IllllllllllilllllllllllilllllilllillllHIHIIHIH‘IHHHIHI 3 1293 020589 This is to certify that the dissertation entitled DOPAMINE RECEPTOR-MEDIATED REGULATION OF TUBEROINFUNDIBULAR DOPAMINERGIC NEURONS presented by Robert Alan Durham has been accepted towards fulfillment of the requirements for PhD. degree in Pharmacology and Toxicology M/A/L/ Major ”pater/£50 Date //'23';7 ........J LEBRASY Michigan waste Univereity nt- PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 11/00 chincmmotnpsspu DOPAMINE RECEPTOR-MEDIATED REGULATION OF TUBEROINFUNDIBULAR DOPAMINERGIC NEURONS By Robert Alan Durham 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 Neuroscience Program 1999 ABSTRACT DOPAMINE RECEPTOR-MEDIATED REGULATION OF TUBEROINFUNDIBULAR DOPAMINERGIC NEURONS By Robert Alan Durham Although it has long been recognized that neurotransmission in most dopaminergic (DA) neurons in the brain is subject to DA autoreceptor-mediated negative feedback, tuberoinfundibular (TDDA neurons in the mediobasal hypothalamus are not responsive to this mode of DA receptor-mediated regulation. This assertion is supported by numerous studies in which DA agonists and antagonists that affect autoreceptor-mediated processes in other DA neurons have no effect on TIDA neurons. The development of DA drugs with greater selectivity for the various DA receptors has revealed that TIDA neurons are subject to DA receptor-mediated regulation, but that this is different in many important ways from that mediated by autoreceptors in other DA neurons. In contrast to autoreceptor-mediated inhibition of neuronal activity, the selective DA D2/3 agonist quinelorane stimulates TIDA neurons. Studies described in this dissertation have utilized a combination of molecular, neurochemical, neuroanatomical and pharmacological experimental approaches to characterize DA receptor-mediated regulation of TIDA neurons. The results. reveal that quinelorane-induced activation of TIDA neurons is mimicked by the D2 receptor agonist PNU95666, but not by the D3 agonist PD128907. Unlike the activation of TIDA neurons by prolactin, this D2 receptor-mediated activation is dependent upon afferent neuronal input to the mediobasal hypothalamus. Although neurotensin and the delta opioid or opioid-1 (delta/0P1) receptor agonist DPDPE can acutely activate these neurons, quinelorane-induced activation of TIDA neurons is not dependent on either neurotensin- or delta/OPl-receptor activation. A role for kappa opioid or opioid-2 (kappa/0P2) receptors was demonstrated by experiments that revealed that quinelorane- induced activation of TIDA neuron is prevented by pre-treatment with the kappa/0P2 antagonist norbinaltorphimine (nor-BN1) and reversed by administration of the kappa/0P2 receptor agonist U-50488. In addition, in Jim mRNA hybridization studies demonstrated that quinelorane reduces the expression of prodynorphin mRN A in the arcuate nucleus. This reduction in dynorphin messenger RNA taken together with the pharmacological data suggests a role for inhibition of dynorphinergic neurons in the DA receptor-mediated activation of TIDA neurons. Experiments utilizing selective D, receptor agonists and antagonists have revealed that activation of D, receptors exerts an inhibitory influence on TIDA neuronal activity. This inhibitory response has also been demonstrated to oppose the stimulatory action of D2 receptors. This finding provides an explanation for the lack of effect of non-selective agonists and antagonists on TIDA neuronal activity. Thus, it is apparent that the activity of TIDA neurons depends on a balance between these opposing D,/D2 receptor influences. Accordingly, TIDA neuronal activity may serve as an important marker for the relative activity of various pharmacological agents at D, and D2 receptors. This work is lovingly dedicated to my wife Cathy. "Res multae filiae congregaverunt divitias tu supergressa es universas." Proverbs 31 :29 iv ACKNOWLEDGMENTS This dissertation represents the combined efforts of many that have provided a fertile environment in which I could develop as a scientist. First and foremost I wish to thank Dr. Keith Lookingland who as my thesis advisor endured much and still had the stamina to help me see this project through to completion. I deeply appreciate all the time and energy he devoted to help me grow as a scientist. A special thanks goes to Dr. Ken Moore who was instrumental in my having the opportunity to study pharmacology at Michigan State University. In addition, I appreciate the time, talent and support of Drs. Peter Cobbett and Cheryl Sisk who faithfully served on my guidance committee. I wish also to pay tribute to Dr. Monty Piercey who served as a member of my committee prior to his untimely death. I will always have fond memories of his enthusiasm for science, great intellect, and the kindness he always extended to me. Along with my professors there were several post-doctoral fellows, and students who often lent their expertise and support. These have included Drs. Jorge Manzanares and Misty Eaton who often gave me much of their time and patience in teaching me the ropes in “the lab of the most loveable & agreeable investigators”. There were also the close supportive relationships with senior students in the lab, Drs. Ed Wagner, John Goudreau, and Annette Fleckenstein as well as my contemporaries Ken Hentschel and Yvonne Will. Thanks to them the memories of long hard hours spent at the bench have faded and all that remains is the fun we had and the laughter shared. A special thank-you to Erika Bronz for providing technical assistance in many of my endeavors. Thanks also to John Johnson who put himself at my disposal for a summer that turned into two years. He exceeded my grandest expectations and readily became a valuable collaborator and friend. I also want to extend my appreciation to the folks in the front office, Dianne Hummel, Mickey Vanderlip and Nelda Carpenter who steered me through the many administrative obstacles so I could concentrate on science. Last but certainly not least, I need to thank my home team. Eric Wing, Dick Latham, and their families have given me the love, support and many prayers needed to see this project through to completion. For my wife Cathy words don’t adequately express how thankful I am to have woman at my side who never fails to inspire. These are the people who were cheering loudest when there didn’t seem to be anyone else in the stadium. Their faithfulness helped me to know that even on the darkest of days the love of Jesus was always at hand. TABLE OF CONTENTS LIST OF TABLES ......................................................................................................................... X1 LIST OF FIGURES ..................................................................................................................... XII CHAPTER ONE ........................................................................................................... 1 NEUROBIOLOGY OF DOPAMINE .................................................................................................... 1 Historical Permective 2 Neumpycbiatric Drug Diccooegy: Serendipiy to Rational Dengn 4 Dopamine receptor familie: (DI and D2) 10 Development of Selective Drug: for Subgpe: of DA Receptor: 13 Anatony and Function of DA Neuronal .S'jrtem: in the Mammalian Brain - - 18 Mesotelencephalic DA Systems ............................................................................................ 18 Hypothalamic DA Systems ................................................................................................... 23 Tubemin/undibular DA Neuron: and tbe Regulation of Prolactin Secretion ...... 28 Dopamine: Tbe Prolactin I n/n'bitog' Factor 30 Prolactin Feedback Regulation of NBA N eumnr ....... . . 32 Aflerent Modulation of'IYDA Neuronal Acting: ..... . 37 Dopamine Receptor Mediated Regulation of TIDA Neurons ........................................ 37 SPECIFIC AIMS ................................................................................................................................. 42 CHAPTER TWO ......................................................................................................... 44 GENERAL METHODS ...................................................................................................................... 44 Animalr 44 Drugs 44 Surgical Procedures 44 Intracerebroventricular Cannulation .................................................................................... 44 Orchidectomy .......................................................................................................................... 45 Neumcbemical ertimation tyneumnal acting! 46 Tissue and Blood Processing ................................................................................................. 48 High Performance Liquid Chromatography ....................................................................... 53 Radioimmunoassay for Prolactin .......................................................................................... 54 Measurement oijnorpbin mRNA bi in ritu Hybridization Hirtocbemirtpl 56 Labeling The mRNA Probe ........................................................ 57 Extraction of the Labeled Probe .......................................................................................... 57 Preparation of Tissue ............................................................................................................. 58 In Situ Hybridization Procedure ........................................................................................... 59 Image Analysis ......................................................................................................................... 61 Dual Immunobtlrtocbemical Determination of For/ Pm and TH. ....... .. 64 Tissue Preparation .................................................................................................................. 65 Dual Immunohistochemistry ................................................................................................ 65 StatirticalAnabue: 71 CHAPTER THREE .................................................................................................... 72 D2 RECEPTORS MEDIATE THE STIMULATORY EFFECTS OF QUINELORANE ON TUBEROINFUNDIBULAR DA NEURONS ...................................................................................... 72 Met/Jodr 74 Drugs ........................................................................................................................................ 74 Result: 75 Discum'on 86 CHAPTER FOUR -- - - ........ --92 D2 RECEPTOR-MEDIATED STIMULATION OF TIDA NEURONS IS DEPENDENT ON AFFERENT NEURONAL INPUT TO THE MEDIOBASAL HYPOTHALAMUS ................................ 92 Metbodr 95 Deafferentation of the Mediobasal Hypothalamus ............................................................ 9S Verification of the Deafferentation ...................................................................................... 96 Rerult: 97 Dtlramion 101 CHAPTER FIVE -- ................................................................................. 107 EVIDENCE THAT ACTIVATION OF DELTA / CPI-RECEPTORS STIMULATES TIDA NEURONS ....................................................................................................................................... 107 Metbodr 108 Drugs ...................................................................................................................................... 108 Result: 109 Dircum'on ..... 1 20 CHAPTER SIX _ -- - - - -- .... ............ 123 ACTIVATION OF D2 RECEPTORS FAILS TO INDUCE THE RELEASE OF A NEUROTRANSMI’I’I‘ER THAT STIMULATES TIDA NEURONS .................................................. 123 Delta/0P1 opioid receptors ............................................................................................... 124 Neurotensin receptors .......................................................................................................... 125 Metbodr . - 127 Drugs - . - -_ . ........................................................................................................... 127 Rerult: ........... 127 Dircum'on 131 CHAPTER SEVEN ................................................................................................... 135 EVIDENCE THAT D2 RECEPTOR-MEDIATED ACTIVATION OF TIDA NEURONS OCCURS VIA DISINHIBITION ...................................................................................................................... 13S Met/)odr ................. 1 38 Drugs ...................................................................................................................................... 1 38 Rem/tr 1 39 Dircum'on 147 CHAPTER EIGHT ................................................................................................... 152 INHIBITION OF TUBEROINFUNDIBULAR DOPAMINERGIC NEURONAL ACTIVITY BY SELECTIVE ACTIVATION OF D1 RECEPTORS ........................................................................... 152 Metbodr 154 Drugs ...................................................................................................................................... 1 54 Rein/t: 1 54 Diram‘ion .- 1 60 CHAPTER NINE ...................................................................................................... 163 OPPOSING ROLES FOR D1 AND D2 DOPAMINERGIC RECEPTORS IN THE REGULATION OF TUBEROINFUNDIBULAR DOPAMINERGIC NEURONAL ACTIVITY .......................................... 163 Met/Md: .. 1 66 Drugs ...................................................................................................................................... 167 Rerult: 1 67 Diramion 1 72 CHAPTER TEN- -- ........ .................................................................... 176 GENERAL SUMMARY AND DISCUSSION ..................................................................................... 176 D2 Rec¢tor—mediated activation of TIDA neuron: - ..... 176 D1 recq)tor-med1'atedinbibition of TIDA neuron: and prolactin :ecretion ....... 181 Interaction: between DI and D2 receptor: in TIDA neuronal regulation ....... -- 185 APPENDIX .................................................................................................................................. 1 89 BIBLIOGRAPHY ....................................................................................................................... 1 92 Table 1 Table 2 Table 3 Table 4 Table 5 Table 6 Table 7 Table 8 Table 9 LIST OF TABLES A comparison of three generations of DA agonists and antagonists by affinity for DA receptors .................................................................................................................... 17 Effects of quinelorane on dopamine and DOPAC/ DA ratios in the median eminence and nucleus accumbens ................................................................................ 78 Effect of PNU-95,666 on dopamine and the DOPAC to dopamine ratio in the median eminence and nucleus accumbens .................................................................. 81 Lack of effect of PD128907 on dopamine and the DOPAC to dopamine ratio in the median eminence and nucleus accumbens ............................................................ 82 Effects of surgical deafferentation on the concentration of dopamine and norepinephrine in the median eminence, and dopamine in the intermediate lobe of the pituitary of male rats one week after surgery ................................................... 98 Lack of effect of DPDPE on DOPAC concentrations in the striatum and intermediate pituitary lobe. .......................................................................................... 11 1 Lack of effect of DPDPE on DOPA concentrations in striatum and intermediate pituitary lobe ..................................................... - ...................................... 1 13 Lack of effect of DPDPE on DOPAC concentrations in striatum and intermediate pituitary lobe ............................................................................................ 1 15 Lack of effect of naltrindole on DOPAC concentrations in the striatum and intermediate pituitary lobe. .......................................................................................... 1 18 Table 10 List of drugs used by generic name chemical name (source) and vehicle solvent used. ............................................................... 1 90 LIST OF FIGURES Figure 1 The structure of dopamine drawn in two conformational extremes designated: alpha rotomer and beta rotomer adapted from Cannon, (1975). ........................... 15 Figure 2 Drawing of a sagittal View of the rat brain modified from Fuxe et a/., (1985) depicting the nigrostriatal DA neuronal pathway. .................................................... 20 Figure 3 Drawing of a sagittal View of the rat brain modified from Fuxe et al., (1985) depicting the mesolimbic DA neuronal pathway. ..................................................... 21 Figure 4 Drawing of a sagittal View of the rat brain modified from Fuxe et al., (1985) depicting the mesocortical DA neuronal pathway .................................................... 22 Figure 5 Drawing of a sagittal View of the rat brain modified from Fuxe et al., (1985) depicting the hypothalamic dopaminergic DA neuronal cell bodies ..................... 27 Figure 6 Prolactin feedback regulation of TIDA neurons .......................................................... 36 Figure 7 Neurochemical events in axon terminals of central dopaminergic neurons ........... 47 Figure 8 The anatomical locations of the nucleus accumbens (shaded), and striatum (cross hatched) tissue samples dissected for neurochemical analysis. ............................... 50 Figure 9 The anatomical location of the median eminence (cross-hatched) tissue sample dissected for neurochemical analysis ........................................................................... 51 Figure 10 A sagittal section through the pituitary gland showing the anatomical location of the neural lobe of the pituitary (shaded) that was removed in order to allow dissection of the intermediate pituitary lobe (cross hatched) for neurochemical analysis ................ - ............................................................................... 52 Figure 11 Scheme for in :itu hybridization with synthetic oligonucleotides. ............................ 60 Figure 12 Computer captured micrograph depicting specific labeling of a cell containing dynorphin mRNA in the arcuate nucleus. ................................................................. 63 Figure 13 Computer-captured image of the ME and ARC in a frontal brain section. ........... 67 Figure 14 Computer-captured micrograph depicting a frontal brain section irnmunocytochemically labeled for both TH and lFos/ F RA. .................................. 69 Figure 15 Dose response effects of quinelorane on DOPAC concentrations in the median eminence and nucleus accumbens of male rats ......................................................... 77 Figure 16 Time course of the effects of PNU-95,666 and PD—128907 on prolactin concentrations in plasma. ................................... 79 Figure 17 Effects of PNU-95,666 and PD-128907 on DOPAC concentrations in the median eminence and nucleus accumbens. ................................................................ 80 Figure 18 Time course of the effects of PNU-95,666 and PD-128907 on DOPAC concentrations in the median eminence. ....... 84 Figure 19 Effects of PD-128907 on DOPA concentrations in the median eminence and nucleus accumbens. ...... . . - ............................................................ 85 Figure 20 Schematic diagram depicting a Halasz knife and the region of the mediobasal hypothalamus circumscribed by the knife—cut procedure which includes median eminence (ME), arcuate nucleus (ARC) and the pituitary. ...................................... 94 Figure 21 Effects of quinelorane on DOPAC concentrations in the median eminence of rats receiving either sham surgery (Sham Control) or surgical deafferentation of the mediobasal hypothalamus (MBH Knife-cut). ............................................................ 99 Figure 22 Effects of quinelorane on DOPAC concentrations in the nucleus accumbens rats receiving either sham surgery (Sham Control) or surgical deafferentation of the mediobasal hypothalamus (MBH Knife-cut). .......................................................... 100 Figure 23 Dose-response of the effects of DPDPE on DOPAC concentrations in the median eminence and nucleus accumbens. .............................................................. 110 Figure 24 Dose-response of the effects of DPDPE on the accumulation of DOPA in the median eminence and nucleus accumbens ............................................................... 112 Figure 25 Time course of the effects of DPDPE on DOPAC concentrations in the median eminence and nucleus accumbens. ............................................................................ 114 Figure 26 Lack of effect of naltrindole on DOPAC concentrations in the median eminence and nucleus accumbens ............................................................................................... 117 Figure 27 Effects of naltrindole on DPDPE-induced increases in DOPAC concentrations in median eminence and nucleus accumbens. ......................................................... 119 Figure 28 Lack of effect of naltrindole on quinelorane-induced increases in DOPAC concentrations in the median eminence and nucleus accumbens. ....................... 129 Figure 29 Lack of effect of SR-48692 on quinelorane-induced Changes in DOPAC concentrations in the median eminence and nucleus accumbens. ....................... 130 Figure 30 Effects of quinelorane on DOPA concentrations in the median eminence and nucleus accumbens of vehicle- and U-50,488-treated rats ..................................... 141 Figure 31 Effects of quinelorane on DOPA concentrations in the median eminence and nucleus accumbens of vehicle-, nor-BNI- and prolactin-treated rats .................. 142 Figure 32 Effects of quinelorane on DOPA concentrations in the median eminence of intact and orchidectomized rats. ................................................................................ 144 Figure 33 Effects of nor-BN1 on DOPA concentrations in the median eminence of intact and orchidectomized rats ............................................................................................ 145 Figure 34 Effects of quinelorane on the expression of dynorphin mRNA in the arcuate (ARC) and ventromedial hypothalamic (VMN) nuclei of the rat. ........................ 146 Figure 35 Dose-response effects of SCH39166 on the concentrations of DOPAC in the median eminence. ........................................................................................................ 156 Figure 36 Time-course of effects of SCH39166 on the concentrations of DOPAC in the median eminence. ........................................................................................................ 157 Figure 37. Effects of SKF38393 on the concentrations of DOPAC in the median eminence and prolactin in plasma. .............................................................................................. 158 Figure 38 SCH391663 blocks the SKF38393-induced decrease in the median eminence DOPAC concentrations. ............................................................................................ 159 Figure 39 Dose—response effect of SKF38393 on DOPAC concentrations in the median eminence quinelorane-treated rats. - ........................................... 169 Figure 40 Effects of apomorphine on DOPAC concentrations in the median eminence of vehicle- and quinelorane-treated rats. ....................................................................... 170 Figure 41 Effect of SKF38393 on quinelorane-induced changes in the number of perikarya in the ARC co-expressing TH-IR and Fos/FRA-IR. ............................................. 171 Figure 42 A schematic representation of the hypothetical relationship between inhibitory D1 and stimulatory D2 receptors on the regulation of TIDA neuronal activity..178 xiv Chapter One Neurobiology of Dopamine Interest in dopaminergic (DA) neurotransmission in the brain over the past forty years has been driven primarily by two seminal Clinical discoveries in neuropharmacology. Firstly, that L-3,4—dihydroxyphenylalanine (L-DOPA) treatment could transiently alleviate the symptoms of Parkinson’s disease by enhancing DA neurotransmission in the basal ganglia, and secondly that blockade of dopamine receptors in limbic centers was found to be associated With the antipsychotic efficacy of neuroleptic drugs. These important discoveries along with advances in molecular biology identifying multiple DA receptors have encouraged the development of drugs with greater selectivity for the various SubtyPes of these receptors. Although it has long been recognized that nellretransmission of most DA neurons in the brain is subject to DA (autoreceptor) mediated negative feedback, tuberoinfundibular DA neurons in the mediobasal hypothalamus are not subject to this mode of DA receptor- Illecliated regulation. This assertion is supported by numerous studies in which DA agonists and antagonists that affect autoreceptor—mediated processes in Qfllét DA neurons have no effect on TIDA neurons. The development of DA (It-“gs with greater selectivity for the various DA receptors has uncovered the possibility that TIDA neurons are subject to DA receptor-mediated regulation. The DA regulation of TIDA neurons is different in many important ways from that mediated by autoreceptors in other DA neurons. Thus, studies described in this dissertation have made use of newer pharmacological tools to better characterize the mechanism of this regulation. Historical Perspective The first reports Of dopamine synthesis were published in 1910 (Mannich and Jacobsohn; Barger and Ewins). However, discovery of a potential physiological role for this compound was overshadowed by comparison to the C10de related biogenic amines epinephrine and norepinephrine. Experiments COil’lparing a series of biogenic amines demonstrated that dopamine had relatively mild vasoconstrictive properties when compared with epinephrine and norepinephrine (Barger and Dale 1910). The eventual discovery of dopamine as a full-fledged neurotransmitter in its OVm right is inexorably linked to the discovery of its immediate precursor DOPA (Funk et al., 1911). During this period the pathway by which these airlines were fonned in vivo received much attention. DOPA was discovered as a byPthuct of this effort It was nearly thirty years later when Holtz et al., (1938) demonstrated the existence of L-DOPA decarboxylase in mammalian kidney. This discovery led Blaschko (1939) to propose that L-DOPA and dopamine were intermediates in the biosynthesis of norepinephrine and epinephrine from L-tyrosine. These studies, in conjunction with the lack of an observable response when tested in various animal models Of adrenergic activity served to strengthen the notion that both L—DOPA and dopamine were merely intermediates in biosynthesis of adrenergic amines. Elevation of dopamine from mere biosynthetic intermediate to neurotransmitter status took an additional 20 years. Blaschko (1957) asserted that dopamine may have some physiological function following studies that examined the content of dopamine in various tissues. The role of dopamine as a biosynthetic intermediate seemed most plausible in Chromaffin tissue where it makes up only 2-3 °/o of the total catecholamine content. However, in adrenergic neurons dopamine levels were comparable to those of norepinephrine (Schfimann 1956) and this suggested a potential, yet undiscovered physiological role for dopamine in these neurons. Why store large quantities of dopamine if it Serves only as a biosynthetic precursor? The experiment demonstrating a unique physiological role for dopamine had been done fifteen years earlier by Holtz and Ctel'lcler (1942) who demonstrated that dopamine administered to guinea pigs Cans ed a vasodilation rather than the expected vasoconstriction (Barger and Dale 1 9 1 0). However, this unique finding went unrecognized because Holtz and Ctetlcier (1942) had suggested that this puzzling effect was the result of dopatnine being metabolized to 3,4-dihydroxyphenylacetaldehyde via ITlotfloamine oxidase. Thus, they dismissed this observation by suggesting that it was the result of a non-specific effect of this metabolite. However, when monoamine oxidase inhibitors became available Hornykiewicz (1958) was able to ConfimI that dopamine and not the phenylacetaldehyde metabolite was the active principle responsible for the vasodilatory effects observed in the guinea pig. This finally confirmed Blaschko’s hypothesis that dopamine has a physiological role that is independent from that of norepinephrine and epinephrine. Neuropsychiatric Drug Discovery: Serendipity to Rational Design Dopamine was defined as a neurotransmitter in the periphery at about the same time the potent antipsychotic chlorpromazine was developed for SChizophrenia, and LDOPA was discovered for the treatment of Parkinson’s disease. The recognition of dopamine as a common denominator in both diseases shifted the focus of this research away from peripheral effects to the fill'1C1:ion DA neurotransmission in the central nervous system. Traditionally drug discovery has been an empirical process that has depended on serendipity coupled with astute clinical observations that have lead to a. greater understanding of biology and the development of new treatments for disease. The discovery of chloropromazine as a tranquilizer and subsequent utility as an antipsychotic agent is a classic example of the traditional drug disc()very process. Henri Laborit discovered the progenitor to chlorpromazine, promethazine after searching for a drug treatment for the prophylaxis of surgical Shock (1949). Promethazine was selected for its histamine blocking activity in hopes that this would be effective in preventing the cardiovascular response to the autonomic nervous system instability. However, some unique secondary Central nervous system actions (hypnotic, analgesic and hypothermic) differentiated this compound from other similar antihistaminergic agents and made it a useful adjuvant to anesthesia. These beneficial secondary properties Were exploited when a series of phenothiazide drugs was evaluated by RhOne- Poulenc pharmacologist Simone Courvoisier in 1950 (Swazey 1977a). The results Of these studies led to the selection of promazine as a lead compound for “eminent of surgical shock (Swazey 1977a). Further experiments demonstrated that the addition of a chlorine group to the phenothiazine ring of promazine resulted in chlorpromazine, a compound that had exceptional activity with low to3(icity (Swazey 1977a). Subsequent Clinical use confirmed that it did indeed have excellent anti-shock activity. However, Laborit also noted its ability to calm at1)ciation with the earlier work on Parkinson’s disease where there is a relatively Simple association between the disappearance of dopamine and the symptoms of tl'le disease. It has long been a tempting analogy to view Parkinson’s disease and Selli20phrenia as opposite sides of the same coin. Even though dopamine is one net"! I! LLEAUJ ..‘....,. n ,0“ ‘ l, 4 ‘Iu&~ ’9'! fr, “1 b 1"] A. 1 r . j L unL 'I‘ d mi“ '1» ill _l AR, 5“ W 6“ 9. I I.