plus MeHg exposure (at times indicated by arrows). In the presence of flumazenil, MeHg still decreased IGABA amplitude and caused a slowing of the decay rate. C, Representative traces depict effects of co-application of 1 pM MeHg plus 10 11M flumazenil (dotted line; Tsjow: 1171.3 ms) on IGABA slow decay time constant (control, dark line; 15.0w: 1007.0 ms). Current amplitudes in response to flumazenil plus MeHg were scaled to those of control. D, Comparative effects of 1 uM MeHg (dark grey bar) and co-application of MeHg plus 10 M flumazenil (hashed bar) on prolongation of Tsjow (control, black bar). Each bar represents the averaged values obtained from 4 - 5 cells. Prolongation of 12510“, produced by MeHg plus flumazenil did not differ significantly from that produced by MeHg alone (p>0.05). 75 A Diazepam + Flu Control Diazepam 603 71 Control Flu 150 s fir Figure 2.10 76 120 s i? <2 0. §L_ 1000 ms Flu + MeHg 600 s n i <12 0. O O O " 1000 ms :1171.3 ms 1000 ms D E100” %% Control MeHg MeHg + Flu Figure 2.10 77 shown in Figure 2.9C-D, under these conditions MeHg plus diazepam (1.0 and 10 M), but not 0.1 pM diazepam, caused a significantly greater prolongation of rslow as compared to MeHg treatment alone (p<0.05). However, this effect did not reach significance for co-application with MeHg. Furthermore, co-application, following MeHg pretreatment, did not alter the time course of [GABA reduction by MeHg (Fig. 2.8). Flumazenil Does Not Alter the Effects of MeHg on [GABA To determine further whether or not the MeHg interacts with the action or binding site of benzodiazepines on the GABAA receptor, the benzodiazepine antagonist, flumazenil was co-applied with MeHg. The rationale was that if MeHg acts at the benzodiazepine site of the receptor, then blocking the site with an antagonist should reduce or prevent the effects of MeHg on IGABA. At the concentration of flumazenil (10 M) used, prolongation of IGABA by diazepam (1 M) was abolished (Fig. 2.10A). Flumazenil, in and of itself, had no effect on [GABA as shown in Figure 2.10B. However, co-application of flumazenil did not prevent MeHg from prolonging 1510“,; the effect of combined treatment was not significantly different from that produced by MeHg alone (Fig. 2.10C-D) (p>0.05). In addition, flumazenil did not alter the time course of IGABA reduction by MeHg as shown in Figure 2.8 (p>0.05). These results show that flumazenil did not interfere with the effects of MeHg on 10A“. 78 DISCUSSION The primary goals of the present study were to characterize the interaction of MeHg with the at subunit-containing GABAA receptor, and specifically to determine whether or not MeHg interacts with the benzodiazepine binding site of the receptor. We found that MeHg (0.1, 1.0, and 10 uM) caused a progressive suppression of IGABA. Kinetically, MeHg prolonged 15.0“, of IGABA in a concentration-independent manner. Furthermore, MeHg did not appear to interact with the diazepam site of the GABAA receptor, as co-application of MeHg (1 uM) and diazepam (1 and 10 11M) did not alter the pattern of effects of diazepam alone or MeHg alone on IGABA, nor did the benzodiazepine antagonist flumazenil prevent or reduce the effect of MeHg on tslow, At all concentrations of MeHg used, the pattern of effects of MeHg on IGABA was similar —the current was suppressed completely and irreversibly. These findings are qualitatively consistent with effects of MeHg on IGABA reported previously but occurred at a much lower range of concentrations. In dorsal root ganglion neurons, high concentrations of MeHg (100 uM) suppressed the peak currents induced by low concentrations of GABA (Arakawa et al., 1991). Similarly, in hippocampal and cerebellar slice, MeHg (20 and 100 uM) gradually decreased lPSPs to complete suppression (Yuan and Atchison, 1997). The time to suppression of [GABA responses differed between concentrations of MeHg tested, such that the time to suppression of responses was prolonged at lower concentrations of MeHg as compared with higher concentrations. The magnitude of effects produced by different concentrations of MeHg, however, did not differ. This is consistent with what is known about the general pattern of effects of MeHg. At each concentration of MeHg tested (0.1, 1.0, and 10 uM), the 79 time-course of reduction of IGABA was approximately linear and rsiow was prolonged to approximately the same extent. The fixed magnitude of effects caused by different concentrations of MeHg suggests that a consistent series of actions takes place once an effective or threshold concentration of MeHg is attained. In addition to causing IGABA suppression, MeHg prolonged the slow component of IGABA decay. GABAA receptor decay is important as it is a major determinant of IPSC duration. The faster component of decay is associated with receptor desensitization, whereas the slower component of decay is typically associated with GABA receptor deactivation. Deactivation rate is influenced by unbinding of GABA from its receptor as well as by transitions of activated receptors between open, closed, and desensitized states (MacDonald et al., 1989a; Twyman et al., 1990; Twyman and MacDonald, 1992). Even though removal of agonist is considered the final step in current deactivation, it can be prolonged by detaining the receptor in the desensitized state (Jones and Westbrook, 1995; Tia et al., 1996; Haas and MacDonald, 1999). Unbinding is prolonged from the desensitized state because the receptor needs first to recover from the desensitized state to reopen prior to unbinding of GABA. Thus, GABAA receptor deactivation can be shaped by several different mechanisms. In the present study, the prolongation of Tsjow by MeHg likely indicates that MeHg slows the rate of unbinding of GABA from its receptor. Alternatively, it could indicate that MeHg produces an accumulation of desensitized GABAA receptors, resulting in prolonged unbinding of GABA from its receptor. MeHg did not appear to affect IGABA fast decay, suggesting that MeHg does not affect GABAA receptor desensitization. However, the present experiments may not have resolved appropriately the fast phases of decay, or desensitization, due to the use of a relatively 80 slow perfusion system. Thus, additional studies, such as recovery from desensitization experiments measured by paired GABA pulses, are needed to determine specifically if MeHg affects GABAA receptor desensitization. MeHg most likely does not compete for the same binding site on the GABAA receptor as does diazepam, because MeHg was able to maintain its effects on IGABA in the presence of diazepam. However, it is also possible that diazepam did not interfere with the actions of MeHg at the GABAA receptor because diazepam and MeHg were not applied at concentrations high enough to saturate the receptor. As a result, the effects of diazepam and MeHg would be additive, as neither drug might have reached its maximum “ceiling” effect. However, even in the presence of flumazenil, a benzodiazepine antagonist, MeHg maintained its effects on IGABA, indicating that the effects of MeHg on IGABA are independent of the actions of diazepam. Although some binding studies showed previously that short-terrn (Corda et al., 1981; Fonfria et al., 2001) as well as long-term administration (Concas et al., 1983) of MeHg enhances benzodiazepine binding to the GABAA receptor, the results from the present study do not necessarily contradict this. Increased binding of diazepam to the GABAA receptor after MeHg exposure, does not necessarily translate into an increase in GABAA receptor function or currents. Instead, increased binding of diazepam in response to MeHg exposure could result from a compensatory response by the cell to increase GABAA receptor binding sites due to antagonism of the GABAA receptor by MeHg. Chronic administration of the benzodiazepine antagonist, R015-1788, to mice caused an upregulation of benzodiazepine binding sites on the GABAA receptor in cortex, cerebellum, and hippocampus (Miller et al., 1989), supporting the idea that an increase in 81 binding sites may result as a compensatory response to antagonist application. Furthermore, experiments by Corda et a1. (1981) and Concas et a1. (1983) involved the use of a whole animal model. The type of model may affect the way in which MeHg interacts with the GABAA receptor, as MeHg acting in vivo may produce compensatory effects in the whole animal not seen in cells in culture, resulting in an upregulation of GABA... receptors in response to MeHg antagonism. In addition, Fonfria et a1. (2001) used a mouse model, as opposed to rat which was used in the present investigation. This is important because effects of MeHg can differ between mice and rats, as was demonstrated by Basu et a1. (2005) in a study in which the inhibition of radioligand binding to muscarinic acetylcholine receptors by MeHg was more pronounced in rat brain as compared with mouse. The present study suggests that MeHg does not interact with the diazepam site of the GABAA receptor, thus alternative binding sites ought to be considered. Results from our lab have indicated recently that MeHg does not bind to the channel pore of the receptor, because stepping the voltage from ~80 mV to +60 mV in the presence of MeHg produced a linear (voltage-independent) I-V relationship (Herden et al., 2004). Furthermore, our lab has shown that MeHg does not appear to bind to the GABA site on the receptor, as time to suppression of IGABA by MeHg was not dependent upon GABA (10 pM — 1 mM) concentration in cortical cells in culture (Herden et al., 2004). Moreover, our observation that MeHg prolongs [GABA slow decay may indicate that MeHg interacts with GABAA receptor extracellular N-terminus structures acting though transmembrane domain (TMl), as these structures have been shown to modulate current deactivation (Bianchi et al., 2001). However, detailed molecular studies using GABAA 82 receptor-containing chimeras or exchange mutants between subunits are needed to determine more accurately the potential binding site for MeHg on the GABAA receptor. In summary, acute bath application of MeHg to rat cortical cells in primary culture irreversibly suppressed IGABA and prolonged Tsjow. Co-application of diazepam or flumazenil either prior to or after MeHg exposure did not alter the patterns of effects of MeHg on IGABA. Our results suggest that MeHg and diazepam appear to have independent effects and that they most likely do not compete for the same binding site on the GABAA receptor in cortical cells. Future studies are needed to determine whether or not MeHg interacts with a site on the GABAA receptor belonging to a known drug or if it has a unique binding site on the receptor. 83 CHAPTER THREE DIFFERENTIAL EFFECTS OF METHYLMERCURY ON GABAA RECEPTOR CURRENT S IN RAT CEREBELLAR GRAN ULE AND CEREBRAL CORTICAL NEURONS IN CULTURE 84 ABSTRACT Granule cells of the cerebellum are particularly sensitive to the inhibitory effects of MeHg on GABAA receptor-mediated synaptic transmission. Suppression of IPSCs by MeHg occurs more rapidly in granule cells than in neighboring Purkinje cells. The mechanism(s) that underlies the differential sensitivity of GABAergic transmission to MeHg in different cerebellar cell types is not known. A possible explanation is the differential expression of a6 subunit-containing GABAA receptors in granule and Purkinje neurons. To test this hypothesis, IGABA were recorded in response to MeHg in cerebellar granule and cerebral cortical cells in culture. Cortical cells were used as a replacement for Purkinje cells as they are more successfully cultured and express the same (11- containing GABAA receptor phenotype as do Purkinje cells (Hansen et al., 2001). IGABA were obtained by means of the whole-cell voltage patch-clamp technique, using a symmetrical chloride concentration. MeHg suppressed completely IGABA in both cell types in a time- and concentration-dependent manner. Suppression in granule cells occurred more rapidly than did suppression in cortical cells. To verify these results, IGABA were recorded in response to MeHg in cerebellar granule cells expressing either (16 or a. subunit-containing GABAA receptors. Suppression of IGABA by MeHg was similar in granule cells containing either subtype of GABAA receptor. Similarly, the effects of MeHg on [GABA were examined in HEK-293 cells expressing either (16 or (:1 subunit- containing GABAA receptors. IGABA suppression by MeHg was comparable in both as and a1 subunit-containing HEK-293 cells. These results suggest that the presence of the (16 subunit alone may not underlie the differential effects of MeHg on [GABA observed in cerebellar granule and cortical neurons; other factors may be involved as well. 85 INTRODUCTION Methylmercury (MeHg) is a wide-spread environmental neurotoxicant. It has a high affinity for sulfhydryl groups numerous on cysteine-containing proteins. Thus, MeHg has the potential to bind to cell membrane proteins and interfere with many cellular processes (Chang, 1977; Atchison and Hare, 1994), including disruption of excitatory and inhibitory synaptic transmission (Yuan and Atchison, 1993, 1995, 1997, 1999, 2003). GABAA receptor-mediated synaptic transmission is inhibited by MeHg. In dorsal root ganglion neurons in culture, high concentrations of MeHg (100 uM) suppress the peak currents induced by GABA (Arakawa et al., 1991). In hippocampal slice, MeHg decreases gradually inhibitory postsynaptic potential (IPSP) amplitudes to complete suppression (Yuan and Atchison, 1995). Suppression of inhibitory synaptic transmission by low concentrations of MeHg occurs earlier than does suppression of glutamatergic synaptic transmission (Yuan and Atchison, 1995, 1997), suggesting that inhibitory synaptic transmission is more sensitive to suppression by MeHg than is excitatory transmission. Interestingly, suppression of inhibitory post synaptic currents (IPSCs) induced by bath-applied MeHg in cerebellar granule cells in brain slice occurs much earlier than does suppression in neighboring Purkinje cells (Yuan and Atchison, 2003). The mechanism(s) by which MeHg differentially affects GABAergic synaptic transmission in cerebellar cells is not known. One possibility for this differential sensitivity to MeHg may be the differential expression of GABAA receptor phenotypes in granule and Purkinje cells. Purkinje cells express the ail-containing receptor. Granule cells are the only neurons in the cerebellum which express the (16 subunit-containing GABAA receptor, although they too express the 86 a; subunit. Expression of the (1., subunit is regulated developmentally both in vivo and in vitro and studies have shown that there is an increased expression of a6 subunits in rats during maturation (Laurie et al., 1992, Thompson et al., 1994). GABAA receptors containing the (16 or a] subunit have unique pharmacological and biophysical properties, including differential sensitivity to agonists, such as benzodiazepines (Luddens and Wisden, 1991; Sieghart, 1992; Makela et al., 1997) and barbiturates (Fisher etal., 1997; Cestari et al., 2000) and to antagonists such as, zinc (Draguhn et al., 1990; Zempel and Steinbach, 1995; Saxena and MacDonald, 1994, 1996), furosemide (Korpi et al., 1995a), and lanthanum (Saxena et al., 1997; Makela et al., 1999). However, whether or not the expression of the (16 subunit-containing GABAA receptor contributes to the differential sensitivity of cerebellar neurons to MeHg remains to be determined. The present study was designed specifically to examine the a subunit specificity of the effects of MeHg by using the whole-cell patch-clamp technique to compare the effects of MeHg on IGABA in cells expressing either a. or 0.6 subunit-containing GABAA receptors. 