LIBRARY Michigan State University PLACE iN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or boioro due due. DATE DUE DATE DUE DATE DUE cmmhi —— EFFECTS OF THE NEUROTOXICANT METHYLMERCURY ON NEURONAL Ca2+ CHANNELS: EFFECTS ON Ca2+ CHANNELS INVOLVED IN TRANSMITTER RELEASE Volume I By Timothy Joseph Shafer A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Pharmacology and Toxicology and Institute for Environmental Toxicology 1991 ABSTRACT EFFECTS OF THE NEUROTOXICANT METHYLMERCURY ON NEURONAL Ca2+ CHANNELS: EFFECIS ON Ca2+ CHANNELS INVOLVED IN TRANSMITTER RELEASE By Timothy Joseph Shafer The ability of the environmental neurotoxicant methylmercury (MeHg) to disrupt neuronal Ca2+ channel function was examined. In rat brain synaptosomes, MeHg (25-125 is M) blocked depolarization-induced 45Ca2+ influx. Increasing the concentration of extracellular Ca2+ ([Ca“L) only slightly antagonized block by MeHg of the Ca2+ channel- mediated component of “Ca influx, whereas block of the Na+ /Ca2* exchange-mediated component of influx was relieved. MeHg also disrupted Ca2+ channel ftmction by altering ionic selectivity and inactivation kinetics of synaptosomal Ca2+ channels. Eflects of MeHg did not depend on the state of synaptosomal Ca2+ channels but were exacerbated by increasing membrane depolarization. MeHg also inhibited [’lfl-nitrendipine binding to synaptosomal preparations. MeHg blocked depolarization induced “Ca2+ influx into differentiated and undiflerentiated rat pheochromocytoma (PC12) cells, and inhibited binding of the Ca2+ channel ligand to -conotoxin (Cg’Ik) to its receptors in this cell line. Studies of effects of MeHg on Ca channel function in PC12 cells using the whole-cell patch voltage clamp technique demonstrated that 5-20 u M MeHg reduced rapidly Ba“ current mediated by N- and btype Ca“ channels in PC12 cells. Increasing [Ba’m antagonized block by MeHg between 10 and 20 mM Ba“ but not between 20 and 30 mM Ba“. MeHg disrupted Ca” channel function in a manner which was voltage- but not state-dependent. Ionic selectivity and inactivation of whole cell current were altered by MeHg. Effects of MeHg on nerve terminal Ca“ channels in intact nerve muscle preparations were examined by measuring effects of MeHg on Ba2*-dependcnt voltage changes in the perineurial sheath of mouse aimgularis stemi preparations. 50-100 u M MeHg blocked the Bah-dependent potential as well as the Na*-dependent component. These results suggest that MeHg: 1) blocks Ca2+ channels in synaptosomal preparations and intact motor nerve terminals, 2) disrupts function of N- and L-type Ca channels in PC12 cells, 3) disrupts Ca2+ channel ionic selectivity and inactivation, and 4) disrupts Ca channel function in a manner which is voltage-, but not state-dependent in synaptosomes and PC12 cells. These actions may contribute to the ability of MeHg to disrupt synaptic transmission as well as other neurotoxic effects of Mel-1g. For Kathryn, Rosie and Loren iv ACKNOWLEDGEMENTS I would like to thank my parents, family and friends for their love, support and encouragement. Also, I would like to thank my wife Jenny for her love, patience and help. These people have contributed to this dissertation in ways that cannot be measured. I would like to thank my advisor, Dr. William Atchison, for his genuine interest in my graduate training. He has provided significant ideas, suggestions and criticisms which shaped this dissertation and has been an excellent teacher of science, laboratory techniques and writing skills. In addition, I would like to express my sincere thanks to my guidance committee for their input, ideas and suggestions, as well as their willingness to share their time to serve on the committee. The time and effort expended by student workers, technicians and other graduate students to help with experiments is appreciated. The excellent technical assistance from Ms. Lynne Adgate and Aizhen Yao and Drs. Paul Levesque and John Spitsbergen is especially appreciated. Finally, I would like to thank Ms. Diane Hummell and Mrs. Gretchen Humphries for their excellent clerical assistance. ‘ PREFACE Some of the data contained in this dissertation has been published previously. Chapter Three appeared as Shafer and Atchison (1989), Chapter Four as Shafer er a1. (1990) and Chapter Five as Shafer and Atchison (1991). TABLE OF CONTENTS LIST OF TABLES ................................................. xi LIST OF FIGURES ........................................... ’ ..... xii LIST OF ABBREVIATIONS .......................................... xv CHAPTER ONE: INTRODUCTION ..................................... 1 A) GENERAL INTRODUCTION ................................. 2 B) BACKGROUND ........................................... 7 1) Methylmercury in humans and the environment. ............... 7 2) Effects of Mel-lg on Synaptic Transmission. .................. 10 3) Synaptic Transmission .................................. 16 4) Types of Ca2+ channel .................................. 18 5) Environmental contaminants, natural toxins and pharmaceutical agents which affect Ca2+ channels ....................... 27 CHAPTER TWO: SYNAPTOSOMES AND PHEOCHROMOCYTOMA CELLS AS MODEL SYSTEMS FOR STUDY on Ca" CHANNELS ............... 32 A) SYNAPTOSOMES AND SYNAPTOSOMAL Ca2+ CHANNELS ...... 33 B) Caz” CHANNEL FUNCTION IN PC12 CELLS ................... 38 CHAPTER THREE: BLOCK OF ”Ca“ INFLUX INTO SYNAPTOSOMES BY METHYLMERCURY: Ca2+ AND Na‘ DEPENDENCE ................ 48 A) ABSTRACT .............................................. 49 B) INTRODUCIION ......................................... 51 C) MATERIALS AND METHODS ............................... 54 D) RESULTS ............................................... 58 E) DISCUSSION ............................................. 72 CHAPTER FOUR: _ CHARACTERIZATION OF THE INTERACTIONS OF METHYLMERCURY WITH Ca2+ CHANNELS IN SYNAPTOSOMES AND PHEOCHROMOCYTOMA CELLS: RADIOTRACER FLUX AND BINDING STUDIES .................................................. 78 A) ABSTRACT .............................................. 79 B) INTRODUCTION ......................................... 82 C) MATERIALS AND METHODS ............................... 86 D) RESULTS ..... . .......................................... 92 E) DISCUSSION ............................................ 115 CHAPTER FIVE: METHYLMERCURY BLOCKS N- AND L-TYPE Cal2+ CHANNELS IN NERVE GROWTH FACTOR-DIFFEREN'I‘IATED PI-IEOCI-IROMOCYTOMA (PC12) CELLS ........................ 127 A) ABSTRACT ............................................. 128 B) INTRODUCTION ........................................ 130 C) MATERIALS AND METHODS .............................. 132 D) RESULTS .............................................. 135 E) DISCUSSION ............................................ 153 CHAPTER SIX: EFFECTS OF METHYLMERCURY ON Na+ CHANNELS IN NEUROBLASTOMA CELLS ................................... 161 A) ABSTRACT ............................................. 162 B) INTRODUCIION ........................................ 163 C) MATERIALS AND METHODS .............................. 165 D) RESULTS .............................................. 167 E) DISCUSSION ............................................ 170 CHAPTER SEVEN: EFFECTS OF METHYLMERCURY ON Na+ AND Ca’*CURRENTs IN INTACT NEUROMUSCULAR JUNCTIONS ....... 174 A) ABSTRACT ............................................. 175 B) INTRODUCTION ........................................ 176 C) MATERIALS AND METHODS .............................. 179 D) RESULTS .............................................. 182 E) DISCUSSION ............................................ 190 CHAPTER EIGHT: SUMMARY AND CONCLUSIONS .................... 194 APPENDIX: TRANSMITTER, ION CHANNEL AND RECEPTOR PROPERTIES OF PHEOCHROMOCYTOMA (PC12) CELLS: A MODEL FOR NEUROTOXICOLOGICAL STUDIES ............................ A) INTRODUCTION ........................................ B) PC12 CELL DIFFERENTTATION. ........................... C) TRANSMITTERS, ION CHANNELS AND RECEPTORS IN PC12 CELLS ............................................... 1) Neurotransmitter Release from PC12 cells. ................. 2) Ion channels in PC12 cells. ............................. 3) Receptors on PC12 cell membranes. ...................... D) CLONAL VARIABILITY AND SUBCLONES OF PC12 ........... E) SUMMARY ............................................. 204 204 205 207 214 218 222 223 LIST OF TABLES Table 1.1 Biophysical and Pharmacological Characteristics of Ca2+ channel Types . . 23 Table 4.1 N on-Linear least squares regression values for apparent inactivation of 45C3“ influx into synaptosomes ............................... 101 Table A1 Activity and Content of Noradrenergic and Cholinergic Synthetic Enzymes and Neurotransmitters in PC12 Cells .............. 211 LIST OF FIGURES Figure 1.1 Effects of stimulus intensity and Ca2+c on EPP amplitude in the presence Figure 1.2. Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5. Figure 3.6. Figure 3.7. Figure 3.8. Figure 4.1. of MeHg. ............................................... 14 Sequence of events during synaptic transmission. .................. 19 Components of Ca2+ influx into synaptosomes. ................... 35 N and L type Ca2+ channels in PC12 cells. ...................... 40 N orepinephrine release from PC 12 cells. ........................ 42 Effects of EtOH on Ca“ channel density in PC12 cells .............. 47 MeHg blocks resting ”Ca“ entry into synaptosomes ................ 59 Influx of “Ca“ into synaptosomes vs duration of depolarization. ...... 60 Influx of 45Ca“ into synaptosomes as a function of protein concentration. ............................................ 61 MeHg blocks depolarization-induced “Ca2+ influx into synaptosomes. . . 63 Block of Ca2+ channels by MeHg is not reversed by increasing [Ca2*]e. . 64 Fast and slow components of influx have differential dependence on N a,. ................................................... 67 In the absence of Na,, block of slow influx by MeHg is much greater in magnitude. .............................................. 68 In the absence of Na,, block of slow influx by MeHg is not reversed by increasing [Caz‘]e .......................................... 71 Block of ”Ca“ influx into synaptosomes vs. time. ................. 93 Figure 4.2. Voltage-dependence of “Ca“ influx into synaptosomes and effects of MeHg. ................................................. 95 Figure 4.3. State-dependence of the actions of MeHg on “Ca2+ influx. .......... 97 Figure 4.4. Effects of MeHg on apparent inactivation of synaptosomal 45Ca2" influx. ................................................. 100 Figure 4.5. Effects of MeHg on the influx of 45Ca“, "SSrz+ and 133Ba2+ into synaptosomes. ........................................... 103 Figure 4.6. Effects of MeHg on the relative influx of 45Ca“, “Sr“ and 1”Ba2+ into synaptosomes. ........................................... 104 Figure 4.7. Influx of ”Ca“ into PC12 cells vs duration of depolarization. ....... 106 Figure 4.8. Influx of “Ca“ into PC12 cells as a function of protein concentration. . 107 Figure 4.9. Effects of nifedipine on depolarization-induced influx of 45Ca“ in PC12 cells ................................................... 108 Figure 4.10. Effects of MeHg on depolarization-induced ”Ca2+ influx into PC12 cells. .................................................. 110 Figure 4.11. Effects of MeHg on nitrendipine binding to synaptosomes .......... .113 Figure 4.12. Effects of MeHg on binding of conotoxin to synaptosomes and PC12 Figure 5.2. Current responses to a series of test potentials. .................. 137 Figure 53. Effects of o-conotoxin and (+)-s202-791 on Ba2+ current in PC12 cells ................................................... 138 Figure 5.4. Effects of MeHg on currents in PC12 cells. ..................... 140 Figure 5.5. Use-dependent effects of MeHg on In. in PC12 cells ............... 142 Figure 5.6. Resting block of In. by MeHg in PC12 cells. .................... 144 Figure 5.7. Effects of MeHg at different holding and test potentials in PC12 cells. . 145 Figure 5.8. Voltage-dependent block by MeHg of In. in PC12 cells is not irreversible. ............................................ 148 Figure 5.9. Ba” can antagonize block of Ca2+ channels by MeHg. ............ 149 Figure 5.10. MeHg alters ionic permeability of Ca2+ channels in PC12 cells. ..... 152 Figure 6.1. Effects of MeHg on Na+ channels in NIB-115 cells. .............. 168 Figure 6.2. Effects of MeHg on time to peak Na+ current. .................. 169 Figure 7.1. Local circuits in the perineurial sheath of motor neurons. .......... 178 Figure 7.2. Diagram of mouse triangulari: stemi muscle. .................... 181 Figure 7.3. Voltage responses recorded from perineurial sheaths. ............. 183 Figure 7.4. Ca2+ channel-mediated component of voltage potentials in the perineurial sheath. ................................................ 186 Figure 7.5. Effects of MeHg on Na‘/Caz* potentials in the perineurial sheath. . . . 187 Figure 7.6. Effects of MeHg on repetitive firing of motor nerves. ............. 189 Figure A]. Vesicular nature of ACh release from PC12 cells. ................ 213 Figure A.2. Voltage-dependent Na‘ channels in PC12 cells. ................. 217 xiv LIST OF ABBREVIATIONS ACh - acetylcholine AN OVA - analysis of variance ATP - adenosine 5’-triphosphate (disodium salt) [Bar'],3 - extracellular Ba2+ concentration Bum - density of binding sites for a ligand CAT - choline acetyltransferase CAMP - cyclic adenosine monophosphate [Ca2+]: - sum of the extracellular concentration of all calcium isotopes [Ca2+], - intracellular Ca2+ concentration Cng - moconotoxin GVIA DA - dopamine DEX - dexamethasone DI-IP - dihydropyridine DRG - dorsal root ganglion EGF - epidermal growth factor EGTA - ethyleneglycol bis-(B—aminoethyl ether)-N,N,N’,N’,-tetraacetic acid EPP - endplate potential EtOH - ethanol FGF - fibroblast growth factor FTX - synthethic funnel-web spider toxin GTP - guanosine triphosphate HEPES - N-2-hydroxyethylpiperazine-N’-2-ethanesulfonic acid IA - whole cell current through the inactivating K+ channel IBa - whole cell current mediated by barium ICa - whole cell current mediated by calcium IK - whole cell current mediated by the delayed rectifier K+ channel INa - whole cell current mediated by sodium IC$0 - concentration of chemical that causes 50% inhibition of the maximal response of the system KD - equilibrium dissociation constant for ligand binding lsd - least significant difference MEPP - miniature endplate potential MeHg - methylmercury ms - milliseconds [Na*],z - extracellular concentration of Na+ N GF - nerve growth factor N MJ - neuromuscular junction pA - picoamperes PC12 cells - pheochromocytoma cells PKA - protein kinase A PKC - protein kinase C PI - phosphoinositide SEM - standard error of the mean TEACl - tetraethylammonium chloride TI'X - tetrodotoxin xvii CHAPTER ONE INTRODUCTION A) GENERAL INTRODUCTION Methylmercury (MeHg) is a neurotoxic heavy metal found as a contaminant throughout the environment, but especially in aquatic ecosystems. The organic component of this molecule imparts to MeHg the ability to bioconcentrate in aquatic organisms and biomagnify within the food chain. Consumption of contaminated seafood has resulted in numerous cases of MeHg-intoxication in human populations. In addition, agricultural use of MeHg has resulted in human intoxication from accidental consumption of MeHg-treated grain. The neurotoxic potential of MeHg was dramatically illustrated by chronic and acute episodes of human exposure in Japan (Takeuchi et al., 1962; 1968) and Iraq (Bakir et aL, 1973), respectively. The characteristic clinical symptoms of MeHg-intoxication observed included cerebellar ataxia, speech and visual impairments, and, in the Iraqi episode, weakness of the extremities (Chang, 1980). Central and peripheral lesions of the nervous system have been observed in cases of MeHg intoxication (Chang and Hartmann, 1972a,b), but the cellular mechanisms responsible for these lesions are unknown. In individuals from the Iraqi episode, neostigmine alleviated clinically observable symptoms of motor weakness produced by MeHg (Rustam et aL, 1975), indicating that MeHg possibly disrupts cholinergic transmission at the motor end- plate. This prompted further in vitro study of the effects of MeHg on neuromuscular transmission. In isolated nerve-muscle preparations, MeHg exerts two prominent presynaptic actions: 1) block of nerve-evoked neurotransmitter release, and 2) 3 increased frequency, followed by complete block of spontaneous neurotransmitter release (Barrett et al., 1974; Juang, 1976; Atchison and Narahashi, 1982). Release of neurotransmitter following invasion of the action potential into the nerve terminal is dependent on influx of Ca2+ through voltage-sensitive Ca2+ channels in the membrane (Katz and Miledi, 1967a,b). Several divalent heavy metal cations which produce effects on evoked and spontaneous transmitter release similar to those of MeHg block the movement of Ca2+ through these channels. Traxinger and Atchison (1987b) demonstrated that, under certain conditions, effects of MeHg on nerve-evoked transmitter release are relieved by increasing the extracellular Ca2+ concentration ([Ca2+]c). MeHg also blocks depolarization-induced 45Ca2“ influx into synaptosomes. Effects of MeHg on 45Ca2+ influx are also partially antagonized by increasing [Ca2+]e (Atchison et 611., 1986). These results indicate that the ability of MeHg to disrupt synaptic transmission at the neuromuscular junction may, in part, be due to a disruption of Ca2+-dependent events at the motor nerve terminal. This dissertation reports results of experiments designed to examine the following hypothesis: disruption of voltage-sensitive Ca2+ channel function by MeHg contributes to block of synaptic transmission observed in neuromuscular preparations. The emphasis of studies undertaken was to characterize effects of MeHg on Ca2+ channels, determine types of Ca2+ channels affected by MeHg, and examine the ability of MeHg to affect Ca2+ channels in intact motor nerve terminals. Entry of Ca2+ into nerve terminals at the neuromuscular junction is difficult to measure directly. Thus, effects of MeHg on channel-mediated 45Ca2+ influx into 4 synaptosomes (nerve terminals isolated from the central nervous system) were examined in initial experiments. In addition, non-channel mediated entry of 45Ca“ was also examined as a measure of the specificity of MeHg for Ca2+ channel function. MeHg readily blocks the channel-mediated components of influx in a manner which is not completely antagonized by increasing [Ca2+]e. MeHg also affects non-channel-mediated components of influx, including Na+ /Ca2+ exchange and depolarization-independent entry of Ca2+. For components of Ca2+ influx not mediated by Ca2+ channels, the ability of increasing [Ca2+]e to antagonize the actions of MeHg and/ or the concentrations at which MeHg effectively blocks entry are different than for effects on voltage-sensitive Ca2+ channels by MeHg (Shafer and Atchison, 1989). Furthermore, the ability of Ca“., to antagonize the block of channel- mediated 45Ca2+ influx by MeHg is decreased when compared to block by divalent cations. When compared to the actions of inorganic Ca2+ channel blockers, among the unique effects of MeHg on Ca2+ channel function in synaptosomes are alteration of the apparent rate of inactivation of synaptosomal Ca2+ channels as well as the ionic permeability of Ca2+, Sr2+ and Ba2+ through the Ca2+ channel. In addition, the effects of MeHg do not depend directly on the state of the channel, but block of 45Ca“ influx is increased in magnitude as a function of the magnitude of the depolarizing stimulus (Shafer et aL, 1990). Because effects of MeHg are expressed in the central as well as the peripheral nervous system, the use of nerve terminal preparations derived from mammalian 5 central neurons provides meaningful and relevant data about effects of MeHg. Continued studies in central nervous system preparations in the future will be necessary to understand completely the mechanisms of MeHg’s neurotoxic effects. Because synaptosomal preparations are heterogeneous with respect to the neurotransmitter content of individual terminals within the preparation, the ability of MeHg to block Ca2+ uptake into synaptosomes may represent a common action of MeHg at all types of terminals, and because Ca2+ entry is required for neurotransmitter release, the ability of MeHg to disrupt this process may be a component underlying its neurotoxicity. However, while these experiments demonstrate that MeHg is capable of producing unique alterations of Ca2+ channel function in a population of isolated nerve terminals, they provide only preliminary evidence to support the stated hypothesis. The purpose of studies performed in the latter portion of this dissertation was to collect data which would provide stronger evidence for or against the hypothesis. Transmitter release from central and peripheral neurons is likely to be mediated at least in part by N- and/ or L-type Ca2+ channels (Fox et aL, 1987a,b) or as yet uncharacterized Ca2+ channel types which inactivate Slowly (Smith and Augustine, 1988; Suszkiw et at, 1989). The hypothesis predicts that MeHg should be capable of blocking one or both of these types of channels if block of Ca“ channels by MeHg contributes to the disruption of synaptic transmission. Therefore, the ability of MeHg to interact with N- and/ or L-type Ca2+ channels was determined using whole-cell patch voltage-clamp techniques to examine effects of MeHg on Ca2+ (or 6 Ba“) current in rat pheochromocytoma (PC 12) cells. PC12 cells express both N - and L-type Ca2+ channels when cultured in the presence of nerve growth factor. The use of patch voltage-clamp techniques allows unambiguous identification of Ca2+ channel types as well as the ability to measure channel function with temporat resolution that approaches physiological rates of channel function. However, this method of examining Ca2+ channel function requires use of supraphysiological concentrations of Ca2+ or other permeant ions such as Ba2+ or Sr“. If MeHg is able to block one or both of the aforementioned types of Ca2+ channel, this would be strong evidence in favor of the hypothesis. The hypothesis was also examined by measuring directly Ca2+ channel function at the terminal regions of motor neurons. Whole cell patch-voltage clamp recording, while allowing for unambiguous determination of Ca2+ channel types, nonetheless measures function of predominantly somatic rather than terminal currents. By precisely placing electrodes in the perineurial sheath near the motor nerve terminal, one can record voltage changes within the perineurial sheath in response to movement of Ba“ through Ca2+ channels in the nerve terminal (Penner and Dreyer, 1986). Measurement of effects of MeHg on Ca2+-dependent voltage waveforms from the perineurial Sheath will provide the most direct assessment of the ability of MeHg to alter Ca2+ channel function in intact motor nerve terminals. MeHg is a highly reactive compound, especially with sulfhydryl groups. As such, MeHg binds to a large variety of proteins and disrupts a large variety of cellular processes, including channel function, mitochondrial function (Levesque and 7 Atchison, 1987; 1988; 1991), microtubule structure (Abe et at, 1975; Sager elf £11., 1983) and protein synthesis (Yoshino et aL, 1966). Furthermore, exposures to MeHg which produce significant pathological changes are also likely to alter numerous cellular functions. Thus, it is important to characterize all events which might contribute to compromised cellular function in the presence of MeHg rather than attempt to define a single critical event which results in altered cellular function (Reuhl, 1987). The plasma membrane, along with its resident channels, receptors and transporter proteins is the initial Site of interaction for all foreign and endogenous compounds with the cell (See Narahashi, 1980) and plays a vital role in cellular homeostasis. In the nervous system, the proper functioning of Na+ and K+ channels is crucial for propagation of action potentials, and Ca2+ channel function is critical to the process of neurotransmitter release. By characterizing the ability of MeHg to disrupt Ca2+ channel function, the effects at this initial site of interaction with the cell will be better understood. The significance of this information may exceed possible effects on neurotransmitter release, as Ca2+ channels are involved in cellular Ca2+ homeostasis; alterations of Ca2+ homeostasis may therefore contribute to disruption of cellular function by MeHg. B) BACKGROUND 1) Methylmercury in humans and the environment. Chronic and acute episodes of human intoxication with MeHg have occurred previously. In the mid- to late- 1950’s, ingestion of MeHg-contaminated shellfish from 8 Minamata Bay in Japan led to 1500 cases of intoxication and 46 deaths (Takeuchi et aL, 1962, 1968). The cause of contamination was due to discharge of organic and inorganic mercurial waste into the bay from a chemical plant producing acetaldehyde from acetylene. Mercury metal used as a catalyst reacted with the acetylene, ultimately forming MeHg as a contaminant. In Iraq, the use of MeHg-treated seed grain for bread (despite government warnings of adverse health effects) resulted in 450 deaths due to acute MeHg poisoning in 1972 (Bakir, et al., 1973). Thus, environmental contamination with MeHg, as well as any significant exposure to MeHg, represent legitimate human health concerns. MeHg still remains a considerable human health hazard, and certian populations, particularly those which consume large omounts of fish, remain at risk (Hansen, 1990). There are two main sources of environmental contamination by MeHg: MeHg itself, and elemental Hg which is converted to MeHg. Because of its antifungal properties, MeHg has been employed by the agricultural, lumber, paper and leather industries, primarily as a preservative. Consequently, it is found as a contaminant in seed grain dressings, agricultural mnoff, and waste-water discharges from these industries. Elemental Hg has been used by the electrical apparatus, instrumentation, paint and chloralkali industries, and is released into the environment via atmospheric venting or discharge along with waste-water. Mercury is also released into the environment during the combustion of fossil fuels. In the aquatic environment, microorganisms can methylate inorganic mercury (Wood et aL, 1968; Ridley et aL, 1977), thus increasing the pool of MeHg in the environment. 9 The conversion of Hg to MeHg in the environment has important toxicological implications. Prior to knowledge of this conversion, inorganic Hg discharged into the environment was considered to be relatively harmless to humans as it is poorly absorbed by the gastrointestinal tract. By contrast, the more lipophilic MeHg tends to bioaccumulate and biomagnify in the food chain, and up to 95% is absorbed by the gastrointestinal tract. MeHg is transported in the plasma and red blood cells and distributes evenly to most tissues, but concentrates in the blood and central nervous system. In rats chronically exposed to MeHg, the spinal dorsal root ganglia accumulated the highest levels of MeHg, while the cerebral cortex and cerebellum also contained high levels of MeHg (Somjen et al., 1973). The blood-brain barrier is easily crossed and also damaged by MeHg (Steinwall and Klatzo, 1966; Steinwall and Olsson, 1969; Chang and Hartmann 1972c). Pathological examination of victims of MeHg toxicity in Minamata indicated that MeHg lesions the calcarine (visual) cortex, as well as cerebellar granule cells and the dorsal root ganglia (Chang, 1977; 1979). Histochemical analysis has shown that MeHg is closely associated with the mitochondria, golgi, endoplasmic reticulum, and the nuclear envelope. In nerves, MeHg is found primarily in myelin sheaths and mitochondria, and causes axonal degeneration (Chang, 1977). The high affinity of MeHg for membranous structures in the neuron may contribute to the toxicity of this compound by disrupting important cellular functions which take place at these sites. 10 MeHg is Slowly excreted; its half-life in humans is approximately 70 days. MeHg slowly decomposes to inorganic Hg, is completely reabsorbed from the bile, and has a low rate of urinary excretion compared to elemental Hg (Magos, 1975). Thus, frequent consumption of MeHg-contaminated food or water can result in significant accumulation of MeHg in the body. Consumption of contaminated seafood has resulted in the accumulation of toxic doses of MeHg in Japanese (Takeuchi et at, 1962; 1968) and Swedish (Berglund et al., 1971) populations. Such episodes have resulted in concern regarding the safety of fish from the Great Lakes. However, mercury levels in fish in the Great Lakes have declined since 1976, and the average Hg concentration in fish in each of the five Great Lakes is below the action level established by the Food and Drug Administration (1.0 mg/kg) and the Michigan Department of Public Health (0.5 mg/kg) (D’Itri, 1985). 2) Effects of MeHg on Synaptic Transmission. The observation that neostigmine alleviated symptoms of motor weakness spurred research into the mechanism underlying this action of MeHg. Neuromuscular preparations such as the rat phrenic nerve-hemidiaphram preparation have served as useful model systems for study of synaptic transmission between nerve and muscle. The neuromuscular junction (NMJ) has been well characterized with respect to its structure, function and pharmacological response to various drugs and toxins. In addition, the synthesis, storage and Caz+-dependent release of acetylcholine (ACh) at the NMJ are similar in many respects to the synthesis, storage and release of other ll neurotransmitters in central and peripheral neurons. As MeHg not only affects neurons which release ACh, but also those neurons which release catecholamines, serotonin, GABA, glutamate and glycine (Borowitz, 1974; Nakazato, et al., 1979; Bondy, et al., 1979; Tuomisto and Komulainen, 1983), information obtained using this preparation may also provide preliminary indications as to how MeHg affects other neurons as well. Following acute bath application to isolated nerve-muscle preparations, MeHg causes a rapid block of nerve-evoked transmitter release and an increase in the frequency, followed by a complete block, of spontaneous neurotransmitter release (Barrett et al., 1974; Juang, 1976; Atchison and Narahashi, 1982; Miyamato,1983; Atchison et al., 1984; Atchison, 1986; Levesque and Atchison 1987; 1988). At the neuromuscular junction, these effects are observed electrophysiologically as alterations of the amplitude of end plate potentials measured in the muscle cell in response to nerve-evoked (EPP) or spontaneous (MEPP) neurotransmitter release. Block of the EPP by 100 11M MeHg occurs rapidly and with no apparent gradual reduction; after 8-20 min the EPP is completely abolished (Atchison and N arahashi, 1982; Atchison et al., 1984; Traxinger and Atchison, 1987b). By contrast, changes in MEPP frequency occur after a latent period, which can be shortened without increasing the peak MEPP frequency observed by increasing the concentration of MeHg (Atchison and N arahashi, 1982) or by K+-induced depolarization (Atchison, 1986). The ultimate block of transmitter release produced by MeHg is not due to depletion of ACh stores; treatment with La3+ after complete block of MEPPs by 12 MeHg results in their restoration (Atchison, 1986; Levesque and Atchison, 1988), indicating that the nerve terminal has not been depleted of ACh-containing vesicles. Although post-synaptic actions of MeHg have not yet been carefully examined, several observations indicate that the block of EPP and increase in MEPP frequency caused by MeHg are the result of actions of MeHg on the presynaptic terminal. The experimental evidence supporting a presynaptic action of MeHg include the following observations: 1) the response of the endplate to iontophoretically applied ACh is unchanged even after 1 hr of exposure to 100 MM MeHg (Atchison and Narahashi, 1982). At this time, neither EPPS nor MEPPS are observed. 2) MEPP amplitude is not affected at the time of complete EPP block by MeHg (Atchison and N arahashi, 1982). These results indicate that the postsynaptic cell still responds to ACh at a time when synaptic transmission is impaired by MeHg. In addition to this experimental evidence, changes in MEPP frequency are thought to be due to presynaptic and not postsynaptic events. Not only do the actions of MeHg appear to be presynaptic, but the differences in the time course of MeHg’S effects on 'EPPs and MEPPS suggest different mechanisms of action by MeHg on each of these forms of release. Indeed, the mechanisms underlying the EPP and MEPP themselves are different. The MEPP is a depolarization of the end-plate membrane resulting from release of one quantum of neurotransmitter and subsequent interaction with receptors located on the postsynaptic membrane (Fatt and Katz, 1952). Although MEPPS are affected by extracellular Ca“, spontaneous release can occur in Ca2*-free solutions. The l3 frequency of MEPP occurence reflects the intraterminal [Ca2+] (Hubbard et al., 1968). Thus, it has been proposed that the actions of MeHg on MEPP frequency are due to a disruption of intraterminal Ca2+ buffering organelles, such as the smooth endoplasmic reticulum and mitochondria, leading to an increase in [Ca2+]i (Atchison et al., 1984; Atchison 1986; Levesque and Atchison, 1987; 1988; 1991). The EPP, however, is the postsynaptic response resulting from synchronous release of many quanta of neurotransmitter following the depolarization of the nerve terminal by an action potential and subsequent entry of extracellular Ca2+ (Katz and Miledi, 1967a,b; Llinas et al., 1981). Under physiological conditions, EPPS require Ca2+e. The rapid and complete block of EPPS by MeHg indicates that MeHg may block conduction of the action potential, electrotonic Spread of depolarization into the terminal and/ or entry of Ca2+ into the nerve terminal. Initially, MeHg-induced block of evoked release at the NMJ is not reversible simply by methods designed to overcome a competitive block of divalent cation influx by MeHg. Block of the EPP by MeHg is not reversed by perfusing the preparation with MeHg-free solutions containing normal or elevated [CaZI] or with MeHg-free solutions containing agents (4-aminopyridine) which prolong Ca2+ entry into the nerve terminal. However, increasing the intensity or duration of the stimulus in MeHg-free solutions does partially and occasionally reverse MeHg-induced block of evoked release (Traxinger and Atchison, 1987b) (Figure 1.1). This indicates that MeHg may affect impulse conduction at either the axonal membrane or the unmyelinated portion of the nerve terminal. Previous studies have indicated that l4 9 e E \u. 4 mM Ca Lu 5 a)” 0..” o MeHg O D lT / MeHg t A ._i 00 CL 3 .MeHg 'T‘V.. e—e—e 2 '\ < Ca 8. LIJ 0 t i 0 6 12 32 40 TIME (min) Figure 1.1 Effects of stimulus intensity and Ca2+e on EPP amplitude in the presence of MeHg. EPPS were evoked in rat hemidiaphram by stimulation of the phrenic nerve at 0.25 Hz throughout the experiment. MeHg was applied continuously for 9 min by which time the EPP was blocked. Immediately following block, the preparation was perfused with a MeHg-free Ca2+ ([Ca“) = 4 mM) solution for 15 min. During this time, EPPS remained blocked. While maintaining the preparation in 4 mM Ca2+ solution, the stimulus intensity was increased (t V); this brought about an immediate return of EPPS. Subsequent rc-exposure of the preparation to MeHg in normal (2 mM) Ca2+ solutions caused EPPS to be abolished within 3 min. Wash of the preparation at this time with MeHg-free, 4mM Ca2+ solution caused a prompt recovery of the EPP. All values are from a single representitive preparation. Modified from: Traxinger and Atchison, T axicol. Appl. Pharmacol. 90:23-33 (1987b). 15 MeHg may interact with Na+ channels or sulfhydryl groups in the membrane which are important in action potential propagation (Shrivistav, et al., 1976; Quandt et al., 1982). An increase in the duration or intensity of stimulation would be expected to overcome a MeHg-induced decrease in excitability in these regions (Konishi, 1985). Under conditions of increased stimulus intensity, re-exposure of the preparation to MeHg resulted in a re-occurrence of block of EPPS. This subsequent block of EPPS could be reversed by increasing [Ca2+], (Traxinger and Atchison, 1987b). Thus, block of EPPS by MeHg may in part be due to block by MeHg of Ca2+ entry associated with neurotransmitter release. This hypothesis is strengthened by the observation that MeHg blocks depolarization-induced 45Ca2+ influx into rat forebrain synaptosomes, and that the block of influx by MeHg over 10 s is antagonized, although not completely overcome by increasing [Ca2"],3 (Atchison et al., 1986). In addition to experimental evidence suggesting that MeHg may block Ca2+ entry associated with neurotransmitter release, one may suspect, based on its similarities in physico-chemical properties (i.e., relatively similar ionic radius, positive charge, ability to gain or lose electrons) (Miller, 1984) with other heavy and/ or transition metals, that MeHg could exert this action. Other heavy metals such as N i2+ (Kita and Van der Kloot, 1973), Cd2+ (F orshaw, 1977; Cooper and Manalis 1983) and Pb2+ (Kostial and Vouk, 1957; Manalis and Cooper, 1973; Silbergeld et al., 1974a,b; Atchison and Narahashi, 1984; Cooper and Manalis, 1983; Biisselberg et al. ,1990), block evoked transmitter release in a fashion Similar to MeHg. However, in contrast to the effects of MeHg, block of synaptic transmission at the NMJ by Pb2+ 16 (Kober and Cooper, 1976; Cooper and Manalis, 1983; Atchison and Narahashi, 1984; Pickett and Bornstein, 1978; 1984), Cd2+ (Nilson and Volle, 1976; Cooper and Manalis, 1984a,b), or Co2+ (Weakly, 1973) is reversed by simply increasing [Ca2+ 1, or washing with normal, metal-free solutions. In synaptosomes, block of 45Ca2+ influx by divalent cations such as Pb“, Ni“, Cd2+ and Hg“ (N achshen, 1984; Atchison et al., 1986), but not MeHg (Atchison et al., 1986; Shafer and Atchison, 1989), is completely reversed by increasing [Ca2+]e concentration. As MeHg has enhanced lipophilicity and a monovalent charge, differences in charge and lipophilicity may underlie the differences in the action of MeHg and inorganic, divalent cations on nerve terminal Ca2+ fluxes. 3) Synaptic Transmission Successful transmission of neuronal signals, whether between nerve and muscle, post-synaptic nerve or other cell type, requires a complex sequence of events. Disruption of any one event, such as block of Na” channels by tetrodotoxin (N arahashi er al. , 1964) or inhibition of acetylcholinesterase by organophosphate insecticides, can lead to deleterious consequences. Many neurotoxic compounds (including heavy metals) disrupt processes at the synapse. Thus, it is conceivable that MeHg may also act at this Site. Transmission at peripheral and central chemical synapses involves the conversion of an electrical signal into a chemical signal in the presynaptic cell, and then, depending on the postsynaptic cell, into an electrical or biochemical (i.e., 2nd l7 messenger) signal. Although the complete sequence of events resulting in neurotransmitter release is not yet known, several processes which take place prior to release have been identified. Action potentials propagated along the nerve axon by movement of Na+ and K+ into and out of the axon through their respective channels (Hodgkin and Huxley, 1952) cause depolarization of the non-myelinated terminal membrane by electrotonic spread into the terminal region (Brigant and Mallart, 1982; Mallart, 1985). This depolarization causes voltage-sensitive Ca2+ channels in the nerve terminal membrane to open, allowing extracellular Ca2+ to move down its electrochemical gradient into the nerve terminal (Katz and Miledi, 1967a,b; Llinas er al., 1981). It is thought that these channels, whose activation is crucial for evoked release, are clustered in areas in which vesicles containing neurotransmitter are arranged in close apposition to the cell membrane (Pumplin et al., 1981; Robitaille et al., 1990; Cohen et al., 1991). Influx of Ca2+ in these areas, or "active zones”, gives rise to locally high [Ca2+]i, resulting in increased interaction of Ca2+ ions with CaZI-binding proteins. This interaction between Ca2+ and binding proteins is thought to trigger directly or indirectly the fusion of synaptic vesicles with the cell membrane, resulting in the release of chemical neurotransmitters into the synaptic cleft (Augustine er al., 1987). The protein or proteins to which Ca2+ binds have yet to be identified, although synapsin I, synexin or other members of the calelectrin family and p38-synaptophysin have been suggested as the macromolecule(s) responsible for coupling Ca2+ entry to neurotransmitter release (Smith and Augustine, 1988). Injection of dephosphorylated synapsin I into the 18 preterminal region of squid giant synapse inhibited neurotransmitter release (Llinas et al., 1985) and reduced spontaneous release (Lin et al., 1990a). Furthermore, depolarization of synaptosomes caused translocation of synapsin I from the particulate (presumably cytoskeletal) fraction to the soluble (cytoplasmic) fraction in a Ca2+ and phosphorylation-dependent manner (Sihra et al., 1989). It has been suggested that this protein may regulate the availability of synaptic vesicles for release (Llinas et al., 1985). However, the exact role of this molecule in neurotransmitter release has yet to be determined. The released transmitter diffuses across the synaptic cleft and binds to receptors located on the membrane of the post-synaptic cell, resulting in an electrical or chemical response in the post-synaptic cell. Continuous activation of the receptor is prevented by re-uptake of the transmitter into the pre-synaptic cell, as in adrenergic synapses, or by enzymatic destruction of the transmitter, as in cholinergic synapses (Figure 1.2). 4) Types of Ca2+ channel Entry of extracellular Ca2+ into the nerve terminal is critical for release of neurotransmitter; N a+ and K" do not support release, nor will significant release occur in Ca2*-fl'ee medium (Katz and Miledi, 1967a,b). Therefore, an understanding of the nature of the Ca2+ channels that may regulate Ca2+ entry associated with neurotransmitter release is important to understand the process of synaptic transmission. However, it is not currently possible to characterize "active zone” Ca2+ l9 Reuotahe A ‘2 [Hydroiysts Messenger 9.Biochemical Changes Figure 1.2. Sequence of events during synaptic transmission. 1) Depolarization of the terminal region by electrotonic spread of the action potential into the terminal region from pro-terminal nodes of Ranvier. 2) Activation of voltage-sensitive Ca2+ channels and influx of Ca2+ into the nerve terminal. 3) Oak-triggered fusion of synaptic vesicles with the terminal membrane. 4) Release of neurotransmitter into the synaptic cleft and 5) difl‘usion to the post-synaptic cell where 6) binding to receptors occurs. Following binding, the transmitter is removed from the cleft by 7) hydrolysis and/or carrier-mediated re-uptake into the presynaptic terminal. Binding of the neurotransmitter to receptor activated channel complexes results in 8) conductance changes and 9) synaptic potentials, whereas binding to other receptor types results in changes in 8) second messingcrs and 9) cellular biochemistry. Modified from: Shepherd, G. H. In: Synaptic organization of the brain. (Shepherd, G. H., ed.). p. 5, (1990). 