‘ common denominator in both diseases the study of etiological factors in schizophrenia is complicated by the lack of an Objective measure to differentiate it from a family of related psychiatric disorders. These difficulties not withstanding, it is clear that the neurobiology of dopamine owes much to the nearly four decades of intensive work that have attempted to illuminate the root causes of these devastating neuropsychiatric diseases. Dopamine receptor families (Di and DZ) Once a distinct neurotransmitter function had been established for dopamine, a significant effort was directed at the study of specific DA binding sites in order to elucidate the various functions of this neurotransmitter. The emerging role for dopamine depletion in the pathogenesis of Parkinson’s disease and the efficacy of DA antagonists in the treatment of affective disorders have provided the fuel that has propelled the basic science of DA receptor pharmacology for nearly 30 years. In addition, these discoveries have moved the search for therapies from a wholly empirical approach to a more rational hypothesis-driven endeavor. Based on their review of a decade of dopamine receptor research, Cools and van Rossum (1976) forwarded the concept that there are two populations of DA receptors in the mammalian brain: excitation- mediating (DAC) and inhibition-mediating (DAi). This notion of a homeostatic balance between DAe - DA, activity accounted for many discrepancies observed 10 in various animal studies and provided a new way of interpreting data from various animal models. A number of investigators accepted and built upon this notion of two distinct receptors for dopamine (Katzman et al., 1977; Spano et al., 1978). This classification was based largely on the finding that dopamine or various DA agonists could stimulate or inhibit the activity of adenylyl cyclase thus altering the levels of cyclic adenosine monophosphate (CAMP) within target cells. In general, there was a good correlation between antipsychotic efficacy and the ability of compounds to block dopamine stimulated increases in CAMP. There were some conspicuous exceptions to this rule since neither butyrophenones such as haloperidol nor substituted benzamides such as sulpiride could block dopamine dependent CAMP production, in spite of their high antipsychotic efficacy (Roufogalis et al., 1976; Spano et al., 1978). However, the advent of radioligand-binding assays Clearly identified a subpopulation of receptors that were targeted by haloperidol (Burt et al., 1976; Seeman et al., 1975). Kebabian and Calne (1979) further extended this idea and named the excitation-coupled receptors D1 and the inhibitory-coupled receptors D2. Thus, identification of two distinct DA receptors explained the earlier conflicting reports by demonstrating that dopamine stimulated CAMP production was mediated by D1 receptors, whereas neuroleptic responses depended upon blockade of D2 receptors. 11 Advances in molecular biological techniques that used the complimentary DNA predicted from known protein sequence and structural information allowed for the discovery of gene sequences and products associated with the super-family of G—protein coupled receptors. These gene-cloning discoveries eventually provided the insight necessary to identify dopamine receptor genes and gene-products as a subset of D1 and D2 families of receptors. Based on the structural similarity of the seven transmembrane spanning sequences for the beta-adrenergic receptor the rat D2 receptor was discovered. Using a similar strategy with the known D2 sequence the D, receptor followed by the D3, D4 and D5 were identified in rapid sequence (Civelli et al., 1993). This information provided a way to clone and then express these dopamine receptors in various model systems. Subsequently, these model systems have allowed for a rigorous examination of DA receptor biochemistry, pharmacology and associated physiological functions. The results of these studies altered the simple concept of two dopamine receptors and replaced it with at least five distinct receptor sub- types. However, these receptors could still be grouped into sub-families based upon identified second messenger systems and gene sequence similarity. Thus, D1 receptors were divided into D, and D5 subtypes, while D2 receptors were comprised of D2, D3, and D4 subtypes. Note that whenever reference is being made to specific subtypes of DA receptors the subscript will be used (e.g. D,). 12 On the other hand when referring to a family of DA receptors no subscript will be used (e.g.Dl). Development of Selective Drugs for Subtypes of DA Receptors One of the limiting factors in the treatment of Parkinson’s disease using DA agonists is the development of psychotic side effects such as disturbing Visual hallucinations (Calne 1978). Similarly, during the treatment of schizophrenia with DA antagonists the emergence of a Parkinson-like syndrome occurs that presents acutely as akinesia and motor slowing. With longer-term treatment these drugs induce the more severe motor impairment of tardive dyskinesia. These extrapyramidal side effects are correlated with greater D2 receptor blockade in the basal ganglia (Nordstrém et al., 1993). The discovery of multiple dopamine receptors has provided hope that development of drugs that interact with a specific subset of these receptors would lead to more effective clinical management of both Parkinson’s disease and schizophrenia with minimal or no unwanted side effects. The process of in developing specific drugs for these diseases has progressed in a stepwise fashion whereby generations of drugs have been developed based on the biological and pharmacological tools of the time. Thus, the first generation of drugs was developed prior to understanding the role of dopamine as a neurotransmitter. Early dopamine agonists (e.g. apomorphine 13 and bromocn'ptine) had been selected empirically because of their utility as adjunctive therapies with L—DOPA in the treatment of Parkinson’s disease. Likewise, the development of first generation DA receptor antagonist chlorpromazine was the result of serendipity as discussed above. The utility of chlorpromazine encouraged the mass screening of chemical libraries of compounds to discover other drugs with similar properties such as haloperidol. In general the first generation of DA agonists and antagonists tend to interact with a variety of other receptors and have no clear preference for a single DA receptor subtype (Table 1). The discovery of multiple dopamine receptors has fostered the development a second generation of DA compounds. The availability this biological information and the technology to readily express and study DA receptors in vitro has allowed for the development of detailed structure activity relationships for large libraries of compounds. It is from these libraries that these partial ergoline compounds such as quinelorane (Foremann et al., 1989) and raclopride (Ogren et al., 1988) were developed because of their preference for the D2 family of receptors. Similarly, the substituted benzazepines SKF38393 (Sibley et al., 1982) and SCH39166 (Chipkin et al., 1988) were developed for their D1 receptor preference (Table 1). Recently, the availability of powerful computers to rigorously compare these structure activity relationships have allowed for the development of a 14 molecular drug design approach. This strategy has been developed to achieve the goal of discovering more selective drugs for the treatment of these neuropsychiatric disorders. Initially a hypothetical model of the target receptor is developed based on the structural Characteristics of known ligands for the receptor. This model is referred to as the “pharmacophore” and describes the structural features of the ideal ligand with theoretical affinity and functional efficacy at a particular receptor. One particular advantage to using this approach to discover new DA ligands has been the observation that many DA ligands incorporate elements of dopamine in their structure. Although dopamine has a somewhat lower affinity for D2 receptors the flexibility of dopamine allows for full efficacy and reasonable affinity at all DA receptors. Dopamine can exist in one of two chemical forms i.e. the alpha rotomer and the beta rotomer (Cannon, 1975; Figure 1). Axis of rotation Alpha Rotomer Beta Rotomer Figure 1 The structure of dopamine drawn in two conformational extremes designated: alpha rotomer and beta rotomer adapted from Cannon, (1975). These conformations are possible due to the free rotation of the aliphatic amine in relation to the catechol moiety. 15 The finding that the alpha rotomer corresponds best with the pharrnacophore of D2-like receptors whereas the beta rotomer corresponds best with Dl-like receptor subtypes provides an explanation for the selectivity of various chemical classes to either of these receptor families. Although this principle cannot account for the selectivity of all chemical substances at dopamine receptors it has provided an organizing framework by which the DA receptor selectivity of various chemical substances may be predicted. A productive approach to discovering compounds with greater selectivity for one receptor subtype has been the use of rigid dopamine analogs that represent dopamine locked in either the alpha- or beta-rotomeric form. Intentionally designing drugs that are locked into or favor the beta-rotomeric conformation has led to the D1 selective agonists SKF38393 (Setler et al, 1978) and dihedrexadine (Mottola et al., 1992). Those that are rigid analogs of the alpha- rotomer seem to be relatively more selective for the D2 family, the partial ergoline quinelorane is a prominent example that has activity at both D2 and D3 receptors (Gackenheirner et al, 1995). The 2—arninotetralins (7 OH-DPAT) and benzoquinolines (PD128907) are ridgid alpha-rotomer analogs that show a preference for D3 receptors. 16 Table 1 A comparison of three generations of DA agonists and antagonists by affinity for DA receptors D1 Family D2 Family D1 D5 D2 D3 D4 First Generation Agonist Apomorphine +/- + +++ ++ +++ Bromocriptine + + +++ +++ + Antagonists Chloropromazine + + +++ ++ ++ Haloperidol + + ++++ ++ +++ SCH23390 ++++ ++++ +/- +/- +/- Second Generation Agonists SKF38393 +++ ++++ + +/- +/- Dihydrexidine ++ ND + + ND Quinperole - ND + /- ++ ++ Quinelorane - ND +++ +++ ND Antagonists SCH39166 + ++ ND + /_ ND ND Raclopride - ND + + + + + + + /- Third Generation Agonists PNU95666 - ND + + + /- - PD128907 - ND + + + + 7OH-DPAT + /- ND + + ++ + + /- ++++, Inhibition constant (K) < 0.5 nM; +++, 0.5 nM < K, < 5 nM; ++, 5 nM < K,< 50 nM; +, 50 nM < K,< 500 nM; +/-, 500 nM < K< 5 11M; -, K>5 11M; ND, not determined; (Modified from Missale et al., 1998) 17 Anatomy and Function of DA Neuronal Systems in the Mammalian Brain A technical advance allowing the visualization of monoarnines in tissue sections via fluorescence microscopy (Falck et al, 1962) allowed for rapid advances in our understanding of the distribution and function of dopamine in the mammalian brain. Soon after this technique was developed, a system for mapping the distribution of DA and noradrenergic cell bodies in the brain was described by DahlstrOm and Fuxe (1964). One of the intriguing features of central DA neurons that distinguishes them from other arninergic systems in the brain is the discrete nature of DA cell body origins and terminal field projections. This differs markedly from the widely diffuse patterns of innervation observed in Cholinergic, 5-hydroxytryptaminergic, and noradrenergic neuronal systems. The discrete nature of its neuronal network has enabled neurobiologists to more readily appreciate the correlation between anatomy, biochemistry and function in these systems. The DA neuronal systems are loosely categorized as being either mesotelencephalic or hypothalamic according to the distribution of their cell bodies and axon terminals. Mesotelencephalic DA Systems The mesotelencephalic DA neuronal systems originate predominantly in the mesencephalon with cell bodies clustered in the par: corrpacta of the 18 substantia nigra and in the ventral tegmental area (Anden et al, 1964; Lindvall and Bjérklund 1974). These brain regions correspond with areas A8, A9 and A,O according to the alpha-numeric system developed by DahlstrOm and Fuxe (1964). These mesotelencephalic DA projections have been sub-divided into three systems based on their terminal field projections: nigrostriatal, mesolimbic and mesocortical. As shown in figure 2, nigrostriatal DA neurons originate in the substantia nigra and project to the caudate-putamen; designated as striatum in the rat. Mesolimbic DA neurons originate in the ventral tegmental area and project to the nucleus accumbens and other subcortical structures such as the lateral septum and olfactory tubercle (Figure 3). Mesocortical DA neurons project to several cortical regions including the cingulate, entorhinal, prefrontal and pyriforrn cerebral cortices (Figure 4). These mesotelencephalic pathways have projections that are highly divergent and subserve different functions within the brain (Bloom 1983). 19 Figure 2 Drawing of a sagittal View of the rat brain modified from Fuxe et al, (1985) depicting the nigrostriatal DA neuronal pathway. The locations of cell bodies are indicated by the triangles, axons are represented as solid lines and dots represent axon terminals and synaptic boutons. Abbreviations: accumbens nucleus (Acb), caudate pummen (CPu), globus pallidus (GP), medial forebrain bundle (mfb), substantia nigra pars compacta (SNC), substantia nigra pars reticulata (SNR), subthalamic nucleus (STh), and olfactory tubercle (TU). 20 Figure 3 Drawing of a sagittal view of the rat brain modified from Fuxe et al, (1985) depicting the mesolimbic DA neuronal pathway. The locations of neuronal cell bodies are indicated by the triangles, axons are represented as solid lines, and dots represent axon terminals and synaptic boutons. Abbreviations: accumbens nucleus (Acb), caudate putamen (CPu), medial forebrain bundle (mfb), olfactory tubercle (I' u). 21 Phi. . EN 154‘ is ‘5 Figure 4 Drawing of a sagittal View of the rat brain modified from Fuxe et al, (1985) depicting the mesocortical DA neuronal pathway. The locations of neuronal cell bodies are indicated by the triangles, axons are represented as solid lines, and dots represent axon terminals and synaptic boutons. Abbreviations: accumbens nucleus (Acb), anterior olfactory nucleus posterior (AOP), caudate putamen (Cpu), lateral habenular nucleus (LHb), lateral septal nucleus (LS), olfactory tubercle (Tu). Electrophysiological evidence suggests that mesotelencephalic DA projections are inhibitory (Bannon et al, 1986; Bunney et al, 1987) and, on the basis of behavioral evidence, subserve a permissive function with regard to emotional responsiveness and initiation of movements (BjOrklund and Lindvall 1983). Perturbations in these systems are implicated in the pathogenesis of Parkinson’s disease (DA hypofunction in nigrostriatal DA system) and schizophrenia (DA hyperfunction in mesolimbic DA system) as discussed in the previous section. In addition these DA neurons have a role in learning acquisition and positive reinforcement, and may play an important role in the mediating the positive reinforcing effects of drugs of abuse (Callahan et al., 1997). Hypothalamic DA Systems Hypothalamic DA neurons are organized into distinct systems with clusters of cell bodies having discrete axonal projections. The terminal fields of these local circuits remain mostly within the hypothalamus but are less well defined than those of the extrahypothalarnic systems. Four DA systems have been characterized in this region: incertohypothalamic, periventricular, periventricular—hypophysial and tuberoinfundibular (TI) DA neurons. The cell bodies of incertohypothalamic DA neurons designated A, 3 are located in the medial zona incerta (DahlstrOm and Fuxe 1964; BjOrklund, et al, 1975). Short fibers from the medial zona incerta project ventromedially to the 23 dorsomedial hypothalamic nucleus, and these fibers have been identified as dendrites (Wagner et al., 1993; Chan-Palay et al., 1984). Projections of the incertohypothalamic DA neurons to the central nucleus of the amygdala, diagonal band of Broca and paraventricular nucleus have been identified by anterograde tracing and neurochemical experiments (Wagner et al, 1995; Eaton et al., 1993a; 1994). In support of these observations, retrograde tracing studies have demonstrated tyrosine hydroxylase positive neuronal projections to the central amygdala , horizontal limb of the diagonal band of Broca and paraventricular nucleus (PVN) originating from the medial zona incerta (Cheung et al, 1998). These tracing studies have provided important clues to the function of incertohypothalamic DA neurons. The horizontal diagonal band contains cell- bodies of gonadotropin-releasing hormone neurons (Merchenthaler et al, 1989) and thus may be the target of these incertohypothalamic DA projections. This hypothesis is strengthened by studies that have implicated the medial zona incerta in the regulation of luteinizing hormone surges and ovulation (Mackenzie et al, 1984; Sanghera et al, 1991). In contrast to either the central amygdala or horizontal diagonal band which receive DA projections from multiple sites, the DA projections to the paraventricular nucleus appear to originate from medial zona incerta. In addition, other studies have demonstrated a close proximity of DA and corticotrophin—releasing hormone neurons in this region (Liposits and 24 Paul] 1989). Taken together with the demonstration DA agonists increase corticotropin—releasing hormone mRNA levels in the PVN, suggests IHDA neuronal involvement in stress-induced activation of the hypothalamic-pituitary- adrenal axis (Cheung et al., 1998). A,4 Periventricular DA cell bodies are distfibuted in the periventricular nucleus along the entire rostrocaudal extent of the third ventricle (van den Pol, Herbst and Powell 1984). Although not well characterized, fibers of periventricular DA neurons in the rostral periventricular nucleus (see Figure 5) are believed to project laterally into adjacent medial preoptic nucleus and anterior hypothalamic area (BjOrklund and Lindvall 1984). The function of periventricular DA neurons is not well characterized, but medial preoptic lesions (Bitran et al., 1988) or injection of DA antagonists into this region (Pehek et al., 1988) inhibits copulatory behavior in male rats suggesting a role for these DA neurons in the regulation of sexual behavior. Periventricular-hypophysial IDA neuronal system consists of a subpopulation of A, 4 DA cells in the periventricular nucleus located ventral to the paraventricular nucleus; these DA neurons project to the pituitary intermediate lobe (Luppi et al., 1986; Goudreau et al., 1992). Dopamine released from these neurons tonically inhibits the release of pro-opiomelanocortin- derived hormones such as alpha—melanocyte stimulating from the intermediate lobe of the pituitary (Millington et al. 1988). 25 Of all the intrahypothalamic DA neurons TIDA neurons are the best studied. Cell bodies designated A,2 (DahlstrOm and Fuxe 1964) for these neurons are located in the arcuate nucleus and short axons project ventrolaterally to the external layer of the median eminence (Bjérklund and Nobin, 1973). These DA neurons function in the regulation of prolactin secretion as discussed in the next section. 26 Flgtlte 5 Drawing of a sagittal view of the rat brain modified from Fuxe et al., (1985) :ePipu'ng the hypothalamic dopaminergic DA neuronal cell bodies. The location of DA cell vod-les is indicated by the filled triangles. Abbreviations: 3V third ventricle, 4V fourth - etltticle, arcuate nucleus (ARC), anterior pituitary lobe (AP), dorsomedial nucleus (DM), rte-mediate pituitary lobe (IL), mamillothalamic tract (mt), median eminence (ME), medial Qua incerta (MZI), neural pituitary lobe (NL), optic chiasm (ox), periventricular nucleus e\fbl), and ventromedial nucleus (V M). 27 Tuberoinfundibular DA Neurons and the Regulation of Prolactin Secretion Unlike the stimulatory effect of hypothalamic releasing factors on the other anterior pituitary hormones the primary action of the hypothalamus upon prolactin secretion is inhibitory. It has long been known that isolated pituitaries grafted into other animals hypersecrete prolactin. Everett (1954) transplanted pituitary glands beneath the kidney capsule and reported on the biological effects of prolactin hypersecretion. Likewise, in studies where the portal blood flow is interrupted by pituitary stalk transection plasma prolactin levels are rapidly elevated (within 30 min) and remain elevated up to 6 months later (N ikitovich- Winer, 1965; Welsch et al., 1971; Kanematsu and Sawyer, 1973; Murai et al., 1 989). Conversely, grafting isolated pituitaries to an intact portal blood supply do not hypersecrete prolactin as they did if grafted elsewhere (Lu and Meites, 1972). That the hypothalamus is the source of the prolactin inhibitory factor carried in portal blood was suggested by the finding that hypothalamic lesions l.Ilduce lactation in rats (Halasz et al., 1962). This possibility was demonstrated more directly by showing that bilateral lesions in the median eminence increase Plasma prolactin levels, whereas lesions in the central nucleus of the amygdala have no effect (Chen et al., 1970). These studies suggested that the mediobasal lJYI’thalamus was the source of or stimulated the release of a prolactin 28 inhibitory factor that controls the secretion of prolactin from the anterior pituitary gland. The notion that the prolactin inhibitory factor originates from the hypothalamus is also supported by a number of in vitro studies that have shown that prolactin secretion in isolated anterior pituitaries is inhibited when exposed to homogenates or acid extracts of hypothalami. These extracts were known to contain releasing factors for follicle-stimulating hormone, luteinizing hormone, thyroid stimulating hormone and adrenocorticotropin in addition to inhibiting factor for prolactin (Pasteels, 1961; Talwaker et al., 1963). Similar in vivo experiments utilizing hypothalamic extracts also reduced plasma prolactin concentrations 1 and 4 hours after injection into non-lactating female rats (Amenmori and Meites, 1970). The possibility that these effects were direct on the pituitary was suggested by Kamberi et al., (1971) who found that infusion of hypothalamic extracts into a portal vessel inhibited prolactin secretion. Talwalker 0” al., (1963) noted that many neurotransmitters found in the hypothalamus (acetylcholine, epinephrine, norepinephrine, serotonin, histamine, oxytocin, and Vas Opressin) had no effect on in vitro prolactin release. A role for biogenic amines in the hypothalamic control of prolactin SeCretion was first suggested by studies in catecholamine-depleted rabbits CIlactin release from anterior pituitaries. Their system made use of 3H-leucine Whith when incubated with isolated pituitaries is incorporated into newly S3’1'1t11esized prolactin. Application of dopamine to these pituitary cultures 31 significantly reduced the level of radiolabeled prolactin in the medium and increased 3‘H-prolactin in the anterior pituitary. Electron micrographic examination of these pituitary cells demonstrated that following dopamine treatment there was a build-up of secretory granules containing prolactin in the rough endoplasmic reticulum (MacLeod and Lehmeyer, 1974). Once dopamine was established as having a direct prolactin inhibiting action, TIDA neurons emerged as an important link in the regulation of prolactin secretion. Anatomical studies performed by Hokfelt and Fuxe, (1972) demonstrated that TIDA neurons projected short axons that coursed ventrally from the arcuate nucleus to terminate in the external layer of the median eminence. Thus, with terminals immediately adjacent to the portal vasculature these DA neurons would be a likely source for the “prolactin-inhibiting” dopamine in the found in the portal circulation. Prolactin Feedback Regulation of TIDA Neurons A number of studies suggested that TIDA neurons participated in the Prolactin-mediated negative feedback of its release from the pituitary (for a review see Moore and Lookingland 1995). An elevation in plasma prolactin via exogenous administration or grafted anterior pituitaries leads to an increase in file synthesis of dopamine in TIDA neurons (Demarest and Moore, 1981; Morgan and Herbert, 1980; Selmanoff 1981). These findings lent support to the 32 hypothesis that feedback regulation of prolactin secretion occurs via stimulation of TIDA neurons. According to this scenario, elevations in plasma prolactin would activate TIDA neurons causing the release of dopamine, which enters hypophysial portal blood. Dopamine is then carried to the anterior pituitary where it activates D2 receptors (Creese, et al., 1977; Caron at al., 1978) and inhibits prolactin secretion from lactotrophs (Ben-Jonathan, 1977; 1985). One of the central questions remaining to be addressed is by what pathway does prolactin act upon TIDA neurons to accomplish this feedback regulation? One possibility is retrograde transport via the long hypophysial portal vessels (Olivier at al., 1977) whereby high levels of prolactin may be delivered to the median eminence and terminals of TIDA neurons. The presence of prolactin receptors in the median eminence is consistent with this hypothesis (Barton at al., 1989). Alternatively, prolactin can be taken up from the general circulation into the brain via transport through the choroid plexus (Clemens and Sawyer 1974; Walsh et al., 1978). Although the precise mechanism by which prolactin activates TIDA neurons is unknown, there is evidence to suggest that this stimulatory effect takes place within the mediobasal hypothalamus. Deafferentation of the Itlediobasal hypothalamus (Hilasz and Pup 1965) has been used as a means to Smdy hypothalamo-pituitary function independent from afferent neuronal 1Ilflllences from the rest of the brain. Indeed, this knife-cut deafferentation 33 procedure has been used successfully to determine the site of prolactin mediated activation of TIDA neurons (Gudelsky et al., 1978). Exogenous administration of prolactin is able to increase TIDA neuronal activity in rats with surgical deafferentation of the mediobasal hypothalamus suggesting that this response is not dependent upon afferent neuronal input to the mediobasal hypothalamm. Rather, prolactin acts within the area circumscribed by the knife-cut to stimulate TIDA neurons. Unlike hypophysectomy or surgical ablation of the mediobasal hypothalamus, this deafferentation procedure has no effect on basal prolactin levels. Thus, tonic inhibitory regulation of prolactin secretion is not dependent upon afferent input to the TIDA neurons from outside the mediobasal hypothalamus (Krulich et al., 1975, Turpen and Dunn 1976, Gudelsky et al., 1 978). Prolactin feedback activation of TIDA neurons is composed of “rapid tonic” and “delayed inductive” components (Demarest et al., 1984; 1986). The “rapid tonic” component regulates the set point for basal circulating prolactin levels by acutely altering TIDA neuronal activity in response to acute changes in Cil’culating prolactin. The “delayed inductive” component of prolactin feedback inVOlves a change in the capacity of TIDA neurons to respond to more Prolonged periods of elevated prolactin. Induced activation does not appear uEltil 12 hr following systemic or central administration of exogenous prolactin 34 (Hokfelt and Fuxe, 1972, Selmanoff at al., 1981; Demarest et al., 1984, 1986) and is neurotensin dependent (I-Ientschel et al., 1998). Thus, as summarized in Figure 6, TIDA neurons provide a key link in the inhibitory feedback regulation of prolactin secretion from the anterior pituitary. When serum prolactin levels are high TIDA neurons are activated and release more dopamine into the hypophysial portal circulation. Dopamine is then carried to the anterior pituitary where it reduces further prolactin secretion via an action at D2 dopamine receptors located on lactotrophic cells. 35 u ' 2 Receptor Anterior Pituitary Figure 6 Prolactin feedback regulation of TIDA neurons. Dopamine (DA) released from terminals of TIDA neurons in the median eminence (ME) is carried via hypophysial portal vasculature to the anterior pituitary. There it binds to D2 receptors on lactotrophs and inhibits prolactin secretion. Prolactin, in turn, feeds back to activate TIDA neurons via a direct action on cell bodies or axon terminals and/ or an indirect action mediated by afferent interneurons. Abbreviations: 3V third ventricle. 36 Afferent Modulation of TIDA Neuronal Activity In addition to prolactin, pharmacological evidence suggests that a number of neurotransmitters are involved in the acute afferent modulation of TIDA neuronal activity. These effects can be either stimulatory (e.g. enkephalin; neurotensin) or inhibitory (e.g. dynorphin) modulators of TIDA activity. Finally, there is accumulating evidence that the activity of these neurons may be subject to DA-receptor—mediated stimulation or inhibition. The goal of the studies contained herein is to examine in detail the DA receptor mediated mechanisms whereby the activity of TIDA neurons is altered and the relationship, if any, to the aforementioned stimulatory and inhibitory modulators. Details will be discussed in the introductions of the respective chapters. Dopamine Receptor Mediated Regulation of TIDA Neurons The responses of DA neuronal systems to a variety of DA agonists and antagonists have been intensively studied. First generation DA agonists such as apomorphine and bromocriptine, and antagonists such as haloperidol acutely modulated the activity of mesotelencephalic and hypothalamic DA neurons except TIDA neurons. DA agonists decrease the firing rate (Bunney et al., 1973), and the synthesis, release and metabolism of dopamine in these neurons (Carlsson, 1975; Roth et al., 1976; Lookingland and Moore, 1984; Imperato and DiChiara, 1985; Imperato et al., 1988) whereas DA antagonists increase these 37 paramaters of neuronal activity. These responses to DA agonists and antagonists are thought to be due to an action at presynaptic DA inhibitory autoreceptors and at postsynaptic DA receptors involved in inhibitory neuronal feedback loops. The regulation of TIDA neurons represents a departure from other mesotelencephalic or hypothalamic DA neurons in that these neurons are not subject to DA receptor-mediated mechanisms. This was established by observations that acute administration of first generation DA agonists (apomorphine, bromocriptine) and antagonists (haloperidol) had no acute effect on the activity of TIDA neurons (Gudelsky and Moore, 1976; 1978; Demarest and Moore, 1979; Lookingland and Moore 1984). Although these neurons are responsive to high doses of these compounds, these responses are delayed and are mediated by prolactin (Moore and Lookingland 1995). Given the numerous observations that DA agonists have no acute effect on the activity of TIDA neurons it was surprising when subsequent studies using the partial ergoline-derived agonists quinpirole and quinelorane (i.e., second generation DA agonists) were shown to activate TIDA neurons (Berry and Gudelsky, 1991; Eaton et al. 1993). This stimulatory effect is present even in female rats where circulating levels of prolactin provide tonic stimulatory feedback onto TIDA neurons (Demarest et al. 1985). In females, the stimulatory effect of quinelorane is masked by decreases in plasma prolactin, which in turn 38 reduces tonic stimulation of TIDA neurons. The net result is no change in the activity of these neurons (Eaton et al., 1993). The ability of quinelorane to induce increases in the activity of these neurons is only apparent when circulating prolactin is immmoneutralized by systemic administration of a prolactin antibody. In these prolactin—neutralized female rats basal TIDA neuronal activity is significantly lower than in normal rabbit serum-treated controls. Under these conditions the stimulatory effect of quinelorane can be observed and is qualitatively similar to that seen in male rats (Eaton et al. 1993). In the studies presented in this dissertation the male rat has been used, in part, to avoid the complications of the tonic stimulatory effect of circulating prolactin on TIDA neurons in female rats. The mechanism by which the D2 agonists such as quinelorane (Foreman at al 1989) activate TIDA neurons appears to be DA receptor mediated, as this stimulatory effect is blocked by second-generation DA receptor antagonists such as raclopride (Eaton et al. 1992; 1993). However, it is not known if the quinelorane-induced activation of TIDA neurons results from its ability to activate D2 or D3 receptors since in vitro studies have confirmed that this agonist binds to both of these receptor subtypes (Gackenheimer et al. 1995). The development of PNU-95,666 a selective D2 agonist (Smith et al. 1996) and PD128907 a selective D3 agonist (Pugsley at al. 1995), i.e., third generation DA agonists, have made it possible to determine the extent to which either the D2 or 39 D3 receptors contribute to the stimulatory effects of quinelorane on TIDA neurons. Stimulatory DA receptor-mediated afferent neuronal regulation of TIDA neurons could occur either by increasing the release of an endogenous stimulator of these neurons or by suppression of a tonically active inhibitory system (disinhibition). This hypothesis is attractive due to the abundant evidence that activation of D2 receptors is coupled to inhibitory cellular events: reduction of CAMP production, or calcium conductance, or increases in potassium conductance (jackson and Westlind-Danielsson 1994). Thus, a mechanism involving disinhibition is invoked to explain how inhibitory D2 receptor activation would induce a net increase in TIDA neuronal activity. For this disinhibition hypothesis to be correct two basic criteria must be met: (1) some neuronal circuit must tonically inhibit TIDA neurons, and (2) exogenous administration of the inhibitory neurotransmitter or its analog should block the Dz-mediated increase in TIDA neuronal activity. At least two inhibitory neurotransmitters; dynorphin and ‘Y-aminobutryic acid (GABA) via actions at kappa/0P2 and GABAA receptors respectively meet this first criterion. The discovery of multiple DA receptor subtypes (Civelli et al., 1993) has spawned the development of more selective DA receptor ligands in order to better assess the functional role of each of the various DA receptor subtypes, m (I EEC- and to identify new drugs that may have potential for treating neuropsychiatrtric disorders. Among these compounds are the substituted benzazepines SKF38393 (Sibley et al., 1982) and SCH39166 (Chipkin et al., 1988) with selectivity for the D1-like receptors. In contrast to the stimulatory effects on TIDA neurons observed after activation of D2-like receptors, the D1 receptor agonist SKF38393 can inhibit activated TIDA neurons (Berry and Gudelsky, 1990). Indeed, SKF38393 is able to inhibit TIDA neurons that had been activated following a variety of treatments: neurotensin, reserpine or haloperidol. Thus, the lack of effect following administration of non-selective agonists and antagonists may be due to opposite effects at D1 (inhibitory) and D2 (stimulatory) receptors. 