87 METHODS Solutions and Chemicals Methylmercuric chloride (MeHg) (ICN Biomedical Inc., Costa Mesa, CA, USA) was dissolved in deionized water to a final concentration of 10 mM, which served as a stock solution. On the day of experiments, MeHg working solutions (0.1, 1, or 10 uM) were diluted in extracellular solution consisting of (in mM): NaCl, 125; CaCl2, 2.0; KCl, 2.5; MgCl2, 1.0; KH2PO4, 1.25; NaHCOg, 26.0, and D-glucose, 20.0, pH-adjusted to 7.4 using 95% 02/5% CO2. In the present studies, three different concentrations of MeHg (0.1, 1, or 10 11M) were used. These concentrations of MeHg were used as they are within the clinically relevant range of concentrations found in victims poisoned with MeHg in Iraq in the 1970’s (Bakir et al., 1973) (see Chapter One). Diazepam, 7-amino- n-butyric acid (GABA), 6-cyano-7-nitroquinoxaline-2,3-dione disodium salt (CNQX), DL-Z-amino-S-phosphonopentanoic acid (APV), glutamine, 4-(2-hydroxyethyl)piperazine- l-ethanesulfonic acid (HEPES), gentamycin, deoxyribonuclease (DNase I), cytosine B-D- arabino-furanoside (Ara-C), adenosine 5'- triphosphate magnesium salt (ATP), ethylene glycol-bis(B-aminoethyl ether)-N, N, N', N',-tetraacetic acid (EGTA), trypsin, poly-L- lysine, picric acid, and formaldehyde (37% solution) were all purchased from Sigma Chemical Co. (St. Louis, MO, USA). Basal Eagle’s Medium, Modified Eagle’s Medium and Dulbecco’s Modified Eagle’s Medium were purchased from Mediatech, Inc. (Hemdon, VA). Heat inactivated horse serum, fetal bovine serum, sodium pyruvate, non- essential amino acids, antimycotic and antibiotics, and Optimem were purchased from GIBCO-Invitrogen Inc. (Carlsbad, CA). Qiagen kits, used for plasmid purification, were purchased from Qiagen Inc. (Valencia, CA) and Fugene 6 was purchased from Roche 88 Molecular Biochemicals (Indianapolis, IN). Secondary antibodies conjugated with tetramethylrhodamine (TRITC) or fluorescein isothiocyanate (FITC) were purchased from Jackson ImmunoResearch Inc. (West Grove, PA). Secondary antibody conjugated with pacific blue was purchased from Invitrogen (Carlsbad, CA). Diazepam was dissolved in dimethyl sulphoxide (DMSO; < 0.01% (v/v) final concentration; a concentration low enough not to affect 10A“) and diluted in ACSF prior to use. Other drugs were dissolved in deionized water and diluted in extracellular solution prior to use. Preparation of Primary Cerebellar Granule Cell Culture Primary cultures of rat cerebellar granule neurons were prepared from 7-day-old Sprague-Dawley rats of either gender (Charles River Laboratories, Wilmington, MA) following procedures described by Gallo et al. (1987). Following extraction of cerebella, cells were digested for 13 min at 37° C with trypsin 0.025 % (w/v) and plated at a density of l x 106 cells/ml on 35-mm Petri dishes (Corning Inc., Corning, NY) coated with poly- L-lysine (0.1 mg/ml). Cells were cultured in Basal Eagle’s Medium supplemented with 10% (w/v) fetal bovine serum, 2 mM glutamine, and 100 ug/ml gentamycin. The final concentration of KCl in the culture medium was adjusted to 25 mM. In order to achieve functional synapse formation (Chen et al., 2000; Prybylowski et al., 2002), at 4 days in vitro the medium was replaced with 5 mM KCl-containing Modified Eagle’s Medium supplemented with 5 mg/ml glucose, 0.1 mg/ml transferrin, 0.025 mg/ml insulin, 2 mM glutamine, and 20 jig/ml gentamycin, in addition to cytosine arabinoside (10 M) to inhibit glial cell proliferation. Cells were maintained in a 37° C incubator with 95% O2 and 5% CO2. Cells were used for recordings 3-8 days after plating, depending upon the 89 aim of the experiment, to allow for maturation and expression of a6 and Ill-subunit containing GABAA receptors (Laurie et al., 1992a). Preparation of Primary Cerebral Cortical Cell Culture Cortical cells were used as a substitute for cerebellar Purkinje cells, as Purkinje cells are extremely difficult to maintain in culture and produce a low yield. Furthermore, they frequently become contaminated with other cell types, particularly glial cells, and MeHg is known to have inhibitory effects on astrocytes (Aschner et al., 2000). Importantly, cortical cells are easier to maintain in culture and they contain the same type of GABAA receptor phenotype as do Purkinje cells (Hansen et al., 2001). Primary cultures of cortical neurons were obtained from newborn Sprague-Dawley rat pups of either gender (Charles River Laboratories, Wilmington, MA) following methods described by Inglefield et a1. (2001). Briefly, following removal of the brain and isolation of cortex, cells were digested for 4.5 min at 37° C with trypsin 0.025 % (w/v) in buffer that contained (in mM): KCl (5.0), KH2PO4 (0.205), NaCl (137.0), Na2HP04 (0.17), D-glucose (5.0), sucrose (59.0), and gentamycin (0.1 mg/ml), pH 7.3. Trypsin was inactivated by addition of the above-described buffer containing 0.016 % (w/v) DNase I for 4.5 min at 37° C. To this mixture warmed Dulbecco’s Modified Eagle’s Medium (DMEM) supplemented with 10 % (w/v) horse serum, 10 mM HEPES, 2 mM glutamine, and gentamycin (0.1 mg/ml) was added and the cells were centrifuged (500xg, 5 min). The resulting pellet was resuspended in DMEM containing DNase I and was centrifuged again. The cell suspension was then resuspended in DMEM and triturated with a Pasteur pipette. Cells were plated at a density of 1 x 106 cells/ dish onto 35-min Petri dishes (Corning Inc., Corning, NY) coated with poly-L-lysine. Cells were 90 maintained in a 37° C incubator with 95% O2 and 5% CO2. DMEM was replaced every two days. Cytosine B-arabinoside (5 pM) was added to cell cultures 72 hr after plating to inhibit glial cell proliferation. Cells were used for recordings 10-14 days after plating. All experiments were done in compliance with NIH and university standards and were approved by the Michigan State University Committee of Animal Use and Care. Culture and Transfection of HEK-293 Cells As described previously (Peng et al., 2002), HEK 293 cells (#CRL-1573; American Type Culture Collection, Rockville, MD) were lightly trypsinized, centrifuged and plated at ~ 50 - 100,000 cells in 2 ml DMEM fortified with 1 mM sodium puruvate, 0.1 mM non-essential amino acids, 10% (w/v) fetal bovine serum, antibiotics and antimycotics and grown overnight at 37° C in 5% C02. The transient expressions were achieved by transfecting the cells one day after plating with 1 pg total DNA (1:1:1 molar ratio of a, B and 7 and 1/5 of a GFP plasmid, all cloned in pCDNA3.1). Plasmids containing rat or], B2 and 72 subunit cDNA were obtained from Dr. C. Czajkowski (University of Wisconsin, Madison,WI) and plasmids containing rat a6 subunit cDNA were obtained from Dr. W. Wisden (University of Heidelberg, Heidelberg, Germany). All plasmids were purified using Qiagen kits (endotoxin free). Optimem (96 p1) was added together with Fugene 6 (3 pl) and DNA (1 pg) to maintain an anion - cation ratio of 1:1. The mixture was incubated for 20 min. and then added to dishes. Green fluorescent cells were seen starting 24 hrs after treatment, and robust expression was detected approximately 40 - 72 hrs post treatment. On the day of recording, the cells were lightly trypsinized, centrifuged and replated at 1/3 - 1/6 dilution on poly-L-lysine coated 91 cover glasses in 35 mm culture dishes at a low density with good spatial separation to permit recordings. Recordings were made 2 - 8 hrs following plating. lmmunocytochemistry Cover slips containing HEK-293 cells were incubated overnight in ice-cold phosphate buffered saline (PBS) containing 0.02% NaN3 to kill cells immediately without permeabilizing their membranes. Following 2 rinses with PBS, HEK-293 cell were labeled overnight in the cold room with primary anti-a; (1:7000) antibody (Novus Biologicals Inc., Littleton, CO) or anti-a6 (1:1000) antibody (Novus Biologicals Inc., Littleton, CO), both generated in rabbits. In contrast, cover slips containing granule cells from 2-8 days in vitro (DIV) and cortical cells (DIV 10) were labeled overnight with anti- (11 (1:7000) or anti-d6 (1:1000) antibody following a 5 min. fixation in ice-cold acetone and 3 rinses in PBS. All cells were incubated with a tetramethylrhodamine (TRITC; 1:500)-conjugated anti-rabbit or fluorescein isothiocyanate (FITC; 1:500)-conjugated anti-rabbit secondary antibody for 2 hrs at 4°C and staining was visualized using a Leitz Laborlux S (Wetzlar, Germany) epifluorescent microscope and a 20x or 60x oil immersion objective. Negative controls included incubation of cells with primary or secondary antibody only, in addition to labeling cells with antibodies against subunits which they do not express (i.e. labeling cortical cells with anti-06 antibody or labeling HEK-293 cells expressing only (16 subunit-containing GABAA receptors with antibody against the a; subunit). No fluorescence was seen in any of the control experiments, except a minute amount of cross-reactivity was detected with the anti-a1 antibody in HEK-293 cells expressing (16 subunit-containing GABAA receptors. This cross-reactivity 92 was not detected in granule cells. Absorption controls with the peptide and antibodies could not be performed, as the peptides were not provided with the antibodies. To confirm that labeling of anti— a] and anti-a6 antibodies was localized to the cell membrane, cover slips containing HEK-293 cells were incubated for 1 hr in Zamboni fixative containing (in ml): picric acid (15), formaldehyde-37% solution (5.5), and PBS (79.5). Following fixation, cover slips were rinsed 2 times with PBS and labeled overnight in the cold room with primary anti-a1 (1:7000) antibody or anti—a6 (1:1000) antibody, both generated in rabbits, or with primary anti-pan cadherin (1:500) antibody, generated in mice. Cells were incubated with (TRITC; l:200)-conjugated anti-rabbit, (FITC; 1:200)-conjugated anti-rabbit, or pacific blue (1:200)-conjugated anti-mouse secondary antibody for 2 hrs at 4°C. Staining was visualized using a Nikon Eclipse 2000-UDiaphot-TMD microscope (Nikon Optics, Tokyo, Japan) and a 20x or 40x objective, as indicated. Whole-cell Recordings Methods are identical to those described in Chapter Two. Data Acquisition and Analysis Methods are identical to those described in Chapter Two. 93 RESULTS MeHg Suppresses [GABA in Cerebellar Granule Cells To determine the subunit specificity of effects of MeHg on 10AM, whole-cell recordings were made from (16 subunit-containing granule cells and a1 subunit-containing cortical cells. First, to verify the GABAA receptor a subunit expression in both cell types, immunocytochemistry was performed using antibodies against the (11 mm; subunit of the receptor. As expected, both granule and cortical cells labeled positively for the a; subunit of the GABAA receptor (Fig. 3.1 A-B), whereas only granule cells labeled positively for the (16 subunit of the receptor (Fig. 3.1 CD). In both cell types the antibody staining appeared to be localized primarily to the cell body, and to a lesser extent in neuronal processes. The cell body is stained in a punctate pattern and in some cells it was evident that the bulk of the staining is in the membrane periphery, suggesting that staining is primarily localized to the membrane surface. Under symmetrical Cl' concentrations, application of a GABA pulse produced an inward current in granule cells, which was gradually abolished by bath application of MeHg (0.1 — 10 pM) (Fig. 3.2). The effects of MeHg never reached a steady-state level less than complete reduction of IGABA (Fig. 3.2 A). Suppression of [GABA by MeHg (1 pM) was not reversible by washing cells for ~ 20 min with a MeHg-free solution or a solution containing D-penicillamine (20 pM), a MeHg chelator (Appendix, Fig. A. 1). The time course of reduction of IGABA by MeHg is shown in Figure 3.2 B. Increasing the concentration of MeHg shortened the time course of suppression. For each concentration of MeHg tested, the time-course of suppression by MeHg was approximately linear so that 50% IGABA reduction occurred at about 50% of the total time to reduction. Time to 94 Figure 3.1. GABAA receptpr a6 and a; subunit expression in cerebellar granule and cerebral cortical cells in culture. Granule (DIV 6) and cortical cells (DIV 10) were labeled overnight with anti-q] (A-B) or anti-a6 (C-D) antibody following acetone fixation. Antibody staining was visualized with TRITC using a Leitz epifluorescent microscope and a 60x oil immersion objective. Note that only cerebellar granule cells express a6 subunit-containing GABAA receptors (C), whereas both granule and cortical cells express a1 subunit-containing receptors (A -B). Images in this thesis are presented in color. 95 Granule Cortical Figure 3.1 96 Figure 3.2. MeHg causes a time- and concentration-dependent suppression of [GABA in granule cells. A, MeHg caused a progressive and complete suppression of IGABA in rat granule cells in primary culture. Whole-cell [GABA were evoked from a holding potential of - 60 mV by a 10 ms pulse of GABA (50 pM, black dot), at intervals of 30 s, in the presence of CNQX (10 pM) and APV (50 pM) in the external solution to block glutamatergic currents. Data were collected before (control) and at various time points during exposure to 1 pM MeHg (6, 8, 12, and 20 min). B, Time course of effects of MeHg on [GABA recorded under identical conditions to those described above at [MeHg] of 0.1 pM (squares), 1.0 pM (triangles), or 10 pM (circles). Data were collected continually before and after MeHg exposure. Each datum point represents the mean value recorded from 3-5 cells. C, Comparative effects of various concentrations of MeHg on time to complete reduction of IGABA. Each bar represents the mean values obtained from 3-5 cells. The times to total reduction of IGABA by MeHg differed significantly between 0.1 and 1.0 pM as indicated by an asterisks (*), and 0.1 and 10 pM MeHg, as indicated by a dagger (1'). No significant difference was observed between 1.0 and 10 pM MeHg (p>0.05). 