20 channels using intracellular or patch electrodes, as this region accounts for only a small portion of the nerve terminal. Therefore, the best method presently available to determine which type or types of Ca2+ channel may be involved in transmitter release is to characterize the types of somatic Ca2+ channel,and then infer which types may be involved in release based on the properties and pharmacological sensitivities of Ca2+ channels and the release process. However, a lack of pharmacological effect on neurotransmitter release may result from inadequate access of drugs or toxins to the channel rather than channel insensitivity. Thus, caution must be observed when making inferences as to channel type(s) critical to neurotransmitter release bases solely on pharmacological sensitivity. Two types of Ca2+ channel have been consistently demonstrated in neurons (Llinas and Yarom, 1981; Carbone and Lux, 1984). One type begins to activate (opens) at membrane potentials between —60 and -40 mV, has a very small unitary conductance, completely inactivates during a sustained voltage pulse and remains inactivated if the membrane potential is held positive to -60 mV. Depending on the nomenclature system used, these channels are termed "Low Voltage Activated" (LVA), "Type I", or "T-type" channels. The second type of Ca2+ channel is activated at membrane voltages of -20 to 0 mV, has a large unitary conductance, and generally does not inactivate during sustained voltage pulses of 100-500 ms. These channels are denoted as "High Voltage Activated" (HVA), "Type II", or "L-type" Ca2+ channels. A third subtype has been described in avian (Nowycky er al., 1985a; Fox et aL, 1987a,b) and mammalian (Kostyuk et al., 1988) embryonic dorsal root ganglion 21 (DRG) cells and in nerve grth factor-differentiated rat pheochromocytoma (PC12) cells (Plummer et al., 1989) as well as many other types of neuronal cells. It is also a high voltage activated channel, but the unitary conductance of this channel is smaller than the "L-type" channel and inactivation of this channel occurs during a sustained voltage pulse. This channel has been termed "N-type". However, the existence of this channel as a separate entity from the L channel has been questioned on the basis of analysis of tail current recordings from chick DRG neurons (Swandulla and Armstrong, 1987) and growth cone currents from PC12 cells (Streit and Lux, 1990). The evidence supporting three separate Ca2+ channel entities has been reviewed recently (Swandulla et al., 1991). Recently, an additional Ca2+ channel type has been described in Purkinje cell neurons of guinea pig cerebellum and in the presynaptic terminal of squid giant synapse (Llinas et al., 1989). The biophysical characteristics of this channel differ from those of T-, N- and L-type channels. In addition, this channel is not sensitive to DHPS or to -conotoxin GVIA, but is sensitive to a polyamine-like toxin from venom of funnel web spiders (Llinas et al., 1989). Using a synthetic funnel-web spider toxin (FTX), proteins have been isolated from mammalian cerebellar tissue which formed FIX-sensitive channels when incorporated into bilayers. Furthermore, the molecular mass of this protein markedly differed from that of other characterized Ca2+ channel proteins (Cherksey er al., 1990) This channel is referred to as the P channel since it was first described in Purkinje cells. Ca2+ current from rat brain mRN A expressed by Xenopus oocytes has a single exponential time course of inactivation of 22 approximately 650 ms and is not sensitive to DHPS nor to -conotoxin GVIA (Leonard et al., 1987), but has recently been reported to be partially sensitive to FTX (Lin et al., 1990b). These results suggest that the channel type expressed from rat mRNA may be similar to the P-channel. The biophysical and pharmacological profiles of the different Ca2+ channel types are summarized in Table 1. Pharmacological differences also exist between the Ca2+ channel subtypes and have been instrumental in distinguishing among the putative subclasses. The dihydropyridine (DI-IP)-class of organic Ca2+ channel agonists and antagonists interact preferentially with L-type channels, whereas N - and T-type channels are unaffected by nanomolar to low micromolar concentrations of these drugs. The peptide, to -conotoxin GVIA, (Cng) acts on a specific population of sensitive N-type channels in mammalian neurons. However, in amphibian and avian preparations, Cng also blocks L-type channels, as defined by voltage sensitivity and current kinetics. In some cases, block of L channels by Cng may be reversible or incomplete. Inorganic divalent cations block all three types of Ca2+ channels. However, the relative potencies of block differ: N and L channels are blocked more readily by Cd2+ than N iz", whereas T channels are blocked more readily by N i2+ than Cd2+ (Tsien et al., 1988). It should be noted that slight differences in electrophysiological properties and / or pharmacological sensitivities, such as the voltage at which current by a channel type increases, or the time constant of inactivation of whole-cell current exist for given Ca2+ channel types from one cell type to the next. In addition, since Ca2+ 23 Table 1.1 Biophysical and Pharmacological Characteristics of Ca2+ channel Types T N L P Activation Range (mV) -70 -10 -10 -30 to ~20 Inactivation Kinetics fast moderate slow slow r=20-50ms r=95ms r>.555 1:? Divalent Cation NI2+ > Cd?“ Cd2+ > N17” Cd" > N?‘ Cdz“, <100pM sensitivity DHPS - - + - to Cng - + + a _ FIX - - - + Single channel conductance (pS) 8 13 25 10-12 (110 mM Ba“) (80 mM Ba“) Data for chick DRG N,L and T channels modified from Tsien er al., TINS 10:431—438, 1988. P-channel data compiled from Llinas eta1., Proc. Natl. Acad. Sci. USA 86: 1689-1693, 1989 and Lin et al., Pmc. Natl. Acad. Sci. USA 87:4538—4542, 1990b. " Mammalian L-type channels are not sensitive to Cng (Plumber er al., Neuron 2:1453-1463, 1989). 24 channel function often depends on the presence of intracellular components (Bean, 1985), channel activty often changes or is lost during whole-cell or cell-free patch recording. This has hampered charactization of Ca2+ channel function and types. Thus, the characteristics attributed to Ca2+ channel types are more descriptive than absolute, as Ca2+ channel characteristics do not always fit neatly into one of the above categories or nomenclature systems. However, it is possible that there is considerable heterogeneity in Ca2+ channel function among different cell types or that additional channel types exist. This is supported by the recent isolation from rat brain of 4 different types of cDNA that code for putative Ca2+ channels; two of which are highly homologous (75% identical residues) with the a 1 subunit of the DI-IP-sensitive Ca2+ channel from skeletal muscle, whereas the other two are much less homologous (50% identical residues) but nonetheless related in sequence (Snutch et al., 1990). Characterization of the primary structure and expression of these cDNAs suggest that several different Ca2+ channel gene families exist which are related in amino acid sequence, but may have subtle differences in pharmacological sensitivity (Starr et al., 1990; Dubel and Snutch, 1990). Despite considerable research characterizing neuronal Ca2+ channels, the channel type(s) involved in neurotransmitter release are still not defined clearly. This is largely due to the technical limitations of current techniques. Intracellular and patch clamp recording require the use of microelectrodes whose diameter is almost as large as the nerve terminal itself. Thus, biophysical characterization of channel types at the active zone is not possible. Exceptions to this problem are the terminals 25 of the neurohypophyseal neurons, which are large enough to allow use of patch- clamp methodology. It has been reported that both L and N type channels are present in these terminals, and that both may contribute to Ca2+ entry associated with transmitter release. Under normal conditions of release, L channels are more likely to be involved in release than N channels, due to rapid inactivation of the latter during trains of action potentials (Lemos and Nowycky, 1989). More recently, however, it has been reported that both Ca2+ channel types in this preparation may be involved in an initial burst of transmitters release, whereas L-type Ca2+ channels are important for sustained release of peptide transmitters (Lim et al., 1990; N owycky, 1991). Although they provide the only mammalian system in which Ca2+ currents can be examined in nerve terminals, these neurons are highly specialized structures which secrete the neurohormones oxytocin and vasopressin. Release of peptide neurotransmitters from these terminals may or may not be truly representative of release of small, classical transmitters from other central and peripheral neurons. Indirect information on the types of Ca2+ channels which mediate Ca2+ influx associated with neurotransmitter release can be obtained by examining the pharmacological sensitivity of neurotransmitter release. At the rat neuromuscular junction, DHP-sensitive Ca2+ channels can have at best a modulatory role in neurotransmitter release (Atchison and O’Leary, 1987), but do not contribute to nerve-evoked release under normal conditions (Atchison, 1989). DHP-sensitive or L- type channels have been implicated in release of substance P from DRG cells (Rane 26 et al., 1987; Holz et al., 1988) and 3H-norepinephrine release from undifferentiated PC12 cells (Kongsamut and Miller, 1986). DHP-insensitive or N-type channels have been implicated as the important channel involved in neurotransmitter release in rat sympathetic neurons (Hirning et al., 1988), peripheral neurons (Perney et al., 1986), nerve growth factor differentiated PC12 cells (Kongsamut and Miller, 1986), various areas of the central nervous system (Dooley et al., 1987a). These data would suggest that N- and/ or L-type Ca2+ channels are responsible for Ca2+ influx associated with neurotransmitter release at many terminals. However, substantial data exists which suggests that another as yet uncharacterized Ca2+ channel subtype may also mediate Ca2+ entry associated with neurotransmitter release. For example, the Ca2+ current associated with neurotransmitter release from the presynaptic terminal of the squid giant synapse activates at relatively hyperpolarized potentials, inactivates slowly, and is insensitive to DHPS and Cng (Charlton and Augustine, 1990). This suggests the possibility that P-type channels may also be involved in transmitter release. In mammalian preparations, Cng and DHPS do not affect release of ACh from motor nerve terminals (Anderson and Harvey, 1987) or aspartate release from hippocampal slices (Mangano et al., 1991). Transmitter release from rat sympathetic ganglion cells is not impaired by DHPS or Cng (Seabrook and Adams, 1989). Species and / or tissue differences in the type Ca2+ channel involved in neurotransmitter release may also exist in mammals, as in vitm Cng produces a greater efiect on transmitter release in smooth muscle preparations from guinea pigs than from rats (Maggi et al., 1988). Wessler er al., 27 (1990) demonstrated in rats that 3H-ACh release from myenteric plexus neurons and the phrenic nerve is not affected by verapamil or diltiazem, whereas Cng reduces release from the neocortex (ICso = 13 nM) and myenteric plexus (ICso = 0.7 nM), but not the phrenic nerve (ICso 2 100 nM). These results suggest that channels which are neither N - nor L-type are involved in neurotransmitter release from some mammalian nerve terminals. Therefore, at many terminals it appears as though N - and/ or L-type channels may mediate Ca2+ entry associated with neurotransmitter release, although additional Ca2+ channel subtypes may also be involved in neurotransmitter release. There is no evidence to date which indicates that the T- type Ca2+ channel is involved in release in any system, but this does not preclude an unknown role of this channel. 5) Environmental contaminants, natural toxins and pharmaceutical agents which affect Ca2+ channels. Ca2+ channel function in neurons is affected by a wide variety of chemical agents including metals, peptides and organic compounds. Inorganic heavy metal cations are effective Ca2+ channel blockers. Lead (Audesirk, 1987 ; Audesirk and Audesirk, 1989) and cadmium, although the latter is not overtly neurotoxic, are environmental contaminants which block Ca2+ channels at concentrations between 1 and 10 micromolar. The transition metals nickel, mercury, copper and zinc are also capable of blocking Ca2+ channels at low micromolar concentrations (Nachshen, 1984). In isolated nerve terminals, block of 45Ca’“ influx by divalent cations such as 28 Pb“, Ni“, Cd2+ and Hg2+ (Nachshen, 1984; Atchison et al., 1986) is overcome by increasing [Ca2+]e. Thus, it is thought that these metals bind to a site in the pore of the Ca2+ channel with high affinity, impeding the flow of Ca2+ through the channel (Hagiwara and Takahashi, 1967; Hagiwara et al., 1974; Lansman et al., 1986; Hess et al., 1986). In addition to inorganic metals, the organic metals methylmercury (Atchison et al., 1986; Shafer and Atchison, 1989) and ethylmercury (Hewett and Atchison, 1990) are capable of blocking 45Ca2+ influx into nerve terminals. Both charge and lipophilicity may be important in determining the characteristics of Ca2+ channel disruption by mercurials, as only positively charged mercurials block Ca2+ channels, and only organic mercurials blocked “Ca2+ influx into synaptosomes in a voltage- dependent manner (Hewett and Atchison, 1990). Several natural toxins which interact Specifically with Ca2+ channels have been isolated from venom preparations. A family of polypeptide toxins with Ca2+ channel activity has been isolated from venom of marine snails of the Conus family (Olivera er al. , 1987). to -Conotoxin GVIA, isolated from the venom of Conus geographus (Olivera er al., 1985 ), has been thoroughly characterized with respect to its action on Ca2+ channels. This compound is a potent inhibitor of non-mammalian Ca2+ channels, binding to and blocking N— and L-type Ca2+ channels irreversibly in sub- picomolar concentrations (Cruz and Olivera, 1986). In mammalian preparations, the action of to -conotoxin GVIA is less clear. It has been suggested that only the N-type Ca“ channel is sensitive to Cng in several preparations (Tsien er al., 1988; Plummer 29 et al., 1989). This peptide blocks neurotransmitter release (Keith et al., 1989) and synaptic transmission (Kamiya er al., 1988; Dutar et al., 1989) in the hippocampus. However, Cng does not affect ACh release at the neuromuscular junction (Anderson and Harvey, 1987), and only partially reduces 45Ca2+ influx into mammalian synaptosomes (Suszkiw et al., 1989). In addition, expression of rat brain mRNA in oocytes results in Ca2+ currents which are insensitive to Cng (Leonard et al., 1987). As discussed in the previous section, these results suggest a unique Ca2+ channel type in these preparations. Maitotoxin, which is produced by the marine dinoflagellate Gambierdiscus toxicus, increases Ca2+ channel function in cultured neuronal cells (Takahashi et al., 1982; 1983). Spider venoms also contain peptides and polyamine-like compounds which block Ca2+ channels. The to -aga-toxins are peptides that have been isolated from the venom of funnel web spider (Agelenopsrll~ aperta) and are potent blockers of Ca2+ channels (Adams et al., 1990). These toxins can be subdivided into three classes based on the ability to inhibit Cng binding in chick synaptosomes; to -AgaI toxins do not affect binding, whereas at -AgaII and to -AgaIII (Venema and Adams, 1990) toxins inhibit Cng binding. (1) -AgaIIIA blocks N- and L-type channels in rat DRG cells (Minty et al., 1990). However, the pharmacological characterization of these compounds has not been completed to date. A polyamine-like toxin which is a potent Ca2+ channel blocker also has been purified from the venom of funnel-web spiders (Llinas et al., 1989). The structure of this toxin has not yet been determined, although several compounds with Similar activities have been synthesized (Cherksey 30 et al., 1989). Pending further characterization of Spider toxins, these compounds may be useful tools in neurobiology. Finally, the endogenous peptide endothelin has been demonstrated to bind to and depolarize dorsal horn neurons (Yoshizawa et al., 1989). Endothelin is a vasoactive peptide which is presumed to increase function of L-type Ca2+ channels in muscle cells. That it has actions in the spinal cord may indicate that there are also endogenous regulators of Ca2+ channel activity in the nervous system as well as the cardiovascular system. Several therapeutic compounds also exert Ca2+ channel block as either a predominant mechanism of therapeutic action, or at least one of several potential mechanisms of therapeutic and/ or toxic action. Prominent among these compounds are several agents used for treatment of cardiovascular disorders such as angina pectoris, hypertension and arrythymias. These so called "Ca2+ channel blockers" also bind readily to CNS preparations (Gould et al., 1985). Therapeutic Ca2+ channel antagonists fall into three chemical classifications: the dihydropyridines (DI-1P) such as nifedipine, the phenylalkylamines such as verapamil and the benzothiazopines, such as diltiazem. DHPS with Ca2+ channel agonist properties have also been synthesized. These are represented by (+ )-Bay K 8644 and (+ )-S-202-791. The DHPS, and presumably the phenylalkylamines and benzothiazopines, interact preferentially with btype Ca2+ channels as opposed to other Ca2+ channel types (Fox et a1. , 1987a,b). Other therapeutic agents whose action is postulated to involve Ca2+ channel block include anticonvulsants such as phenytoin (Twombly et al., 1988), the barbiturates (Blaustein and Ector, 1975; Werz and Mcdonald, 1985; Gross and 31 Mcdonald, 1988; Gunderson et al., 1988), as well as the benzodiazepines (Leslie, 1987) and certain general anesthetics such as halothane (Takenoshita and Steinbach, 1991). The neuroleptic chlorpromazine blocks Ca2+ influx into synaptosomes (H085 and Forrnaniak, 1984) and Ca2+ currents in neuroblastoma cells (Ogata and Narahashi, 1990). Finally, the aminoglycoside antibiotics may also interact with neuronal Ca2+ channels and this interaction may contribute to the toxicity of these therapeutic agents. In synaptosomes, neomycin and polymyxin block synaptosomal Ca2+ channels in a manner which is reversed by increasing extracellular [Ca2+] (Atchison et al., 1988) and aminoglycoside antibiotics compete with binding of radiolabelled m-conotoxin GVIA (Knaus et al., 1987). Low concentrations of ethanol (EtOH) decrease 4SCaz“ influx into synaptosomes (Friedman et al., 1980; Leslie et al., 1983; Skattebol and Rabin, 1987) and block Ca2+ currents in neuroblastoma cells (Twombly et al., 1990). Chronic exposure to low millimolar concentrations of EtOH results in an increase in depolarization-induced 45Ca2" influx into PC12 cells. This increase in Ca2+ influx is presumably channel mediated, as it is blocked by DHPS and inorganic divalent cations (Greenberg et al., 1987a). In addition, chronic exposure to ethanol alters the binding of Ca2+ channel ligands to PC12 cells (Marks et al., 1989). CHAPTER TWO SYNAPTOSOMES AND PHEOCHROMOCYTOMA CELLS AS MODEL SYSTEMS FOR STUDY OF Ca2+ CHANNELS 32 33 A) SYNAPTOSOMES AND SYNAPTOSOMAL Ctr2+ CHANNELS Because of the limitations imposed by the small size of most nerve terminals, direct electrophysiological study of channel function is severely restricted. Synaptosomal suspensions are preparations containing large numbers of pinched-off nerve terminals which allow for biochemical study of channel function and neurotransmitter release in vitro. Synaptosomes are derived by homogenization of central nervous system tissue followed by differential centrifugation to separate synaptosomal material from other CNS components. Synaptosomes are spherical particles 1 - 2 pm in diameter and contain intracellular organelles including mitochondria, endoplasmic reticulum and synaptic vesicles (Gray and Whittaker, 1962). Many metabolic pathways and synthetic pathways remain intact in synaptosomes, including glycolysis, Kreb’s cycle and neurotransmitter synthesis (Suszkiw and O’Leary, 1982; Eder-Colli et al., 1989). Synaptosomes when incubated in glucose-containing medium remain viable for biochemical studies of metabolic function, channel function and neruotransmitter release for several hours after isolation (Bradford, 1975). The activity of Na"/K+ ATPase is very high in synaptosomal preparations, allowing for maintenance of a membrane potential between -30 and -80 mV (Blaustein and Goldring, 1975 ; Suszkiw et al., 1986). Synaptosomes also contain voltage-sensitive Na‘, K+ and Ca2+ channels (Nelson et al., 1983; Nelson, 1985; 1986) and release neurotransmitter in a Ca2*-dependent manner following depolarization (Blaustein, 1975; Floor, 1983; Leslie, er al., 1985). This preparation has been widely employed in studies of channel function (N achshen 34 and Blaustein, 1979; 1980; 1982), neurotransmitter synthesis and release, receptor binding (see Chapter four), and regulation of metabolic processes (Blaustein et al., 1978) at the nerve terminal. Results of experiments using this preparation have contributed Significantly to knowledge of processes that occur at this portion of the neuron. Depolarization-dependent Ca2+ influx into synaptosomes occurs in two distinct phases (Nachshen and Blaustein, 1980; Suszkiw and O’Leary, 1983; Nachshen, 1985): a "fast" phase, which inactivates after 1-2 s of depolarization and a "Slow" phase which remains active for 20-90 s after depolarization (Figure 2.1). The slow phase of influx may be mediated by a Ca2+ channel and a reversed N a+ / Ca2+ exchange mechanism; the latter may predominate, as removal of Na+e largely reduces "slow" influx (Coutinho et al., 1984; Turner and Goldin, 1985; Suszkiw et al., 1986). The fast phase of influx, which is associated with neurotransmitter release (Drapeau and Blaustein, 1983; Floor, 1983; Suszkiw and O’Leary, 1983; Leslie et al., 1985), is presumably mediated by voltage-dependent Ca2+ channels and is not dependent on Na“,. In many ways, 4sCa“ influx into synaptosomes correlates with Ca“ channel function in other systems. Depolarization of synaptosomes by K” results in influx of 45Ca“ in a manner which is related to the [10],. Preng that [K+]e/ [10]i (Blaustein and Goldring, 1975) determines the membrane potential, approximately 20-40 mV depolarization is required to initiate 45’Ca“ influx (Nachshen and Blaustein, 1980; Suszkiw er al., 1986; 1989; Adam-Vizi and Ligeti, 1986). Predepolarization of the synaptosomal suspension in the presence or absence Ca“ 032+ Figure 2.1. Components of Ca2+ influx into synaptosomes. Schematic diagram of a synaptosome containing mitochondria (M; and synaptic vesicles (SV). Depolarization of the membrane (A EM) results in fast, channel-mediated Ca + influx (FAST) and slow influx (SLOW). The slow component is mediated by reversed Na” /Ca2+ exchange and slowly- or non-inactivating Ca "’ channels. 36 of Ca2+ eliminated the fast component (Nachshen and Blaustein, 1980). Ca2+ channels in synaptosomes also undergo both voltage- and Ca2+-dependent inactivation (Suszkiw et al., 1986; 1989). In addition to Ca“, Sr2+ and Ba2+ are also permeant through these Ca2+ channels; the relative influx of Caz“:Sr2":Ba2+ for the fast component of synaptosomal influx is 6:3:2 (Nachshen and Blaustein, 1982). Incorporation of synaptosomal Ca2+ channels into lipid bilayers indicated that the single channel conductance was gBa2+ > gCa“ =gSr2+ (Nelson, 1986). This is similar to cardiac Ca2+ channels, for which whole cell current indicated that the relative premeability was Ca2+ > Sr2+ > Ba“, but single channel conductance measurements indicated that gBa2+ >gCa2+ (Hess et al., 1986). Like Ca2+ channels in other systems, fast 45Ca2+ influx into synaptosomes is sensitive to block by inorganic divalent and trivalent cations. The most potent blockers of influx are La“, Pb“ and Cd“, which have Ki values of less than 10 11M. Other metals including Co“, Ni“, Hg2+ and Zn2+ also block the fast component of infux, although higher concentrations (Ki = 10 - 100 nM) are required (Nachshen, 1985). The organic Ca2+ channel blocker verapamil also decreases 45Ca2+ influx into synaptosomes (N achshen, 1985). Single-channel recordings of synaptosomal Ca2+ channels in lipid bilayers also confirm that La”, Cd2+ and verapamil are potent blockers of synaptosomal Ca2+ channels (Nelson, 1985 ). Block of Ca2+ channels by verapamil suggests the possibility that L-type channels may be involved in Ca2+ influx. However, this is a subject of considerable controversy, as DHPS do not have consistent effects on ”Ca2+ influx in synaptosomal preparations (see below). 37 Fast influx has been associated with release of Substance P (Floor, 1983), acetylcholine (Suszkiw and O’Leary, 1983) and dopamine (Drapeau and Blaustein, 1983, Leslie et al., 1985) from synaptosomes. Thus, the fast phase couples Ca2+ entry and subsequent neurotransmitter release from nerve terminals. The slow component of ”Ca2+ influx into synaptosomes differs in many respects from the fast component. It does not inactivate for up to 90 5 during depolarization and is much less sensitive to La3+ than the fast phase (N achshen and Blaustein, 1980). Furthermore, the relative flux of Caz“:Sr2“:Ba2+ for the slow component is 6:3:1 (Nachshen and Blaustein, 1982), indicating that this component has different properties of selectivity than the fast component. Finally, this component is extremely sensitive to [N a"]e (Blaustein and Oborn, 1975). Inasmuch as preparation and incubation of synaptosomes in Na”-free solutions drastically reduces influx via this component, it has been proposed that a reversed Na+/Ca2+ exchange contributes significantly to this phase of influx (Blaustein and Oborn, 1975; Akerman and N icholls, 1981; Coutinho et al., 1984; Turner and Goldin, 1985; Suszkiw et al., 1986). This does not preclude involvement of Ca2+ channels in this component, as small fluxes are measured in the absence of Na+ (Suszkiw et al., 1986; Shafer and Atchison, 1989) and DHPS have been reported to decrease this component of influx in Na+-containing solutions (Martinez-Serrano et al., 1989). Initial experiments in this dissertation examine effects of MeHg on synaptosomal “Ca2+ influx Thus, information regarding Ca2+ channel types in this system is important. Synaptosomes are likely to contain more than one type of Ca2+ 38 channel, but the types of Ca2+ channel present have not been identified. DHP antagonists bind to synaptosomal preparations with high affinity (Rampe, 1984, Turner and Goldin, 1984; Suszkiw et al., 1986; Dooley et al., 1987b; Massieu and Tapia, 1988). However, synaptosomal ”Ca2+ influx and / or neurotransmitter release have been variously reported to be sensitive (Turner and Goldin, 1985 ; Dunn, 1988; Martinez-Serrano, 1989) or insensitive (Nachshen and Blaustein, 1979; Daniel] et al., 1983; Reynolds et al., 1986; Rivier et al., 1987; Suszkiw et al., 1986; 1989; Massieu and Tapia, 1988) to block by DHP antagonists. Additionally, in mammalian synaptosomes, 4SCa“ influx is only partially sensitive to block by to -conotoxin GVIA (Suszkiw et al., 1989). Furthermore, the activation, inactivation (Suszkiw et al., 1989) and pharmacological sensitivity (Suszkiw et al., 1989; Lundy et al., 1991) of 45Ca2+ influx into synaptosomes suggest that these preparations contain a high voltage activatid channel which is neither N- nor L-type. In some respects, Ca2+ influx into synaptosomes has characteristics similar to current mediated by P-channels. However, further characterization is required to verify the type or types of Ca2+ channels present in synaptosomal preparations. B) Ca” CHANNEL FUNCTION IN PC12 CELLS PC12 cells contain many voltage-sensitive channel types important to neuronal function. These include tetrodotoxin- (TT'X) sensitive and -insensitive N a” channels (Rudy ct al., 1987), voltage-sensitive K” channels (Pun and Behbehani, 1989), and Ca2+ channels. PC12 cells have been utilized extensively to study the functions, 39 expression and regulation of Ca2+ channels. Furthermore, this cell line has been employed for numerous studies of neuronal function. AS such, many biochemical and phsyiological processes of this cell line are well understood. Neurotransmitter release and receptor function in PC12 cells are reviewed in the Appendix. By contrast to synaptosomes, the type of Ca2+ channels present in PC12 cells are much better defined. It is generally agreed that NGF-differentiated PC12 cells contain both DHP-sensitive L- (Figure 2.2, top panel) and to -conotoxin GVIA- sensitive N- (Figure 2.2, bottom panel) type Ca2+ channels (Plummer et al., 1989; Janigro et al., 1989; Usowicz et al., 1990; Shafer and Atchison, 1991). The presence of T-type Ca2+ channels in differentiated PC12 cells also has been reported (Garber et al., 1989), although pharmacological characterization is required to confirm the presence of T-type channels in PC12 cells. I The existence of DHP-sensitive (L-type) channels in both undifferentiated and NGF-differentiated cells was established by both ion flux assays (Stallcup, 1979; Toll, 1982; Takahashi and Ogura, 1983; Takahashi et al., 1985) and electrophysiological studies (Streit and Lux, 1987; Plummer et al., 1989; Janigro er al., 1989; Usowicz gt a], 1990; Shafer and Atchison, 1991). The presence of DHP-insensitive (N-type) channels in PC12 cells has been more difficult to establish. Release of 3I-I-NE from PC12 cells grown in NGF was less sensitive to DHPS than was release from undifferentiated PC12 cells (Takahashi er al., 1985; Kongsamut and Miller, 1986), although 4"’Caz“ influx remained largely DHP-sensitive (Kongsamut and Miller, 1986). Depolarization of undifferentiated PC12 cells by K" resulted in Ca2+-dependent 4o Figure 2.2. N and L type Ca2+ channels in PC12 cells. DHP—sensitive and ~insensitive single Ca2+ channel current in differentiated PC12 cells recorded using the cell-attached patch voltage-clamp technique. (a) current records from a patch containing one 'L-type” and one "N-type" channel. Four consecutive sweeps are shown for each set of depolarizations from two difierent holding potentials, as indicated in the voltage protocol above the sweeps. (b) Current traces from a patch containing a single "N-type' channel In each case, the DHP agonist (+)-(S)-202-791 is present at 1 uM, and Ba2+ (110 mM) is the charge carrier. From Plummet er at. M 2:1453-1463 (1989). 41 release of transmitter which is sensitive to DHP Ca2+ channel agonists and antagonists (Figure 2.3). Differentiation with NGF resulted in decreased sensitivity of neurotransmitter release to DHP drugs (Figure 2.3B,C). Indeed, in most 45Ca2+ flux studies in PC12 cells, no DHP-insensitive components were reported with the exception of a small (< 10%) component in one study (Shafer et al., 1990). Although this latter study and studies of neurotransmitter release indicated expression of DHP- insensitive or N-type Ca2+ channels in differentiated PC12 cells, they did not provide conclusive evidence that N-type channels were indeed being expressed. By contrast to radiolabel flux experiments, expression of DHP-insensitive or N -type channels has been demonstrated in NGF-differentiated PC 12 cells using both whole-cell (Plummer et al., 1989; Usowicz _e_t al, 1990; Shafer and Atchison, 1991) and cell-attached (Plummer et al., 1989) patch voltage-clamp techniques. In undifferentiated PC12 cells, small N -type current components have been reported (Plummer er al., 1989; Janigro et al., 1989). In addition to flux and electrophysiological measurements on Ca“ channel function, numerous studies of Ca2+ channel ligand binding have been performed to identify and characterize pharmacologically the Ca2+ channel subtypes expressed in PC12 cells. The most commonly used ligands are the DHP (+)-PN 200-110 and the peptide to -conotoxin GVIA. Differentiation of PC 12 cells increases the number of conotoxin binding sites by 2-3 fold and increases the number of DHP binding Sites by less than 2 fold (Usowicz 51 a1, 1990). In PC12 cells, conotoxin is specific for N- type channels, as there is no interaction between DHP binding and conotoxin binding 42 > “g 20' 1 if?) 15' < 5 ‘5' 5" 10' i" 1' : \E g g :5”, “‘1. I'- 5 ’ J’ \‘E. I. O 5 10 15 20 25 g TIME (min) 250[ B ’3 120[ C -e- -NGF xi" \ I, \\ I ' \ _. zoo. / \\ 5‘ 80. O r \ I E “ " ‘1 a O 150. o . 100 e g ~ ~4 0 0.010.1 1 10 BAY K8644 (III) NITRENDIPINE (I‘M) Figure 2.3. Nore inephrine release from P012 cells. Release of [3H]-norepincphrine from PC12 cells. A.) Release of [ -norepinephrine from PC12 cells expressed as a percentage of tissue stores present at the beginning of the particular fraction. The first two 5 min. fractions were release in non-depolarizing K" solutions, with or without incubation of drugs. The third fraction was release induced by depolarization with 70 mM K+ solution. The final two fractions are release in non-depolarizing solutions. (diamond) 70 mM K+ alone, (square) 70 mM K+ plus 10" M Bay K 8644, (circle) 70 mM K+ plus 1045 M nitrendipine, (triangle) 70 mM K+ Ca2+-tree plus 02 mM EGTA. B. and C.) Dose- response relationship for efl’ects of Bay K 8644 (B) and nitrendipine (C) on release of NE from N GF- differentiated (open circle) and undifierentiated (solid circle) PC12 cells. Results are expressed as a percentage of control-stimulated release in depolarizing solutions. Values are the meanzi: SEM of 6-12 quJeriments. Modified from Kongsamut and Miller, Proc. Natl. Acad. Sci. (U.S.A.) 83:2243-2247 (1986). 43 (Usowicz et al., 1990), and conotoxin blocks only the inactivating component of N- type current (Plummer er al., 1989). In an NGF-insensitive PC12 mutant line, no changes in Ba2+ current or conotoxin binding were observed in response to NGF (Usowicz et aL, 1990). In addition, high concentrations of ouabain (0.1 mM), which did induce morphological differentiation of the mutant line, did not induce changes in Ca2+ channel expression (Usowicz et al., 1990). These results suggest that NGF- differentiation induces increased expression of N-type Ca2+ channels in PC12 cells, and that morphological differentiation and changes in Ca2+ channel expression are not necessarily co-requisite events. As indicated by binding studies, the proportion of N- to L- type channels changes with differentiation by N GF. PC12 cells grown in absence of N GF express predominantly L-type Ca2+ channels, whereas those treated with N GF exhibit a substantial increase in the expression of N-type Ca2+ channels without a concomitant decrease in expression of L-type channels (Streit and Lux, 1987; Plummer et al., 1989; Usowicz 31 a1, 1990). Ca2+ channels are regionally distributed in differentiating PC12 cells; growth cones contain a higher density of Ca2+ current than neurites and Ca2+ current Shows more inactivation in grth cones than in the soma (Streit and Lux, 1989). Streit and Lux conclude that the increase in the inactivating component of current (which resembles currents mediated by N -type channels) with N GF treatment may result from changes in the distribution, but not types (eg. N and L) of Ca2+ channels in PC12 cells (Streit and Lux, 1990). However, this conclusion is based on 44 whole cell recordings and does not consider pharmacological sensitivity to conotoxin of Ca2+ currents in PC12 cells nor single channel current kinetics. The apparent differential expression of Ca2+ channel subtypes in PC12 cells is important for two reasons. First, N— and L-type Ca2+ channels have been suggested to modulate or possibly even to mediate Ca2+ entry associated with neurotransmitter release from several neuronal preparations (Atchison and O’Leary 1987; Rane et al., 1987; Holz et al., 1988; Atchison, 1989; Lemos and Nowycky, 1989). Second, one can control to some extent, by use or omission of NGF, the proportion and distribution of N- vs L-type channels in PC12 cell cultures. This change induced by NGF may help to define whether N- and L-type channels are indeed separate entities or whether the biophysical and pharmacological differences arise from different "functional states” of one channel type. There is some evidence that L-type Ca2+ channels in PC12 cells are regulated by second messengers, including protein kinase C (PKC) and G-proteins. Activators of PKC reduce DHP-sensitive ”Caz" influx into PC12 cells (Harris et al., 1986; Di Virgilio et al., 1986), while inhibitors of PKC have little effect on influx (Greenberg er al. , 1987b). Calmidazolium and trifluoroperazine, inhibitors of calmodulin, decrease influx of 45Ca“ into PC12 cells, indicating that calmodulin or calmodulin- dependent enzymes may also regulate Ca2+ channels (Greenberg et al., 1987b). G- proteins may also be involved in Ca2+ channel regulation. Stable GTP analogs increase the displacement PN 200-1 10 binding by the DHP agonist Bay K 8644. Pertussis toxin inhibits this action of Bay K 8644, as well as Bay K 8644-induced 45 potentiations of depolarization-dependent 4SCaz“ influx into PC 12 cells (Bergamaschi et aL, 1990). Chronic K*-induced depolarization or treatment with ionomycin, a Ca2+ ionophore, reversibly decreased depolarization-induced 45Ca2“ influx and the number of [3H]-nitrendipine (a DHP antagonist) binding Sites (DeLorme et al., 1988). Elevations in [Ca2+], may be involved in Ca2+ channel down regulation, although [Caz‘i]i returned to normal levels while Ca2+ channel function remained depressed (DeLorme and M‘Gee, 1988). One second messenger which does not appear to affect Ca2+ channel function in PC12 cells is cAMP (Rabe et al., 1982; Nishizawa et al., 1990). Further characterization of the regulation of Ca2+ channels by second messengers is needed to understand more fully the role of Ca2+ channels in neurotransmitter release and other cellular functions in PC12 cells. In general, PC12 cells have not been used extensively as a tool in neurotoxicological studies, although the interactions of several neurotoxic agents with Ca2+ channels have been examined. PC12 cells have been used to characterize interactions of the neurotoxicant methylmercury (MeHg) with Ca2+ channels. MeHg blocks 4"Ca2+ influx into undifferentiated and N GF-differentiated PC 12 cells. In addition, MeHg alters the binding of conotoxin to intact PC12 cells (Shafer et al., 1990). In more recent studies, acute effects of MeHg on current carried by Ba2+ (13,) in NGF-differentiated PC12 cells have been examined. MeHg blocks both N- and L- type Ca2+ current in PC12 cells in a manner which is voltage- but not State- dependent (see chapter 5). The effects of MeHg on Ca2+ channels are not reversed by washing with MeHg-free solutions, but are antagonized by increasing extracellular I Hi ml ll... I' :- "‘\ ,r the 46 [Bah] (Shafer and Atchison, 1991). These most recent results suggest that the effects of MeHg on 13, in PC12 cells mirror the actions of MeHg on synaptosomal 4sCa“ influx (Atchison et al., 1986; Shafer and Atchison, 1989; Shafer et al., 1990). EtOH also alters Ca2+ channel function in PC12 cells. Chronic exposure to low millimolar concentrations of EtOH results in an increase in depolarization-induced 45Ca2+ influx into PC12 cells. This increase in Ca2+ influx is presumably channel mediated, as it is blocked by DHPS and inorganic divalent cations (Greenberg et al., 1987a). Furthermore, chronic exposure to EtOH increased the number of (+ )-PN 200-110 binding sites in PC12 cell membranes (Figure 2.4). This action was not reduced by co-treatment with Bay K 8644, indicating that increased channel number was not the result of decreased Ca“ influx into PC12 cells (Marks et al., 1989). Further study of the effects of EtOH on Ca2+ channel numbers in PC12 cells may provide insight into the mechanisms of regulation of expression of Ca2+ channels in PC12 cells. Chronic exposure (3 days) to immunoglobulin G isolated from serum of patients with Lambert-Eaton Myasthenic Syndrome has also been demonstrated to decrease 45Ca2“ influx into PC12 cells (Lang, et al., 1989). The effects of volatile anesthetics (Kress er al., 1987), maitotordn (Takahashi et al., 1982; 1983; Choi et al., 1990) and a- latrotoxin (Vicentini and Meldolesi, 1984; Meldolesi et al., 1986) on Ca2+ channel fUnction have also been examined using PC12 cells. Thus, PC12 cells are clearly useful models for the study of interactions between neurotoxic compounds and Ca2+ channels. 47 .p (A N -A PN 200-110 bound (truer/10‘ cells) 0 O l L I l 100 200 300 400 500 PM zoo-110 (pM) Figure 2.4. Effects of EtOH on Ca2+ channel density in PC12 cells. Saturable binding of the DHP Antagonist (+)-[3H]PN 200-110 to intact PC12 cells grown for 6 days in the absence (open circles) or presence (closed circles) of 200 mM ethanol. Data shown are the average of triplicate values from a single representative experiment. From: Marks er at. W 53:168-172 (1989). CHAPTER THREE BLOCK OF 45C3“ INFLUX INTO SYNAPTOSOMES BY NIETHYLMERCURY: Ca2+ AND Na+ DEPENDENCE 48 49 ABSTRACT Block of Ca2+ influx into isolated nerve terminals by the neurotoxicant methylmercury (MeHg) was studied for its dependence on extracellular Ca2+ and Na‘. Depolarization-independent entry of ”Ca“ was determined in rat forebrain synaptosomes incubated in 5 mM K+ solution. 4SCaz” influx was Similarily measured following 1 ("fast" phase) or 10 5 ("total") of elevated K+ - (41.25 mM) induced depolarization, or during 10 S of elevated K+-induced depolarization after synaptosomes had been predepolarized for 10 s in Ca2+- and MeHg-free solutions ("slow" phase). In 5 mM K+ solutions, MeHg concentrations of 125 11M and greater Significantly reduced synaptosomal 45Ca2+ accumulation measured during 1 or 10 S of incubation. In K*-depolarized synaptosomes, the estimated [CS0 for block of "total", "fast" and "slow" 45Ca2+ influx was 75 nM: 250 uM MeHg reduced influx by approximately 90%. The ability of [Ca2+],: to antagonize block was tested by increasing [Ca2+]e from 0.01 to 1.15 mM in the presence and absence of MeHg. When compared to control, 50 11M MeHg reduced "total" influx of 45Ca""' by 270% and reduced "fast" influx by 20-60% at all [Ca2+], tested. At [Ca2+], concentrations of 0.01-0. 15 mM, MeHg (50 nM) reduced "slow" influx by 75-90%, but did not affect Slow influx at higher [Ca2+]e (2 0.30 mM). When the dependence of block of 45Ca2+ influx on [Na‘i]c was tested, equivalent levels of inhibition were caused by MeHg (25 nM) for fast influx by synaptosomes in Na+-containing and NaI-free solutions. Conversely, 25 11M MeHg reduced slow influx by 26.1 14.5 and 96.01-4.0% in Nai- containing and N a“-free solutions, respectively. In Na*-free solutions, block of the 50 slow phase of 45Ca2+ influx was not overcome by increasing [Ca2+]c. Results of this study indicate that 1) MeHg affects Ca2+ influx into synaptosomes in the absence of depolarization; 2) block of the fast phase of 45Ca2+ influx by MeHg is only weakly antagonized by Ca”; and 3) block of fast 45Ca2“ influx was independent of [Na+]e; 4) block and reversal of block of the slow phase of 45Ca2+ influx by MeHg are highly Na+-dependent. 5 1 INTRODUCTION The hypothesis states that effects of MeHg on voltage-sensitive Ca2+ channels which are involved in neurotransmitter release at the motor nerve terminal may contribute to disruption of neurotransmitter release produced by MeHg. This would be best tested by simultaneously measuring Ca2+ entry and neurotransmitter release in intact motor nerve terminals. However, several limitations discussed in Chapter One preclude direct measurement of function of Ca2+ channels in intact motor nerve terminals. Although Ca2+ channel function in intact nerve terminals is difficult to examine using conventional electrophysiological techniques, biochemical measurement of Ca2+ channel function can be accomplished easily by measuring 45C2?“ flux into isolated nerve terminals, or synaptosomes. Thus, experiments to examine the ability and nature of MeHg interaction with nerve terminal Ca2+ channels were performed using synaptosomal preparations. Heavy metals are known to disrupt Ca2+ influx into nerve terminals leading to decreased neurotransmitter release. Divalent heavy metal cations such as Co“, Cu“, Zn“, Cd“, Pb2+ and Hg2+ block the influx of “Ca2+ into K+-depolarized synaptosomes in a reversible manner (N achshen, 1984; Suszkiw et al., 1985; Atchison et al., 1986). Micromolar concentrations of monovalent MeHg also block 45Ca2+ influx into Ki-depolarized synaptosomes (Atchison et al., 1986). In contrast to inorganic divalent cations, MeHg blocked 45Ca2" influx into synaptosomes during 10 s of depolarization in a manner which was not completely antagonized by increasing [C324]: (Atchison er al., 1986). However, the ability of increasing [Ca2+], to 52 antagonize block by MeHg of fast and slow components of 45Ca“ influx was not examined. Studies at the neuromuscular junction demonstrated that after increasing stimulus duration and/ or intensity, the subsequent block of nerve-evoked endplate potentials (EPPS) by MeHg could be reversed by increasing [Caz‘i]e (Traxinger and Atchison, 1987b). Since increasing [Ca2"]c can, under the conditions described above, reverse the effects of MeHg at the neuromuscular junction, and antagonize block of synaptosomal 45Ca2” influx during 10 S of K"-induced depolarization (total influx), experiments were designed to examine the ability of increasing [Ca2+], to antagonize block by MeHg of the fast and slow phases of Ca2+ influx into synaptosomes. When examining block by MeHg of the fast and slow phases of 45Ca2+ influx into synaptosomes, consideration must be given to the mechanisms mediating Ca2+ influx during each of the two phases. As mentioned previously, the fast phase is mediated by voltagesdependent Ca2+ channels, while the slow phase is mediated largely by a reversed Na+ /Ca2+ exchange mechanism. Because the two phases differ in their dependence on [Na*]c, it is possible that the block of ”Ca2+ influx by MeHg during the fast and slow phases may also exhibit a differential dependence on [Na+]¢. Thus, experiments were designed to test the dependence of block of each phase by MeHg on [Na+]e. MeHg is also capable of forming mercaptides with sulfhydryl groups (Klaassen, 1985) and reacting with amino, carbonyl and hydroxyl groups (Bidstrup, 1964), all of which can be constituents of membranes or membrane proteins. Because of its reactivity, MeHg can affect a variety of membrane-associated functions. MeHg alters 53 the resting membrane potential and blocks the action potential in the squid giant axon (Huneeus-Cox er al., 1966), and increases its threshold for excitation (Shrivastav et al., 1976). In addition, MeHg alters Na+- and Ki-channel function (Quandt et al., 1982; Shrager, 1977) and permeability (Passow and Rothstein, 1960). At rest, permeability of the membrane to Ca2+ may be mediated in part by random openings of Ca“ channels, active transport proteins or non-specific permeability. Therefore, effects of MeHg on depolarization-independent entry of 45Ca2" into synaptosomes were examined as a measure of the specificity of MeHg for Ca2+ channel-mediated transport. 54 MATERIALS AND METHODS Materials. Methylmercuric acetate was obtained from Pfaltz and Bauer, Inc. (Stamford, CT) and was dissolved in 4% (v/v) glacial acetic acid to make a 2 mM stock solution. Aliquots of this solution were diluted in the appropriate buffers and adjusted to pH 7.4 at room temperature (25 °C) to yield the final concentrations of MeHg indicated in the text of chapters 3 and 4. N-methyl—glucamine, ethyleneglycol- bis-(B-aminoethyl ether) N,N,N’,N’-tetraacetic acid (EGTA), and choline chloride were obtained from Sigma Chemical Co. (St. Louis, MO). N-2- Hydroxyethylpiperazine—N’-2-ethanesulfonic acid (HEPES) was purchased from United States Biochemical Corporation (Cleveland, OH). 