41 Specific Aims The overall goal of studies described in this dissertation is to characterize the mechanisms underlying DA receptor-mediated regulation of TIDA neurons, with special emphasis on the roles of inhibitory D1 and stimulatory D2 receptors in this process. For comparison, the responses of DA neurons comprising the nigrostriatal and mesolimbic systems will also be examined in key experiments. An underlying theme in these studies is that anatomically distinct populations of DA neurons are differentially regulated by DA receptors. The development of DA agonists and antagonists that are highly selective for D1 and D2 receptor families has made it possible to distinguish these differences. A combination of molecular, neurochemical, neuroanatomical and pharmacological experimental approaches will be used to address the following specific aim$: 1. Determine the D2 receptor subtype that mediates the quinelorane- mediated activation of TIDA neurons Determine if quinelorane-induced activation of TIDA neurons is mediated by stimulatory delta/0P1 or neurotensin receptors Determine if quinelorane—induced activation of TIDA neurons is mediated by the release of an endogenous stimulator of TIDA neurons Determine if quinelorane-induced activation of TIDA neurons is mediated by inhibitory kappa/0P2 receptors Characterize D1 receptor-mediated regulation of TIDA neurons 42 6. Determine if the activation of inhibitory D1 receptors opposes the stimulatory effects of quinelorane on TIDA neurons. 43 Chapter Two General Methods Animals All experiments were performed in male Long-Evans rats purchased from Harlan Breeding Laboratories (Indianapolis, IN) and maintained in a temperature- (22i1°C) and light-controlled room (lights on between 06:00 h and 20:00 h) with food (T eklad Rodent Diet) and tap water provided ad libitum. Drugs Drugs are referred to by their generic names throughout the text. A listing of chemical names, vehicle solvents and sources for these compounds can also be found in Table I of the APPENDIX. Dosages of all drugs are calculated based on either the salt or free base. Routes and times of drug administration are indicated in the legends of the appropriate figures. Surgical Procedures Intracerebroventricular Cannulation Rats receiving intracerebroventricular (i.c.v.) injections were anesthetized via intraperitoneal (imp) administration of 3 ml Equithesin/ kg body wt. ' Once anesthesia was evident the surgical site was shaved and cleaned with betadine scrub and 75% ethyl alcohol. Rats were then assessed for surgical anesthesia by 44 pinching a toe; a lack of reflex movement was judged as sufficient anesthesia to begin the procedure. Rats were then placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA) and stainless steel guide cannulae were implanted into right lateral cerebral ventricle 5 days prior to the experiment. For this procedure the incisor bar on the stereotaxic frame was set 2.4 mm below the intraaural line (Konig and Klippel, 1963). A 23-gauge stainless steel guide cannula was implanted 1.4 mm lateral to the midline at bregma and lowered 3.2 mm below the dura matter surface. This guide cannula was then anchored to the skull with stainless-steel screws and denture acrylic. On the day of the experiment, i.e.v. injections were carried out by connecting a 10 pl Hamilton microsyringe to a 30-gauge stainless steel injector which protruded 1 mm below the tip of the guide cannula and into the right lateral ventricle. Drugs or their vehicles were manually injected at an approximate rate of 3-5 pl per minute (total volume 3-5 Ill). The placement of each cannula was verified histologically and only those animals with a cannula track in the right lateral ventricle were included in the data analysis. Orchidectomy Orchidectomy was performed in some experiments to assess the role, if any, of gonadal hormones on the effects of pharmacological manipulations on TIDA neuronal activity. Rats were anesthetized with diethylether seven days 45 prior to the experiment and the surgical site was shaved, and scrubbed with betadine and 75% ethyl alcohol. Testicles were removed by making a single incision along the midline of the lower abdomen through the skin and muscle layers. The testicles were exteriorized, ligated with silk suture material and then excised. The muscle layer was closed with sutures and the skin was closed with wound clips. Neurochemical estimation of neuronal activity Within axon terminals of DA neurons there is tight coupling between synthesis, release and metabolism to insure that adequate levels of releasable neurotransmitter are maintained (Figure 7). Dopamine is synthesized from dietary tyrosine, which is then converted to DOPA by the rate limiting enzyme tyrosine hydroxylase. DOPA is then rapidly decarboxylated by the action of L- aromatic amino acid decarboxylase thereby forming a pool of cytoplasmic dopamine which is either packaged into storage vesicles for future release or metabolized by monoamine oxidase to form 3,4-dihydroxyphenylacetic acid (DOPAC). This coupling between dopamine release, and de novo synthesis and metabolism affords two methods to neurochemically estimate DA neuronal activity. The first method measures the concentration of the DOPAC in terminal regions of DA neurons as an index of neuronal activity. Increased neuronal activity (synthesis and release) is reflected as increases in free intra-neuronal 46 dopamine. Monoamine oxidase has access to the dopamine in these unprotected non-vesicular pools and converts it to DOPAC. Procedures which increase or decrease DA neuronal activity will produce a concomitant change in the concentration of this metabolite (Lookingland et al., 1987). Post-Synaptic Tyrosine-TH->DOPA-DO>DA ® Cell Body or Dendrite Tyrosine I 1 If. - Tyrosine-TH->DOPA-Do; W—> DA w Figure 7 Neurochemical events in axon terminals of central dopaminergic neurons. Dashed lines represent inhibitory feedback loops. Abbreviations: DA (dopamine), DC (DOPA decarboxylase), DOPA (3,4~dihydroxyphenylalanine), DOPAC (3,4~dihydroxyphenylacetic acid), MAO (monoamine oxidase), and TH (tyrosine hydroxylase). Median Eminence Anterior Pituitary The second method of indexing neuronal activity involves estimating the synthesis of dopamine. Synthesis of new dopamine can be estimated by measuring DOPA accumulation following administration of an L-aromatic amino acid decarboxylase inhibitor (Carlsson et al., 1972). This estimate is based on the tight coupling between tyrosine hydroxylase activity and its end-product inhibition by intra-neuronal free dopamine. Thus, following DOPA 47 decarboxylase inhibition procedures which increase or decrease neuronal activity produce a corresponding change in the rate of DOPA accumulation (Demarest et al, 1981). In experiments described in this dissertation the concentration of DOPAC or accumulation of DOPA 30 min after decarboxylase inhibition will be measured as indices of DA neuronal activity. Tissue and Blood Processing Following in viva manipulations rats were killed by decapitation. Brains and pituitaries were rapidly removed from the skull, frozen on aluminum foil placed over dry ice. Frontal brain sections (600 um thick) were prepared in a cryostat at -9°C. Ten adjacent serial sections were collected beginning at approximately 2.2 mm anterior to bregma and ending at approximately 3.8 mm posterior to bregma (Pardnos and Watson 1986). These sections were placed on glass slides and re- frozen. The slides were then transferred to the stage of a stereo microscope and the nucleus accumbens, striatum and median eminence were dissected using a modification (Lookingland and Moore, 1984) of the method of Palkovits (1973). The nucleus accumbens was dissected from sections #1 and #2 by means of a 21 gauge hypodermic needle, the striatum from section #2 using an 18 gauge hypodermic needle (Figure 8), and the median eminence from sections #8 and #9 using an oval shaped 18 gauge needle (Figure 9). The pituitaries were dissected according to the method of Lookingland et al., (1985). An 18 gauge 48 needle was used to first remove the neural lobe and then the intermediate lobe was collected by separating it from the underlying anterior pituitary (Figure 10). Tissue samples were placed in 100 ill (nucleus accumbens and striatum) or 50 pl (median eminence and intermediate lobe of the pituitary) of a 0.1 M phosphate- citrate buffer (pH 2.5) containing 15% methanol and stored at -20°C until assayed. 49 SE CTION #2 Figure 8 The anatomical locations of the nucleus accumbens (shaded), and striatum (cross hatched) tissue samples dissected for neurochemical analysis. Each frontal section was 600 11m thick and section #1 was from approximately 2.2 mm and section #2 was from approximately 1.6 mm anterior to bregma respectively. Figures were modified from Paxinos and Watson, 1986. 50 SECTION #9 Figure 9 The anatomical location of the median eminence (cross-hatched) tissue sample dissected for neurochemical analysis. Each frontal section was 600 [1m thick and section #8 was taken from approximately 2.56 mm and section #9 from approximately 3.3 mm posterior to bregma respectively. Figures were modified from Paxinos and Watson, 1986. 51 v o'o’o‘o » “vs 3 . ‘v's l nte I'm ed I ate v o'o’o’o’o’t’o'o’o'o'o'o'o'o'o'o'o'o’o’o"9 - ° - satAt.tsVIN)”.fifololofofofofotozoto}, P IIU 'tary L0be .. . A\, Anterior Pituitary Figure 10 A sagittal section through the pituitary gland showing the anatomical location of the neural lobe of the pituitary (shaded) that was removed in order to allow dissection of the intermediate pituitary lobe (cross hatched) for neurochemical analysis (Modified from Paxinos and Watson, 1986). 52 High Performance Liquid Chromatography On the day of the assay, tissue samples were thawed, sonicated for 3 s (Sonicator Cell Disruptor, Heat Systems-Ultrasonic, Plainview, NY) and centrifuged for 30 s in a Beckman 152 Microfuge. Contents of DOPAC, DOPA, dopamine and norepinephrine in supernatants were determined by high performance liquid chromatography coupled with electrochemical detection as described previously (Lindley et al., 1990). The amounts of these catecholamines in tissue samples were measured by comparing peak heights (determined by a Hewlett Packard Integrator, Model 3395) with those obtained from external standards run on the same day. The standards contained a mixture of the above analytes each at a concentration of 0.5 or 1.0 ng/50 111. The lower limit of sensitivity of this assay for these compounds was 1-5 pg/ sample. Fifty microliters of each sample supernatant was injected onto a C18 reverse-phase analytical column (Slim spheres; 250m in length; Biophase ODS, Bioanalytical systems, West Lafayette IN) preceded by a pre—column filter cartridge (5 [1m spheres; 30mm). Following separation analyte signals from the nucleus accumbens, striatum, or median eminence samples were determined using a electrochemical detector (LC4A; Bioanalytical systems, West Lafayette IN) equipped with a TL-5 glassy carbon electrode set at +0.75 V relative to an Ag/AgCl reference electrode. Because samples from the intermediate lobe of 53 the pituitary contain lower levels of these catecholamines, they were analyzed using a single coulometric electrode conditioning cell in series with dual electrode analytical cells (ESA, Bradford MA). The conditioning cell electrode was set a + 0.4 V and the analytical electrodes were set at +0.12 V and — 0.4 V respectively, relative to the Ag/AgCl reference electrode. The high performance liquid chromatography mobile phase consisted of a 0.1 M phosphate/ citrate buffer (pH 2.65) containing 0.1 mM ethylenediaminetetraacetic acid, 0.03-0.045°/o sodium octylsulfate and 15 - 25% methanol. Depending on the condition of the column, components of the mobile phase were adjusted slightly to maintain separation of the analytes of interest and to minimize total retention times (Chapin et al., 1986). Tissue pellets from each sample were dissolved in 100 or 200 ill of 1.0 N sodium hydroxide depending on the size of the tissue punch and assayed for protein (Lowry et al., 1951). All neurochemical data have been reported as concentrations (ng/ mg protein) to correct for variability in tissue sample size. Radioimmunoassa y for Prolactin In some studies plasma prolactin measurements were made as an indirect physiological index of dopamine release from TIDA neurons, or to determine the effect of various pharmacological manipulations on D2 receptors in the pituitary gland. Immediately after rats were decapitated, trunk blood was 54 toicc 310:6: PAM tbes :I: 24 collected into glass tubes containing heparin and 150 mg of bacitracin, and stored on ice. Following collection, tubes were centrifuged for 30 min at 4°C, plasma was separated from red cells and stored in sodium citrate containing tubes at -20°C until the day of the assay. Plasma prolactin concentrations were determined by a double-antibody radioirnmunoassay. This assay made use of reagents and procedures supplied by Drs. A. F. Parlow and S. Raiti, NIADDK National Hormone and Pituitary Program. NIDDK rat prolactin (RP—3) was use as the standard and rat prolactin (rat PRL-L6; NIDDK) was labeled with 125Iodine (Amersham Arlington Heights, IL) using the chloramine-T technique (Chard, 1990). Rat plasma samples were incubated with anti-rat prolactin (anti-rat PRL S-9 NIDDK) and 125I-labeled prolactin at 4° C for 24 hours. On the next day the precipitating secondary antibody (goat anti-rabbit gama globulin) was added to the samples and incubated for an additional 24 hours at 4° C. The following day the mixture was centrifuged, the supernatant decanted and the precipitate was counted in a gamma counter (Micromedic model 4/ 600). Using a 100 pl aliquot of plasma the lower limit of sensitivity for prolactin in these assays was 100—200 pg/ sample, and the inter-assay coefficient of variability was approximately 10%. 55 Measurement of Dynorphin mRNA by in situ Hybridization Histochemistry In catecholaminergic neurons the cellular machinery for synthesis, storage and release of neurotransmitters is located in the axon terminals. Thus, in neurochemical determinations of neuronal activity it is useful to make measurements by sampling tissue from a brain region corresponding to the axon terminal field projections of the neurons of interest. However, peptidergic neurotransmitters are synthesized in neuronal perikayra. Using sensitive immunohistochemical methods it is possible to qualitatively determine the presence of such neuropeptides and determine their precise location. However, it is exceedingly difficult to make quantitative assessments that reflect changes in neuronal activity in a way that was described above for DA systems. In situ hybridization histochemistry is a method whereby the mRNA transcripts for a particular protein can be measured as a quantitative index of the activity of specific peptidergic neurons. The underlying theory is that changes in neuropeptide release from axon terminals are accompanied by corresponding changes in do: nova synthesis of mRNA transcripts in the neuronal perikarya. Thus, in situ hybridization provides anatomical and quantitative information about the activity of peptidergic neurons. 56 COII GG Gil tit 193: i"): 5 . fdcl Labeling The mRNA Probe The sequence of the 47 base synthetic oligomer probe was complementary to bases 862-909 of the rat prodynorphin gene [5’GC TAT GGG GGC TI’C CT G CGG CGC ATT CGC CCC AAG CTT AAG TGG GAC3’] (Civelli at al., 1985). Purified oligodeoxynucleotides were 3’ end-labeled with 3SS-dATP using terminal deoxynucleotidyl transferase (T (H) (Young at al., 1986). Briefly, 1 111 of 5 11M probe was reacted with 4 ill of TdT enzyme (25U/tll; Boehringer Manheim; Indianapolis, IN), 10 [ll of 5X DNA tailing buffer containing 0.5 M potassium cacodylate, pH 7.2, 10 mM cobalt chloride and 1 mM dithiothreitol, 5 ill of 35 S—dATP (1389 Ci/mmol; New England Nuclear; Boston MA) and 30 [ll sterile diethyl pyrocarbonate (DEPC)-treated water for 5 min at 37°C. Extraction of the Labeled Probe The labeling reaction was stopped by addition of 400 ill of Tris-EDTA (TE) buffer containing 10 mM Tris HCL/EDTA, pH 7.5 and 2 [1.1 yeast tRNA (25 mg/ ml) The labeled probe was extracted with the addition of 400 ml phenol/chloroform/isoamyl alcohol (50:49:1). The mixture was vortexed then centrifuged and the recovered aqueous phase was extracted with the addition of 400 ml of chloroform : isoamyl alcohol (24:1). The recovered aqueous phase was then precipitated with the addition of cold absolute ethyl alcohol containing 57 0.3M sodium acetate on ice for 45 min and then centrifuged for 45 min at 4°C. The supernatant was discarded and the pellet was allowed to air dry. The air dried pellet was dissolved in 100 ml TE buffer containing 1 [ll of 1 M dithiothreitol (DTT) and stored at 4°C until the hybridization procedure was done. Preparation of Tissue Following in viva experimental manipulation rats were euthanized by decapitation, brains were removed and rapidly frozen on aluminum foil covered dry ice (see page 48). Frontal sections (25 m thick) through the arcuate nucleus in these fresh frozen brains were prepared in a microtome cryostat (IEC Minotome, International Equipment Corp.) and affixed directly to gel-coated slides. These slides were then placed in a vacuum/ desiccator at 4°C to allow them to dry. Slides were stored in boxes containing desiccant at -80°C until the day of the hybridization procedure. All solutions used in pre-hybridization are made using sterile DEPC-treated water. On the day of hybridization slides were warmed to room temperature loaded into autoclaved racks and fixed for 5 min using 4% formaldehyde in sterile phosphate buffered saline (PBS; pH 7.4). Slides were then rinsed 2 times with PBS to re-hydrate the tissue. Once re- hydration was complete tissues were treated with acetic anhydride 0.25% in 0.1 M triethanolamine (TEA) / 0.9% saline (pH 8.0) for 10 min. Acetylation of the 58 tissue via acetic anhydride reduces non-specific binding. Sections were then dehydrated in serial ethanol solutions (70% 80% 90% and 100%), de-fatted for 5 min in absolute chloroform, re—hydrated in graded ethanol solutions (100% and 95%) and air dried. In Situ Hybridization Procedure Labeled probe was added to hybridization buffer containing: 50 % formamide, 500 ug/ ml sheared ss DNA (Sigma Chemical St. Louis, MO), 250 mg/ ml yeast tRNA, 4X saline-sodium citrate (SSC), 1X Denhardt’s solution, and 10% dextran sulfate to reach a concentration of 20,000 cpm/ [.11. Approximately 25 ul of this hybridization solution was added to each section and incubated under a sterile coverglass on humidified trays at 37°C for 18 hr. After hybridization, coverslips were removed and slides rinsed in 1X SSC and then washed in each of the following: 50% formamide in 2X SSC at 40°C (4 X 15 min), 1X SSC at room temperature (2 X 30 min). Tissues were then dehydrated in serial ethanol solutions (50, 70, 90 and 95%) containing 300 mM arnoniurn acetate then in 100% ethanol. Air dried slides were then dipped in NTB-Z liquid photographic emulsion (Eastman Kodak; Rochester, NY) diluted 1:1 with water for autoradiographic analysis. Slides were exposed at -20°C for 7 days. After development slides were rinsed, counter-stained with nissil stain, dehydrated and coverslipped to facilitate 59 regional localization of exposed silver grains in the photographic emulsion. A summary of the in situ hybridization histochemistry procedure is depicted in Figure 11. Dynorphin (DYN) oligonucleotide probe 19:17 + TDT + [358] *dATP* i .. = ‘ _ A*A*A* Label DYN oligonucleotide probe with Cut sections and mount terminal deoxynucleotidyl transferase onto gel-coated slides Apply probe solution onto slides, hybridize overnight at 37°C Dissolve probe in hybridization buffer /@ @/ Post hybridization 63 63 A/wash and dehydrate Dip slides in photographic emulsion (NTB—2), expose and develop Figure 11 Scheme for in situ hybridization with synthetic oligonucleotides. 60 Image Analysis Specific labeling of dynorphin mRNA was visualized using dark field optics (Micro Video Instruments, Avon MA) connected with a Leitz LaborLux S microscope (Leica, Germany). Quantitative analysis of autoradiographic grain density was done using the microscope mounted video camera model CCD-725 (Dage-MTI, Michigan City, IN) coupled to a LG3 video card (Scion Corp. Frederick MD) on a Macintosh 8800 PowerPCTM computer. NIH image (ver. 1.6) and an automated grain counting macro was used to quantify the number of exposed silvergrains in the emulsion overlying brain sections. The labeling ratio of a cell was defined as the ratio of exposed silver grains atop its perikarya to the background number of exposed silver grains in an adjacent cell-sized area of the neuropil (Popper at al., 1992a; 1992b). An individual cell was considered to be specifically labeled when the labeling ratio was greater than or equal to three (Figure 12). Slides were coded such that the treatment groups were not known until after the data analysis was complete. Brain sections at approximately the same level in the hypothalamus -2.30 mm anterior/ posterior to bregma (Paxinos and Watson 1986) were chosen macroscopically. Ten specifically labeled cells in each section were counted and background determinations were made to obtain a 61 labeling ratio for each cell. These ten determinations were then averaged to obtain a labeling ratio for each brain (11 = 6 / treatment group). 62 Figure 12 Computer captured micrograph depicting specific labeling of a cell containing dynorphin mRNA in the arcuate nucleus. The arrow indicates a cell specifically labeled by the dynorphin mRNA probe observed using both darkfield optics (left panel) and brightfield optics right panel. 63 Dual lmmunohistochemical Determination of Folera and TH The c-fos gene encodes the F 03 transcription factor which is involved with the modulation of gene activity in response to cell surface signals (Curran and Franza 1988). Although neuronal Fos is expressed at very low levels under baseline conditions, the level of this protein peaks 60 to 90 min after stimulation and the magnitude of its induction reflects the number of activated neurons recruited by the stimulus (Lee at al., 1992; Verbalis at al, 1991). Fos/Fra expression complements neurochemical estimates of TIDA activity since changes in the synthesis or metabolism of dopamine in the median eminence do not provide information about the responsiveness of distinct subpopulations of TIDA neurons. The arcuate nucleus has been subdivided into two distinct populations of tyrosine hydroxylase immmoreactive (PH-IR) neurons (dorsomedial and ventrolateral). These distinctions are made based upon phenotypic differences in size, location of their perikarya and neurochemistry (Everitt at al., 1986). In studies described in chapter 7 of this dissertation, the activity of these two groups of TH positive cells were compared by measuring the expression of Fos/FRA in TH-immunoreactive cell bodies located in the dorsomedial and ventrolateral arcuate nucleus (Figure 13). 64 it: mi. buff pm the: m . BID biou B. Tissue Preparation The numbers of TH-IR neurons containing Fos/FRA-IR nuclei were determined using dual irnmunohistochemistry. After drug administration, rats were anesthetized with Equithesin (3 ml/ kg; i.p.) and perfused through the aorta with 0.9%saline (4°C), followed by 4% paraformaldehyde in 0.1 M phosphate buffer. Brains were removed from the skull and post-fixed overnight in 4% paraformaldehyde buffer. The brains were then cryoprotected by immasing them in 0.1 M phosphate buffer containing 20% sucrose at 4°C for 2448 hours. Frontal sections (36 um) through the medial arcuate nucleus (Cheung at al., 1997a) were placed in 24 well cell culture plates for immunohistochemical staining procedures. Dual Immunohistochemistry These sections were incubated in a primary sheep anti-Fos oncoprotein antiserum (1 :2500; Cambridge Research Biochemicals) for 40 hours at 4°C. The primary antiserum was localized using donkey anti-sheep antibody, the avidin- biotin complex (Vector Laboratories) and diaminobenzadine subtrate with nickel-intensification. The sections were then incubated in primary mouse anti- TH antiserum (1:1000; Incstar Co.) for 40 hours at 4°C. The primary antiserum was localized with horse anti-mouse antibody and rhodamine—tagged avidin 65 (Vector Laboratories). Thus, double-labeled neurons appear as red cells (T H) with dark nuclei (F 05 / F RA) depicted in Figure 14). 66 Figure 13 Computer—captured image of the ME and ARC in a frontal brain section. Panel A.: Depicts the distribution of TH-IR neurons in the ARC located at 2.6 mm posterior to bregma (Paxinos and Watson, 1986). 3V (third ventricle). The inset is shown as panel B. Panel B: Computer captured image showing the subdivision of the ARC into dorsomedial (DM) and ventrolateral (VL) regions. 67 68 Figure 14 Computer-captured micrograph depicting a frontal brain section immunocytochemically labeled for both TH and Fos/FRA. Panel A: Bright-field depiction of a computer captured image showing Fos/FRA-IR neuclei Panel B: The same computer- captured image using fluroescence to depict TH-IR neurons. Arrows correspond to cell bodies where TH- and Fos/FRA-IR are co-localized in the DM-ARC.. Arrows correspond to perikaryia in which Fos/FRA-IR and TH-IR are co-localized in the DM-ARC. 3V, third ventricle. 69 70 it w L 1? mt U11 ii: for pt: Statistical Analyses In neurochemistry experiments data were tested to determine if they met the basic requirements of an analysis of variance (AN OVA). That is, that these data fell into a normal distribution and demonstrated equality of variance between groups. Data not normally distributed were analyzed by a non- parametric analysis of variance using Kruskal—Wallace and Dunn’s test for multiple compansons. Data that violated the assumption of equal variance were transformed using a logarithmic transformation to reduce the between groups variance. Data that were inherently non-parametric (i.e., exposed siver grain counts for in situ hybridization procedures or cell counts in immunohistochemical procedures) a square root transformation was performed to convert the data to a normal distribution and / or analyzed by non-parametric procedures. When only two treatment groups were compared significant differences were determined by a Students t—test or if there were more than two groups an AN OVA was used to determine significant differences among treatment groups. Differences were considered significant if the probability that the null hypothesis was rejected due to error was less than 5%. 71 [1le iii“ teu AUDI it: he in m (Cg-.3 ‘ u gr‘ (abet Chapter Three 0;. Receptors Mediate the Stimulatory Effects of Ouinelorane on Tuberoinfundibular DA Neurons The lack of an inhibitory response to quinelorane in TIDA compared to mesolimbic DA neurons is not surprising since it has been well established that TIDA neurons lack the DA autoreceptor-mediated responses (Demarest and Moore, 1979) characteristic of other mesotelencephalic and hypothalamic DA neuronal systems (Lookingland and Moore, 1984). Due to the presence of these autoreceptors, the activity of most DA neuronal systems is inhibited following administration of the DA agonist apomorphine and activated when exposed to the DA antagonist haloperidol. The novel stimulatory effect induced by quinelorane (Eaton et al., 1993) is intriguing in that it suggests a heretofore- unrecognized mechanism whereby the activity of TIDA neurons can be regulated by DA receptors. Although quinelorane has an advantage over first generation compounds of greater selectivity for D2 versus D1 receptor families, it has been shown to bind avidly to both D2 and D3 receptor subtypes within the D2 subfamily (Gackenheirner et a1 1995). Thus, it is not clear which receptor subtype is mediating the response to quinelorane. Newer, even more selective agonsts have been developed in an effort to better distinguish the functional roles of D2 and D3 receptors. PNU-95,666 72 01 do [0' 51!: iii To in: HQ mi exhibits a high degree of selectivity for the D2 receptor compared to the D1, D3, or D4 receptors (Smith et a1 1996) whereas, PD128907 has an approximately 18 fold greater affinity for the D3 versus D2 receptor (Pugsley et a1 1995), and at dosages up to 300 pg/ kg this compound acts selectively at D, receptors (Pugsley at al 1995; Bristow at al 1996). The availability of these more selective agonists could be used to determine which DA receptor subtype is mediating the quinelorane-induced stimulation of TIDA neurons. Administration of selective D2 or D3 antagonists in conjunction with quinelorane could be an alternative strategy to determine which DA receptor subtype mediates the stimulatory action of quinelorane. Since these pharmacological tools were not available selective agonists were used in experiments described in this chapter. The purpose of these experiments was to establish the DA receptor subtype that mediates the stimulatory response of TIDA neurons to quinelorane. To this end, the effects of PNU-95,666 and PD128907 on the activity of TIDA neurons were compared in order to determine which selective agonist most closely mimicked the stimulatory effect of quinelorane. For comparative purposes the effects of these compounds on the activity of mesolimbic DA neurons and plasma prolactin concentrations were also assessed. The hypothesis to be tested is that the stimulatory effect of quinelorane on TIDA neurons is 73 1111 DR PX mediated by D2 receptors. If this is the case then PNU-95,666 should generate a similar response to quinelorane, whereas PD128907 should have no effect. Methods Care and handling of the rats, and neurochemistry procedures for determining the dopamine synthesis and metabolism in median eminence and nucleus accumbens were carried out as described in the General Methods. The radioimrnunoassay for plasma prolactin was carried out as described above in the General Methods. The lower limit of detection in this assay was 120 pg/ 100 pl of plasma and the intraassay variability was approximately 6%. Drugs Quinelorane (Dr. M. Niedenthal The Eli Lilly Company; Indianapolis IN), PNU-95,666 (Dr. P. VonVoigtlander, The Pharmacia-Upjohn Company, Kalamazoo, MI), and 3—hydroxybenzylhydrazine dihydrochloride (NSD 1015, Sigma Chemical Company, St Louis, MO, USA) were dissolved in 0.9 % saline. PD128907, (Dr. M. D. Davis, Parke Davis Pharmaceutical Research, Ann Arbor, MI) was dissolved in distilled water. Drugs were administered as indicated in the legends of the appropriate figures; doses of drugs were calculated as the respective salts. 