97 A 20 min —> GABA 0 12min —-> 8min —> 6 min 400 DA 2000 ms Control (0 min) —> B |GABA (°/o Control) Time (min) Figure 3.2 98 10 ass [MeHg] (MM) L O O O O 6 4 2 Figure 3.2 C 2:5 cozonoom ES 9 9:: 99 Figure 3.3. MeHg causes a time- and concentration-dependent suppression of [GABA in cortical cells. A, MeHg caused a progressive and complete suppression of IGABA in rat cortical cells in primary culture. Whole-cell [GABA were recorded under identical conditions to those described in Figure 3.2. Data were collected before (control) and at various time points during exposure to 1 pM MeHg (5, 8, 11, 18, and 37 min). B, Time course of effects of MeHg on IGABA recorded at [MeHg] of 0.1 pM (open squares), 1.0 pM (filled triangles), or 10 pM (open circles). Data were collected continually before and after MeHg exposure. Each datum point represents the mean value recorded from 3- 5 cells. C, Comparative effects of different concentrations of MeHg on time to complete reduction of IGABA. Each bar represents the mean (:1: SEM) values obtained from 3-5 cells. The times to total reduction of IGABA by MeHg differed significantly between 0.1 and 1.0 pM as indicated by an asterisks (*), and 0.1 and 10 pM MeHg, as indicated by a dagger (1'). No significant difference was observed between 1.0 and 10 pM MeHg (p>0.05). D, Comparative effects of MeHg on suppression of [GABA in cortical (dark bars) and granule cells (light bars) in culture. IGABA was suppressed significantly faster (p<0.05) in granule cells as compared with cortical cells for 1.0 and 10 pM MeHg, as indicated by the asterisks. 100 GABA 37 min 18 min 11 min 8 min 5 min Control —> 1000 m5 (0 min) |GABA (% Control) Time (min) Figure 3.3 101 _ _ 0 0 0 0 4 2 C 3:: cozosoom ES 9 95 [MeHg] (11M) l:l Granule W Cortical D 60‘ 3:5 cozonoom ES 2 9:: [MeHg] (MM) Figure 3.3 102 total reduction of IGABA by MeHg differed significantly between each concentration of MeHg tested (p<0.05) (Fig. 3.2 C). At 10 pM MeHg, the highest concentration tested, complete reduction of [GABA occurred within approximately 17 min. MeHg Suppresses [GABA in Cerebral Cortical Cells Similar to the effects of MeHg observed in cerebellar granule cells, bath application of MeHg (0.1 - 10 pM) produced a gradual and complete suppression of IGABA in cortical cells (Fig. 3.3 A). The time course of reduction of IGABA by MeHg in cortical cells is shown in Figure 3.3 B. Decreasing the concentration of MeHg prolonged the time course of suppression. For each concentration of MeHg tested, the time-course of reduction by MeHg was approximately linear. Time to total reduction of [GABA by MeHg differed significantly between each concentration of MeHg tested (p<0.05) (Fig. 3.3 C). At 10 pM MeHg, the highest concentration tested, complete reduction of IGABA occurred within approximately 30 min. Interestingly, at 1 and 10 pM MeHg, the time to suppression of [GABA was significantly less in granule cells as compared with cortical cells (Fig. 3.3 D), however this difference was not observed at 0.1 pM MeHg. In addition, stepping the voltage from ~80 mV to +60 mV in the presence of MeHg produced a linear current-voltage relationship in granule cells (Fig. 3.4 A) as well as in cortical cells (Fig. 3.4 B), suggesting that suppression of IGABA by MeHg is voltage- independent in both cell types. 103 Figure 3.4. Effects of MeHg on IGABA are voltage-independent in granule and cortical cells. A, Whole-cell IGABA were recorded from cerebellar granule cells before (control) and after 5 min. exposure to 1 pM MeHg at holding potentials ranging from -80 to +60 mV. Each datum point represents the mean value recorded from 3-4 cells. MeHg caused a reduction in IGABA amplitude in a linear fashion, such that it did not change at different holding potentials. B, IGABA were recorded from cerebral cortical cells before (control) and after 10 min. exposure to 1 pM MeHg at holding potentials ranging from - 80 to +60 mV. Each datum point represents the mean value recorded from 3 cells. MeHg caused a reduction in [GABA amplitude in cortical cells in a voltage-independent manner. 104 .s U" 1 Control A O r ” MeHg I I I l O I 01 I ...L C I _'. 01 If -80 -60 -40 -20 0 20 40 60 Voltage (mV) Normalized Current (pA) > OD O 20 . Control Normalized Current (pA) w -80 so -40 -2o 0 20 40 so Voltage (mV) Figure 3.4 105 Figure 3.5. GABAA receptor an and (:1 subunit expression in cerebellar granule cells at different days in culture. Granule cells (DIV 2-8) were labeled overnight with anti- 0.; or anti-(16 antibody following acetone fixation. Antibody staining was visualized with TRITC (anti-a1) or FITC (anti-06) using a Leitz epifluorescent microscope and a 60x oil immersion objective. A, (11 subunit-containing GABAA receptors are expressed in granule cells grown in vitro throughout days 2-8 (DIV 2-8). B, 016 subunit-containing receptors are not expressed in granule cells until DIV 6 and 8. 106 DIV 2 DIV 4 1 1 pm 38 pm DIV 6 DIV 8 38 pm 15 pm Figure 3.5 107 B Anti-a6 DIV 2 DIV 4 38 pm 38 pm DIV 6 DIV 8 Figure 3.5 108 m ‘ Figure 3.6. Effects of MeHg on granule cells at different days in culture. Time course of effects of MeHg on IGABA were obtained under identical conditions to those described in Figure 3.1. IGABA were recorded from granule cells grown in culture for 4 or 6-8 days in the presence of [MeHg] of 0.1 pM (A), 1.0 pM (B), or 10 pM (C). Data were collected continually before and after MeHg exposure. Each datum point represents the mean value recorded from 3-4 cells. D, Comparative effects of different concentrations of MeHg on time to complete suppression of IGABA in granule cells grown for 4 or 6-8 days in culture. Each bar represents the mean (:1: SEM) values obtained from 3-4 cells. The times to total reduction of IGABA by MeHg did not differ significantly between cells grown for 4 or 6-8 days in culture. 109 > 100 ” IGABA Reduction (% Control) CD 0 0.1 pM MeHg —l— DIV4 + DIV 6-8 I l 40 — 20 " 0 0 2 100 — 9 E _ o 80 O 23 so - c .9 *5 40 z '8 a: 20 < 2 _o 0 Figure 3.6 8 16 24 32 40 Time (min) 1.0 pM MeHg ——I— DIV4 + DIV 6-8 6 12 18 24 Time (min) 110 10 14M MeHg C + DIV 6-3 _ _,_ DIV4 000000 00000 99:50 at cozozumm <96. n\\\\\\\\\\\\\\\\\\\km 000000 55555 D Eé 838.com ES 2 9:: 6 am e m .015 F 111 The Effects of MeHg Do Not Differ in Granule Cells Expressing Only a; or a Combination of a1 and a6 Subunit-Containing GABAA Receptors Following cerebellar granule cell culture, cells were fixed and labeled with antibodies against GABAA receptor on or (16 subunits at different days in culture to determine the time-course of a subunit expression. As seen in Figure 3.5 A, expression of on subunits was observed on the first day of antibody labeling (day 2 in culture) and continued through to the last day of labeling (day 8 in culture). In contrast, expression of a6 subunits was not observed until day 6 in culture, and expression continued through to day 8 in culture (Fig. 3.5 B). The staining for either antibody appeared to be primarily localized to the cell body in a punctate pattern. In some cells, the focal plane shows that much of the staining is localized to the periphery of the cell body, suggesting cell surface labeling. Staining was also observed, to a lesser degree, in neuronal processes. GABAA receptor-mediated currents were examined from cells from day 4 in culture to represent til-containing GABAA receptor responses, and from day 6-8 in culture to represent (16- containing responses. At each day in culture tested, MeHg produced a gradual and complete reduction of IGABA (Fig. 3.6). This suppression was concentration- and time- dependent. The time course of suppression by MeHg did not differ between days 4 and 6-8 in culture as seen in Figures 3.6 A-D. For 10 pM MeHg, the highest concentration tested, the average time to suppression of IGABA in granule cells from day 4 in culture was 20.8 :1: 2 min, and for cells from 6-8 days in culture it was 21.3 :1: 3 min. 112 MeHg Suppresses [GABA in HEK-293 Cells Expressing (11 or (16 Subunit-Containing GABAA Receptors HEK-293 cells were used in these experiments, as these nonexcitable cells are commonly used for heterologous expression of membrane proteins including GABAA receptors (Saxena, 2000; Bianchi and MacDonald, 2002; Hinkle and MacDonald, 2003; J ones-Davis et al., 2005). First, to verify that HEK-293 cells expressed the appropriate subunit subtype of interest, cells were treated and labeled with antibody against an or (16 subunit—containing GABAA receptors as seen in Figure 3.7. As expected, only HEK-293 cells transfected with cDN A for the a; subunit-containing GABAA receptor labeled positive for the a] subunit and not the (16 subunit (Fig. 3.7 A-B). Moreover, HEK-293 cells transfected with cDN A for the (16 subunit-containing GABAA receptor labeled positive for the (16 subunit and not the (:1 subunit (Fig. 3.7 CD). A slight amount of fluorescence to anti-d1 was detected for the (16 subunit-containing cells (Fig. 3.7 C), suggesting that non-specific binding or a small amount of cross-reactivity for the anti-a1 antibody occurred, perhaps with auxiliary subunits such as [32 or 72. The pattern of staining of HEK-293 cells for both antibodies appeared punctate and primarily localized to the periphery of the cell body, suggesting membrane surface staining. To confirm membrane surface staining, co-localization experiments were conducted in HEK-293 cells with a primary anti—pan cadherin (membrane protein) antibody and a primary anti-a6 or anti-a1 antibody (Fig. 3.8 — 3.10). As shown in Figure 3.8, HEK-293 cells transfected with mRNA for (xi—containing GABAA receptors in addition to green fluorescence protein (GFP), stained positive for GFP (Fig. 3.8 B). Note that the transfection was efficient in approximately 25 % of cells. Most cells, both transfected and untransfected, stained 113 Figure 3.7. GABAA receptor «6 and a; subunit expression in HEK-293 cells. HEK- 293 cells were labeled overnight with anti-a1 or anti-r16 antibody following killing by NaN3, to prevent membrane perrneabilization. Antibody staining was visualized with TRITC using a Leica epifluorescent microscope and a 60x oil immersion objective. HEK-293 cells expressing (11 subunit-containing GABAA receptors labeled positive for the (:1 subunit (A) and negative for the (:6 subunit (B). HEK-293 cells expressing a6 subunit-containing GABAA receptors labeled negative for the on subunit (C) and positive for the (16 subunit (D). Note, a slight amount of fluorescence was detected in C, suggesting that a small amount of cross-reactivity or non-selective binding for the anti-u] antibody may exist. 114 Anti-0t6 Figure 3.7 115 Figure 3.8. GABAA receptor membrane localization in (11 subunit-containing HEK- 293 cells. HEK-293 cells were labeled overnight with anti-a1 or anti-pan cadherin antibody following fixation. Antibody staining was visualized with TRITC or pacific blue using a Nikon epifluorescent microscope and a 20x or 40x (as indicated) objective. HEK-293 cells expressing a1 subunit-containing GABAA receptors labeled positive for GFP protein (A) in approximately 25% of cells (B). All cells labeled positive for the membrane protein cadherin (pacific blue) (C). A lesser number of cells stained positive for the on subunit (TRITC) (D). Co-localization of cadherin and the al subunit- containing GABAA receptor in these cells is evident at 20x (E) and (40x) (F) magnification. 116 GFP GFP + Brightfield Cadherin 0‘1 30 pm Cadherin + a1 Cadherin + a1 (40 x) Figure 3.8 117 Figure 3.9. GABAA receptor membrane localization in an subunit-containing HEK- 293 cells. HEK-293 cells were labeled overnight with anti-a6 or anti-pan caherin antibody following fixation. Antibody staining was visualized with TRITC or pacific blue using a Nikon epifluorescent microscope and a 20x or 40x (as indicated) objective. HEK-293 cells expressing r16 subunit-containing GABAA receptors labeled positive for GFP protein (A) in approximately 25% of cells (B). All cells labeled positive for the membrane protein cadherin (pacific blue) (C). A lesser number of cells stained positive for the (16 subunit (TRITC) (D). Co-localization of cadherin and the cur, subunit-containing GABAA receptor in cells is evident at 20x (E) and (40x) (F) magnification. 118 HEK-oi6 .- GFP + Brightfield C Cadherin Cadherin + a6 Cadherin + a6 (40 X) Figure 3.9 119 Figure 3.10. Control Secondary antibody staining in HEK-293 cells. HEK-293 cells were labeled overnight with PBS not containing primary antibody following fixation. Secondary antibody staining (TRITC or pacific blue) was visualized using a Nikon epifluorescent microscope and a 20x objective. HEK-293 cells expressing a1 subunit- containing GABAA receptors labeled positive for GFP protein in approximately 30% of cells (A). Secondary antibody staining was also detected lightly in a few transfected cells (pacific blue) (B). HEK-293 cells expressing a; subunit-containing GABAA receptors labeled positive for GFP protein in approximately 25 % of cells (C). Slight secondary antibody staining was detected in a few transfected cells (TRITC) (D). 120 Controls GFP + Brightfield Pacific blue only C GFP + Brightfield TRITC only Figure 3.10 121 Figure 3.11. Effects of diazepam on [GABA in HEK-293 cells. Whole-cell [GABA were evoked from a holding potential of - 60 mV by a 10 ms pulse of GABA (500 pM, black dot), at intervals of 30 5, under similar recording conditions to those described in Figure 1. Current traces were collected before (control) and after exposure to diazepam (1 pM) at times indicated by arrows. A, Diazepam prolonged IGABA slow decay rate in HEK-293 cells expressing (11 subunit-containing GABAA receptors as can be seen readily by the superimposed control and diazepam traces (right). B, Diazepam did not affect IGABA slow decay rate in HEK-293 cells expressing a6 subunit-containing GABAA receptors. 122 I- A (in-containing Control 605 Diazepam Overlay N O 0 Control ti: 0 r Normalized Current (pA) .L ~80 ~60 ~40 ~20 0 Voltage (mV) W N O I O is o Normalized Current (pA) 6: ti) ~80 ~60 ~40 ~20 0 Voltage (mV) Figure 3.13 132 20 40 Control 60 20 40 DISCUSSION The primary aim of these studies was to determine if the differential expression of a6 subunit-containing GABAA receptors in cerebellar granule and Purkinje neurons underlies the differential sensitivities to MeHg observed in these cells. To test this, the whole-cell patch-clamp technique was used to record GABAA receptor-mediated currents in the presence or absence of MeHg in cerebellar granule (cg-containing) and cerebral cortical (oi-containing) cells in culture and in HEK-293 cells transfected with cDNA for either a1~ or a6containing GABAA receptors. MeHg (0.1 — 10 pM) suppressed IGABA in a time- and concentration-dependent manner in each cell type investigated. Suppression of IGABA occurred more rapidly in granule neurons as compared to cortical neurons. However, time—to-suppression of [GABA by MeHg did not differ in granule neurons expressing either a6~containing or Ill-containing GABAA receptors. Furthermore, effects of MeHg on IGABA were examined in HEK-293 cells transfected with cDNA for the a6- or Oil-containing GABAA receptor. The effects of MeHg on IGABA did not differ in HEK- 293 cells expressing either GABAA receptor subtype. These results suggest that the presence of the (16 subunit alone may not be the underlying reason for the differential effects of MeHg on GABAergic currents observed in cerebellar granule and Purkinje neurons; other factors be contribute as well. Previous studies conducted in slice preparations and primary cultures of rat neurons have revealed that MeHg impairs inhibitory GABAergic neurons in hippocampus and cerebellum (Yuan and Atchison, 1995, 1997, 2003; Xu and Atchison, 1997). GABAergic synaptic transmission was shown to be more sensitive to the effects of MeHg as compared with gluamatergic synaptic transmission (Yuan and Atchison, 133 2003). Importantly, some cells types appear to be particularly sensitive to the effects of MeHg. Suppression of inhibitory post synaptic currents (IPSCs) in cerebellar granule cells, for instance, occurs much earlier than does suppression in neighboring Purkinje cells (Yuan and Atchison, 2003). Thus, GABAA receptors in the cerebellum appear to be particularly sensitive to the inhibitory effects of MeHg. The exact mechanism(s) by which MeHg differentially interferes with GABAergic function in different cell types are not known. One possibility for the enhanced sensitivity of GABAergic responses to MeHg in granule cells, as compared with Purkinje cells, is their differential expression of the orig-subunit containing GABAA receptor. GABAA receptors containing the (16 subunit, as compared with non-d6 containing ones, are more sensitive to inhibition by other heavy metals such Zn+2 (Draguhn et al., 1990; Saxena and MacDonald, 1994, Zempel and Steinbach, 1995) and La’“3 (Saxena et al., 1997, Makela et al., 1999) and MeHg has been shown to increase concentrations of intracellular Zn’2 (Denny and Atchison, 1994). To test this possibility, IGABA were recorded in response to MeHg and compared in granule cells and in cortical cells (used as a replacement for Purkinje cells). IGABA were more sensitive to the effects of MeHg in granule cells than in cortical cells, as IGABA were inhibited more rapidly in granule cells than in cortical cells. These findings are qualitatively consistent with effects of MeHg on IGABA reported previously in cerebellar slice, but occurred at much lower concentrations of MeHg (Yuan and Atchison, 1999, 2003). The results of these experiments support the possibility that the (L6 subunit may contribute to the differential effects of MeHg on IGABA observed in granule and Purkinje cells. 