4S’CaClz, 133BaCl2 and 858er (specific activities 15-47, 11 and 5.8 mCi/mg, respectively) were purchased from New England Nuclear Co. (Boston, MA). For incubation of synaptosomes, normal physiological saline contained (mM): NaCl, 145;KC1, 5; MgClz, 1; d-glucose, 10; and HEPES, 10. For 45Ca“ influx experiments, elevated Ki solution contained (mM): NaCl, 72.5; KCl, 77.5; MgClz, 1; CaClz, 0.04; d-glucose, 10; and HEPES, 10. Normal K+ (5 mM) solution contained the same constituents and amounts as did elevated K+ solution except that [K*] was lowered to 5 mM by replacement of KCl with 72.5 mM choline chloride. Ca2+-free solutions contained no added CaClz, but may have contained residual amounts of Ca2+ (1-2 I: M). In Na*-free experiments, N aCl was replaced in all solutions by choline chloride on an equimolar basis. Quenching solution contained (mM): KCl, 5; MgClz, 2; EGTA, 1; d-glucose, 10; HEPES, 10; and N-methylglucamine, 145. N -methylglucamine, which replaces Na+ 55 in the buffer solution, does not permeate the voltage-dependent N a“ channel (I-Iille, 1971). All solutions were adjusted to pH 7.4 at room temperature (25 °C). Synaptosomal preparation. Synaptosomes were prepared from forebrains of male Sprague-Dawley rats (Harlan, 175-200 gm) using a modification of the method of Gray and Whittaker, (1962) as described by Atchison et al. (1986, 1988). Forebrains were homogenized in 0.32 M sucrose and centrifuged for 10 min at 1000xg. The pellet was discarded and the supernatant was centrifuged at 17,500xg for 20 min. This pellet was resuspended in fresh 0.32 M sucrose and layered onto a discontinuous sucrose gradient of 0.8 and 1.2 M sucrose. After centrifuging for 2 h at 69,000xg, the synaptosome-enriched fraction was removed from the interface between the 0.8 and 1.2 M sucrose layers and diluted slowly 1:1 (v/v) with physiological saline. This step facilitates return of the synaptosomal preparation to physiological osmolarity. The synaptosomes were repelleted by centrifugation at 10,000xg for 10 min and then resuspended to an approximate concentration of 20—40 mg/ml in physiological saline (Lowry et al., 1951). This step returns the synpatosomal preparation to physiological osmolarity and removes sucrose from the solution. As the synaptosomal preparation was performed at 0-4 °C, synaptosomal suspensions were incubated in a 37 °C water bath for 20 min prior to their use. 45Ca2+ influx measurements. 4SCaz“ influx was measured as described by Nachshen and Blaustein (1980; 1982) by adding 501:1 of synaptosomal suspension (approximately 20 mg protein/ ml) to 50 pl of either elevated (77.5 mM) or normal (5 mM) [K‘] solution containing 0.06 mM 45Ca2+ and MeHg, when appropriate. 56 Actual volume of synaptosomes in the suspension is presumed to be negligable, as protein concentrations are low. Therefore, the final K+ concentration of the depolarizing solution was 41.25 mM. After 1 s (fast phase) or 10 s (total influx) of incubation, 2 ml of ice-cold quench solution was added to stop 45Ca2+ influx. A metronome was used to facilitate measurements at 1 s. Predepolarized (slow phase) influx was measured by adding 25 pl of synaptosomal suspension (approximately 40 mg protein/ml) to 25 ul of Ca2+- and MeHg-free elevated [K*] solution for 10 s, followed by the addition of 50 ul of elevated [K+] solution containing 0.06 mM of “Ca2+ and MeHg (when appropriate). After an additional 10 s, the reaction was quenched as above. A stopwatch was used to measure 10 5 intervals. The final extracellular concentration of Ca2+ plus 45Ca2“ ([Ca2*]c) was always 50 MM, unless Otherwise noted. In experiments in which [Ca2+], was increased, the molar ratio of 4“(’Caz‘VCaz“ was maintained at a constant value. Immediately following addition of quenching solution, samples were filtered through Millipore filters (0.45 pm) and washed with two 5 ml aliquots of ice-cold quenching solution. Filters were placed into scintillation vials containing 1.5 ml of Triton X-100/ HCl solubilizer and 10 ml of scintillation cocktail was added after 10 min. Radiotracer content was determined in a Searle Mark III scintillation counter with a 70% efficiency for “Ca“, or in a Searle model 1197 gamma counter with efficiencies of 47 and 53% for 85Sr2+ and 133Ba“, respectively. Depolarization- induced “Ca“ influx was determined by subtracting the average of three replicate values obtained in 5 mM K“ solution from the average of three replicate values 57 obtained in 41.25 mM K” solutions and converting the net value to nmol ”Ca2+ /ug of synaptosomal protein (Lowry et al., 1951). Statistical analysis. All dose-response data were analyzed using a randomized block analysis of variance (ANOVA) and Dunnett’s one-tailed t-test (Steele and Torrie, 1980). Data from experiments in which [Ca2+], was increased were analyzed using a split-plot ANOVA and, if significant differences were found, comparisons were made using the least significant difference test. Comparisons between data obtained in NaI-containing and Nai-free solutions were made using a paired Student’s t-test. For all comparisons, statistical significance was defined by ps0.05. 58 RESULTS Effects of MeHg on depolarization-independent entry of ”Ca2+ into synaptosomes. Because of the ability of MeHg to alter a variety of membrane- associated functions, its ability to disrupt those mechanisms which help to regulate cytosolic Ca2+ in synaptosomes in the absence of depolarization was investigated. Synaptosomes were incubated in normal (5 mM) K+ solutions in the presence of MeHg and 45Ca“ for 1 or 10 s, and then assessed for 45Ca“ content. Compared to untreated controls, MeHg had no significant effect on depolarization-independent 4sCa2+ entry into synaptosomes at concentrations up to 50 I: M, but significantly reduced 45Ca2+ entry at concentrations of 125 and 250 p. M (Figure 3.1). Thus, during exposures of 1 and 10 s, 45Ca2+ entry into synaptosomes is suppressed, but only by high concentrations of MeHg. As noted in the Methods section, this component of influx is subtracted from depolarization-induced influx. Thus, this effect did not contribute to block of depolarization-induced influx by MeHg. Effects of MeHg on depolarization-induced 6Ca2+ influx into synaptosomes. Depolarization of synaptosomes with elevated [K*] solutions resulted in two components of influx (Figure 3.2). A rapid component which inactivated after approximately 2 s and a slower component which saturated after approximately 20 3 Furthermore, influx was linearly related to the concentration of protein in the sYIlaptosomal preparation (Figure 3.3). The effects of increasing concentrations of MeHg on total 45Ca2“ influx into sYnaptosomes are shown in Figure 3.4A. In contrast to the effects of MeHg described :9” SiA g x D .e i a x e 3- 1* CL 0’ g .. O E 5 I 1 - e B m sol § E *- 8 20- 9 . at- 'O 5 ‘5'0 ‘560 [MeHg] (1.1M) Figure 3.1. MeHg blocks resting 4L5Ca2+ entry into synaptosomes. Efiects of MeHg on depolarization— independent 45Ca2+ influx into synaptosomes during 1 s (A) or 10 s (B). Synaptosomes were incubated in 5 mM K" solutions for 1 or 10 s in the presence of 5Ca2+ and various concentrations of MeHg. Values are the mean 1- SEM of three or five separate experiments, respectively. Data points not connected indicate MeHg-free control values and the asterisks (’) indicate values which are significantly less than control (150.05). INFLUX 20 “’Ca 312%: it In, lift] 1 “Win 6O 'o 70 . >5 1; ./.\"O .1 u. -' ,. z 2 +- o. 35 ‘ ./ °‘ or .8 g. 0’./ g 0,; - t: V O 5 10 30 45 60 TIME (sec) Figure 3.2. Influx of “Ca“ into synaptosomes vs duration of depolarization. 45Ca2“ influx into synaptosomes induced by depolarization in elevated [K+] solutions as a function of time. Two phases are illustrated: a fast component which lasts approximately 2 s, and a slower component which lasts up to 60 s. The values shown are the mean i SEM of three experiments, where the values for any one experiment are the average of three replicates. 61 1500 1000 500 ' / . 0 100 200 300 0 “Ca“lNFLUX (0pm x 10’) pg PROTEIN/5 O uL Figure 3.3. Influx of ”Ca2+ into synaptosomes as a function of protein concentration. Values from a single representative experiment comparing 45Ca2+ influx with protein content of the synaptosomal suspension. The line is the linear least-squares regression of influx on protein content. Values for influx are the average of three replicates. 62 for depolarization-independent influx of Ca2+, depolarization-induced influx was more prominently suppressed by MeHg. At concentrations above 5 [J M, MeHg caused a concentration-dependent suppression of 45Ca2+ influx which reached approximately 100% inhibition with 250 11M MeHg. Concentrations of MeHg 2 25 11M significantly depressed 45Ca“ influx when compared to untreated controls. An estimated 1C5o value for MeHg effects on total influx is 75 BM. These results are similar to those of Atchison et a1. (1986). When effects of MeHg on the two distinct phases of depolarization-induced 45Ca2“ influx were examined, tracer influx was again reduced in a concentration- dependent manner by MeHg. Increasing concentrations of MeHg caused a dose- dependent suppression of 45Ca2+ influx during the fast phase (Figure 3.4B) which reached a maximal effect of approximately 90% with 250 [1. M MeHg. Nearly identical effects of MeHg were observed on the slow phase of ”Ca2+ influx (Figure 3.4C). The estimated ICso value for MeHg is also approximately 75 u M for both the fast and slow phases of 45Ca”’ influx. Antagonism or block of ”Ca2+ influx in response to MeHg by increasing [Ca2+],. The effect of increasing [Ca2+], from 0.01 to 1.15 mM on block of total 45Ca2“ influx by MeHg is shown in Figure 3.5A. This concentration of MeHg approximated the estimated 1C,0 value. Block of 45Ca2+ influx by 50 u M MeHg was decreased from approximately. 98% to 75% by increasing [Can]e from 0.01 to 1.15 mM. However, at each [Ca2+], tested, there was significantly less influx of ”Ca“ in the presence of 50 11M MeHg than for the corresponding untreated control. 63 50' 45C0 UPTAKE (7: Control) 50' 5 50 500 MW Figure 3. 4. Meng blocks depolarization-induced 6Ca2+ influx into synaptosomes. Efiects of MeHg on influx of4 ”Ca into rat forebrain synaptosomes during 10 (TOTAL, A) and 1 (FAST, B) s of depolarization with 41.25 mM KCl, or during 10 s of depolarization which was preceded by a 10 s predepolarization in the absence of Ca2+ and MeHg to inactivate the fast phase (SLOW, C). Synaptosomes In physiological saline solution (5 mM K+) were added to a reaction mixture which contained 77.5 mM KCl, 40 ll M CaCl,, 0 .06 mM 45CaCl,, and MeHg. Final concentrations of MeHg ranged from 2.5 to 2.50 nM. Influx In 5 mM K+ solutions was subtracted from that m 41.25 mM K+ as depolarization-independent background. All values are the mean 1: SEM of four to six difierent experiments. The values for any individual experiment are the average of three replicates. When SE bars are not shown, the SE is smaller than the size of the symbol. The asterisks (‘) indicate results that are significantly less than control values (1:50.05). r“ ' .0. K 5 L5 0 '2' a. 01.0 E" '2 50.5 45m: UPTAKE mm Figure 3.5. Block of Ca2+ channels by MeHg is not reversed by increasing [Ca2+]e. Dependence on [Ca2+]e of 45Ca2+ influx into synaptosomes in the absence (solid circles) and presence (open circles) of 50 p. M MeHg during A) 10 s, or B) 1 s of K+-induced depolarization, or C) 10 s of depolarization following a 10 s predepolarization in the absence of [Ca2+]e. 5Ca2+ influx during TOTAL, FAST and SLOW phases was measured as described in Figure 3.4. [Ca2+]e ranged from 0.01 to 1.2 mM, and the mole to mole ratio of 45Ca2+ to Ca2+ was held constant at 0.22. Values are the mean: SEM of five to seven different experiments. The values for any individual experiment are the average of three replicates. When SE bars are not shown, the SE is smaller than the size of the symbol. The asterisks (‘) indicate results that are significantly less than control values (50.05). 65 Figure 3.5B depicts the effect of increasing [Ca2”]g on block of the fast phase of 45Ca“ influx by 50 11M MeHg. Once again, in the presence of MeHg, increasing [Ca2+]: did not overcome the block of depolarization-dependent 4S’Ca“ influx.MeHg reduced 45Ca2+ influx by approximately 50% at all [Ca2+]c tested, except 0.01 and 0.62 mM. With the exception of the value for 0.62 mM [Ca2+]c, 45Ca2+ influx at each [Caz‘i]e tested was significantly lower in the presence of MeHg, when compared to the respective control values. In contrast to the results observed for both total and fast influx, block of the slow phase of 45Ca2+ influx was reversed completely by increasing [Ca2+],3 (Figure 3.5C). At [Ca2+], below 0.3 mM, 50 11M MeHg caused significantly less influx of 45Ca2+ when compared to control values, reducing 45Ca2+ influx by 50 to 90%. However, above 0.3 mM [Ca2+]c, ”Ca2+ influx returned to control levels despite the presence of MeHg. It should be noted that the sum of the fast and slow components does not equal the amount of total 45Ca2* influx. Two factors account for this. The first is that the fast phase is largely inactivated by the 10 s predepolarization step when Slow influx is measured. However, the fast phase does contribute an unknown amount to the 10 s of 45Ca“ influx measured for the total component. The second factor is that total influx is measured over the first 10 s of depolarization, whereas the slow component of influx is measured over seconds 10-20 of depolarization following a 10 s predepolarization. Small differences in the slow component of influx may therefore exist in the two intervals during which total and slow influx are measured. 66 Dependence of MeHg-induced block or “Ca2+ influx on [Nat], The dependence of block of 45C3“ influx on [Na‘i]c in each phase of influx was examined by using synaptosomes incubated in solutions in which NaCl had been replaced by choline chloride on an equimolar basis. MeHg (50 11M) caused nearly complete suppression of 45Ca2+ influx into synaptosomes incubated in N ai-containing and N ai- free solutions (data not Shown). Figure 3.6A depicts effects of 25 11M MeHg on the fast phase of 45Ca“ influx into synaptosomes incubated in Na”-containing and Na+- free solutions. Synaptosomal 4'"Caz" influx was significantly reduced to approximately 50% of the MeHg-free control value in both Na*-containing and Nai-free solutions. Figure 3.63 depicts effects of 25 pM MeHg on the slow phase of 45Ca2+ influx in Na+-containing and Na*-free solutions. In the absence of MeHg (open bars), influx of ”Ca2+ into synaptosomes incubated in Na+-free solutions is reduced greatly compared to that in Na*-containing solution. Since Slow phase 45Ca2+ influx into synaptosomes may be due largely to reversal of electrogenic Na+ /Ca2+ exchange (Coutinho et al., 1984; Turner and Goldin, 1985; Suszkiw et al., 1986), removal of Na” from the incubation medium would deplete the synaptosomes of Na“, eliminating or greatly reducing Ca2+ uptake via Na+ /Ca2+ exchange. In the presence of 25 1: M MeHg, ”Ca2+ influx into Na” containing and Na*-free synaptosomes was reduced to values significantly less than their respective MeHg-free controls. The comparative blocking effect of 25 11M MeHg in both N a+-containing and N a”-free solutions in shown for both fast and slow phases of 45Ca2+ influx in Figure 3.7. In the slow phase, 25 BM MeHg inhibited 26.1245 and 96.0:4.0% of control 67 l5 UCONTROL A 72 MeHg it? _C_> X '0 .. .E 2 ill * O E. or 5 - 5:" O E 5 O 45 to B x E . :1 3O ' O o 5,? I5 - i O NORMAL No-Free Figure 3.6. Fast and slow components of influx have differential dependence on Nae. Dependence of block of 4sea“ influx into syna tosomes by MeHg on [Na+]e during A) 1 s (FAST) of K+-induced depolarization or B) 10 s of K -induced depolarization following a 10 s predepolarization in the absence of Ca2+ and MeHg (SLOW). Synaptosomes were incubated in normal Na+ (145 mM) or NaI- free (NaCl replaced by choline chloride) solutions in the absence (open bars) or presence (striped bars) of 25 [1M MeHg. Values shown are the mean 1 SEM of five individual experiments. Values for any individual experiment are the average of three replicates. The asterisks (’) indicate results that are significantly less than control values (50.05). 68 IOO r 50 - \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ x 1., We % REDUCTION FROM CONTROL SLOW FAST Figure 3.7. In the absence of Na, block of slow influx by MeHg is much greater in magnitude. Reduction of “Ca2+ influx from control values by 25 u M MeHg during the slow and fast phases of influx in normal (open bar) and NaT-free (striped bars) solutions. All other conditions are the same as those in Figure 3.4. The asterisk (‘) indicates a significantly geater percent reduction by 25 u. M MeHg in Na+-free solution than in normal Na+-containing solution (50.05). 69 tracer influx in synaptosomes incubated in Na*-containing and Na*-free solutions, respectively. Thus, block of the slow phase of 45Ca2+ influx was significantly more pronounced when Na“c was deleted from the solutions. For the fast phase, 25 BM MeHg inhibited control tracer influx by 44.6:5.1% in Na*-containing and by 51.81-63% in Na+-free solutions, respectively. This result is consistent with the Na+- independence of the fast phase of 45Ca2+ influx. Complete antagonism of block by MeHg of slow 45Ca2+ influx by increasing [Ca2+], may be due to competitive inhibition of the various components of influx associated with this phase, or to an irreversible block of one component by MeHg which is counteracted by increasing [CaZi]e due to a lack of effect of MeHg on some other component(s) of influx. For example, MeHg may not affect the exchange- mediated component of the slow phase of influx, but may reduce or block the other component(s) of this phase of influx. This type of action by MeHg would enable an operating exchange component to induce a functional antagonism of MeHg-induced block of the other component(s) of influx. Thus, the true efficacy of the blocking action of MeHg on the other component(s) of slow influx can only be observed when the exchange-mediated component is made inoperative, such as by the removal of Na+ from the medium. As [Ca2+]c is increased, the amount of Ca2+ uptake by the reversed Na+ /Ca2+ exchanger is greater due to the increased driving force. If [Ca2+], is increased to a high enough level, the resulting uptake may completely mask the block by MeHg of other component(s) of influx. If MeHg blocks slow influx by this mechanism, block of 45Ca2+ influx should not be reversed when Na+ is removed from 70 the medium. This hypothesis was tested and the results are illustrated in Figure 3.8. At [Ca2+]e = 0.15 mM, “Ca2+ influx was 24:07 x 10'5 and 06:06 x 10'5 nmol 4S‘Ca“ /ug protein in the absence and presence of 25 it M MeHg, respectively. When [Caz‘i]e was increased to 1.15 mM, the values for 45’Ca2+ influx in the absence and presence of 25 11M MeHg were 16.1:4.9 x 10—5 and 05:05 x 105 nmol 4SCa“ /p. g protein, respectively. Thus, MeHg-induced block of 45Ca2“ influx was not antagonized by increasing [Cay]e in the absence of Na+c. 71 [Céle (mM) * ZOi 8&015 . EJIJ5 4506 UPTAKE § (nmol/pg protein xIO-Si) CON MeHg Figure 3.8. In the absence of Nae, block of slow45 influx by MeHg is not reversed by' Increasing [Ca2+]. Effects of Increasing [Ca +e] on eblock of slow 452+Ca influx by MeHg (5 I1» in Na -free solutions. 452+Ca influx was measured In solutions which contained 0.15 or 1.15 mM Ca + .All other conditions are as in Figure 3. 4. Values shown are the mean: SEM of five individual experiments. Values for any individual experiment are the average of+ three replicates. The asterisk (‘) indicates results that are significantly increased by Increasing [Ca2 +c] from 0.15 to 1.15 mM (50. 05). 72 DISCUSSION This study confirms previous results showing a dose-dependent suppression of depolarization-induced ”Ca2+ influx into synaptosomes by MeHg, and partial antagonism of MeHg-induced block of total influx by increasing [Ca2+]c (Atchison et al., 1986). In addition, the present experiments characterize further the mechanisms of blocking action of MeHg on Ca2+ influx into nerve terminals. More specifically, these results suggest that: 1) high concentrations of MeHg depress depolarization- independent 45Ca2" entry into synaptosomes; 2) block of the fast and slow phases of 45Ca’" influx by MeHg differ with respect to their sensitivity to antagonism by increasing [Ca2+]e; 3) the magnitude of block of the Slow phase by MeHg is increased in the absence of Na+,; and 4) block of the slow phase is not antagonized by increasing [Caz‘i]e when synaptosomes are incubated in Na+-free solutions. Incubation of synaptosomes in non-depolarizing solution containing MeHg resulted in decreased 45Ca2“ entry after 1 and 10 s when compared to MeHg-free controls. These results indicate that high concentrations of MeHg reduce the permeability of synaptosomes for Ca“. MeHg increases K+ permeability in yeast (Passow and Rothstein, 1960) and increases leakage conductance in squid axon (Shrivastav er al., 1976). Interestingly, MeHg has been shown to increase Ca2+ permeability in synaptosomes (Komulainen and Bondy, 1987), but only after long periods of incubation. The decrease in Ca2+ permeability produced by MeHg in the present experiments may be due specifically to block of Ca2+ entry via random openings of channels, a decrease in the frequency of random openings resulting from 73 a K*-induced hyperpolarization, or an effect on membrane Ca2+ transport systems. Given the ability of MeHg to block depolarization-dependent Ca2+ influx via Ca2+ channels, it is possible that MeHg may also block Ca2+ entry via random openings of channels in nonodepolarized membranes. Alternatively, MeHg alters Na“ and K+ gating mechanisms in neuroblastoma cells (Quandt et aL, 1982) and may therefore alter Ca2+ channel function indirectly. If MeHg were to increase K+ permeability in synaptosomes as it does in yeast (Passow and Rothstein, 1960), then hyperpolarization of the membrane would result in a decreased frequency of random openings of Ca2+ channels, thus decreasing the permeability of synaptosomes for Ca“. The lack of alteration by MeHg of decay of miniature endplate potential (MEPP) frequency (Traxinger and Atchison, 1987a) and the ability of increasing [Ca2+]: to reverse block of the Slow phase of depolarization-induced influx indicate that membrane Ca2+ transport systems, such as Na+ /Ca2+ exchange may be less susceptible to disruption by MeHg. Finally, MeHg causes release of Ca2+ from intraterminal stores, such as the mitochondria (Levesque and Atchison, 1987; 1988; 1990). Increases in intrasynaptosomal [Ca2+] may decrease random opening of Ca2+ channels resulting from Ca2"-dependent channel inactivation (Eckert and Chad, 1984). Thus, MeHg may reduce Ca2+ permeability by decreasing Ca2+ entry via random openings of channels, or indirectly by decreasing the frequency of random openings of Ca2+ channels. In contrast to its effects on depolarization-independent “Ca2+ entry into synaptosomes, these results indicate that MeHg is a potent and effective inhibitor of 74 depolarization-induced 4SCa“ influx. Atchison et al., (1986) have shown previously that MeHg reduces influx of 45Ca2+ into synaptosomes and that increasing [Ca2+]c only partially antagonizes block of total influx of 45Ca2+ by MeHg. Inasmuch as total influx consists of at least two distinct components, the effect of increasing [Ca2+]c on MeHg-induced block of each phase was examined to determine the mechanism of interaction of MeHg with Ca2+ channels, as well as exchange processes. Block of the fast phase of influx by MeHg was not significantly antagonized when [Ca2+]e was increased. Thus, MeHg appears to block the voltage-dependent Ca2+ channels which mediate fast influx in a non-competitive, irreversible manner. Fast influx has been associated with release of Substance P (Floor, 1983), acetylcholine (Suszkiw and O’Leary, 1983) and dopamine (Drapeau and Blaustein, 1983, Leslie et al., 1985) from synaptosomes. Thus, the fast phase couples Ca2+ entry and subsequent neurotransmitter release from nerve terminals. The ability of increasing [Ca2+]e to antagonize block of fast influx by MeHg is consistent with the ability of increased [Ca2"]e to cause partial restoration of EPPS at the neuromuscular junction following block by MeHg (Traxinger and Atchison, 1987b). In isolated nerve terminals, then, MeHg appears to block Ca2+ channels which are associated with depolarization-dependent Ca2+ entry and subsequent neurotransmitter release. Complete restoration of EPPS following block by MeHg is also dependent on increased duration and / or intensity of stimulus, indicating that MeHg may exert additional effects on membrane excitability at the nonmyelinated portion of the neuron or along the axonal membrane (Traxinger and Atchison, 1987b). It is possible 75 that MeHg acts at both sites simultaneously or that block of Ca2+ channels follows diminished membrane excitability by MeHg. Increasing [Ca2+], resulted in reversal of block of the slow phase of influx by MeHg. This result is by comparison much different from those observed with total or fast influx. However, given the mechanism by which slow Ca2+ influx occurs in synaptosomes, this result is neither contradictory nor unexpected. Because the slow phase of influx may be mediated by both a channel component and a reversed Na*/Ca2+ exchange, an explanation of the actions of MeHg on this phase of influx cannot be derived simply on the basis of the results of this experiment, as MeHg may affect either non-inactivated Ca“ channels, or Na+ /Ca2+ exchange, or both. Na+ /Ca2+ exchange in synaptosomes is extremely sensitive to the internal and external concentrations of Na+ (Blaustein and Oborn, 1975; Akerman and Nicholls, 1981; Nachshen et al., 1986) and is thought to be reversed in synaptosomes due to their tendency to become Na+-loaded (Coutinho et al., 1984). When the Na+- dependence of the blocking action of MeHg was examined, MeHg block of fast 4'"Ca2“ influx was unafiected by the absence of extracellular Na‘. This is consistent with the Nai-independent, channel-mediated mechanism of fast influx. The large decrease in slow influx observed in MeHg-free synaptosomes in the absence of Na” is also consistent with previous results indicating that the slow phase is mediated largely by reversed Na“ /Ca2+ exchange, which can be abolished by the removal of Na“c (Coutinho et al., 1984; Turner and Goldin, 1985; Suszkiw et aL, 1986). 76 When Na+ /Ca2+ exchange is inactivated by removal of N a+, inhibition of the remaining component of slow influx by MeHg is increased greatly. Therefore, MeHg may exert a preferential effect on the channel-mediated component of this phase, which could be offset by an unaffected Nai /Ca2+ exchange in the presence of Na”. At the neuromuscular junction, MeHg did not alter the rates of decay of frequency of miniature endplate potentials (MEPPS) after repetitive Stimulation with Sr2+ and Ba“ (T raxinger and Atchison, 1987a). MEPP frequency is directly related to free intraterminal divalent cation concentration, and the inability of MeHg to alter the rate of decay of MEPP frequency suggests that MeHg may not alter the ability of nerve terminal buffering systems, including Na+ /Ca2+ exchange, to clear divalent cations from the cytoplasm. The inability of increasing [Ca2+] to reverse block of slow influx in the absence of N a: is consistent with an irreversible, non-competitive block of the channel portion of slow influx. In the presence of Naie this is relieved by a functional, but reversed N a+ / Ca2+ exchange. Thus, block of both fast and slow influx by MeHg occurs via an irreversible block of the channel components which mediate each phase. This is a unique action of MeHg when compared to block of Ca2+ channels by divalent cations such as Hg“, Pb“, Zn2+ and Cd“, in that block of synaptosomal Ca2+ influx by these ions is reversed by increasing [Ca2+], (Nachshen, 1984; Suszkiw et al., 1985). Whether this difference is due to the reactivity, increased lipophilicity, or monovalent charge of MeHg is not presently known. 77 In summary, low concentrations of MeHg block channel-mediated 45Ca2+ influx into KI odepolarized synaptosomes in a manner which is only partially antagonized by increasing [Ca2+]e, whereas only high concentrations of MeHg affect Na”/Ca2+ exchange. Furthermore, non-depolarized influx of 45Ca2+ is only depressed by high concentrations of MeHg. These results suggest a unique pattern of interaction of this monovalent, neurotoxic organic cation with Ca2+ channels involved in neurosecretion. CHAPTER FOUR CHARACTERIZATION OF THE INTERACTIONS OF METHYLMERCURY WITH Ctr2+ CHANNELS IN SYNAPTOSOMES AND PHEOCHROMOCYTOMA CELLS: RADIOTRACER FLUX AND BINDING STUDIES 78 79 ABSTRACT The interaction of methylmercury (MeHg) with neuronal Ca2+ channels in rat forebrain synaptosomes, and dihydropyridine (DHP)-sensitive Ca2+ channels in rat pheochromocytoma (PC12) cells was examined using radiotracer flux assays and radioligand binding analyses. In synaptosomes, the influx of 45Ca2+ was used to examine the voltage- and state-dependence of block of Ca2+ channels by MeHg as well as the effects of MeHg on apparent inactivation of 45Ca2+ influx. In addition, the differential influx of 45Ca“, 85Sr“ and 133Ba2+ was used to examine the effect of MeHg on the ionic selectivity of synaptosomal Ca2+ channels. The ability of MeHg to block 45Ca2“ influx via a DHP-sensitive Ca2+ channel was examined in PC12 cells. Effects of MeHg on binding of [3H]-nitrendipine in synaptosomes and [mu-w -conoto- xin GVIA (Cng) in synaptosomes and PC12 cells were measured. In synaptosomes, MeHg blocked 45Ca2+ influx in a voltage-dependent manner, as increasing the extracellular K” concentration increased the magnitude of block by 100 1.1M MeHg. When synaptosomes were incubated for 1 s in either a non-depolarizing or a depolarizing solution prior to measuring 1 s of depolarization-induced 4SCaz“ influx, the potency and efficacy of the block of 45Ca“ influx by MeHg were similar. Thus, block of Ca2+ channels by MeHg does not appear to be state-dependent. To determine the kinetics of apparent inactivation of 45C3“ influx, synaptosomes were predepolarized in Ca2+-free, high [K‘] solution for intervals varying from 1-10 s prior to measuring 1 s of KI-induced 45Ca“ influx. When compared to control, MeHg (100 I: M) altered the rate consth for apparent inactivation and decreased the fraction 80 of 45Ca“ influx which does not inactivate. Influx of 45Ca“, 8SSrZ“ and 133Ba2+ during 1 S of depolarization was blocked in a dose-dependent manner by MeHg with estimated 1C,0 values of 125, 150 and > 150 uM for ”Ca“, 8588* and 133m“, respectively. In triple-label experiments MeHg decreased influx of all three divalent cations: the effect on 858T2+ influx was significantly greater than on 45Ca2“ and 1”Ba2+ influx. The relative flux of radiolabeled Ca2+:Sr2+:Ba2+ was altered from ap- proximately 6:2:3 to 6:1:3 in the presence of 100 uM MeHg. In undifferentiated and nerve grth factor (NGF)-differentiated PC12 cells, K*-induced 45Ca2“ influx was blocked by the DHP nifedipine with an approximate ICso value of 5 nM. MeHg reduced 45Ca2+ influx in PC12 cells with an estimated ICso value of 50 11M, and 125 RM MeHg reduced uptake by > 90%. [3H]-nitrendipine bound to synaptosomes with high affinity in normal and elevated [K+] solutions. In normal [K+] solutions, Scatchard analysis of the binding data resulted in a KD value of 630: 160 pM and a BM value of 130:40 frnol / mg protein. In the presence of 100 u. M MeHg, the values for KD and BM were 25202630 pM and 200130 fmol/ mg protein, respectively. Results obtained in the absence and presence of MeHg in elevated [K*] buffers were not significantly different from those obtained in normal [Ki] buffers. [mI]-m - conotoxin GVIA bound to a single high affinity site on synaptosomes and PC 12 cells with a half saturation occurring in the subnanomolar range. 100 u M MeHg did not significantly alter binding of Cng in synaptosomes, but did alter Cng binding in PC 12 cells by decreasing the amount of Cng bound at all concentrations of Cng tested. These results indicate that MeHg: 1) blocks synaptosomal Ca2+ channels in 81 a voltage- but not state-dependent manner; 2) alters synaptosomal Ca2+ channel ionic selectivity and inactivation kinetics; 3) blocks Ca2+ influx via DHP-sensitive Ca2+ channels in PC12 cells; and 4) alters the binding properties of DHPS in synap- tosomes, and Cng in PC12 cells, but not synaptosomes. 82 INTRODUCTION Preliminary experiments demonstrate that MeHg does block Ca2+ channels in isolated nerve terminals, and indicate that effects of MeHg on channels are difficult to overcome (Atchison et al., 1986; Shafer and Atchison, 1989). However, these experiments provide little information regarding mechanisms by which MeHg interferes with Ca2+ channel function, or the types of Ca2+ channels affected by MeHg. However, results of preliminary experiments indicate that there may be interesting differences in the interactions of MeHg and divalent metal cations with neuronal Ca2+ channels. MeHg and divalent metals produce similar disruptions of synaptic transmission at the NMJ and impede the movement of Ca2+ through voltage—dependent Ca2+ channels. Micromolar concentrations of MeHg (Atchison et al., 1986; Shafer and Atchison, 1989) or inorganic divalent cations (Pb2+, Ni2+ and Cd“) (Nachshen, 1984) block 1 s (”fast" phase) of depolarization-induced influx of 45Ca“ into synaptosomes. Increasing [CaH]e concentration readily reverses block of "fast" influx by divalent cations (Nachshen, 1984), but not by MeHg (Atchison et al., 1986; Shafer and Atchison, 1989). MeHg differs in two important respects from divalent inorganic Ca2+ channel blockers: enhanced lipophilicity imparted by the methyl group, and monovalent rather than polyvalent charge. Because data suggest that MeHg disrupts Ca2+ entry into the nerve terminal by interacting directly with Ca2+ channels, and because chemical differences between MeHg and other heavy metal Ca2+ channel blockers 83 may give rise to differences in action on Ca2+ channels, experiments were designed to characterize in detail the nature of the interaction of MeHg with nerve terminal Ca2+ channels. Specifically, experiments were designed to determine whether MeHg blocks Ca2+ channels in a voltage- and/or state-dependent manner, and whether MeHg is capable of altering the kinetics of apparent inactivation and/or ionic selectivity of synaptosomal Ca2+ channels. The observation that multiple types of Ca2+ channels exist in nerve terminals (Penner and Dreyer, 1986; Miller, 1987; Atchison and O’Leary, 1987) raises the question of the identity of Ca2+ channels which are blocked by MeHg. As only some types of Ca2+ channels have been associated with neurotransmitter release, determination of the types of Ca2+ channels affected by MeHg is crucial to determining whether or not effects on Ca2+ channels may contribute to the effects of MeHg on transmission at the NMJ. Since the type or types of Ca2+ channels in synaptosomal preparations are not well characterized, rat pheochromocytoma (PC 12) cells were chosen as an alternative model system for studies of the effects of MeHg on Ca2+ channels. Under normal culture conditions, PC 12 cells express predominantly dihydropyridine (DHP)-sensitive Ca2+ channels (Takahashi and Ogura, 1983) known as "L" or "long opening" type (Nowycky et aL, 1985a,b). When cultured in the presence of NGF, the expression of DHP-insensitive, N-type channels is greatly increased without decreases in the expression of the DHP-sensitive channels. Thus, PC12 cells provide a means by which to examine the effects of MeHg on a 84 homogenous and/ or defined population of Ca2+ channel types which are associated with neurotransmitter release. Finally, the effects of MeHg on the binding of two well-characterized Ca2+ channel antagonists were examined. Although the pharmacological binding potency of Ca2+ channel antagonists does not necessarily correlate with their functional effects, competitive binding assays nevertheless provide a measure of the ability of MeHg to compete with known Ca2+ channel antagonists for their binding sites on nerve membranes. The DHP nitrendipine has been employed in numerous binding and flux Studies in synaptosomes (Rampe et al., 1984; Turner and Goldin, 1985; Suszkiw et al., 1986) and PC12 cells (Toll, 1982) to examine the properties of binding sites which are presumed to be located on or associated with "L"-type Ca2+ channels. Therefore, the effects of MeHg on the binding of [3H]-nitrendipine could provide additional information regarding the ability of MeHg to interact with binding sites associated with this type of Ca2+ channel. to -Conotoxin GVIA (Cng), a peptide isolated from the venom of the marine snail Conus geographus, reduces Ca2+ influx in chick synaptosomes (Rivier er al., 1987) and binds to sites on synaptosomes and PC12 cells which appear to be distinct from the DHP binding site (Abe et al., 1986; Cruz and Olivera, 1986; Marqueze et al., 1988; Sher er al., 1988). Therefore, effects of MeHg on binding of Cng were examined in synaptosomes and PC12 cells to determine whether MeHg may interact with these sites instead of, or as well as, DHP binding sites. Inasmuch as Cng has been reported to inhibit transmitter release from and 45Ca“ influx into isolated nerve 85 endings, and to bind at distinct sites from the DHPS, its use would permit a differential analysis of binding of MeHg to sites spatially associated with perhaps two different Ca2+ channels, or two separate binding Sites on Ca2+ channels. 86 MATERIALS AND METHODS The materials and solutions, procedure for isolation of synaptosomes and general method for measuring radiotracer influx are described in detail in Chapter three. The methods described in this chapter are specific experimental procedures and manipulations required for the experiments in Chapter four. Unless Specifically stated, all solutions contained the same constituents and concentrations as those described in Chapter three. The K+ dependence of 45Ca2+ influx was measured in the presence and absence of MeHg (100 nM) by incubating synaptosomes for 2 s in Na-free K“ solutions containing 45Ca2+. Choline chloride in normal [K‘] solution was replaced by KC] to achieve the various K+ concentrations used. To obtain a more accurate estimate of influx at the lower K+ concentrations used, influx was measured for 2 s at all K” concentrations. The state-dependence of block of the "fast" phase of 4S’Caz" influx by MeHg was determined by measuring 1 s of 45Ca“ influx in synaptosomes which had been exposed to MeHg under conditions in which the population of Ca2+ channels would be primarily in the "resting" state, or under conditions in which the population of Ca2+ channels would be primarily in the "open" or "inactivated" State. To "open" or "inactivate" Ca2+ channels, synaptosomes were predepolarized in 41.25 mM K” solution for 10 s by adding 25 ill of synaptosomal suspension (5 mM K”) to 25 ul of elevated [K‘] (77.5 mM) solution containing MeHg. Ca2+ channels in the resting state were exposed to MeHg in 5 mM K” solution for 10 s by adding 25 ul of 87 synaptosomes (5 mM K") to 25 pl of normal (5 mM) [K*] solution containing MeHg. Following the 10 s exposure to MeHg, 1 s of 4"Ca“ influx was measured by addition of 50 ul of K+ solution (77.5 or 113 mM K") containing 0.06 mM of 45Ca“ and MeHg, so that the final K“ concentration of all solutions was equal to 59.4 mM. One s after the addition of K+ solution, 45Ca2" influx was stopped by the addition of 2 ml of quenching solution. In order to maintain osmolarity conditions equal to those in other experiments and to increase [K‘] to 113 mM, [Na+] was lowered to 36 mM in all K+ solutions, being replaced by KCl (113 mM K+ solution) or choline chloride (physiological saline and 77.5 mM K“ solutions) on an equimolar basis. The apparent inactivation of depolarization-induced Ca2+ influx was determined by measuring 1 s of 45Ca2+ influx after various intervals of predepolariza- tion (0-10 s) in nominally CaZI-free medium to inactivate Ca2+ channels (Suszkiw et al., 1986). Briefly, 200 pl of Ca2+-free, elevated (77.5 mM) [K”] solution with or without MeHg were combined with 200 ul of synaptosomal suspension (5 mM K”). Following 1 to 10 S of predepolarization, 400 pl elevated [K+] solution containing 0.02 mM “Ca2+ and MeHg were added to the mixture. After an additional 1 s, 3 ml of ice-cold quenching solution were added to stop 4f’Ca“ influx. The final concentration of MeHg before the addition of quenching solution was 100 u M. Non-inactivated influx (1 s) was measured by combining 200 ul of synaptosomal suspension with 600 ill of elevated [K‘] solution containing MeHg (100 nM) and ”Ca2+. Synaptosomes were retained on Millipore filters and 4"‘Ca“ influx was determined as described in Chapter three. 88 PC12 Cell Experiments. PC12 cells were cultured at 37°C in a humidified 5% CO2 atmosphere using Dulbecco’s modified Eagle medium (pH 7.4, GIBCO) containing 10% horse serum, 5% fetal bovine serum, and 2 mM HEPES. PC12 cells were differentiated by culturing for 7 days in medium containing 100 ng/ml nerve growth factor (NGF) isolated from mouse submaxillary gland. The NGF was a gift from Dr. Steven Heidemann at Michigan State University. Cells were harvested by suspension in normal (5 mM) [K+] PC12 medium and centrifugation for 10 min at 5000 x g. This Step was repeated twice to wash completely the grth medium from the cells. The effects of nifedipine (Sigma, St. Louis, MO) and MeHg on depolarization- induced 45Ca2“ influx were measured using the same method as for synaptosomes, except that elevated [K‘] PC12 medium contained 100 mM K+ so that the final concentration of K“ used for depolarization of the PC12 cells was 52.5 mM. Elevated (100 mM) and normal (5 mM) [K‘] PC12 medium used in influx studies with PC12 cells have been described by Stallcup (1979) and are similar to those described for synaptosomes in Chapter three. Influx was measured over 2 min. PC 12 cells were not preincubated with nifedipine or MeHg, therefore the only exposure of the PC12 cells to MeHg or nifedipine was the 2 min during which 4SCa“ influx was measured. The final concentration of Ca2+ was 1 mM. During filtrations, the cells were retained on Whatrnan GF/B glass fiber filters. Binding Experiments. The equilibrium binding of [TU-nitrendipine (specific activity, 70 Ci / mmol) was measured in synaptosomes in the following manner. For 89 any given experiment, 2 ml of a stock solution containing [3H]-nitrendipine and MeHg (100 nM), when appropriate, was prepared in normal [K‘] solution. Binding of [am-nitrendipine was initiated by addition of 20 [J] of synaptosomal suspension (250-350 mg protein) to 500 pl of stock solution. After 1 hr of incubation at 25°C, the solution was filtered rapidly through Whatman GF/B glass fiber filters and rinsed with three 5-ml aliquots of quenching solution. To determine total ligand concentra- tion, a SO-ul aliquot was taken from the remaining 500 pl of stock solution. Non- specific binding was measured in the presence of 1 11M nifedipine. Radioactivity remaining on filters was determined as described for 45Ca2+ influx experiments, except that the scintillation counter contained a 3H quench curve and program to convert cpm to dpm. The average of triplicate values were used for Scatchard analysis (1949) to determine the equilibrium dissociation constant (KD) value for binding of Pin-nitrendipine and the density of binding sites (Bum). As with all experiments using DHPS, these procedures were carried out in the dark. To measure the binding of 3Iii-nitrendipine under depolarizing conditions, the exact same procedure was followed using elevated [K‘] solution in place of normal [K‘] solution. Binding of [mu-to -conotoxin GVIA (specific activity 2200 Ci/mmol) to PC12 cells and synaptosomes was measured under non-depolarizing conditions by methods similar to those described for nitrendipine, except that PC12 cell medium was used for binding experiments in PC12 cells. In addition, the effects of pretreatment of synaptosomes with MeHg on Cng binding were examined in the following manner. Once isolated, synaptosomes were resuspended in 5 ml of physiological saline 90 containing 100 RM MeHg and incubated at room temperature for 10 min. Following the incubation, synaptosomes were repelleted and washed in MeHg-free physiological saline by centrifugation at 10,000 x g for 10 min and subsequent resuspension. Synaptosomes were then incubated for 25 min at room temperature (25°C) and 10 Ill aliquots were used to measure the binding of [1”I]-3; __________________ 8‘ 0 IO 20 30 TIME OF DEPOLARIZATION (seC) Figure 4.1. Block of 45Ca2+ influx into synaptosomes vs. time. Block of 45Ca2+ influx into synaptosomes by MeHg as a function of time of exposure in Na-containing (solid symbols) or Na-free medium (Open symbols). 45Ca2+ inflrrx was measured by the addition of 50 211.1 of synaptosomes to 50 pl of elevated (77.5 mM) or normal (5 mM) [K ] solutions containing 45 C and 50 i4 M MeHg. Influx was stopped after various intervals of time by the addition of 2 ml of quenching solution. Influx In normal [K ] solutions was subtracted 2to determine depolarization-induced 452Ca influx. The results are expressed as a percentage of 452+Ca infirnr In MeHg-free solutions, and are the mean :1: SEM of three or four separate experiments, respectively. Values for any particular experiment are the average of three replicates. The asterisk (‘) indicates values that are significantly different than (p < .05, one-sided t-test) the value at 1 s in Na-containing (significantly greater) or N a-free (significantly less) synaptosomes. After 30 s of depolarization, the values for influx of 45Ca2+ into synaptosomes in the absence of MeHg were 1.99 x 058 and 7.85 1 2.22 nmol/pg protein x 10'4 in Na-free and Na-containing solutions, respectively. 