74 R. 10 C0 Results The differential effects of incremental doses of quinelorane on DOPAC concentrations in the median eminence and nucleus accumbens of male rats are depicted in Figure 15. Quinelorane elicited a dose-related increase in DOPAC concentrations in the median eminence, with significant increases attained at doses of 50 and 100 pg/ kg. On the other hand, quinelorane reduced DOPAC concentrations in the nucleus accumbens, even at the lowest dose employed (10 ug/ kg) In contrast, quinelorane had no effect on dopamine concentrations so that the ratio of DOPAC to dopamine increased in the median eminence and decreased in the nucleus (T able 2). The hallmark of drugs with D2 agonist properties is inhibition of prolactin secretion from the anterior pituitary (Ben-Jonathan, 1985). Consistent with the effect of other D2 agonists 10 mg PNU-95,666/ kg significantly reduced plasma prolactin levels by 30 min and this effect lasted at least 120 min (Figure 16). The D3 receptor selective dose of PD128907 (300 ug/ kg, Pugsley at al, 1995), failed to alter plasma levels of prolactin. However, PNU-95,666 induced a dose-related increase in DOPAC levels in the median eminence and reduced DOPAC concentrations in the nucleus accumbens (Figure 17). PNU-95,666 failed to alter dopamine concentrations in these brain regions (Table 3) so that DOPAC to dopamine ratios increased in the median eminence and decreased in the nucleus 75 accumbens. In contrast, PD128907, the selective D3 agonist, failed to alter DOPAC (Figure 17) or dopamine in the median eminence or nucleus accumbens. Thus, DOPAC to dopamine ratios were unchanged in these regions as well (Table 4). 76 MEDIAN EMINENCE - 1:: Vehicle ’5 — Quinelorane ’63 * § 21 - * Q. E’ 3, 14 - T 5 O < o. 7- O o O VEH 10 50 100 49 ' NUCLEUS ACCUMBENS e 42 - E; T 2 35 - * * 0' * “E” 28 - B: 5 21 - 2 a. 14 - O o 7- 0 VEH 10 50 100 Figure 15 Dose response effects of quinelorane on DOPAC concentrations in the median eminence and nucleus accumbens of male rats. Rats were injected with various doses of quinelorane (10, 50 or 100 pg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. Columns represent the means and vertical lines 1 SEM of 8 determinations of DOPAC concentrations in the median eminence and nucleus accumbens of vehicle- (open columns) or quinelorane-treated (solid columns) rats. *p < 0.05, values from quinelorane-treated rats that are significantly different from vehicle-treated controls 77 Table 2 Effects of quinelorane on dopamine and DOPAC/ DA ratios in the median eminence and nucleus accumbens. Rats were injected with various doses of quinelorane (10, 50 or 100 ug/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. Values in each cell represent the means i 1 SEM of 8 determinations. *p < 0.05, values from quinelorane-treated rats that are significantly different from vehicle-treated controls. Median Eminence Nucleus Accumbens gfidgfk; Dopamine Doggon Dopamine 001122431321 Vehicle 124.7 i 7.7 0.11 r 0.02 80.2 :t 2.5 0.46 i 0.02 10 141.3 i 8.7 0.11 :1: 0.01 93.1 r 3.5 0.30 r 0.01* 50 128.1 i 5.8 0.14 i: 0.01 95.2 i: 5.0 0.33 :l: 001* 100 125.9 :1: 6.9 0.15 :t 001* 85.2 :1: 6.7 0.35 :t: 001* 78 10- A PNU-95,666 o PD-128907 a 8' s U) c_u .2- 6 a s, .5 4- ‘5 E e D. 2 . 0. I; 0 30 60 90 120 240 360 480 MIN AFTER INJECTiON Figure 16 Time course of the effects of PNU-95,666 and PD128907 on prolactin concentrations in plasma. Rats were decapitated at various times after being injected with PNU-95,666 (10 mg/ kg so) or PD128907 (0.3 mg/ kg i.p.) 30 minutes after their respective vehicles (time 0). Symbols represent means and vertical lines 1 SEM. of plasma prolactin concentration in 5-7 rats; where no vertical line is depicted 1 SEM is less than the radius of the symbol. Solid symbols represent those values that are significantly different (p < 0.05) from time 0 controls. 79 ffhuflICCLc c-rPh\-rh:\ C(Qfibc 25.9:th U.E\CC» 0(10G it h h ' Oi] 14+ ll Median Eminence Nucleus Accumbens _ [:1 Vehicle , E 28 . _ PNU-95666 g 24 j § 3 20 j g; 5. 15 ‘. E c B V 12 ‘ s, 2 ‘ o (L 8 ‘ < O ‘ D. o 4 — 8 0 - _ VEH 1 3 10 VEH 1 3 1o Dose (mg/kg) :wehicle g 25 _PD128907 jg 70 (D § 20 § 60 1:. °- 50 O) O) E 15 E 40 < < n_ 5 0. 10 O 8 D o VEH 3 30 300 0°39 (“g/k9) Figure 17 Effects of PNU-95,666 and PD128907 on DOPAC concentrations in the median eminence and nucleus accumbens. Rats were decapitated 30 minutes after being injected so with incremental doses of PNU-95,666 or its vehicle, or 60 minutes after being injected i.p. with incremental doses of PD128907 or its vehicle. Columns (open-vehicle; solid-drug) represent means and vertical lines 1 SEM. of DOPAC concentrations in median eminence or nucleus accumbens from 7-9 rats. *, Values in drug—treated rats that are significantly different (p < 0.05) from vehicle-treated controls. 80 Table 3 Effect of PNU-95,666 on dopamine and the DOPAC to dopamine ratio in the median eminence and nucleus accumbens. Rats were decapitated 30 minutes after being injected s.c. with incremental doses of PNU-95,666 or its vehicle. Values represent means i 1 SEM. of dopamine, or the ratio of DOPAC to dopamine concentrations in median eminence or nucleus accumbens from 7-9 rats. *, Values in drug-treated rats that are significantly different (p < 0.05) from vehicle-treated controls. Median Eminence Nucleus Accumbens PNU-95,666 Do . DOPAC/DA DO . c DOPAC/ DA (mg/kg) Pwm‘e Ratio 1mm“ Ratio Vehicle 196.9 i 11.7 0.09 :I: 0.01 110.8 i: 2.7 0.48 :I: 0.01 1 219.4 i 13.2 0.09 :I: 0.01 108.6 :I: 5.7 0.41 :1: 0.01* 3 204.1 :1: 14.9 0.12 i 0.01* 107.9 d: 5.9 0.37 :I: 0.01* 10 190.8 i 16.1 0.12 i 0.01* 121.7 i 4.5 0.34 :I: 0.01* 81 Table 4 Lack of effect of PD128907 on dopamine and the DOPAC to dopamine ratio in the median eminence and nucleus accumbens. Rats were decapitated 60 minutes after being injected i.p. with incremental doses of PD128907 or its vehicle. Values represent means i 1 SEM. of dopamine, or the ratio of DOPAC to dopamine concentrations in median eminence or nucleus accumbens from 7-9 rats. PD128907 Median Eminence Nucleus Accumbens Vehicle 146.1 i 8.3 0.10 i: 0.01 112.9 1: 4.4 0.48 i 0.01 3 144.9 i 7.9 0.10 i 0.01 114.3 1' 2.9 0.51 :1: 0.01 30 151.4 :I: 10.5 0.09 i 0.01 122.6 1: 7.2 0.47 :i: 0.01 300 146.1 i 12.2 0.09 i 0.01 115.3 t 3.9 0.47 :I: 0.03 82 A similar pattern emerged in the time course study (Figure 18). PNU— 95,666 caused a prompt (within 30 min) and a sustained (more than 120 min) increase in DOPAC concentrations in the median eminence, while at the highest dose employed (300 ug/ kg) PD128907 failed to alter the DOPAC content in the median eminence or nucleus accumbens at any of the times examined. The lack of effect of PD128907 raised concerns that the drug sample had lost activity or was injected inappropriately. Thus, an additional experiment was designed based upon reports that PD128907 reduced the rate of DOPA accumulation in the nucleus accumbens (Pugsley et al, 1995; Bristow at al, 1996). In our hands, PD128907 selectively reduced the accumulation of DOPA in the nucleus accumbens 30 minutes after the administration of NSD1015, but had no effect on DOPA accumulation in the median eminence (Figure 19). Although there were no objective measures of motor activity made in these studies a characteristic pattern of increased motor activity and stereotyped behavior was observed in rats receiving either quinelorane or PNU-95,666 compared to vehicle control injections. On the other hand, administration of PD128907 induced a period of reduced motor activity compared to vehicle controls. 83 I“! l"..' rlrle C . n U I u '1‘ 30- A PNU-95,666 25 - O PD-128907 20- 10( MM DOPAC (11me protein) 0 30 60 90 120 240 360 480 MIN AFI'EFI INJECTION Figure 18 Time course of the effects of PNU-95,666 and PD128907 on DOPAC concentrations in the median eminence. Rats were decapitated at various times after being injected with PNU-95,666 (10 mg/ kg, s.c.) or PD128907 (0.3 mg/kg i.p.) 30 minutes after their respective vehicles (Time 0). Symbols represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence of 5-8 rats; where no vertical line is depicted 1 S.E.M. is less than the radius of the symbol. Solid symbols represent those values that are significantly different (p < 0.05) from time 0 controls. 84 Artsvuotn. 0E\0£v (0.00. Median Eminence Nucleus Accumbens 15- 70. E E ,9 12- .9 56- r 9 9 o. g 9- T g) 42- E 5. E 28- < < o. 3. CL 14. 8 8 0 0 VEH 3 30 300 VEH 3 30 300 Figure 19 Effects of PD128907 on DOPA concentrations in the median eminence and nucleus accumbens. Rats were decapitated 30 min after administration of NSD1015 and 60 min after i.p. administration of incremental doses of PD128907 or its vehicle. Columns (open-vehicle; solid-drug) represent means and vertical lines 1 S.E.M. of DOPAC concentrations in median eminence or nucleus accumbens from 7-9 rats. *, Values in drug- treated rats that are significantly different (p < 0.05) from vehicle-treated controls. 85 Dis: Hill N Discussion The results presented here reveal that quinelorane activates TIDA neurons while inhibiting other ascending DA systems such as the mesolimbic DA neurons that tenninate in the nucleus accumbens. These findings are consistent with those of Berry and Gudelsky ( 1991) and Eaton at al, ( 1993) who demonstrated that second generation DA agonists (e.g., quinpirole and quinelorane) with increased selectivity for D2 receptors induce a novel stimulatory response in TIDA neurons. On the other hand, these agonists induce the typical inhibitory response of first generation agonists (e.g., bromocriptine, and apomorphine) observed in other DA neurons via an action at autoreceptors. The present results comparing the effects of these more selective compounds with those of quinelorane reveal that the D2 selective agonist PNU- 95,666 closely mimicked the effects of quinelorane on plasma prolactin, TIDA and mesolimbic DA neuronal activity. Whereas the D3 selective agonist PD128907 was without effect on plasma prolactin or TIDA neuronal activity, but did inhibit the activity of mesolimbic DA neurons. One of the difficulties in making extrapolations about in viva receptor functionality from DA agonists that have been defined primarily by in vitm receptor binding assays, is the influence that intrinsic efficacy and receptor 86 1r Ititli' 100 these were prol RCC reserve for each of the target receptor subtypes (i.e. D2 and D,) has on functional activity in viva (Kenakin, 1993). Therefore, it is important to verify the efficacy of these selective agonists in viva if possible. To this end, plasma prolactin levels were used as an in viva measure of functional activity at the D2 receptor, since prolactin secretion is regulated by the action of dopamine on D2 receptors located on pituitary lactotrophs (Ben-Jonathan at al 1989). Thus, DA agonists with in viva activity at D2 receptors will lower plasma prolactin concentration by a direct action on pituitary lactotrophs. In these studies quinelorane and PNU- 95,666 reduced plasma prolactin concentration, while PD128907 was without effect. These results confirm that both quinelorane and PNU-95,666 but not PD128907 had in viva efficacy at the D2 receptor. The subjective behavioral observations made in these experiments are in accord with studies that have objectively reported decreases in motor activity following administration of a D3 selective dose of this drug (Pugsley at al 1995; Bristow et al, 1996). In spite of the lack of a rigorous assessment of in viva D3 receptor function these results confirm that PD128907 exerted a behavioral effect that has been attributed to the D3 receptor at a dose that had no measurable in viva efficacy at Dzreceptors (i.e. plasma prolactin concentration). The D2 selective agonist PNU-95,666 mimics the effects of quinelorane on TIDA neurons as measured by either synthesis or metabolism of dopamine in the median eminence, whereas the D3 selective dose of PD128907 (up to 300 87 mg/ kg) is without effect These findings support the hypothesis that TIDA neurons are subject to a novel D2 receptor-mediated stimulatory influence even though they lack autoreceptor—mediated inhibition characteristic of other DA neuronal systems. (Roth and Elsworth, 1996). These results also indicate that TIDA neurons lack a regulatory mechanism mediated by the D3 receptor. The D2 agonist PNU-95,666 lowered the concentration of DOPAC in the nucleus accumbens at all dosages tested and these data are consistent with an action at D2 autoreceptors on mesolimbic DA neurons. Interpretation of the neurochemical results following administration of the D3 agonist PD128907 is complicated by the discrepancy between a lack of effect on DOPAC concentration, and a significant reduction in DOPA accumulation in the nucleus accumbens. One possible explanation for the lack of effect of PD128907 on mesolimbic DA metabolism could be that the compound was injected inappropriately or had lost activity. However, observations that rats given the drug exhibited a behavioral change consistent with reported effect of the D3 agonist (i.e. decreased motor activity) argues against either of these possibilities. It therefore seems unlikely that failure to inject the drug correctly would explain the lack of effect on dopamine metabolism in these studies. This discrepancy between D, receptor-mediated effects on synthesis and metabolism of dopamine in these neurons may be explained by one or two potentially related phenomena. Autoreceptors found on DA neurons can be 88 111651 hm WEE vets add 030 dos selc inhi} didn it) .- alliOn located on cell bodies, dendrites or on axon terminals. These receptors function to maintain neurotransmitter homeostasis by regulating synthesis and/ or release of dopamine within the neurons upon which they are located (Elsworth and Roth 1997). The relative contribution of D2 and D3 receptors in the regulation of dopamine release and synthesis has yet to be demonstrated unequivocally in mesolimbic DA neurons. Studies using the D3-prefer1ing agonist 7-OH-DPAT have suggested that D3 receptors function to modulate the synthesis of dopamine in these neurons (Aretha et a1 1995). However this assertion is weakened by the observation that 7OH-DPAT may not be as selective for D3 versus D2 in vivo as determined by in vitm affinity studies (Burris at al, 1995). In addition, Seabrook and colleagues (1995) have suggested that claims of D3 receptor functionality based on selective agonists should be demonstrated at dosages that are without effect on plasma prolactin, and be reversible with selective D3 antagonists. Aretha at al, (1995) met this first criterion by showing that the selective D3 antagonist UH232 could block the dopamine synthesis inhibiting effect of 7—OH-DPAT on mesolimbic DA neurons. Although they didn’t rule out involvement of D2 receptors by measuring plasma prolactin levels, this finding suggests that D3 receptors may serve as the synthesis modulating autoreceptors in these neurons. The density of D2 receptors in the ventral tegmental area is much greater than that for D3 receptors (Richtand at al 1995). Thus, greater affinity for 89 receptors in vitm may not have any functional relevance in viva if the number of D2 receptors far exceeds that of D3 receptors. This may be highly relevant for mesolimbic DA neurons with cell bodies located in the ventral tegmental area. Based on these observations, it could be postulated that the efficacy of PDl 28907 to affect changes in dopamine homeostasis would necessarily be less than that of PNU-95,666. Thus, the discrepancy between a lack of effect on dopamine metabolism and an inhibitory effect on dopamine synthesis may reflect a lower functional efficacy for the D3 receptor in overall DA receptor- Inediated autoregulation. However, the inhibition of dopamine synthesis by PDT 28907 in these studies agrees with the findings of others (Pugsley er a1 1995; Bristow et a1 1996) that show similar changes in dopamine synthesis. Taken together, these results suggest that dopamine synthesis may be a more sensitive aSSfily of the D3 receptor-mediated effect on mesolimbic DA neurons. However, deSpite this small discrepancy it is clear that these data are in accord with the hYPOthesis that both D2 and D3 receptors can function as autoreceptors in IneSOlimbic DA neurons. Although quinelorane binds to both D2 and D3 receptors, these results demonstrate that D2 and not D3 receptors mediate the increase in TIDA ne“tonal activity following administration of this second-generation DA agonist. DeVelopment of more selective DA agonists such as quinelorane have 90 uncovered a previously unrecognized role for dopamine in the regulation of TIDA neurons. 91 Chapter Four D, Receptor-Mediated Stimulation of TIDA Neurons is Dependent on Afferent Neuronal Input to the Mediobasal Hypothalamus Pharmacological evidence reviewed in Chapter Two suggests that the activity of TIDA neurons can be modulated by a D2 receptor-mediated mechanism. Activation of D2 receptors with agonists (quinelorane or PNU- 95,666) stimulates rather than inhibits TIDA neurons which is not consistent with a known action mediated by DA autoreceptors located on axon terminals, cell bodies or dendrites of these neurons. The inhibitory nature of cellular responses typically coupled to D2 receptors, e.g. inhibition of adenylyl cyclase, reduced calcium conductance, or increased potassium conductance (Liu et al, 1994), suggests that the direct response to D2 receptor activation would necessarily be inhibitory. More likely the stimulatory effect of quinelorane represents the effect of D2 receptor inhibition of an inhibitory neuronal pathway (disinhibiton) one or more neurons removed from TIDA neurons. Thus, an investigation of this neuronal circuit and at least a rudimentary understanding about the location of D2 receptors involved will be an important first step toward defining the mechanism of this response. Deafferentation of the mediobasal hypothalamus (Halasz and Pup 1965) has been used to isolate the hypothalamo-pituitary axis from the rest of the 92 bit the Sul det me 101 R11 Ill thj life the L801 C011 brain. This technique was developed to study the functional capacity and role of the mediobasal hypothalamus in control of trophic hormone secretion. Subsequently, this knife—cut deafferentation procedure was also used to determine the site of prolactin mediated activation of TIDA neurons (Gudelsky at al, 1978). In these experiments exogenous administration of prolactin is able to increase TIDA neuronal activity in rats with surgical deafferentation of the mediobasal hypothalamus suggesting that the stimulatory effects of prolactin are not dependent upon afferent neuronal input to the mediobasal hypothalamus. Rather, prolactin acts within the area circumscribed by the knife-cut to stimulate TIDA neurons. The purpose of these experiments of these was to determine the extent to which the D2 receptor-mediated activation of TIDA neurons is dependent on afferent neuronal input from outside the mediobasal hypothalamus. To this end, the effects of quinelorane were tested in rats bearing a complete deafferentation of the mediobasal hypothalamus. The knife-cut procedure was employed to isolate the mediobasal hypothalamus and TIDA neurons from afferent neuronal connections to the rest of the brain (Figure 20). 93 'IV ,’1 /A13 _~ y I \, ,’I M21." , \ I ‘ _‘—v‘y \ \ I” H PeVN’ l'-v~ ‘ I I'V' av I ‘V,\“ l ,—- v \ . ~ I r=1.5mm Figure 20 Schematic diagram depicting a Hillasz knife and the region of the mediobasal hypothalamus circumscribed by the knife—cut procedure which includes median eminence (ME), arcuate nucleus (ARC) and the pituitary (denoted by -. Abbreviations: third ventricle (3V), anterior pituitary (AP), dorsomedial nucleus (DM), intermediate pituitary lobe (IL), medial zona incerta (MZI), neural pituitary lobe (NL), optic chiasm (ox), periventricular nucleus (PeVN), and ventromedial nucleus (V M) The hypothesis tested was that D2 receptor mediated activation of TIDA neurons is dependent upon afferent neuronal input arising from outside the mediobasal hypothalamus. If the D2 receptors that mediate the stimulatory effect on TIDA neurons reside outside the area circumscribed by the knife-cut, then the effects of quinelorane should be blocked in deafferentated rats. 94 Met Dea' met and 5161 bit W35 T0< 3m! Methods Deafferentation of the Mediobasal Hypothalamus Seven days prior to the day of the experiment Long-Evans rats receiving mediobasal hypothalamic knife—cut lesions were anesthetized with Equithesin (3 ml/ kg i.p.) Once general anesthesia was apparent, the surgical site was shaved and scrubbed with betadine and 75% ethyl alcohol. Rats were then placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA, USA) with the incisor bar 8 mm below the intraaural line (Konig and Klippel 1963). A small incision was made along the midline of the skull and the skin retracted so that a 4 mm x 4 mm opening in the skull could be made using a small cutting drill (Dremel Moto- Tool model 380-5, Racine, WI). A modified Hilasz knife (height 1.5 mm, radius 1.5 mm; (Figure 20) was attached to a stereotaxic carrier and lowered in the midsagittal plane to the base of the skull 1.3 mm posterior to bregma. The anterior portion of the mediobasal hypothalamic deafferentation was accomplished by rotating the knife twice; 90° to the left and then to the right. The knife was then moved in a posterior direction 1.5mm, rotated 180° and moved hack 1.5 mm in the anterior direction. The knife was then rotated back to midline and removed from the brain at the point of entry. Sham surgery consisted of lowering the knife 3.0 mm below the dura mater surface without rotation. After surgery the bone was sealed with bone wax and the skin closed 95 tit hi1 Ve hi2 prt the deg inc me Hi?! HOS with surgical wound clips. The position and efficacy of the mediobasal hypothalamic lesions was verified in each rat by histological and neurochemical criteria described below. Only those animals that met both criteria were included in the study. Verification of the Deafferentation In neurochemistry experiments rats were killed by decapitation and the brains were removed and rapidly frozen over dry ice. Frontal sections were prepared from these frozen brains as detailed in the General Methods. During the preparation of these frontal sections the mediobasal hypothalamic deafferentation was visually evaluated and rats in which the knife-cut was incomplete, inappropriately placed, or had evidence of necrosis of the mediobasal hypothalamus were eliminated from the study. In addition several neurochemical measures were taken to verify that deafferentation of the mediobasal hypothalamus was complete. Dopamine and norepinephrine were measured by high performance liquid chromatography coupled to electrochemical detection from median eminence and intermediate pituitary lobe samples. Dopamine and norepinephrine served as indirect indices of DA and noradrenergic innervation to these regions. 96 Results The effectiveness of the mediobasal hypothalamic deafferentation was verified by several neurochemical measures. The result of deafferentation on dopamine and norepinephrine in the median eminence is shown in Table 5. After deafferentation median eminence dopamine decreased by 26% and norepinephrine by 62%. In contrast deafferentation produced a much larger decrease (45%) in dopamine from samples in the intermediate lobe of the pituitary (Table 5). As shown in Figure 21, deafferentation of the mediobasal hypothalamus had no effect on the concentration of DOPAC in the median eminence per re, but blocked quinelorane-induced increases in this dopamine metabolite. On the other hand, the knife-cut lesion had no effect on the ability of quinelorane to decrease DOPAC concentrations in the nucleus accumbens (Figure 22). 97 Table 5 Effects of surgical deafferentation on the concentration of dopamine and norepinephrine in the median eminence, and dopamine in the intermediate lobe of the pituitary of male rats one week after surgery. One week following sham or deafferentation surgery, rats were sacrificed by decapitation. Values represent means i 1 SEM of the concentrations of dopamine and norepinephrine from 17 sham surgery control or mediobasal hypothalamic (MBH) knife-cut male rats. Sham Control MBH Knife-Cut (n=17) (11:17) % Change Median Eminence Dopamine 171.83 :I: 7.62 126.23 :I: 14.67 - 26 % Norepinephrine 29.95 i 2.35 11.43 i 0.87 _ 62 0/0 Intermediate Pituitary Lobe Dopamine 15.01 i 1.18 8.31 :I: 1.29 - 45 °/o Norcpinephrine 2.14 :t 0.22 1.54 :1: 0.22 —28 % 98 :2 Vehicle 2° ' 2° ' — Quinelorane * ’E‘ E E 15 - '§ 15 - l 9 9 D. O. E.” T E’ a, 10 - a, 10 ~ 5 5 0 0 E E O 5 ‘ o 5 * o o 0 0 Sham Control MBH Knife-cut Figure 21 Effects of quinelorane on DOPAC concentrations in the median eminence of rats receiving either sham surgery (Sham Control) or surgical deafferentation of the mediobasal hypothalamus (MBH Knife-cut). One week following sham or deafferentation surgery, rats were injected with either quinelorane (100 pg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/kg; i.p.) 60 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 5-12 determinations of DOPAC concentrations in the median eminence of vehicle- (open columns) and quinelorane-treated (solid columns) rats. *p < 0.05, values from quinelorane- treated rats that are significantly different from vehicle-treated controls. 99 1:! Vehicle — Quinelorane 35 - 35 T A A T 5 28+ * ,g 28 ~ * 2 2 9 e O. D. g, 21 " CE» 21 '1 Bi 3» 5 14 5 14 o 1 o ‘ g a o 7 1 8 7 - O 0 Sham Control MBH Knife-cut Figure 22 Effects of quinelorane on DOPAC concentrations in the nucleus accumbens rats receiving either sham surgery (Sham Control) or surgical deafferentation of the mediobasal hypothalamus (MBH Knife-cut). One week following sham or deafferentation surgery, rats were injected with either quinelorane (100 pg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg, ip.) 60 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 5-12 determinations of DOPAC concentrations in the nucleus accumbens of vehicle- (open columns) and quinelorane-treated (solid columns) rats. *p < 0.05, values from quinelorane- ' treated rats that are significantly different from vehicle-treated controls. 100 Discussion Data presented in the previous chapter demonstrated that TIDA neurons are activated following administration of quinelorane or the D2 selective agonist PNU-95,666, but not after administration of the D3 selective agonist PD128907. These pharmacological data indicate that the activity of TIDA neurons is subject to modulation by an action at D2 receptors. Since these neurons lack autoreceptors and the pattern of increased neuronal activity after a DA agonist is inconsistent with autoreceptor-mediated phenomena, TIDA neurons are not likely the source of dopamine that could mediate this response. This study was undertaken to test the hypothesis that the D2 receptor-mediated effect arises from outside the mediobasal hypothalamus. Deafferentation of the mediobasal hypothalamus was verified by both neurochemical and histological methods, and used to determine the dependence of quinelorane-induced activation of TIDA neurons on neuronal inputs from outside the mediobasal hypothalamus. Both objective and subjective measures demonstrated that deafferentation of the mediobasal hypothalamus effectively isolated TIDA neurons from the rest of the brain. Neurochemical measurements provided evidence that the knife-cut lesion disrupted afferent neurons originating from outside and terminating within or passing through the mediobasal hypothalamus. Periventricular hypophysial DA neurons with cell bodies located in the periventricular nucleus 101 have axons that course thorough the mediobasal hypothalamus and tenninate within the intermediate lobe of the pituitary (Goudreau 1992). Dopamine levels were reduced in the intermediate pituitary lobe in lesioned rats suggesting that periventricular hypophysial DA neurons had been disrupted. Surgical deafferentation of the mediobasal hypothalamus disrupts afferent noradrenergic neuronal inputs from the pons medulla to the median eminence (Fuxe 1965) and results in a 50-60% depletion of norepinephrine concentrations with in this region (Gudelsky at al, 1978). The 62% loss of NE in lesioned rats compared to control is consistent with these earlier findings and is evidence that the knife-cut was complete and disrupted these noradrenergic projections. While the remaining 38% of norepinephrine left in the median eminence after deafferentation could indicate that the lesion was not complete, it is possible that norepinephrine in this instance may come from a peripheral source. Alper at al, (1980) found that surgical ligation of the superior cervical ganglion reduces norepinephrine levels in the median eminence by 22%. This finding suggests that some of the remaining norepinephrine in the median eminence of mediobasal hypothalamic deafferentated rats may come from these autonomic noradrenergic neurons. Neurochemical measurements were also used to insure that TIDA neurons were not disrupted by surgical deafferentation of the mediobasal hypothalamus. There was no change in the basal activity of TIDA neurons. 