134 To confirm the results obtained in granule and cortical cells, the effects of MeHg on IGABA were investigated in granule cells only at different postnatal days as these cells can developmentally express either (16- and/or (ll—containing GABAA receptors. The effects of MeHg on IGABA did not differ in granule neurons expressing either Go- containing or (ii-containing GABAA receptors. [GABA were suppressed by MeHg in a time- and concentration-dependent manner in granule cells containing either subtype of GABAA receptor, suggesting that the (16 subunit alone does not underlie the differential sensitivity of [GABA to MeHg in granule and cortical cells. Other factors, such as differential GABAA receptor auxiliary subunit expression in these two cell types, may be involved. In addition to the on and (:6 subunits, cerebellar granule cells express GABAA receptors containing either [32 or [13 subunits in combination with 72, 73, or 6 subunits (Laurie et al., 1992). Cortical cells (and Purkinje cells), on the other hand, express mainly the (1613272 subtype of GABAA receptor (Wisden et al., 1996). Variability in GABAA receptor subunit composition can alter the responsiveness of the receptor to drugs. For instance, the 06-6 subtype of receptor has an increased affinity for GABA and is extremely sensitive to inhibition by Zn+2 (Saxena and MacDonald, 1994). Thus, Zn+2 may indirectly contribute to inhibition of the (1.5-6 subunit-containing receptor. In addition, sensitivity of the receptor to agents such as barbiturates has been shown to be dependent of the identity of the [3 subunit, as the [33 subunit makes the GABAA receptor insensitive to the effects of pentobarbital (Cestari et al., 2000). Furosemide inhibition is enhanced when the [31 subunit of the GABAA receptor is replaced with either the B2 or [33 subunit subtypes (Thompson et al., 1999). Similarly, replacement of the [3, subunit of the 135 receptor with the B2 or [33 subunit, results in enhanced sensitivity to potentiation by ethanol (Mihic et al., 1997) or etomidate, a general anesthetic (Belelli et al., 1997). To examine the effects of MeHg on (16 or (xi-containing receptor-mediated currents in isolation, whole-cell recordings were also conducted in HEK-293 cells containing either (16 or (11 subtype of GABAA receptor. Results showed that effects of MeHg on [GABA suppression did not differ between HEK-293 cells containing either subtype of GABAA receptor. Time to suppression of long». by MeHg (1 and 10 pM) occurred at similar times in Go or (xi-containing HEK-293 cells. These findings suggest that differential expression of a6 subunit-containing GABAA receptors in granule and Purkinje cells does not underlie the differential sensitivities of IGABA to MeHg observed in granule and Purkinje cells. Furthermore, as both subtypes of GABAA receptor in HEK-293 cells contained the same [32 and 72 subunits, it suggests that sensitivity to MeHg by these subunit subtypes does not differ. An alternate explanation for the lack of difference in sensitivity to MeHg observed between 016- and ail-containing HEK-293 cells, is the specific sequence of chNA used. Plasmids (pRK5) containing GABAA-a6 cDN A (Luddens et al., 1990) were used for HEK-cell transfections in these studies. When cDNA was released from its vector using digestion, several bands were detected, including a 1.3 kb band, which was determined to be the correct (116-containing cDNA insert (Hajela, personal communication). Expression of the (116-containing GABAA receptor was verified immunocytochemically and pharmacologically, by evaluating the responsiveness to diazepam. However, after completion of the experiments described in this paper, it was observed that the same GABAA-a6 cDNA injected into Xenopus oocytes, produced GABAergic responses that did not appropriately respond to the Cl‘ 136 channel blocker, niflumic acid. Instead of being inhibited by niflumic acid, GABAergic responses in oocytes expressing the 016- subunit were potentiated. Therefore, the (16- containing cDN A insert sequence used in these studies may not have been spliced at precisely the right site, or the receptor isoform simply responded differently when expressed in oocytes as compared with in HEK-293 cells. In sum, acute bath application of MeHg to rat cerebellar granule and cerebral cortical cells completely and irreversibly suppressed 10AM. IGABA in granule cells were more sensitive to the effects of MeHg than were IGABA in cortical cells. However, IGABA in granule cells expressing 0.5-containing GABAA receptors were approximately equally as sensitive to the effects of MeHg as [GABA in (xi-containing cells. Moreover, suppression of [GABA by MeHg was similar in HEK-293 cells expressing either 0.6- or al- containing receptors. These results suggest that the expression of the (16 subunit alone does not underlie the differential effects of MeHg on IGABA observed in cerebellar granule and Purkinje neurons; additional factors may be involved as well. Further chimeric studies are needed to determine the relevant contribution, if any, of each GABA receptor subunit and its subtypes, to the differential sensitivity of GABAAergic responses to MeHg in cerebellar cells. 137 CHAPTER FOUR SUMMARY and DISCUSSION 138 A. Summary of Research The studies described in this thesis were aimed at characterizing the interactions of MeHg with the GABAA receptor and at determining if the differential expression of (16- containing GABAA receptors in cerebellar granule and Purkinje neurons underlies the differential sensitivities to MeHg observed in these cells. This is due to the observations that MeHg inhibits GABAergic synaptic transmission, and does so more rapidly in cerebellar granule than in Purkinje cells (Yuan and Atchison, 1999, 2003). A possible explanation for this differential effect of MeHg on [GABA may be the differential expression of a6 subunit-containing GABAA receptor in these cell types. To address this hypothesis, 1 performed whole-cell voltage-patch clamp experiments in cerebellar granule and cerebral cortical cells in primary culture and in HEK-293 cells expressing specific GABAA receptor subunits of interest. I showed that GABAA receptor-mediated currents in cerebral cortical cells in culture are completely and irreversibly suppressed by MeHg (0.1, 1.0, and 10 pM) in a time- and concentration-dependent manner. Because an effect less than complete suppression of current was not seen, the variable measured was time to suppression. This is a general response of cells to MeHg (Arakawa et al., 1991; Yuan and Atchison, 1993, 1995, 1997, 1999, 2003; Xu et al., 1997). The magnitude of effects produced by different concentrations of MeHg did not differ. At each concentration of MeHg tested (0.1, 1.0, and 10 pM) the time-course of suppression of [GABA was approximately linear and complete suppression was always attained (if recordings could be maintained long enough). This is consistent with the effects of MeHg reported on GABAergic synaptic transmission in granule and Purkinje cells in the cerebellar slice preparation (Yukun and Atchison, 2003). In cerebellar slice, the magnitude of changes in 139 frequency or amplitude of postsynaptic currents did not change, irrespective of the concentration of MeHg used. The fixed magnitude of effects caused by different concentrations of MeHg suggests that a consistent series of actions takes place once an effective or threshold concentration of MeHg is attained. I also showed that MeHg affects the kinetics of IGABA by prolonging the slow decay rate. These effects were concentration-independent, in that at each concentration of MeHg tested, 15.0,. was enhanced to approximately the same extent. GABAA receptor current decay is important as it is a major determinant of IPSC duration (Twyman et al, 1990; MacDonald and Twyman, 1992; Jones and Westbrook, 1995). The faster component of decay is associated with receptor desensitization, whereas the slower component of decay is typically associated with GABA receptor deactivation (MacDonald et al., 1989a; Twyman et al., 1990; Twyman and MacDonald, 1992). Deactivation rate is influenced by unbinding of GABA from its receptor as well as by transitions of activated receptors between open, closed, and desensitized states. Even though removal of agonist is considered the final step in current deactivation, it can be prolonged by detaining the receptor in the desensitized state (Jones and Westbrook, 1995; Tia et al., 1996; Haas and MacDonald, 1999). Unbinding is prolonged from the desensitized state because the receptor needs first to recover from the desensitized state to reopen prior to unbinding of GABA. Thus, GABAA receptor deactivation can be shaped by several different mechanisms. In the present study, the prolongation of Tslow by MeHg likely indicates that MeHg slows the rate of unbinding of GABA from its receptor. Alternatively, it could indicate that MeHg produces an accumulation of desensitized GABAA receptors, resulting in prolonged unbinding of GABA from its receptor. To 140 investigate the effects of MeHg receptor desensitization, further studies are needed, such as ones investigating the effects of MeHg on the recovery of the GABAA receptor from desensitization using paired GABA pulses applied in the absence and in the presence of MeHg. Additional studies at the single channel level would also be useful in examining the effects of MeHg on GABAA receptor kinetics in more detail. These experiments should be performed with the use of a fast perfusion system, as a relatively slow perfusion system, such as the one used to perform experiments in this thesis, may not appropriately resolve the fast phases of decay, or desensitization. For the experiments described in this dissertation, physiological saline solution was delivered via gravity into the extracellular solution at a rate of approximately 0.45 mllmin and the 10~90% rise time for solution exchange was ~ 15 8. Furthermore, I have shown that the effects of MeHg are voltage-independent, suggesting that MeHg does not interact with the pore of the GABAA receptor. 1 have shown that the effects of MeHg on IGABA are GABA concenUation-independent, suggesting that MeHg does not compete with GABA for its binding site on the receptor. Moreover, MeHg prolonged IGABA decay and completely suppressed IGABA in the presence of diazepam (10 pM) or the benzodiazepine antagonist, flumazenil (20 pM), suggesting that MeHg does not compete with diazepam for its binding site on the GABAA receptor. Thus, MeHg may interact with a unique site on the receptor or one belonging to another known drug. It is also possible that MeHg may have non-specific effects on the GABAA receptor. To address the second goal of these studies, I examined and compared the effects of MeHg on IGABA in cerebellar granule (rm-containing) and cortical cells (al- 141 containing) in culture. Suppression of IGABA occurred more rapidly in granule neurons as compared to cortical neurons, suggesting that (1.5-containing granule cells are more sensitive to the effects of MeHg. To verify these findings in granule cells, which can express either subtype of GABAA receptor, I initially attempted to distinguish the two GABAA receptor subtypes by pharmacological means. By using the benzodiazepine, diazepam, I predicted that I would observe differences in IGABA responsiveness between cells expressing 016- and non- (1.5-containing receptors. However, in almost all cells tested, whole-cell IGABA were similarly potentiated by diazepam. Thus, it was determined that it would be extremely difficult to differentiate the receptor subtypes pharmacologically in granule cells, as they contain GABAA receptors that have a wide variety of subunit combinations. Importantly, these subunit arrangements can consist of a1, (16, or a mixture of a; and (16, in addition to other subunit variations, producing receptors with a multitude of different pharmacological profiles. Consequently, immunocytochemistry was used to identify a1- and (1.5-containing granule cells. As GABAA receptor subtypes are expressed developmentally, granule cells were grown for different days in culture and their responsiveness to MeHg was compared at DIV 4 (a1 expressing) and at DIV 6—8 (a; and a6 expressing). Subunit subtype expression was determined using antibodies against each a subunit of interest. Time-to-suppression of IGABA by MeHg did not differ in granule neurons expressing either dig-containing or III-containing GABAA receptors. To confirm the results obtained in native cerebellar granule neurons, effects of MeHg on IGABA were examined in HEK-293 cells transfected with cDNA for the a6~ or (Ii-containing GABAA receptor. The effects of MeHg on IGABA did not differ in HEK-293 cells expressing either GABAA receptor subtype. The results obtained in cerebellar granule and cortical neurons 142 and those reported in HEK-293 cells suggest that the presence of the (16 subunit alone may not be the underlying reason for the differential effects of MeHg on GABAergic currents observed in cerebellar granule and Purkinje neurons. . Relationship to Previous Work In hippocampal slice, GABAergic inhibitory synaptic transmission is more sensitive to suppression by MeHg than is excitatory transmission, because suppression of inhibitory synaptic transmission by MeHg occurrs before suppression of glutamatergic synaptic transmission (Yuan and Atchison, 1995, 1997). These studies also showed that MeHg first increased excitatory postsynaptic potential (EPSP) amplitudes prior to their suppression. This early enhancement of EPSP amplitude prior to inhibition was shown to occur as a result of MeHg suppressing GABAA receptor-mediated inhibitory synaptic transmission prior to suppressing excitatory transmission (Yuan and Atchison, 1997). In the cerebellum, GABAergic synaptic transmission is also more sensitive to the effects of MeHg as compared with glutamatergic synaptic transmission (Yuan and Atchison, 2003). Importantly, these effects of MeHg on GABAergic synaptic transmission in cerebellum were cell type preferential. IPSCs were more sensitive to suppression by MeHg in granule cells than in Purkinje cells (Yuan and Atchison, 2003). Similarly in this thesis, in primary neurons in culture, MeHg suppressed IGABA more rapidly in cerebellar granule cells (org-containing) than in cerebral cortical cells (cl-containing). In both cell types, MeHg suppressed IGABA completely and irreversibly in a time- and concentration- dependent manner. Granule cells were more sensitive to the effects of MeHg than were cortical cells, as [GABA were inhibited more rapidly in granule cells than in cortical cells. 