94 (open symbols) of 100 11M MeHg. Influx in 5 mM K+ has not been subtracted from these data. The magnitude of depolarization caused by each K+ concentration can be estimated by the relationship AVm = 60 log [K+]H/[K+]L, where an is the change in membrane potential and [K”] is the K+ concentration of the normal (L) and elevated (H) [K*] solutions (Blaustein and Goldring, 1975). Thus, the voltage-dependence of influx can be estimated by this method. The K+ concentrations used in this experiment will result in depolarizations of 0 to 55 mV from the resting membrane potential, which under Similar conditions has been estimated to be approximately -80 mV (Suszkiw et al., 1986). In the absence of MeHg, synaptosomal 45Ca2+ influx activated at [K+] above 10 mM, and increased in a linear manner as [K‘] was increased to 41.3 mM. In a separate experiment, higher concentrations did not increase “Ca2+ influx further (data not shown). Thus, 45Ca2+ influx in synaptosomes activates at K“ concentrations above 10 mM and is maximal at K+ concentrations of approximately 41 mM. These results are consistent with those of Suszkiw et al. (1989). In the presence of 100 11M MeHg, ”Ca2+ influx increased only gradually as the [K‘] was increased, being significantly less than control values at each respective K+ concentration. In panel B of Figure 4.2, the percent inhibition of 45Ca“ influx by MeHg is plotted as a function of [Ki]. At low K+ concentrations, MeHg blocked approximately 20% of the 45Ca2+ influx. As the [K*] was increased, block of uptake by 100 u M MeHg increased to approximately 50% at the highest K+ concentration tested. The block of 45Ca“ influx in 36 mM or 95 fl - O - - O 3.50 2.25 " ' / . “. “Ca iNFLUx x10... ("moi/no protein) ; U —/3 0 T __-66 2 x 1 . é); ___________ (I, ______ it 1‘ 1.00 *" L_t , ‘ l ‘ 1 1 n I 5 10 50 60 r B a a II .1 it: _ I ' lz" :-§ 40 [/11 I E I 1 A o L ‘ l l n . 1 a a c c 5 10 50 “(‘1 (mM) Figure 4.2. Voltage-dependence of 45Ca2+ influx into synaptosomes and reflects of MeHg. A) Effects of [K ] on influx of 4'SCa“ into synaptosomes in the absence (solid symbols) and presence of 100 uM MeHg (open symbols). Syna ynaptosomes were incubated for 2 s in N a-free bufiers containing ”Ca and various concentrations of K. Infirnr was stopped by the addition of 2 ml of quenching solution. The asterisk (‘) indicates values of influx In the presence of MeHg which are significantly less than (p< .05) their respective MeHg-free control value. B) Magnitude of block of 45Ca2+ influx by 100 uM MeHg at difi’erent K+ concentrations used in part A. The values marked by an asterisk (‘) indicate significantly greater block (p<.0 5) by MeHg than the block in 5 mM KL All values are the mean: SEM of four experiments, where each experimental value is the average of three replicates. 96 higher concentrations of K+ was significantly greater than that observed in 5 mM K“ solutions. State-dependence Experiments. Block of Ca2+ channels in excitable membranes frequently exhibits so-called "state-dependence", wherein the blocking affinity of the antagonist is dependent upon the configuration of the channel. Inasmuch as the blocking affinity of certain divalent cations for 45Ca2+ influx into synaptosomes has been shown to be "state-dependent" (Nachshen, 1985), experiments were designed to determine whether the degree of block by MeHg depended on prior utilization of the channel. This was accomplished by incubating synaptosomes for 10 s in the presence of MeHg in either 5 mM or 41.3 mM K” solutions, prior to measuring 1 S of depolarization-induced ”Ca2+ influx. These K+ concentrations are below and above the threshold for activation of 4"Ca“ influx into synaptosomes, respectively. Thus, in synaptosomes incubated in 41.3 mM K+, MeHg had access to the open and inactivated states of the Ca2+ channel, while in synaptosomes incubated in 5 mM K+ solution, MeHg had access to Ca2+ channels in the resting state. After incubation in 5 mM K+ solution, MeHg caused a dose-dependent suppression of 1 s of 45Ca2+ influx which reached 84.5 15.0% (mean a SEM, n =6) reduction with 250 it M MeHg; the estimated 1C5o value was 75 BM (Figure 4.3, filled circles). When MeHg had access to open or inactivated Ca2+ channels prior to measuring 1 s of influx, a dose-dependent suppression of 45Ca’“ influx was also observed (Figure 4.3, Open circles). The estimated 1C,o value for MeHg was also 75 I: M and inhibition was maximum (89.414.5%, mean a SEM, n=5) at 250 uM MeHg. Although ”Ca2+ 97 120 r —‘—' Resting --O-- Open I x ... _, D .3 90 \I E! :3 °‘~~ L.\’. .2.. ‘0: 60 i ”(1) ------------ 13’ 0 (U «.0 ES 30 - V o - - - _ 2 10 [MeHg] (pM) Figure 4.3. State-dependence of the actions of Mei-lg on “Ca2+ influx. Influx measured during 1 s of K+ -induced depolarization following predepolarization for 10 s in the presence of MeHg (open symbols) or following 10s of incubation with MeHg in a non-depolarizing solution (5 mM K”) (solid symbols). Values are the mean: SEM of six different emeriments. For any given experiment, the value was the average of three replicates. 98 influx at low concentrations of MeHg was reduced to a greater extent in predepolar- ized synaptosomes than in non-depolarized synaptosomes, there was no significant difference in the reduction of 45Ca“ influx under the two conditions. Thus, block of 45Ca“ influx by MeHg was not altered when MeHg was allowed access to different states of the Ca2+ channel. Inactivation Experiments. Predepolarization of synaptosomes in the absence of added Ca2+ resulted in a time-dependent decrease in 45Ca2“ influx into synaptosomes when 45Ca2+ -containing solutions were subsequently added (Figure 4.4). This decrease is presumably due to voltage-dependent inactivation of synaptosomal Ca2+ channels. The ability of MeHg to alter this process in a quantifiable manner was examined using the following model for apparent inactivation in synaptosomes: Q = qle'kit + qz; where Q is the fraction of Ca2+ influx remaining after time t of predepolarization, k is the rate constant for apparent inactivation, and q1 and qz are the fractions of Ca2+ influx which do and do not inactivate, respectively. This relationship has been shown previously (Suszkiw et aL, 1986) to describe the apparent inactivation of Ca“ influx in synaptosomes under nearly identical experimental conditions. To validate the use of this model, the goodness of fit of these data to the above model was examined by non-linear least squares regression analysis. For control experiments, correlation coefficient values ranged from 0.77 to 0.94. The goodness of fit of these data to several other potential models for inactivation was also examined, including a model for the exponential decay of a voltage-activated current (Belluzzi er al., 1985) which has been used recently to describe the apparent 99 inactivation of Ca2+ influx in synaptosomes incubated in Na+-free medium (Suszkiw et al., 1989). Correlation coefficient values for goodness of fit of these data to other models were less satisfactory than those for the model chosen. Using the CRUNCH statistical program, non-linear, least-squares regression analysis was performed on the data fiom each individual experiment to determine ki, ql and q2. The average values for ki, ql and q2 for control and MeHg-treated synaptosomes are listed in Table 1 and yield the lines in Figure 4.4 when used in the above function. Significant differences in ki, ql and q2 in MeHg-treated (n=3) and control (n = 5) synaptosomes were tested for using a one-sided Student’s t-test (pg .05). When MeHg-treated synaptosomes were compared to untreated control synaptosomes, the values for ki, ql and q2 were significantly different. Thus, MeHg alters the kinetics of apparent inactivation of 45Ca’“ influx in synaptosomes by altering the rate of apparent inactivation and increasing the fraction of influx which inactivates. It should be noted that in the presence of MeHg, inactivation of 45Ca2+ influx was essentially complete after only 1 s of predepolarization, thus the value for the rate constant ki only reflects the apparent constant for the rate of inactivation. Ionic Selectivity Experiments. For dose-response curves, solutions contained a final concentration of 2.4 mM unlabelled Ca2+, Sr2+ or Ba“, and 1 uCi of ”Ca2+, 85Sr2+ or 133Ba“, respectively. MeHg reduced depolarization-induced influx of 45Ca2“ and 8SSrz” into synaptosomes in a dose-dependent manner, with estimated ICso values of 125 and 150 uM, respectively (Figure 4.5). Although the ICso value for block of “Ca2+ uptake by MeHg is slightly higher than previously reported (Shafer and 100 X D .1 L E _.l < . Z \ Q . . l- 2 g .2 i m T T LI. 0 O O . l . L . O 4 8 1 2 PREDEPOLARIZATION (seC) Figure 4.4. Effects of MeHg on apparent inactivation of synaptosomal “Ca“ influx. Effects of MeHg (100 u. M) on the rate of apparent inactivation of depolarization-induced 45Ca2+ influx. Synaptosomes (200 #1) were predepolarized for varying intervals by addition of 200 #1 of Can-free, elevated [KW-solution containing (open symbols) or free of (solid symbols) MeHg. Following predepolarization, 400 pl of elevated [K+] solution containing MeHg (if appropriate) and 1 uCi of 45Ca2+ were added. Influx was stopped after 1 s by the addition of 3 ml of quenching solution. Fractional influx is the amount of 45Ca + influx remaining after a designated interval of predepolarization, relative to 1 s of 45Ca“ influx in the absence of predepolarization. The values shown are the mean :1: SEM of five (MeHg-free) and three (MeHg) different experiments. Values for any particular experiment are the average of three replicates. Curves are drawn using the average non-linear least squares values listed in Table 1. 101 Table 4.1 Non-Linear least squares regression values for apparent inactivation of 45Ca2+ influx into synaptosomesa ki (5’1) (11 (12 Controlb 050: 0.10 0.66: 0.02 038: 0.04 MeHgc 6.33: 0.08‘ 0.74! 0.02‘ 0.26: 0.2‘ aNon-linear least squares regression was used to fit data to the function Q = qle'kit + 92 where k! is the rate constants, ql is the fraction of influx which inactivates and 92 is the fraction of influx which does not inactivate. bData from five difierent experiments were used to calculate regression values which are given as the mean :9: SEM. cData from three different experiments using 100 [1M MeHg were used to calculate regression values which are given as the mean: SEM. ‘Significantly different from control values (1:50.05). 102 Atchison, 1989), the concentration of extracellular calcium used in this experiment was 2.4 mM, compared to 0.05 mM Ca2+ in previous experiments. In contrast to Ca“ and Sr“, the influx of 133Ba2+ was only slightly inhibited by concentrations of MeHg as high as 250 uM (Figure 4.5). The differential sensitivity of 45Ca“, 8581'“, and 133Ba“ fluxes to block by MeHg suggests that MeHg alters the ionic selectivity of the Ca2+ channel for these ions. However, the decreased potency of MeHg to block Ba2+ influx may not be due to a reduced blocking action of MeHg on Ba“ influx, but to an increased influx of Ba“ into synaptosomes due to lack of Ca2+-induced Ca2+ channel inactivation by Ba2+ (Eckert and Chad, 1984). In order to control for differences in Ca2+ channel inactivation, triple-label studies were designed in which the influx of 45Ca2+, 8SSr“ or 133Ba2+ into synaptosomes was measured in solutions which contained 0.1 mM 811* and Ba“, 0.02 mM Ca2+ and 0.5 mM Mg“. The effects of MeHg (100 nM) on the relative influx of the three divalent cations are shown in Figure 4.6. To correct for differences in concentration, the values have been normalized by dividing the influx of each ion by its concentration in the extracellular medium (Nachshen and Blaustein, 1982). In the absence of MeHg, the ratio of influx of 4sCazi’SSrsza was approximately 6:2:3. When 100 uM MeHg was present in the medium, the influx of 45Ca“ and 1338a“ were 61129.4 and 72.6: 11.6% of their respective control values, but the influx of 8SSr“ was only 44.1 t 6.9% of its influx into synaptosomes in the absence of MeHg. This was a statistically significant difference when compared to the effects of MeHg on 45Ca“ and 1333a2+ influx into 103 150 _ .._.‘_ 058'. _-O_- 10:88 _._ 4803 X g 100- D C .1 ° - ll. 0 Z "" 3‘: 50 - o - AAA. - . n A A kLA. - - 5 ‘IO 50 100 [MeHg] (IJM) Figure 4.5. Efl'ects of MeHg on the influx of 45Ca2+, ”Sr” and 1338a” into synaptosomes. Influx was measured by addition of 50 ul of synaptosomal suspension to 50 pl of elevated (77.5 mM) or normal (5 mM) [Kl'] solution containing MeHg and 1 [.5 Ci of tracer. Influx was stopped after 1 s by the addition of 2 ml of quenching solution. Depolarization-induced influx was determined by subtracting the influx of radiolabel in normal +] solution. The values shown are the mean: SEM of four (“Ca“), three (8583+) and five (1338a +) different experiments. Values for any particular experiment are the average of three replicates. 104 1.20 P - CONTROL El 100 pM MeHg. X D _l [2'- 0.80 ~ 57% UJ ..>_ E 0.40 ~ .J UJ (I *44% 0.00 ‘ 4608 858'. 13388 RADIOISOTOPE Figure 4.6. Effects of Meflgzg; the relative influx of 45Ca“, ”Sr“ and 133Ba2+ into synaptosomes. Influx of 45Ca2"’, 8SSr“ or a2+ during 1 s of K+-induced influx was measured in the absence (solid bars) or presence (open bars) of 100uM MeHg. K+ solutions contained (mM): SrClz, 0.1; BaClz, 0.1; CaClz, 0.02; M302, 0.05 and 1 u Ci of ”Ca“, 3539*, or 1333a“. The results have been normalized by dividing the influx of each cation by its external concentration. The values shown are the mean 3: SEM of three to five separate experiments. Values for any particular experiment are the average of three replicates. The asterisk (‘) indicates an effect of MeHg on 8581"?+ influx which is significantly diflerent from the effect of MeHg on 45'01“ and 1338a2+ influx (p< .05). 105 synaptosomes. Thus, MeHg altered the relative permeability of 4sCaz‘gssrsza to 6:1:3. “Ca2+ Influx into PC12 Cells. Influx of 45Ca2+ into PC12 cells by depolariza- tion in 52.5 mM K+ gradually increased as time of depolarization was increased, reaching a maximal level after about 1.5 to 2 min of depolarization (Figure 4.7). In addition, influx was a linear function of protein content of the PC12 cell suspension (Figure 4.8). The effects of nifedipine, a Ca2+ channel antagonist, on 2 min of 45C3“ influx into PC12 cells are shown in Figure 4.9. In two experiments, the approximate 1C5o value for block of 45Ca2+ influx by nifedipine in non-differentiated PC12 cells was 5 nM, and 100 nM nifedipine reduced influx of 45Ca2+ by greater than 95 % (solid symbols). This is consistent with previous studies which have shown that Ca2+ influx in undifferentiated PC12 cells is largely DHP-sensitive (T011, 1982; Takahashi and Ogura, 1983; Kongsamut and Miller, 1986), therefore further repetitions were deemed unnecessary. In NGF-differentiated cells, nifedipine reduced depolarization-dependent 45Ca2+ influx with an estimated 1C50 of approximately 5 nM (n=3) (Figure 4.9, open circles). When compared to untreated controls, nifedipine significantly reduced ”Ca“ influx at all concentrations tested in differentiated PC12 cells. Although 45Ca2+ influx was largely DHP-sensitive, 10.4 i 0.9% of control influx remained in cells treated with 100 nM nifedipine, possibly indicating a DHP-insensi- tive component. Higher concentrations of nifedipine were not tested. In undifferentiated PC12 cells, MeHg reduced 45Ca2+ influx with an ICso of approximately 50 u M, and 250 u M MeHg blocked 45C3“ influx completely (Figure 106 X 3 20[ 3 x 2E o/I\o 9- 10 ' o—O/l 1' . t. a /11 1 0 = ,o 3 O ’* E 0L - . 7, O 100 200 300 600 TIME (sec.) Figure 4.7. Influx or ”Ca“ into PC12 cells vs duration of depolarization. ”Ca2+ influx into PC12 cells induced by depolarization in elevated [K’] solutions as a function of time. The values shown are the mean 1 SEM of three experiments, where the values for any one experiment are the average of three replicates. 107 O 250 500 4"Ca“INFLUX (nmolx 1 0") pg PROTEIN/5 O |JL Figure 4.3. Influx of 4sea” into PC12 cells as a function or protein concentration. Values from a single representative experiment comparing 45Ca2“ influx with protein content of the PC12 cell suspension. The line is the linear least-squares regression of influx on protein content. Values for influx are the average of three replicates from a single representative experiment. 108 100 P ’ -0- -NGF x :2 75 " --O-- +NGF D O E “E z o 50 - "" O (U . 0 °\0 is ID v 25 ' '0 ..... g V ‘\ I \ \\‘* ox? 0 A AA]. AA Al ; A-AAA. [NIFEDIPINE] (mm Figure 4.9. Efl'ects of nifedipine on depolarization-induced influx of 4E’Ca“ in PC12 cells. Influx was measured by the addition of 50 pl of nonodifferentiated (open symbols) and NGF-difi'erentiated (solid symbols) PC12 cell suspension in normal (5 mM) [10], PC12 medium to 50 pl of elevated (100 mM) [K+] or normal [K+] PC12 medium containing 1 u. Ci of 4""Ca“ and various concentrations of nifedipine. The values shown are the mean of two (non-differentiated) and mean 1 SEM of three (NGF-differentiated) different experiments. Values for any particular experiment are the average of three replicates. For experiments using NGF-differentiated cells, SE bars are not shown if the SEM was smaller than the symbol size. The asterisk (‘) indicates results that are significantly less than control (p< .05). 109 4.10, filled circles). In NGF-differentiated PC12 cells, nearly identical results were obtained (Figure 4.10, open circles). MeHg concentrations of 50 uM and above significantly reduced 45Ca2+ influx in undifferentiated and NGF-differentiated PC12 cells when compared to untreated controls. Thus, MeHg blocks a DHP-sensitive Ca2+ channel in PC12 cells with nearly identical potency to its block of Ca2+ entry in synaptosomes. In addition, in two out of three experiments, MeHg completely blocked 45Ca2" influx into NGF-differentiated PC 12 cells. This may indicate an action of MeHg on a DHP-insensitive Ca2+ channel in PC12 cells, but would require electrophysiological experiments to confirm. MeHg did not alter the depolarization- independent entry of ”(332+ into undifferentiated or N GF-differentiated PC12 cells. In addition, concentrations of MeHg as high as 250 llM did not affect the viability of PC12 cells as measured by Trypan Blue exclusion, even after 1 hr of exposure (data not shown). Thus, the effects of MeHg on “Ca2+ influx into PC12 cells appear to be due to disruption of voltage-dependent, channel-mediated Ca2+ influx and not to overt cytotoxicity to PC12 cells. Nitrendipine Binding. The binding of [3H]-nitrendipine to synaptosomes reached equilibrium after 10-15 min and was stable for 90 min (results not shown). When the binding of [am-nitrendipine (concentrations ranged from 0.18 to 3 nM) to synaptosomes was measured in normal (non-depolarizing) [K‘] solutions, a single, high-affinity binding site was observed (Figure 4.11). Non-specific binding accounted for approximately 50% of total binding. Scatchard analysis of the data (n=3) yielded a KD value of 630: 160 pM and an apparent BM value of 130140 fmol/mg of 110 100 - P :3 75 ” -O- -NGF 3‘ o -l :5 ’ -°0--+NGF a s ._ o 50 - m I 0 e3 3 25 . r \\* . “5 ..... *- 0 “U' - --_-_-. X...” 5 10 50 100 [MeHg] (uM) Figure 4.10. Effects or MeHg on depolarization-induced ”Ca2+ influx into PC12 cells. 4Sea“ influx into non-difl'erentiated (solid symbols) and N GF-difl'erentiated (open symbols) PC12 cells was measured as described in Figure 4.9. The values shown are the mean :9: SEM for three separate experiments. Values for any particular experiment are the average of three replicates. When SE bars are not shown, the SE is smaller than the size of the symbol. The asterisk (‘) indicates results that are significantly less than control (p< .05). 111 protein. Hill slope was approximately equal to 1. These values are consistent with values reported by others (Rampe et al., 1984; Turner and Goldin, 1985; Suszkiw et al., 1986) for [3H]-nitrendipine binding in synaptosomes. In the presence of 100 uM MeHg, the binding of [3H]-nitrendipine was also linear, however, the value for K0 was increased significantly over the control KD value to 25201630 pM (n=3). The apparent Bum was 200130 fmol/mg of protein in the presence of 100 MM MeHg. A valid determination of BM could not be made due to the nitrendipine concentration range used, therefore the BM value obtained in the presence of MeHg is only an estimate and little direct information about the effects of MeHg on the number of binding sites can be gained by comparison to control Bmax values. Under depolarizing conditions, similar results were obtained. In the absence of MeHg, the KO value for the binding of [3H]-nitrendipine was 11001300 pM. Although KD was higher under depolarizing conditions, it was not significantly different from the KB value in normal [K‘] solutions. In the presence of MeHg, the KB value in elevated [K+] solutions was 533012100 pM and was significantly higher than the respective control value. Apparent BM values in the absence and presence of 100 uM MeHg were 200123 and 320174 frnol/ mg, respectively. Thus, MeHg decreased the affinity of binding of [am-nitrendipine to synaptosomes, indicating that it may act at or near the nitrendipine binding site. As depolarization did not alter the characteristics of [3H]nitrendipine binding to synaptosomes, the results of these experiments are reported here, but are not illustrated in Figure 4.11. 112 m-Conotoxin Binding. The specific binding of [mu-to -conotoxin GVIA to synaptosomes in the absence and presence of MeHg is shown in Figure 4.12A. Non- specific binding accounted for approximately 20-30% of total binding. In the absence of MeHg, Cng bound to synaptosomes with an apparent half-saturation of 0.1 to 0.2 nM. The maximum density of binding sites estimated by extrapolation of the saturated component to the Y-axis is between 300 and 450 frnol/ mg protein. These results are similar to results for Cng binding to synaptosomal membranes from rat and chick brain (Abe et al., 1986; Cruz and Olivera, 1986; Marqueze et al., 1988). In the presence of 100 uM MeHg, binding of Cng to synaptosomes was not significantly different from that of control. Because the actions of MeHg at the neuromuscular junction (Traxinger and Atchison, 1987b) and in synaptosomes are not readily reversible (Shafer and Atchison, 1989), synaptosomes were pretreated with MeHg prior to measuring Cng binding. This pretreatment did not change the characteristics of Cng binding (results not shown). In undifferentiated PC12 cells, specific binding of [mm-o -conotoxin in the absence of MeHg was saturable (Figure 4.12B). The apparent half saturation was less than 0.1 nM Cng. The apparent BM was approximately 25-50 pmol / mg protein. In the presence of 100 llM MeHg, the amount of [mike -conotoxin specifically bound was reduced significantly when binding was measured at Cg'I‘x concentrations between 10 and 500 pM. These results indicate that MeHg interacts with the receptor for to -Cng on PC12 cells, but not on synaptosomes. 113 0.07 a A CONTROL UJ 1 L 0.05 \ 0 § 0. 0.02 CD 0.00 5 27 48 7o BOUND (pM) Figure 4.11. Effects of MeHg on nitrendipine binding to synaptosomes. Specific binding of [sin-nitren- dipine to synaptosomes in the absence (solid triangles) and presence (Open circles) of 100 u. M MeHg in normal [K+] solutions. Values shown are the average of three replicates from a single representative experiment» 114 0-60 i o CONTROL A 0 MeHg I 0.40 r (T --------- i /o O 20 l ”x 1 o g/x.’ A" 1 A“ 0.00 '1’ 0.0 0.1 0.2 0.3 0.4 0.5 B i /— i e l 50' i I 25 . _/ 2 Cng BOUND (pmol/mg protein) 0 . . 0.0 0.1 0.2 0.3 0.4 0.5 [Cng] (nM) r0; *fi-l Figure 4.12. Effects of Mei-lg on binding of conotoxin to synaptosomes and PC12 cells. Specific binding of [mun-conotoxin GVIA to synaptosomes (A) and PC12 cells (B) in the absence (open symbols) and presence (solid symbols) of 100 p M MeHg. Values shown are the mean 1 SEM of three experiments, where the values for any given experiment are the average of three replicates. When SE bars are not shown, the SE is smaller than the size of the symbol. The asterisks (‘) indicate values which are significantly different from their respective MeHg-free control (p< .05). 115 DISCUSSION The goal of these experiments was to characterize the effects of MeHg on nerve terminal Ca2+ channel properties using synaptosomes and to attempt to identify the type or types of neuronal Ca2+ channels with which MeHg interacts in a defined population using PC12 cells. These results suggest that MeHg: 1) blocks 45Ca2+ influx into synaptosomes in a voltage-dependent manner; 2) blocks Ca2+ influx through synaptosomal Ca2+ channels in a manner which does not depend on the configuration (open, resting or inactivated) of the Ca2+ channel; 3) acts on synaptosomal Ca2+ channels to alter both the kinetics of apparent inactivation and ionic selectivity; 4) is capable of blocking a DHP-sensitive Ca2+ channel in PC12 cells, and may also block a DHP-insensitive Ca2+ channel in NGF-differentiated PC12 cells; and 5) alters nitrendipine and Cng binding to high-affinity sites on synaptosomes and PC 12 cells, respectively. The block of 45Ca2“ influx by MeHg increases as a function of time of exposure in Na-free medium, but not in N a+-containing meduim. Under the former conditions, Na+ /Ca2+ exchange does not contribute significantly to 45Ca“ influx; therefore influx occurs predominantly via Ca2+ channels (Suszkiw et al., 1986; Shafer and Atchison, 1989). These results indicate that MeHg may block more Ca2+ channels as the time of exposure increases or alter the inactivation rate of Ca2+ channels or both. MeHg differs from inorganic, divalent Ca2+ channel blockers such as Cd“, Pb“, Hg2+ in that it is monovalent and has increased lipophilicity imparted by the 116 methyl group. The ability of MeHg to block 45Ca2+ influx in a voltage-dependent manner implies that the monovalent charge on MeHg may be important for the interaction of this compound with Ca2+ channels. More recent results suggest however, that lipophilicity may be critical for voltage-dependent actions of mercurials, as only organic mercurials block synaptosomal 45Ca2+ influx in a voltage-dependent manner (Hewett and Atchison, 1990). Voltage-dependent block of cardiac Ca2+ channels by inorganic divalent cations is thought to arise due to binding of the cation to sites within the pore of the channel (Hess et al., 1986; Lansman et al., 1986). Thus, these results suggest that MeHg may act at a site within the channel. It should be noted that the voltage-dependence of block could be due to non- specific interactions of MeHg with the membrane rather than specific interaction with Ca2+ channels. Sulfhydryl reactive agents such as mercuric chloride and parahydroxymercuribenzoate depolarize the membrane (Huneeus-Cox et aL, 1966). If MeHg were to cause significant depolarization, it might also give rise to a "voltage- dependent" block of influx, as it may antagonize the block by MeHg at lower K+ concentrations by causing depolarization and Opening more Ca2+ channels than in the control synaptosomes. In squid axon membranes, concentrations of MeHg less than 100 p. M do not cause depolarization (Shrivastav et al., 1976) and in neuroblastoma cells 40 u M MeHg results in only 4 mV of depolarization after 6 min of exposure (Quandt et al., 1982). Thus, MeHg does not cause significant depolarization of the membrane in these preparations. However, in guinea pig synaptosomes, 100 u M MeHg caused a depolarization of the membrane of 36 mV 117 after 1 min of exposure (Kauppinen et al., 1989). Lower concentrations of MeHg may be less effective at depolarizing the synaptosomal membrane, as 10 llM MeHg produces a maximal depolarization of 15 mV (Hare and Atchison, 1990). In the experiments reported in the present study, exposure to MeHg was extremely short (2 5). Therefore, although it is possible that MeHg causes some membrane depolarization, it is unlikely that the magnitude of depolarization would be great enough to affect the results significantly. Access to the resting or open and inactivated states of synaptosomal Ca2+ channels did not affect the potency or efficacy of block of 45Ca2+ influx by MeHg. Different association and dissociation rates from the resting, open and inactivated states of ionic channels is thought to underlie state-dependent block of IN, by local anesthetics (Strichartz, 1973; Courtney, 1975 ; Hille, 1977) and the action of certain Ca2+ channel blockers (Lee and Tsien, 1983). Results of the present study suggest that MeHg does not have an increased affinity for any particular state of the Ca2+ channel, as its blocking effects were equivalent following access to the channel in the presence and absence of predepolarization. This is distinct from the blocking action of DHPS which block 1c. in a state-dependent manner in cardiac myocytes (Uehara and Hume, 1985) and the possible state-dependence of Ni“, La3+ and verapamil in synaptosomes (Nachshen, 1985 ). It should be noted that synaptosomal preparations may contain multiple types of Ca2+ channels which may have different rates of inactivation. Thus, it is possible that these experiments have examined effects of MeHg on two or more subpopulations of Ca“ channels and that MeHg has equal 118 affinity for each subpopulation. However, apparent inactivation in these experiments was best described by a single exponential process, which is consistent with previous results (Suszkiw et al., 1986; 1989). Therefore, the results of the state-dependent experiments have been interpreted as though the synaptosomal Ca2+ channel population responded to depolarization in a homogeneous manner. Predepolarization of synaptosomes in the presence of MeHg significantly in- creased the rate of apparent inactivation of 45Ca“ influx. Inactivation of synaptosomal Ca2+ channels with time results in decreased “Ca2+ influx; MeHg may decrease influx further by blocking open/ active Ca2+ channels, giving rise to an increase in the rate of apparent inactivation. Intuitively, one might expect that the rate constants for normal inactivation and MeHg-induced block of Ca2+ channels would differ, resulting in two exponential components of inactivation, unlike the case observed in the present study. Nevertheless, the results of the present experiment cannot rule out this hypothesis definitively. Alternatively, these results may indicate an ability of MeHg to hasten the process of Ca2+ channel inactivation. MeHg may act directly on the Ca2+ channel to increase the rate of apparent inactivation or may act indirectly to increase inactivation of Ca2+ channels by causing release of Ca2+ from intracellular stores (Kauppinen et al., 1989), resulting in increased [Ca2+]i and Ca2+-dependent inactivation of Ca2+ channels. This latter mechanism, although possible, is unlikely to contribute largely to the increased rate of apparent inactiva- tion of 45Ca2+ influx in this experiment. It requires that MeHg diffuse through the synaptosomal membrane and interact with the mitochondria to cause Ca2+ release. 119 Ca2+ must then diffuse back to the membrane and interact with Ca2+ channels to cause inactivation. This sequence of events would have to take place extremely rapidly, as apparent inactivation in the presence of MeHg occurs in less than 1 s. In addition, the synaptosomes have been prepared, incubated in and predepolarized in nominally Ca2+-free buffers. A redistribution of Ca2+ from the mitochondria would be expected, lowering the overall mitochondrial Ca2+ content. In addition to increasing the rate of apparent inactivation, MeHg also decreases the fiaction of 45Ca2+ influx which does not inactivate during the predepolarization period. N on-inactivating 45Ca2“ influx in synaptosomes is comprised of both a channel-mediated component and a reversed Na‘ /Ca2+ exchange component (Coutinho et al., 1984; Suszkiw et al., 1986; Shafer and Atchison, 1989). Nerve terminal N a+ / Ca2 + exchange appears to have decreased sensitivity to block by MeHg (T raxinger and Atchison, 1987b; Shafer and Atchison, 1989); therefore, if MeHg caused significant Ca2+ release from mitochondria, it could decrease this phase of uptake by decreasing the driving force for exchange, or reversing exchange. Alternatively, the effect of MeHg on the non-inactivating portion of Ca2+ influx suggests an action of MeHg on a non-inactivating population of Ca2+ channels. This decrease may be the result of either a direct block of these channels, or increased Ca2+-dependent inactivation due to MeHg-induced Ca2+ release from intracellular stores. MeHg also altered the relative influx of ”Ca2+, 85Sr2+ and 133Ba2+ into synaptosomes via Ca2+ channels. 85812+ influx through Ca2+ channels was more 120 sensitive to block by MeHg than was influx of 45Ca2“ or 133Ba“. This represents a unique action by MeHg on Ca2+ channels when compared with other divalent cations such as Ni“, which will block the influx of Ca2+, Sr2+ and Ba2+ without altering the relative influx of these three cations (Nachshen and Blaustein, 1982). Although these experiments do not define the site of action of MeHg, they suggest that MeHg may affect the function of the selectivity filter of the Ca2+ channel. It has been suggested that ionic selectivity in cardiac Ca2+ channels is due to differences in the affinity of divalent cations for binding sites within the pore of the channel, rather than a physical structure which "filters" ions which can enter the channel (Hess et al., 1986; Lansman et al., 1986). If this were the case for synaptosomal Ca2+ channels as well, then these results suggest that Sr2+ is less effective at displacing MeHg from this site than Ca2+ or Ba“. In addition, these results would be consistent with voltage- dependent effects of MeHg. The ability of nifedipine to block the influx of 45Ca“ into PC12 cells is consistent with the action of other DHP antagonists on Ca2+ influx in this cell line (Toll, 1982; Takahashi and Ogura, 1983; Kongsamut and Miller, 1986) and indicates that depolarization-induced 45Ca2“ influx into these cells is mediated largely by DHP- sensitive or "L'Ltype Ca2+ channels. In addition, in the present study, approximately 10% of 45Ca’“ influx in differentiated PC12 cells remained after exposure to 100 nM nifedipine, indicating a DHP-insensitive portion of “Ca2+ influx. Although it would require electrophysiological experiments to confirm, the remaining influx may represent influx via an "N"-type conductance in PC12 cells. MeHg blocked 4sCaz“ 121 influx into both non-differentiated and NGF-differentiated PC 12 cells at concentrations similar to those which block 45Ca2” influx into synaptosomes (Shafer and Atchison, 1989). In two out of three experiments, MeHg completely blocked 4S’Ca“ influx into NGF-differentiated PC12 cells. This indicates that MeHg may be capable of blocking a DHP-insensitive Ca2+ channel as well as a DHP-sensitive channel. The ability of MeHg to block DHP-sensitive or "L"-type Ca2+ channels in PC12 cells is important in light of recent studies which indicate that these channels may be responsible for, or at least modulate Ca2+ entry associated with release of some neurotransmitters from neurons (Perney et al., 1986; Atchison and O’Leary, 1987; Rane et al., 1987; Holz et al., 1988; Atchison, 1989) and PC12 cells (Takahashi et al., 1985 ; Kongsamut and Miller, 1986). In addition, DHP-insensitive Ca2+ conductances have been reported to mediate in part release of [3H]-norepinephrine from NGF-differentiated PC12 cells (Kongsamut and Miller, 1986). Thus, MeHg is capable of blocking at least one and possibly two types of Ca2+ channels which are associated with neurotransmitter release. This action may contribute in large measure to the observation that MeHg depresses Caz+-dependent transmitter release (Traxinger and Atchison, 1987b). High affinity binding of DHPs has been demonstrated in synaptosomes (Turner and Goldin, 1985; Suszkiw er al., 1986; Dooley et al., 1987b; Dunn, 1988). However, the relationship between binding of DI-IPs and effects of DHPs on 45Ca“ influx into synaptosomes is unclear. Several studies (Daniell et al., 1983; Suszkiw et al., 1986; 1989; Reynolds et al., 1986; Rivier et al., 1987) indicate that Ca2+ influx into . 122 synaptosomes is insensitive to block by DHP antagonists, whereas, other studies (Turner and Goldin, 1985; Dunn, 1989; Martinez-Serrano et al., 1989) indicate that a portion of Ca2+ influx in synaptosomes is sensitive to DHP antagonists. In either case, the concentrations Of DHP required to block 45Ca2+ influx are in great excess of those at which DHPs bind specifically to synaptosomes. In present study, [fin-nitrendipine bound to a single class of high-affinity sites in rat forebrain synaptosomes. The values for KD and Bmu correlate well with values observed by others for nitrendipine binding in synaptosomes (Turner and Goldin, 1985; Suszkiw et al., 1986). DHP agonists and antagonists are thought to bind to Ca2+ channels more readily when the channel is in the open or inactivated state, respectively (Uehara and Hume, 1985 ). However, depolarization of the synaptosomes did not readily affect the affinity of nitrendipine binding or the maximum number of binding sites. The binding affinity of the DHP antagonist (+ )-PN200-110 to synaptosomes was decreased by KI-induced depolarization (Dunn, 1989). Although other explanations are possible, the lack of effect of depolarization on the affinity of DHP binding would be consistent with the suggestion that DHP binding sites in synaptosomal preparations are associated with fragments of postsynaptic membrane which remain attached to the synaptosomes (Suszkiw et al., 1989). These fragments could not maintain a membrane potential and hence would not respond to changes in K" concentration. The affinity of nitrendipine binding to synaptosomes was decreased in the presence of a concentration of MeHg which readily blocks ”Ca2+ influx in synaptosomes (Shafer and Atchison, 1989) and DHP-sensitive influx in PC12 123 cells. The ability of MeHg to compete with nitrendipine for a high-affinity binding site in synaptosomes is consistent with its ability to block DHP-sensitive channels in PC12 cells. Experiments to demonstrate specific binding of nitrendipine to intact PC12 cells were unsuccessful, as appreciable specific binding was not observed. The author is unaware of any reports of specific binding of nitrendipine to intact PC12 cells. Caution must be observed in correlating the results of binding studies in synaptosomes with flux studies in PC12 cells and concluding that MeHg blocks a DHP-sensitive Ca“ channel in synaptosomes. According to one model for the movement of Ca2+ through DHP-sensitive cardiac Ca2+ channels, higher concen- trations of Ca2+ are necessary to observe a Ca“ current (Hess and Tsien, 1984) than are normally employed in experiments with synaptosomes. Thus, fluxes measured in these experiments may not necessarily be mediated by DHP-sensitive Ca2+ channels. As shown in Figure 4.5, MeHg will block 45Ca2+ influx into synaptosomes at con- centrations of Ca2+ high enough to mediate flux via btype channels. However, the contribution, if any, of “Ca“ flux into synaptosomes mediated by L-type channels under these conditions has not been determined. The polypeptide, to -conotoxin GVIA, binds to Ca2+ channels at a site which is distinct from the DHP binding site (Abe et al., 1987; Cruz and Olivera, 1986; Marqueze et al.; 1988; Sher et al., 1988) and in some axon terminals is thought to be associated with transmitter release. As opposed to nitrendipine, MeHg did not alter the binding of Cng to synaptosomes. This indicates that MeHg does not interact bit ICC rec: 117121; eXper recap) the oz C000(0) 124 with this site in synaptosomes, or that Cng displaces MeHg from this site. The effects of MeHg on 45Ca2+ uptake in synaptosomes are only partially reversible by increasing [Ca2+]c (Shafer and Atchison, 1989), and at neuromuscular junctions are only reversible under certain conditions (Traxinger and Atchison, 1987b). Thus, it seemed likely that MeHg is tightly or irreversibly bound to receptors on the cell surface. Consequently, experiments were designed to test whether pretreatment with MeHg might alter Cng binding to its receptor in synaptosomes. Results of these experiments suggest that MeHg does not interact with the Cng binding site in synaptosomes, as Cng binding was not altered by pretreatment with MeHg or in the presence of MeHg. In contrast to the results of synaptosomal experiments, MeHg did alter the binding of Cng to PC12 cells. The amount of Cng bound to PC12 cells was decreased significantly in the presence of MeHg at concentrations which effectively block 45Ca2“ entry into this cell line. Thus, MeHg apparently interacts with the Cng receptor in PC12 cells but not on synaptosomes. This suggests that the Cng receptor, and hence perhaps the Ca2+ channels, in synaptosomes and PC12 cells are different. to -Conotoxin is thought to interact with "N" and possibly "L" type channels in mammalian neurons (Reynolds et al., 1986). In both synaptosomes and PC12 cells, experimental results indicate the presence of only a single class of conotoxin receptor, thus it is possible that Cng binds to one channel type in synaptosomes and the other type in PC12 cells. However, it is difficult to interpret the results of conotoxin binding experiments, as the type(s) of Ca2+ channel(s) in synaptosomes are 125 unknown, and, in PC12 cells, only "N" type conductances are sensitive to block by Cng (Plummer et al., 1989). While the latter indicates that m -conotoxin GVIA may not block "L"-type channels in PC12 cells, it does not preclude binding to this channel type from taking place. Thus, it is to determine which Ca2+ channel subtype Cng is binding to in these experiments. Alternatively, Cng may be binding to a site in synaptosomes and/or PC12 cells which is not associated with Ca2+ channels. Much higher concentrations of Cng are required to block 45Ca2+ influx in mammalian synaptosomes than in avian or amphibian synaptosomes (Suszkiw et al., 1987), in spite of the presence of high affinity binding sites for Cng. Thus, interpretation of these data beyond the conclusion that MeHg interacts with the Cng receptor in PC12 cells, but not synaptosomes, would be purely speculative. MeHg is a very reactive heavy metal and alters a number of neuronal functions. Although effects of MeHg on Ca2+ channels are probably not solely responsible for the neurotoxic actions of MeHg, the rapid action of MeHg on Ca2+ influx into synaptosomes indicates that nerve terminal Ca2+ channels may be an initial site of interaction of MeHg with neurons. In summary, the results of this study suggest that MeHg interacts with nerve terminal Ca2+ channels in a manner unique from inorganic divalent cations or DHPs. MeHg clearly interacts with a DHP-sensitive Ca2+ channel in PC12 cells, but the relationship of this interaction to MeHg’s effects on 45Ca’“ influx in synaptosomes, and synaptic transmission at the neuromuscular junction is unclear. In addition, MeHg does not block Ca2+ channels in a state-dependent manner as do DHPs. MeHg is also unique from other 126 neurotoxic heavy metals in its ability to alter both the apparent inactivation kinetics and ionic selectivity of fluxes through the Ca2+ channel. Differences in lipophilicity and charge may underlie the unique blocking action of MeHg on Ca2+ channels. EFFECTS OF THE NEUROTOXICANT METHYLMERCURY ON NEURONAL on2+ CHANNELS: EFFECTS ON on“ CHANNELS INVOLVED IN TRANSMITTER RELEASE Volume H By Timothy Joseph Shafer A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHI]..OSOPHY Department of Pharmacology and Toxicology and Institute for Environmental Toxicology 1991 CHAPTERFIVE METHYLMERCURY BLOCKS N- AND L-TYPE Ca“ CHANNELS IN NERVE GROWTH F ACTOR-DIFFERENTIATED PHEOCHROMOCYTOMA (PC12) CELLS 127 128 ABSTRACT Effects of methylmercury (MeHg) on whole cell Ba2+ currents in rat pheochromocytoma (PC12) cells were examined. Based on biophysical characteristics and sensitivity to to -conotoxin GVIA (Cng) and dihydropyridine (DHP) agonists and antagonists, voltage-activated Ba2+ currents in PC12 cells were mediated by N- and L-type Ca2+ channels. Addition of MeHg (10p M) to the extracellular solution caused a rapid and complete block of current carried by 20 mM Ba“. The rate of block of 1,3,I by MeHg increased in a concentration-dependent manner between 1 and 20 ll M. Increasing the frequency of stimulation from 0.1 to 0.4 Hz facilitated block of IBa by MeHg. A 2 min application of 10 pM MeHg in the absence of stimulation also reduced 18, by approximately 80%. Thus, block of 13‘ by MeHg is not state- dependent. Additionally, MeHg blocked 13. when the membrane holding potential was -40, -70 and -90 mV, indicating that both N- and L—type Ca2+ channels are blocked by MeHg. Block of In. by MeHg was voltage-dependent at a membrane holding potential of -40 mV, but not at holding potentials of -70 and -90 mV. Decreasing the extracellular concentration of Ba2+ ([Ba2*]e) from 20 mM to 10 mM increased the magnitude of block by MeHg from 45.6 to 77.3%. Increasing [Ba2"]e to 30 mM mused no further antagonism of block. Block of IE, by MeHg was not reversed by washing with MeHg-free solution. The ionic permeability of PC12 cell Ca2+ channels was Ca2+ = Sr2+ > Ba“. In the presence of MeHg, all three divalent cations were equally permeant. These results indicate that: 1) MeHg blocks N - and L-type Ca2+ channels in PC12 cells at low micromolar concentrations; 2) block of 129 ' Ca2+ channels by MeHg occurs regardless of the state of the Ca2+ channel or membrane holding potential; 3) block of L-type Ca2+ channels by MeHg is voltage- dependent; 4) decreasing the extracellular [Ba2*] exacerbates the block by MeHg, but washing cells in MeHg-free solution did not reverse block of Ca2+ channels by MeHg; 5) the ionic permeability of Ca2+ channels in PC12 cells is altered by MeHg. 130 INTRODUCTION Results of experiments presented in previous chapters confirm important differences in the action of MeHg on synaptosomal 45Ca2+ influx and neuromuscular transmission when compared to divalent cations. While increasing [Ca2+]e readily reverses block by divalent cations of 45Ca2+ influx into synaptosomes (Nachshen, 1984; Suszkiw et al. , 1985 ) and nerve-evoked transmitter release at the neuromuscular junction (Nilson and Voile, 1976; Kober and Cooper, 1976; Cooper and Manalis, 1983), the effects of MeHg on synaptosomes (Atchison et al., 1986; Shafer and Atchison, 1989) and at the neuromuscular junction (Traxinger and Atchison, 1987b) are only partially antagonized by increasing [Ca2+]e. MeHg also alters the ionic selectivity and apparent rate of inactivation of 45Ca“ influx in synaptosomes and blocks synaptosomal 45Ca2" influx in a manner which is voltage-, but not state- dependent (Shafer et al., 1990). Such effects are not typically observed with inorganic divalent cations. Furthermore, MeHg blocks depolarization-induced influx of 45Ca2+ into PC12 cells and synaptosomes with similar potency and efficacy (Shafer et al., 1990) and alters binding of specific Ca2+ channel antagonists to their receptors on synaptosomes and PC12 cells. These data provide preliminary indications that MeHg interacts with Ca2+ channel types which are associated with neurotransmitter release. However, as noted in the previous chapter, the results of binding experiments do not allow for unambiguous determinations of channel types due to occasional lack of correlation of functional effects with specific binding. In addition, the kinetic resolution of 131 putative channel-mediated events using flux measurements is markedly slower than for electrophysiological measurements of channel activity in single cells. Thus, experiments were designed to make unambiguous determinations of Ca2+ channel types affected by MeHg and to examine the effects of MeHg on channel function using techniques with more physiologically relevant time resolution. Calcium or barium currents from PC12 cells cultured in nerve grth factor- (NGF) containing medium are mediated by separate channel entities with distinct kinetics and pharmacological sensitivities resembling the L- and N-type Ca2+ channels (Plummer et al., 1989; Usowicz et al., 1990) described by Fox et al. (1987a,b). These two channel types have been associated with neurotransmitter release in several different preparations (Atchison and O’Leary 1987; Rane et al., 1987; Holz et al., 1988; Atchison, 1989; Lemos and Nowycky, 1989). In the present study, PC12 cells and the patch voltage-clamp technique (Hamill et al., 1981) were used to examine directly the ability of MeHg to interact with N - and/ or L-type Ca2+ channels and to examine characteristics of Ca2+ channel block by MeHg such as reversibility of block, state and voltage-dependence of block, and effects of MeHg on Ca2+ channel inactivation and ionic permeability. 