102 However, the 24% decrease in the median eminence dopamine content might indicate that some TIDA neurons originating in more rostral portions of the arcuate nucleus were disrupted by the deafferentation procedure. Dopamine is a precursor of norepinephrine therefore the loss of noradrenergic neurons following knife-cut lesions may account for some of the loss of median eminence dopamine. However, the number of DA neurons innervating the median eminence far outnumber the noradrenergic neurons in this region so this loss would be very small It is important to note that the loss of dopamine in the median eminence was substantially less than in the intermediate lobe where DA neurons had clearly been disrupted. Physiological changes were also evident after the deafferentation procedure. Hilasz and Pup (1965) reported that rats displayed a syndrome of diabetes insipidus after complete mediobasal hypothalrnic deafferentation. Diabetes insipidus is characterized by excessive drinking and urinating due to the disruption of antidiuretic hormone-secreting neurons that arise from the supraoptic and paraventricular nuclei in route to the posterior pituitary. Increased urinary output and compensatory drinking is due to the loss of antidiuretic hormone-mediated water reabsorption in the renal collecting duct. Although no objective measure of water balance was recorded in the present studies, it was apparent from the need to refill the water bottles and change the bedding more frequently that animals receiving knife-cut lesions were drinking 103 and urinating significantly more than the sham-treated control rats. Taken together, these observations suggest that TIDA neurons are relatively spared compared to noradrenergic afferent projections to the median eminence, or periventricular hypophysial DA or antidiuretic hormone neurons that pass through the mediobasal hypothalamus in route to the pituitary gland. The purpose of this experiment was to determine if activation of TIDA neurons by D2 receptor-mediated events is dependent upon afferent neuronal inputs that arise from outside the mediobasal hypothalamus. This possibility was tested by examining the effects of quinelorane in rats where the TIDA neurons (and associated mediobasal hypothalamus) was isolated from the rest of the brain. The results of this experiment revealed that the stimulatory effect of quinelorane on TIDA neurons is dependent upon afferent neuronal input to the mediobasal hypothalamus. The inability of quinelorane to increase median eminence DOPAC concentration in rats with a knife-cut lesion is not likely due to the disruption of TIDA neurons because the surgical procedure has no effect on the basal activity of these neurons. In addition, Gudelsky at al, (1978) demonstrated that TIDA neurons were still responsive the effects of prolactin after this deafferentation procedure. Nor is the lack of effect due the surgical stress-induced alterations in the pharmacokinetics of quinelorane, since mediobasal hypothalamic deafferentation did not alter the inhibitory response of mesolimbic DA neurons to this D2 agonist. Rather, these results indicate that 104 activation of D2 receptors on neurons residing outside the area circumscribed by the mediobasal hypothalamic knife-cut provides stimulatory neuronal input to TIDA neurons. Diencephalic DA neurons are organized into distinct systems with clusters of cell bodies that have discrete axonal projections (Dahlstrom and Fuxe 1964; Bjorklund at al, 1975). The terminal fields of these local circuits remain mostly within the hypothalamus but are less well defined than those of the ascending mesotelencephalic DA systems. Any DA neurons outside the region of the knife-cut including hypothalamic (A13 or A14) or mesotelencephalic DA neurons (A8 - A10) could be the source of dopamine that mediates this stimulatory response. The lack of definitive information about DA neuronal circuits that may be involved in the regulation of TIDA neurons make it difficult to more specifically define the mechanism whereby this regulation could occur. Interestingly, electrical stimulation of the dorsomedial nucleus (the site of A13 dendrites) increases the concentration of DOPAC in the median eminence (Gunnet and Moore 1985), suggesting a role for incertohypothalamic DA neurons in this process. An alternative to this interpretation of the results might suggest that afferent input to the mediobasal hypothalamus is not mediating the stimulatory Dz-mediated effect on TIDA neurons but rather plays a permissive role. That is, the loss of afferent input to the mediobasal hypothalamus leads to the loss of the 105 neuronal circuit within this brain region that more directly mediates the response. If this scenario were true, then D2 receptors that mediate the stimulatory effect of quinelorane should be located within the mediobasal hypothalamus. But, localization studies have failed to demonstrate the existence of significant levels of D2 receptors in this brain region. Autoradiographs of D2 receptor binding or measurement of mRNA for the D2 receptor, have not demonstrated significant levels of the D2 receptor within the mediobasal hypothalamus (Bouthenet at al, 1987; 1991). The most straightforward interpretation of the results presented here support the hypothesis that quinelorane induces an increase in TIDA neuronal activity by an action at D2 receptors that is mediated indirectly by afferent neuronal inputs that arise from outside the mediobasal hypothalamus. 106 Chapter Five Evidence That Activation of Delta/OP1-Fleceptors Stimulates TIDA Neurons Three pharmacologically distinct subtypes of opioid receptors have been identified in the central nervous system: delta, kappa and mu (Goldstein and James 1984; Martin 1984; and Paterson at al., 1983) recently renamed OP1, OP2 and OP3 respectively (Dhawan at al, 1996). Studies of the regulation of DA neurons by subtypes of opioid receptors have been dependent upon the availability of specific agonists and antagonists for these receptors. Much of what is known about the effects of opioids on DA neurons has been based upon studies using compounds that act at kappa/OP2 or mu/OP3 receptors. Activation of mu/OP3 receptors stimulates mesolimbic, nigrostriatal (Di Chiara and Imperato, 1988) and incertohypothalamic DA neurons (Lookingland and Moore, 1985), while inhibiting TIDA neurons (Deyo et al, 1979; Gudelsky and Porter 1979; Alper at al, 1980). On the other hand, activation of kappa/OP2 receptors inhibits the basal activity of periventricular DA neurons and pharmacologically—stimulated activities of mesolimbic DA, nigrostriatal DA and TIDA neurons (Manzanares at al, 1991). Comparatively less is known about the role of delta/0P1 receptors in the regulation of DA neuronal systems in the brain. Administration of the 107 delta/OP1-selective enkephalin analog [D-Pen2,D-Pen5]enkephalin (DPDPE) increases the release of dopamine in the nucleus accumbens (Spanagel et al, 1990), but has no effect in the striatum (Mulder at al, 1989). No information is available regarding the effects of delta/0P1 receptor activation or blockade on the activities of hypothalamic DA neurons. The purpose of these studies was to determine the dose-response and time course effects of the delta/0P1 selective agonist DPDPE and antagonist naltrindole on neurochemical estimates of the activity of TIDA neurons. For comparison purposes the effects of these treatments on the activity of mesolimbic, nigrostriatal, and periventricular- hypophysial DA neurons were also evaluated. Methods Surgical care and handling of rats for implantation of i.c.v. guide cannula, and neurochemistry procedures for determining the synthesis and metabolism of dopamine in the median eminence, nucleus accumbens, striatum and intermediate pituitary lobe were carried out as described in the General Methods. Dmgs The delta/0P1 agonist DPDPE (Peninsula Laboratories, Belmont, CA, USA) and antagonist naltrindole hydrochloride (Donated by Dr. James H. Woods University of Michigan Medical School, Ann Arbor, MI, USA) were dissolved in distilled water. NSD 1015 was dissolved in 0.9% saline. Drugs were 108 administered as indicated in the appropriate figure legends and dosages were calculated as the respective salts. Results Dose response effects of DPDPE on DOPAC concentrations in the median eminence and nucleus accumbens are depicted in Figure 23. Intracerebroventricular administration of DPDPE effected a dose-related increase in DOPAC concentrations in the median eminence and nucleus accumbens, but was without effect in either the striatum or intermediate lobe of the pituitary (Table 6). Similarly, DPDPE increased the accumulation of DOPA in the median eminence and nucleus accumbens (Figure 24), but again was without effect in either the striatum or intermediate lobe of the pituitary (Table 7). The ability of DPDPE to increase the concentrations of DOPAC in terminals of TIDA and mesolimbic DA neurons was observed again when the time course of this delta/0P1 receptor agonist was determined. Figure 25 shows that i.c.v. administration of DPDPE caused a rapid increase in DOPAC concentrations in the nucleus accumbens (by 15 min) and median eminence (by 30 min), but again was without effect in either the striatum or intermediate pituitary lobe (Table 8). 109 Median Eminence _a _A 0' m _L I0 DOPAC (ng/mg protein) 0) (O (A) O 0.0 0.3 1.0 3.0 10.0 Nucleus Accumbens DOPAC (ng/mg protein) a '8 O 0.0 0.3 1.0 3.0 10.0 Dose of DPDPE (pg/rat) Figure 23 Dose-response of the effects of DPDPE on DOPAC concentrations in the median eminence and nucleus accumbens. Rats were injected i.c.v. with either DPDPE (0.3, 1, 3, 10 ug/rat) or its water vehicle (5 til/rat) 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in tissues obtained from 6-8 rats. *, values from DPDPE—treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated (open columns) controls. 110 Table 6 Lack of effect of DPDPE on DOPAC concentrations in the striatum and intermediate pituitary lobe. Rats were injected i.c.v. with either DPDPE (0.3, 1, 3, 10 ug/ rat) or its water vehicle (5 ul/ rat) 30 min pn'or to decapitation. Values represent the means :1: 1 SEM of DOPAC concentrations (ng/ mg protein) in tissues obtained from 6-8 rats. DPDPE Striatum Intermediate Pituitary Lobe _£E£./m) Vehicle 41.1 :i: 1.9 1.70 i 0.14 0.3 42.0 :t 1.7 1.88 i- 0.10 1.0 41.9110 1.92i0.16 3.0 44.3 :1: 2.1 1.73 i 0.08 10 47.4 :t 2.1 1.64 :1: 0.11 111 Median Eminence DOPA (ng/mg protein) 0.0 0.3 1.0 3.0 10.0 Nucleus Accumbens 32 ' * * 0.0 0.3 1.0 3.0 10.0 Dose of DPDPE (ug/rat) Figure 24 Dose-response of the effects of DPDPE on the accumulation of DOPA in the median eminence and nucleus accumbens. Rats were injected i.c.v. with either DPDPE (0.3, 1, 3, 10 ug/ rat) or its water vehicle (5 p.1/ rat) and killed by decapitation 30 min later. All rats were injected with NSD 1015 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPA concentrations in tissues obtained from 6-8 rats. *, values from DPDPE-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated (open columns) controls. 112 Table 7 Lack of effect of DPDPE on DOPA concentrations in striatum and intermediate pituitary lobe. Rats were injected i.c.v. with either DPDPE (0.3, 1, 3, 10 ug/ rat) or its water vehicle (5 [.11/ rat) 30 min prior to decapitation. Values represent the means :1: 1 SEM of DOPA concentrations (ng/ mg protein) in tissues obtained from 6-8 rats. DPDPE Striatum Intermediate Lobe Ara/rat) Vehicle 23.7 i 0.6 1.88 i 0.11 0,3 22.7 :t 1.8 1.77 i‘ 0.12 1.0 21.0 :t 0.8 1.82: 0.19 3.0 22.3 :1: 0.7 1.81 i 0.07 10 22.7 i 0.8 1.90 i: 0.04 113 Median Eminence 24 - 20‘ 16* 12- DOPAC (ng/mg protein) o 1530 so 120 240 Nucleus Accumbens 7s - Sad/\% 45- 4 30‘ DOPAC (ng/mg protein) 15‘ 0 1530 60 120 240 Min After DPDPE Figure 25 Time course of the effects of DPDPE on DOPAC concentrations in the median eminence and nucleus accumbens. Rats were injected i.c.v. with DPDPE (10 ug/ rat) and killed by decapitation either 15, 30, 60, 120 or 240 min later. Control rats were injected i.c.v. with water vehicle (5 ul/ rat) and killed by decapitation 30 min later. Symbols represent means and vertical lines 1 SEM of DOPAC concentrations in tissues from 5-9 rats. When vertical lines are not visible 1 SEM is smaller than the radius of the symbol. Solid symbols represent values from DPDPE-treated rats that are significantly different (p < 0.05) from vehicle-treated controls. 114 Table 8 Lack of effect of DPDPE on DOPAC concentrations in striatum and intermediate pituitary lobe. Rats were injected i.c.v. with DPDPE (10 ug/ rat) and killed by decapitation either 15, 30, 60, 120 or 240 min later. Control rats were injected i.c.v. with water vehicle (5 111/ rat) and killed by decapitation 30 min later. Values are reported as means :I: 1 SEM of DOPAC concentrations (ng/ mg protein) in tissues from 5-9 rats. its: Striatum IntermedlizgeePituitary DPDPE Vehicle 38.5 3:19 1.50 i014 15 39.9 i1.7 1.52 i010 30 39.1 :10 1.52 i016 60 41.7 i- 2.1 1.52 i 0.08 120 37.6 i 1.1 1.60 i008 240 35.5 i 2.1 1.67 i 0.11 115 Intracerebroventricular administration of the selective delta/0P1 receptor antagonist naltrindole had no effect on DOPAC concentrations in either the median eminence or nucleus accumbens (Figure 26), nor did it alter the concentrations of DOPAC in the striatum or intermediate pituitary lobe (Table 9). On the other hand, naltrindole completely blocked DPDPE-induced increases in DOPAC concentrations in both the median eminence and nucleus accumbens (Figure 27). 116 Median Eminence -‘N .A 000030100100 1 EL 1 _A-L DOPAC (11ng protein) A 0.0 0.3 1.0 3.0 Nucleus Accumbens N (A) A O) k» 0) on O DOPAC (ng/mg protein) ‘16 O 0.0 0.3 1.0 3.0 Dose of Naltrindole (pg/rat) Figure 26 Lack of effect of naltrindole on DOPAC concentrations in the median eminence and nucleus accumbens. Rats were injected i.c.v. with either naltrindole hydrochloride (0.3, 1 and 3 pg/ rat) or its water vehicle (5 ul/ rat) 45 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in tissues from 8-9 rats 117 Table 9 Lack of effect of naltrindole on DOPAC concentrations in the striatum and intermediate pituitary lobe. Rats were injected i.c.v. with either naltrindole (0.3, 1 and 3 ug/ rat) or its water vehicle (5 pl/ rat) 45 min prior to decapitation. Values represent the means i 1 SEM of DOPAC concentrations (ng/ mg protein) in tissues from 8-9 rats. DPDPE Striatum Intermediate Pituitary Lobe Ali's/m) Vehicle 237 i 0.6 1.88 i 0.11 0.3 22.7 i1.8 1.77 i 0.12 1.0 21.0 i 0.8 1.82 2'.”- 0.19 3.0 22.3 i 0.7 1.81 i 0.07 10 I 22.7 i 0.8 1.90 i 0.04 118 Median Eminence 13- ,E t 1: Vehicle #3 ‘5‘ — DPDPE D. 12 e “e” * ’ a, 9 5 0 6j E O 3' o 0 Vehicle Naltrindole so Nucleus Accumbens 70'1 a 60* 501 zoj 10- DOPAC (rig/mg protein) 8 Vehicle Naltrindole Figure 27 Effects of naltrindole on DPDPE-induced increases in DOPAC concentrations in median eminence and nucleus accumbens. Rats were injected i.c.v. with either a mixture of DPDPE (1O |.l.g/ rat and naltrindole hydrochloride (3 ug/ rat), or DPDPE (10 ug/ rat) alone and killed by decapitation 45 min later. Control rats were injected ic.v. with water vehicle (5 Lll/ rat) 45 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in tissues of 6-8 rats. *, values form DPDPE-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated (open columns) controls. 119 Discussion The results of these studies indicate that i.c.v. administration of the delta/0P1 selective agonist DPDPE increases the activity of mesolimbic and TIDA neurons, but fails to alter the activity of either nigrostriatal or periventdcular—hypophysial DA neurons. The stimulatory effect of DPDPE on mesolimbic DA neurons is in agreement with a previous report (Spanagel et al, 1990) demonstrating that the DPDPE-induced increases of mesolimbic DA neuronal activity is blocked by the selective delta/0P1 antagonist naltrindole. These results indicate that this action is mediated through delta/0P1 receptors. Kappa/OP2 and mu/OP3 receptor agonists inhibit and stimulate respectively, the activities of both mesolimbic and nigrostriatal DA neurons in concert (Manzanares et al, 1991; Di Chiara and Imperato, 1988). The selective stimulatory effect of DPDPE on mesolimbic but not nigrostriatal DA neurons represents a departure from effects observed after administration of kappa/OP2 or mu/OP3 opioid agonists. The present study represents the first in viva demonstration that the activity of TIDA neurons is increased following activation of delta/0P1 receptors with DPDPE. These results reveal that delta/0P1 receptor-mediated regulation of DA neurons is not limited to mesolimbic DA neurons. Indeed, the efficacy for delta/OPl-mediated increases in neuronal activity appears to be 120 greater for TIDA compared to mesolimbic DA neurons, as evidenced by a greater magnitude of increase in DOPAC concentrations in the median eminence compared to the nucleus accumbens following DPDPE administration. On the other hand, DPDPE has no effect on the activity of periventricular-hypophysial DA neurons that terminate in the intermediate lobe of the pituitary, and in this respect, these neurons resemble the DA neurons of the nigrostriatal system. A possible explanation for the lack of effect of DPDPE on nigrostriatal and periventriwlar—hypophysial DA neurons is that delta / 0P1 receptors are already maximally occupied by an endogenous ligand. But, blockade of delta/0P1 receptors with naltrindole had no effect on the activities of periventricular-hypophysial, nigrostriatal, mesolimbic or TIDA neurons suggesting that there is no tonic regulation of these DA neurons by endogenous delta/0P1 selective neuropeptides. Thus, the lack of effect of DPDPE on nigrostriatal and periventricular-hypophysial DA neurons is not due to pre- existing maximum stimulation by endogenous delta/ 0P1 ligands. There is substantial evidence for interactions between enkephalinergic and mesotelencephalic DA neuronal systems. Exogenous application of dopamine to rat striatal slices in vitro induces an efflux of MET-enkephalin that is blocked by pretreatment with D2 antagonist and mimicked by apomorphine but not D1 selective agonists (Pasinetti et al, 1984). Administration of quinpirole induces a 121 stimulatory effect on enkephalin gene expression and release of enkephalin in the striatum in viva (Bannon et al., 1989). In addition Pickel et al, (1992) have provided anatomical evidence for these DA / enkephalinergic interactions. Delta/0P1 receptors have a more restricted distribution than other subtypes of opioid receptors. Dense areas of delta/0P1 receptors have been located in olfactory-related areas, neocortex, striatum, nucleus accumbens and amygdala, whereas few receptors have been located in the hypothalamus (Mansour at al, 1988; Desjardins et al, 1990). While the stimulatory action of DPDPE on mesolimbic DA neurons may occur at the receptors located in nucleus accumbens, the absence of delta/0P1 receptors in regions containing cell bodies or terminals of TIDA neurons suggests that the stimulatory action of DPDPE on these neurons may occur indirectly, possibly via afferent neurons. The studies described in this chapter represent the first in viva demonstration that the activity of TIDA neurons like that of mesolimbic DA neurons can be acutely activated by DPDPE via an action at delta/0P1 receptors. 122 Chapter Six Activation of 02 Receptors Fails to Induce the Release of A Neurotransmitter That Stimulates T IDA Neurons Results from studies presented in Chapters Two and Three demonstrated that activity of TIDA neurons is regulated via a modulatory pathway that involves activation of D2 receptors, and that Dz-mediated stimulation TIDA neurons is indirect and mediated by afferent neuronal input to the mediobasal hypothalamus. In order to explore the nature of these afferent neurons two basic hypotheses were formulated. The first of these states that activation of D2 receptors induces the release of a neurotransmitter that stimulates TIDA neurons. A second (but not mutually exclusive) hypothesis states that activation of TIDA may be the result of inhibition of an endogenous substance that tonically inhibits TIDA neurons. These possibilities are examined in this and the subsequent chapter. In order for the first hypothesis to be correct two basic criteria would have to be met; 1) administration of the putative stimulatory neurotransmitter or its receptor agonist should activate TIDA neurons, and 2) pharmacological blockade of receptors for these neurotransmitters should prevent the D2 receptor-mediated activation of TIDA neurons. At least two classes of receptors 123 for endogenous neuropeptides have been identified which fullfill this first criterion. Delta/0P1 opioid receptors Experiments described in chapter 4 demonstrated that DPDPE acting via delta/0P1 receptors acutely stimulates TIDA neurons. This suggests the possibility that endogenous enkephalins acting selectively at delta/0P1 receptors may be involved in quinelorane-induced activation of these neurons. In support of this theory there is substantial evidence for interactions between enkephalinergic and mesotelencephalic DA neuronal systems. Indeed, exogenous application of dopamine to rat striatal slices in vitra induces an efflux of MET-enkephalin that is blocked by pretreatment with a D2 antagonist and mimicked by nonselective agonist apomorphine but not D1 selective agonists (Pasinetti at al, 1984). In viva the D2 receptor agonist quinpirole increases preproenkephalin gene expression in the stn'atum, whereas D1 agonists are ineffective (Bannon at al, 1989). In addition, Pickel at al, (1992) have provided anatomical evidence for DA / enkephalinergic interactions in mesotelencephalic DA neurons, but there are no studies describing similar DA / enkephalinergic interactions in the mediobasal hypothalamus. 124 Neurotensin receptors The tridecapeptide neurotensin was initially isolated from bovine hypothalamus (Carraway et al, 1973). Subsequently, neurotensin has been found to be widely distributed throughout the central nervous system (U hl et al., 1977) but is found in highest concentration in nucleus accumbens, preoptic area and mediobasal hypothalamus (Kobayashi et al, 1977). Of particular interest is the wealth of pharmacological, anatomical and neurochemical evidence that neurotensin interacts with central DA neurons (N emeroff 1986). The neuroleptic-like effects attributed to exogenous administration of neurotensin have consequently focused much of the attention upon these interactions in the mesolimbic and nigrostriatal DA neurons. (Nemeroff 1980; Ervin at al, 1981; Blaha at al, 1992). The dopamine inhibiting property of endogenous neurotensin was directly tested by Wagstaff et al, (1994) who demonstrated that methamphetamine-induced dopamine release was facilitated in rats pretreated with either neurotensin antiserum or SR48692, the non-peptide antagonist of neurotensin receptors. Quinpirole, a second-generation D2 agonist, can increase neurotensin release in the anterior nucleus accumbens and lateral caudate nucleus (Wagstaff at al, 1996). Thus, elevated DA neuronal activity would induce neurotensin release that, in turn, inhibits DA neurons. 125 Neurotensin imrnunoreactive cell bodies have also been localized to mediobasal hypothalamic regions; including the arcuate nucleus (Kahn at al, 1980) which contains the perikarya of TIDA neurons. Additional observations have provided direct evidence that neurotensin activates TIDA neurons (Berry and Gudelsky 1990; Pan at al, 1992), and that this increased activity was coincident with a reduction in the release of prolactin from the anterior pituitary (Pan at al, 1992). The discovery of SR48692, a non-peptide antagonist of neurotensin receptors (Gully at al, 1993) has made it possible to further test the physiological role of neurotensin in the regulation of TIDA neurons and prolactin secretion. The goal of experiments presented in this chapter is to test the hypothesis that endogenous enkephalins via delta/0P1 receptors and/ or neurotensin via neurotensin receptors are involved in D2 receptor-mediated increases in TIDA neuronal activity. To this end, the effect of quinelorane on DOPAC concentrations in the median eminence were determined in rats pretreated with either the delta/0P1 receptor antagonist naltrindole, or the neurotensin receptor antagonist SR48692. For comparative purposes the concentration of DOPAC in the nucleus accumbens was also measured. If release of either enkephalins or neurotensin mediate the stimulatory effects of quinelorane, then administration of these selective antagonists should block quinelorane-induced increases in the concentration of DOPAC in the median eminence. 126 IMnhods Care and handling of rats and brain tissue used in these experiments followed the procedures for surgical implantation of i.c.v. guide cannula, and neurochemistry procedures for measuring the concentration of DOPAC in the median eminence and nucleus accumbens, were performed as described in the General Methods. [Mugs SR48692 (Provided by Dr. Danielle Gully Sanofi Recherche, Toulouse, France) was dissolved in 10% dimethyl-sulfoxide (DMSO), naltrindole hydrochloride (Donated by Dr. James H. Woods University of Michigan Medical School, Ann Arbor, MI, USA) was dissolved in distilled water. Quinelorane dihydrochloride was dissolved in 0.9% saline. Drugs were administered as indicated in the appropriate figure legends and dosages were calculated as the respective salts. Resuns Figure 28 depicts the effect of quinelorane alone or when co-administered with naltrindole on the concentration of DOPAC in the median eminence and nucleus accumbens. These results show that naltrindole failed to block the quinelorane-induced increase in the concentration of DOPAC in the median eminence. Similarly, quinelorane-induced decreases in the concentration of 127 DOPAC in the nucleus accumbens were not altered by co-administration of naltrindole. The effect of co-administration of the selective neurotensin antagonist SR48692 on quinelorane-induced changes in the concentration of DOPAC in the median eminence and nucleus accumbens are shown in Figure 29. Co- administration of SR48692 had no effect on the quinelorane-induced increase in the concentration of DOPAC in the median eminence or the decrease in the concentration of DOPAC in the nucleus accumbens. 128 Median Eminence C: Vehicle . — Quinelorane d” m—t * d U" A N 9-1 DOPAC (rig/mg protein) .4 o — Vehicle Naltrindole Nucleus Accumbens 21- T 18" a T 15‘ 121 61 DOPAC (rig/mg protein) Vehicle Naltrindole Figure 28 Lack of effect of naltrindole on quinelorane-induced increases in DOPAC concentrations in the median eminence and nucleus accumbens. Rats were injected with quinelorane (100 ug/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg i.p.), and naltrindole hydrochloride (3 ug/ rat; i.c.v.) or its water vehicle (3 )11/ rat) 60 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in tissues sampled from 7-8 rats. *, values from quinelorane-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated (open columns) controls. 129 DOPAC (rig/mg protein) DOPAC (ng/mg protein) '0 d .5 m _L 0" _tL N Median Eminence :21 Vehicle — Quinelorane * Vehicle SIR-48692 Nucleus Accumbens T T Vehicle SR48692 Figure 29 Lack of effect of SR48692 on quinelorane-induced changes in DOPAC concentrations in the median eminence and nucleus accumbens. Rats were injected with quinelorane (100 ug/ kg, i.p.) or its 0.9% saline vehicle (1 ml/kg i.p.), and SR48692 (100 [lg/kg; i.p.) or its 10 °/o DMSO vehicle (1 ml/ kg, i.p.) 60 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in tissues sampled from *, Values from quinelorane-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated (open columns) controls. 130 Discussion Quinelorane-induced activation of TIDA neurons is mediated by D2 receptors and dependent upon afferent neuronal input arising from outside the mediobasal hypothalamus. There are a number of stimulatory neuropeptides that could serve as intermediaries in this pathway most prominently enkephalins and neurotensin. As demonstrated in the previous chapter pharmacological activation of delta/0P1 receptors with the selective agonist DPDPE activates TIDA neurons. However, little is known about endogenous DA / enkephalinergic interactions in the hypothalamus of male rats. In female rats during lactation and suckling there is an increase in the expression of enkephalin in tyrosine hydroxylase positive neurons in the arcuate nucleus (Merchenthaler 1993; 1994; Merchenthaler at al, 1995). In this instance enkephalins are thought to help maintain elevated prolactin during lactation and suckling by suppressing the tendency of TIDA neurons to become activated thereby reducing prolactin secretion in response to this hyperprolactinemic state. Experiments in this chapter were performed to test the hypothesis that activation of TIDA neurons is dependent upon endogenous activation of delta/OP1-receptors. The observation that naltrindole does not alter the stimulatory effect of quinelorane on TIDA neurons indicates that endogenous release of enkephalins acting at delta/0P1 receptors do not mediate this 131 stimulatory response. Thus, there is no evidence to suggest that quinelorane activates TIDA neurons via an action on enkephalinergic neurons. Comparatively more is known about neurotensin and DA neuronal interactions in the mediobasal hypothalamus. Administration of neurotensin directly into the third cerebral ventricle reduces the concentration of prolactin in plasma via a dopamine dependent process (Koenig at al, 1982). In addition, Gudelsky at al, (1989) have established that exogenous administration of neurotensin induces a dose-related increase in the activity of TIDA neurons in male rats. Similar results have been obtained in female rats where the increase in TIDA activity following neurotensin administration was correlated with reductions in the concentration of plasma prolactin (Pan at al, 1992). Hentschel and coworkers (1998) have shown that SR48692 can block the stimulatory effects of prolactin on TIDA neuronal activity. In these studies SR48692 had no effect per re but did block the haloperidol—induced prolactin- mediated activation of these neurons. Taken together, these results suggest that the stimulatory effect of prolactin on TIDA neurons may be mediated, in part, by endogenous neurotensin. These observations coupled with the finding that D2—like agonists induce the release of neurotensin in the telencephalon raised the possibility that endogenous release of neurotensin mediates the Dz-induced increase in TIDA neuronal activity. 132 The failure of the neurotensin antagonist SR48692 to modify the stimulatory effect of quinelorane on TIDA neurons argues against the involvement of neurotenisn in this stimulatory pathway. One possible explanation for the ineffectiveness of SR48692 could be that the antagonist does not interact with the neurotensin receptor that mediates the activation of TIDA neurons. Studies utilizing a variety of in vitm and in viva assays have demonstrated that although SR48692 can antagonize neurotensin-induced hypomotility in the rat, it fails to block neurotensin-induced hypothermia and analgesia (Dubuc at al, 1994). These disparate effects are suggestive of at least two neurotensin receptor subtypes, one of which is SR48692 insensitive. On the other hand, SR48692 can block the haloperidol-induced prolactin-mediated activation of TIDA neurons, suggesting that this effect is mediated, in part, by endogenous neurotensin (Hentschel at al, 1998). This observation establishes that prolactin feedback activation of TIDA neurons is mediated by a SR48692 sensitive neurotensin receptor and by extension is independent of the stimulatory effects mediated by D2 receptor agonists. For comparison purposes the effects of these treatments on the activity of mesolimbic DA neurons were also determined. The results reveal that blockade of either delta/0P1 receptors or neurotensin receptors has no effect on quinelorane induced inhibition of these neurons. These results indicate that 133 neither endogenous enkephalins nor neurotensin are involved in the D2 receptor— mediated inhibitory feed back of these neurons. Taken together, the results of these studies demonstrate that neither endogenous NT nor enkephalins acting at NT- or delta/OPl-receptors, respectively, mediates the quinelorane-induced increase in TIDA neuronal activity. 