143 These findings are qualitatively consistent with effects of MeHg on IGABA reported previously in cerebellar slice, but occurred at much lower concentrations of MeHg (Yuan and Atchison, 1999, 2003). Differences in concentrations of MeHg needed to produce inhibition in cerebellar slice as compared with cells in culture may be due to the difference in thickness of the preparations (Meacham et al., 2005). The slice preparation is much thicker, resulting in a limited surface area available for accumulation of MeHg. To reach its target cells, MeHg must diffuse through the slice, encountering a greater number of sites of interaction for MeHg. However, for cells grown in culture, most cells have a majority of their surface area directly exposed to MeHg in the bathing solution, thus a much lower concentration is needed to produce effects. These findings suggest that the differential expression of the (16 subunit may contribute to the differential sensitivities of IGABA to MeHg in granule and cortical or Purkinje cells. However, additional data in this thesis do not support this hypothesis. IGABA in cerebellar granule cells expressing (1.5-subunit containing receptors were equally as sensitive to inhibition by MeHg (0.1, 1.0, and 10 pM) as were IGABA in granule cells not expressing dis-containing receptors. It is important to keep in mind that cells expressing (1.5-containing receptors also express Ill-containing ones, and a1 and (:6 subunits are often co-expressed in the same receptor. Thus, complete isolation of (16 responses was most likely not achieved. In the HEK-293 cell expression system, responses mediated solely by (11 or (1.5-containing GABAA receptors can be recorded. Findings from this thesis revealed that IGABA in HEK-293 cells were abolished completely and over a similar time- course by MeHg (0.1, 1.0, and 10 pM) in both 016- or ail-containing cells. Effects of MeHg on IGABA were voltage-independent in cells containing either receptor phenotype. 144 These results suggest that the presence simply of the dis-containing GABAA receptor may not underlie the differential effects of MeHg on IGABA observed in cerebellar granule and Purkinje neurons, but that other factors are likely involved. One contributing factor may be the differential expression of auxiliary GABAA receptors subunits in these cells. Other factors may include the alteration of intracellular concentrations of Ca+2 by MeHg. This is important because within the cerebellum, granule cells express the highest density of cholinergic muscarinic (M3) receptors as compared with other cell types (Neustadt et al., 1988). These receptors are involved in synaptic transmission between granule and Purkinje cells, in that acetylcholine increases the glutamate-mediated excitation of Purkinje cells (Takayasu et al., 2003). M3 receptors play an important role in increasing [Can] from intracellular stores, as they activate 1P3 (inositol triphosphate) receptors resulting in release of Ca+2 into the cytoplasm from the endoplasmic reticulum (Kandel et al., 2000), Notably, M3 receptors have been shown to be upregulated by MeHg (Coccini et al., 2000), and in cerebellar granule cells in culture, a decrease in M3 receptors protects against neurotoxicity by MeHg (Limke et al., 2004). Another contributing factor may include the alterations of intracellular concentrations of Zn+2 by MeHg. This is important because Zn+2 can be released from glutamate-containing vesicles (Assaf and Chung, 1984; Howell et al., 1984). Granule cell axon terminals that from synapses with Purkinje cells have dis-containing GABAA receptors, which are extremely sensitive to inhibition by Zni2 (Gallo et al., 1981). Thus, Zn+2 my be co-released with glutamate from granule cell axons onto Purkinje cells, and may feed back onto dis-containing GABAA receptors in granule cells and inhibit them. 145 A small number of studies have found that administration of MeHg, both short term and long term, enhances benzodiazepine binding in rodent brain via interaction with the GABAA receptor (Corda et al., 1981; Concas et al., 1983; Fonfria et al., 2001). However, Komulainen and Tuomisto (1985) showed that both acute as well as chronic exposure to MeHg does not affect the number of benzodiazepine receptors labeled with [3H]flunitrazepam in rat cerebellum. To determine if MeHg interacts with the benzodiazepine site of the GABAA receptor, studies in this thesis were performed to examine the effects of MeHg on IGABA in the presence or absence of diazepam. In the presence of diazepam, the effects of MeHg on [GABA were maintained, even following pretreatment with diazapem. The actions of MeHg on IGABA were also unaffected by treatment with the benzodiazepine antagonist flumazenil. These results suggest that MeHg interacts at a site on the GABAA receptor different from that used by diazepam. They are consistent with those reported by Komulainen and Tuomisto (1985). However, these results do not necessarily contradict those described by others (Corda et al., 1981; Concas et al., 1983; Fonfria et al., 2001). This is because increased binding of diazepam to the GABAA receptor after MeHg exposure, does not necessarily indicate that MeHg alters the function of the receptor in response to diazepam. Instead, increased binding of diazepam in response to MeHg exposure could occur as result of the cell trying to compensate for the antagonistic affects of MeHg by increasing GABAA receptor benzodiazepine binding sites. In addition, exposure to MeHg could alter GABAA receptor subunit expression, perhaps increasing expression of benzodiazepine sensitive subunits, such as on. 146 C. Possible Mechanisms Underlying the Differential Effects of MeHg on Cerebellar Granule and Purkinje Neurons The circuitry of the cerebellum may play a role in the differential effects of MeHg-induced neurotoxicity observed in granule and Purkinje cells. Excitatory mossy fibers release glutamate on granule cell dendrites, which results in the propagation of action potentials down granule cell axons, or parallel fibers. These axons form excitatory synapses with Purkinje cells and Golgi cells. Upon Golgi cell activation by granule cell parallel fibers, Golgi cells release GABA onto the synapse between granule cells and mossy fibers, forming an inhibitory feed-back loop. MeHg is known to increase Ca"2 concentrations within cells (Denny et al., 1993; Hare et a1, 1993; Denny and Atchison, 1996; Marty and Atchison, 1997; Limke and Atchison, 2002) which may result in enhanced spontaneous GABA release from Golgi, stellate, and basket cells, and increased glutamate release from granule cell parallel fibers. Because the membrane excitability of neurons is highly dependent on the combined activity from both excitatory and inhibitory inputs, disruption of GABAergic synaptic transmission in cells can be damaging. Loss of inhibition can cause an imbalance in excitability to occur, resulting in cell overexcitability. Thus, because GABAergic transmission is inhibited preferentially in granule cells by MeHg (Yuan and Atchison, 2003), it may be possible that disruption of the granule cell-Golgi cell inhibitory feedback loop could occur. As a result, granule cells would have diminished abilities to respond to the inhibitory input from Golgi cells, and in turn become overexcited by glutamatergic input from mossy fibers. In addition, MeHg could suppress dig-containing GABAA receptor-mediated responses, resulting of impairment of tonic inhibition mediated by this receptor subtype. GABAA receptors 147 containing (16 subunits can also contain 6 subunits and are localized at the extrasynaptic membrane of granule cells, typically at synapses with Golgi cell axons (Nusser et al., 1996; Nusser et al., 1998). This is important because one of the main contributors in the regulation of granule cell excitability is the tonic inhibitory Cl' conductance. The Golgi cell granule cells synapse is specialized to promote transmitter spillover, as this synapse is within a glomerulus (Hamori and Zsentagothai, 1966; Jakab and Hamori, 1988). Glomeruli are specialized synapses that become ensheathed by glial cells, forcing neighboring dendrites to share a limited space, thus facilitating the build-up of neurotransmitter (Fig. 4.1). Each glomerulus contains glutamatergic mossy fiber terminals surrounded by numerous granule cell dendrites intermingled with Golgi cell axon terminals that form GABAergic synapses with granule cells. GABA released at Golgi to granule cell synapses is confined by the glial sheath and is removed by uptake into Golgi terminals and by surrounding glia (Itouji et al., 1996). As a result, any GABA that escapes from the synaptic cleft is likely to spill over (Rossi et al., 1998) onto both postsynaptic and extrasynaptic sites on granule cell dendrites, as both types of synaptic sites are expressed within the glomerulus (Nusser et al., 1995; 1996). Increased Ca+2 concentrations caused by MeHg may initially contribute to enhanced GABA release from Golgi cells and increased GABA spillover. Extrasynaptically located (16 (and 6)~ containing receptors are likely to be activated by low spillover concentrations of GABA, 148 Figure 4.1. Drawing showing a cerebellar glomerulus. This specialized synapse is ensheathed by glial cells and contains glutamatergic mossy fiber terminals surrounded by numerous granule cell dendrites. It also contains Golgi cell axon terminals that form GABAergic synapses with granule cells. Figure 4.1, borrowed from Kandel et al., 2000. 149 Glomerulus Mossy fiber terminal Golgi cell axon Granule cell dendrite Mossy fiber Figure 4.1. 150 as these receptors have a high affinity for GABA and do not readily desensitize (Saxena et al., 1994; Tia et al., 1996b). Spillover GABA acting at (1.5-containing receptors results in persistent opening of GABAA receptor channels producing a tonic form of inhibition (Nusser et al., 1998). Phasic inhibition produced by direct activation of granule cells by Golgi cells is thought to produce more synchronous rhythmic firing, and tonic inhibition is thought to reduce synchronicity and rhythmicin of firing patterns of granule and Golgi cells (De Schutter and Maex 1996; De Schutter, 2002; Maex and De Schutter, 2003). MeHg antagonizes both (16 and (xi-containing GABAA receptors, which could produce a reduction in tonic and phasic inhibition, resulting in heightened levels of excitability of granule cells (Fig. 4.2.). Increased intracellular Ca+2 concentrations caused by MeHg may further contribute to increased glutamate release from granule cell parallel fibers. Enhanced glutamate could also occur in parallel with increased Zn+2release, as both are contained and released from glutamatergic synaptic vesicles (Assaf and Chung, 1984; Howell et al., 1984; Xie and Smart, 1991). As a result, Zn+2 can inhibit a66~containing GABAA receptors in granule cells, which are extremely sensitive to its effects, contributing to the suppression of tonic inhibition. However, (rm-containing receptors in Purkinje cells, which have a reduced sensitivity to Zn”, are less likely to be affected. The excessive glutamatergic output from granule cell parallel fibers onto Purkinje neurons could be detrimental in regulation of cerebellar output, as the mossy fiber-granule cell- Purkinje cell pathway through the cerebellum has been shown to be the operational path, as it correlates with and controls moment-to-moment behavior (Kandel et al., 2000). Under normal conditions, Purkinje cells are excited by glutamatergic input from granule cell axons and climbing fiber terminals, and are inhibited by GABAergic input from 151 Figure 4.2. Diagram depicting the proposed mechanism of interaction of MeHg with GABAA receptors of granule cells to alter cerebellar excitation. A. Climbing fibers (CF) release glutamate and excite Purkinje cells (PC). Mossy fibers (MF) release glutamate and excite both granule cells (GC) and Golgi cells (GOL). GCs release glutamate and and excite GOL and PC. Zn+2 may also be co-released with glutamate, and it may feedback to inhibit extrasynaptic GABAA receptors. GOL release GABA and form an inhibitory feedback loop with GC. GCs express both do and on containing GABAA receptors, whereas PC express only (xi-containing ones. PC express only AMPA-type glutamate receptors, whereas GCs and GOL express both AMPA and NMDA subtypes of +211 and receptor. M3 receptors are abundant in GCs and play a role in increasing [Ca enhancement of glutamate release. Nearby astroctyes (A) contain glutamate transporters that remove glutamate from the synaptic space. PC provide the only output from the cerebellar cortex; it is inhibitory. For simplicity, basket and stellate intemeurons have been excluded. B. MeHg increases [Can], and release of glutamate and GABA, and inhibits both (16 and a1-containing GABAA receptors as well as astrocyte glutamate transporters. As a result, GCs are unable to respond to GOL inhibition and become overexcited. GCs, in turn, overexcite PC, thus enhancing the inhibitory output of the cerebellar cortex. 152 engage . 0.05). 169 500 100 10 wwwmmo EEC 558.35 2 met. GABA (11M) Figure A.2. 170 BIBLIOGRAPHY Adkins CE, Pillai GV, Kerby J, Bonnert TP, Haldon C, McKeman RM, Gonzalez JE, Oades K, Whiting PJ and Simpson PB (2001) Alpha4beta3delta GABA(A) receptors characterized by fluorescence resonance energy transfer-derived measurements of membrane potential. J Biol Chem 276(42):38934~9. Akabas MH, Kaufmann C, Archdeacon P and Karlin A (1994) Identification of acetylcholine receptor channel-lining residues in the entire M2 segment of the alpha subunit. Neuron l3(4):919~27. Akabas MH, Kaufmann C, Cook TA and Archdeacon P (1994) Amino acid residues lining the chloride channel of the cystic fibrosis transmembrane conductance regulator. J Biol Chem 269(21): 14865-8. Akk G and Steinbach JH (2000) Activation and block of recombinant GABA(A) receptors by pentobarbitone: a single-channel study. Br J Pharmacol 130(2):249~58. Amin J, Dickerson IM, and Weiss DS (1994) The agonist binding site of the gamma- aminobutyric acid type A channel is not formed by the extracellular cysteine loop. Mol Pharmacol 45(2):3 17-23. Amin J and Weiss DS (1993) GABA(A) receptor needs two homologous domains of the beta-subunit for activation by GABA but not by pentobarbital. Nature 366(6455):565~9. Amin-Zaki L, Elhassani S, Majeed MA, Clarkson TW, Doherty RA and Greenwood M (1974) Intra-uterine methylmercury poisoning in Iraq. Pediatrics 54(5):587~95. Arakawa O, Nakahiro M and Narahashi T (1991) Mercury modulation of GABA- activated chloride channels and non-specific cation channels in rat dorsal root ganglion neurons. Brain Res 551(1-2):58-63. Aschner M and Allen JW (2000) Astrocytes in methylmercury, ammonia, methionine sulfoximine and alcohol-induced neurotoxicity. Neurotoxicology 21(4):573~9. 