132 MATERIALS AND METHODS Chemicals and solutions. Stock solutions (500p M) of methylmercuric chloride (ICN Biomedicals, Plainview, NY) were prepared weekly in extracellular solution containing the proper concentration and species of divalent cation, from which test solutions were prepared daily. Stock solutions of nifedipine (Sigma, St. Louis, MO) and (+)-(S)-202-791 (generously donated by Sandoz Ltd., Switzerland) in 95 % ethanol were diluted in extracellular solution to the required concentrations (less than 0.05% ethanol after addition to extracellular solution). 0) -Conotoxin GVIA was purchased from Peninsula Laboratories (Belmont, CA). Unless otherwise noted, extracellular solution contained (mM): HEPES (20), MgCl2 (1.0), d-glucose (10), NaCl (115), and TEACl (20) pH 7.3. This solution was mixed with appropriate volumes of 100 mM BaClz, CaCl2 or SrCl2 to yield the final concentration of divalent cation indicated in the text (usually 20 mM BaClz). Tetrodotoxin (TTX, 5 nM) purchased from Sigma (St. Louis, MO) was added to block voltage-sensitive Na+ channels (Narahashi et aL, 1964). Internal (pipette) solution contained (mM): I-IEPES (10), MgC12(1), d-glucose (10), TEACl (10), EGTA (5), CsCl (125) and ATP (Sigma, St. Louis, MO) (2), pH 73. TBA", Ba“ (Hagiwara et al., 1978; Armstrong et al., 1982) and Cs+ were used in external and / or internal solutions to suppress K+ current in PC 12 cells. ATP was added to internal solutions to prevent time-dependent rundown of current amplitude (Bean, 1985). 133 Culture of PC12 cells. PC12 cells were cultured at 37°C in a humidified 10% CO2 atmosphere using Dulbecco’s modified Eagle medium (pH 7.3, GIBCO) containing 10% horse serum, 5% fetal bovine serum and 2 mM HEPES. Inasmuch as certain antibiotics have been shown to be capable of blocking Ca2+ channels (Atchison et al., 1988), antibiotics were omitted from the culture medium. PC12 cells were differentiated by addition of 100 ng/ml of 7s NGF. To maintain consistency from experiment to experiment, cells from the 15th passage from our receipt of the cell line were used for recording Ba2+ currents. Current recordings. Cells were plated into 35 mm polystyrene dishes and cultured for 5-7 days in 2 ml of medium containing NGF. Prior to recording Ba2+ current, the culture medium was removed and cells were washed 2-3 times and then covered with 1 ml of extracellular solution. Test compounds were prepared in extracellular solutions such that addition of a 500 pl aliquot to the dish would result in the final concentrations of test compound indicated in the text and figure legends. This method ensured rapid mixing of solutions and maintained osmolarity of the extracellular solution. Voltage-activated Ba2+ currents in PC12 cells were recorded using the whole cell patch voltage-clamp method (Hamill et al., 1981). Resistance of patch electrodes (1.2 mm glass, World Precision Instruments, New Haven, CT) was between 2-6 M0 . The patch clamp circuit consisted of an Axon Instruments CV-1 headstage and Axopatch 1B patch clamp. Pulse protocols were generated and current responses recorded on-line using a Compaq 386 micro computer and the pClampm program interfaced to the Axopatch 13 via a Axon Instruments TL-l interface board. 134 For analysis of current-voltage relationships, a 255 ms pulse to each indicated test potential was applied from a holding potential of -40, -70 or -90 mV. Leak current was measured at the appropriate holding potential by pulses which were equal in duration, but opposite in polarity and V2 the magnitude of the test potentials. To examine the reversibility of MeHg-induced block of In. by washing in MeHg-free solutions, inward (holding potential -90 mV, test potential + 10 mV, 255 ms pulses) and leak currents were recorded from PC 12 cells in the absence of MeHg. MeHg was then added to the extracellular solution for 1 min and the currents were recorded again. All but approximately 100 - 200 ll] of extracellular solution were then removed from the dish and replaced with a 1 ml aliquot of MeHg-free extracellular solution. This was repeated 3 times to wash the cells and current was recorded again. Current responses were filtered at 10 KHz using a 4-pole lowpass Bessel filter. Following scaling, linear leak and capacitive currents were subtracted from gross current using the pClampTM program. Pulse protocols and filter settings for other experimental protocols are described in the Results section. Series resistance remaining after compensation was less than 2 M0 , as 80% or greater compensation was commonly achieved. Recordings which contained indications of inadequate space clamp, such as pronounced shoulders during current activation, were not included in analysis. All recordings were made at room temperature (22-25°C). A minimum of 3 cells were recorded from for any given experiment and dishes of cells were discarded after addition of test solutions. 135 RESULTS Description of In. in PC12 cells. The pharmacological and biophysical properties of Ba“ current in differentiated PC12 cells were characterized in order to confirm the identity of current components presumably mediated by N - and/ or L- type Ca2+ channels (Plummer et al., 1989; Usowicz et al., 1990). Ba2+ currents elicited by a test potential to + 10 mV from holding potentials of -40 (Figure 5.1A), -70 (Figure 5.1B), or -90 mV (Figure 5.1C) contained non-inactivating and inactivating components (-70 and -90 mV only). The time constant, 1: , for the inactivating component of currents at a test potential of + 10 mV averaged 225 1 86 ms (n= 8) and 175139 (n= 15) ms in cells held at -90 and -70 mV, respectively. A family of Ba2+ currents is illustrated in Figure 5.2. The holding potential was -90 mV and test potentials began at ~50 mV and were increased in 10 mV increments to +70 mV. Components of Ba2+ current in PC12 cells are sensitive to block by the N- channel antagonist Cng (Figure 5.3A). The inactivating component of current was apparently eliminated, while a non-inactivating component remained. Nifedipine (20 nM), a DHP Ca2+ channel antagonist, substantially reduced 18, (data not shown), whereas (+)-(S)-202-791 (<5 uM), a pure DHP Ca2+ channel agonist (Hof et al., 1985), enhanced 13, during voltage steps (Figure 53B) and induced tail currents upon repolarization (Figure 5.3, inset). Thus, Ba“ currents recorded from differentiated PC12 cells cultured in this laboratory have pharmacological and biophysical profiles consistent with currents mediated by N- and L-type Ca2+ channels, respectively. 136 -40— mV _70___ ~90—’ _ l — — Figure 5.1. inactivating and non-inactivating components of Ba“ current in PC12 cells. Ba2+ current in diflerentiated PC12 cells in response to a voltage step to + 10 mV from membrane holding potentials of ~40 (A), ~70 (B) or ~90 (C) mV. Command potentials shown below current traces were 255 ms long. linear components of leak and capacitance have been subtracted. Current responses were filtered at 10 1012. 137 Figure 5.2. Current responses to a series of test potentials. Current responses in a differentiated PC12 cell to test potentials between ~50 and +70 mV when the magnitude of the test potential is increased in 10 mV increments. Membrane holding potential was ~90 mV and test potentials were 255 ms in duration. Current responses were filtered at 10 KHz and linear components of leak and capacitance current have been subtracted. 138 C ng CON 100 pA L— 30 ms [— 8202-791 B "i CON W 8202-791 1 3°PAL_ 30ms Figure 5.3. Effects oral-conotoxin and (+)~8202~791 on Ba2+ current in PC12 cells. A) Efiect of Cng (16 MM) on 13. in a difierentiated PC12 cell. The current response is to a 255 ms step to + 10 mV from a holding potential of ~90 mV. Current response is filtered at 10 KHz. B) Effect of (+)~(S)~202-791 (< 5 nM) on 13. in a difierentiated PC12 cell. The current response is to a 755 ms step to + 10 mV from a holding potential of ~70 mV. Inset: Tail currents recorded upon repolarization to ~70 mV from a 20 ms step to + 10 mV prior to and after addition of (+)~(S)~202-791. Scale bars are 400 pA and 7 ms. All current responses were filtered at 10 KHz and linear components of leak and capacitance were subtracted. Responses shown are the average of 5 trials prior to and after addition of Cng or (+)~(S)~ 202-791. 139 Efl‘ect of MeHg on In. in PC12 cells. The effects of MeHg on membrane properties and currents were examined using the pulse protocol shown in Figure 5.4A. A depolarizing pulse was delivered to elicit inward current, followed by a hyperpolarizing pulse so that the leak current in the cell could be measured. Linear components of leak and capacitive current were not subtracted fi'om these records. Thus, effects of MeHg on inward current, presumably mediated by Ba“, leak current and capacitive current can be examined in consecutive current traces. Addition of MeHg (10 nM) to the extracellular solution resulted in a rapid block of inward current (Figure 5.4A), but did not alter membrane leak current nor membrane capacitance, indicating that the effect of MeHg is not due simply to disruption of membrane electrotonic properties. When this depolarizing pulse was repeated once every 5 5, addition of increasing concentrations of MeHg to the extracellular solution resulted in a concentration-dependent increase in the rate and magnitude of block of 13,, following leak subtraction (Figure 5.4B). Ba2+ current amplitude in PC12 cells often increased during the first 1-1.5 min after adjusting whole cell capacitance and series resistance compensation controls (Figures 5.4B and 5.5, data prior to MeHg addition), after which time current amplitude remained stable for beyond 10 min. This initial increase in current may be associated with exchange of cytosolic constituents and pipette solutions, resulting in block of K+ channels by Cs+ and TEA+ and/ or phosphorylation of Ca2+ channels by ATP-dependent mechanisms. To ensure that effects of MeHg were examined on stable Ba2+ current, cells were held at rest for 2 min before recording 13. in all 140 O) O o E C _m u. 0 Z 9. l- 0 < 0: LL 0 100 200 300 TIME (s) Figure 5.4. Efl'ects of MeHg on currents in PC12 cells. A) Effect of 10 u. M MeHg on gross inward and leak current measured using whole cell patch voltage-clamp techniques on a diflerentiated PC12 cell. As shown in the command potential protocol below the current traces, membrane potential was clamped at ~70 mV and current responses to the designated voltage steps (100 ms ea.) were recorded. This - voltage protocol was repeated once every 5 5. Current traces shown were recorded immediately before addition of MeHg (CON), and 25, 50 and 100 s after addition of MeHg to the recording solution. Current responses were filtered at 1 KHz and leak current has not been subtracted. B) Efiect Of increasing concentrations of MeHg on peak Ba2+ current measured in response to the pulse protocol in A after leak subtraction. Pulses were repeated once every 5 s and recording medium containing MeHg was added after 100 s, as indicated by the arrow. Final concentrations of MeHg were: 0 (upright triangles), 1 (solid circles), 2.5 (solid diamonds), 5 (solid squares), 10 (open circles) or 20 (inverted triangles) u. M. Peak current in each trace was normalized to peak current measured prior to the addition of MeHg. Values shown are the average of peak meat from 4—8 different cells and average standard error values were 23% of mean values throughout the experiment. 141 subsequent experiments, except for those designed to examine frequency-dependence of block. In addition, all experiments were completed within 10 min of adjusting patch circuit controls. Use-dependence of block of In. by MeHg. The ability of MeHg to interact with Ca2+ channels in PC12 cells in a use-dependent manner was assessed by examining effects of MeHg on Ba2+ currents elicited by stimulating PC12 cells at frequencies between 0.1 and 0.4 Hz. At the fastest rate of stimulation, the time required for MeHg (5 ll M) to block approximately 70% of the current (approximately 3 min) was less than the time required for MeHg to block current at slower rates (approximately 6 min) of stimulation (Figure 5.5). Thus, increasing the frequency of stimulation, and hence use of Ca2+ channels, appears to facilitate slightly their block by MeHg. Although increasing the frequency of stimulation appears to facilitate block of Is. by MeHg, Ba2+ current was reduced appreciably in MeHg-containing solutions even at relatively low rates of stimulation (0.1 Hz). In addition, block of synaptosomal 45Ca2“ influx by MeHg is not state-dependent (Shafer et al., 1990). Thus, experiments were designed to determine if activation of Ca2+ channels is essential for the blocking action of MeHg. Accordingly, effects of MeHg on In. in PC12 cells were examined following a period of inactivity during which MeHg was added to the extracellular solution. The protocol was as follows: a series of 10 voltage steps from ~70 to + 10 mV (255 ms duration) was delivered at 3 5 intervals. Leak current, which was subtracted after scaling and averaging, was then measured by a series of 20 steps from ~70 to ~110 mV. After measurement of leak current, 5 142 FRACTION OF I at 100 s O 200 400 600 TIME (s) Figure 5.5. Use-dependent effects of MeHg on In. in PC12 cells. Effect of increasing the rate of stimulation on peak inward current in the absence (solid symbols) or presence of 5 11M MeHg (open symbols) in differentiated PC12 cells. The pulse protocol used in Figure 5.4A was repeated at frequencies of 0.01 (squares), 0.2 (triangles) and 0.4 (circles) Hz. Peak current in each trace was normalized to peak current measured prior to the addition of MeHg. Values shown are the mean of 4-8 different cells and average standard error values were 24% of mean values throughout the experiment. Current responses were filtered at 1 KHz and leak current has been subtracted. Control values for 0.2 and 0.4 Hz were not difl'erent from values for 0.1 Hz and are not illustrated. 143 or 10 uM MeHg was added to the extracellular solution, and 2 min later the voltage protocol described above was repeated. When no MeHg was added to the extracellular solution, no significant changes were observed in peak or end current (results not shown). After a 2 min exposure to 5 uM MeHg in the absence of stimulation, the average amplitude of the first current response was reduced when compared with the average amplitude of the first current response in the absence of MeHg. Peak and end currents were reduced by an average of 32% and 26%, respectively (Figure 5.6, bottom panel). Addition of 10 uM MeHg resulted in significant reductions (paired t-test, p < 0.05) of 77 and 70% for peak and end currents, respectively, when compared to control values (Figure 5.6, bottom panel). In addition, 10 u M MeHg caused complete elimination of the inactivating component of whole-cell current (Figure 5.6, top panel). in none of four cells to which 10 uM MeHg was added could an estimate of r be obtained for the first current response in the presence of MeHg, and no current responses containing inactivating components were observed. Holding Potential- and Voltage-Dependence of Block by MeHg. Current- voltage relationships for 13, in PC12 cells were obtained from holding potentials of ~40, ~70, and ~90 mV in the presence and absence of 10 uM MeHg (Figure 5.7). At all three holding potentials, a 2 min exposure to MeHg decreased the magnitude of peak Ba2+ current elicited at each test potential. There was no readily apparent holding potential-dependence of the action of MeHg. Currents elicited by depolarizing steps from a holding potential of ~40 mV were blocked in a voltage- 144 AWWI‘ IBcr M“ 200 pA L— 30ms ICON .Mel-lg "oi PEAK CURRENT { 390 - 600 300 INWARD Ba CURRENT AMP. (pMUJ 5pm 109“ Figure 5.6. Resting block of 13‘ by MeHg in PC12 cells. A) In“ in a differentiated PC12 cell recorded prior to (CON) and after MeHg (10 u M) was present in the recording solution for 2 min. B) Amplitude of inward Ba2 current recorded before (solid bars) and after (striped bars) a 2 min. exposure to 5 or 10 it M MeHg. The asterisk (‘) indicates a significant reduction in current amplitude in the presence of MeHg when compared to control values. (Paired t-test, P< .05). Values shown are the mean: SEM of 7 and 4 diflerent cells exposed to 5 or 10 u. M MeHg, respectively. Current responses are to a 255 ms voltage step to + 10 mV from a membrane holding potential of ~70 mV and were filtered at 10 KHz. Linear components of leak and capacitance current have been subtracted. 145 TEST POT. (mV) -4O -20 O 20 4O 60 Tl/‘o — UN- / A \‘O‘N '0---o- L ‘0’ /; A 4' T 3/' < ‘75 3/ Q h.p.=-40 v '150 -60-40 '20 O 20 4O 60 8O ’0 ”k? \5. T 5’ B \3 \°‘ ' I T I,~I/’ / ‘ ,o’ \I‘ <,\\C"°’ '/ ‘" °°“ \--200 /3/ -0- Mai-lg “\l. l/ /..hp =-70 ./. -400 -60-40 ~20 O 20 4O 60 so - R /,_ ‘k.\\‘a\ ’5/25 C “It. {,9 -/' (V -/' inv/ ;/ 4501/ °\" T h.p.--90 . Ba -20mM LI/ I I -700 PEAK CURRENT Figure 5.7. Efl'ects of MeHg at different holding and test potentials in PC12 cells. Current-voltage relationships for 13. in differentiated PC12 cells recorded in the absence (filled symbols) and presence (open symbols) of 10 uM MeHg. Voltage steps (255 ms) to the test potentials (TEST POT.) indicated in the figure were applied from holding potentials of ~40 (A), ~70 (B) and ~90 (C) mV. Current responses were filtered at 10 KHz and following subtraction of leak and capacitance current, peak current was measured. Results shown are the mean: SEM of peak current from 4 to 8 cells, each cell received all test potentials. 146 dependent manner by MeHg. The magnitude of block by MeHg increased from 23 to 98% as the test potential was increased from ~20 to +40 mV.Block of 18a by MeHg was not strongly voltage-dependent at holding potentials of ~70 and ~90 mV. At the holding potential of ~90 mV, the magnitude of reduction of 18a by MeHg at the most weakly and most strongly depolarizing test potentials (~30 and + 70 mV) was greater than the magnitude of reduction of IBa by MeHg at test potentials which elicit peak IBa (0 and + 10 mV). MeHg did not alter the test potentials at which activation of [Ba or peak IBa were observed. However, MeHg caused a negative shift in the apparent reversal potential of In, at all holding potentials. To determine whether irreversible, voltage-dependent block of Ca2+ channels by MeHg occurred, block of IBa by 5 and 10 u M MeHg was examined using a holding potential of ~40 mV; test pulses (255 ms) were increased by 10 mV increments from ~ 30 to +50 mV and then decreased in 10 mV increments back to ~30 mV (Figure 5.8). If MeHg blocks Ca2+ channels in an irreversible, voltage—dependent, manner, one would expect a greater magnitude of block as test pulses are decreased from +50 to - 30 mV than that observed when test potentials were increased from ~30 to +50 mV. Current amplitudes and effects of MeHg were similar in each case, indicating that MeHg does not bind irreversibly to Ca2+ channels in a voltage-dependent manner. Effects of [Ba’*]e on block of In. by MeHg. Previous studies in synaptosomes indicated that at non-saturating [Ca2*]e, increasing [Ca2+]: antagonizes somewhat the magnitude of block of “Ca2+ influx by MeHg (Atchison et al., 1986; Shafer and Atchison, 1989). Therefore, it seemed prudent to determine if the action of MeHg 147 on In. depended on [Ba2+]c. Decreasing [Baz‘i]¢ resulted in decreases in the control current as well as currents measured after 2 min exposure to 10 uM MeHg (Figure 5.9). At a test potential of + 10 mV, MeHg caused a 77.3% reduction when compared to control current for solutions containing 10 mM Ba2+ and a 45.6% for those containing 20 mM Ba“. In 30 mM Ba2+ solutions, inward currents were not significantly larger than in 20 mM Ba2+ solutions, indicating that these concentrations saturate Ca2+ channels in PC12 cells. In 30 mM Ba2+ solutions, exposure to 10 NM MeHg reduced IBa by 49.8% (results not shown). Thus, [Ba2+]¢ antagonizes the blocking action of MeHg, but is not capable of overcoming completely the effects of MeHg. As in the above current-voltage relationships, MeHg also caused a negative shift in the apparent reversal potential for 18., and an outward current was observed in the presence of MeHg when [Ba2*]e= 10mM. Reversibility of MeHg-induced Block by Wash. To determine if MeHg acted on Ca2+ channels in a fashion similar to inorganic divalent cations or if MeHg acted on Ca2+ channels in an irreversible fashion, the reversibility of MeHg-induced block of In. in PC12 cells by washing cells with MeHg-free solutions was examined. Effects on 13, of exposure to 10 uM MeHg were not reversed by washing with MeHg-free solutions in 4 attempts (results not shown). In one cell, block of In. by 5 uM MeHg was also not reversed by washing with MeHg-free solutions (results not shown). In contrast, block of In, by 1 u M Cd2+ was reversed readily by washing with Cd2+-free solutions (results not shown). The ability of micromolarconcentrations of Cd2+ to block In. is consistent with previous results in cultured cells (Boland and Dingledine, 148 TEST POT.(mV) -40‘-20 0 20 4O 60 .9 Q. 7§§ 7,6 \ \ V ' , Ntin‘L~l/rl I . \ A) l .b O I N O O N O .5 O O) O PEAK CURRENT ( J \\7 Y .. ”£0171;pr may Figure 5.8. Voltage-dependent block by MeHg of in. in PC12 cells is not irreversible. Current-voltage relationship for In. in differentiated PC12 cells in the absence (filled symbols) and presence (open symbols) of 5 or 10p. M Mei-lg. From a holding potential of ~40 mV, test pulses (255 ms) were delivered in increasing increments (circles) of 10 mV from ~30 to +50 mV, followed by decreasing increments (triangles) of 10 mV from +50 to ~30 mV. Current responses were filtered at 10 KHz. Following subtraction of leak and capacitance current, peak current was measured. Results shown are the mean 1 SEM ofpeakcurrentfrom4ceils. 1: Q. Ce. 149 TEST POT. (mV) -60- 40 ~20 o 20 4o 60 30 _th._¢__Tr - O”° 0\._2 O 0 ”0,4; _/ A \3 \(Pho’ .0 _/o < A \ /e/. 0.. 3 ~ . V --o- CON - 3 5 of ./' IE ‘0'” MeHg \. h.p.=-9 0 $ [Ba]=1 OmM g - 7 00 O ~6£‘-:40 ~20 0 20 40 60 - 80' x B 3 \\ ll’fi’ O < 1L 1 x i/ E \\ 1 i/ i 3/ l O \ 2’ T/ ' - 3 5 O /I ’\ h. p. =-9 0 'i\ \l/ [Ba]=20mM - 7 00 Figure 5.9. Ba2+ can antagonize block of Ca2+ channels by MeHg. Current-voltage relationships for 13. in difl’erentiated PC12 cells recorded in the absence (solid circles) and presence (open circles) of 10 M MeHg in recording solutions containing 10 (A) or 20 mM BaC12 (B). Voltage steps (255 ms) to the test potentials (TEST POT.) indicated in the figure were applied from a holding potential of ~90 mV. Current responses were filtered at 10 IQ-Iz and following subtraction of leak and capacitance current, peak current was measured. Results shown are the mean 1 SEM of peak current from 4 to 8 cells, each cell received all test potentials. 150 1990), as is the ability of washing with Cd2*~free solutions to reverse the block of IBa in cultured cells (Boland and Dingledine, 1990) and block Of ”Ca“ influx into synaptosomes (Nachshen, 1984) by Cd“. Thus, effects of MeHg on Ca2+ channels are not readily reversible. Ionic permeability of Ca2+ channels in PC12 cells. Currents mediated by Ba2+ (Plummer et al., 1989; Usowicz et al., 1990) and Ca2+ (Garber et al., 1989) have been measured in NGF-differentiated PC12 cells. However, the author is unaware of any previous reports comparing the permeability of PC12 cell Ca2+ channels for divalent cations (Ca2+, Ba2+ and Sr“). MeHg altered the ionic selectivity of synaptosomal Ca2+ channels by decreasing influx of 8SSr“ to a greater extent than influx of radiolabeled Ca2+ and/ or Ba2+ (Shafer et al., 1990). Thus, currents mediated by Ca2+ channels using 10 mM Ca“, Ba2+ or Sr2+ as the charge carrier in extracellular solutions were examined and effects of MeHg on ionic permeability of PC12 cell Ca2+ channels were characterized. To prevent differences in cell size fiom influencing the results, recordings were made using solutions of Ca2+, Ba2+ or Sr2+ which were labelled as unknowns, and results from recordings made from 46 cells were pooled. Only after current records had been analyzed was the identity of the divalent cation in the extracellular solution revealed. The permeability sequence of Ca2+ channels in differentiated PC12 cells was Ca2+ = Sr“ > Ba2+ when the holding potential was ~90 mV (Figure 5.10). This observation was consistent throughout several cell passage numbers (10, 11 and 15) and days in culture (4, 5, 7 and 12) in the presence of NGF. MeHg (10 nM) reduced current mediated by all three divalent cations. However, the 151 apparent ionic selectivity of Ca2+ channels for Ca2+ and Sr2+ over Ba“ was eliminated in the presence of MeHg. 152 TEST POT. (mV) -60-40 -20 O 20 4O 60 80 '__I~.V \\L\o I ‘I/5 .// _o--1OIIIMB|-\\H I./O., [Kl i\/ ’ -A-10mMCa ‘{ ‘ 441’ L \ ./ ---10nwlst . \ I \., -60-4O ~20 O 20 40 60 80 PEAK CURRENT (pA) -200 Figure 5.10. MeHg alters ionic permeability or Ca2+ channels in PC12 cells. Current-voltage relationships for current mediated by 10 mM Ca2+ (triangles), Sr2+ (squares) or Ba2+ (circles) in the absence (upper panel) or presence (lower panel) of 10 u. M MeHg. Voltage steps (255 ms) to the test potentials (TEST POT.) indicated in the figure were applied from holding potentials of ~90 mV. Current responses were filtered at 10 KHz and following subtraction of leak and capacitance current, peak current was measured. Results shown are the meant: SEM of peak current from 14-15 and 6 cells in the absence and presence of MeHg, respectively, each cell received all test potentials. 153 DISCUSSION Results of the present study indicate that 1) NGF-differentiated PC12 cells cultured in this laboratory contain both N- and L-type Ca2+ channels; 2) low micromolar concentrations of MeHg block IBa in PC12 cells and the rate of block is concentration-dependent; 3) increasing the frequency of channel stimulation may facilitate block of 13, by MeHg, but channel use is not required for the blocking action of MeHg; 4) block of 13, by MeHg is not holding potential-dependent; 5) block of 18a by MeHg is antagonized, but not completely overcome, by increasing [Ba2+]c; 6) block of 13:: by MeHg is voltage-dependent when the holding potential is ~ 40 mV, but is not strongly voltage-dependent when the holding potential is ~70 or ~90 mV, 7) block of In, by MeHg is not reversed by washing with MeHg-free solutions; 8) the apparent ionic selectivity of whole cell current mediated by Ca2+ channels in PC12 cells is Ca2+ = Sr2+ > Ba“, and MeHg alters the apparent ionic selectivity of Ca2+ channels in PC12 cells. These results suggest that block of 18a in differentiated neuron-er cells mirrors block of radiotracer flux into synaptosomes. Moreover, the blocking properties of MeHg appear to be unique from inorganic divalent metals. NGF~differentiated PC12 cells express both N- and L-type Ca2+ channels (Plummer et al., 1989; Usowicz _e_t a1, 1990). The presence of T~type channels in PC12 cells has also been reported (Garber et al., 1990). Due to variability between PC12 subclones it is important to characterize the biophysical and pharmacological properties of Ca2+ channel types present prior to making conclusions about effects of uncharacterized chemicals on Ca2+ channels in these cells. Ba2+ current from cells 154 cultured in the presence of NGF in this laboratory contained no inactivating component when held at ~40 mV, and contained inactivating and non-inactivating components when held at more hyperpolarized membrane potentials. The average time constant for decay of the inactivating portion of the current was somewhat slower than that reported previously for N-type channels and current in PC12 cells (Plummer et al., 1989), although many individual cells contained currents with r values of around 95 ms. Furthermore, Ba2+ current from cells used in these experiments exhibited characteristic responses to DHP L-channel type agonists and antagonists and were blocked by to ~conotoxin GVIA, which affects exclusively N-type channels in PC12 cells (Plummer, et al., 1989; Usowicz et al., 1990). Thus, based on both the biophysical and pharmacological characteristics of whole cell current, these results clearly indicate the presence of N- and L-type Ca2+ channels in PC12 cells cultured in this laboratory. Block of Ba2+ current in PC12 cells is rapid and complete upon addition of 10 ll M MeHg to the extracellular solution. Leak and capacitative currents in PC12 cells were not affected by MeHg, indicating that effects of MeHg are relatively specific for channel function. Both N~ and L-type channels in PC12 cells are affected by MeHg, as MeHg blocks Ba2+ current at all holding potentials tested. These results confirm results Of previous binding experiments which indicated that MeHg interacts with N- and Ltype channels in PC12 cells and synaptosomes by interfering with Cng and DHP binding, respectively (Shafer et al., 1990). The concentration-dependency for block of 13. occurred at lower concentrations (1-20 u M) of MeHg than did block of 155 4S’Ca“ influx (50—100 nM) by MeHg into PC12 cells (Shafer gt .al, 1990) and synaptosomes (Shafer and Atchison, 1989). Although this discrepancy does not appear extremely large, one must consider differences in divalent cation concentration used for electrophysiological (10-20 mM Ba“) compared to flux (0.05 - 1 mM Ca2+) measurements. There are small but significant differences in the techniques used to measure Ca2+ influx and Ba2+ current; namely continuous and comparatively long (1-10 5) intervals of depolarization yet brief exposure (only 1-10 5) to MeHg for flux studies as opposed to short (100-250 ms), repeated step depolarizations during a several minute period of exposure to MeHg for Ba2+ current measurements. Higher concentrations of MeHg may be necessary to block Ca2+ channels due to the much shorter overall lengths of exposure to MeHg during flux assays. Block of Na+ channels by local anesthetics (Strichartz, 1973; Courtney, 1975; Hille, 1977) and Ca2+ channels by organic Ca2+ channel antagonists (Uehara and Hume, 1985) has been described as “use-dependent", indicating that these compounds bind preferentially to the open or inactivated state of the channel. The results of the present experiments suggest that MeHg effectively blocks Ca2+ channels in the resting state, yet the blocking action can be facilitated by channel use. Thus, the action of MeHg on Ca2+ channels is not state-dependent. This is consistent with results obtained using synaptosomes (Shafer et al., 1990). DHP Ca2+ channel blockers substantially reduce 10, in the absence of prior stimulation and have use—dependent actions only during high-frequency stimulation (Lee and Tsien, 1983; Uehara and 156 Hume, 1985). By contrast, the actions of verapamil and D-600 are strongly use- dependent, whereas diltiazem exhibits both resting and use-dependent block of ICa (Lee and Tsien, 1983; Uehara and Hume, 1985). Furthermore, multivalent inorganic cations (Nachshen, 1985) act on synaptosomal Ca2+ channels in a use-dependent manner. MeHg, like the organic Ca2+ channel blockers, is lipophilic, but it is also similar to divalent cations in size and is charged. Thus, MeHg may have access to the Ca2+ channel by more than one route, including diffusion through the membrane and entry through the channel pore. Perhaps the combination of charge and lipophilicity of MeHg may account for its ability to block Ca2+ channels in the resting state, while increasing frequency of stimulation facilitates the block. Addition of MeHg in the absence of stimulation consistently eliminated the inactivating component of current. This action of MeHg is unique compared to that of Cd2+ and Ni“, which do not alter the time course of decay of 13, in a dorsal root ganglion cell line (Boland and Dingledine, 1990). N—type Ca2+ channels in PC12 cells may be more sensitive to block by MeHg than L-type channels. Alternatively, MeHg may impair the ability of N-type channels to inactivate during the voltage step. This latter mechanism would also require a simultaneous reduction in single channel conductance or number of open Ca2+ channels (open probability) by MeHg to account for the decreased current amplitude when MeHg is present. Either mechanism is consistent with the results of the present experiments; single channel recordings would be required to distinguish between them. MeHg altered the inactivation kinetics of synaptosomal 45Ca2“ influx; the rate constant for inactivation 157 was increased and the amount of non-inactivating influx was decreased (Shafer et al., 1990). Thus, MeHg produces analogous actions on inactivating and non-inactivating current in PC12 cells and synaptosomes. MeHg blocked Ba2+ currents elicited by depolarizing steps to positive potentials from holding potentials of ~40, ~70 and -90 mV. The blocking action of MeHg was not strongly voltage-dependent at holding potentials of ~70 and ~90 mV, except that the degree of block was less at test potentials near the peak of the current-voltage relationship when the holding potential was ~90 mV. This may be related to the ability of Ba2+ to antagonize the blocking action of MeHg, as the driving force for Ba2+ entry through Ca2+ channels and the number of channels in the open state is greatest at these test potentials. Block of IBa was voltage-dependent at ~40 mV. Presumably, there is very little N-type current at this holding potential due to steady-state inactivation of N channels (Fox et al., 1987a,b). Thus, MeHg appears to block L-type Ca2+ channels in a voltage-dependent manner, a result consistent with voltage-dependent block of MeHg on synaptosomal ”Ca2+ influx (Shafer et aL, 1990). The hydrophobic nature of MeHg may be important in this regard, as organic, but not inorganic mercurials block 45Ca2“ influx into synaptosomes in a voltage-dependent manner (Hewett and Atchison, 1990). Additionally, MeHg caused a negative shift in the apparent reversal potential of In, at all holding potentials and for all Ba2+ concentrations tested. Under the experimental conditions used, the reversal potential for In, is theoretically positive to + 80 mV. An actual outward Ba2+ current is not expected under the present 158 experimental conditions, due to the presence of EGTA in the internal solution. We cannot eliminate the possibility that MeHg interferes with block of K+ channels by TEA+ or Cs“, but does not block K+ channels itself, thus giving rise to outward currents in the presence of MeHg. Outward currents have also been observed in isolated cardiac myocytes (Lee and Tsien, 1982) and in bovine adrenal chromaffin cells (Fenwick et al., 1982) when K+ or Cs+ permeate Ca2+ channels at highly positive test potentials. However, Ca2+ channel blockers such as D600 or Cd2+ effectively block these outward currents (Lee and Tsien, 1982). Thus, any potential outward Cs+ currents should be blocked by MeHg. The present data do not suggest any particular mechanism underlying this effect of MeHg. Washing PC12 cells in MeHg-free solutions did not reverse the block of 13, by MeHg in any of the cells tested, although block of IBa by Cd2+ was reversed readily by washing. Furthermore, extracellular Ba2+ can antagonize, but not completely overcome the blocking actions of MeHg. This is consistent with effects of increasing [Ca2+]c on the effects of MeHg on synaptosomal ”Ca2+ influx (Shafer and Atchison, 1989). These results are also consistent with previous experiments at the neuromuscular junction which indicate that the effects of MeHg on synaptic transmission are not readily reversed by washing with MeHg-free solutions and can only be partially reversed by increasing [Ca2“],3 (Traxinger and Atchison, 1987b). Block of neuromuscular transmission by Pb“ (Cooper and Manalis, 1983; Atchison and Narahashi, 1984; Pickett and Bornstein, 1984), Cd“ (Nilson and Voile, 1976; Cooper and Manalis, 1984a,b), or Co2+ (Weakly, 1973) is reversed by increasing 159 [Ca2+]c or washing with normal, metal-free solutions. This represents a fundamental difference in the action of MeHg when compared to divalent heavy metals. The mechanism by which MeHg interacts with Ca2+ channels in an essentially irreversible manner is unknown, although reactivity of MeHg with sulfhyde groups on membrane proteins and/ or the enhanced lipophilicity of MeHg due to the presence of the methyl group could be involved. Ionic permeability ratios in PC12 cells have not been previously reported. Based on whole cell currents, the permeability of PC12 cell Ca2+ channels was Ca2+ = Sr2+ > Ba“. This ionic permeability sequence was somewhat unexpected, as Ba2+ usually is not the least permeant of the three cations based on single channel recordings (Tsien et al., 1988). Addition of MeHg to the extracellular solution resulted in a loss of apparent ionic selectivity of Ca2+ channels, a result consistent with the ability of MeHg to alter ionic selectivity of synaptosomal 45Ca2+ influx (Shafer et al., 1990). In conclusion, MeHg interacts rapidly with N- and L-type Ca2+ channels in PC12 cells at low micromolar concentrations. Given the ability of MeHg to interfere with binding of Ca2+ channel antagonists (Shafer et al., 1990), pharmacological methods alone cannot be used to examine the effects of MeHg on individual Ca2+ channels subtypes. Further characterization of the effects of MeHg on the individual Ca2+ channel subtypes will require single channel recording techniques. Concentrations of MeHg which block In, in PC12 cells are consistent with those which block evoked endplate potentials at the neuromuscular junction (Atchison and Natal inpOi have area net) pret to b dbr 160 Narahashi, 1982). The ability of MeHg to block both N- and L-type channels is important in this respect. At the neuromuscular junction, L-type Ca2+ channels can have at least a modulatory role in neurotransmitter release (Atchison and O’Leary, 1987), but do not contribute to nerve-evoked release under normal conditions (Atchison, 1989). L-type channels have also been implicated in release of substance P fi'om chick dorsal root ganglion cells (Rane et al., 1987; Holz et al., 1988) and 3H~ norepinephrine release from undifferentiated PC12 cells (Kongsamut and Miller, 1986). In rat sympathetic neurons (Hirning et al., 1988), peripheral neurons (Perney et al., 1986), NGF-differentiated PC12 cells (Kongsamut and Miller, 1986), various areas of the central nervous system (Dooley et al., 1987a) and presumably rat motor nerve terminals (Atchison, 1989), N-type Ca2+ channels are implicated as the predominant type involved in neurotransmitter release. Given the ability of MeHg to block channel types which are associated with neurotransmitter release, block of Ca2+ channels at the nerve terminal may be an important component of its action to disrupt synaptic transmission. CHAPTER SIX EFFECTS OF METHYLMERCURY ON Na+ CHANNELS IN NEUROBLASTOMA CELLS 161 Eff NC cc] 162 ABSTRACT Effects of MeHg on whole-cell Na+ currents am) in neuroblastoma NIB-115 and NG~108 cells were examined using patch-voltage clamp techniques. In neuroblastoma cells, 10 EM MeHg reduced INa by approximately 50% after a 2 min exposure. MeHg did not alter the activation time Of Na+ current, nor the time course of whole-cell current inactivation. Na+ currents could still be recorded after 8 min of exposure to MeHg. Effects of MeHg on NG~108 cells did not differ from those of MeHg on N1E~ 115 cells. These results suggest that MeHg interacts readily with Na+ channels in neuroblastoma cells. patent dbserl Howe stimu? in the (Hun 1976] redth sulfll Na+ Na’ new 163 INTRODUCTION Results of Traxinger and Atchison (1987b) indicate that MeHg disrupts action potential conduction as well as Ca2+ channel function at the N MI. The focus of this dissertation has been directed towards possible effects of MeHg on Ca2+ channels. However, propagation of the action potential in mammalian motor nerves, the stimulus for Ca2+ channels to open, depends on the function of Na+ and K+ channels in the plasma membrane. MeHg blocks the action potential in squid giant axon (Huneeus-Cox et al., 1966) and increases its threshold for excitation (Shrivistav et al., 1976). Na” currents in neuroblastoma cells are decreased by MeHg while only small reductions in K“ current are observed (Quandt et al., 1982). Furthermore, the sulfhydryl agents p-chloromercuriphenylsulfonic acid and inorganic mercury reduced Na" current in crayfish axons (Shrager, 1977). Thus, in addition to Ca2+ channels, Na’r channels are also a potential target for MeHg action in mammalian motor nerves. Alterations in Na+ channel function by MeHg may therefore indirectly affect Ca2+ channel function by altering patterns of cellular excitability. In previous investigations of the effects of MeHg on action potential propagation and Na+ channel function, concentrations of MeHg between 25 and 100 uM were required to alter Na“ channel function (Shrivistav et al., 1976; Quandt et al., 1982), whereas concentrations between 5 and 20 u M were required to block Ba2+ current in PC12 cells (Shafer and Atchison, 1991). However, studies of effects of MeHg on N a“ channel function have employed primarily intracellular voltage-clamp techniques, while effects of MeHg on Ca2+ channel function has been examined using 164 whole-cell patch voltage-clamp techniques. Thus, it is difficult to compare these results directly due to the differences in techniques employed to study channel function. Therefore, experiments were designed to examine effects of MeHg on Na+ currents measured using the whole-cell configuration of the patch voltage-clamp technique. By doing so, effects of MeHg on Na+ channels will be described under conditions very similar to those for which effects of MeHg on Ca2+ channels have been well characterized. This will provide additional information concerning the ability of MeHg to affect a second type Of voltage-gated channel in neuronal cells. eh'ci 165 MATERIALS AND METHODS Na+ current recordings. Voltage-activated Na“ currents were recorded fi'om undifferentiated mouse neuroblastoma NIB-1 15 cells (provided by Dr. P. Cobbett of Michigan State University) or CAMP-differentiated (7 days) NG~108 cells using the whole cell configuration of the patch voltage-clamp technique (Hamill et al., 1981). These cell lines were used to examine effects of MeHg on Na+ current rather than PC12 cells because consistent expression of Na+ current could not be achieved from available stocks of PC12 cells. External solution contained (mM): HEPES (20), MgC12(1.0), d-glucose (10), NaCl (115), and TEACl (20) pH 7.3. Internal or pipette solution contained (mM): HEPES (10), MgCl2 ( 1), d-glucose (10), TEACl (10), EGTA (5), CsCl (125) and ATP (2), pH 7.3. To record Na+ currents, resistance of patch electrodes (1.2 mm glass) was between 2-6 megohms and the patch clamp circuit consisted of an Axon Instruments CV-l headstage and Axopatch 1B patch clamp. Pulse protocols were generated and current responses recorded on-line using a Compaq 386 personal computer and the pClampTM program interfaced to the Axopatch 1B via an Axon Instruments TL-1 interface board. Current responses were filtered at 10 kHz. Linear leak and capacitive current components were subtracted using a computer-generated protocol. This protocol functions by using a series of small amplitude subpulses to compute leak and capacitive current expected during a larger test pulse. Ideally, the subpulses elicit no voltage-gated current. Hence the leak and capacitive current measured during the subpulses can be averaged and scaled to that expected from the test pulse. 