134 Chapter Seven Evidence that Dz Receptor-Mediated Activation of TIDA Neurons Occurs via Disinhibition Results presented in the previous chapter demonstrated that quinelorane- induced activation of TIDA neurons is not mediated by the release of endogenous neurotensin or enkephalin. Alternatively, stimulatory D2 receptor- mediated afferent neuronal regulation of TIDA neurons could occur via suppression of a tonically active inhibitory system. This disinhibition hypothesis is attractive due to the abundant evidence that activation of D2 receptors is coupled to inhibitory cellular events such as inhibition of adenylyl cyclase and reduced calcium conductance, or increases in potassium conductance (Jackson and Westlind-Danielsson 1994). Thus, a mechanism involving disinhibition is invoked to explain how inhibitory D2 receptor activation would induce a net increase in TIDA neuronal activity. For this disinhibition hypothesis to be correct two criteria must be met: 1) TIDA neurons must be tonically inhibited, and 2) exogenous administration of an inhibitory neurotransmitter or a receptor agonist should block D2 receptor-mediated regulation of TIDA neurons. The inhibitory neurotransmitter dynorphin via an action at kappa/OP2 receptors is known to meet the first criterion. 135 In the male rat, the activity of TIDA neurons is tonically inhibited by dynorphinergic neurons; an action that is mediated by kappa/OP2 receptors. Administration of the kappa/OP2 receptor-selective agonist U-50,488 has no dose- or time-related effect on either the accumulation of DOPA or concentrations of DOPAC in the median eminence (Manzanares at al, 1991; 1992a). Whereas, administration of the selective kappa/OP2 antagonist nor- binaltorphirnine (nor-BNI) induces an increase in the accumulation of DOPA in median eminence that is reversed by U-50,488 (Manzanares et al, 1992a). Furthermore, immunoneutralization of either the active fragment dynorphin],8 or full-length dynorphin].17 induces an increase in the concentration of DOPAC in the median eminence that was reversed by administration of U-50,488 (Manzanares at al, 1992b) These data indicate that the activity of TIDA neurons in male rats is tonically inhibited by endogenous dynorphin acting at kappa/OP2 receptors. Tonic inhibition of TIDA neurons in the male rat is also dependent upon the presence of androgens (Gunnet at al, 1986; Toney at al, 1991), and Manzanares at al, (1992a) have demonstrated that this inhibitory action of androgens is mediated, in part, by endogenous dynorphin acting at kappa/OP2 receptors. The purpose of the experiments described in this chapter is to test the hypothesis that inhibition of tonically active dynorphinergic neurons is the mechanism that accounts for Dz-receptor-mediated activation of TIDA neurons. 136 To this end, the effects of the kappa/OPZ-selective agonist U-50,488 and antagonist norBNI were tested for their ability of modify quinelorane-induced changes in the synthesis of dopamine in the median eminence. For comparison, neurochemical measurements were also made in the nucleus accumbens. Since gonadal androgens are known to be important for maintenance of the tonic dynorphinergic inhibition of TIDA neurons, an additional series of experiments was carried out in orchidectomized males. In situ hybridization studies were also conducted to determine if quinelorane-induced increases in TIDA neuronal activity were accompanied by a reduction in the activity of dynorphinergic neurons in the hypothalamus. If a reduction in dynorphin release and subsequent activation of inhibitory kappa/OP2 receptors is involved in this change in TIDA activity then administration of the kappa/OP2 agonist U- 50,488 should re-instate this inhibitory tone and block the stimulatory effects of quinelorane. On the other hand, the blockade of kappa/0P1 receptors with nor- BNI should acutely stimulate TIDA neurons, but prevent further activation by quinelorane. There was concern that a lack of quinelorane effect in nor-BNI treated rats could be attributed to reaching a ceiling in the maximal possible activation of TIDA neurons. Therefore some rats in these studies were pre-treated with prolactin. The prolactin mediated activation of TIDA neurons is known to occur by a mechanism different from thst of quinelorane, and these treatments 137 should be additive. Since maintenance of inhibitory dynorphinergic tone is gonadal steroid dependent, orchidectomy alone should increase basal TIDA activity and block both nor-BNI- and quinelorane-induced changes in median eminence DOPA accumulation. Finally, if quinelorane induced changes are the result of inhibition of a population of dynorphinergic neurons then expression of dynorphin mRNA should be reduced after quinelorane treatment. Methods Surgical care and handling of rats for implantation of i.c.v. guide cannula, gonadectomies, and neurochemistry procedures were carried out as described in the General Methods. The level of dynorphin mRNA expression in the arcuate and ventromedial nuclei was measured by in .ritu hybridization procedures described in General Methods. Drugs U-50,488 HCl (Dr. P. VonVoigtlander, The Pharmacia and Upjohn Company, Kalamazoo, MI), nor-BNI HCl, (RBI, Natic, MA) and rat prolactin (Dr. A.F. Parlow, NIDDK-rPRL-B—S-SIAFP Lot # APP-3697A) were dissolved in distilled water. Quinelorane dihydrochloride and NSD1015 were dissolved in 0.9 % saline. Drugs were administered as indicated in the legends of the appropriate figures; doses of drugs were calculated based on the weight of the free base or salt form of each drug. 138 Results As shown in Figure 30, activation of kappa/OP2 receptors with U-50,488 (10 mg/ kg, s.c.) had no effect per re, but prevented quinelorane-induced increases in the concentration of DOPA in the median eminence. In contrast, U-50,488 did not alter quinelorane-induced decreases in the concentration of DOPA in the nucleus accumbens. Blockade of kappa/OP2 receptors with nor-BNI increased median eminence DOPA accumulation, and prevented the stimulatory effects of quinelorane on dopamine synthesis in this region (Figure 31). Administration of prolactin also increased median eminence DOPA accumulation, but did not alter the ability of quinelorane to further stimulate dopamine synthesis in the median eminence. On the other hand, neither nor-BNI nor prolactin pretreatment altered quinelorane-induced decreases in DOPA concentration in the nucleus accumbens. Orchidectomy increased the basal concentration of DOPA in the median eminence but had no effect on the quinelorane-induced increases in DOPA concentration in this region (Figure 32). As shown in Figure 33, nor-BNI increased DOPA concentrations in the median eminence of gonadally-intact and orchidectomized rats. Orchidectomy alone increased the concentration of DOPA compared to intact controls but did not prevent the nor—BNI induced increase in DOPA accumulation. 139 Specific hybridization of 3’sS-labeled cells in the arcuate and ventromedial nuclei is consistent with the presence of dynorphin containing neuronal cells within these brain regions. Administration of quinelorane (100 lig/ kg, i.p.) significantly reduced the dynorphin mRNA labeling ratio in the arcuate nucleus but was without effect in the ventromedial hypothalamic nucleus (Figure 34). 140 Median Eminence .A. U" * 1:1 Saline — Quinelorane —L N DOPA (rig/mg protein) H Vehicle U50488 15 . Nucleus Accumbens 121 DOPA (rig/mg protein) Vehicle U50488 Figure 30 Effects of quinelorane on DOPA concentrations in the median eminence and nucleus accumbens of vehicle— and U-50,488-treated rats. Animals were injected with either U—50,488 (10 mg/ kg; s.c.) or its distilled water vehicle (1 ml/ kg, s.c.) 65 min prior to decapitation and with either quinelorane (100 pg/ kg, i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. All rats were injected with NSD 1015 (100 mg/ kg; i.p.) 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 6-9 determinations of DOPA concentrations in the median eminence and nucleus accumbens of vehicle- (open columns) or quinelorane- (solid columns) treated rats. * P < 0.05, values from quinelorane-treated rats that are significantly different from vehicle-treated controls. 141 Figure 31 Effects of quinelorane on DOPA concentrations in the median eminence and nucleus accumbens of vehicle-, nor-BNI- and prolactin-treated rats. Rats were injected with either nor-BNI (12.5 ug/ rat; i.c.v) or its distilled water vehicle (3 pl/ rat; i.c.v.) 120 min prior to decapitation, or with either prolactin (10 pg/ rat; i.c.v) or its distilled water vehicle 12 hours prior to decapitation. Vehicle-, nor-BNI- and prolactin-treated rats were injected with either quinelorane (100 ug/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. All rats were injected with NSD 1015 (100 mg/ kg; i.p.) 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 6-9 determinations of DOPA concentrations in the median eminence and nucleus accumbens of vehicle- (open columns) or quinelorane- (solid columns) treated rats. * P < 0.05, values from quinelorane treated rats that are significantly different from vehicle-treated controls. # P < 0.05, values from nor-BNI and prolactin- treated rats that are significantly different from vehicle-treated controls. 142 Median Eminence 2‘ ‘ 1:23 Saline * — Quinelorane 18 . # r: l # .9 15- e . J5 O. a: 12 1 ti 5 9 ‘ E o 5 ‘ o 3 l o __ _ Vehicle NOR-BNI Prolactin Nucleus Accumbens 36 A 30 - CD ‘9‘ 24 J ”I“ Q. E’ E, 18 I * * * o- 12 ' O o 6 . 0 — Vehicle NOR-BNI Prolactin 143 28 1 a 1:: Saline A 24 ‘ — Quinelorane : E 20 - T 9 t a a) 16 - it 5 12 ~ E I o 3 ' D 4 4 0 Intact Orchidectomy Figure 32 Effects of quinelorane on DOPA concentrations in the median eminence of intact and orchidectomized rats. Rats were injected with either quinelorane (100 itg/ kg i.p.) or its 0.9 0/o saline vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. All rats were injected with NSD 1015 (100 mg/ kg; i.p.) 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 7-11 determinations of DOPA concentrations in the median eminence of vehicle- (open columns) or quinelorane- (solid columns) treated-rats. * P < 0.05, values from quinelorane treated rats that are significantly different from vehicle-treated controls. ** P < 0.05, values from orchidectomized rats that are significantly different from intact controls. 144 N .a l_—_l Saline * 18 J _ Nor-BNI E 9 15 1 9 O. a g: 12 - I l | E 9- < T o. o 5 ‘ o 3 .. 0 Intact Orchidectomy Figure 33 Effects of nor-BNI on DOPA concentrations in the median eminence of intact and orchidectomized rats. Rats were injected with either nor-BNI (1 2.5)1g/ rat; i.c.v.) or its distilled water vehicle 5 |J.l/ rat; i.c.v.) 120 min prior to decapitation. All rats were injected with NSD 1015 (100 mg/ kg; i.p.) 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of 7-8 determinations of DOPA concentrations in the median eminence of vehicle- (open columns) or nor-BNI- (solid columns) treated rats. * p < 0.05, values from nor-BNI-treated rats that are significantly different from vehicle-treated controls. ** p < 0.05, values from orchidectomized rats that are significantly different from intact controls. 145 :3 Saline - Quinelorane 4i Labeling Ratio DYN mRNA/Cell ARC VMN Figure 34 Effects of quinelorane on the expression of dynorphin mRNA in the arcuate (ARC) and ventromedial hypothalamic (VMN) nuclei of the rat. Quinelorane (100 pg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) was administered to rats 90 min prior to decapitation. Columns represent means and vertical lines 1 SEM. of the labeling ratio/ cell determined in 6 rats. * P < 0.05, values of quinelorane-treated rats (solid column) that are significantly different from saline-treated rats (open column). 146 Discussion Although TIDA neurons lack the DA autoreceptor-mediated regulation that has been well characterized in the mesolimbic DA system, these neurons are subject to activation that is mediated by DA D2 receptors. D2 receptor mediated regulation of mesolimbic DA neurons is effected by autoreceptors located on neurons which, either directly or indirectly via interneurons, provide reciprocal innervation back to these DA neurons (Oades and Halliday 1987). Pharmacological activation of D2 receptors inhibits the activity of mesolimbic DA neurons by autoreceptor mediated prevention of dopamine release and afferent neuronal inhibition of impulse conduction. In agreement, the results of the present study demonstrate that activation of D2 receptors following administration of quinelorane inhibits mesolimbic DA neurons terminating in the nucleus accumbens. Unlike mesolimbic DA neurons, TIDA neurons are not inhibited by quinelorane; this finding is consistent with the conclusion that these neurons lack D2 receptor-mediated afferent neuronal feedback inhibition (Demarest and Moore 1979). Rather, results of experiments described in this dissertation have demonstrated that activation of D2 receptors on neurons residing outside the mediobasal hypothalamus provides stimulatory neuronal input to TIDA neurons. In addition, this afferent stimulatory input is not dependent upon 147 endogenous activation of delta / OP1 , or neurotensin receptors. Experiments in this chapter were designed to test the hypothesis that D2 receptor-mediated activation of TIDA neurons is the result of disinhibition. The results reveal that D2 receptor—mediated activation of TIDA neurons occurs, at least in part, via a mechanism involving kappa / OP2 receptors. If quinelorane attenuates kappa / OP2 receptor-mediated tonic inhibition of TIDA neurons by preventing dynorphin release (Manzanares at al, 1992a; 1992b) then: 1) administration of an kappa/OP2 agonist should block the actions of quinelorane, and 2) administration of an kappa/OP2 antagonist should mimic the effects of quinelorane by stimulating TIDA neurons, and prevent any further stimulation by quinelorane. The results presented here indicate that this is the case. The inability of quinelorane to further increase median eminence DOPA accumulation in nor-BNI-treated rats was not due to limitations in the maximal rate of dopamine synthesis (i.e., ceiling effect). Indeed, administration of prolactin increased median eminence DOPA accumulation but did not alter the ability of quinelorane to further stimulate dopamine synthesis in this region. Quinelorane was active when co-administered with nor-BNI since DOPA accumulation in the nucleus accumbens was reduced following administration of this agonist. Therefore, these results are consistent with the hypothesis that D2 receptor-mediated activation of TIDA neurons occurs via 148 disinhibition that is dependent upon inhibition of tonically active dynorphinergic neurons. Contrary to our hypothesis, the results of these expedments also reveal that Dz-mediated increases in TIDA neuronal activity are independent of the effects of gonadal steroids. This conclusion is based on the observation that removal of gonadal androgens by orchidectomy increased basal TIDA neuronal activity but had no effect on either quinelorane— or nor-BNI induced increases in the accumulation of DOPA. Earlier studies have demonstrated that TIDA neurons are subject to tonic inhibition by the action of dynorphin (Manzanares at al, 1992b) and that this tonic inhibition is dependent, in part, upon the presence of gonadal androgens (Manzanares at al, 1992a). In vitra experiments demonstrating that release of dynorphin from hypothalamic slices is lower in tissue from orchidectomized versus gonadally-intact or orchidectomized testosterone-treated male rats further support these in viva findings (Almeida at al, 1987). It is possible that two populations of tonically active dynorphinergic neurons modulate the activity of TIDA neurons. One that is responsive to circulating gonadal androgens and another that is insensitive. Taken together, these results suggest that an androgen insensitive population of dynorphinergic neurons is responsive to the inhibitory actions of the D2 receptor agonists. The results of the present study also confirm that in male rats the basal activity of mesolimbic DA neurons terminating in the nucleus accumbens is not 149 regulated by kappa/OP2 receptors (Manzanares at al, 1992a; Manzanares at al, 1992b). Indeed, neither activation (with U-50,488) nor blockade (with nor-BNI) of kappa/OP2 receptors had any effect on nucleus accumbens DOPA accumulation in control rats. The inability of either U-50,488 or nor-BNI to block the inhibitory effects of quinelorane in the nucleus accumbens reveals that neither kappa/OP2 receptors nor dynorphinergic neurons participate in D2 receptor-mediated regulation of mesolimbic DA neurons. These experiments also demonstrate that the DA D2 receptor-mediated activation of TIDA neurons is independent of that mediated by prolactin on these neurons. Prolactin-induced activation is delayed (requires 12 hours), is not dependent upon afferent neuronal inputs to the mediobasal hypothalamus and is mediated by endogenous activation of neurotensin receptors (Hokfelt and Fuxe, 1972; Gudelsky at al, 1978; Hentschel et al, 1998). In contrast, the D2 receptor- mediated activation is acute (occurs within an hour), dependent upon afferent neuronal input to the mediobasal hypothalamus, and independent of neurotensin. Activation of D2 receptors specifically reduces the capacity of dynorphinergic neurons in the arcuate nucleus to synthesize new dynorphin. This conclusion is based on the observation that quinelorane selectively reduces the expression of dynorphin mRNA in the arcuate nucleus, but is without effect on the expression of dynorphin mRNA in ventromedial nucleus. These data 150 support the hypothesis that activation of TIDA neurons via D2 receptors is mediated, in part, by the inhibition of tonically active dynorphinergic neurons. The reduction in dynorphin output from this population of neurons, in turn, reduces the activation of inhibitory kappa/OP2 receptors thereby disinhibiting TIDA neurons. 151 Chapter Eight Inhibition of Tuberoinfundibular Dopaminergic Neuronal Activity by Selective Activation of D1 Receptors Previous chapters in this dissertation have dealt with D2 receptor- mediated modulation of TIDA neurons. In this chapter the focus of attention will be shifted to consider the role that D1 receptors may play in regulating TIDA neuronal activity. The discovery that antipsychotic efficacy was highly correlated with antagonism of D2 receptors (Creese at al, 1976; Seeman at al, 1976) stimulated the development of many D2 selective compounds. In contrast, there were relatively fewer D1 receptor selective compounds. When SKF 38393 was initially described it had a unique profile that was attributed to D1 receptor selectivity (Setler at al, 1978). Although this agonist failed to alter plasma prolactin in viva it was shown to mimic dopamine-stimulated adenylyl cyclase in rat striatal homogenates. Subsequently, Sibley at al (1982) used radioligand displacement studies to demonstrate that SKF 38393 bound preferentially to D1 receptors. The first Dl-selective antagonist was SCI-123390 (Iorio at al, 1983). This compound has been widely used as a tool to study the anatomical distribution and function of D1 receptors. Reported interactions with SHT2 receptors in vitra and and its short half-life in viva were serious limitations to its clinical utility 152 (Billard at al., 1984; Barnett at al, 1986). These problems led to the development of a second generation D1 antagonist SCH39166 which has a longer half-life and greater selectivity for D1 receptors (Chipkin at al, 1988; McQuade at al, 1991). Acute administration of the D1 receptor agonist SKF38393 has been shown to inhibit “activated” TIDA neurons (Berry and Gudelsky 1991), but little information is available regarding the effects of D1 receptors on the basal activity of these neurons. Thus, the purpose of the present series of experiments was to characterize the role of D1 receptors in the regulation of TIDA neurons. To this end, the effects of the selective D1 receptor agonist SKF38393 (Sibley at al, 1982) and D1 receptor antagonist SCH39166 (Chipkin at al, 1988) were examined on the activity of TIDA neurons as determined by measuring the concentrations of DOPAC in the median eminence. In addition, the effects of these treatments on plasma prolactin were also determined. These experiments were designed to test the hypothesis that TIDA neurons are subject to inhibitory modulation via D1 receptors. If this hypothesis is true then SKF38393 should lower DOPAC concentrations in the median eminence and increase plasma prolactin whereas, pre-treatment with SCH39166 should block or attenuate these SKF38393-induced changes. 153 Methods Brain tissue for neurochemical analysis and blood for radioimmuno assay of prolactin were handled as described in the General Methods section. The lower limit of prolactin detection in these assays was 150 pg/ 100 pl of plasma and the interassay coefficient of variation was 10%. Drugs SCH39166 (Dr. A. Barnett Scherng-Plough, Bloomfield, NJ) was dissolved in 50% ethanol (v/v), while SKF 38393 (Research Biochemicals International, Natick, MA) was dissolved in 0.1% ascorbic acid. Both drugs were administered as indicated in the appropriate figure legends, and dosages were calculated based on their respective salts. Resufls The effect of incremental doses of the D1 antagonist SCH39166 on DOPAC concentration in the median eminence is shown in Figure 35. SCH39166 induced a dose-related increase in the concentration of DOPAC in the median eminence that was significant at 1 and 3 mg/ kg. Administration of 3 mg SCH39166/ kg induced a time-related increase in median eminence DOPAC concentration that was significant by 1 h and returned to control levels by 16 h (Figure 36). In contrast, the D1 agonist SKF38393 significantly reduced the concentration of DOPAC in the median eminence at all doses tested ( Figure 154 37A), and significantly increased plasma prolactin levels at these doses ( Figure 37B). As shown in Figure 38 administration of 20 mg SKF38393/ kg reduced the concentration of DOPAC in the median eminence and increased plasma prolactin. Pre-treatment with 3 mg SCH39166/ kg increased the concentration of DOPAC in the median eminence per re and attenuated both the SKF38393 induced decrease in this metabolite and the increase in plasma prolactin. 155 36- 301 24- 18 - ._T_ 12‘ DOPAC (ng/mg protein) VEH 0.3 1.0 SCH39166 (mg/kg) Figure 35 Dose-response effects of SCH39166 on the concentrations of DOPAC in the median eminence. Rats were injected with either 0.3, 1 or 3 mg SCH39166/ kg; s.c. or its 50% ethanol vehicle (1 ml/ kg, s.c.) 120 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence from 7-9 rats. *Values from SCH39166-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated rats(open column). 156 18- 3.? 93 O 512- U) E \ C) 5 2 6- D. O D 0 III I I I 012 4 8 16 Hours after SCH39166 Figure 36 Time-course of effects of SCH39166 on the concentrations of DOPAC in the median eminence. Rats were injected with SCH39166 (3 mg/ kg, s.c.) at 1, 2, 4, 8 or 16 h or its 50% ethanol vehicle (1 ml/ kg; s.c.) 1 h prior to decapitation (time 0). Circles represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence from 7-9 male rats. The filled circle represents a value that is significantly different (p < 0.05) from time 0. 157 DOPAC (ng/mg protein) Prolactin (ng/ml) VEH 5.0 10.0 20.0 SKF38393 (mg/kg) Figure 37. Effects of SKF38393 on the concentrations of DOPAC in the median eminence and prolactin in plasma. Rats were injected with 5,10 or 20 mg SKF38393/kg; i.p. or its 0.1% ascorbic acid vehicle (1 ml/ kg; i.p.) 60 min prior to decapitation. Panel A: Columns represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence from 7-8 male rats. *Values from SKF38393-treated rats (solid columns) that are significantly different (p <0.05) from vehicle-treated controls (open column). Panel B: columns represent means and vertical lines 1 SEM of plasma prolactin concentrations in 7-8 male rats. *Values from SKF38393-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated controls (open column). 158 A 25 . E Saline A - SKF38393 .E 2 20 - 9 ._'L. .. a _'r_ g 15 . * B: 5 1o .. O E o 5 ‘ o O — Vehicle SCH39166 B 12 - Prolactin (ng/ml plasma) 0) Vehicle SCH39166 Figure 38 SCH391663 blocks the SKF38393-induced decrease in the median eminence DOPAC concentrations. Rats were injected with SCH 39166 (3 mg/ kg; s.c.) or its 50 % ethanol vehicle (1 ml/ kg, s.c.) 120 min, and with either SKF38393 (20 mg/ kg) or its 0.1% ascorbic acid vehicle (1 ml/ kg, i.p.) 60 min prior to decapitation. Panel A: columns represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence from 7- 8 rats. Panel B: columns represent means and vertical lines 1 SEM of plasma prolactin concentrations from 7-8 rats.* Values from SKF38393-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated controls (open columns). ** Values from SCH39166+SKF38393-treated rats (solid columns) that are significantly different from SKF38393+vehicle-treated controls (solid columns). 159 Discussion The D1 antagonist SCH39166 induces a stimulatory effect on TIDA neurons as demonstrated by the dose- and time-related increase in the median eminence DOPAC concentration. By definition an antagonist binds to a receptor with high affinity but lacks intrinsic activity or efficacy. This is the case for SCH39166 (Chipkin 1988; Wamsley et al, 1991; Tice at al, 1994). The conclusion drawn when administration of an antagonist induces an increase in activity as shown for SCH39166 is that the drug is interfering with tonic inhibitory effects mediated by the D1 receptor. This finding supports the hypothesis that the D1 receptor plays an inhibitory role in the regulation of TIDA neuronal activity and suggests that this inhibition is tonically active. These inhibitory effects are rather unexpected given the findings that direct effects of D1 receptor activation are usually stimulatory i.e., increases in CAMP and intracellular Ca";+ concentrations (jakson and Westlind-Danielsson 1994). Thus, the direct effect of D1 receptor activation might be expected to be stimulatory. While it is possible that the D1 receptor mediating this effect is coupled via an inhibitory second messenger system, it is more likely the result of an indirect action. Thus, D1 receptors could act indirectly by tonically activating a neuronal system that inhibits TIDA neurons. In this way, the D1 receptor- mediated inhibition would be analogous to the D2 receptor-mediated activation 160 ...'- ,d-c-j of TIDA which also occurs indirectly via inhibition of dynorphinergic neurons (Chapter 6). Incremental doses of the D1 receptor agonist SKF38393 significantly reduced the basal activity of TIDA neurons as demonstrated by the reduction in the concentration of DOPAC in the median eminence. Although D1 receptor- mediated inhibition of TIDA neurons is tonic, the further reduction in activity observed after SKF 38393 suggests that this inhibition is not maximal. Although, Berry and Gudelsky (1990) reported that there was only a trend towards reduction in activity, SKF 38393 at these doses did not significantly reduce tyrosine hydroxylase activity in TIDA neurons except in one experiment where the 20 mg/ kg dose reduced dopamine synthesis in the median eminence. These results not withstanding, the biological importance of the statistically significant decrease in the metabolic activity of TIDA neurons after administration of the D1 agonist can be appreciated by looking at the effects on plasma prolactin. At all dosages tested SKF38393 significantly increased the concentration of prolactin in plasma. This finding is in agreement with others who have reported similar elevations in prolactin after administration of D1 agonists (Saller and Salama 1989). Since there is no evidence to support a direct effect of SKF38393 on lactotrophs in the anterior pituitary (due to absence of D1 receptors; Cocchi et al, 1987), the elevation in prolactin is taken as functional evidence for inhibition of TIDA neurons which regulate prolactin release. 161 Pro-treatment with the D1 antagonist blocked the SKF38393-induced decrease in both metabolic and functional TIDA neuronal activity. This is demonstrated by the blockade of SKF38393-induced changes in the concentration of DOPAC in the median eminence and prolactin in plasma. These agonist/ antagonist studies suggest that the effects of SCH39166 and SKF38393 are both mediated by D1 receptors. However, these compounds induce opposite effects on the activity of TIDA neurons and this leaves open to question whether or not these effects are mediated at the same D1 receptor site. However, in vitro binding studies as well as in viva data suggest that at these doses these drugs are specific for the D1 receptor family (Sibley et al, 1982; Chipkin et a], 1988). Although, there is a possibility that these drugs could be working on D1 receptors in different sites or possibly D1 versus D5 receptors within the D1 receptor family, but the pharmacological tools to distinguish between D1 and DS receptors are not yet available. In conclusion, the results of studies presented in this chapter reveal that TIDA neurons are inhibited following activation of D1 receptors, and that under basal conditions these neurons are tonically regulated by inhibitory D1 receptors. 162 lg.» Chapter Nine Opposing Roles for 01 and 02 Dopaminergic Receptors in the Regulation of Tuberoinfundibular Dopaminergic Neuronal Activity. The major ascending nigrostriatal and mesolimbic DA neuronal systems in the brain are regulated by DA receptor-mediated mechanisms; i.e. DA agonists reduce, whereas DA antagonists increase impulse flow and the neurochemical activities of these neurons (Bunney et al, 1987). These effects are chiefly mediated by inhibitory D2 and / or D3 DA autoreceptors located on axon terminals, dendrites and neuronal cell bodies which function to maintain a homeostatic balance between impulse flow-induced neurotransmitter release and its replenishment by de nova synthesis (Elsworth and Roth, 1997). In contrast, TIDA neurons lack DA autoreceptors, and are unresponsive to acute administration of non-selective DA agonists (e.g. apomorphine and bromocriptine) and antagonists (e.g. haloperidol; Moore 1987). Utilization of newer more selective agonists which differentiate between the D1 and D2 receptor subfamilies (Schwartz et al, 1992;]ackson and Westlind- Danielsson, 1994) has revealed that TIDA neurons are subject to regulation by DA receptors. Indeed, acute administration of quinelorane with preferential affinity for the D2 receptor family increases the activity of TIDA neurons (Eaton 163 5' ‘3; et al, 1993). Since apomorphine also interacts with D2 receptors it is curious that this compound has no effect on the activity of TIDA neurons (Demarest and Moore 1979). One possible explanation for this observation is that apomorphine also interacts with both known receptors in the D1 family (D, and D5). Currently available agonists and antagonists do not distinguish between these two receptors. In addition, the D5 receptor is poorly expressed in rat brain when compared with the D, receptor (Meador-Woodruff at al, 1992; Tiberi at al, 1991). For these reasons the inibitory effect on TIDA neurons as documented by Berry and Gudelsky (1991) and confirmed in the previous chapter of this dissertation are likely mediated by the D, receptor in rats. Thus, opposing effects mediated by D, and D2 receptors on TIDA neuronal activity may lead to no net measurable change when both receptors are activated. The purpose of experiments described in this chapter was to test the hypothesis that the lack of response of TIDA neurons to non-selective DA agonists and antagonists is due to the simultaneous activation of inhibitory D, and stimulatory D2 receptors. To test this hypothesis, the effects of quinelorane, in the presence of either the selective D, receptor agonist SKF 38393 or the non- selective DA agonist apomorphine were examined on the concentration of DOPAC in the median eminence. If the lack of effect of non-selective DA drugs on TIDA neurons is due to opposite effects from simultaneous activation 164 1F. .. of D, )5 and D2 receptors then SKF 38393 should reverse quinelorane-induced increases in the activity of TIDA neurons. Likewise, apomorphine should block the stimulatory effect of quinelorane on these neurons. Changes in the expression of immediate early gene products (i.e., c-Fos and F os-related antigens (FRA’s)) represent an additional index of change in the activity of specific neuronal populations. The appearance of these transcription factors has been used as markers to identify neurons in which activity has changed in response to drug treatments or alterations in physiological state (Hoffman et al, 1993). Thus, immunocytochemical localization of Fos/FRA can be used to visualize individual neurons that are active. One disadvantage is the lack of specificity of this index since changes in Fos expression have been demonstrated in response to a wide array of stimuli in a variety of cell types (Hughes and Dragunow 1995). Thus, in a mixed neuronal population it is important to specifically identify the neuronal cell type in which changes in Fos/FRA expression occur. Two populations of tyrosine hydroxylase—immunoreactive (TH-IR) neurons have been described in the arcuate nucleus (ARC). These have been identified based upon their neurochemical phenotypes, and the size and location of their cell bodies in the dorsomedial (DM) and ventrolateral (VL) ARC (Everitt at al, 1986). The expression of Fos/FRA in TH-IR neurons in these sub- 165 Tl 'r divisions of the ARC is subject to differential regulation by gonadal steroids. In females, estrogen stimulates Fos/FRA in the DM-ARC whereas in males testosterone inhibits the expression of Fos/FRA in both the DM- and VL-ARC (Cheung at al, 1997b). The purpose of studies in this chapter is to determine if pharmacological manipulations that induced neurochemical changes in TIDA neurons also cause similar changes in immediate early gene expression in these neurons. To this end, the effects of quinelorane alone and in combination with SKF 38393 were determined on the co-localization of TH- and Fos/FRA-IR in the dorsomedial and ventrolateral subdivisions of the arcuate nucleus. If the pharmacological effects of these DA agonists alter the activity of TIDA neurons at the level of the genome then quinelorane should activate and SKF 38393 reduce the number of neurons expressing F 05 / FRA-IR. Methods The activity of TIDA neurons was determined by measuring dopamine metabolism (concentration of DOPAC) in the terminals DA neurons in the median eminence. In addition, dual irnmunohistochemistry for TH and Fos/FRA was used to quantify the frequency with which TH-IR and Fos/Fra— IR were co-localized in neurons of the arcuate nucleus. Details of these procedures can be found in the General Methods. 