171 Aschner M, Yao CP, Allen JW and Tan KH (2000) Methylmercury alters glutamate transport in astrocytes. Neurochem Int 37(2~3):199~206. Assaf SY and Chung SH (1984) Release of endogenous Zn2+ from brain tissue during activity. Nature 308(5961):734~6. Atchison WD (1986) Extracellular calcium-dependent and ~independent effects of methylmercury on spontaneous and potassium-evoked release of acetylcholine at the neuromuscular junction. J Pharmacol Expt Ther 237 :672-80. Atchison WD and Hare MF (1994) Mechanisms of methylmercury-induced neurotoxicity. FASEB J 8(9):622-29. Atchison WD and Narahashi T (1982) Methylmercury-induced depression of neuromuscular transmission in the rat. Neurotoxicology 3:37-50. Axtell CD, Cox C, Myers GJ, Davidson PW, Choi AL, Cernichiari E, Sloane-Reeves J, Shamlaye CF and Clarkson TW (2000) Association between methylmercury exposure from fish consumption and child development at five and a half years of age in the Seychelles Child Development Study: an evaluation of nonlinear relationships. Environ Res 84(2):71~80. Bakir F, Damluji S, Amin-Zaki L, Murtadha M, Khalidi A, Al-Rawi NY, TikritiS, Dhahir HI, Clarkson TW, Smith JC and Doherty RA (1973) Methylmercury poisoning in Iraq. Science 181:230~40. Barberis A, Cherubini E and Mozrzyma SJW(2000) Zinc inhibits miniature GABAergic currents by allosteric modulation of GABA(A) receptor gating. J Neurosci 20:8619-27. Barker JL, Behar T, Li YX, Liu QY, Ma W, Maric D, Marie 1, Schaffner AE, Serafini R, Smith SV, Somogyi R, Vautrin JY, Wen XL and Xian H (1998) GABAergic cells and signals in CNS development. Perspect Dev Neurobiol S(2-3):305~22. Barker JL, Harrison NL, Lange GD and Owen DG (1987) Potentiation of gamma- aminobutyric-acid-activated chloride conductance by a steroid anaesthetic in cultured rat spinal neurones. J Physiol (Land) 386:485-501. 172 Bastian AJ and Thach WT (1995) Cerebellar outflow lesions: a comparison of movement deficits resulting from lesions at the levels of the cerebellum and thalamus. Ann Neurol 38(6):881~92. Basu N, Stamler CJ, Loua KM, and Chan HM (2005) An interspecies comparison of mercury inhibition on muscarinic acetylcholine receptor binding in the cerebral cortex and cerebellum. Toxicol Appl Pharmacol 205:71-6. Baumann SW, Baur R and Sigel E (2001) Subunit arrangement of gamma-aminobutyric acid type A receptors. J Biol Chem 276(39):36275~80. Belelli D, Lambert J], Peters JA, Wafford K and Whiting PJ (1997) The interaction of the general anesthetic etomidate with the gamma-aminobutyric acid type A receptor is influenced by a single amino acid. Proc Natl Acad Sci U S A 94(20):] 1031-6. Bellantuono C, Reggi V, Tognoni G and Garattini S (1980) Benzodiazepines: clinical pharmacology and therapeutic use. Drugs 19(3): 195-219. Ben-Ari Y (2002) Excitatory actions of GABA during development: the nature of the nurture. Nat Rev Neurosci 3(9):728~39. Bera AK, Chatav M and Akabas MH (2002) GABA(A) receptor M2~M3 loop secondary structure and changes in accessibility during channel gating. J Biol Chem 277(45):43002-10. Bianchi MT and MacDonald RL (2001) Mutation of the 9' leucine in the GABA(A) receptor gamma2L subunit produces an apparent decrease in desensitization by stabilizing open states without altering desensitized states. Neuropharmacology 41:737- 44. Bianchi MT, Haas KF and MacDonald RL (2001) Structural determinants of fast desensitization and desensitization-deactivation coupling in GABA(A) receptors. J Neurosci 15: 1 127-36. Bianchi MT, MacDonald RL (2002) Slow phases of GABA(A) receptor desensitization structural determinants and possible relevance for synaptic function. J Physiol (Lond) 544(1):3-18. 173 Bloom FE and Iversen LL (1971) Localizing 3H-GABA in nerve terminals of rat cerebral cortex by electron microscopic autoradiography. Nature 229(5287):628-30. Blount P and Merlie JP (1990) Mutational analysis of muscle nicotinic acetylcholine receptor subunit assembly. J Cell Biol 111(6 Pt 1):2613-22. Boileau AJ, Evers AR, Davis AF and Czajkowski C (1999) Mapping the agonist binding site of the GABA(A) receptor: evidence for a beta-strand. J Neurosci l9(12):4847-54. Brandon NJ, J ovanovic JN, Smart TG and Moss SJ (2002) Receptor for activated C kinase-1 facilitates protein kinase C-dependent phosphorylation and functional modulation of GABA(A) receptors with the activation of G-protein-coupled receptors J Neurosci 22(15):6353-61. Brickley SG, Cull-Candy SG and Farrant M (1996) Development of a tonic form of synaptic inhibition in rat cerebellar granule cells resulting from persistant activation of GABA(A) receptors. J Physiol (Land) 497 :753-759. Callachan H, Cottrel] GA, Hather NY, Lambert J], N ooney JM and Peters JA (1987) Modulation of the GABA(A) receptor by progesterone metabolites. Proc R Soc Lond B Biol Sci 231(1264):359-69. Castoldi AF, Coccini T, Ceccatelli S and Manzo L (2001) Neurotoxicity and molecular effects of methylmercury. Brain Res Bull 55:197—203. Cestari IN, Min KT, Kulli JC and Yang J (2000) Identification of an amino acid defining the distinct properties of murine beta] and beta3 subunit-containing GABA(A) receptors. J Neurochem 74(2):827-38. Chang LW (1977) Neurotoxic effects of mercury. Environ Res l4(3):329-73. Chang LW, Desnoyers PA and Hartmann HA (1972) Quantitative cytochemica] studies of RNA in experimental mercury poisoning 1. Changes in RNA content. J Neuropathol Exper Neurol 31:489-501. 174 Chang Y, Wang R, Barot S and Weiss DS (1996) Stoichiometry of a recombinant GABA(A) receptor. J Neurosci 16:5415-24. Choi BH (1989) The effects of methylmercury on the developing brain. Prog Neurobiol 32(6):447-70. Choi BH, Lapham LW, 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):7l9~33. Claisse D, Cossa D, Bretaudeau-Sanjuan J, Touchard G and Bombled B (2001) Methylmercury in molluscs along the French coast. Mar Pollut Bull 42(4):329-32. Clarkson TW (1993) Mercury major issues in environmental health. Environ Health Perspect 100:31-8. Coccini T, Randine G, Candura SM, Nappi RE, Prockop LD 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. Concas A, Corda MG, Salis M, Mulas ML, Milia A, Corongiu PP and Biggio G (1983) Biochemical changes in the rat cerebellar cortex elicited by chronic treatment with methyl mercury. T oxicol Lett 18:27-33. Corda MG, Concas A, Rossetti Z, Guarneri P, Corongiu F P and Biggio G (1981) Methyl mercury enhances [3H]diazepam binding in different areas of the rat brain. Brain Res 229:264-9. Cordier S, Gare] M, Mandereau L, Morce] H, Doineau P Gosme-Seguret S, Josse D, White R and Amiel-Tison C (2002) Neurodevelopmental investigations among methylmercury-exposed children in French Guiana. Environ Res 89(1): 1-1 1. Costa E and Guidotti A (1979) Recent studies on the mechanism whereby benzodiazepines facilitate GABA-ergic transmission. Adv Exp Med Biol 123:371-8. 175 Costa LG, Aschner M, Vitalone A, Syversen T and Soldin OP (2004) Developmental neuropathology of environmental agents. Annu Rev Pharmacol T oxicol 44:87-110. Crump KS, Kjellstrom T, Shipp AM, Silvers A and Stewart A (1998) Influence of prenatal mercury exposure upon scholastic and psychological test performance benchmark analysis of a New Zealand cohort. Risk Anal l8(6):701~13. Dalziel JE, Cox GB, Gage PW and Bimir B (2000) Mutating the highly conserved second membrane-spanning region 9' leucine residue in the alpha(1) or beta(1) subunit produces subunit-specific changes in the function of human alpha(1)beta(1) gamma-aminobutyric Acid(A) receptors. Mol Pharmacol 57(5):875-82. Davidson P W, Myers G J, Cox C, Shamlaye C F, Marsh DO, Tanner MA, Berlin M, Sloane-Reeves J, Cemichiari E, Choisy O, Choi A, and Clarkson T W (1995) Longitudinal neurodevelopmental study of Seychellois children following in utero exposure to methylmercury from matema] fish ingestion outcomes at 19 and 29 months. Neurotoxicology 16:677-88. Davidson P W, Myers G J, Cox C, Axtell C, Shamlaye C, Sloane-Reeves J, Cemichiari E, Needham L, Choi A, Wang Y, Berlin M, and Clarkson T W (1998) Effects of prenatal and postnatal methylmercury exposure from fish consumption on neurodevelopment: outcomes at 66 months of age in the Seychelles Child Development Study. JAMA 280:701-7. De Schutter E (2002) Cerebellar cortex: computation by extrasynaptic inhibition? Curr Biol 12(10):R363~5. De Schutter E and Maex R (1996) The cerebellum: cortical processing and theory. Curr Opin Neurobiol 6(6):759-64. DeLorey TM and Olsen RW (1992) Gamma-aminobutyric acidA receptor structure and function. J Biol Chem 267(24): 16747-50. Denny MF and Atchison WD (1994) Methylmercury-induced elevations in intrasynaptosoma] zinc concentrations an 19F-NMR study. J Neurochem 63:383-6. 176 Denny MF and Atchison WD (1996) Mercurial-induced alterations in neuronal divalent cation homeostasis. Neurotoxicology l7(l):47-61. Denny MF, Hare MF and Atchison WD (1993) Methylmercury alters intrasynaptosomal concentrations of endogenous polyvalent cations. Toxicol Appl Pharmacol 122:222-32. Didier M, Xu M, Berrnan SA, Saido TC and Bursztajn S (1997) Involvement of three glutamate receptor epsilon subunits in the formation of N-methyl-D-aspartate receptors mediating excitotoxicity in primary cultures of mouse cerebellar granule cells. Neuroscience 78(4):1 129-46. Draguhn A, Verdoom TA, Ewert M, Seeburg PH and Sakmann B (1990) Functional and molecular distinction between recombinant rat GABA(A) receptor subtypes by Zn2+- Neuron 5:781-88. Edwards JR, Marty MS and Atchison WD (2005) Comparative sensitivity of rat cerebellar neurons to dysregulation of divalent cation homeostasis and cytotoxicity caused by methylmercury. Toxicol Appl Pharmacol 208:222-32. EPA (1997) Mercury report to Congress. Washington, DC: US Environmental Protection Agency, Office of Air Quality Standards. EPA (2000) The mercury research strategy. Washington, DC: US Environmental Protection Agency, Office of Research and Development. Eto K and Takeuchi T (1978) A pathological study of prolonged cases of Minamata Disease. With particular reference to 83 autopsy cases. Acta Pathol Jpn 28(4):565-84. Fisher JL and MacDonald RL (1998) The role of an a subtype M2-M3 his in regulating inhibition of GABA(A) receptor current by zinc and other divalent cations. J Neurosci 18:2944-53. Fisher JL, Zhang J and MacDonald RL (1997) the role of alphal and alpha6 subtype amino-terminal domains in allosteric regulation of ~aminobutyric acida receptors. Mol Pharmacol 52:714-24. 177 Fonfria E, Rodriguez-Farre E and Sunol C (2001) Mercury interaction with the GABA(A) receptor modulates the benzodiazepine binding site in primary cultures of mouse cerebellar granule cells. Neuropharmacology 41:819-33. Forsyth DS Casey V, Dabeka RW and McKenzie A (2004) Methylmercury levels in predatory fish species marketed in Canada. Food Addit Contam 21(9):849-56. Gallo V, Kingsbury A, Balazs R and Jorgensen OS (1987) The role of depolarization in the survival and differentiation of cerebellar granule cells in culture. J Neurosci 7(7):2203~ l 3. Gallo V, Levi G, Raiteri M and Coletti A (1981) Enhancement by GABA of glutamate depolarization-induced release from cerebellar nerve endings. Brain Res 205(2):431-5. Gao B and Fritschy J-M (1995) Cerebellar granule cells in vitro recapitulate the in vivo pattern of GABA(A)-receptor subunit expression. Develop Brain Res 88: 1-16. Gerstenberger SL and Dellinger J A (2002) PCBs, mercury, and organochlorine concentrations in lake trout, walleye, and Whitefish from selected tribal fisheries in the Upper Great Lakes region. Environ Toxicol 17:513-9. Gilbertson M (2004) Male cerebral palsy hospitalization as a potential indicator of neurological effects of methylmercury exposure in Great Lakes communities. Environ Res 95:375-84. Grandjean P, Weihe P, White RF and Debes F (1998) Cognitive performance of children prenatally exposed to "safe" levels of methylmercury. Environ Res 77(2): 165-72. Grandjean P, Weihe P, White RF, Debes F, Araki S, Yokoyama K, Murata K, Sorensen N, Dahl R and Jorgensen PJ (1997) Cognitive deficit in 7-year-old children with prenatal exposure to methylmercury. Neurotoxicol Teratol l9(6):417-28. Grandjean P, White RF, Nielsen A, Cleary D and de Oliveira Santos EC (1999) Methylmercury neurotoxicity in Amazonian children downstream from gold mining. Environ Health Perspect 107(7):587-91. 178 Grosman C, Salamone FN, Sine SM and Auerbach A (2000) The extracellular linker of muscle acetylcholine receptor channels is a gating control element. J Gen Physiol 116(3):327-40. Haas KF and MacDonald RL (1999) GABA(A) receptor subunit gamma2 and delta subtypes confer unique kinetic properties on recombinant GABA(A) receptor currents in mouse fibroblasts. J Physiol (Land) 514:27-45. Hamill OP, Borrnann J and Sakmann B (1983) Activation of multiple-conductance state chloride channels in spinal neurones by glycine and GABA. Nature 305(5937):805-8. Hamill OP, Marty A, Neher E, Sakmann B and Sigworth FJ (1981) Improved patch- clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Eur J Physiol 391(2):85-100. Hamori J and Szentagothai J (1966) Participation of Golgi neuron processes in the cerebellar glomeruli: an electron microscope study. Exp Brain Res 2(1):35-48. Hanchar HJ, Dodson PD, Olsen RW, Otis TS and Wallner M (2005) Alcohol-induced motor impairment caused by increased extrasynaptic GABA(A) receptor activity. Nature Neurosci 8:339-45. Hansen SL, Ebert B, Fjalland B and Kristiansen U (2001) Effects of GABA(A) receptor partial agonists in primary cultures of cerebellar granule neurons and cerebral cortical neurons reflect different receptor subunit compositions. Br J Pharmacol l33(4):539-49. Hare MF and Atchison WD (1995) Methylmercury mobilizes Ca2+ from intracellular stores sensitive to inositol 1.4.5-trisphosphate in NG108-15 cells. J Pharmacol Exp Ther 272: 1016-23. Hare MF and Atchison WD (1995) Nifedipine and tetrodotoxin delay the onset of methylmercury-induced increase in [Ca2+]i in NG108-15 cells. T oxicol Appl Pharmacol 135:299-307. Hare MF, McGinnis KM and Atchison WD (1993) Methylmercury increases intracellular concentrations of Ca2+ and heavy metals in NG108-15 cells. J Pharmacol Exp Ther 266:1626-35. 