166 The magnitude of the subpulses is determined by dividing the magnitude of the test pulse by the number of subpulses. For these experiments, a series of 10 subpulses was used to measure leak. MeHg was added to 1 ml of bath solution in 500 ill aliquots to yield the final concentration indicated in the text. 167 RESULTS Effects of MeHg on Na+ current in neuroblastoma cells. Addition of MeHg to the extracellular solution rapidly reduced Na+ current in neuroblastoma cells. A 2 min resting exposure of neuroblastoma cells to 10 pM MeHg reduced peak inward Na” current by 20 ~ 50% at all test potentials examined (Figure 6.1A). However, in contrast to effects of MeHg on whole-cell Ba2+ current (Shafer and Atchison, 1991), MeHg did not alter the inactivation kinetics of whole-cell Na+ current. This is clearly apparent if current recordings obtained from the same cell in the absence and presence of MeHg are compared after scaling the MeHg-treated current response to the same amplitude as the control response. (Figure 6.1, inset). MeHg did not affect t , the rate constant for inactivation, at test potentials between ~40 and +40 mV (Figure 6.1B). Also, MeHg did not alter significantly the time to peak amplitude of Na+ current in neuroblastoma cells (Figure 6.2). This is consistent with effects of other sulfhydryl reagents on Na” channels (Shrager, 1977). Na+ current could be recorded from neuroblastoma cells after 8 min of exposure to 10 EM MeHg (not shown). This concentration of MeHg completely blocks In. in PC12 cells after 3-5 min exposure. Effects of MeHg on Na+ currents in NG~108 cells were not different from those on N1E~115 cells (results not shown). Blocking actions of MeHg may be altered by administration of hyperpolarizing prepulses (results not shown). 168 A TEST POTENTIAL (mV) -60 -4O -20 O 20 4O I,“ N1E-115 cells A a E ‘~..$\ I I T ‘$ v . “‘0‘ ,I. 0 -|» ’,-O""?'” [- “~. I ,-.-o" /' Z I o--..O----<> ‘ g o ' 3 0 O I E I I .. I/. 8 .\’\,,/' h.p.= - 9 O - 5 5 0 B '8 e 9 - 9 Z ‘\ 9 2 6 D I “X. L— O 2 NR '5 3 m“ < .m- . .2. Kg 3 m < I- -40 -2O 0 20 40 TEST POTENTIAL (mV) Figure 6.1. Effects or MeHg on Na“ channels in N1E-115 cells. A. Peak inward Na+ current-voltage relationship in N 1E~115 cells recorded prior to (filled circles) and following a 2 min exposure to 10 u. M MeHg (open circles). B. Time constant (tau) for inactivation of whole-cell Na+ current at test potentials used in A measured prior to (filled circles) and following a 2 min exposure to 10 p. M MeHg (open circles). All current responses were recorded using 10 ms test potentials from a holding potential of ~90 mV. ReSponses were filtered at 10 kHz and values are the mean: SEM of six cells, each of which was treated with MeHg. Inset: Top pair of traces are current responses to a test potential of ~10 mV prior to (filled circles) and following a 2 min exposure to 10 u M MeHg (open circles). The lower trace is the MeHg-treated response scaled to the same magnitude as the control response. Note that the kinetics of decay do not appear to be altered by Mel-lg. Scale bars for the tOp pair of traces are 200 pA and 1 ms. The current responses shown are the average of ten responses. 169 2.0 ‘ -50 -20 1O 40 TIME To PEAK I", (ms) TEST POT. (mV) Figure 6.2. Efl’ects of MeHg on time to peak Na+ current. The time required for inward Na+ current to reach peak amplitude following a depolarizing stimulus. The values are the mean: SEM of 5 (solid circles, control) or 3 cells treated with 10 u M MeHg. 170 DISCUSSION Results of these experiments indicate that low micromolar concentrations of MeHg disrupts N a+ channel flmction in clonal cells. Several previous studies indicated that MeHg disrupts action potential propagation and Na+ channel function (Shrivistav et al., 1976; Quandt et al., 1982), but effects of MeHg on Na+ channel function had not been examined using patch-voltage clamp techniques. Shrivistav et al., (1976) reported that concentrations of 25-100 llM MeHg decreased N 3" current by 50% in squid axons after 15 min of exposure in the bath solution. Quandt et al., (1982) reported that a 3 min exposure to 20 uM MeHg reduced Na+ current by approximately 10% in N1E~115 cells. Higher concentrations of MeHg (60 nM) produced greater reductions (80 ~ 100%) in Na+ current (Quandt et al., 1982). Results of present experiments indicate that 10 RM MeHg rapidly and effectively blocks Na+ current in neuroblastoma cells. It is possible that the lower temperatures (10°C) employed by Quandt et al. (1982) may account for some of the differences in the effects of MeHg on neuroblastoma cells. Changes in temperature may alter channel gating (Hille, 1984) or may influence fluidity of the cell membrane; should target sites important for MeHg-disruption of Na‘ channel function be located within the membrane, changes in fluidity may alter the ability of MeHg to reach these sites. In addition to Na+ current, MeHg also blocks Ca2+ current in PC12 cells at concentrations between 1~20 uM (Shafer and Atchison, 1991). This is unique when compared to inorganic Ca2+ channel antagonists, in that MeHg affects two different channel types with equal avidity. Although the present data (including effects of 171 MeHg on Ca2+ channels) do not suggest any particular mechanism responsible for this ability of MeHg, at least three mechanisms could account for this effect. First, like Na+, MeHg is monovalent. However, the Na” channel is not permeanated to positively charged methylated amines (Hille, 1984), and therefore may not be permeanated by MeHg. In addition, mercury is a transition metal; this class of metals includes Hg“, Cd“, Ni“, Pb2+ and Zn2+ (Nachshen, 1984), which block Ca2+ channels by interactions with binding sites within the channel pore (Lansman et al., 1986; Hess et al., 1986). Thus, MeHg may enter the pore of both Na+ and Ca2+ channels, resulting in block. Second, other Ca2+ channel antagonists, including the DHPS, have also been shown to block N a+ channels in cardiac myocytes (Yatani and Brown, 1985), although the concentrations of DHP required to block Na+ channels were higher than those which block Ca2+ channels. PD 122860, is a DHP compound which stimuluates Na‘ channel function and blocks Ca2+ channel function (Haleen et al., 1989). In addition, veratridine and grayanotoxin block Ca2+ and Na“ channels in the same concentration range in N1E~115 cells (Romey and Lazdunski, 1982). This has led to the suggestion that similarities in the molecular structures of Na+ and Ca2+ channels may underlie these interactions (Hosey and Lazdunski, 1988). There is a high degree of sequence homology between the a subunit of the voltage- dependent Na+ channel and the a 1 subunit of the DHP-sensitive Ca2+ channel from skeletal muscle. The greatest area of homology exists within four repeated domains, each containing six transmembrane segments, one of which presumed to be responsible for voltage-sensitivity (Campbell et al., 1988). Perhaps MeHg, which is 172 lipophilic, interacts with these regions in both Na” and Ca2+ channels. The possible effect of MeHg on steady-state inactivation of Na+ channels is interesting in this regard. Recently, four putative Ca2+ channel cDNA molecules which are homologous to the a 1 subunit of the DHP-sensitive Ca2+ channel from heart and skeletal muscle have been isolated from rat brain (Snutch et al., 1990). Finally, ethanol has also been reported to block both Na+ (Harris and Bruno, 1985) and Ca2+ channels (Greenberg et al., 1987a; Marks et al., 1989) in different systems. Thus, disruption of lipid environment may also reduce channel function. MeHg alters both Na+ and Ca2+ channel function, however, there appear to be differences in the characteristics of block of Na+ and Ca2+ channels by MeHg. The most striking difference is the lack of effect of MeHg on inactivation kinetics of whole-cell Na” current. In PC12 cells, 10 11M MeHg consistently eliminated the inactivating component of Ba2+ current (Shafer and Atchison, 1991). Second, the time courses may be different; block of Na+ channels by MeHg was not complete after 8 min, whereas block of Ca2+ channels by 1011M MeHg was complete after 3-5 min. Although this difference was not statistically significant, the phenomenon was consistent. In conclusion, MeHg interacts readily with Na+ channels in neuroblastoma cells at concentrations which also block Ca2+ channels under similar conditions. The ability of MeHg to interact with Na+ channels at low micromolar concentrations is consistent with previous reports (Huneeus-Cox et al., 1966; Shrivistav et al., 1976; Quandt et al., 1982) as well as effects of MeHg at the neuromuscular junction 173 (Traxinger and Atchison, 1987b). The ability of MeHg to interact with more than one channel type at similar concentrations is unique when compared to inorganic Ca2+ channel antagonists. Furthermore, in contrast to effects on whole-cell Ba2+ current, MeHg reduced whole cell Na+ current without producing apparent alterations of N a” channel inactivation kinetics. Thus, in addition to Ca2+ channels, Na“ channels may also be an important site of MeHg action in neuronal membranes. CHAPTER SEVEN EFFECTS OF METHYLMERCURY ON Na'r AND Ca2+ CURRENTS IN INTACT NEUROMUSCULAR JUNCTIONS 174 175 ABSTRACT The ability of MeHg to block presynaptic Ca2+ and N a+ channels in intact neuromuscular junctions was examined in mouse triagularis stemi motor nerves. Potential changes arising from Na+ and Ca2+ channel function could be recorded from the perineurial sheath surrounding motor neurons when K+ channels were blocked by TEA and 3, 4~diaminopyridine. Two components of the Ca2+~dependent potential were observed; a fast component of approximately 60 ms duration and a long-lasting component of approximately 100 ~ 800 ms duration. MeHg (100 nM) blocked both Na+ and Ca2+ components rapidly. In 2 of 5 preparations exposed to 50 uM MeHg, the Ca2+ channel-mediated component was blocked prior to block of the Na+ channel-mediated component. In the remaining three preparations, both Na*~ and Ca2+~dependent potentials were blocked at similar times. Following block by MeHg, perfusing the preparation in MeHg-free solutions did not result in recovery of Na+ or Ca2+ channel activity; nor did increases in intensity and/or duration of stimulus to the intercostal nerves. In the presence of K" channel blockers, repetitive firing of nerves in response to a single stimulus was observed in 20 ~ 30% of the triangulan's preparations; in two preparations treated with MeHg, repetitive firing decreased prior to block of the stimulus-induced N a+ / Ca2+ potentials. These results suggest that MeHg blocks Na+ and Ca2+ channels in intact neuromuscular preparations. 176 INTRODUCTION The results of experiments presented in previous chapters provide strong evidence that MeHg disrupts function of Ca2+ channels in nerve terminals and blocks Ca” channel types which are thought to be involved in neurotransmitter release. Effects of MeHg on synaptosomal 4S’Caz“ influx and on Ba2+ currents through somatic N- and L-type Ca2+ channels in PC12 cells are comparable in many respects; MeHg alters ionic selectivity and inactivation, blocks Ca2+ channels in a voltage- but not state-dependent manner, and still exerts blocking effects in spite of increasing [Ca2+], These results suggest either that the same types of Ca2+ channels are present in each preparation, or that MeHg interacts with a variety of Ca2+ types in a similar fashion. Regardless, these results support the hypothesis that disruption of nerve-evoked neurotransmitter release by MeHg is in part the result of block of Ca2+ channels at the presynaptic terminal by MeHg. However, experiments performed previously have not measured effects of MeHg on Ca2+ channel function in intact motor nerve terminals. Thus, one cannot be certain that MeHg is indeed blocking Ca2+ channels in intact motor nerve terminals. Furthermore, MeHg also blocks Na+ channels in Cllltured cells (Chapter 6, Quandt et al., 1982) and may alter Na” channel flmction in intact nerves (Traxinger and Atchison, 1987b). Therefore, experiments were designed to examine effects of MeHg on Na+ and Ca2+ channel function in intact m0tor nerves by measuring effects of MeHg on Na*~ and Can-dependent potentials in the perineurial sheath of mouse motor nerves. 177 The perineurium is a layer of flattened cells which surrounds mammalian nerve trunks, forming a relatively impermeant barrier to small molecules and ions. Because of this property, the perineurium functions as an electrical insulator around the nerve trunk (Figure 7.1). Axial currents which move from the most distal nodes of Ranvier into the nerve terminal form local circuits in the perineurium by returning to their node of origin via the perineurial space (Gundersen et al., 1982; Mallart, 1985). By placing a microelectrode inside the perineurial sheath near motor nerve terminal regions, voltage changes resulting from inward Na+ currents in preterminal regions and inward Ca2+ currents in nerve terminals can be recorded simultaneously (Mallart, 1985; Penner and Dreyer, 1986). Furthermore, since the perineurium contains multiple nerve fibers, the signal to noise ratio is increased over "loose patch" recordings (Brigant and Mallart, 1982; Mallart and Brigant, 1982) of Na" and Ca2+ currents from individual terminals (Penner and Dreyer, 1986). Thus, effects of MeHg on Na+ and Ca2+ channels were examined simultaneously by using perineurial recording techniques in intact motor nerve terminals of mouse triangular-lls stemz' muscle. 178 -5-9 """"""" \ -------- :1:8-------‘-___-p a_’_'_.--i/ ______ jif‘ _ _______ T3 3.5:, no; / n L / n L Figure 7 .1. Local circuits in the perineurial sheath of motor neurons. Top, simplified diagram of a motor nerve terminal, with the terminal on the left. The current pathways are indicated by the dashed lines. hn, heminodal region; n, node of Ranvier; p, perineurium; my, myelin sheath; a, axon. Bottom, simplified equivalent electrical circuit. A corresponds to the external medium, B indicates the perineurial space. Rm, membrane transverse resistance of the terminal; Rm, membrane transverse resistance at the nodes; R and R. are the longitudinal resistances of the perineurial space and the axon, respectively. The arrows indicate direction of current flow. Modified from: Mallart, J. Physiol. 368565-575 (1985). 179 MATERIALS AND METHODS Na“ and Ca2+ voltage recordings from the perineurial sheath. Potential changes arising from Na+ and Ca2+ channel function were recorded from the perineurial sheath of motor neurons in mouse triangulanlr stemi muscle (McArdle et al., 1981). The muscle and intercostal nerves were dissected from the ribcage of adult male Swiss-Webster mice (30-50 g). Following cervical dislocation, the ribcage was separated from the spinal column and bathed in oxygenated physiological solution consisting of (mM): NaCl(135); KCl (5); CaCl2 (2); MgCl2 (1); HEPES (14) and d- glucose (11), pH 7.4. This solution was replaced completely every 3-5 min while the overlying intercostal muscles and ribs were removed. Upon completion of the dissection (Figure 7.2), the muscle was immediately transferred to a perfusion chamber and perfused at a rate of 1~2 m1 / min with oxygenated physiological solution. To measure Na+~ and Ca2+ (Bard-dependent potentials in the perineurial sheath, the muscle was perfused with solution containing (mM): NaCl (118); TEACl (20); BaCl2 (5); MgCl2 (1); HEPES (14); d-glucose (11); 3,4-diaminopyridine (0.5) and d-tubocurarine chloride (0.05), pH 7.4. The solution was saturated with 95 /5% Oz/COZ. TEACl and 3,4-diaminopyridine were added to suppress K” channel activity and d-tubocurarine was added to block post-synaptic ACh receptors, thereby eliminating muscle twitch. The intercostal nerves were stimulated at 0.1 Hz by a suction electrode; typical threshold values were 15-30 mV and 15-30 its duration. Potential changes in the perineurial sheath were measured by glass microelectrodes (1.2 mm O.D., 10 MD) filled with physiological solution. Signals were recorded using 180 a CV~1 headstage and an Axopatch 1D patch clamp amplifier in the current clamp mode. Voltage responses were stored on FM tape (Vetter Instruments) for later analysis using a chart recorder (Gould Instruments) and/ or by digitization at a rate of 1 point every 20 us and subsequent storage on a Compaq personal computer. 181 0 'l‘l'i'-,:H-‘/ O I' {In . W/ e , Figure 7.2. Diagram of mouse niangdan': and muscle. The muscle, located on the inner surface of the ribcage inserts into the sternum and the intercalations of the ribs. The overlying ribs and intercostal muscles have been removed. Branches of the 2nd, 3'“, and 4‘“ intercostal nerves, which innervate the triangulalis, are visible. 182 RESULTS Effects of MeHg on Na" and Ca2+(Ba2*)~dependent voltage changes in the perineurial sheath. Placing a glass microelectrode inside the perineurial sheath of nerves in the triangularis stemi muscle resulted in a negative potential of ~2 to -5 mV, consistent with previous reports (Mallart, 1985 ; Penner and Dreyer, 1986; Anderson and Harvey, 1988). When perfusing the muscle with normal saline solution, stimulation of the nerve was followed closely (approximately 0.8 ms) by a negative deflection in voltage (Figure 7.3A); a small fast component (arrow) slightly preceded a larger component of longer duration. Previously, it has been demonstrated that the first component arises from Na+ currents in the nodal and heminodal portions of the nerve, whereas the second component arises from outward K+ current in the nerve terminal (Brigant and Mallart, 1982; Mallart, 1985; Penner and Dreyer, 1986; Anderson and Harvey, 1988). When the solution perfusing the muscle contained 20 mM TEA and 500 uM 3,4-diaminopyridine, the potential consisted of a fast negative deflection followed by a longer-lasting positive deflection (Figure 7.3B): demonstrated to arise from nodal Na+ and terminal Ca2+ currents, respectively (Brigant and Mallart, 1982; Mallart, 1985; Penner and Dreyer, 1986; Anderson and Harvey, 1988). Addition of TTX (5 11M) to the perfusing solution blocked the Na+ component of the potential (results not shown). Two components of positive potential were observed: a fast transient of approximately 60 ms in duration (Figure 7.3C), and a long-lasting (100 ~ 800 nls) potential of varying duration (Figure 7.3D,E). These results are consistent with those 183 Figure 73. Voltage responses recorded from perineurial sheaths. A) Na+ /I(+ voltage potential recorded from perineurial sheaths near nerve terminal regions. The arrow indicates the shoulder arising from Na+ current in the nodal and heminodal regions. B) Na" /Ca2+ voltage potential recorded in the presence of TEA and 3,4-diaminopyridine in 5 mM Ba2+ solutions. C-E) Na" Ca2+ voltage potentials from the same preparation demonstrating rapid (C) and long-lasting (D, E) Ca + responses. Responses are not averaged or summed and were recorded in response to intercostal nerve stimulation at a rate of 0.1 Hz. Scale bars are 035 mV and 2 ms (A), 0.5 mV and 4 ms (B), 0.5 mV and 200 ms (C-E). 184 of Penner and Dreyer (1986) in which two types of Ca2+ potentials were demonstrated in mouse motor nerve terminals; presumably due to the presence of two different Ca2+ channel subtypes. Microscopic resolution was limited (15OX) to an extent which prevented placement of electrodes within the finest branches of the perineurial sheath, which contain 3-4 nerves (Penner and Dreyer, 1986). Under these conditions the number of nerves contributing to potentials cannot be determined. Therefore, quantitative assessment of effects of MeHg on potential amplitude provide little definitive information concerning effects of MeHg on individual nerves or terminals. Amplitude scale bars have been included in some figures to provide information concerning relative amplitudes. The amplitude of the Ca2+ channel- dependent potential typically was between 0.5 and 1.5 mV in the absence of MeHg, depending on the preparation. The positive portion of the potential was dependent on the presence of divalent cations in the perfusing solution. Preparations which were dissected in 2 mM Ca2+ and then perfused with nominally Ca2+~ and Ban-free solution exhibited small and short duration Ca2+ potentials, presumably due to residual Ca“ in the muscle tissue (Figure 7.4A,B, left trace). Addition of 5 mM Ba2+ to the perfusing solution increased both the magnitude and duration of the Ca2+ potential (Figure 7.4A,B, right trace). Furthermore, "puffing" 1 mM cobalt into the perfusion chamber decreased the amplitude and duration of only the positive potential in a reversible manner (Figure 7.4B). These results are consistent with results of Mallart (1985) and Penner and Dreyer (1986) in which these potentials were demonstrated to arise from 185 Ca2+ channel function in nerve terminals. Addition of MeHg (100 nM) to the perfusing solution caused a reduction in amplitude and eventual block of both the Na” and fast Ca2+ potentials after a delay of 3-4 min (Figure 7.5). Both Na” and fast Ca2+ potentials were decreased simultaneously in three preparations. In contrast to the results of Traxinger and Atchison (1987b), washing with MeHg-free solution at the time of block did not reverse effects of MeHg; even after 15-20 min. In addition, increasing stimulus duration (up to 5 times threshold) and/or intensity (up to 10 times threshold) at the time of block and/or during wash also were ineffective in reversing block of either the Na+ or the Ca2+ channel-mediated response by MeHg. 50 11M MeHg also rapidly decreased both Na+ and fast Ca2+ potentials in 4 out of 5 preparations. However, in 2 preparations, the fast Ca2+ component appeared to be affected prior to the Na+ component. In three preparations (one had been exposed to 100 p M MeHg), the long-lasting component of the positive potential was decreased prior to block of Na” and fast Ca2+ potentials. These results suggest a somewhat earlier or more specific effect of MeHg on nerve terminal Ca2+ channels. Thus, MeHg is capable of blocking Ca2+ currents in intact motor nerve terminals. In addition, effects of MeHg on specific components of perineurial waveforms indicate that the decreases in signal amplitude are not due to disruption of the insulating properties of the sheath by MeHg. As with 100 uM MeHg, these effects were not reversible by washing the preparation in MeHg-free solutions or by increasing stimulus intensity and/ or duration. Thus, block by MeHg of Na+ channels, and possibly of Ca“ channels, appears to be irreversible under these conditions. 186 Figure 7.4. Ca2+ channel-mediated component of vol potentials in the perineurial sheath. When perfusing with solutions containing no added Ca2+ or Ba +, the Ca2+ channel-mediated component of the potential is small and of short duration (A and B, left traces). When 5 mM Ba“ is added to the perfusing solution (A), or "puffed" into the perfusion chamber (B), the Ca2+ channel-mediated component of the potential increases in amplitude as well as duration ( ' t traces). Scale bars are 9 msand025mVinAand0.125mVinB. C) WhileperfusinginSmMBa + solutionilrnMCo2+ was "puffed“ over the muscle. The voltage traces were recorded prior to addition of Co I (CON), in the presence of Co2+ (Co), and during washout of C02” (WASH). Scale bars are 025 mV and 20 ms. 187 Figure 7.5. Efl'ects or MeHg on NaVCa“ potentials in the perineurial sheath. Nat /Ca2+ potentials recorded prior to (A) and 0.16 (B), 1 (C), 2 (D), 3 (E) 4 (F) and 5 (G) min after addition of 100 pM MeHg. Scale bars are 0.5 mV and 40 ms, except for B; for which the temporal bar is 80 ms. Note the loss of the long-lasting component of the Ca2+ potential after 1 min. 188 When using K+ channel blocking agents to unmask Ca2+ channel activity in the nerve terminal, repetitive firing of neurons often occurs due to lack of repolarization of the neuronal membrane by outward K+ currents. Other investigators have used small concentrations of local anesthetics fo prevent repetitive firing (Brigant and Mallart, 1982; Mallart, 1985 ; Penner and Dreyer, 1986; Anderson and Harvey, 1987; 1988). Since MeHg may affect axonal Na‘ channels and decrease excitability (Shrivistav et al., 1976; Quandt et al., 1982; Traxinger and Atchison, 1987b), use of local anesthetics might mask or confound possible effects of MeHg on Na+ channel function in intact nerves. Thus, no pharmacological manipulations were undertaken in these experiments to suppress repetitive firing of motor neurons. Repetitive firing was observed in approximately 20 to 30% of triangularis stemi preparations (Figure 7.6A,B), only 2 of which were treated with MeHg. Addition of MeHg (50 nM) to these preparations resulted in a suppression of repetitive firing prior to block of stimulus-evoked N a” and fast Ca2+ potentials (Figure 7.6C-F). 189 L..l..'..l1 ‘- l... . i '. . I. '; . n, L1 I'T 5:1 .l 1L ”fiflqummn i ‘_ :_ l i I i i, l: IV I | I :‘ ll .. I . . .-.. '_.. . . . |.. . . : z -‘ " - . .. 1.. . . . .._. . .. . l . . ._....__i _: .. . _;_ . .. .. - .. . 1 _. .-..... .. . . .. , .. ._ . z i I ' , 3 . ’ ' ; I i r I 5 I (1 .;... ._.:. ...-...... ._ -.-. .1. . 1-121.: ' - i it“ I . D W WWWNTNMWNWWMTWWHJ I "~l _ I , I III i ——_I-- [T‘Iéfl— ..--‘_- .4. -... ' :rt::é'.:--.é-zl..*ij ...! I I t , j I I : I55. :I .0. .I -- , . , r 'r': ”T‘ "'“" Figure 7.6. Efl'ects of Mei-lg on repetitive firing of motor nerves. An example of repetitive firing in a motor nerve recorded prior to (A,B) and 20 (C), 70 (D), 80 (E), and 90 (F) s after the addition of 50 #M MeHg. Note also the loss of the long-lasting portion of the Ca2+ response in this example. Scale bar is 700 ms. 190 DISCUSSION Results of these experiments indicate that MeHg: 1) appears to block both N a” and Ca2+ currents in intact neuromuscular preparations, and 2) under the conditions used in this study, blocks presynaptic Na+ currents in a manner which is not reversible by increasing stimulus intensity or duration. At concentrations of 50 and 100 NM, MeHg blocked the inward Na+~ dependent component of the voltage potential. In addition, MeHg suppressed repetitive firing of motor nerves prior to significant reductions of stimulus-induced Na+ or Ca“ potentials. Repetitive firing of the nerve is a function of membrane excitability, and may occur under conditions of these experiments due to lack of repolarization of the nerve by outward K+ current before Na+ channels recover fiom inactivation. The ability of MeHg to suppress repetitive firing is consistent with the results of Traxinger and Atchison (1987b) which indicated that MeHg decreased neuronal excitability. Block of a portion of the population of N a+ channels by MeHg may prevent repetitive firing by decreasing the number of N 3+ channels available for opening immediately after an action potential propagates through a given region of the nerve. Alternatively, MeHg has been demonstrated to depolarize neuronal membranes (Shrivistav et “al., 1976; Hare and Atchison, 1990). Should this occur in intact axons as a result of MeHg exposure, one might expect a decrease in nerve excitability; depolarization might result in steady-state inactivation of Na‘ channels. Therefore, fewer channels would be available for activation. 191 In addition to blocking the Na” potential, MeHg also reduced the fast Ca2+ potential recorded from the perineurial sheath. Because of effects on Na+ potentials, decreases in Ca2+ potentials might simply have been due to failure of the action potential to invade the terminal and activate Ca2+ channels. However, in two preparations, the Ca2+ channel-mediated potentials were reduced prior to block of the Na+ channel-mediated potentials. In addition, the long-lasting component of the Ca2+ potential was blocked prior to block of the Na+ and fast Ca2+ potential in three preparations. Under these conditions, the action potential successfully invaded nerve terminals, since the fast Ca2+ potential was unaffected. These results indicate that in addition to N a” channels, Ca2+ channels in the presynaptic terminal are blocked by MeHg. Effects of MeHg on neurotransmitter release may be due to effects on both Na+ and Ca2+ channels in the presynaptic nerve; consistent with predictions of Traxinger and Atchison (1987b). The fast component of the Ca2+ potential is presumed to be associated with neurotransmitter release (Mallart, 1985; Penner and Dreyer, 1986), although the Ca2+ channel type(s) which contribute to this component have not yet been determined. This component is not affected by organic Ca2+ channel blockers (Penner and Dreyer, 1986) or to ~conotoxin GVIA (Anderson and Harvey, 1987), but is blocked by millimolar concentrations of divalent cations (Penner and Dreyer, 1986). Thus, the pharmacological characteristics of this component cannot be classified by the T, N, and L nomenclature system (N owycky et al., 1985a). The long- lasting component of the potential is not thought to be coupled directly to 192 neurotransmitter release (Penner and Dreyer, 1986). However, no evidence exists to preclude this mechanism from contributing to nerve-evoked release. This component is sensitive to verapamil and diltiazem but not DHPS (Penner and Dreyer, 1986). Thus, the Ca2+ channel types mediating this component are unknown. Although MeHg blocks both N- and L-type Ca2+ channels in PC12 cells, the ambiguity over the types of Ca2+ channels present in the motor nerve terminal makes it difficult to determine whether or not MeHg is acting on N- and L-type channels in this preparation. Traxinger and Atchison (1987b) reported that block of EPPS in rat hemidiaphram preparations could be reversed by washing in MeHg-free solutions and increasing the intensity and/or duration of the stimulus to the phrenic nerve. In addition, following reversal of initial MeHg-induced block, a subsequent block by MeHg was relieved by increasing [Ca2+]c. In the present experiments, the effects of MeHg on Na” and Ca2+ potentials were not relieved by washing in MeHg-free solutions and / or by. increasing stimulus intensity or duration. The disparity of results from this study with those of Traxinger and Atchison (1987b) may be due to the conditions employed in this study to unmask Ca2+ channel-mediated components of voltage changes in the perineurial sheath. Block of K+ channels by TEA and 3,4- diaminopyridine will result in depolarization of the axonal membrane; should MeHg further depolarize the membrane, decreases in excitability may result from Na+ channel inactivation. In addition, the high concentration of Ba“ used to measure Ca2+ channel function may also stabilize the axonal membrane (Frankenheuser and 193 Hodgkin, 1957; reviewed by Hille, 1984). Indeed, Traxinger and Atchison (1987b) reported that increasing [Ca2+]: along with addition of 4-aminopyridine prevented reversal of MeHg-induced block of EPPS when stimulus intensity was increased. Under the conditions of this study, therefore, it might be expected that MeHg- induced decreases in nerve excitability may be difficult to overcome. In conclusion, effects of MeHg on Na+ and Ca2+ channel function in intact motor nerves and terminals have been examined for the first time. Consistent with predictions from in vitro systems, the results indicate that MeHg disrupts function of both channel types in intact nerve terminals. The ability of MeHg to block Ca2+ channels in intact motor nerve terminals supports the hypothesis that effects of MeHg on Ca2+ channel function may contribute to block of neurotransmitter release by MeHg. CHAPTER EIGHT SUMMARY AND CONCLUSIONS 194 195 This dissertation has presented results of experiments designed to examine the following hypothesis: block of Ca2+ channels by MeHg may contribute to MeHg~ induced block of synaptic transmission at the motor nerve terminal. This hypothesis was based on the experimental observations that block of synaptic transmission at the neuromuscular junction (Traxinger and Atchison, 1987b), and of depolarization- induced 45Ca2” influx into synaptosomes by MeHg could, under certain conditions, be antagonized by extracellular calcium (Atchison et al., 1986). Specifically, experiments were designed to examine how MeHg disrupts Ca2+ channel function and to determine types of Ca2+ channels affected by MeHg: particular emphasis has been placed on Ca2+ channels located in nerve terminals and on Ca2+ channel types which are presumed to mediate neurotransmitter release from central and/ or peripheral terminals. The small size and close anatomical association of the motor nerve terminal with muscle fibers severely restrict the ability to measure directly Ca2+ channel function in intact NMJ preparations. Thus, out of necessity, some experiments were performed on isolated nerve terminals or cultured cell lines. Based on results of these experiments, one putative mechanism by which MeHg effects neurotransmitter release at the N MJ is now better understood. Preliminary experiments characterized the ability and nature of block by MeHg of various components of Ca“ influx into isolated nerve terminals. MeHg blocked voltage-sensitive Ca2+ channels in synaptosomes; the component of ”Ca“ influx which has been associated specifically with neurotransmitter release from synaptosomes (Blaustein, 1975; Floor, 1983; Leslie et al., 1985) was dramatically 196 reduced by MeHg (Atchison et al., 1986; Shafer and Atchison, 1989) Block of this component was only partially relieved by increasing [Ca2+]: (Shafer and Atchison, 1989). Thus, in isolated nerve terminals, MeHg disrupts, in a potent and efficacious manner, function of Ca2+ channels demonstrated to mediate Ca2+ influx responsible for neurotransmitter release. The ability of MeHg to block Ca2+ channels in a heterogeneous preparation may represent a common action of MeHg at all nerve terminals. Furthermore, elucidation of the mechanisms underlying MeHg neurotoxicity will require more information regarding effects of MeHg in the central nervous system. Inasmuch as alterations in Ca2+influx may contribute to acute effects of MeHg on neurotransmitter release at the N M], these data suggest that alterations of Ca2+ channel function should be considered in future studies of effects of MeHg in the central nervous system. Further characterization of MeHg effects on synaptosomal Ca2+ influx demonstrated that MeHg altered Ca2+ channel functions including ionic selectivity and inactivation. In addition, block of Ca2+ channels by MeHg was demonstrated to be voltage~ but not state-dependent. MeHg also altered binding of specific Ca2+ channel ligands to their receptor sites in synaptosomes and PC12 cells (Shafer et al., 1990). These effects of MeHg were unique when compared to effects of divalent inorganic cations such as Cd“, Co2+ and Pb2+ (Nachshen, 1984; Suszkiw et al., 1985). The ability of MeHg to alter binding of to ~conotoxin GVIA and the DHP nitrendipine suggested that MeHg may interact with receptor sites associated with N ~ and L-type Ca2+ channels, respectively. 197 The role of voltage-sensitive Ca2+ channels in mediating neurotransmitter release is crucial. To date, there are three well characterized Ca2+ channel types; N, L and T (Tsien, 1988). The N- and L-type channels have been proposed to regulate neurotransmitter release from a variety of central, peripheral and clonal nerve cells. The patch-voltage clamp recording technique was used to examine the ability of MeHg to block N- and L-type Ca2+ channels in the clonal cell line PC12; both channel types were blocked rapidly by low micromolar concentrations of MeHg (Shafer and Atchison, 1991). These experiments demonstrated two important aspects of block of Ca2+ channels by MeHg; that MeHg blocks Ca2+ channel types which are thought to mediate neurotransmitter release fiom central (Dooley et al., 1987a) and peripheral nerve terminals (Rane et al., 1987; Holz et al., 1988; Atchison, 1989), and that characteristics of block of somatic and synaptosomal Ca2+ channels by MeHg are similar in many respects. The ability of MeHg to affect Ca2+ channels in intact motor nerve terminals was confirmed by examining effects of MeHg on voltage changes in the perineurial sheath arising from Ca2+ channel function in mouse motor nerve terminals. MeHg blocked Ca2+ potentials as well as N a" potentials recorded from the perineurium. This result confirms directly that MeHg disrupts Ca2+ channel function in intact motor nerve terminals, consistent with predictions from experiments in synaptosomes and PC 12 cells. Thus, results of experiments within this dissertation clearly support the hypothesis: MeHg disrupts function of Ca2+ channel types which are associated with neurotransmitter release (Shafer and Atchison, 1991) and alters function of Ca2+ 198 channels located in the terminals of central (Atchison et al., 1986; Shafer and Atchison, 1989; Shafer et al., 1990) and motor neurons (chapter 7). While no direct evidence has been presented htat Mehg disrupts channel function by entering the channel pore, the facilitation of block of In, by MeHg by increasing the frequency of stimuluation as well as the voltage-dependence of block suggest that MeHg may interact within the channel pore. The ionic radii of ions which carry current through Ca2+ channels range from 1.06 A for Ca2+ to 1.43 A for Ba2+ while Hg2+ has an ionic radius of 0.93 A The hydration energies are 382, 316 and 441 Kcal/mol, respectively (Cooper et al., 1984). Divalent mercury (Hg2+) has an atomic raduis of 1.49 A (Miller, 1984). Given that ionic radius will decrease by addition of a positive charge and hydration energy will decrease, one might expect that the ~Hg+ portion of MeHg would be on the order of ionic size and hydration to allow it to enter the Ca2+ channel pore. The N a” channel is permeable to organic ions such as guanidium and aminoguanidium, which are much larger in ionic radii than N a“. Interestingly, methylated ions cannot pass through the Na+ channel (Hille, 1984). Thus, it is not inconceivable that MeHg may enter the pore of the Ca2+ channel, resulting in block of Ca2+ flux through the channel. These results also provide direction for future investigations of the neurotoxic effects of MeHg. This includes but is not limited to: studies of effects of MeHg on neurotransmitter release, further examination of effects on channel function, effects of MeHg on Ca2+ channel functions other than neurotransmitter release, and how 199 effects on Ca2+ channels may relate to neurotoxicity observed during chronic exposure to MeHg. Although effects of MeHg on neurotransmitter release have been demonstrated clearly at the neuromuscular junction, relatively little information exists concerning effects of MeHg on neurotransmitter release from the central nervous system (See: McKay et aL, 1986). Since effects of MeHg on Ca2+ channels in both central and peripheral terminals have been characterized, the ramifications of these effects on neurotransmitter release can be examined in more detail. Of particular importance is how alterations in Ca2+ channel function produced by MeHg impact MeHg-induced effects on neurotransmitter release. Release of neurotransmitters from the nerve terminal is thought to follow transient, localized elevations of [Ca2"]i to near micromolar or micromolar levels (see: Smith and Augustine, 1988). Effects of MeHg on the rate and magnitude of this increase would be expected to result from block of Ca2+ channels. However, MeHg may have additional effects on neurotransmitter release independent of its effects on Ca2+ channel function (Atchison et al. , 1986; Traxinger and Atchison, 1987a). It has also been suggested that in addition to Ca2+ entry into the nerve terminal, voltage changes (beyond those which evoke Ca2+ entry) may be required for neurotransmitter release (Parnas et al., 1986; 1989). Therefore alterations of membrane potential by MeHg (Hare and Atchison, 1990) may also contribute to block of the exocytotic process. Future studies of effects of MeHg on the release processes which occur after Ca2+ entry into the nerve terminal will provide additional information concerning the role of Ca2+ 200 channel block by MeHg, as well as possibly providing a link between effects of MeHg on Ca2+ channels and other neuronal systems. Also important with respect to neurotransmitter release is the ability of MeHg to block both N ~ and L~type Ca2+ channels. Present interpretation of pharmacological and electrophysiological studies of Ca2+ channels involved in transmitter release suggest that N-type channels mediate release of small clear vesicles containing classical transmitters whereas L~type channels mediate release of dense-core vesicles containing peptide neurotransmitters (Hirning et al., 1988; Smith and Augustine, 1988). Neuropeptides often are involved in modulation of neuronal activity, resulting in long term changes. Thus, the ability of MeHg to block L~type Ca2+ channels may be an effect that has important ramifications during chronic exposures to MeHg. Effects of MeHg on T~type Ca2+ channels have not been examined; this channel has been associated with dendritic action potentials and pacemaker activity in the CNS (Tsien, et al., 1988). The ability of MeHg to affect the N- and L~type Ca2+ channels, as well as the ability of MeHg to affect Na” channel function and neuronal excitability (Traxinger and Atchison, 1987b), suggest that possible effects of MeHg on T~type channels be considered in studies of MeHg effects in the CNS. In addition to the role of Ca2+ channels in neurotransmitter release, effects of MeHg on channel function suggest topics for further study. The actions of MeHg on Ca2+ channels are unique in many ways from the actions of heavy metals such as Cd“, Co“, Pb“, Hg2+ and Ni“; which also block Ca2+ channels (Nachshen, 1984). One subject of interest is how difierences in charge and lipophilicity affect the ability 201 of mercurials to alter channel function. Recent studies suggest that only positively charged organic and inorganic mercurials are able to block Ca2+ channels, and only lipophilic mercurials block Ca2+ channels in a voltage-dependent manner (Hewett and Atchison, 1990). MeHg also alters Ca2+ channel ionic selectivity and inactivation (Shafer et al., 1990; Shafer and Atchison, 1991), which suggests that interactions between MeHg and Ca2+ channels differ from those of inorganic Ca2+ channel blockers. Further characterization of effects of MeHg on Ca2+ channel function at the single channel and biochemical level will provide additional information regarding the mechanisms by which MeHg produces these effects. In addition to effects on Ca2+ channels, this dissertation has presented evidence that MeHg produces effects on Na” channels in clonal cell lines and in motor nerve axons at similar concentrations to effects on Ca2+ channels. These results are consistent with previous studies of MeHg action (Quandt et al., 1982; Traxinger and Atchison, 1987b). Alterations of N a+ channel function by MeHg would certainly affect synaptic transmission in intact NMJ preparations (T raxinger and Atchison, 1987b; chapter 8), as block of impulse conduction by MeHg would prevent neurotransmitter release. However, due to the importance of Na“ channels in neuronal excitability, these effects of MeHg may also have implications for other aspects of MeHg toxicity. Thus, future experiments on ionic channel function may include examinations of common mechanisms of interaction of MeHg with different channel types as well as how effects on different ionic channels in the membrane may contribute to disruption of cellular function by MeHg. 202 MeHg alters a number of cellular processes including Ca2+ channel function. It has been established that MeHg also alters mitochondrial function (Levesque and Atchison, 1987; 1988; 1991) as well as Na+ channel function (Quandt et al., 1982; chapters 7 and 8) and nerve terminal membrane potential (Hare and Atchison, 1990). The reactivity of MeHg with sulfhydryl groups as well as amino and carboxyl groups makes it unlikely that block of synaptic transmission, or other neurotoxic effects of MeHg, are the result of disruption of a single process. Thus, it is important to characterize all of the effects of MeHg and try to synthesize sequences of events which contribute to toxicity. In addition to effects on neurotransmitter release, MeHg-induced block of Ca2+ channels in neuronal cells may have other effects on cellular function. For example, Ca2+ channels provide for transient, localized elevations in [Ca2+]i involved in grth cone elongation (Streit and Lux, 1987; 1989; 1990) or activation of Can-dependent enzymes (Kennedy, 1989). Regardless, Ca2+ channels are one cellular component involved in regulation of [Ca2+]i; other components include mitochondria, endoplasmic reticulum and other proteins and organelles (Meldolesi et al., 1988). Thus, effects of MeHg on Ca2+ channels (Atchison et al., 1986; Shafer and Atchison, 1989; 1991; Shafer et al., 1990), coupled with efiects on mitochondrial function (Levesque and Atchison, 1987; 1988; 1990) and membrane depolarization (Hare and Atchison, 1990) suggest the possibility that alterations in Ca2+ regulation may occur in response to MeHg exposure. Disruption of Ca2+ regulation might produce a cascade of events leading ultimately to cell death (Choi, 1988). 