166 [hugs The D2 selective agonist quinelorane (Dr. M.M. Foreman, The Eli Lilly Co., Indianapolis, IN) dissolved in 0.9% saline, D, agonist SKF38393 (Research Biochemicals International, Natic, MA) was dissolved in 0.1% ascorbic acid, and the non-selective agonist apomorphine (Sigma Chemical Co. St Louis, MO) was dissolved in distilled water. All drugs were administered as indicated in the appropriate figure legends. Dosages for all drugs were calculated based on their respective salts. Resufls Given the inhibitory effects of SKF 38393 on DOPAC concentration in the median eminence this drug was tested for its ability to block increases in this metabolite following administration of the D2 receptor agonist quinelorane. As shown in Figure 39, administration of 2 mg SKF38393/ kg completely reversed the quinelorane-induced increase in median eminence DOPAC, and 20 mg SKF 38393 / kg significantly reduced DOPAC concentration compared to vehicle controls. To determine if simultaneous activation of D1 and D2 receptors could explain the lack of effect on the concentration of DOPAC in the median eminence following administration of apomorphine, rats were treated with quinelorane 100 lig/ kg and apomorphine 2 mg/ kg. As shown in figure Figure 40, apomorphine had no effect per re, but blocked the ability of quinelorane to 167 Ti increase DOPAC concentration in the median eminence. There was a significant interaction between quinelorane and apomorphine on median eminence DOPAC concentrations in this study (p < 0.05). The number of cells containing both TH-IR and Fos/FRA-IR was quantified and the results are presented in Figure 41. SKF 38393 (20 mg/ kg, i.p.) had no effect per re but was able to block the quinelorane—induced increase in the number of TH-IR cells expressing Fos/FRA in the DM-ARC whereas, none of these drug treatments had any effect in the VL-ARC. 168 15- !: Saline E 12 - - Qurnelorane 0) § 3 9 - ——T— E B) 5 . O 6 < n. 8 3‘ 0 Vehicle 02 2 20 SKF38393 (mg/kg) Figure 39 Dose-response effect of SKF38393 on DOPAC concentrations in the median eminence quinelorane-treated rats. Rats were injected with quinelorane (100 ilg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg; i.p.) at 60 min, and with 0.2, 2 or 20 mg SKF38393/ kg, i.p., or its 0.1% ascorbic acid vehicle (1 ml/ kg, i.p.) 60 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence sampled from 6—8 rats. *Values from quinelorane-treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated controls (open column). “Values from SKF38393-treated rats that are significantly different from vehicle-treated controls. 169 12- A 1: Vehicle .E _ Apomorphine t s 9. 9 T a ** E’ B: 6‘ T 5 0 3‘. 3. O D 0 Saline Quinelorane Figure 40 Effects of apomorphine on DOPAC concentrations in the median eminence of vehicle- and quinelorane-treated rats. Rats were injected with quinelorane (100 ug/ kg; i.p.) or its 0.9% saline vehicle (1 mg/ kg; i.p.) at 60 min, and with apomorphine (2 mg/ kg; s.c.) or 0.9% saline 30 min prior to decapitation. Columns represent means and vertical lines 1 SEM of DOPAC concentrations in the median eminence sampled from 6-8 rats. *Values from quinelorane-treated rats that are significantly different (p < 0.05) from saline-treated controls. ** Values for apomorphine-treated rats (solid columns) that are significantly different (p < 0.05) mm quinelorane-treated rats (open columns). 170 DM-ARC 30- * a) =Vehicle SE 24., _Ouinelorane 6:11 8E 59$ 12' 'as 3% :13 6‘ T o 30‘ VL-ARC 3E, 24« its 3%18- it ' 12‘ %é we. 212 6‘ T o fi—DL— Vehicle SKF38393 Figure 41 Effect of SKF38393 on quinelorane-induced changes in the number of perikarya in the ARC co-expressing TH-IR and Fos/FRA-IR. Rats were injected with quinelorane (100 ilg/ kg; i.p.) or its 0.9% saline vehicle (1 ml/ kg i.p.) at 90 min, and with SKF38393 (20 mg/ kg; i.p.) or its 0.1 °/o ascorbic acid vehicle (1 ml/ kg; i.p.) 90 min prior to perfusion. Columns represent means and vertical lines 1 SEM of the number of TH-IR neurons co- expressing Fos/FRA-IR in the DM- or VL-ARC of 5 male rats. *Values from quinelorane- treated rats (solid columns) that are significantly different (p < 0.05) from vehicle-treated controls (open columns). 171 Discussion Administration of the D, agonist blocked quinelorane-induced increases in metabolic activity of TIDA neurons. These findings are in agreement with those of Berry and Gudelsky (1991) who demonstrated that D, receptors function to inhibit TIDA neurons activated by a variety of stimuli. Indeed, activation of D, receptors reduced both the basal activity of these neurons as well as quinelorane- induced increases. Thus, it appears that the opposite effects of D1 and D2 agonists on TIDA neurons cancel one another when activated simultaneously. These data lend support to the basic thesis of this chapter i.e., that non-selective agonists such as apomorphine have no effect on the activity of TIDA neurons because of concurrent activation of both D1 and D2 receptors. One possible way to test this hypothesis would be to examine the effects of apomorphine in combination with a D, antagonist. Following blockade of D1 receptors, apomorphine should behave as a D2 receptor selective agonist similar to quinelorane. The problem with this approach is that D, antagonists elevate TIDA neuronal activity due to the tonic D, receptor-mediated inhibitory influences described in the previous chapter. An alternative approach to address this question would be to activate the D2 receptors by pretreatment with quinelorane and then test the ability of apomorphine to alter the activity of TIDA neurons. In this scenario apomorphine should behave as a D, agonist 172 "i since quinelorane will already be interacting with D2 receptors. This was the case, and under these conditions apomorphine had no effect per re, but reversed quinelorane-induced increases in the activity of TIDA neurons. Taken together, these results suggest that the lack of effect of non-selective agonists on TIDA neurons is due to the opposing roles for D, and D2 receptors in the regulation of the activity of their activity. The neurochemical information suggests that two converging pathways modulate the activity of TIDA neurons. 1) The stimulatory path mediated by the action of dopamine at D2 receptors is dependent upon afferent neuronal input to the mediobasal hypothalamus and this effect is mediated, in part, by reducing tonic inhibition of TIDA neurons via the action of dynorphin at kappa/OP2 receptors. 2) The inhibitory pathway is mediated by an action of dopamine at D1 receptors, perhaps by an indirect mechanism that is dependent upon the augmentation of an inhibitory system. Based on this information it is tempting to speculate that inhibitory dynorphinergic neurons may serve as the target that mediates these inhibitory D, )5 receptor effects. Thus, both the stimulatory and inhibitory actions of dopamine on TIDA activity could be mediated indirectly via D2 receptor inhibition and D, )5 stimulation of dynorphinergic neurons. Neurochemical changes observed in the terminals of TIDA neurons were reflected in changes in Fos/FRA expression in TH-IR neurons of the DM-ARC with the caveat that only increases and not decreases in TIDA activity were 173 Ti observed with this technique. While quinelorane increases the expression of these immediate early gene products in TH-IR cells, SKF 38393 was without effect per re, yet it completely blocked the quinelorane-induced increase in Fos/FRA expression in these neurons. There are several related proteins that the fos antiserum recognizes (Hoffman et al, 1994) and these authors have demonstrated that in addition to the rapid inducible form of F as there are related forms of this immediate early gene (FRA’s) that are constitutive. Thus, with selection of the appropriate F 05 antiserum (Cambridge Research Biochemicals as used here) for identification of F RA irnmunocytochemical localization of these proteins permits not only assessment of stimulated activity, but also suppressed function in these neurons. However the low basal activity of TIDA neurons in males may account for our inability to measure decreases in F 05 / FRA expression after administration of the D, agonist. Although these studies failed to demonstrate the inhibitory effect of D, )5 agonists immediate early gene expression was useful for demonstrating the stimulatory effect induced by D2 receptor activation. These results suggest, that the opposing roles of D, and D2 receptors in the regulation of TIDA neurons extend to alterations in gene expression and (perhaps more interestingly) that only the TIDA neurons originating in the dorsomedial portion of the arcuate nucleus are responsive to these treatments. These data suggest that in addition to the neurochemical changes observed in the median eminence the activation of 174 is DA receptors alters c-fos gene transcription, and synthesis of c-fos mRNA and FRA proteins. These changes may in turn be related to regulation of tyrosine hydroxylase in these neurons via the AP1 promoter. These results also suggest that the activation of TIDA neurons occurs at perikarya or dendrites rather than axon terminals. In conclusion, data in this chapter support the hypothesis that the activity of TIDA neurons can be modulated by the action of DA agonists at inhibitory D, receptors and stimulatory D2 receptors. These opposing effects provide an explanation for the lack of effect non-selective DA agonists on TIDA neuronal activity. 175 kn. Chapter Ten General Summary and Discussion This body of work has aimed to characterize DA receptor-mediated regulation of TIDA neurons. Although these neurons are not regulated by pre- or post-synaptic DA receptor mediated negative feedback in the way that other DA neurons are, the development of drugs that distinguish between the D1 and D2 receptor families has uncovered a heretofore unrecognized mode of regulation of these neurons. These studies have served to demonstrate that D2 receptor activation stimulates, whereas D, receptor activation inhibits TIDA neurons. D2 Receptor-mediated activation of TIDA neurons The schematic in Figure 42 depicts the hypothesized relationship between D2 receptors and the intervening steps that lead to activation of TIDA neurons. It was interesting to discover that the mechanism involves disinhibition of TIDA neurons mediated by inhibition of tonically active dynorphinergic influences. These inhibitory neurons are likely spared after mediobasal hypothalamic deafferentation since the basal activity of TIDA neurons was not altered by this procedure. In addition, in ritu hybridization experiments reveal that quinelorane decreases prodynorphin mRNA in the arcuate nucleus while having no effect on dynorphin message in the nearby ventromedial hypothalamic nucleus. Taken 176 together, these data suggest that the location of the dynorphinergic neurons mediating this response is within the region circumscribed by the mediobasal hypothalamic deafferentation. 177 D, Inhibition D2 Activation I».m.~.~.\_. TIDA + Prolactin /\ DA Figure 42 A schematic representation of the hypothetical relationship between inhibitory D1 and stimulatory D2 receptors on the regulation of TIDA neuronal activity. Abbreviations; (DA) dopamine, (DYN) dynorphin, (DMN) dorsomedial hypothalamic nucleus, (VMN) ventromedial hypothalamic nucleus. 178 Gonadal steroids are known to modulate the activity of TIDA neurons. In male rats androgens tonically inhibit TIDA neurons (Gunnet et al, 1986; Toney et al, 1991) and this effect is mediated, in part, by the actions of dynorphin at kappa/OP2 receptors. Administration of either the kappa/OP2 receptor antagonist nor-BNI or anti-dynorphin antibodies induces an acute increase in the activity of TIDA neurons in intact male but not female rats (Manzanares et al., 1992B). This effect was completely reversed by the co- administration of the kappa/OPZ-selective agonist U50,488 (Manzanares at al, 1992A). The lack of kappa/OP2 mediated regulation of TIDA neurons in females was initially of some concern, since our laboratory (Eaton at al, 1993) had previously demonstrated that the quinelorane-induced activation of TIDA occurs in both male and female rats. In females, the ability of quinelorane to suppress circulating levels of prolactin (via a direct action at the anterior pituitary) indirectly reduces the activity of TIDA neurons and thereby masks the stimulatory action of this drug on these neurons. Thus, if the mechanism of quinelorane—induced activation of TIDA neurons is mediated by inhibition of dynorphinergic neurons, then a similar arrangement should exist in females. Interestingly, the activity of TIDA neurons is reduced after ovariectomy and this effect is mediated by the loss of estrogen-dependent positive feedback by prolactin (T oney at al., 1992). In this situation the kappa/OP2 receptor 179 3T: II} antagonist nor-BNI is capable of activating these neurons (Manzanares at al, 1992B). Taken together, these data suggest that ovariectomy and the resulting reduction in positive prolactin feedback unmasks a latent dynorphinergic inhibitory pathway. This possibility is consistent with the observation that treatment with prolactin antisera renders female rats sensitive to the stimulatory effects of quinelorane (Eaton et al, 1993). The discovery that there is a gonadal steroid insensitive population of dynorphinergic neurons that can account for the disinhibition of TIDA neurons after activation of D2 receptors in male rats suggests that this mechanism might also occur in female rats as well. Future experiments could address this possibility by making use of immunohistochemistry and / or in rim hybridization techniques that could allow for the dynamic neuroanatomical mapping of dynorphinergic neurons in the mediobasal hypothalamw. Thus, a hypothesis that quinelorane activates TIDA neurons via a population of dynorphinergic neurons in both male and female rats that is insensitive to the effects of gonadal steroids could be tested. If this hypothesis is correct then quinelorane should reduce the activity of dynorphinergic neurons. This effect ought to be evident in both male and female castrated rats. 180 in 'l' D1 receptor-mediated inhibition of TIDA neurons and prolactin secretion The observation that the D, antagonist SCH39166 activates TIDA neurons suggests that D, receptor-mediated inhibition is tonically active. Considering the lack of effect of mediobasal hypothalamic deafferentation on the basal activity of TIDA neurons in male rats (Barton et al, 1989, Chapter Three) this finding suggests that the effect of this inhibitory circuit is located with in the area circumscribed by the deafferentation. If D, receptor-mediated inhibition of these neurons was dependent upon afferent input arising from outside the mediobasal hypothalamus, then deafferentation should disrupt this inhibitory tone and increase the basal activity of these neurons. Deafferentation studies could be employed to explicitly test this hypothesis, as was done in the D2 stimulatory pathway. If this hypothesis is correct then deafferentation should have no effect on the ability of SCH39166 to increase the activity of TIDA neurons. In the studies describing D, inhibitory regulation of TIDA neurons, decreases in the activity of these neurons coincided with elevated plasma prolactin levels. This finding was taken as functional evidence that the reduced activity of TIDA neurons was biologically significant and is consistent with numerous studies that have demonstrated that endogenous dopamine released from TIDA neurons is a prolactin inhibiting factor by an action at D2 receptors on the anterior pituitary (Ben-Jonathan, 1985). 181 Enhanced prolactin secretion following activation of stimulatory D, receptors on the anterior pituitary could be an alternate explanation for these findings. Many studies have associated the action of D, receptor agonists with elevated prolactin levels. A number of in vitra studies have suggested that dopamine under appropriate conditions may stimulate prolactin secretion by an action at stimulatory DA receptors located on the anterior pituitary. Indeed, this effect has been demonstrated in primary cultures of anterior pituitaries from both male and female rats (Shin, 1977; Denef at al, 1980; Kramer and Hopkins 1982; Burris at al, 1991; Hill at al, 1991). This was most elegantly demonstrated in studies where application of the non-selective agonist apomorphine to cultured pituitary cells inhibited prolactin secretion into the media. In contrast, administration of the antagonist haloperidol had no effect unless combined with apomorphine and this combination induced a 30% increase in prolactin release (Denef at al, 1980). Thus, blocking the inhibitory D2 agonist actions of apomorphine effectively converted it into a D, agonist. These findings were taken as evidence of a second DA binding site that was unique from that coupled to inhibition of prolactin secretion. Burris et al, (1991) postulated that this receptor was related to the D2 receptor since the highly selective D2 antagonist eticlopride (Seeman and Ulpian, 1988) was able to block both the inhibitory and stimulatory actions of dopamine on prolactin secretion in this system. However, the DA agonist bromocriptine alone or in the 182 presence of the D2 antagonist eticlopride failed to elicit an increase in the stimulation of prolactin in these cells. Whereas in earlier studies bromocriptine alone or in the presence of haloperidol increased the rate of prolactin secretion. Much effort has also been directed at finding the “prolactin-stimulating” DA receptor responsible for activation of prolactin secretion in vitra. In studies utilizing prolactin secreting GH4C, cells expressing D2, D,’ or D5 receptors have demonstrated that either D, or DS receptors will positively couple to prolactin secretion, whereas the D2 receptor exhibits only an inhibitory effect on prolactin secretion in these cells (Porter et al, 1994). While this study demonstrates that D, receptors are competent to mediate this response in cell culture systems, the assertion that these findings are relevant to the secretion of prolactin in viva is weakened by the failure to unequivocally demonstrate D, receptors in the anterior pituitary in rim. Thus, the prolactin stimulatory effect of dopamine in these systems may be an artifact of removing pituitary cells from their normal in rim environment possibly leading to the abnormal expression of a D, receptor that responds to physiological levels of dopamine by increasing rather than inhibiting prolactin secretion. If this is the case these culture studies may be more relevant to the mechanisms underlying trans formation of normal anterior pituitary lactotrophs to prolactin secreting tumor cells. Messenger RNA for the D5 receptor has been measured in tissue from the anterior pituitary (Porter at al, 1994) and this could be the receptor that is 183 expressed in these culture systems. However, the absence of functional D, or D5 receptor proteins in the rat anterior pituitary (Cocci at al, 1987; Albert at al, 1997) argues against involvement of these receptors in the in viva regulation of prolactin secretion. Rather, the absence of compelling evidence for a stimulatory DA receptor on the anterior pituitary suggests that the D, agonists and antagonists are having their effect on prolactin secretion via a central action. It is interesting to note that in sheep, central administration of SKF38393 increases prolactin secretion (Curlewis et al, 1994), possibly via an action in the hypothalamic ventromedial nucleus (Curlewis et al, 1995a). It has been postulated that D, receptors located in this region mediate photoperiod-induced changes in prolactin secretion in this species (Curlewis at al, 1995b). In rats the ventromedial nucleus contains a very low density of D1 receptors as identified by specific irnmunocytochernical localization studies (Fremeau et al, 1991). Furthermore, there is no evidence that DA neurons terminate in this region. Thus, D1 receptors located here may be present but “non-functional” in rats that, unlike sheep, do not respond to changes in photoperiod. Alternatively the site of action of these D1 receptor agonists and antagonists could be determined by comparing the effects of deafferentation with hypophysectomy. If the site of action is within the mediobasal hypothalamm rather than on the pituitary itself, then hypophysectomy alone should block both the stimulatory effects of 184 ii, SCH39166 and inhibitory effects of SKF 38393. On the other hand, these effects should not be disrupted by deafferentation of the mediobasal hypothalamus. Interactions between D1 and D2 receptors in TIDA neuronal regulation Researchers have long appreciated that there are important interactions between the DA receptor subtypes found in the brain. Evidence for synergistic/ cooperative interactions have been demonstrated in behavioral studies where the full expression of stereotyped behaviors observed with non- selective DA agonists can only be duplicated when D, and D2 receptor agonists are combined (Walters et al, 1987). Parallel studies examining the effects of these agents on neuronal firing rates in the striatum, and nucleus accumbens add further support to a synergistic/ cooperative interaction between D, and D2 receptors (Walters et a1, 1987). Results obtained in studies descn'bed in this dissertation lend evidence to there being an opposing interaction between D, and D2 receptor families in the regulation of TIDA neuronal activity. The discovery of drugs with greater selectivity for the various members of the DA receptor family has uncovered a heretofore-unrecognized pattern of DA regulation in these neurons. Furthermore, these results suggest that concurrent activation of dopamine D, and dopamine D2 receptors nullify their respective actions on TIDA neurons, which likely accounts for the inability of mixed D, / D2 receptor agonists to have an acute effect on these hypothalamic 185 dopamine neurons. Thus, the activity of TIDA neurons is dependent upon the relative balance of D, vs. D2 receptor influences. This balance may be important in the regulation of prolactin in certain physiological states. There is a circadian rhythmicity in the secretion of prolactin that is associated with the rapid eye movement stage of sleep (Franz 1978). In addtion, stressful situations are known to elevate prolactin (Gala 1990). Although the physiological role for prolactin in these instances is poorly understood it has been postulated that it may play a role in the maintenance of normal immune system function (Gala 1990). In either of these examples elevated prolactin may be the result of a shift in the balance of DA activity to favor the D, inhibitory path thereby leading to elevated prolactin. If this hypothesis is true then local application of D, agonists into the ventromedial nucleus ought to inhibit TIDA neurons and increase prolactin secretion. In addition, the activity of DA neurons in this nucleus ought to be elevated in association with this sleep-related elevation in prolactin. The shift in balance toward D2 stimulation of TIDA activity is independent of prolactin mediated feedback. The activation of these neurons by prolactin results from a 12-hour delay that is dependent upon the activation of neurotensin receptors (Hentschel et al, 1998). Quinelorane-induced activation of the D2 stimulatory pathway is acute and not dependent on the action of neurotensin. In fact, the stimulatory effects of prolactin and quinelroane on 186 TIDA neuronal activity are additive and this suggests an independent pharmacological mechanism of action as shown in Figure 42. The balance of DA activity may shift toward the D2 stimulatory pathway in situations following suppression of these neurons during physiological states where prolactin has been elevated i.e. stress, lactation, pre-ovulatory prolactin surge. The source of dopamine in this situation may be the neurons in the dorsomedial hypothalamic nucleus as electrical stimulation of this region has been shown to elevate TIDA neuronal activity (Gunnett et al, 1988). Interestingly, this response to electrical stimulation occurred even under conditions of hyperprolactinemia induced by administration of haloperidol. These results suggest that as in the D2 stimulatory pathway electrical stimulation of the dorsomedial nucleus is independent of prolactin induced activation of TIDA neurons also. Neurochemical experiments have not been useful for measuring changes in DA synthesis and metabolism in the ventromedial and dorsomedial nuclei. This problem may be the result of these systems being too small or the release of dopamine too minute to be discernable by this method. In this situation, dynamic neuroanatomical mapping by dual irnmunohistochemistry for FRA’s and TH may be useful in determining the relevance of DA neurons in the ventromedial or dorsomedial nuclei in the regulation of TIDA neurons. Alternatively, this technique may suggest other brain regions that may be important in the physiological regulation of these neurons. 187 T‘I'r' ‘t The results of studies described in this dissertation suggest that as drugs with greater selectivity for DA receptors become available it will be important to determine their effect on TIDA neurons in order to anticipate any side-effects due to alterations in prolactin secretion. Finally, this dopamine receptor mediated regulation of TIDA neuronal activity may allow for an easy way to determine the relative D1/D2 properties of new DA drugs as they are F‘ developed. , 188 APPENDIX 189 Table 10 List of drugs used by genedc name chemical name (source) and vehicle solvent used. Generic Name Chemical name (Source) Solvent Apomorphine 4H-Dibenzo[de,g]quinoline-10,1 1-diol,5,6,6a,7- 0.9% saline tetrahydro-6-methyl- ,hydrochloride (Sigma Chemical Co., St. Louis, MO) DPDPE [D-Pen2, D-Pen5]enkephalin (Peninsula Laboratories, Distilled water Belmont, CA.) Naltrindole 17-cyclopropyl-methyl-6,7,2’,3’-indol morphinan (Dr. Distilled water James H. Woods Department of pharmacology, University of Michigan Medical school, Ann Arbor, MI) Neurotensin PGlu-Leu-Tyr-Glu-Asn-Lys-Pro-Arg-Arg-Pro-Tyr-Ile- Distilled water Lou-OH (Sigma chemical Co., St. Louis, MO) Nor-BNI Nor-binaltorphimine hydrochloride hydrate (Research 0.03 % saline Biochemical International, Natic, MA) NSD 1015 3-Hydroxybenzylhydrazine dihydrochloride (Aldrich Distilled water Chemical Co., Milwaukee, WI) PD128907 [R-(+)-trans-3,4,4a,10b-tetrahydro-4—propyl— 2H,5H- Distilled water [1]benzopyrano[4,3-b]-1,4-oxazin-9-ol] (Dr. M. D. Davis Parke-Davis Research, Ann Arbor MI) PNU-95,666 5(R)-(methylamino)-2,4,5,6-tetrahydro~1H- 0.9% saline imidazo[4,5,1-ij]-quinolin-2- onemaleate (Dr. P. Von Voigtlander, The Upjohn Co., Kalamazoo, MI) Quinelorane Trans-(-)-5,5a, 6,7,8,9,9a, 10-octahydro-6 0.9% saline propylpyrirnido[4,5-nguinolin-2-amine dihydrochlorde (Dr. M. M. Foreman, The Eli Lilly Co., Indianapolis, IN) SCH 39166 [(-)-trans-6,7,6a,a,9,13b-hexahydro-3-chloro-2-hydroxy- 50% ethyl N-methyyl-SH benzo[d]naptho- {2,1-b}azepine (Dr. A. alcohol v/v Barnett Schereng—Plouhg, Bloomfield, NJ) SKF 38393 (i)-1-phenyl-2,4,5—tetrahydro-(1H)—3benzapine-7,8-diol 0.1% ascorbic hydrochloride (Research Biochemicals International, acid Natic, MA) SR48692 Tricyclo[3.3.1.13,7]decane-2-carboxylic acid,2- [[[1-(7- 10% DMSO chloro—4-quinolinyD-5- (2,6-dimethoxyphenyD-1H- pyrazol-3-yl]carbonyl]amino] (Dr. Danielle Gully Sanofi Recherche, Toulouse, France.) 190 Generic Name Chemical name (Source) Solvent U-50,488 Trans-(i)—3,4-dichloro-N-Methyl-N—(2- [1- Distilled water pyrrolindinyl] cyclohexyl) benzenacetamide (Dr. P. Von Voigtlander, The Upjohn Co., Kalamazoo, MI) 191 BIBLIOGRAPHY 192 Almeida, O. F., Nikolarakis, K. E., and Herz, A.: Significance of testosterone in regulating hypothalamic content and in vitro release of beta-endorphin and dynorphin. Journal of Neurochemistry 49 (3): 742-747, 1987. Alper, R H., Demarest, K. T., and Moore, K. E.: Effects of surgical sympathectomy on catecholamine concentrations in the posterior pituitary of the rat. Experientia 36 (1): 134-135, 1980. Amenomori, Y., Chen, C. L., and Meites, J.: Serum prolactin levels in rats during different reproductive states. Endocrinology 86 (3): 506-10, 1970. Andén, N. E., Carlsson, A., Dahlstrom, A., Fuxe, K., Hillarp, N. A., and Larsson, K.: Demonstration and mapping out of nigro-neostriatal dopamine neurons. Life Sciences 2: 448-458, 1964. Angrist, B., Lee, H. K., and Gershon, S.: The antagonism of amphetamine-induced symptomatology by a neuroleptic. AmJ Psychiatry 131 (7): 817-9, 1974. Aretha, C. W., Sinha, A., and Galloway, M. P.: Dopamine D3-preferring ligands act at synthesis modulating autoreceptors. Journal of Pharmacology and Experimental Therapeutics 274 (2): 609-613, 1995. Baldessarini, R. J.: Dopamine receptors and clinical medicine, ed. by K. A. Neve and R. L. Neve, pp. 457-498, Humana Press, Totowa, 1997. Bannon, M. J., Lee, J. M., Giraud, P., Young, A., Affolter, H. U., and Bonner, T. I.: Dopamine antagonist haloperidol decreases substance P, substance K, and preprotachykinin mRNAs in rat striatonigral neurons.J Biol Chem 261 (15): 6640-2, 1986. Barger, G., and Dale, H. H: Chemical structure and sympathomirnetic action of amines. Journal of Physiology 41: 19-59, 1910. Barger, G., and Ewins, A. J.: Some phenolic derivatives of b-phenylethylamine. : 2253-2261, 1910. Barnett, A., Ahn, H. S., Billard, W., Gold, E. H., Kohli,J. D., Glock, D., and Goldberg, L. 1.: Relative activities of SCH 23390 and its analogs in three tests for D1/ DA1 dopamine receptor antagonism. EurJ Pharmacol 128 (3): 249-53, 1986. Barton, A. C., Demarest, K. T., Lookingland, K. J., and Moore, K. E.: A sex difference in the stimulatory afferent regulation of tuberoinfundibular dopaminergic neuronal activity. Neuroendocrinology 49 (4): 361-366, 1989. Barton, A. C., Demarest, K. T., and Moore, K. E.: The anterior hypothalamus provides stimulatory input to tuberoinfundibular dopamine neurons which is not mediated by prolactin. Neuroendocrinology 47 (5): 416-423, 1988. Barton, A. C., Lahti, R. A., Piercey, M. F ., and Moore, K. E.: Autoradiographic identification of prolactin binding sites in rat median eminence. Neuroendocrinology 49 (6): 649- 653, 1989. Ben-Jonathan, N.: Dopamine: A prolactin-inhibition hormone. Endocrine Reviews 6 (4): 564-589, 1985. 