179 Herden CJ, Yuan Y and Atchison WD (2004) Effects of methylmercury (MeHg) on GABA(A) receptor-mediated currents in rat cortical cells in culture. Toxicologist 78: 1 122. Hevers W and Lttddens H (1998) The diversity of GABA(A) receptors. Pharmacological and electrophysiological properties of GABA(A) channel subtypes. Molec Neurobiol 18:35-86. Hewett SJ and Atchison WD (1992) Effects of charge and lipophilicity on mercurial- induced reduction of 45Ca2+ uptake in isolated nerve terminals of the rat. Toxicol Appl Pharmacol 113(2):267-73. Hosli E and Hosli L (1995) Autoradiographic studies on the uptake of 3H-noradrenaline and 3H-serotonin by neurones and astrocytes in explant and primary cultures of rat CNS: effects of antidepressants. Int J Dev Neurosci 13(8):897-908. Howell GA, Welch MG and Frederickson CJ (1984) Stimulation-induced uptake and release of zinc in hippocampal slices. Nature 308(5961):736-8. Huang CS and Narahashi T (1996) Mercury chloride modulation of the GABA(A) receptor-channel complex in rat dorsal root ganglion neurons. Toxicol Appl Pharmacol 140:508-20. Huang CS and Narahashi T (1997) The role of G proteins in the activity and mercury modulation of GABA-induced currents in rat neurons. Neurapharmacalagy 36(11- 12): 1623-30. Huang CS and N arahashi T (1997) The role of phosphorylation in the activity and mercury modulation of GABA-induced currents in rat neurons. Neurapharmacalagy 3601-12): 1631-40. Hunter D and Russell D (1945) Focal cerebral and cerebellar atrophy in a human subject due to organic mercury compounds. Neurol Neurosurg Psychiat 17:235-41. Hunter D, Bomford RR and Russell DS (1940) Poisoning by methyl mercury compounds. Q J Med 35: 193-218. 180 Irn WB, Binder JA, Dillon GH, Pregenzer JF, Im HK Altman RA (1995) Acceleration of GABA-dependent desensitization by mutating threonine 266 to alanine of the alpha 6 subunit of rat GABA(A) receptors. Neurosci Lett l86(2-3):203-7. Inglefield JR, Mundy WR and Shafer, TJ (2001) Inositol 1,4,5-triphosphate receptor- sensitive Ca2+ release, store-operated Ca2+ entry, and CAMP responsive element binding protein phosphorylation in developing cortical cells following exposure to polychlorinated biphenyls. J Pharmacol Exp Ther 297 :762-73. Inoue M, Oomura, Y, Yakushiji, T, and Akaike, N (1986) Intracellular calcium ions decrease the affinity of the GABA receptor. Nature (324): 156-8. Itouji A, Sakai N, Tanaka C and Saito N (1996) Neuronal and glial localization of two GABA transporters (GATl and GAT3) in the rat cerebellum. Brain Res Mol Brain Res 37(1-2):309-16. Jackel C, Kleinz R, Makela R, Hevers W, Jezequel S, Korpi ER and Luddens H (1998) The main determinant of furosemide inhibition on GABA(A) receptors is located close to the first transmembrane domain. Eur J Pharmacol 357(2-3):251-6. Jakab RL and Hamori J (1988) Quantitative morphology and synaptology of cerebellar glomeruli in the rat. Anat Embryol (Berl) 179(1):81-8. Jechlinger M, Pelz R, Tretter V, Klausberber T and Sieghart W (1998) Subunit composition and quantitative importance of heterooligomeric receptors GABA(A) receptors containing (16 subunits. J Neurosci 18:2449-57. Jones A, Korpi ER, McKeman RM, Pelz R, Nusser Z, Makela R, Mellor JR, Pollard S, Bahn S, Stephenson FA, Randall AD, Sieghart W, Somogyi P, Smith AJH and Wisden W (1997) Ligand-gated ion channel subunit partnerships GABA(A) receptor (16 subunit gene inactivation inhibits 6 subunit expression. J Neurosci 17:1350-62. Jones MV and Westbrook GL (1995) Desensitized states prolong GABA(A) channel responses to brief agonist pulses. Neuron 15:181-91. Jones-Davis DM and MacDonald RL (2003) GABA(A) receptor function and pharmacology in epilepsy and status epilepticus. Curr Opin Pharmacol 3(1): 12-8. 181 Kandel E, Schwartz J, and Jessell, T (2000) Principles of Neural Science 4‘h ed. McGraw-Hill, New York. Kardos J (1999) Recent advances in GABA research. Neurochem Int 34(5):353-8. Karlin A and Akabas MH (1995) Toward a structural basis for the function of nicotinic acetylcholine receptors and their cousins. Neuron 15(6): 1231-44. Kash TL, Jenkins A, Kelley JC, Trudell JR and Harrison NL (2003) Coupling of agonist binding to channel gating in the GABA(A) receptor. Nature 421(6920):272~5. Kehrig HA, Malm O, Akagi H, Guimaraes JR and Torres J P (1998) Methylmercury in fish and hair samples from the Balbina Feservoir Brazilian Amazon. Environ Res 77:84- 90. Keinanen K, Wisden W, Sommer B, Werner P, Herb A, Verdoom TA, Sakmann B and Seeburg PH (1990) A family of AMPA-selective glutamate receptors. Science 249(4968):556-60. Khan ZU, Gutierrez A and De Blas AL (1994) The subunit composition of a GABA(A)/benzodiazepine receptor from rat cerebellum. J Neurochem 63:371-4. Khan ZU, Gutierrez A and De Blas AL (1996) The (11 and a6 subunits can coexist in the same cerebellar GABA(A) receptor maintaining their individual benzodiazepine-binding specificities. J Neurochem 66:685-91. Kimelberg HK, Pang S, and Treble DH (1989) Excitatory amino acid-stimulated uptake of 22Na“ in primary astrocyte cultures. J Neurosci 9(4):1 141-9. Knoflach F, Benke D, Wang Y, Scheurer L, Luddens H, Hamilton B], Carter DB, Mohler H and Benson JA (1996) Pharmacological modulation of the diazepam-insensitive recombinant gamma-aminobutyric acidA receptors alpha 4 beta 2 gamma 2 and alpha 6 beta 2 gamma 2. Mol Pharmacol 50(5): 1253-61. 182 Kolbaev SN, Sharonova IN, Vorobjev VS and Skrebitsky VG (2002) Mechanisms of GABA(A) receptor blockade by millimolar concentrations of furosemide in isolated rat Purkinje cells. Neurapharmacalogy 42:913-21. Komulainen H and Tuomisto J (1985) 3H-dopamine uptake and 3H-haloperidol binding in striatum after administration of methyl mercury to rats. Arch Toxicol 57:268-71. Korpi ER and Luddens H (1993) Regional gamma-aminobutyric acid sensitivity of t- butylbicyclophosphoro[35S]thionate binding depends on gamma-aminobutyric acid A receptor alpha subunit. Mol Pharmacol 44:87-92. Korpi ER, Kuner T, Seeburg PH and Luddens H (1995) Selective antagonist for the cerebellar granule cell-specific 7-aminobutyric acid Type A receptor Mal Pharmacol 47 :283-9. Korpi ER, Kleingoor C, Kettenmann H and Seeburg PH (1993) Benzodiazepine-induced motor impairment linked to point mutation in cerebellar GABA(A) receptor. Nature 361(6410):356-9. Kostial K and Landeka M (1975) The action of mercury ions on the release of acetylcholine from presynaptic nerve endings. Experientia 31(7):834-5. Landrigan PJ, Schechter CB, Lipton JM, Fahs MC and Schwartz J (2002) Environmental pollutants and disease in American children: estimates of morbidity, mortality, and costs for lead poisoning, asthma, cancer, and developmental disabilities. Environ Health Perspect 110(7):721-8. Laurie DJ, Seeburg PH and Wisden W (1992) The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. 11. Olfactory bulb and cerebellum. J Neurosci 12(3): 1063-76. Laurie DJ, Wisden W and Seeburg PH (1992) The distribution of thirteen GABA(A) receptor subunit mRN As in the rat brain. III. Embryonic and postnatal development. J Neurosci 12(11):4151-72. 183 Le Novere N, Corringer PJ and Changeux JP (2002) The diversity of subunit composition in nAChRs evolutionary origins physiologic and pharmacologic consequences. J Neurobiol 53(4):447-56. Levenberg KA (1944) Method for the solution of certain problems in least squares. Quart Appl Math 2:164-8. Levesque PC and Atchison WD (1991) Disruption of brain mitochondria] calcium sequestration by methylmercury. J Pharmacol Exp Ther 256(1):236-42. Leyshon-Sorland K, Jasani B, Morgan AJ (1994) The localization of mercury and metallothionein in the cerebellum of rats experimentally exposed to methylmercury. Histochem J 26(2): 161-9. Limke TL and Atchison WD (2002) Acute exposure to methylmercury opens the mitochondrial permeability transition pore in rat cerebellar granule cells. Toxicol Appl Pharmacol l78(1):52~6l. Limke TL, Bearss JJ and Atchison WD (2004) Acute exposure to methylmercury causes Ca2+ dysregulation and neuronal death in rat cerebellar granule cells through an M3 muscarinic receptor-linked pathway. Toxicol Sci 80(1):60-8. Luddens H and Wisden W (1991) Function and pharmacology of multiple GABA(A) receptor subunits. Trends Pharmacol Sci l2(2):49-51. Luddens H, Pritchett DB, Kohler M, Killisch I, Keinanen K, Monyer H, Sprengel R and Seeburg PH (1990) Cerebellar GABA(A) receptor selective for a behavioural alcohol antagonist. Nature 346(6285):648-51. Ma J Y and Narahashi T (1993) Enhancement of gamma-aminobutyric acid-activated chloride channel currents by lanthanides in rat dorsal root ganglion neurons. Neurosci 13:4872-9. MacDonald RL and Barker JL (1978) Different actions of anticonvulsant and anesthetic barbiturates revealed by use of cultured mammalian neurons. Science 200(4343):775-7. 184 MacDonald RL and Olsen RW (1994) GABA(A) receptor channels. Annu Rev Neurosci 17:569-602. MacDonald RL and Twyman RE (1992) Kinetic properties and regulation of GABA(A) receptor channels. Ion Channels 3:315-43. MacDonald RL, Rogers CJ and Twyman RE (1989) Kinetic properties of the GABA(A) receptor main conductance state of mouse spinal cord neurones in cell culture. J Physiol (Land ) 410:479-99. MacDonald RL, Rogers CJ and Twyman RE (1989) Barbiturate regulation of kinetic properties of the GABA(A) receptor channel of mouse spinal neurones in culture. J Physiol (Land) 417:483-500. Maconochie DJ, Zempel JM and Steinbach JH (1994) How quickly can GABA(A) receptors open? Neuron 12(1):61~71. Maex R and De Schutter E (2003) Resonant synchronization in heterogeneous networks of inhibitory neurons. J Neurosci 23(33): 10503-14. Mahaffey KR, Clickner RP, Bodurow CC (2004) Blood organic mercury and dietary mercury intake: National Health and Nutrition Examination Survey, 1999 and 2000. Environ Health Perspect 112(5):562-70. Magos L (1975) Letter; Variation of biological half-life of methylmercury in man. Arch Environ Health 30(7):371. Makela R, Uusi-Okari M, Homanics GE, Quinlan JJ, Fierstone LL, Wisden W and Korpi ER (1997) Cerebellar 7-aminobutyric acid type A receptors pharmacological subtypes revealed by mutant mouse lines. Mal Pharmacol 53:380-8. Makela R, Uusi-Oukari M, Oja SS, Alho H, Anghelescu I, Klawe C, Luddens H and Korpi ER (1999) Furosemide action on cerebellar GABA(A) receptors in alcohol- sensitive ANT rats. Alcohol 19(3): 197-205. Makela R, Wisden W and Korpi ER (1999) Loreclezole and La3+ differentiate cerebellar granule cell GABA(A) receptor subtypes. Eur J Pharmacol 367(1): 101-5. 185 Manalis RS and Cooper GP (1975) Evoked transmitter release increased by inorganic mercury at frog neuromuscular junction. Nature 257(5528):690~1. Marquardt D (1963) An algorithm for least-squares estimation of nonlinear parameters. SIAM J Appl Math 11:431-41. Marsh DO, Clarkson TW, Myers GJ, Davidson PW, Cox C, Cemichiari E, Tanner MA, Lednar W, Shamlaye C, Choisy O, Horeau C and Berlin M (1995) The Seychelles study of fetal methylmercury exposure and child development. Neurotoxicalagy 16:583-96. Marsh DO, Turner MD, Smith JC, Allen P and Richdale N (1995) Fetal methylmercury study in a peruvian fish-eating population. Neurataxicalagy 16:717-26. Marty MS and Atchison WD (1997) Pathways mediating Ca2+ entry in rat cerebellar granule cells following in vitro exposure to methyl mercury. Toxicol Appl Pharmacol 147:319-30. Marty MS and Atchison WD (1998) Elevations of intracellular Ca2+ as a probable contributor to decreased viability in cerebellar granule cells following acute exposure to methylmercury. Toxicol Appl Pharmacol 150:98-105. McKeown-Eyssen GE and Ruedy J (1983) Methylmercury exposure in northern Quebec. 1. Neurologic findings in adults. Am J Epidemiol 118(4):461-9. McKeown-Eyssen GE, Ruedy J and Neims A (1983) Methyl mercury exposure in northern Quebec. 11. Neurologic findings in children. Am J Epidemiol 118(4):470—9. Meacham CA, Freudenrich TM, Anderson WL, Sui L, Lyons-Darden T, Barone S Jr, Gilbert ME, Mundy WR and Shafer TJ (2005) Accumulation of methylmercury or polychlorinated biphenyls in in vitro models of rat neuronal tissue. Toxicol Appl Pharmacol 205(2): 177-87. Mellor JR and Randall A (1997) Frequency-dependent actions of benzodiazepines on GABA(A) receptors in cultured murine cerebellar granule cells. J Physiol (Land) 503:353-69. 186 Mellor JR and Randall AD (1998) Voltage-dependent deactivation and desensitization of GABA responses in cultured murine cerebellar granule cells. J Physiol (Land) 506(2):377-90. Mihalek RM, Banerjee PK, Korpi ER, Quinlan JJ, Firestone LL, Mi ZP, Lagenaur C, Tretter V, Sieghart W, Anagnostaras SG, Sage JR, Fanselow MS, Guidotti A, Spigelman 1, Li Z, DeLorey TM and Olsen RW, and Homanics GE (1999) Attenuated sensitivity to neuroactive steroids in gamma-aminobutyrate type A receptor delta subunit knockout mice. Proc Natl Acad Sci U S A 96(22): 12905-10. Mihic SJ, Ye Q, Wick MJ, Koltchine VV, Krasowski MD, Finn SE, Mascia MP, Valenzuela CF, Hanson KK, Greenblatt EP, Harris RA and Harrison NL (1997) Sites of alcohol and volatile anaesthetic action on GABA(A) and glycine receptors. Nature 389(6649):385~9. Miller LG, Greenblatt DJ, Roy RB, Gaver A, Lopez F and Shader RI (1989) Chronic benzodiazepine administration. III. Upregulation of gamma-aminobutyric acidA receptor binding and function associated with chronic benzodiazepine antagonist administration. J Pharmacol Exp Ther 248: 1096-1 101. Mintz 1M, Adams ME and Bean BP (1992) P-type calcium channels in rat central and peripheral neurons. Neuron 9(1):85-95. Moss SJ and Smart TG (1996) Modulation of amino acid-gated ion channels by protein phosphorylation. Int Rev Neurobiol 39: 1-52. Miura K and Imura N (1989) Mechanism of cytotoxicity of methylmercury. With special reference to microtubule disruption. Biol Trace Elem Res 21:313-6. Miura K, Koide N, Himeno S, Nakagawa I, and Imura N (1999) The involvement of microtubular disruption in methylmercury-induced apoptosis in neuronal and nonneuronal cell lines. Toxicol Appl Pharmacol. 160(3):279-88. Murata K, Weihe P, Renzoni A, Debes F, Vasconcelos R, Zino F, Araki S, Jorgensen PJ, White RF and Grandjean P (1999) Delayed evoked potentials in children exposed to methylmercury from seafood. Neurataxicol T eratal 21(4):343-8. 187 N agaya N and MacDonald RL (2001) Two gamma2L subunit domains confer low Zn2+ sensitivity to ternary GABA(A) receptors. J Physiol (Land) 532217-30. Narahashi T, Ma J Y, Arakawa O, Reuveny E and Nakahiro M (1994) GABA receptor- channel complex as a target site of mercury, copper, zinc, and lanthanides. Cell Mal Neurobiol 14:599-621. 