203 In conclusion, interactions of MeHg with voltage-dependent Ca2+ channels have been examined by the experiments contained in this thesis. Disruption of Ca“ channel function by MeHg has been observed in isolated and intact nerve terminals. In addition, MeHg alterats specific Ca2+ channel functions and disrupts funtion of two Ca2+ channel types which are presently associated with neurotransmitter release. Although further study of the effects of MeHg on neurotransmitter release are needed, evidence has been presented which supports the hypothesis that block of neurotransmitter release at the NMJ is, in part, the result of block of voltage- dependent Ca2+ channels which mediate Ca2+ entry into the nerve terminal. Furthermore, results of these experiments provide additional knowledge concerning interaction of MeHg with these critical membrane-associated molecules; which may contribute to our understanding of the mechanisms of neurotoxicity of this important environmental compound. APPENDIX APPENDIX 204 APPENDIX TRANSMITTER, ION CHANNEL AND RECEPTOR PROPERTIES OF PHEOCHROMOCYTOMA (PC12) CELLS: A MODEL FOR NEUROTOXICOLOGICAL STUDIES A) INTRODUCTION The use of isolated nerve terminals (synaptosomes) and intact nerve-muscle preparations to measure effects of MeHg on nerve terminal Ca2+ channels represent model systems in which have been extensively used to study neuronal Ca2+ channel channel function and therefore have been well characterized with respect to their strengths and limitations. By contrast, pheochromocytoma (PC12) cells are not of neuronal origin, and measurement of Ca2+ channel function in this cell line using patch voltage-clamp techniques assesses function of somatic or growth cone Ca2+ channels. However, because of the unique properties of PC12 cells, data collected under defined conditions may be applicable to neurons. PC12 cells have been a popular tool for neurobiological studies and represent a potentially useful tool for neurotoxicological studies as well. This chapter will focus on use of PC12 cells as model systems for studies of channel function as well as neurotransmitter release and receptor function. PC12 is a clonal cell line derived from a catecholamine-secreting tumor (pheochromocytoma) of rat adrenal medulla (Greene and Tischler, 1976). When cultured under normal conditions, these cells resemble adrenal chromaffin cells in 205 morphology as well as many aspects of physiology and biochemistry. However, when cultured in the presence of nerve growth factor (NGF), PC12 cells differentiate to resemble sympathetic neurons morphologically. The process of differentiation is accompanied by physiological and biochemical changes which ultimately result in neuron-like function in PC12 cells. In the absence or presence of NGF, PC12 cells release dopamine (DA), norepinephrine (NE) and/ or acetylcholine (ACh) in response to depolarization or activation of nicotinic or muscarinic ACh receptors. In addition, PC12 cells possess membrane—bound receptors, including receptors coupled to guanine nucleotide-binding proteins (G-proteins) and Na”, K+ and Ca2+ channels. As such, PC12 cells provide a useful model for studying: 1) processes associated with neuronal differentiation, 2) synthesis, storage and release of neurotransmitters, 3) flmction and regulation of ion channels and 4) interactions of compounds with membrane-bound receptors. The ability of PC12 cells to differentiate in response to NGF allows for selective expression of certain proteins and for comparisons of responses in undifferentiated, chromaffin-like and differentiated, neuron-like cells. B) PC12 CELL DIFFERENTIATION. Inclusion of NGF in the culture medium elicits a characteristic response from PC12 cells. Within 24 hours after its addition, the cells Stop dividing and begin forming neuronal processes (Greene and Tischler, 1976). After 4 to 14 days, the cell bodies have increased in size and grown long processes which form extensive networks. This process of morphological differentiation is accompanied by changes in the expression of a number of proteins including Na+ /K+ 206 ATPase (Inoue et al., 1988), voltage-sensitive N a+ channels (Dichter et al., 1977; Reed and England, 1986; Rudy et al., 1987) and increased choline acetyltransferase (CAT) activity resulting in increased ACh content (Greene and Rein, 1977b). Membrane electrical excitability and response of PC12 cells to ACh are increased by N GF-induced differentiation (Dichter et al., 1977). In addition, Changes in Ca2+ channel expression (Takahashi et al., 1985) decrease the susceptibility of neurotransmitter release to blockade by dihydropyridine-(DHP) type Ca2+ channel antagonists (Kongsamut and Miller, 1986). Thus, the differential expression and/ or function of proteins, including ion channels and receptors, can be induced by NGF- treatment of PC 12 cells. Other substances such as fibroblast growth factor (FGF) also induce differentiation of PC12 cells (for review see Fujita et al., 1989). However, the changes in enzyme and channel activity induced by FGF and other differentiating agents differ from those induced by NGF (Garber et al., 1989; Matsuoka et al., 1989; Damon et al., 1990; Pollock et al., 1990). Removal of NGF from the medium results in a rapid reversal of the process of differentiation (Greene and Tischler, 1976). The ability of PC12 cells to differentiate and de—differentiate in the presence or absence of NGF has made these cells a popular model for studying processes associated with N GF receptor activation and neuronal development and maturation. This use of PC12 cells has been reviewed elsewhere (Guroff, 1985; Fujita, et al., 1989). Clearly, there is great potential for use of this cell line in examining how neurotoxic compounds affect critical processes during neuronal development. 207 One other aspect of PC12 cell differentiation deserves mention here. Both adrenal chromaffin cells and sympathetic neurons are derived from embryonic neural crest cells. By contrast to embryonic neural crest cells (Doupe et aL, 1985; Anderson and Axel, 1986; Naranjo et al., 1986), Greene and Tischler (1976) reported that glucocorticoids such as dexamethasone (DEX) do not prevent the actions of N GF on PC12 cells. However, more recently, PC12 cells have been reported to respond to sodium butyrate (Byrd et al., 1987) by differentiating to an adrenal chromaffin-like phenotype. Upon butyrate treatment, PC12 cells undergo slight morphological changes without neurite extension and produce greater amounts of opioid peptide neurotransmitters (Byrd et al.,1987). Co-culture of PC12 cells with adrenal medullary endothelial cells induces PC12 cells to differentiate into an adrenal chromaffin-like phenotype which is unresponsive to N GF (Mizrachi, et al., 1990). Thus, PC12 cells respond to some stimuli by differentiating to an adrenal chromaffin-like phenotype, and other stimuli by differentiating to a neuron-like phenotype. C) 'I'RANSMI'ITERS, ION CHANNELS AND RECEPTORS IN PC12 CELLS. l) Neurotransmitter Release from PC 12 cells. As previously mentioned, PC12 cells synthesize and release DA, NE (Greene and Tischler, 1976; Greene and Rein, 1977a) and ACh (Greene and Rein, 1977b), but not epinephrine (Greene and Tischler, 1976). These three transmitters are all released in response to depolarization (Table 1). In addition, undifferentiated and NGF-differentiated PC12 cells release catecholamines in a Ca2+~dependent fashion 208 in response to nicotinic (Greene and Rein, 1977c) and muscarinic (Rabe et al., 1987) cholinergic agonists. The apparent decrease in activity of enzymes and amounts of neurotransmitters following treatment with NGF is most likely due to increased expression of new proteins induced by NGF, not to a decrease in the expression of enzymes associated with catecholamine synthesis (Greene and Tischler, 1976). The membranes of PC 12 cells contain a saturable NE transporter which is sensitive to cocaine and the tricyclic antidepressant desmethylimiprarnine (Greene and Rein, 1977a). NE and DA are most likely stored in vesicles, as exposure to reserpine depletes cellular catecholamine levels (Greene and Rein, 1977a). Thus, the mechanisms of synthesis, storage and re-uptake of catecholamines in PC 12 cells are similar to those of catecholaminergic neurons. In undifferentiated PC 12 cells, synthesis of NE, but not DA, changes with the cell cycle, being maximal during the G2 phase. Release of catecholamines by depolarization is maximal during G1, whereas carbamylcholine-induced release is maximal during the S and G2 phases. Cell cycle-specific changes in 45Ca“ influx induced by K+ depolarization and muscarinic receptor activation correlate with neurotransmitter release (Koike and Takashima, 1986). This characteristic of undifferentiated PC12 cells is clearly different from neurons, which do not divide. Permeabilization of PC 12 cells provides more direct access to release mechanisms. In staphylococcal a ~toxin permeabilized cells, Mg-ATP, calmodan or proteins containing ~SH or ~OH groups are not required for Ca2+~dependent release 209 (Ahnert-Hilger and Gratzl, 1987). Furthermore, release of catecholamines from a ~ toxin or digitonin-permeabilized cells is enhanced by phorbol esters (Pozzan et al., 1984; Ahnert-Hilger et al., 1987; Carroll et al., 1990) and reduced in a pertussis toxin~ sensitive fashion by an activator of guanine binding proteins, GTPy S (Ahnert-Hilger et al., 1987). These results indicate that protein kinase C (PKC) and G-proteins, respectively, may modulate catecholamine release from PC 12 cells. However, the role of G-proteins in modifying catecholamine release from PC12 cells is not yet clearly defined. GTPy S stimulated only Ca2+~independent NE release from digitonin- permeabilized PC12 cells, whereas an inhibitor of guanine binding proteins cyclase, GTPBS, reduced only the GTPy S component of release. Thus, the component of release affected by cyclic guanine nucleotides is not essential to NE release (Carroll et al., 1990). Furthermore, the opposite actions of GTPyS in a ~toxin and digitonin- permeabilized cells indicate that the agent used to permeabilize cells may influence G-protein-dependent pathways. DA release evoked by depolarization and / or nicotinic ACh receptor activation is accompanied by increases in cyclic AMP (CAMP). Moreover, forskolin, an activator of adenylate-cyclase, enhances depolarization-(Rabe et al., 1982; Baizer and Weiner, 1985) or carbachol-induced release (Baizer and Weiner, 1985). Thus, CAMP may also modulate catecholamine release in PC12 cells. However, the action of CAMP presumably is on a component of the release mechanism which functions after Ca2+ entry, as forskolin and/or its analogs either do not affect 4sCa2+ entry (Rabe et al., 210 1982), or block voltage-dependent Ca2+ channels in a CAMP-independent manner (Nishizawa et al., 1990) Effects of neurotoxic compounds on catecholamine synthesis and release have been studied in PC12 cells. Chronic exposure to ethanol (EtOH) decreased (muscarinic) ACh-induced release of [3H]-NE from PC12 cells without affecting NE uptake or ACh binding. However, EtOH did decrease the rise in [Ca2+]i induced by stimulation of muscarinic receptors on PC12 cells (Rabe and Weight, 1988). Concentrations of DA and NE in PC12 cells were decreased by a 6 h. treatment with 1 to 100 ppm Arochlor 1254 (Seegal et al., 1989). Thus, PC12 cells may provide a useful in vitro system for examining mechanisms of chemical action on catecholamine levels and/or release processes in neurons. In addition to catecholamine neurotransmitters, PC12 cells also synthesize and release ACh. This ability to synthesize and release two distinct classes of transmitters (ACh and catecholamines) permits PC12 cells to be used for differential studies of effects of neurotoxicants on the release of unique transmitters. In contrast to its effect on biosynthetic enzymes for catecholamines, NGF-differentiation increased CAT activity (Table 1). Treatment with retinoic acid instead of NGF resulted in an increase in CAT activity without morphological differentiation (Matsuoka et a1. , 1989). Undifferentiated PC12 cells have a high-affinity (Km = 12 u ), Na+~dependent choline transporter which is relatively insensitive to hemicholinium~3 (Melega and Howard, 1981). This differs from normal cholinergic neurons for which the high~ affinity choline transporter is sensitive to 211 Table A1 Activity and Content of Noradrenergic and Cholinergic Synthetic Enzymes and N eurotransmitters in PC12 Cells Undifferentiated NGF-Difl'erentiated Enzymes Enzyme Activity (pmol/min/ mg protein) Tyrosine Hydroxylase 39 1 5 10 1 1 Dopa decarboxylase 770 1 99 130 1 15 Dopamine fl ~hydroxylase 806 1 84 161 1 19 Choline Acetyltransferase‘ 440 1 17 520 :I: 20 N eurotransmitters Neurotransmitter Content (pmol/mg protein) Dopamine 16.61 1.7 4.41 0.4 Norepinephrine 6.11 6 1.51 0.2 Epinephrine <0.15 <0.15 Acetylcholine‘ 1335 1 77 5594 1 637 PC12 cells were treated with 50 mg/ml of NGF for 14 (noradrenergic properties) or 22 (cholinergic properties) days prior to measuring enzyme activity and neurotransmitter content. All values are expressed as mean 1 SEM and protein content is total cellular protein content. .Values from: Greene and Rein, Nature (Land) 268: 349-352, 1977b. All other values from: Greene and Tischler, PIOC. Natl. Acad. Sci. (U.S.A.) 73: 2424—2428, 1976. 212 hemicholinium~3. In undifferentiated PC12 cells, ACh and catecholamines are stored in dense-core vesicles, but based on reserpine sensitivity and electron microscopy, the two transmitters are not co~localized (Schubert and Kleir, 1977). Evidence that release of ACh from PC12 cells is quantal was demonstrated by co~culture of PC12 cells with a skeletal muscle clonal cell line, L~6. After as few as 3 days of co~culture in NGF-containing medium, a ~bungarotoxin~ and d-tubocurarine-sensitive miniature endplate-potentials could be recorded from L-6 myotubes (Schubert et al., 1977). Howard and co~workers loaded PC12 cells with 2H4~choline and used combined gas chromatography-mass spectrometry to measure levels of endogenous 2ITO-ACh and newly synthesized 2H4-ACh in the presence and absence of vesamicol, which inhibits uptake of ACh into vesicles (Marshall, 1970). Their results clearly indicate that newly synthesized ACh in PC12 cells is stored in vesicles (Figure A.1A) and evoked-release of 2H4~ACh (Figure A.1B)is predominantly vesicular in nature (Melega and Howard, 1984; Howard at al., 1986). Thus, in these respects, release of ACh from PC12 cells mirrors normal release from native cholinergic terminals. Effects of neurotoxins on cholinergic properties of PC12 cells have not been widely studied. The effects of tetanus toxin on ACh release from PC12 cells indicate that this neurotoxin inhibits ACh release at subnanomolar concentrations (Sandberg et al., 1989a) and that tetanus toxin may inhibit ACh release by interfering with cyclic-GMP metabolism (Sandberg, et al., 1989b). 213 120 l . (Eek, g \ \k, \t x E ‘, f\i o o \ \ * 4o - \ I !\ \ \ 0 .01 1 Io [VESAMICOL] (uti) Specific E] I'HJAChII‘HJAChI .cnvay 0 r P w 18 % of VESAMICOL-FREE CONTROLS TOTAL ACh EVOKED CONTENT RELEASE Figure A.1. Vesicular nature of ACh release from PC12 cells. Effects of vesamicol (AH5183) on the cellular and vesicular content of acetylcholine and on acetylcholine release in PC12 cells. A.) PC12 cells were incubated for 30 min with 10 p. M [3H]choline in the presence of vesamicol. Acetylcholine (ACh) was extracted from the cells (triangles) and from a vesicle fraction (circles) of a cell homogenate, and the levels of [2H4]ACh (solid symbols) and native [ZHOJACh (open symbols) were measured. The results are expressed as the percentage of control values. Values shown are the mean 1 SD for triplicate incubations. B.) Release of ACh from vesamicol-treated cells during 5 min incubations. Total cellular ACh content is the sum of the [2H01ACh or the [2H4]ACh released to the buffers during a 5 min incubation and the amount of ACh remaining in the cells after the incubations. Specific activity is the ratio of [2H4]-ACh/ total ACh. The results are expressed as a percentage of control values for cells that were incubated in the absence of vesamicol. The values are the mean 1 SD for quadruplicate incubations. Modified from: Howard er al., In: Calcium, Neumnal Function and Transmitter Release. (R. Rahamimofi and B. Katz, eds). p. 282 (1986). 214 2) Ion channels in PC12 cells. K+ Channels. PC12 cells also express multiple K+ conductances and channel subtypes. Several groups have reported the presence of a large Ca2*~activated K+ conductance in PC12 cells (Hoshi and Aldrich, 1988a; Auguste et al., 1989; Pun and Behbehani, 1990). This current is blocked readily by agents which decrease Ca2+ influx into PC12 cells, such as inorganic divalent cations (Pun and Behbehani, 1990). More than one type of Ca2+~activated K+ channel may be present in PC12 cells, as 86Rb+ efflux is only completely blocked by a combination of tetraethylammonium and apamin (Schmid-Antomarchi et al., 1986), a polypeptide toxin from bee venom which blocks some types of CaZI-activated K+ channels. PC12 cells express large amounts of the apamin receptor after as little as 4 days in culture. Interestingly, after 7 days in culture in the presence of NGF, the number of apamin receptors/ mg protein was 10 fold less than in undifferentiated cells (Schmid-Antomarchi et al., 1986). Thus, by contrast to N a“ and Ca2+ channels, N GF treatment does not increase the expression of this channel type. Apamin was used in affinity labelling experiments in an attempt to identify components of the apamin receptor in PC12 cells. 12$I~labeled derivatives of apamin bound to a 30 kilodalton receptor in PC12 cell membranes (Auguste er al., 1989). Since this cell line expresses large amounts of the apamin receptor, it may be useful in isolating the genetic sequences which code for this type of Ca2+-activated K+ channel. Whole-cell currents similar to those mediated by rapidly inactivating (IA) (Rogawski et al., 1988; Pun and Behbehani, 1990) and delayed rectifier (Ix) K+ 215 channels (Rogawski, et al., 1988; Hoshi and Aldrich, 1988a) have also been reported for PC12 cells. Phencyclidine (1~100 n M) and tetraethylammonium chloride (20 mM) block IK (Rogawski, et al., 1988), whereas IA is inhibited by 2 mM 4- aminopyridine (Rogawski et al., 1988). Hoshi and Aldrich (1988a,b) have characterized four voltage-dependent K+ channels which may underlie whole-cell K” currents in PC12 cells. The kinetics of three of these channels are similar to those of delayed rectifier K“ channels, whereas the kinetics of the fourth channel are similar to kinetics of the inactivating K+ channel. However, the channels in PC12 cells also have characteristics which are unique from those of channels which mediate IK and I A (Hoshi and Aldrich, 1988a,b). Thus, it cannot be assumed that K+ channels in PC12 cells behave in a fashion identical to those described elsewhere. Finally, it has been demonstrated that the peptide galanin inhibits DA secretion from PC12 cells by activating a K" conductance (de Weille et al., 1989). Galanin had no effect on L~type Ca2+ channels, but its action on K+ conductances was inhibited by pretreatment with pertussis toxin. These results suggest that K+ channels and the galanin receptor may be linked by a G-protein in PC12 cells (de Weille et al., 1989). Whether or not the galanin response is mediated by a voltage- sensitive or receptor operated channel is presently unknown. By contrast to Ca2+ channels, the effects of N GF ~differentiation on voltage- sensitive K+ channels have not yet been widely studied. Thus, it is unclear at this point whether “or not changes in distribution or composition of K+ channel types play 216 an important role in, or are affected by the process of differentiation of PC12 cells. Toxicological studies utilizing K+ channels in PC12 cells are also lacking. Na+ channels. Like Ca2+ channels, expression of voltage-gated Na+ channels is increased by differentiation of PC12 cells with NGF. TTX-sensitive, Ca2+- independent action potentials in PC12 cells differentiated with NGF for 2 weeks were observed by Dichter et al. (1977). Undifferentiated PC12 cells express functional Na+ channels, but differentiation with N GF increased TTX-sensitive influx of ”Na” by approximately 12 fold (Reed and England, 1986). Induction of a TTX-insensitive N a+ flux by NGF has also been reported (Rudy et al., 1987). The mechanisms underlying N GF-induced increases in Na“ channel function are presently unclear. Increases in Na+ channel function by NGF were mimicked by increases in cAMP, and responses to N GF and CAMP were blocked by an inhibitor of CAMP-dependent protein kinase (PKA), suggesting that the expression of Na+ channels in response to NGF is in part mediated by PKA (Kalman et al., 1990). However, Pollock et al. (1990) reported that CAMP did not stimulate neurite outgrowth and decreased INa in PC12 cells, whereas FGF was as effective as NGF in increasing IN, In addition to NGF, treatment with DEX also increased the expression of Na” current in PC12 cells (Figure A.2, Garber et al., 1989), although this response appears to be inconsistent. Treatment with DEX and NGF produced less of an increase in Na+ current than treatment with N GF alone, and despite treatment with NGF and / or DEX, some PC12 cells failed to express measurable whole-cell N a“ conductances (Figure A.2) (Garber et al., 1989). Furthermore, Pollock 217 A serum 3. ~50toOmV C. m 1.33: Figure A.2. Voltage-dependent Na+ channels in PC12 cells. Representative families of the voltage- dependent Na+ currents from untreated, dexamethasone-(DEX), nerve gowth factor- (N GF), and DEX/NGF-treated cells recorded using the whole-cell patch voltage-clamp technique. In each treatment group, some cells (A) expressed no or little Na+ current while others (B) expressed appreciable amounts of N a+ current. C, Peak current-voltage (I / V) curves were obtained from cells which did not express the Na+ current (black squares) and those which expressed the Na+ current (circles). As K+ was often used in internal solutions, outward K" currents are present in some families of current records. From: Garber et al. Mm 9:3976-3987 (1989). 218 et al., (1990) reported that DEX failed to increase IN“, and inhibited NGF-induced increases in IN, in PC12 cells. Finally, several other treatments including DEX, mitotic inhibitors and dimethyl sulfoxide, which induced physiological differentiation in neuroblastoma cell lines, were less effective at increasing N a+ current than N GF . In addition, ACh responses induced by N GP in PC12 cells were not necessarily coupled to increases in N a“ channel function (Ifune and Steinbach, 1990). Further study is needed to understand clearly the regulation of Na“ channel expression in PC 12 cells. 3) Receptors on PC12 cell membranes. Another potential use of PC12 cells in neurotoxicology is for the study of interactions of compounds with neurotransmitter receptors or receptor / ligand-activated channels. Both nicotinic (Greene and Rein, 1977b) and muscarinic ACh receptors (Jumblatt and Tischler, 1982; Cross at al., 1984) are expressed in PC12 cell membranes, as well as adenosine receptors (Guroff et aL, 1981). A recent report also suggests that PC12 cell membranes may also contain N-methyl-D-aspartate receptors which stimulate phosphoinositide hydrolysis, increased [Ca2+], and NE release (Kurozumi et al., 1990). Adenosine or adenosine receptor agonists stimulate adenylate cyclase resulting in increased tyrosine hydroxylase activity in PC12 cells (Rabe and M‘Gee, 1983; Noronha~Blob et al., 1986). Radioligand binding studies have demonstrated that adenosine receptors in PC12 cells are of the A2 subtype (Williams et al., 1987). Short pretreatment of PC12 cells with adenosine increased depolarization-evoked release 219 of [3H]-NE and [3H]~ACh without altering 45Ca2“ influx or depolarization as measured by “RE efflux through nicotinic ACh receptor channels (Rabe and M‘Gee, 1983). It has been suggested that adenosine or adenosine nucleotides may act via A2 receptors in a paracrine or autocrine fashion (Roskoski and Roskoski, 1989). ATP released during exocytosis of NE— or ACh-containing vesicles may be degraded to adenosine and stimulate A2 receptors. In PC12 cells, the activity of adenosine is terminated by re-uptake (Roskoski and Roskoski, 1989). Activation of muscarinic ACh receptors in undifferentiated PC12 cells results in elevations in inositol triphosphate (1P3) and free [Ca2+]i (Vincentini et al., 1985), followed by neurotransmitter release (Rabe et al., 1987). Inoue and Kenimer (1988) have suggested that a receptor-activated Ca“ channel may mediate muscarinic receptor-induced neurotransmitter release. Muscarinic agonists increased 45Ca2+ influx, and neurotransmitter release was not blocked by Ca2+ channel blockers. Moreover, the influx of 4“"Ca2" and neurotransmitter release were inhibited by pertussis toxin, indicating the possible involvement of G-proteins (Inoue and Kenimer, 1988). However, neurotransmitter release and phosphoinositide (PI) hydrolysis stimulated by muscarinic receptor activation appear to be separate events. In Ca2+~free solutions, methacholine stimulated PI hydrolysis but not NE release. By contrast, activators of PKC inhibited PI turnover but not NE release (Takashima and Kenimer, 1989). Similarly, activation of muscarinic receptors in PC12 cells also results in an inhibition of CAMP which is a separate event from P1 hydrolysis (Horwitz, 1989). However, there is no evidence that these actions are mediated by 220 different receptor subtypes (Horwitz, 1989). Thus, stimulation of the muscarinic ACh receptor in PC12 cells may also stimulate cellular processes in addition to neurotransmitter release. Characterization of muscarinic receptor subtypes in undifferentiated PC12 cells by pharmacological sensitivity of [3H]~NE release and [3H]~telenzipine (a selective antagonist for the M1~type muscarinic receptor) binding indicates that PC12 cells contain more than one type of muscarinic receptor (Bt'inish et al., 1990). However, the predominant receptor subtype clearly is not of the M2 or M3 category, and also may not be of the M1 subtype (Michel et al., 1989; Bdnish et al., 1990). It has been proposed that the muscarinic ACh receptor in PC12 cells may be of the M4 subtype (Michel et al., 1989). Further study is needed to define better the type of muscarinic ACh receptor(s) present on PC12 cells. Effects of organophosphate insecticides on muscarinic ACh receptors in PC12 cells indicate that 50 uM soman decreased the number of muscarinic receptors after 24 h. of exposure. However, the insecticides did not inhibit binding of radiolabeled muscarinic antagonists, nor did atropine prevent organophosphate-induced decreases in muscarinic receptor numbers. Furthermore, decreases in muscarinic receptors induced by organophosphates were qualitatively different fiom those produced by the ACh agonist carbamylcholine. Thus, organophosphate insecticides decrease the numbers of muscarinic ACh receptors by a mechanism other than agonist-induced receptor desensitization (Viana et aL, 1988). 221 Stimulation of nicotinic and muscarinic ACh receptors results in a rapid increase in PKC activity. However, this response appears to be largely the result of nicotinic receptor activation, as muscarinic antagonists only Caused small reductions of carbachol-induced PKC activity (Messing et al., 1989). Furthermore, the nicotinic response was dependent on Ca2+ influx via voltage-dependent Ca2+ channels, as the DHP antagonist nifedipine reduced PKC activity to the level observed during muscarinic receptor activation (Messing et al., 1989). Recent binding studies of nicotinic ACh receptors in PC12 cells using [3H]-ACh and a ~bungarotoxin indicate that there may be heterogeneous expression of nicotinic receptor subtypes (Lukas, 1990). M°Gee and co~workers have used PC 12 cells to examine agonist-induced down~ regulation of the nicotinic ACh receptor (Robinson and McGee, 1985). The nicotinic ACh receptor is not the target of CAMP-dependent phosphorylation in PC12 cells (M‘Gee and Liepe, 1984), but high concentrations of forskolin can impair receptor function by anesthetic-like effects (M°Hugh and McGee, 1986). Furthermore, nicotinic ACh receptors are not influenced by prolonged depOlarization or elevations in [Ca2+]i (DeLorme and McGee, 1988). Adenylate cyclase does not appear to be involved in desensitization of nicotinic ACh receptors in PC12 cells, as forskolin and its analogs do not markedly alter agonist-induced desensitization of nicotinic ACh receptors in PC12 cells (Nishizawa et al., 1990). Further study of agonist-induced desensitization of nicotinic ACh receptors is needed to elucidate the mechanisms responsible for desensitization in PC12 cells. However, enough information exists about the types 222 and functions of ACh receptors in PC12 cells to allow examination of interactions between neurotoxins and these receptor types. D) CLONAL VARIABILITY AND SUBCLONES OF PC12 When using any model system, one must know the weaknesses and limitations of the model in order to maximize its usefulness while minimizing the likelihood of error or overinterpretation of results. One limitation of the PC12 cell line is that the number of passages or subcultures from the original isolation of the line is unknown. Second, PC12 cell cultures have a tendency to become heterogenous with respect to cell morphology after approximately 20 passages (Guroff, 1985). Thus, it is often difficult to compare directly results from different laboratories due to differences in passage number. This is most apparent for studies of the regulation of Na+ channel expression, and undoubtedly limits the use of this line for screening and/ or regulatory purposes. However, PC12 cells remain an excellent model for mechanistic studies of disruption of neuronal function. By routinely keeping track of the number of passages from receipt of the cell line, one can maintain internal consistency by using the same "passage number" for experiments and can begin a new culture from a frozen stock when the number of passages approaches 20. A number of subclones of PC12 exist due to spontaneous and/ or induced mutations. Heterogeneity in NE and DA synthesis and responses to glucocorticoids has been reported (Koike and Takashima, 1984). Other clones vary in their ability to respond to N GP or in the type of response induced by N GF (see Guroff, 1985). 223 Presently, the subclones of PC12 are used primarily in studies of the action of NGF. However, they may also represent a useful tool in neurotoxicology as well. E) SUMMARY PC12 cells have been widely used for studies of neuronal biochemistry and physiology. In addition to responses to NGF, the neurotransmitter, ion channel and receptor properties of this cell line have been well characterized. A well characterized and understood model system is a prerequisite for mechanistic studies of the action of toxic compounds. PC12 cells represent a potentially useful model for the study of mechanisms of action of neurotoxic compounds on neuronal differentiation and maturation, synthesis, storage and release of DA, NE and ACh, as well as the function of voltage-sensitive Na”, K+ and Ca2+ channels. With respect to Ca2+ channels, the subtypes of Ca2+ channels expressed in this cell line are extremely well characterized by ion flux, binding and electrophysiological techniques. In addition, the composition of Ca2+ channel subtypes can be altered predictably by NGF-differentiation, permitting accurate electrophysiological assessment of effects of putative toxicants on these important membrane macromolecules. Several receptor types, including nicotinic and muscarinic ACh receptors and adenosine receptors are also becoming well characterized in PC12 cells. Thus, this cell line, which has been used only sparingly in neurotoxicological studies to date, may in the future become an extremely useful model system. REFERENCES REFERENCES ABE, T., HAGA, T. and KUROKAWA, M. (1975). Blockage of axonal transport and depolymerization of assembled microtubules by methylmercury. Brain Res. 86:504-508. ABE, T., KOYANO, K., SAISU, H., NASHIUCHI, Y. and SAKAKIBARA, S. (1986). Binding of Q'COIIOIOXIII to receptors associated with the voltage-sensitive calcium channel. Neurosci. Lett. 71:203-208. ADAMS, M. E., BINDOKAS, V. P., HASEGAWA, L. and VENEMA, V. J. (1990). (t) ~Agatoxins: Novel calcium channel antagonists of two subtypes from funnel web spider (Agelenopsis aperra) venom. J. Biol. Chem. 265:861-867. ADAM-VIZI, V. and LIGETI, E. (1986). Calcium uptake of rat brain synaptosomes as a function of membrane potential under different depolarizing conditions. J. Physiol. 372:363-377. AHNERT—I-IILGER, G., BRAUT'IGAM, M. and GRATZL, M. (1987). Ca2*~stimulated catecholamine release from alpha-toxin-permeabilized PC12 cells: biochemical evidence for exocytosis and its modulation by protein kinase C and G proteins. Biochem. 26:7842-7848. 225 AHNERT-HILGER, G. and GRATZL, M. (1987). Further characterization of dOpamine release by permeabilized PC12 cells. J. Neurochem. 49:764~770. AKERMAN, K. E. 0. and NICHOLLS, D. G. (1981). Ca2+ transport by intact synaptosomes. The voltage-dependent Ca 2+ channel; re-evaluation of the role of sodium/ calcium exchange. Eur. J. Biochem. 117:491-497. ANDERSON, D. J. and AXEL, R. (1986). A bipotential neuroendocrine precursor whose choice of cell fate is determined by N GF and glucocorticoids. Cell 47: 1079-1090. ANDERSON, A. J. and HARVEY, A. M. (1987). m-Conotoxin does not block the verapamil-sensitive calcium channels at mouse motor nerve terminals. Neurosci. Lett. 82:177-180. ARMSTRONG, C. M., SWENSON, R. P. and TAYLOR, S. R. (1982). Block of squid axon K channels by internally and externally applied barium ions. J. Gen. Physiol. 80:663-682. ATCHISON, W. D. (1986). Calcium-dependent and-independent effects of methylmercury on spontaneous and potassium-evoked release of acetylcholine at the neuromuscular junction. J. Pharmacol. Exp. Ther. 237 :672-680. 226 ATCHISON, W. D. (1989). Dihydropyridine-sensitive and ~inscnsitive components of acetylcholine release from rat motor nerve terminals]. Pharmacol. Exp. Then 251:672-678. ATCHISON, W. D. and NARAHASHI, T. (1982). Methylmercury induced depression of neuromuscular transmission in the rat. Neurotoxicology 3:37-50. ATCHISON, W. D., CLARK, A. W. and NARAHASHI, T. (1984). Presynaptic effects of methylmercury at the mammalian neuromuscular junction. In: Cellular and molecular neurotoxicology (Narahashi, T., ed.). Raven Press, New York, pp. 23-43. ATCHISON, W. D. and NARAHASHI, T. (1984). Mechanism of action of lead. Neurotoxicology 5:267-284. ATCHISON, W. D., JOSHI, U. and THORNBURG, J. E. (1986). Irreversible supression of calcium entry into nerve terminals by methylmercury. J. Pharmacol. Exp. Ther. 238:618-624. ATCHISON, W. D. and O’LEARY, S. M. (1987). Bay K 8644 increases release of acetylcholine at the murine neuromuscular junction. Brain Res. 419:315-319. 227 ATCHISON, W. D., ADGATE, L. and BBAMAN, C. M. (1988). Effects of antibiotics on uptake of calcium into isolated nerve terminals]. Pharmacol. Exp. Ther. 245:394-401. AUDESIRK, G. (1987). Effects of in vitro and in vivo lead exposure on voltage- dependent calcium channels in Lymnaea stagnalis. Neurotoxicol. 8:579-592. AUDESIRK, G. and AUDESIRK, T. (1989). Effects of in vitro lead exposure on voltage-sensitive calcium channels differ among cell types in Lymnaea stagnalis. Neurotoxicol. 10:659-670. AUGUSTE, P., HUGUES, M. and LAZDUNSKI, M. (1989). Polypeptide constitution of receptors for apamin, a neurotoxin which blocks a class of Can-activated K+ channels. FEBS Lett. 243350154. AUGUSTINE, G. J., CHARLTON, M. P. and SMITH, S. J. (1987). Calcium action in synaptic transmitter release. Ann. Rev. Neurosci. 10:633-693. BAKIR, F., DAMLUJI, S. F., AMIN-ZAKI, L., MURTADHA, M., KHALIDI, A., AL-RAWI, N. Y., TIKRITI, S., DHAHIR, H. 1., CLARKSON, T. W., SMITH, J. C. and DOHERTY, R. A. (1973). Methylmercury poisoning in Iraq. Science 181:230-241. 228 BAIZER, L. and WEINER, N. (1985). Regulation of dopamine release from PC12 pheochromocytoma cell cultures during stimulation with elevated potassium or carbachol. J. Neurochem. 44:495-501. BARRETT, J ., BOTZ, D. and CHANG, D. B. (1974). Block of neuromuscular transmission by methylmercury. Behavioral toxicology. In: Early detection of occupational hazards. (Zintaras, C., Johnson, B. L. and de Groot, I., eds). US. Dept. of Health, Education, and Welfare. 5:277-287. BEAN, B. P. (1985). Two kinds of calcium channels in canine atrial cells. J. Gen. PhysioL 86:1-30. BELLUZZI, 0., SACCHE, O. and WANKE, E. (1985). A fast transient outward current in rat sympathetic neurone studied under voltage-clamp conditions. J. Physiol. (Lona) 358:91-108. BERGAMASCHI, S., TRABUCCHI, M., BA'ITAINI, F., PARENTI, M., SCHETTINI, G., MEUCCI, O. and GOVONI, S. (1990). Modulation of dihydropyridine-sensitive calcium channels: A role for G proteins. Eur. Neurol. 30:16-20. 229 BERGLUND, F., BERLIN, M., BIRKE, G., VON FULER, U., FRIBERG, L., HOLMSTEDT, B., JONSSON, E., RAMELC., SKERFVING, S., SWENSSEN, A. and TEJNING, S. (1971). Methylmercury in fish. A toxicologic epidemiologic evaluation of risks. Report from an expert group. Nord. Hyg. Tidskr. Suppl. 4. BIDSTRUP, P. L. (1964). Toxicity of mercury and its compounds. Elsevier Publishing Co., Amsterdam-London-New York. p.74. BLAUSTEIN, M. P. (1975). Effects of potassium, veratridine and scorpion venum on calcium accumulation and transmitter release by nerve terminals in vitm. J. Physiol. 247:617-655. BLAUSTEIN, M. P. and GOLDRING, B. M. (1975). Membrane potentials in pinched-off presynaptic nerve terminals monitored with a flourescent probe: Evidence that synaptosomes have potassium diffusion potentials. J. Physiol. (Land) 247:589-615. BLAUSTEIN, M. P. and ECTOR, A. C. (1975). Barbiturate inhibition of calcium uptake by depolarized nerve terminals. Mol. Pharmacol. 11:369-378. 230 BLAUSTEIN, M. P. and OBORN, C. J. (1975). The influence of sodium on calcium fluxes in pinched-off nerve terminals in vitro. J. PhysioL (Land) 247:657-686. BLAUSTEIN, M. P., RATZLAFF, R. W. and KENDRICK, N. K. (1978). The regulation of intracellular calcium in presynaptic nerve terminals. Ann. NY. Acad. Sci. 307:195-211. BOLAND, L. M. and DINGLEDINE, R. (1990). Multiple components of both transient and sustained barium currents in a rat dorsal root ganglion cell line. J. PhysioL (Land. ) 420223-245. BONDY, S. C., ANDERSON, C. L, HARRINGTON, M. E. and PRASAD, K. N. (1979). The effects of organic and inorganic lead and mercury on neurotransmitter high-affinity transport and release mechanisms Environ. Res. 19:102-111. BONISCH, H., BOER, R., DOBLER, M. and SCHUDT, C. (1990). Pharmacological characterization of muscarine receptors of PC12 (rat phaeochromocytoma) cells. Naunyn-Schmiedeberg. Arch Pharmacol. 341:158-164. BOROWI'IZ, J. L. (1974). Mechanism of adrenal catecholamine release by divalent mercury. Toxicol. Appl. Pharmacol. 28:82-87. 231 BRADFORD, H. F . (1975). Isolated nerve terminals as an in vitna preparation for study of dynamic aspects of transmitter metabolism and release. In-Handbook of psychopharmacology. (Iverson, L. L, Iverson, S. D. and Snyder, S. H., eds.) Plenum Press, New York, pp 191-252. BRIGANT, J. L and MALLART, A. (1982). Presynaptic currents in mouse motor nerve endings. J. Physiol. (Land. ) 333:619-636. BUSSELBERG, D., EVANS, M. L, RAHMANN, H. and CARPENTER, D. O. (1990). Lead inhibits the voltage-activated calcium current of Aplysia neurons. Toxicol. Lett. 5:51-57. BYRD, J. C., NARANJO, J. R. and LINDBERG, I. (1987). Proenkephalin gene expression in the PC12 pheochromocytoma cell line. Endocrinology 121:1299- 1305. CAMPBELL, K., P., LEUNG, A. T. and SHARP, A. H. (1988). The biochemistry and molecular biology of the dihydropyridine-sensitive calcium channel TINS. 11:425-430. CARBONE, E. and LUX, L D. (1984). A low voltage-activated, fully inactivating Ca channel in vertebrate sensory neurones. Nature (Land. ) 310:501-503. 232 CARROLL, A. G., RHOADS, A. R. and WAGNER, P. D. (1990). Hydrolysis- resistant GTP analogs stimulate catecholamine release from digitonin- permeabilized PC12 cells. J. Neurachem. 55:930-936. CHARLTON, M. P. and AUGUSTINE, G. J. (1990). Classification of presynaptic calcium channels at the squid giant synapse: neither T-, L- nor N-type. Brain Res. 525:133-139. CHANG, L W. (1977). Neurotoxic effects of mercury. A review. Env. Res. 14:329- 373. CHANG, L W. (1979). Pathological effects of mercury poisoning. In: The biochemistry of mercury in the environment. (Nriagu, ed.) Elsevier/North Holland Biomedical Press. CHANG, L W. (1980). Mercury. In: Experimental and Clinical Neurotoxicology (Spencer, P. S. and Schaumburg, H. H., eds.). The Williams and Wilkins Co., Baltimore. pp. 508-526. CHANG, L W. and HARTMANN, H. A. (1972a). Ultrastructural studies of the nervous system after mercury intoxication. 1. Pathological changes in the nerve cell bodies. Acta Neuropath. 20:122-138. 233 CHANG, L W. and HARTMANN, H. A. (1972b). Ultrastructural studies of the nervous system after mercury intoxication. 11. Calcium ion-induced swelling. Biochem. J. 95:378-386. CHANG, L W. and HARTMANN, H. A. (1972c). Blood-brain barrier dysftmction in experimental mercury intoxication. Acta Neurapath. 21:179-184. CHERKSEY, B. D., SUGIMORI, M. and LLINAS, R. R. (1990). Isolation of a calcium channel from mammalian cerebellar tissue using synthetic FTX. Neurasci. Abstr. 16:956. CHOI, D. W. (1988). Calcium-mediated neurotoxicity: relationship to specific channel types and role in ischemic damage. TINS. 11:465-468. CHOI, O. H., PADGETI‘, W. L, NISHIZAWA, Y., GUSOVSKY, F ., YASUMOTO, T. and DALY, J. W. (1990). Maitotoxin: Effects on calcium channels, phosphoinositide breakdown, and arachidonate release in pheochromocytoma PC12 cells. Mol. Pharmacol. 37:222-230. COHEN, M. w. JONES, o. T. and ANGELIDES, K. J. (1991). Distribution of Ca2+ channels in frog motor nerve terminals revealed by fluorescent m -conotoxin. J. Neurosci. 11:1032-1039. 234 COOPER, G. P. and MANALIS, R. S. (1983). Influence of heavy metals on synaptic transmission: A Review. Neurotoxicology. 4:69-84. COOPER, G. P. and MANALIS, R. S. (1984a). Interactions of lead and cadmium on acetylcholine release at the frog neuromuscular junction. Taxical. App. Pharm. 74:411-416. COOPER, G. P. and MANALIS, R. S. (1984b). Cadmium: effects on transmitter release at the frog neuromuscular junction. Eur. J. Pharmacol. 99:251-256. COOPER, G. P., SUSZKIW, J. B. and MANALIS, R. S. (1984). Presynaptic effects of heavy metals. In: Cellular and molecular neurotoxicology. (Narahashi, T., ed.). Raven Press, New York. pp. 1-21. COURTNEY, K. R. (1975). Mechanism of frequency dependent inhibition of sodium currents in frog myelinated nerve by the lidocaine derivative GEA 968. J. Pharmacol. Exp. Ther. 195:225-236. COUTINHO, O. P., CARVALHO, C. A. M. and CARVALHO, A. P. (1984), Calcium uptake related to K*-depolarization and Na+ /Ca2" exchange in sheep brain synaptosomes. Brain Res. 290:261-271. 235 CROSS, A. J ., JOHNSON, J. A., FRITH, C. and TAYLOR, G. R. (1984). Muscarinic cholinergic receptor in a rat phaeochromocytoma cell line. Biochem. Biophys. Res. Comm. 119:163-167. CRUZ, L J. and OLIVERA, B. M. (1986). Calcium channel antagonists. m -Conotoxin defines a new high affinity site. J. Biol. Chem. 261:6230-6233. DAMON, D. H., D’AMORE, P. A. and WAGNER, J. A. (1990). Nerve growth factor and fibroblast growth factor regulate neurite outgrowth and gene expression in PC12 cells via both protein kinase C- and CAMP-independent mechanisms. J. Cell Biol. 110:1333-1339. DANIELL, 1. c., BARR E. M. and LESLIE, s. w. (1983). 45Ca2+ uptake into rat whole brain synaptosomes unaltered by dihydropyridine calcium antagonists. J. Neurachem. 41:1455-1459. DELORME, E. M. and MCGEE, R. (1988). Effects of prolonged depolarization on the nicotinic acetylcholine receptors of PC12 cells. J. Neurochem. 50:1248- 1252. 236 DELORME, E. M., RABE, C. S. and MCGEE, R. (1988). Regulation of the number of functional voltage-sensitive Ca+ + channels on PC12 cells by chronic changes in membrane potential. J. Phannacal Exp. Ther. 244:838-843. DE WEILLE, J ., FOSSET, M., LAZDUNSKI, M. and SCHMID-ANTOMARCHI, H. (1989). Galanin inhibits dopamine secretion and activates a potassium channel in pheochromocytoma cells. Brain Res. 485:199-203. DICHTER, M. A., TISCHLER, A. S. and GREENE, L A. (1977). Nerve growth factor-induced increase in electrical excitability and acetylcholine sensitivity of a rat pheochromocytoma cell. Nature (Land) 268:501-504. D’ITRI. F. Concentrations of contaminants if fishes of the great lakes. In: Toxic contamination in large lakes. Volume 2: Impact of toxic contamination on fisheries management. (Schmidtke, N. W., ed.). Lewis Publishers, Chelsea. pp 51-84. DI VIRGILIO, F., POZZAN, T., WOLLHEIM, C. B., VICENTINI, L M. and MELDOLESI, J. (1986). Tumor promoter phorbol myristate acetate inhibits Ca2+ influx through voltage-gated Ca2+ channels in two secretory cell lines, PC12 and RINmSF. J. Biol. Chem. 261:32-35. 237 DOOLEY, D. J., LUPP, A. and HERTI‘ING, G. (1987a). Inhibition of central neurotransmitter release by m-conotoxin GVIA, a peptide modulator of the N-type voltage-sensitive calcium channel. Naunyn-Schmiedeberg Arch. Pharmacol. 336:467-470. DOOLEY, D. J., MAHLMANN, H., BRENNER, O. and OSSWALD, H. (1987b). Characterization of the dihydropyridine binding sites of rat neocortical synaptosomes and microvessels. J. Neurochem. 49:900-904. DOUPE, A J., LANDIS, S. C. and PATTERSON, P. H. (1985). Environmental influences in the development of neural crest derivatives: Glucocorticoids, growth factors and chromaffin cell plasticity. J. Neurosci. 5:2119-2142. DRAPEAU, P. and BLAUSTEIN, M. P. (1983). Initial release of [3Hldopamine from rat striatal synaptosomes: Correlation with calcium entry. J. Neurosci. 