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. J Comp Neural 272(3):317-30. Nicol] RA and Wojtowicz J M (1980) The effects of pentobarbital and related compounds on frog motoneurons. Brain Res 19l(1):225-37. Nierenberg DW, Nordgren RE, Chang MB, Siegler RW, Blayney MB, Hochberg F, Toribara TY, Cemichiari E and Clarkson T (1998) Delayed cerebellar disease and death after accidental exposure to dimethylmercury. NEJM 338: 1672-6. Nusser Z, Roberts J DB, Baude A, Richardson JG and Somogyi P (1995) Relative densities of synaptic and extrasynaptic GABA(A) receptors on cerebellar granule cells as determined by quantitative immunogold method. J Neurosci 15:2948-60. N usser Z, Sieghart W and Somogyi P (1998) Segregation of different GABA(A) receptors to synaptic and extrasynaptic membranes of cerebellar granule cells. J Neurosci 18: 1693-1703. Nusser Z, Sieghart W, Stephenson FA and Somogyi P (1996) The (16 subunit of the GABA(A) receptor is concentrated in both inhibitory and excitatory synapses on cerebellar granule cells. J Neurosci 16:103-14. O'Kusky JR and McGeer EG (1985) Methylmercury poisoning of the developing nervous system in the rat: decreased activity of glutamic acid decarboxylase in cerebral cortex and neostriatum. Brain Res 353:299-306. O'Kusky JR and McGeer EG (1989) Methylmercury-induced movement and postural disorders in developing rat: high-affinity uptake of choline, glutamate, and gamma- 188 aminobutyric acid in the cerebral cortex and caudate-putamen. J Neurochem 53:999- 1006. Olsen RW (1981) The GABA postsynaptic membrane receptor-ionophore complex. Site of action of convulsant and anticonvulsant drugs. Mal Cell Biochem 39:261-79. Olsen RW (1987) GABA-drug interactions. Prag Drug Res 31:223-41. Olsen RW and Avoli M (1997) GABA and epileptogenesis. Epilepsia 38(4):399-407. Olsen RW and Tobin AJ (1990) Molecular biology of GABA(A) receptors. FASEB J 4(5): 1469-80. Olsen RW, Yang J, King RG, Dilber A, Stauber GB, Ransom RW (1986) Barbiturate and benzodiazepine modulation of GABA receptor binding and function. Life Sci 39(21): 1969-76. Owens DF and Kriegstein AR (2002) Is there more to GABA than synaptic inhibition? Nat Rev Neurosci 3(9):715~27. Peng S, Hajela RK and Atchison WD (2002) Effects of methylmercury on human neuronal L-type calcium channels transiently expressed in human embryonic kidney cells (HEK-293). J Pharmacol Exp Ther 302(2):424-32. Pollard S, Duggan MJ, Stephenson FA (1993) Further evidence for the existence of alpha subunit heterogeneity within discrete gamma-aminobutyric acidA receptor subpopulations. J Biol Chem 268(5):3753~7. Poltl A, Hauer B, Fuchs K, Tretter V and Sieghart W (2003) Subunit composition and quantitative importance of GABA(A) receptor subtypes in the cerebellum of mouse and rat. J Neurochem 87(6): 1444-55. Pritchett DB, Luddens H and Seeburg PH (1989) Type I and type II GABA(A)- benzodiazepine receptors produced in transfected cells. Science 245(4924): 1389-92. 189 Puia G, Ducic I, Vicini S and Costa E (1993) Does neurosteroid modulatory efficacy depend on GABA(A) receptor subunit composition? Receptors Channels 1(2): 135-42. Puia G, Mienville JM, Matsumoto K, Takahata H, Watanabe H, Costa E and Guidotti A (2003) On the putative physiological role of allopregnanolone on GABA(A) receptor function. Neuropharmacology 44(1):49-55. Puia G, Santi MR, Vicini S, Pritchett DB, Purdy RH, Paul SM, Seeburg PH and Costa E (1990) Neurosteroids act on recombinant human GABA(A) receptors. Neuron 4(5):759~ 65. Quirk K, Gillard NP, Ragan CI, Whiting PJ, and McKeman RM (1994) Model of subunit composition of gamma-aminobutyric acid A receptor subtypes expressed in rat cerebellum with respect to their alpha and gamma/delta subunits. J Biol Chem 269(23) 1 6020-8. Randall A and Tsien RW (1995) Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J Neurosci 15(4):2995~3012. Rice DC (1995) Neurotoxicity of lead, methylmercury, and PCBs in relation to the Great Lakes. Environ Health Perspect 103:71-87. Rodier PM (1995) Developing brain as a target of toxicity. Environ Health Perspect 103(6):73-6. Rossi P, De Filippi G, Armano S, Taglietti V and D'Angelo E (1998) The weaver mutation causes a loss of inward rectifier current regulation in premigratory granule cells of the mouse cerebellum. J Neurosci 18:3537-47. Sakmann B, Hamill OP and Bormann J (1983) Patch-clamp measurements of elementary chloride currents activated by the putative inhibitory transmitter GABA and glycine in mammalian spinal neurons. J Neural Trans Suppl 18:83-95. Sarafian TA (1993) Methyl mercury increases intracellular Ca2+ and inositol phosphate levels in cultured cerebellar granule neurons. J Neurochem 61(2):648-57. 190 Saxena NC and MacDonald RL (1994) Assembly of GABA(A) receptor subunits role of the delta subunit. J Neurosci 14:7077-86. Saxena NC and MacDonald RL (1996) Properties of putative cerebellar 7-aminobutyric acid A receptor isoforms. Mal Pharm 49:567-79. Saxena NC, Neelands TR and MacDonald RL (1997) Contrasting actions of lanthanum on different recombinant Gamma-aminobutyric acid receptor isoforms expressed in L929 fibroblasts. Mal Pharmacol 51(2):328-35. Schmid G, Bonanno G, Raiteri L, Sarviharju M, Korpi ER and Raiteri M (1999) Enhanced benzodiazepine and ethanol actions on cerebellar GABA(A) receptors mediating glutamate release in an alcohol-sensitive rat line. Neuropharmacalagy 38: 1273-9. Schofield PR, Darlison MG, Fujita N, Burt DR, Stephenson FA, Rodriguez H, Rhee LM, Ramachandran J, Reale V, and Glencorse TA (1987) Sequence and functional expression of the GABA A receptor shows a ligand-gated receptor super-family. Nature 328(6127):221-7. Schulz DW and MacDonald RL (1981) Barbiturate enhancement of GABA-mediated inhibition and activation of chloride ion conductance: correlation with anticonvulsant and anesthetic actions. Brain Res 209(1): 177-88. Segal M and Barker JL (1984) Rat hippocampal neurons in culture Potassium conductances. J Neuraphysial 51: 1409-33. Semyanov A, Walker MC, Kullmann DM and Silver RA (2004) Tonically active GABA(A) receptors modulating gain and maintaining the tone. Trends Neurosci 27 :262- 9. Shafer TJ, Contreras ML and Atchison WD (1990) Characterization of interactions of methylmercury with Ca2+ channels in synaptosomes and pheochromocytoma cells radiotracer flux and binding studies. Molec Pharmacol 38:102-13. Shamlaye CF, Marsh DO, Myers GJ, Cox C Davidson PW, Choisy O, Cemichiari E, Choi A, Tanner MA and Clarkson TW (1995) The Seychelles child development study 191 on neurodevelopmental outcomes in children following in utero exposure to methylmercury from a maternal fish diet background and demographics. Neurataxicalogy l6(4):597-6l2. Sieghart W (1992) GABAA receptors: ligand-gated Cl‘ ion channels modulated by multiple drug-binding sites. Trends Pharmacol Sci 13:446-50. Sieghart W ( 1995) Structure and pharmacology of gamma-aminobutyric acidA receptor subtypes. Pharmacol Rev 47(2): 181-234. Sigel E, Baur R, Kellenberger S and Malherbe P (1992) Point mutations affecting antagonist affinity and agonist dependent gating of GABA(A) receptor channels. EMBO J ll(6):2017-23. Sirois JE and Atchison WD (1996) Effects of mercurials on ligand- and voltage-gated ion channels a review. Neurataxicalagy 17:63-84. Smart TG and Constanti A (1990) Differential effect of zinc on the vertebrate GABA(A)- receptor complex. Br. J Pharmacol 99:643-54. Smart TG, Moss SJ, Xie X and Huganir RL (1991) GABA(A) receptors are differentially sensitive to zinc dependence on subunit composition. Br J Pharmacol 103: 1837-9. Steuerwald U, Weihe P, Jorgensen PJ, Bjerve K, Brock J, Heinzow B, Budtz-Jorgensen E and Grandjean P (2000) Maternal seafood diet methylmercury exposure and neonatal neurologic function. J Pediatr l36(5):599-605. Study RE and Barker JL (1981) Diazepam and (~~)-pentobarbital: fluctuation analysis reveals different mechanisms for potentiation of gamma-aminobutyric acid responses in cultured central neurons. Proc Natl Acad Sci U S A 78(11):7180-4. Sumikawa K and Gehle VM (1992) Assembly of mutant subunits of the nicotinic acetylcholine receptor lacking the conserved disulfide loop structure. J Biol Chem 267(9):6286-90. Syversen TLM (1981) Effects of methyl mercury on protein synthesis in vitro. Acta Pharmacol Toxicol 49:422-6. 192 Takayasu Y, Iino M, Furuya N and Ozawa S (2003) Muscarine-induced increase in frequency of spontaneous EPSCs in Purkinje cells in the vestibulo-cerebellum of the rat. J Neurosci 23(15):6200-8. Takeuchi T, and Eto K (1999) The pathology study of Minamata Disease: a tragic story of water pollution. Kyushu University Press, INC. Okinawa, Japan. Takeuchi T, Norikawa N, Matsumoto H and Shiraishi Y (1962) A pathological study of Minamata Disease in Japan. Acta Neuropathol 2:40-57. Takeuchi T, Eto K and Tokunaga H (1989) Mercury level and histochemical distribution in a human brain with Minamata Disease following a long-term clinical course of twenty- six years. Neurataxicolagy 10(4):651-7. Takeuchi T, Matsumoto H, Sashi M, Kambara T, Shiraishi Y, Hirata Y, Nobukiro M and Ito H (1968) Pathology of Minamatma disease. Kumamata Med J 34:521-4. Thompson CL and Stephenson FA (1994) GABA(A) receptor subtypes expressed in cerebellar granule cells: a developmental study. J Neurochem 62:2037-44. Thompson SA, Arden SA, Marshall G, Wingrove PB, Whiting PJ and Wafford KA (1999) Residues in transmembrane domains I and II detemrine gamma-aminobutyric acid type A receptor subtype-selective antagonism by furosemide. Mol Pharmacol 55(6):993-9. Thompson SA, Whiting PJ and Wafford KA (1996) Barbiturate interactions at the human GABA(A) receptor: dependence on receptor subunit combination. Br J Pharmacol 117(3):521-7. Tia S, Wang JP, Kotchabhakdi N and Vicini S (1996) Distinct deactivation and desensitization kinetics of recombinant GABA(A) receptors. Neurapharmacalagy 35:1375-82. Toxicological Effects of Methylmercury (2000) Committee on the Toxicological Effects of Methylmercury. Board on Environmental Studies and Toxicology Commission on Life Sciences. National Research Council National Academy Press: Washington, DC. 193 Trasande L, Schechter CB, Haynes KA and Landrigan PJ (2006) Mental retardation and prenatal methylmercury toxicity. Am J Ind Med 49(3): 153-8. Traxinger DL and Atchison WD (1987) Reversal of methylmercury-induced block of nerve-evoked release of acetylcholine at the neuromuscular junction. Toxicol Appl Pharmacol 90(1):23-33. Twyman RE and MacDonald RL (1992) Neurosteroid regulation of GABA(A) receptor single-channel kinetic properties of mouse spinal cord neurons in culture. J Physiol 456:2]5-45. Twyman RE, Green RM and MacDonald RL (1992) Kinetics of open channel block by penicillin of single GABA(A) receptor channels from mouse spinal cord neurones in culture. J Physiol 445:97-127. Twyman RE, Rogers CJ and MacDonald RL (1990) Intraburst kinetic properties of the GABA(A) receptor main conductance state of mouse spinal cord neurones in culture, J Physiol 423: 193-220. Varju P, Katarova Z, Madarasz E and Szabo G (2001) GABA signalling during development: new data and old questions. Cell Tissue Res 305(2):239-46. Vicini S and Mienville JM (1987) Costa E. Actions of benzodiazepine and beta-carboline derivatives on gamma-aminobutyric acid-activated Cl- channels recorded from membrane patches of neonatal rat cortical neurons in culture. J Pharmacol Exp Ther 243(3):1195-201. Vicini S, Ferguson C, Prybylowski K, Kralic J, Morrow AL and Homanics GE (2001) GABA(A) receptor alpha] subunit deletion prevents developmental changes of inhibitory synaptic currents in cerebellar neurons. J Neurosci 21:3009-16. Vicini S, Losi G and Homanics GE (2002) GABA(A) receptor delta subunit deletion prevents neurosteroid modulation of inhibitory synaptic currents in cerebellar neurons. Neurapharmacolagy 43(4):646-50. Vicini S, Mienville JM, and Costa E (1986) Flunitrazepam action on the GABA-Cl ionophore complex in rat cortical neurons in culture: a patch-clamp study. Clin Neurapharmacol 9(4):395-7. 194 Weihe P, Hansen JC, Murata K, Debes F, Jorgensen P, Steuerwald U, White RF and Grandjean P (2002) Neurobehavioral performance of Inuit children with increased prenatal exposure to methylmercury. Int J Circumpolar Health 61(1):41-9. Weis IM (2004) Mercury concentrations in fish from Canadian Great Lakes areas of concern: an analysis of data from the Canadian Department of Environment database. Environ Res 95:341-50. Westbrook GL and Mayer ML (1987) Micromolar concentrations of Zn2+ antagonize NMDA and GABA responses of hippocampal neurons. Nature 328(6131):640-3. Westh-Hansen SE, Witt MR, Dekerrnendjian K, Liljefors T, Rasmussen PB and Nielsen M (1999) Arginine residue 120 of the human GABA(A) receptor alpha 1, subunit is essential for GABA binding and chloride ion current gating. Neurarepart 10(11):2417- 21. World Health Organization (1993) Evaluation of certain food additives and contaminants. Forty-first report of the joint FAO/WHO expert committee on food additives. WHO Technical Report Series No. 837 (Geneva: WHO). Wieland HA, Luddens H and Seeburg PH (1992) A single histidine in GABA(A) receptors is essential for benzodiazepine agonist binding. J Biol Chem 267(3): 1426-9. Wisden W and Seeburg PH (1992) GABA(A) receptor channels from subunits to functional entities. Curr Opin Neurobiol 2(3):263~9. Wisden W, Korpi ER and Bahn S (1996) The cerebellum a model system for studying GABA(A) receptor diversity. Neuropharmacalagy 35:1139-60. Wohlfarth KM, Bianchi MT and MacDonald RL (2002) Enhanced neurosteroid potentiation of ternary GABA(A) receptors containing the delta subunit. J Neurosci 22(5): 1541-9. Wood JM, Kennedy FS and Rosen CG (1968) Synthesis of methyl-mercury compounds by extracts of a methanogenic bacterium. Nature 220(163):173-4. 195 Xie XM and Smart TG (1991) A physiological role for endogenous zinc in rat hippocampal synaptic neurotransmission. Nature 349(6309):521-4. Xu M and Akabas MH (1996) Identification of channel-lining residues in the M2 membrane-spanning segment of the GABA(A) receptor alphal subunit. J Gen Physiol 107(2): 195-205. Xu YF and Atchison WD (1997) Methylmercury blocks 7- anrinobutyric acid (GABA(A) current and induces a nonspecific inward current in rat cerebellar granule cells. Sac Neurosci Abstr 23: 373. Yuan Y and Atchison WD (1993) Disruption by methylmercury of membrane excitability and synaptic transmission of CA1 neurons in hippocampal slices of the rat. Toxicol Appl Pharmacol 120:203-15. Yuan Y and Atchison WD (1994) Comparative effects of inorganic divalent mercury methylmercury and phenylmercury on membrane excitability and synaptic transmission of cal neurons in hippocampal slices of the rat. Neurataxicalagy 15:403-12. Yuan Y and Atchison WD (1995) Methylmercury acts at multiple sites to block hippocampal synaptic transmission. J Pharmacol Exp Ther 275:1308-16. Yuan Y and Atchison WD (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 282264-73. Yuan Y and Atchison WD (1999) Comparative effects of methylmercury on parallel- fiber and climbing-fiber responses of rat cerebellar slices. J Pharmacol Exp Ther 288: 1015-25. Yuan Y and Atchison WD (2003) Methylmercury differentially affects GABA(A) receptor-mediated spontaneous IPSCs in Purkinje and granule cells of rat cerebellar slices. J Physiol (Land) 550: 191-204. 196 Yuan Y and Atchison WD (2005) Methylmercury induces a spontaneous transient slow inward chloride current in Purkinje cells of rat cerebellar slices. J Pharmacol Exp Ther 313:751-764. Yuan Y, Otero-Montanez J KL, Yao A, Herden CJ, Sirois J E, Atchison WD (2005) Inwardly rectifying and voltage-gated outward potassium channels exhibit low sensitivity to methylmercury. Neurataxicalagy 26:439-54. Zempel JM and Steinbach JH (1995) Neonatal rat cerebellar granule and Purkinje neurons in culture express different GABA(A) receptors. Eur J Neurosci 7(9): 1895-905. Zhang S-P, Natsukari N, Bai G, Nichols RA and Weiss B (1993) Localization of the multiple calmodulin messenger RNAs in differentiated PC12 cells. Neuroscience 55:571- 82. Zhu W], Wang JF, Corsi L and Vicini S (1998) Lanthanum-mediated modification of GABA(A) receptor deactivation, desensitization and inhibitory synaptic currents in rat cerebellar neurons. J Physiol ( Land) 51 l(3):647-61. Zhu W], Wang JF, Krueger KB and Vicini S (1996) Delta subunit inhibits neurosteroid modulation of GABA(A) receptors. J Neurosci l6(21):6648~56. 197