3:703-713. DUBEL S. J. and SNUTCH, T. P. (1990). Primary structure and expression of the rat brain class-B calcium channel. Neurasci. Abstr. 16:1174. 238 DUTAR, P., PASCOL O. and LAMOUR, Y. (1989). m-Conotoxin GVIA blocks synaptic transmission in the CA1 field of the hippocampus. Eur. J. Pharmacol. 174:261-266. DUNN, S. M. J. (1988). Multiple calcium channels in synaptosomes: Voltage dependence of 1,4-dihydropyridine binding and effects on function. Biochemistry 27:5275-5281. ECKERT, R. and CHAD, J. E. (1984). Inactivation of Ca channels. Prog. Biophys. Molec. Biol. 44:215-269. EDER-COLLI, L, BRIAND, P. A., PELLEGRINELLI, N. and DURANT, Y. (1989). A monoclonal antibody raised against plasma membrane of cholinergic nerve terminals of the Torpedo inhibits choline acetyltransferase activity and acetylcholine release. J. Neurochem. 53:1419-1427. FATT, P. and KATZ. B. (1952). Spontaneous subthreshold activity at motor nerve endings. J. Physiol. (Land) 117:109-128. FENWICK, E. M., MARTY, A. and NEHER, E. (1982). Sodium and calcium channels in bovine chromaffin cells. J. Physiol. 331:599-635. 239 FLOOR, E. (1983). Substance P release from K+-depolarlzed synaptosomes at one second resolution. Brain Res. 279:321-324. FORSHAW, P. J. (1977). The inhibitory effect of cadmium on neuromuscular transmission in the rat. Eur. J. Pharmacol. 42:371-377. FOX, A. P., NOWYCKY, M. C. and TSIEN, R. W. (1987a). Kinetic and pharmacological properties distinguishing three types of calcium currents in chick sensory neurones. J. Physiol. (Land) 394:149- 172. FOX, A. P., NOWYCKY, M. C. and TSIEN, R. W. (1987b). Single channel recordings of three types of calcium channels in chick sensory neurons. J. Physiol. (Land) 394: 173-200. FRANKENHEUSER, B. and HODGKIN, A. L. (1957). The action of calcium on the electrical properties of squid axons. J. Physiol. (Land). 137:218-244 FRIEDMAN, M. B., ERIKSON, C. K. and LESLIE, S. W. (1980). Effects of acute and chronic ethanol administration on whole mouse brain synaptosomal calcium influx. Biochem. Pharmacol. VOL:1903-1908. 24o FUJITA, K., LAZAROVICI, P. and GUROFF, G. (1989). Regulation of the differentiation of PC12 pheochromocytoma cells. Environ. Health Perspective 1989; 80: 127-142. GARBER, S. S., HOSHI, T. and ALDRICH, R. W. (1989). Regulation of ionic currents in pheochromocytoma cells by nerve growth factor and dexamethasone. J. Neurosci. 9:3976-3987. GOULD, R. J ., MURPHY, K. M. and SNYDER, S. H. (1985). Autoradiographic localization of calcium channel antagonist receptors in rat brain with 3H- nitrendipine. Brain Res. 330:217-223. GRAY, E. G. and WHITAKER, V. P. (1962). The isolation of nerve endings from brain: An electron microscopic study of cell fragments derived by homogenization and centrifugation. J. Anat. (Land) 96:79-87. GREENBERG, D. A., CARPENTER, C. L and MESSING, R. O. (1987a) Ethanol-induced component of 45Ca2+ uptake in PC12 cells is sensitive to Ca2+ channel modulating drugs. Brain Res. 410: 143-146. 241 GREENBERG, D. A., CARPENTER, C. L. and MESSING, R. O. (1987b). Lectin-induced enhancement of voltage-dependent calcium flux and calcium channel antagonist binding. J. Neurochem. 48:888-894. GREENE, L. A. and TISCHLER, A. S. (1976). Establishment of a noradrenergic clonal line of rat adrenal pheochromocytoma cells which respond to nerve growth factor. Proc. Natl. Acad Sci. (USA) 73:2424-2428. GREENE, L. A. and REIN, G. (1977a). Release, storage and uptake of catecholamines by a clonal cell line of nerve growth factor (NGF) responsive pheochromocytoma cells. Brain Res. 129:247-263. GREENE, L. A. and REIN, G. (1977b). Synthesis, storage and release of acetylcholine by a noradrenergic pheochromocytoma cell line. Nature (Land) 268:349-351. GREENE, L A and REIN, G. (1977c). Release of [3H]Norepinepherine from a clonal line of pheochromocytoma cells (PC12) by nicotinic cholinergic stimulation. Brain Res. 138:521-528. 242 GROSS, R. A. and MACDONALD, R. L (1988). Barbiturates and nifedipine have different and selective effects on calcium currents of mouse DRG neurons in culture: A possible basis for differing clinical actions. Neurology 38:443-451. GUNDERSEN, C. B., KATZ, B. and MILEDI, R. (1982). The antagonism between botulinum toxin and calcium in motor nerve terminals. Proc. R. Soc. B. 216:369-376. GUNDERSON, C. B., UMBACH, J. A. and SCHWARTZ, B. E. (1988). Barbiturates depress currents through human brain calcium channels studied in Xenopus oocytes. J. Pharmacol. Exp. Ther. 247:824-829. GUROFF, G., DICKENS, G., END, D. and LONDOS, C. (1981). The action of adenosine analogs on PC12 cells. J. Neurochem. 37:1431-1439. GUROFF, G.‘ (1985). PC12 cells as a model of neuronal differentiation. In: Cell culture in the neurosciences. (Bottenstein, J. E. and Sato, G, eds). Plenum Press, New York. pp 245-272. HAGIWARA, S. and TAKASHIMA, K. (1967). Surface density of calcium ions and calcium spikes in the barnacle muscle fiber membrane]. Gen. Physiol. 50:583- 601. 243 HAGIWARA, S., FUKUDA, J. and EATON, D. (1974). Membrane currents carried by Ca, Sr and Ba in barnacle muscle during voltage clamp. J. Gen. Physiol. 63:564-578. HAGIWARA, S., MIYAZAKI, S., MOODY, W. and PATTAK, J. (1978). Blocking effects of barium and hydrogen ions on the potassium current during anomalous rectification in the starfish egg. J. Physiol. (Land) 279:167-185. HALEEN, S. J., STEFFEN, R. P., SIRCAR, I., MAJOR, T. C., TAYLOR, M. D., PUGSLEY, T. A. and WEISHARR, R. E. (1989). PD122860: A novel dihydropyridine with sodium channel stimulating and calcium channel blocking properties. J. Pharmacol. Exp. Ther. 250:22-30. HAMILL, O. P., MARTY, A. SAKMANN, B. and SIGWORTH, F. J. (1981). Improved patch-clamp techniques for high resolution current recording from cell and cell-free membrane patches. Pflfigers Arch. 391:85-100. HANSEN, J. C. (1990). Human exposure to metals through consumption of marine food: A case study of exceptionally high intake among Greenlanders. In: Heavy metals in the marine environment (Furness. A. W. and Rainbow, P. S., eds). CRC Press, New York. pp 227-243. 244 HARE, M. F. and ATCHISON, W. D. (1990). Differential effects of mercurials on synaptosomal membrane potentials. Neurosci. Abstr. 16:445. HARRIS, R. A. and BRUNO, P. (1985). Effects of ethanol and other intoxicant- anesthetics on voltage-dependent sodium channels of brain synaptosomes. J. Pharmacol. Exp. Ther. 232:401-406. HARRIS, K. M., KONGSAMUT, S. and MILLER, R. J. (1986). Protein kinase C mediated regulation of calcium channels in PC-12 pheochromocytoma cells. Biochem. Biophys. Res. Commun. 134: 1298- 1305. HESS, P. and TSIEN, R. w. (1984). Mechanism of ion permeation through Ca2+ channels. Nature (Land) 309:453-456. HESS, P., LANSMAN, J. B. and TSIEN, R. W. (1986). Calcium channel selectivity for divalent and monovalent cations. Voltage and concentration dependence of single channel current in ventricular heart cells. J. Gen. Physiol. 88:293-319. HEWE'IT, S. J. and ATCHISON, W. D. (1990). Calcium channel blocking actions of charged and uncharged mercurials in rat forebrain synaptosomes. Neurosci. Abstr. 16:512. 245 HILLE, B. (1971). The permeability of the sodium channel to organic cations in myelinated nerve. J. Gen. Physiol. 58:599-619. HILLE, B. (1977). Local anesthetics: Hydrophilic and hydrophobic pathways for the drug-receptor reaction. J. Gen. Physiol. 69:497-515. HILLE, B. (1984). Ionic channels of excitable membranes. Sinauer Associates Inc., Sunderland, MA. HIRNING, L D., FOX, A. P., MCCLESKEY, E. W., OLIVERA, B. M., THAYER, S. A., MILLER, R. J. and TSIEN, R. W. (1988). Dominant role of N-type Ca2+ channels in evoked release of norepinephrine from sympathetic neurons. Science 239:57-61. HODGKIN, A. L and HUXLEY, A. F. (1952). A quantative description of membrane current and its application to conduction and excitation in the nerve. J. Physiol. (Land) 117:500-544. HOF, R. P., RUEGG, U. T., HOF, A. and VOGEL, A. (1985). Stereoselectivity at the calcium channel: Opposite action of the enantiomers of a 1,4- dihydropryidine. J. Cardiovasc. Phann. 7 :689-693. 246 HOLZ, G. G., DUNLAP, K. and KREAM, R. M. (1988). Characterization of the electrically-evoked release of substance P from dorsal root ganglion neurones: Methods and dihydropyridine sensitivity. J. Neurosci. 8:463-471. HORWITZ, J. (1989). Muscarinic receptor stimulation increases inositol-phospholipid metabolism and inhibits cyclic AMP accumulation in PC12 cells. J. Neurochem. 53:197-204. HOSEY, M. M. and LAZDUNSKI. (1988). Calcium channels: Molecular pharmacology, structure and regulation. J. Membrane Biol. 104:81-105. HOSHI, T. and ALDRICH, R. W. (1988a). Voltage-dependent K" currents and underlying single K+ channels in pheochromocytoma cells. J. Gen. Physiol. 91:73-106. HOSHI, T. and ALDRICH, R. W. (1988b). Gating kinetics of four classes of voltage- dependent K+ channels in pheochromocytoma cells. J. Gen. Physiol. 91:107- 131. HOSS, W. and FORMANIAK, M. (1984). Calcium channel activity in rat brain synaptosomes: Effects of neuroleptics and other factors regulating phosphorylation and transmitter release. Neurochem. Res. 9:109-120. 247 HOWARD, B. D., MELEGA, W. P., KOIDE, M., BUTLER, C. M., ZHANG, M. and KHAW, M. (1986). Neurotransmitter release from PC12 cells. In: Calcium, neuronal function and neurotransmitter release. (Rahamimoff, R. and Katz, B, eds.) Martinus-Nijhoff Inc., Boston. pp. 279-297. HUBBARD, J. I., JONES, S. F. and LANDAU, E. M. (1968). On the mechanism by which calcium and magnesium affect the spontaneous release of transmitter from mammalian motor nerve terminals. J. PhysioL (Land) 194:353-380. HUNEEUS-COX, F., FERNANDEZ, H. L and SMITH, B. H. (1966). Effects of redox and sulfliydryl reagents on the bioelectric properties of the giant axon of the squid. Biophys. J. 6:675-689. IFUNE, C. K. and STEINBACH, J. H. (1990). Regulation of sodium currents and acetylcholine responses in PC12 cells. Brain Res. 506:243-248. INOUE, K. and KENIMER, J. G. (1988). Muscarinic stimulation of calcium influx and norepinepherine release in PC12 cells. J. Biol. Chem 263:8157-8161. INOUE, N., MATSUI, H. and HATANAKA, H. (1988). Nerve growth factor induces a Na+,K*-ATPase in a nerve cell line. J. Neurochem. 50:230-236. 248 JANIGRO, D., MACCAFERRI, G. and MELDOLESI, J. (1989). Calcium channels in undifferentiated PC 12 rat pheochromocytoma cells. FEBS Lett. 255:398-400. JUAN G, M. S. (1976). An electrophysiological study of the action of methylmercury chloride and mercuric chloride on the sciatic nerve-sartorius muscle preparation of the frog. Toxicol. Appl. Pharmacol. 35:339-348. JUMBLATI‘, J. E. and TISCHLER, A. S. (1982). Regulation of muscarinic ligand binding sites by nerve growth factor in PC12 pheochromocytoma cells. Nature (Land) 297:152-154. KALMAN, D., WONG, B., HORVAI, A. E., CLINE, M. J. and O’LAGUE, P. H. (1990). Nerve grth factor acts through CAMP-dependent protein kinase to increase the number of sodium channels in PC12 cells. Neuron 2:355-366. KATZ, B. and MILEDI, R. (1967a). The timing of calcium action during neuromuscular transmission. J. Physiol. (Land) 189:535-544. KATZ, B. and MILEDI, R. (1967b). A study of synaptic transmission in the absence of nerve impulses. J. Physiol. (Land) 192:407-436. 249 KAUPPINEN, R. A., KOMULAINEN, H. and TAIPALE, H. (1989). Cellular mechanisms underlying the increase in cytosolic free calcium concentration induced by methylmercury in cerebrocortical synaptosomes from guinea pig. J. Pharmacol. Exp. Ther. 248:1248-1254. KAMIYA, H., SWADA, S. and YAMAMOTO, C. (1988). Synthetic to -conotoxin blocks synaptic transmission in the hippocampus in vitro. Neurosci. Lett. 91:84- 88. KEITH, R. A., MANGANO, T. J. and SALAMA, A. I. (1989). Inhibition of N- methyl-D-aspartate- and kaniac acid-induced neurotransmitter release by to - conotoxin GVIA. Br. J. Pharmacol. 98:767-772. KENNEDY, M. B. (1989). Regulation of neuronal function by calcium. TINS 12:417- 419. KITA, H. and VAN DER KLOOT, W. (1973). Action of Co and Ni at the frog neuromuscular junction. Nature (Land) 245:52-53. KLAASSEN, C. D. (1985). Heavy metals and heavy metal antagonists. In: The pharmacological basis of therapeutics. (Gilman, A. G., Goodman, L. 8., 250 Randall, T. W., and Murad, F., eds.). MacMillan Publishing Co., New York, pp. 1611-1614. KN AUS, H., STRIESSNIG, J ., KOZA, A. and GLOSSMANN, H. (1987). Neurotoxic aminoglycoside antibiotics are potent inhibitors of [mI]-omega-conotoxin GVIA binding to guinea-pig cerebral cortex membranes. Naunyn- Schmiedeberg. Arch. Pharmacol. 336:583-586. KOBER, T. E. and COOPER, G. P. (1976). Lead competitively inhibits calcium-dependent synaptic transmission in the bullfi'og sympathetic ganglion. Nature (Land) 262:704-705. KOH(E, T. and TAKASHIMA, A. (1984). Clonal variability of PC12 pheochromocytoma cells with respect to catecholamine biosynthesis. J. Neurochem. 42:1472-1475. KOIKE, T. and TAKASHIMA, A. (1986). Cell cycle-dependent modulation of biosynthesis and stimulus-evoked release of catecholamines in PC12 pheochromocytoma cells. J. Neurochem. 46: 1493-1500. 251 KOMULAINEN, A. and BONDY, S. C. (1987). Increased free intrasynaptosomal Ca2+ by neurotoxic organometals: Distinctive mechanisms. T oxicoL Appl. Pharmacol. 88:77-86. KONISHI, T. (1985). Electrical exciatbility of motor nerve terminals in the mouse. J. Physiol. (Land). 366:411-421. KONGSAMUT, S. and MILLER, R. J. (1986). Nerve grth factor modulates the drug sensitivity of neurotransmitter release from PC12 cells. Proc. Natl. Acad Sci. (U.S.A.) 83:2243-2247. KOSTIAL, K. and VOUK, V. B. (1957). Lead ions and synaptic transmission in the superior cervical ganglion of the cat. Br. J. Pharmacol. Chemother. 12:219-222. KOSTYUK, P. G., SHUBA, Y. M. and SAVCHENKO, A. N. (1988). Three types of calcium channels in the membrane of mouse sensory neurones. Pflr’igers Arch. 411:661-669. KRESS, H. G, ECKHARDT-WALLASCH, E., TAS, P. W. L and KOSCHEL K (1987). Volatile anesthetics depress the depolarization-induced cytoplasmic calcium rise in PC12 cells. FEBS Lett. 221:28-32. _r ‘J‘L‘r-nl“ 252 KUROZUMI, K., MURAYAMA, T. and NOMURA, Y. (1990). Generation of inositol phosphates, cytosolic Ca2+ and secretion of noradrenaline in PC12 cells treated with glutamate. FEBS Lett. 270:225-228. LANG, B., VINCENT, A., MURRAY., N. M. F. and NEWSOM-DAVIS, J. (1989). Lambert-Eaton myasthenic syndrome immunogloban G inhibition of Ca2+ flux in tumor cells correlates with disease severity. Ann. Neurol. 25:265-271. LANSMAN, J. B., HESS, P. and TSIEN, R. W. (1986). Blockade of current through single calcium channels by Cdz”, Mg“ and Ca2+. Voltage and concentration dependence of calcium entry into the pore. J. Gen. Physiol. 88:321-347. LEE, K. S. and TSIEN, R. W. (1982). Reversal of current through calcium channels in dialysed single heart cells. Nature (Land) 297:498-501. LEE, K. S. and TSIEN, R. W. (1983). Mechanism of calcium channel blockade by verapamil, D-600, diltiazem and nitrendipine in single dialyzed heart cells. Nature (Land) 302:790-794. LEMOS, J. R. and NOWYCKY, M. C. (1989). Two types of calcium channels coexist in peptide-releasing vertebrate nerve terminals. Neuron 2:1419-1426. 253 LEONARD, J. P., NARGIOT, J ., SNUTCH, T. P., DAVIDSON, N. and LESTER, H. A. (1987). Ca channels induced in Xenopus oocytes by rat brain mRNA. J. Neurosci. 7 :875-881. LESLIE, S. W. (1987). Calcium channels. Interactions with ethanol and other sedative-hypnotic drugs. In: Recent developments in alcoholism. Volume 5. (Galanter, M., ed.). Plenum Publishing, New York. pp 289-302. LESLIE, S. W., BARR, E., CHANDLER, J. and FARRAR, R. P. (1983). Inhibition of fast- and slow-phase depolarization-dependent synaptosomal calcium uptake by ethanol. J. Pharmacol. Exp. Ther. 225:571-575. LESLIE, S. W., WOODWARD, J. J. and WILCOX, R. E. (1985). Correlation of rates of calcium entry and endogenous dopamine release in mouse striatal synaptosomes. Brain Res. 325:99-105. LEVESQUE, P. C. and ATCHISON, W. D. (1987). Interactions of mitochondrial inhibitors with methylmercury on spontaneous quantal release of acetylcholine. ToxicoL App. Pharmacol. 87:315-324. 254 LEVESQUE, P. C. and ATCHISON, W. D. (1988). Effect of alteration of nerve terminal Ca2+ regulation on increased spontaneous quantal release of acetylcholine by methylmercury. Toxicol. Appl. Pharmacol. 95:55-65. LEVESQUE, P. C. and ATCHISON, W. D. (1991). Disruption of brain mitochondrial calcium sequestration by methylmercury. J. Pharmacol. Exp. Ther. 256:236-242. LIM, N. F., NOWYCKY, M. C. and BOOKMAN, R. J. (1990). Direct measurement of exocytosis and calcium currents in single vertebrate nerve terminals Nature (Land) 334:449-451. LIN, Y. W., SUGIMORI, M., LLINAS, R. R., MCGUINNESS, T. L and GREENGARD, P. (19903). Effects of synapsin I and calcium/calmodulin- dependent protein kinase H on spontaneous neurotransmitter release in the squid giant synapse. Proc. Natl. Acad Sci. (U.S.A.). 87:8257-8261. LIN, Y. W., RUDY, B. and LLINAS, R. R. (1990b). Funnel-web spider venom and a toxin fraction block calcium current expressed from rat brain mRNA in Xenopus oocytes. Proc. Natl. Acad Sci. (U.S.A.). 87:4538-4542. 255 LLINAS, R. and YAROM, Y. (1981). Properties and distribution of ionic conductances generating electroresponsiveness of mammalian inferior olivary neurones in vitro. J. Physiol. (Land) 315:569-584. LLINAS, R., STEINBERG, I. Z. and WALTON, K. (1981). Presynaptic calcium currents in squid giant synapse. Biophys. J. 33:289-322. LLINAS, R., MCGUINNESS, T. L, LEONARD, C. S., SUGIMORI, M. and GREENGARD, P. (1985). Intraterminal injection of synapsin I or calcium/calmodulin—dependent protein kinase II alters neurotransmitter release at the squid gaint synapse. Proc. Natl. Acad Sci. USA 82:3035-3039. LLINAS, R., SUGUMORI, M., LIN, J. W. and CHERKSEY, B. (1989). Blocking and isolation of a calcium channel from neurons in mammals and cephalopods utilizing a toxin fraction (FI‘X) from funnel-web spider poison. Proc. Natl. Acad Sci. (U.S.A.) 86:1689-1693. LOWRY, O. H., ROSEBROUGH, N. J ., FARR, A. L. and RANDALL, R. J. (1951). Protein measurement with the Folin phenol reagent. J. Biol. Chem. 193:265-275. 256 LUKAS, R. J. (1990). Heterogeniety of high-affinity nicotinic [3H]-acetylcholine binding sites. J. Pharmacol. Exp. Ther. 253:51-57. LUNDY, P. M., FREW, R., FULLER T. W. and HAMILTON, M. G. (1991). Pharmacological evidence for an w‘CODOtOXIn, dihydropyridine-insensitive neuronal Ca2+ channel. Eur. J. Pharmacol. 206:61-68. MAGGI, C. A., PATACCHINI, R., SANTICIOLI, P., LIPPE, I. T., GIULIAMI, S., GEPPETI‘I, P., DEL BIANCO, E., SELLERI, S. and MELI, A. (1988). The effect of omega conotoxin GVIA, a peptide modulator of the N-type voltage sensitive calcium channels, on motor responses produced by activation of efferent and sensory nerves in mammalian smooth muscle. Naunyn- Schmiedeberg Arch. Pharmacol. 338:107-113. MAGOS, L. (1975). Mercury and mercurials. Br. Med Bull. 33:241-245. MALLART, A. (1985). Electric current flow inside perineurial sheaths of mouse motor nerves. J. Physiol. (Land) 368:565-575. MALLART, A. and BRIGANT, J. L (1982). Presynaptic currents in mouse motor endings. J. Physiol. (Land) 333:619-636. 257 MANALIS, R. S. and COOPER, G. P. (1973). Presynaptic and postsynaptic effects of lead at the neuromuscular junction. Nature (Land) 243:354-356. MANGAN OS, T. J .,PATEL, J ., SALAMA, A. I. and KEITH, R. A. (1991). Inhibition of K+-evoked [3H]D-aspartate release and neuronal calcium influx by verapamil, diltiazem and dextromethorphan: evidence for non-L/non-N voltage-sensitive calcium channels. MARKS, S. S., WATSON, D. L, CARPENTER, C. L, MESSING, R. O. and GREENBERG, D. A. (1989). Comparative effects of chronic exposure to ethanol and calcium channel antagonists in calcium channel antagonist receptors in cultured neural (PC12) cells. J. Neurochem. 53:168-172. MARQUEZE, B., MARTIN—MOUTOT, N., LLEVESQUE, C. and COURAND, F. (1988). Characterization of the cti-conotoxin binding molecule in rat brain synaptosomes and cultured neurons. Mol. Pharmacol. 34:87-90. MARSHALL, I. G. (1970). Studies on the blocking action of 2-(4-phenyl piperidino) cyclohexanol (AH5183). Br. J. Pharmacol. 38:503-516. MARTINEZ-SERRANO, A., BOGONEZ, E., VITORICA, J. and SATRUSTEGUI, J. (1989). Reduction of K+-stimulated 4sCa influx in synaptosomes with age 258 involves inactivating and non-inactivating calcium channels and is correlated with temporal modifications in protein dephosphorylation. J. Neurochem. 52:576-584. MASSIEU, L and TAPIA, R. (1988). Relationship of dihydropyridine binding sites with calcium-dependent neurotransmitter release in synaptosomes. J. Neurochem. 51:1184-1189. MATSUOKA, l., MIZUNO, N. and KURIHARA, K. (1989). Cholenergic differentiation of clonal rat pheochromocytoma cells (PC12) induced by retanoic acid: Increase of choline acetyltransferase activity and decrease of tyrosine hydroxylase activity. Brain Res. 502:53-60. MCARDLE, J ., AN GAUT-PETIT, D., MALLART, A., FAILLE, L and BRIGANT, J. L (1981). Advantage of the triangularis sterni muscle of the mouse for investigations of synaptic phenomena. J. Neurosci. Meth. 4: 109-113. MCHUGH, E. M. and MCGEE, R. (1986). Direct anesthetic-like effects of forskolin on the nicotinic acetylcholine receptors of PC12 cells. J. Biol. Chem. 261:3103- 3106. 259 MCGEE, R. and LEIPE, B. (1984). Acute elevation of cyclic AMP does not alter the ion-conducting properties of the neuronal nicotinic acetylcholine receptor of PC12 cells. Mol. Pharmacol. 26:51-56. MCKAY, S. J., REYNOLDS, J. N. and RACZ, W. J. (1986). Effects of mercury compounds on the spontaneous and potassium-evoked release of [3H]- dopamine from mouse striatial slices. Can. J. Physiol. Pharmacol. 64: 1507-1514. MELDOLESI, J ., SCHEER, H., MADEDDU, L and WANKE, E. (1986). Mechanism of action of a -latrotoxin: the presynaptic stimulatory toxin of the black widow spider venom. Trends in Pharmacological Sci. 7: 151-155. MELDOLESI, J ., VOLPE, P. and POZZAN, T. (1988). The intracellular distribution of calcium. TINS 11:449-452. MELEGA, W. P. and HOWARD, B. D. (1981) Choline and acetylcholine metabolism in PC12 secretory cells. Biochem. 20:4477-4483. MELEGA, W. P. and HOWARD, B. D. (1984). Biochemical evidence that vesicles are the source of the acetylcholine released from stimulated PC12 cells. Proc. Natl. Acad Sci. (U.S.A.) 81:6535-6538. 260 MESSING, R. O., STEVENS, A. M., KIYASU, E. and SNEADE, A. B. (1989). Nicotinic and muscarinic agonists stimulate rapid protein kinase C translocation in PC12 cells. J. Neurosci. 9:507-512. MICHEL, A. D., STEFANICH, E. and WHITING, R. L (1989). PC12 phaeochromocytoma cells contain an atypical muscarinic receptor binding site. Br. J. Pharmacol. 97:914-920. MILLER, F. M. (1984). Chemistry, Structure and dynamics. M‘Graw-Hill Book Company, St. Louis. pp 177, 748. MILLER, R. J. (1987). Multiple calcium channels and neuronal function. Science 235:46-52. MINTZ, J. M., VENEMA, V. J., ADAMS, M. B. and BEAN, B. P. (1990). T h e funnel-web speder toxin in -Aga-IIIA blocks N- and L-type calcium channels in neurons and cardiac muscle. Neurosci Abstr. 16:956. MIYAMATO, M. D. (1983). Hg2+ causes neurotoxicity at an intracellular site following entry through Na and Ca channels. Brain Res. 267:375-379. MIZRACHI, Y., NARANJO, J. R., LEVI, B., POLLARD, H. B. and LELKES, P. I. (1990). PC12 cells differentiate into chromaffin-cell like phenotype in 261 coculture with adrenal medullary endothelial cells. Proc. Natl. Acad Sci. (U.S.A.) 87:6161-6165. NACHSHEN, D. A. (1984). Selectivity of the Ca binding site in synaptosome Ca channels. Inhibition by multivalent metal cations. J. Gen. Physiol. 83:941-967. NACHSHEN, D. A. (1985). The early time course of potassium-simulated calcium uptake in presynaptic terminals isolated from rat brain. J. Physiol. (Land) 361:25 1-268. NACHSHEN, D. A. and BLAUSTEIN, M. P. (1979). The effects of some organic "calcium antagonists" on calcium influx in presynaptic nerve terminals. Mol. Pharmacol. 16:579-586. NACHSHEN, D. A. and BLAUSTEIN, M. P. (1980). Some properties of potassium-stimulated calcium influx in presynaptic nerve endings. J. Gen. Physiol. 76:709-728. NACHSHEN, D. A. and BLAUSTEIN, M. P. (1982). Influx of calcium, strontium, and barium in presynaptic nerve endings. J. Gen. Physiol. 79: 1065-1087. 262 NACHSHEN, D. A., SANCHEZ-ARMASS, S. and WEINSTEIN, A. M. (1986). The regulation of cytosolic calcium in rat brain synaptosomes by sodium-dependent calcium efflux. J. Physiol. (Land) 381:17-28. NAKAZATO, Y., ASANO, T. and OHGA, A. (1979). The in vitro effect of mercury compounds of noradrenaline output from guinea pig vas deferens. Toxicol. Appl. Pharmacol. 48:171-177. NARAHASHI, T. (1980). Nerve membrane as a target of environmental neurotoxicants. In: Experimental and clinical neurotoxicology. (Spencer, P. S., Schaumburg, H. H., eds.). Williams and Wilkins, Inc. Baltimore. pp. 225-238. NARAHASIH, T., MOORE, J. W. and SCOTT, W. R. (1964). Tetrodotoxin blockage of sodium conductance increase in lobster giant axons. J Gen. Physiol. 47:965-974. NARANJO, J. R., MOCCHETTI, I., SCHWARTZ, J. P. and COSTA, E. (1986). Permissive effect of dexamethasone on the increase of proenkephalin mRNA induced by depolarization of chromaffin cells. Proc. Natl. Acad Sci. (U.S.A.) 83:1513-1517. 263 NELSON, M. T., ROUDNA, M. and BAMBERG, E. (1983). Single channel K+- current measurements from brain synaptosomes in lipid bilayers. Am. J. Physiol. 245:C151-C156. NELSON, M. T. (1985). Effects of blockers and permeant ions on the properties of single calcium channels from brain. In: proteins of excitable membranes. (Hille, B. and Fambrough, D., eds.). John Wiley and Sons, New York. pp 215- 230. NELSON, M. T. (1986). Interactions of divalent cations with single calcium channels from rat brain synaptosomes. J. Gen. Physiol. 87:201-222. NILSON, R. and VOLLE, R. L (1976). Blockade by cadmium (CdH) at the frog neuromuscular junction. Fed Proc. 35:696. NISHIZAWA, Y., SEAMON, K. B., DALY, J. W. and ARONSTAM, R. S. (1990). Effects of forskolin and analogues on nicotinic receptor-mediated sodium flux, voltage-dependent calcium flux and voltage-dependent rubidium efflux in pheochromocytoma cells. Cellular Molec. Neurobiol. 10:351-368. 264 NORONHA-BLOB, L, MARSHALL, R. P., KINNIER, W. J. and U’PRICHARD, D. C. (1986). Pharmacological profile of adensoine A2 receptor in PC12 cells. Life Sci. 39:1059-1067. NOWYCKY, M. C. (1991). Kinetic properties of L and Nt-type Ca2+ channels and their ability to promote exocytosis in neurohypophysial terminals. Biophys. J. 59:276a. N OWYCKY, M. C., FOX, A. P. and TSIEN, R. W. (1985a). Three types of neuronal calcium channels with different calcium agonist sensitivity. Nature (Land) 316:440-443. NOWYCKY, M. C., FOX, A. P. and TSIEN, R. W. (1985b). Long opening mode of gating of neuronal calcium channels and its promotion by the dihydropyridine calcium channel agonist Bay K 8644. Proc. Natl. Acad Sci. (U.S.A.) 82:2178-2182. OGATA, N. and NARAHASHI, T. (1990). Potent blocking action of chlorpromazine on two types of calcium channels in cultured neuroblastoma cells. J. Pharmacol. Exp. Ther. 252:1142-1149. 265 OLIVERA, B. M., GRAY, W. R., ZEIKUS, R., MCINTOSH, J. M., VARGA, J., RIVIER, J., DESANT OS, V. and CRUZ, L J. (1985). Peptide neurotoxicns from fish-hunting cone snails. Science (Wash. D. C.) 230:1338-1343. OLIVERA, B. M., CRUZ, L J., DESANTOS, V., LECI-IEMINANT, G. W., GRIFFIN, D., ZEIKUS, R., MCINTOSH, J. M., GALYEAN, R., VARGA, J ., GRAY, W. R. and RIVIER, J. (1987). Neuronal calcium channel antagonists. Discrimination between calcium channel subtypes using to- conotoxin from Conus magnus venom. Biochemistry 26:2086-2090. PARNAS, H., DUDEI, J. and PARNAS, I. (1986). Neurotransmitter release and its facilitation in crayfish. VII. Another voltage dependent process beside Ca entry controls the time course of phasic release. Pflugers Arch. 406:121-130. PARNAS, H., HOVAV, G. and PARNAS, I. (1989). Effect of Ca2+ diffusion on the time course of neurotransmitter release. Biophys. J. 55:859-874. PASSOW, H. and ROTHSTEIN, A. (1960). The binding of mercury by the yeast cell in relation to changes in permeability. J. Gen. Physiol. 43:621-633. PENNER, R. and DREYER. F. (1986). Two different presynaptic calcium currents in mouse motor nerve terminals. Pflu'gers Arch. 406: 190-197. 266 PERNEY, T. M., HIRNING, L D., LEEMAN, S. E. and MILLER, R. J. (1986). Multiple calcium channels mediate neurotransmitter release from peripheral neurons. Proc. Natl. Acad Sci. (U.SA.) 83:6656-6659. PICKETT, J. B. and BORNSTEIN, J. C. (1978). Lead competitively inhibits calcium-dependent evoked release at rat neuromuscular junctions. Neurology 28:335. PICKETT, J. B. and BORNSTEIN, J. C. (1984). Some effects of lead at mammalian neuromuscular junction. Am. J. Physiol. 246:C271-C276. PLUMMER, M. R., LOGOTHETIS, D. E. and HESS, P. (1989). Elememtary properties and pharmacological sensitivities of calcium charmels in mammalian peripheral neurons. Neuron 2:1453-1463. POLLOCK, J. D., KREMPIN, M. and RUDY, B. (1990). Differential effects of N GF , FGF, ECF, cAMP, and dexamethasone on neurite outgrowth and sodium channel expression in PC12 cells. J. Neurosci. 8:2626-2637. POZZAN, T., GATTI, G., DOZIO, N., VICENTINI, L M. and MELDOLESI, J. (1984). Ca2*-dependent and independent release of neurotransmitters from PC12 cells: A role for protein kinase C activation? J. Cell Biol. 99:628-638. 267 PUMPLIN, D. W., REESE, T. S. and LLINAS, R. (1981). Are the presynaptic membrane particles the calcium channels? Proc. Natl. Acad Sci. (U.S.A.) 78:7210-7213. PUN, R. Y. K. and BEHBEHANI, M. M. (1989). A rapidly inactivating Ca2+ -dependent K+ current in phenochromocytoma cells (PC12) of the rat. Eur. J. PhysioL 415:425432. QUANDT, F. N., KATO, E. and NARAHASHI, T. (1982). Effects of methylmercury on electrical responses of neuroblastoma cells. Neurotoxicology 3:205-220. RABE, C. S., SCHNEIDER, J. and MCGEE, R. (1982). Enhancement of depolarization-dependent neurosecretion from PC 12 cells by forskolin-induced elevation of cyclic AMP. J. Cyclic Nucl. Res. 8:371-384. RABE, C. S. and MCGEE, R. (1983). Regulation of depolarization-dependent release of neurotransmitters by adenosine: Cyclic AMP-dependent enhancement of release from PC12 cells. J. Neurochem. 41:1623-1634. RABE, C. S., DELORME, E. and WEIGHT, F. F. (1987). Muscarine-stimulated neurotransmitter release from PC12 cells. J. Pharmacol. Exp. Ther. 243:534- 541. 268 RABE, C. S. and WEIGHT, F. F. (1988). Effects of ethanol on neurotransmitter release and intracellular free calcium in PC12 cells. J. Pharmacol. Exp. Ther. 243:417-422. RAMPE, D., JANIS, R. A. and TRIGGLE, D. J. (1984). Bay K 8644, a 1,4-dihydropyridine Ca2+ channel activator: Dissociation of binding and fimctional effects in brain synaptosomes. J. Neurochem. 43:1688-1692. RANE, s. G., HOLTZ, G. G. and DUNLAP, K (1987). Dihydropyridine inhibition of neuronal calcium current and substance P release. Pfliigers Arch. 409:361-366. REED, LR. and D. ENGLAND. (1986). The effect of nerve grth factor on the development of sodium channels in PC12 cells. Biochem. Cell Biol. 64:1153-1159. REUHL, K. R. (1987). Neuropathology of organometallic compounds: Review of selected literature. In: N eurotoxicants and neurobiological function: Effects of organoheavy metals. (Tilson, H. A. and Sparber, S. B., eds.) John Wiley and Sons, Inc., New York. pp 117-136. rm 'fflp 269 REYNOLDS, I. J ., WAGNER, J. A., SNYDER, S. H., THAYER, S. A., OLIVERA, B. M. and MILLER, R. J. (1986). Brain voltage-sensitive calcium channels - differentiated by Q'COHOLOXIH fraction GVIA. Proc. Natl. Acad Sci. (U.S.A.) 83:8804-8807. RIDLEY, W. P. DIZIKES, L. J. and WOOD, J. M. (1977). Biomethylation of toxic elements in the environment. Science 197:329-332. RIVIER, J ., GAIfiEAN, R., GRAY, W. R., AZIMI-ZONOOZ, A., MCINTOSH, J. M., CRUZ, L J. and OLIVERA, B. M. (1987). Neuronal calcium channel inhibitors. Synthesis of to -conotoxin 45Ca uptake by synaptosomes. J. Biol. Chem. 262:1194-1198. ROBINSON, D. and MCGEE, R. (1985). Agonist-induced regulation of the neuronal nicotinic acetylcholine receptor of PC12 cells. Mol. Pharmacol. 27:409-417. ROBITAILLE, R., ADLER, E. M. and CHARLTON, M. P. (1990). Strategic location of calcium channels at transmitter release sites of frog neuromuscular synapses. Neuron 5:733-739. 270 ROGAWSKI, M. A., PIENIEK, M., SUZUKI, S. and FFRENCH-MULLEN, J. M. H. (1988). Phencyclidine selectively blocks the sustained voltage-dependent potassium conductance in PC12 cells. Brain Res. 456:38-48. ROSKOSKI, R. and ROSKOSKI, L. M. (1989). Adenosine receptor activation and the regulation of tyrosine hydroxylase activity in PC12 and PC18 cells. J. Neurochem. 53:1934-1940. ROMEY, G. and LAZDUNSKI, M. (1982). Lipid-soluble toxins thought to be specific for Na“ channels block Ca2+ channels in neuronal cells. Science 297:79-80. RUDY, B., KTRSCHENBAUM, B., RUBENSTEIN, A. and GREENE, L A. (1987). Nerve growth factor increases the number of functional Na channels and induces TTX-resistant Na channels in PC12 pheochromocytoma cells. J. Neurosci. 7 :1613-1625. RUSTAM, H., VON BURG, R., AMIN-ZAKI, L and EL HASSANI, S. (1975). Evidence for a neuromuscular disorder in methylmercury poisioning. Clinical and electrophysiological findings in moderate to severe cases. Arch. Environ. Health 30:190-195. 271 SAGER, P. R., DOHERTY, R. A. and OLMS'IED, J. B. (1982). Interaction of methylmercury with microtubules in cultured cells and in vitro. Exp. Cell Res. 146: 127-137. SANDBERG, K., BERRY, C. J. and ROGERS, T. B. (1989a). Studies on the intoxication pathway of tetanus toxin in the rat pheochromocytoma (PC12) cell line. Binding, internalization, and inhibition of acetylcholine release. J. Biol. Chem. 264:5679-5686. SANDBERG, K., BERRY, C. J. and ROGERS, T. B. (1989b). A role for cGMP during tetanus toxin blockade of acetylcholine release in the rat pheochromocytoma (PC12) cell line. J. Neurosci. 9:3946-3954. SCATCHARD, G. (1949). The attractions of proteins for small molecules and ions. Ann. NY. Acad Sci. 51:660-672. SCHMID-ANTOMARCHI, H., HUGUES, M. and LAZDUNSKI, M. (1986) Properties of the apamin-sensitive Ca2+-activated K+ channel in PC12 pheochromocytoma cells which hyper-produce the apamin receptor. J. Biol. Chem. 261:8633-8637. . 9 272 SCHUBERT, D. and KLIER, F. G. (1977). Storage and release of acetylcholine by a clonal cell line. Proc. Natl. Acad Sci. (U.S.A.) 74:5184-5188. SCHUBERT, D., HEINEMANN, S. and KIDOKORO, Y. (1977). Cholenergic metabolism and synapse formation by a rat nerve cell line. Proc. Natl. Acad Sci. (U.S.A.) 74:2579-2583. SEABROOK, G. R. and ADAMS D. J. (1989). Inhibition of neurally-evoked transmitter release by calcium channel antagonists in rat parasympathetic ganglia. Br. J. Pharmacol. 97:1125-1136. SEEGAL, R. F., BROSCH, K., BUSH, B., RI'I‘Z, M. and SHAIN, W. (1989). Effects of Arochlor 1254 on dopamine and norepinephrine concentrations in pheochromocytoma (PC-12) cells. Neurotoxicology 10:757-764. SHAFER, T. J. and ATCHISON, W. D. (1989). Block of 45Ca uptake into synaptosomes by methylmercury: Ca2+- and N a+-dependence. J. Pharmacol Exp. Ther. 248:696-702. SHAFER, T. J. and ATCHISON, W. D. (1991). Methylmercury blocks of N- and L- type Ca2+ channels in nerve growth factor-differentiated pheochromocytoma cells. J. Pharrhacol. Exp. Ther. 258:697-702. 273 SHAFER, T. J., CONTRERAS, M. L and ATCHISON, W. D. (1990). Characterization of interactions of methylmercury with Ca2+ channels in synaptosomes and pheochromocytoma cells: Radiotracer flux and binding studies. Mol. Pharmacol. 38:102-113. SHEPHARD, G. M. (1990). Introduction to synaptic circuits. In: The synaptic organization of the brain. (Shepherd, G. M., ed.). Oxford University Press, New York. p 5. SHER, E., PANDIELLA, A. and CLEMENTI, F. (1988). to-Conotoxin binding and effects on calcium channel function in human neuroblastoma and rat pheochromocytoma cell lines. FEBS Lett. 235:178-182. SIHRA, T. S., WANG, J. K., GORELICK, F. S. and GREENGARD, P. (1989). Translocation of synapsin I in response to depolarization of isolated nerve terminals. Proc. Natl. Acad Sci. (U.S.A.) 86:8108-8112. SHRAGER, P. (1977). Slow sodium inactivation in nerve after exposure to sulfhydryl blocking reagents. J. Gen. Physiol. 69:183-202. SHRIVASTAV, B. B., BRODWICK, M. S. and NARAHASHI, T. (1976). Methylmercury: Effects on electrical properties of squid axon membranes. Life Sci. 18:1077-1082. 274 SILBERGELD, E. K., FALES, J. T. and GOLDBERG, A. M. (1974a). Evidence for a junctional effect of lead on neuromuscular function. Nature (Land) 247:49-50. SILBERGELD, E. K., FALES, J. T. and GOLDBERG, A. M. (1974b). The effects of inorganic lead on the neuromuscular junction. Neuropharmacal. 13:795-801. SKATTEBOL, A. and RABIN, R. A. (1987). Effects of ethanol on 45Ca2+ uptake in synaptosomes and PC12 cells. Biochem. Pharmacol. 36:2227-2229. SMITH, S. J. and AUGUSTINE, G. J. (1988). Calcium ions, active zones and synaptic transmitter release. TINS 11:458-464. SNUTCH, T. P., LEONARD, J. P., GILBERT, M. M., LESTER, H. A. and DAVIDSON, N. (1990). Rat brain expressis a heterogeneous family of calcium channels. Proc. Natl. Acad Sci. (U.S.A.). 87:3391-3395. SOMJEN, G. G., HERMAN, S. P., KLEIN, R., BRUBAKER, P. E., BRINER, W. H., GOODRICH, J. K., KRIGMAN, M. R.,and MESEMAN, J. K. (1973). The uptake of methylmercury (”Hg) in different tissues rleated to its neurotoxic effects. J. Pharmacol. Exp. Ther. 187:602-611. 275 STALLCUP, W. B. (1979). Sodium and calcium fluxes in a clonal nerve cell line. J. Physiol. (Land) 286:525-540. STARR, T. V., PRYSTAY, W. A. and SNUTCH, T. P. (1990). Primary structure and expression of class-A calcium channel. J. Neurosci. 16:1174. STEELE, R. G. D. and TORRIE, J. H. (1980). Principles and Procedures of Statistics. McGraw-Hill. New York. pp. 99-160. STEINWALL, O. and KLATZO, I. (1966). Selective vulnerability of the blood brain barrier in chemically induced lesions. Acta Neuropathol. Exp. Neurol. 25:542- 559. STEINWALL O. and OLSSON, Y. (1969). Selective vulnerability of the blood brain barrier in chemically induced lesions. Acta Neurol. Scand 45:351-361. STREIT, J. and LUX, H. D. (1987). Voltage dependent calcium currents in PC12 growth cones and cells during N GF-induced cell growth. Pflu'gers Arch. 408:634-641. STREIT, J. and LUX, H. D. (1989). Distribution of calcium currents in sprouting PC12 cells. J. Neurosci. 9:4190-4199. 276 STREIT, J. and LUX, H. D. (1990). Calcium current inactivation during nerve- growth-factor-induced differentiation of PC12 cells. Pfliigers Arch. 416:368-374. STRICHARTZ, G. R. (1973). The inhibition of sodium currents in myelinated nerve by quaternary derivatives of lidocaine. J. Gen. Physiol. 62:37-57. SUSZKIW, J. B. and O’LEARY, M. E. (1982). Differential labeling of depot and active acetylcholine pools in nondepolarized and potassium-depolarized rat brain synaptosomes. J. Neurochem. 38:1668-1675. SUSZKIW, J. B. and O’LEARY, M. E. (1983). Temporal characteristics of potassium-stimulated acetlycholing release and inactivation of calcium influx in rat brain synaptosomes. J. Neurochem. 41:868-873. SUSZKIW, J. B., TOTH, G., MURAWSKY, M. M. and COOPER, G. P. (1985). Effects of Pb2+ and Cd2+ on acetylcholine release and Ca movements in synaptosomes and subcellular fractions from rat brain and Torpedo electric organ. Brain Res. 323:34-46. SUSZKIW, J. B., O’LEARY, M. E., MURAWSKY, M. M. and WANG, T. (1986). Presynaptic calcium channels in rat cortical synaptosomes: Fast-kinetics of 277 phasic calcium channel inactivation, and relationship to nitrendipine receptors. J. Neurosci. 6: 1349-1357. SUSZKIW, J. B., MURAWSKY, M. M. and FORTNER, R. C. (1987). Heterogeneity of presynaptic calcium channels revealed by species differences in the sensitivity of synaptosomal 45Ca entry to to -conotoxin. Biochem. Biophys. Res. Commun. 145:1283-1286. SUSZKIW, J. B., MURAWSKY, M. M. and SHI, M. (1989). Further characterization of phasic calcium influx in rat cerebrocortical synaptosomes: Inferences regarding the calcium channel type(s) in nerve endings. J. Neurochem. 52:1260- 1269. SWANDULLA, D. and ARMSTROMG, C. M. (1987). Fast-deactivating calcium channels in chick sensory neurons. J. Gen. PhysioL 92:197-218. SWANDULLA, D., CARBONE, E. and LUX, H. D. (1991). Do calcium channel classifications account for neuronal calcium channel diversity. TINS 14:46-51. TAKAHASHI, M., OHIZUMI, Y. and YASUMOTO, T. (1982). Maitotoxin, a Ca2+ channel activator candidate. J. Biol. Chem. 257:7287-7289. 278 TAKAHASHI, M. and OGURA, A. (1983). Dihydropyridines as potent calcium channel blockers in neuronal cells. FEBS Let. 152:191-194. TAKAHASHI, M., TATSUMI, M., OHIZUMI, Y. and YASUMOTO, T. (1983) Ca2+ channel activating function of maitotoxin, the most potent marine toxin known, in clonal rat pheochromocytoma cells.J. Biol. Chem. 258:10944-10949. TAKAHASHI, M., TSUKUI, H. and HATANAKA, H. (1985). Neuronal differentiation of Ca2+ channel by nerve growth factor. Brain Res. 341:381-384. TAKASHIMA, A. and KENIMER, J. G. (1989). Muscarinic-stimulated norepinephrine release and phosphoinositide hydrolysis in PC12 cells are independent events. J. Biol. Chem. 264:10654-10659. TAKENOSHITA, M. and STEINBACH, J. H. (1991). Halothane blocks low-voltage- activated calcium current in rat sensory neurons. J. Neurosci. 11:1404-1412. TAKEUCHI, T., NORIKAWA, N., MATSUMOTO, H. and SHRIAISHI. Y. (1962). Pathological study of minamata disease in Japan. Acta Neuropathol. 2:40-57. 279 TAKEUCHI, T., MATSUMOTO, H., SASHI, M., KAMBARA, T., SHIRAISHI, Y., HIRATA, Y., NOBUKIRO, M. and ITO, H. (1968). Pathology of Minimata disease. Kumamato Med J. 34:521-524. TOLL, L (1982). Calcium antagonists; high-affinity binding and inhibition of calcium transport in a clonal cell line. J. Biol. Chem 257: 13189-13192. TRAXINGER, D. L. and ATCHISON, W. D. (1987a). Comparative effects of divalent cations on the methylmercury-induced alterations of acetylcholine release. J. Pharmacol. Exp. Ther. 240:451-459. TRAXINGER, D. L and ATCHISON, W. D. (1987b). Reversal of methylmercury- induced block of nerve-evoked release of acetylcholine at the neuromuscular junction. Toxicol. Appl. Pharmacol. 90:23-33. TSIEN, R. W., LIPSCOMBE, D., MADISON, D. V., BLEY, K. R. and FOX, AP. (1988). Multiple types of neuronal calcium channels and their selective modulation. Trends in Neurosci. 10:431-438. TUOMISTO, J. and KOMULAINEN, H. (1983). Release and inhibition of uptake of 5-hydroxytryptamine in blood platelets in vitro by copper and methyl mercury. Acta Pharmacol. Et Toxicol. 52:292-297. 280 TURNER, T. J. and GOLDIN, S. M. (1985). Calcium channels: Identification and pharmacological characterization. High affinity blockade by organic Ca channel blockers. J. Neurosci. 5:841-849. TWOMBLY, D. A, YOSHII, M. and NARAHASHI, T. (1988). Mechanisms of calcium channel block by phenytoin. J. Pharmacol. Etp. Ther. 246:189-195. TWOMBLY, D. A, HERMAN, M. D., KYE, c. H. and NARAHASHI, T. (1990). Ethanol effects on two types of voltage-dependent calcium channels. J. Pharmacol. Exp. Ther. 254:1029-1037. UEHARA, A. and HUME, J. R. (1985). Interactions of organic calcium channel antagonists with calcium channels in single frog atrial cells. J. Gen. Physiol. 85:621-647. USOWICZ, M. M., PORZIG, H. BECKER, C. and REUTER, H. (1990). Differential expression by nerve grth factor of two types of Ca2+ channels in rat phaeochromocytoma cell lines. J. Physiol. 426:95-116. 281 VENEMA, V. J. and ADAMS, M. E. (1990). New to-Agatoxin calcium channel antagonists from funnel web spider venom active in the avain and mammalian brain. Neurosci Abstr. 16:956. VIANA, G. B., DAVIS, L H. and KAUFFMANN, F. C. (1988). Effects of organophosphates and nerve grth factor on muscarinic receptor binding number in rat pheochromocyotoma PC12 cells. Toxicol. Appl. Pharmacol 93:257-266. VICENTINI, L M. and MELDOLESI, J. (1984). Alpha-latrotoxin of black widow spider venom binds to a specific receptor coupled to phosphoinositide breakdown in PC12 cells. Biochem. Biophysical. Res. Comm. 121:538-544. VICENTINI, L M., AMBROSINI, A., DI VIRGILIO, F., POZZAN, T. and MELDOLESI, J. (1985). Muscarinic receptor-induced phosphoinositide hydrolysis at resting cytosolic Ca2+ concentration in PC12 cells. J. Cell Biol. 100: 1330- 1333. WEAKLY, J. N. (1973). The action of cobalt ions on neuromuscular transmission in the frog. J. Physiol. (Land) 234:597-612. 282 WERZ, MA. and MACDONALD, R. L (1985). Barbiturates decrease voltage- dependent calcium conductance of mouse neurons in dissociated cell culture. Mol. Pharmacol. 28:269-277. WESSLER, 1., DOOLEY, D. J., WERHAND, J. and SHHLEMMER, F. (1990). Differential effects of calcium channel antagonists (on -conotoxin GVIA, nifedipine, verapamil) on the electrically-evoked release of[3H] acetlycholine from the myenteric plexus, phrenic nerve and neocortex of rats. Naunyn- Schmiedeburg Arch. Pharmacol. 341:288-294. WILLIAMS, M., ABREU, M., JARVIS, M. F. and NORONHA-BIDB, L (1987). Characterization of adenosine receptors in the PC12 pheochromocytoma cell line using radioligand binding: Evidence for A2 selectivity. J. Neurochem. 48:498-502. WOOD, J. M., KENNEDY, F. S. and ROSEN, C. G. (1968). Synthesis of methly- mercury compounds by extracts of a methanogenic bacterium. Nature (Land) 220:173-174. YATANI, A. and BROWN, A. M. (1985) The calcium channel blocker nitrendipine blocks sodium channels in neonatal rat cardiac myocytes. Circulation Res. 56:868-875. 283 YOSHINO, Y., MOZAI, T. and NAKAO, K. (1966). Biochemical changes in the brain of rats poisoned with an alkyl mercuric compound, with special reference to the inhibition of protein synthesis in brain cortex slice. J. Neurochem. 13:1223-1230. YOSHIZAWA, T., KIMURA, S., KANAZAWA, 1., UCHIYAMA, Y., YANAGISAWA, M. and Y. MASAKI. (1989). Endothelin localizes in the dorsal horn and acts on the spinal neurones: Possible involvement of dihydropyridine-sensitive calcium channels and substance P release. Neurosci. Lett. 102:179-184. STRTE UNIV. MICHIGAN "WWW" ”II 13112 L 13 ll llllllllllll RnRIEs WW 93009 39472