193 Ben-Jonathan, N., Arbogast, L. A., and Hyde, J. F .: Neuroendocrine regulation of prolactin release. Progress in Neurobiology 33: 399-447, 1989. Ben-Jonathan, N., Oliver, N. C., Winer, H. J., Mica], R S., and Porter, J. C.: Dopamine in hypophysial portal plasma of the rat during the estrus cycle and thorught pregnancy. Endocrinology 100: 452-458, 1977. Berry, S. A., and Gudelsky, G. A.: D1 receptors function to inhibit the activation of tuberoinfundibular dopamine neurons. Journal of Pharmacology and Experimental Therapeutics 254 (2): 677-682, 1990. Berry, 8. A., and Gudelsky, G. A.: Effect of D2 dopamine agonists on tuberoinfundibular dopamine neurons. Neuropharmacology 30: 961-965, 1991. Bertler, A., and Rosengren, E.: Occurrence and distribution of dopamine in brain and other tissues. Experientia 15: 10-11, 1959. Billard, W., Ruperto, V., Crosby, G., Iorio, L. C., and Barnett, A.: Characterization of the binding of 3H-SCH 23390, a selective D-1 receptor antagonist ligand, in rat striatum. Life Sci 35 (18): 1885-93, 1984. Birge, C. A., Jacobs, L. 8., Hammer, C. T., and Daughaday, W. H.: Catecholamine inhibition of prolactin secretion by isolated rat adenohypophyses. Endocrinology 86 (1): 120-130, 1970. Bitran, D., Hull, E. M., Holmes, G. M., and Lookingland, K. J.: Regulation of male rat copulatory behavior by preoptic incertohypothalamic dopamine neurons. Brain Res Bull 20 (3): 323-31, 1988. BJorklund, A., and Lindvall, O.: Dopamine-containing system in the CNS. In Classical Transmitters in the CNS Part 1, ed. by A. Bjéirklund and T. Ht'ikfelt, vol. 2, pp. 55-122, Elsevier, Amsterdam, 1984. Bjérklund, A., Lindvall, O., and Nobin, A.: Evidence of an incerto-hypothalarnic dopamine neurone system in the rat. Brain Research 89 (1): 29-42, 1975. Bjorklund, A., Moore, R. Y., Nobin, A., and Stenevi, U.: The organization of tubero- hypophyseal and reticulo-infundibular catecholamine neuron systems in the rat brain. Brain Research 51: 171-191, 1973. Bjorklund, A., and Nobin, A.: Florescence histochemical and microspectroflurometric mapping of dopamine and noradrenaline cell groups in the rat diencephalon. Brain Research 51: 171-191, 1973. Blaha, C. D., and Phillips, A. G.: Pharmacological evidence for common mechanisms underlying the effects of neurotensin and neuroleptics on in vivo dopamine efflux in the rat nucleus accumbens. Neuroscience 49 (4): 867-77, 1992. Blaschko, H.: The specific action of L-DOPA decarboxylase. Journal of Physiology (London) 96: 50P-51P, 1939. Blaschko, H.: Metabolism and storage of biogenic amines. Experientia 13: 9-12, 1957. Bloom, F.: Catecholamines II: CNS aspects. In The biochemical basis of neuropharmacology 5‘“ ed., ed. by J. R. Cooper, F. E. Bloom, and R. H. Roth, Oxford University Press, New York, 1986. 194 1,: _. —. Bouthenet, M. L., Martres, M. P., Sales, N., and Schwartz, J. C.: A detailed mapping of dopamine D-2 receptors in rat central nervous system by autoradiography with [1251]iodosulpride. Neuroscience 20 (1): 117-155, 1987. Bouthenet, M. L., Souil, E., Martres, M. P., Sokoloff, P., Giros, B., and Schwartz, J. C.: Localization of dopamine D3 receptor mRNA in the rat brain using in situ hybridization histochemistry: comparison with dopamine D2 receptor mRNA. Brain Research 564 (2): 203-219, 1991. Bowers, M. B., Jr.: Central dopamine turnover in schizophrenic syndromes. Arch Gen Psychiatry 31 (1): 50-4, 1974. Bowery, B., Rothwell, L. A., and Seabrook, G. R.: Comparison between the pharmacology of dopamine receptors mediating the inhibition of cell firing in rat brain slices through the substantia nigra pars compacta and ventral tegmental area. British Journal of Pharmacology 112 (3): 873-880, 1994. Bristow, L. J., Cook, G. P., Gay, J. C., Kulagowski, J. J., Landon, L., Murray, F., Saywell, K. L. H., P.H., Young, L., and Hutson, P. H.: The behavioural and neurochemical profile of the putative dopamine D3 receptor agonist, (+)-PD 128907, in the rat. Neuropharmacology 35: 285-294, 1996. Bunney, B. S., Sesack, S. R., and Silva, N. L.: Midbrain dopaminergic systems: neurophysiology and electrophysiological pharmacology. In Psychopharmacology: The Third Generation of Progress, pp. 127-139, Raven Press, New York, 1987. Bunney, B. S., Walters, J. R., Roth, R. H., and Aghajanian, G. K.: Dopaminergic neurons: effect of antipsychotic drugs and amphetamine on single cell activity.J Pharmacol Exp Ther 185 (3): 560-71, 1973. Burris, T. P., Stringer, L. C., and Freeman, M. E.: Pharmacologic evidence that a D2 receptor subtype mediates dopaminergic stimulation of prolactin secretion from the anterior pituitary gland. Neuroendocrinology 54 (2): 175-83, 1991. Callahan, P. M., de la Garza, R., 2nd, and Cunningham, K. A.: Mediation of the discriminative stimulus properties of cocaine by mesocorticolimbic dopamine systems. Pharmacol Biochem Behav 57 (3): 601-7, 1997. Calne, D. B.: Parkinsonism. Clinical and neuropharmacologic aspects. Postgrad Med 64 (2): 82-8, 1978. Cannon, J. G.: Chemistry of dopaminergic agonists. Adv Neurol 9: 177-83, 1975. Carlsson, A.: The occurrence, distribution and physiological role of catecholamines in the nervous system. Pharmacolocical Reviews 11: 490-493, 1959. Carlsson, A.: Drugs acting through dopamine release. Pharmacol Ther [B] 1 (3): 401-5, 1975. Carlsson, A., Lindquist, M., T., M., and Waldbeck, B.: On the presence of 3 hydroxytyramine in brain. Science 127: 471, 1958. Carraway, R., and Leeman, S. E.: The isolation of a new hypotensive peptide, neurotensin, from bovine hypothalami. Journal of Biological Chemistry 248 (19): 6854-6861, 1973. 195 F . £1- Chan-Palay, V., Zaborszky, L., Kohler, C., Goldstein, M., and Palay, S. L.: Distribution of tyrosine-hydroxylase-irnmunoreactive neurons in the hypothalamus of rats. J Comp Neur01227 (4): 467-96, 1984. Chapin, D. 8., Lookingland, K. J., and Moore, K. E.: Effects of mobile phase composition on retention times for biogenic amines and thier precursors and metabolites. Current Separation 7: 68-70, 1986. Chard, T.: An Introduction to radioirnmunoassay and related techniques, Elsevier, Amsterdam, 1990. Chen, C. L., Amenomori, Y., Lu, K. H., Voogt, J. L., and Meites, J.: Serum prolactin levels in rats with pituitary transplants or hypothalamic lesions. Neuroendocrinology 6 (4): 220- 7, 1970. Cheung, S., Ballew, J. R., Moore, K. E., and Lookingland, K. J.: Contribution of dopamine neurons in the medial zona incerta to the innervation of the central nucleus of the amygdala, horizontal diagonal band of Broca and hypothalamic paraventricular nucleus. Brain Res 808 (2): 174-81, 1998. Cheung, S., Johnson, J. D., Moore, K. E., and Lookingland, K. J.: Dopamine receptor- mediated regulation of expression of Fos and its related antigens (FRA) in somatostatin neurons in the hypothalamic periventricular nucleus. Brain Res 770 (1 -2): 176-83, 1997a. Cheung, S., Will, Y. M., Hentschel, K., Moore, K. E., Lookingland, K. J: Role of gonadal steroids in determining sexual differences in expression of Fos-related antigens in tyrosine hydroxylase- irnmunoreactive neurons in subdivisions of the hypothalamic arcuate nucleus. Endocrinology 138 (9): 3804-10, 1997b. Chiodo, L. A.: Dopamine-containing neurons in the mammalian central nervous system: electrophysiology and pharmacology. Neuroscience and Biobehavioral Reviews 12 (1): 49-91, 1988. Chipkin, R. E., Iorio, L. C., Coffin, V. L., McQuade, R. D., Berger, J. G., and Barnett, A.: Pharmacological profile of SCH39166: A Dopamine D1 selective benzonaphthazepine with potential antipsychotic activity. Journal of Pharmacology and Experimental Therapeutics 247 (3): 1093-1101, 1988. Civelli, O., Bunzow, J. R., and Grandy, D. K.: Molecular diversity of the dopamine receptors. Annual Review of Pharmacology and Toxicology 33: 281-307, 1993. Civelli, O., Bunzow, J. R., Zhou, Q. Y., and Grandy, D. K.: The diversity of the dopamine receptors. NIDA Research Monographs 126: 23-33, 1992. Civelli, O., Douglass, J., Goldstein, A., and Herbert, E.: Sequence and expression of the rat prodynorphin gene. Proc Natl Acad Sci U S A 82 (12): 4291-5, 1985. Clemens, J. A., and Sawyer, B. D.: Identification of prolactin in cerebrospinal fluid. Exp Brain Res 21 (4): 399-402, 1974. Cocchi, D., Ingrassia, S., Rusconi, L., Villa, 1., and Muller, E. E.: Absence of D1 dopamine receptors that stimulate prolactin release in the rat pituitary. European Journal of Pharmacology 142 (3): 425-429, 1987. 196 lav.- Colthorpe, K. L., Anderson, S. T., Martin, G. B., and Curlewis, J. D.: Hypothalamic dopamine D1 receptors are involved in the stimulation of prolactin secretion by high environmental temperature in the female sheep.J Neuroendocrinol 10 (7): 503-9, 1998. Cools, A. R., and Van Rossum, J. M.: Excitation-mediating and inhibition-mediating dopamine-receptors: a new concept towards a better understanding of electrophysiological, biochemical, pharmacological, functional and clinical data. Psychopharmacologia 45 (3): 243-54, 1976. Creese, I., Burt, D. R., and Snyder, S. H.: Dopamine receptor binding predicts clinical and pharmacological potencies of antischizophrenic drugs. Science 192 (4238): 481-3, 1976. Creese, I., Schneider, R., and Snyder, S. H.: 3H-Spiroperidol labels dopamine receptors in pituitary and brain. EurJ Pharmacol 46 (4): 377-81, 1977. Crow, T. J., Johnstone, E. C., Deakin, J. F ., and Longden, A.: Dopamine and schizophrenia. Lancet 2 (7985): 563-6, 1976. Crow, T. J.,Johnstone, E. C., and McClelleand, H. A.: The coincidence of schizophrenia and Parkinsonism: some neurochemical implications. Psychol Med 6 (2): 227-33, 1976. Curlewis, J. D., Clarke, 1. J., and McNeilly, A. S.: Dopamine D1 receptor analogues act centrally to stimulate prolactin secretion in ewes.J Endocrinol 137 (3): 457-64, 1993. Curlewis, J. D., Thiery, J. C., and Malpaux, B.: Effect of hypothalamic infusion of a dopamine D1 receptor antagonist on prolactin secretion in the ewe. Brain Res 697 (1- 2): 48-52, 1995. Curlewis,J. D., Thiery,]. C., and Malpaux, B.: Evidence for dopamine D1 receptor-mediated stimulation of prolactin secretion in ewes under long daylength. J Reprod Fertil Suppl 49: 539-43, 1995. Dahlstrém, A., and Fuxe, K.: Evidence of the existence of monoamine containing neurons in the central nervous system. 1. Demonstration of monoamines in the cell bodies of brainstem neurons. Acta Physiol. Scand. 232: 1-55, 1964. Delay, J., Deniker, P., and Harl, J. M.: Utilisation en thérapeutique psychiatrique d'une phenothiazine d'action centrale elective. Ann. Med. Psychol. 110: 112-131, 1952. Demarest, K. T., Alper, R. H., and Moore, K. E.: Dopa accumulation is a measure of dopamine synthesis in the median eminence and posterior pituitary. Journal of Neural Transmission 46 (3): 183-193, 1979. Demarest, K. T., and Moore, K. E.: Comparison of dopamine synthesis regulation in the terminals of nigrostriatal, mesolimbic, tuberoinfundibular and tuberohypophyseal neurons. Journal of Neural Transmission 46 (4): 263-277, 1979. Demarest, K. T., and Moore, K. E.: Sexual differences in the sensitivity of tuberoinfundibular dopamine neurons to the actions of prolactin. Neuroendocrinology 33 (4): 230-234, 1981. Demarest, K. T., Riegle, G. D., and Moore, K. E.: Prolactin-induced activation of tuberoinfundibular dopaminergic neurons: evidence for both a rapid 'tonic' and a delayed 'induction' component. Neuroendocrinology 38 (6): 467-475, 1984. 197 11w Demarest, K. T., Riegle, G. D., and Moore, K. E.: Hypoprolactinemia induced by hypophysectomy and long-terrn bromocriptine treatment decreases tuberoinfundibular dopaminergic neuronal activity and the responsiveness of these neurons to prolactin. Neuroendocrinology 40 (5): 369-376, 1985. Demarest, K. T., Riegle, G. D., and Moore, K. E.: The rapid 'tonic' and the delayed 'induction’ components of the prolactin-induced activation of tuberoinfundibular dopaminergic neurons following the systemic administration of prolactin. Neuroendocrinology 43 (3): 291-9, 1986. Denef, C., Manet, D., and Dewals, R.: Dopaminergic stimulation of prolactin release. Nature 285 (5762): 243-6, 1980. Desjardins, G. C., Brawer,J. R., and Beaudet, A.: Distribution of mu, delta, and kappa opioid receptors in the hypothalamus of the rat. Brain Res 536 (1 -2): 114—23, 1990. Deutch, A. Y., Lewis, D. A., Whitehead, R. E., Elsworth,J. D., Iadarola, M. J., Redmond, D. E. J., and Roth, R. H.: Effects of D2 dopamine receptor antagonists on fos protein expression in the striatal complex and entorhinal cortex of the nonhuman primate. Synapse (3): 182-191, 1996. Dhawan, B. N., Cesselin, F., Raghubir, R., Reisine, T., Bradley, P. B., Portoghese, P. S., and Hamon, M.: International Union of Pharmacology. XII. Classification of opioid receptors. Pharmacol Rev 48 (4): 567-92, 1996. DiChiara, G., Tanda, G., and Carboni, E.: Estimation of in-vivo neurotransmitter release by brain rnicrodialysis: The issue of validity. Behavioral Pharmacology 7 (7): 640-657, 1996. Donoso, A. 0., Bishop, W., and Mch, S. M.: The effects of drugs which modify catecholamine synthesis on serum prolactin in rats with median eminence lesions. Proc Soc Exp Biol Med 143 (2): 360-363, 1973. Eaton, M. J., Gopalan, C., Kim, E., Lookingland, K. J., and Moore, K. E.: Comparison of the effects of the dopamine D2 agonist quinelorane on tuberoinfundibular dopaminergic neuronal activity in male and female rats. Brain Research 629 (1): 53-58, 1993. Eaton, M. J., Lookingland, K. J., and Moore, K. E.: Effects of the selective dopaminergic D2 agonist quinelorane on the activity of dopaminergic and noradrenergic neurons projecting to the diencephalon of the rat. Journal of Pharmacology and Experimental Therapeutics 268 (2): 645-652, 1994. Eaton, M. J., Moore, K. E., and Lookingland, K. J.: Differential effects of the D2 receptor agonist quinelorane on the secretion of prolactin and alpha-melanocyte-stimulating hormone. Life Sciences 53 (2): 107-112, 1993. Eaton, M. J., Tian, Y., Lookingland, K. J., and Moore, K. E.: Comparison of the effects of remoxipride and raclopride on nigrostriatal and mesolimbic dopaminergic neuronal activity and on the secretion of prolactin and alpha-melanocyte-stimulating hormone. Neuropsychopharmacology 7 (3): 205-211, 1992. 198 Elsworth, J. D., and Roth, R. H.: Dopamine autoreceptor pharmacology and function. In The Dopamine Receptors, ed. by K. A. Neve and R. L. Neve, pp. 223-265, Humana Press, Totowa, 1997. Ervin, G. N., Birkemo, L. S., Nemeroff, C. B., and Prange, A. J., Jr.: Neurotensin blocks certain amphetamine-induced behaviours. Nature 291 (5810): 73-6, 1981. Everett, J. W.: Luteotrophic function of autografts of the rat hypophysis. Endocrinology 54: 685-690, 1954. Everitt, B. J., Meister, B., Hékfelt, T., Melander, T., Terenius, L., Rokaeus, A., Theodorsson- Norheim, E., Dockray, G., Edwardson, J., and Cuello, C.: The hypothalamic arcuate nucleus-median eminence complex: irnmunohistochemistry of transmitters, peptides and DARPP-32 with special reference to coexistence in dopamine neurons. Brain Res 396 (2): 97-155, 1986. Falck, B., Hillarp, N.-A., Thieme, G., and Torp, A.: Fluorescence of catecholamines and related compounds condensed with formaldehyde. Journal of Histochemistry and Cytochemistry 10: 348-354, 1962. Farley, I. J., Price, K. S., and Hornykiewicz, O.: Dopamine in thelimbic regions of the human brain: normal and abnormal. Adv Biochem Psychopharmac0116: 57-64, 1977. Foreman, M. M., Fuller, R. W., Hynes, M. D., Gidda, J. S., Nichols, C. L., Schaus, J. M., Komfeld, E. C., and Clemens, J. A., and Clemens, J. A.: Preclinical studies on quinelorane a potent and highly selective D2-dopaminergic agonist. Journal of Pharmacology and Experimental Therapeutics 250 (1): 227-235, 1989. Funk, C.: Synthesis of dl-3:4~dihydroxyphenylalanine. Chemical Society Journal (London). 99(11): 554-557, 1911. Fuxe, K.: The distribution of monoamine nerve terminals in the central nervous system. Acta Physiol. Scand. 64: suppl. 247, 1965. Fuxe, K., Agnati, L. F., Kalia, M., Goldstein, M., Andersson, M., & Hirfstrand, K.: Dopaminergic systems in brain and pituitary. In: E. Fliickiger, E. E. Miiller, & M. O. Thorner (Eds), The Dopaminergic System (pp. 23-35). New York: Springer Verlag 1985. Gackenheimer, S. L., Schaus, J. M., and Gehlert, D. R.: [3H]-quinelorane binds to D2 and D3 dopamine receptors in the rat brain. Journal of Pharmacology and Experimental Therapeutics 274 (3): 1558-1565, 1995. Gala, R. R.: The physiology and mechanisms of the stress-induced changes in prolactin secretion in the rat. Life Sciences 46 (20): 1407-1420, 1990. Gibbs, D. M., and Neill, J. D.: Dopamine levels in hypophysial stalk blood in the rat are sufficient to inhibit prolactin secretion in vivo. Endocrinology 102 (6): 1895-900, 1978. Goldstein, A., and James, I. F.: Site-directed alkylation of multiple opioid receptors. II. Pharmacological selectivity. Mol Pharmacol 25 (3): 343-8, 1984. Goudreau, J. L., Lindley, S. E., Lookingland, K. J., and Moore, K. E.: Evidence that hypothalamic periventricular dopamine neurons innervate the intermediate lobe of the rat pituitary. Neuroendocrinology S6 (1): 100-105, 1992 199 Gudelsky, G. A., Annunziato, L., and Moore, K. E.: Localization of the site of the haloperidol-induced, prolactin- mediated increase of dopamine turnover in the median eminence: studies in rats with complete hypothalamic deafferentations. Journal of Neural Transmission 42 (3): 181-192, 1978. Gudelsky, G. A., Berry, S. A., and Meltzer, H. Y.: Neurotensin activates tuberoinfundibular dopamine neurons and increases serum corticosterone concentrations in the rat. Neuroendocrinology 49 (6): 604-609, 1989. Gudelsky, G. A., Koenig, J. I., Simonovic, M., Koyama, T., Ohmori, T., and Meltzer, H. Y.: Differential effects of haloperidol, clozapine, and fluperlapine on tuberoinfundibular dopamine neurons and prolactin secretion in the rat. Journal of Neural Transmission 68 (3-4): 227-240, 1987. Gudelsky, G. A., and Porter, J. C.: Morphine- and opioid peptide-induced inhibition of the release of dopamine from tuberoinfundibular neurons. Life Sci 25 (19): 1697-702, 1979. Gully, D., Canton, M., Boigegrain, R., Jeanjean, F., Molimard, J. C., Poncelet, M., Gueudet, C., Heaulme, M., Leyris, R., Brouard, A., and al, e.: Biochemical and pharmacological profile of a potent and selective nonpeptide antagonist of the neurotensin receptor. Proceedings of the National Academy of Science USA 90 (1): 65-69, 1993. Gunner, J. W., Lookingland, K. J., and Moore, K. E.: Effects of gonadal steroids on tuberoinfundibular and tuberohypophysial dopaminergic neuronal activity in male and female rats. Proc Soc Exp Biol Med 183 (1): 48-53, 1986. Gunnet,J. W., and Moore, K. B: Effect of electrical stimulation of the ventromedial nucleus and the dorsomedial nucleus on the activity of tuberoinfundibular dopaminergic neurons. Neuroendocrinology 47 (1): 20-26, 1988. Hilasz, B., and Pupp, L.: Hormone secretion of the anterior pituitary gland after physical interruption of all nervous pathways to the hypophysiotrophic area. Endocrinology 77: 553-562, 1965. Halasz, B., Pupp, L., and Uhlarik, S.: Journal of Endocrinology 25: 147-, 1962. Hentschel, K., Cheung, S., Moore, K. E., and Lookingland, K. J.: Pharmacological evidence that neurotensin mediates prolactin-induced activation of tuberoinfundibular dopamine neurons. Neuroendocrinology 68 (2): 71-6, 1998. Hoffman, G. E., Le, W. W., Abbud, R., Lee, W. S., and Smith, M. 8.: Use of Fos-related antigens (FRAs) as markers of neuronal activity: FRA changes in dopamine neurons during proestrus, pregnancy and lactation. Brain Research 654 (2): 207-215, 1994. Hoffman, G. E., Lee, W. S., Smith, M. S., Abbud, R., Roberts, M. M., Robinson, A. G., and Verbalis, J. G.: c-Fos and Fos-related antigens as markers for neuronal activity: perspectives from neuroendocrine systems. NIDA Research Monographs 125: 117- 133, 1993. Hékfelt, T., and Fuxe, K.: Effects of prolactin and ergot alkaloids on the tubero- infundibular dopamine (DA) neurons. Neuroendocrinology 9: 100-122, 1972. 200 Holtz, P., and Crender, K.: Die enzymatische entstehung von oxytyramin irn organismus und die physiologische bedeutung der dopadecarboxylase. Naunyn-Schmiedeberg's Arch. Exp. Path. Pharmak. 200: 356-388, 1942. Holtz, P., Heise, R., and Liidtke, K.: Naunyn-Schmiedeberg's Arch. Exp. Path. Pharmak. 191: 87-118, 1938. Hornykiewicz, O.: The action of dopamine on the arterial blood pressuer of the guinea pig. British Journal of Pharmacology 13: 91-94, 1958. Hughes, P., and Dragunow, M.: Induction of irnmediate-early genes and the control of neurotransmitter-regulated gene expression within the nervous system. Pharmacology Reviews 47 (1): 133-178, 1995. Imperato, A., and Di Chiara, G.: Dopamine release and metabolism in awake rats after systemic neuroleptics as studied by trans-striatal dialysis. Journal of Neuroscience 5 (2): 297-306, 1985. Imperato, A., Tanda, G., Frau, R., DiChiara, G., and Di Chiara, G.: Pharmacological profile of dopamine receptor agonists as studied by brain dialysis in behaving rats. Journal of Pharmacology and Experimental Therapeutics 245 (1): 257-264, 1988. Iorio, L. C., Barnett, A., Leitz, F. H., Houser, V. P., and Korduba, C. A.: SCH 23390, a potential benzazepine antipsychotic with unique interactions on dopaminergic systems. J Pharmacol Exp Ther 226 (2): 462-8, 1983. Jackson, D. M., and Westlind-Danielsson, A.: Dopamine receptors: Molecular biology, biochemistry and behavioural aspects. Pharmacology and Therapeutics 64: 291-369, 1994. Johnston, C. A., Demarest, K. T., and Moore, K E.: Cycloheximide disrupts the prolactin- mediated stimulation of dopamine synthesis in tuberoinfundibular neurons. Brain Research 195 (1): 236-240, 1980. Johnstone, E. C., Crow, T. J., Frith, C. D., Carney, M. W., and Price, J. S.: Mechanism of the antipsychotic effect in the treatment of acute schizophrenia. Lancet 1 (8069): 848-51, 1978. Kahn, D., Abrams, G. M., Zimmerman, E. A., Carraway, R., and Leeman, S. E.: Neurotensin neurons in the rat hypothalamus: an immunocytochemical study. Endocrinology 107 (1): 47-54, 1980. Kamberi, I. A., Mica], R. 8., and Porter, J. C.: Pituitary portal vessel infusion of hypothalamic extract and release of LH, FSH and prolactin. Endocrinology 88 (6): 1294-9, 1971. Kanematsu, S., Hilliard, J., and Sawyer, C. H.: Effect of reserpine on pituitary prolactin content and its hypothalamic site of action in the rabbit. Acta Endocrinology (Kbh.) 44: 467-474, 1963. Kanematsu, S., and Sawyer, C. H.: Elevation of plasma prolactin after hypophysial stalk section in the rat. Endocrinology 93 (1): 238-41, 1973. Katzman, R., Makman, M. H., Ahn, H. S., Mishra, R. K., and Gardner, E.: Effects of drugs and lesions on dopamine-stimulated adenylate cyclase: evidence for different classes of dopamine receptors. Trans.Am.Neurol.Assoc. 102:76-9: 76-79, 1977. 201 Kebabian, J. W., and Calne, D. B.: Multiple receptors for dopamine. Nature 277 (5692): 93- 96, 1979. Kenakin, T. P.: Drug Receptor Theory. In Pharmacologic Analysis of Drug-Receptor Interaction, pp. 1-38, Raven Press, New York, 1993. Kline, N. S.: Presidential Address. In Psychopharrnacology; a review of progress, 1957-1967; proceedings of the 6th annual meeting of the American College of Neuropsychopharmacology (pp. 1-3). D. H. Efron. (Ed), Washington, D.C.: US. Govt. Printing Office 1968. Kobayashi, R. M., Brown, M., and Vale, W.: Regional distribution of neurotensin and somatostatin in rat brain. Brain Research 126 (3): 584-588, 1977. Koch, Y., Lu, K. H., and Meites, J.: Biphasic effects of catecholamines on pituitary prolactin release in vitro. Endocrinology 87 (4): 673-675, 1970. Koenig, J. I., Mayfield, M. A., McCann, S. M., and Krulich, L.: On the prolactin-inhibiting effect of neurotensin. The role of dopamine. Neuroendocrinology 35 (4): 277-281, 1982. Konig, J. F. R., and Klippel, R. A.: . In The Rat Brain: A Stereotaxic Atlas of the Forebrain and Lower Parts of the Brain Stem, Krieger Huntington, New York, 1963. Kramer, I. M., and Hopkins, C. R.: Studies on the kinetics of dopamine-regulated prolactin secretion. Mol Cell Endocrin0128 (2): 191-8, 1982. Krulich, L., Hefco, E., and Aschenbrenner, J. E.: Mechanism of the effects of hypothalamic deafferentation on prolactin secretion in the rat. Endocrinology 96 (1): 107-118, 1975. Laborit, H.: Sur l'utisation de certaines agents pharmacodynamiques 5 action neuro- végétative en periode per- et postopératoire. Acta Chir. Belg. 48: 485-492, 1949. Lee, W. S., Abbud, R., Smith, M. 8., and Hoffman, G. E.: LHRH neurons express cJun protein during the proestrous surge of luteinizing hormone. Endocrinology 130 (5): 3101-3, 1992. Lieberman, A. N ., and Goldstein, M.: Bromocriptine in Parkinson disease. Pharmacological Reivews 37: 217-227, 1985. Lieberman, A. N., Goldstein, M., and Leibowitz, M.: Treatrnent of advanced Parkinson disease with pergolide. Neurology 31: 675-682, 1981. Lindley, S. E., Gunnet, J. W., Lookingland, K. J., and Moore, K. E.: 3,4- Dihydroxyphenylacetic acid concentrations in the intermediate lobe and neural lobe of the posterior pituitary gland as an index of tuberohypophysial dopaminergic neuronal activity. Brain Research 506 (1): 133-138, 1990. Lindvall, O., Bjorklund, A., Moore, R. Y., and Stenevi, U.: Mesencephalic dopamine neurons projecting to neocortex. Brain Res 81 (2): 325-31, 1974. Liposits, Z., and Paull, W. K.: Association of dopaminergic fibers with corticotropin releasing hormone (CRIB-synthesizing neurons in the paraventricular nucleus of the rat hypothalamus. Histochemistry 93 (2): 119-27, 1989. 202 Liu, Y. F., Civelli, O., Grandy, D. K., and Albert, P. R.: Differential sensitivity of the short and long human dopamine D2 receptor subtypes to protein kinase C. Journal of Neurochemistry 59 (6): 2311-2317, 1992. Liu, Y. F., Civelli, 0., Zhou, Q. Y., and Albert, P. R: Cholera toxin-sensitive 3',5'-cyclic adenosine monophosphate and calcium signals of the human dopamine-D1 receptor: selective potentiation by protein kinase A. Molecular Endocrinology 6 (11): 1815-1824, 1992. Liu, Y. F., Jakobs, K. H., Rasenick, M. M., and Albert, P. R.: G protein specificity in receptor-effector coupling. Analysis of the roles of G0 and Gi2 in GH4C1 pituitary cells.J Biol Chem 269 (19): 13880-6, 1994. Lookingland, K. J., Farah, J. M., Jr., Lovell, K. L., and Moore, K. E.: Differential regulation of tuberohypophysial dopaminergic neurons terminating in the intermediate lobe and in the neural lobe of the rat pituitary gland. Neuroendocrinology 40 (2): 145-151, 1985. Lookingland, K. J., and Moore, K. E.: Dopamine receptor-mediated regulation of incertohypothalamic dopaminergic neurons in the male rat. Brain Research 304 (2): 329-338, 1984. Lowry, O. H., Rosebrough, N. J., Farr, A. L., and Randall, R. J.: Protein measurement with the folin phenol reagent. Journal of Biological Chemistry 193: 265-275, 1951. Lu, K H., and Meites, J.: Effects of L-dopa on serum prolactin and PIF in intact and hypophysectomized, pituitary-grafted rats. Endocrinology 91 (4): 868-72, 1972. Luppi, P. H., Sakai, K., Salvert, D., Berod, A., and Jouvet, M.: Periventricular dopaminergic neurons terminating in the neuro- intermediate lobe of the cat hypophysis. J Comp Neur01244 (2): 204-12, 1986. MacKenzie, F. J., Hunter, A. J., Daly, C., and Wilson, C. A.: Evidence that the dopaminergic incerto-hypothalarnic tract has a stimulatory effect on ovulation and gonadotrophin release. Neuroendocrinology 39 (4): 289-95, 1984. MacLeod, R. M.: Regulation of prolactin secretion. In Frontiers in Neuroendocrinology, ed. by L. Martini and W. F. Ganong, vol. 4, pp. 170-194, Raven Press, New York, 1976. MacLeod, R. M., and Lehmeyer, J. E.: Studies on the mechanism of the dopamine-mediated inhibition of prolactin secretion. Endocrinology 94 (4): 1077-1085, 1974. Mailman, R. B., Nichols, D. E., and Tropsha, A.: Molecular drug design and dopamine receptors, ed. by K. A. Neve and R. L. Neve, pp. 105-133, Humana Press, Totowa, 1997. Mannich, C., and Jacobsohn, W.: Uber oxyphenyl-alkylamine und dioxyphenyl-alkylamine. Ber. Deutsch. Cehm. Ges. 43: 189-197, 1910. Mansour, A., Khachaturian, H., Lewis, M. E., Akil, H., and Watson, S. J.: Anatomy of CNS opioid receptors. Trends Neurosci 11 (7): 308-14, 1988. Manzanares, J., Lookingland, K. J., and Moore, K. E.: Kappa opioid receptor-mediated regulation of dopaminergic neurons in the rat brain. Journal of Pharmacology and Experimental Therapeutics 256 (2): 500-505, 1991. 203 Manzanares, J., Wagner, E. J., LaVigne, S. D., Lookingland, K. J., and Moore, K. E.: Sexual differences in kappa opioid receptor-mediated regulation of tuberoinfundibular dopaminergic neurons. Neuroendocrinology 55 (3): 301-307, 1992. Manzanares, J., Wagner, E. J., Lookingland, K. J., and Moore, K. E.: Effects of irnmunoneutralization of dynorphin1_17 and dynorphin1_8 on the activity of central dopaminergic neurons in the male rat Effects of irnmunoneutralization of dynorphin1- 17 and dynorphin1- 8 on the activity of central dopaminergic neurons in the male rat. Brain Research 587 (2): 301-305, 1992. Martin, W. R.: A steric theory of opioid agonists, antagonists, agonist-antagonists, and partial agonists. NIDA Res Monogr 49: 16-23, 1984. Matthysse, S.: Dopamine and the pharmacology of schizophrenia: the state of the evidence.J Psychiatr Res 11: 107-13, 1974. McQuade, R. D., Duffy, R. A., Coffin, V. L., Chipkin, R. E., and Barnett, A.: In vivo binding of SCH 39166: a D-1 selective antagonist.J Pharmacol Exp Ther 257 (1): 42-9, 1991. Meites, J., Lu, K. H., Wuttke, W., Welsch, C. W., Nagasawa, H., and Quadri, S. K.: Recent studies on functions and control of prolactin secretion in rats. Recent Prog Horm Res 28: 471-526, 1972. Merchenthaler, 1.: Induction of enkephalin in tuberoinfundibular dopaminergic neurons during lactation. Endocrinology 133 (6): 2645-2651, 1993. Merchenthaler, 1.: Induction of enkephalin in tuberoinfundibular dopaminergic neurons of pregnant, pseudopregnant, lactating and aged female rats. Neuroendocrinology 60 (2): 185-193, 1994. Merchenthaler, I., Culler, M. D., Petrusz, P., Flerko, B., and Negro-Vilar, A.: Immunocytochemical localization of the gonadotropin-releasing hormone- associated peptide portion of the LHRH precursor in the hypothalamus and extrahypothalamic regions of the rat central nervous system. Cell Tissue Res 255 (1): 5-14, 1989. Merchenthaler, I., Lennard, D. E., Cianchetta, P., Merchenthaler, A., and Bronstein, D.: Induction of proenkephalin in tuberoinfundibular dopaminergic neurons by hyperprolactinemia: the role of sex steroids. Endocrinology 136 (6): 2442-2450, 1995. Missale, C., Nash, S. R., Robinson, S. W.,Jaber, M., and Caron, M. G.: Dopamine receptors: from structure to function. Physiol Rev 78(1), 189-225, 1998. Montagu, K. A.: Catechol compounds in rat tissues and in brains of different animals. Nature 180: 244-245, 1957. Moore, K. E.: Hypothalamic dopaminergic neuronal systems. In Psychopharmacology: The Third Generation of Progress, ed. by H. Y. Meltzer, pp. 127-140, Raven Press, New York, 1987. Moore, K. E., and Lookingland, K. J.: . In Psychopharmacology: the Fourth Generation of Progress, ed. by F. E. Bloom and D. J. Kupfer, Raven Press, New York, 1994. Morgan, W. W., Herbert, D. C., and Pfeil, K. A.: The effect of hypophysectomy and subsequent prolactin replacement or of elevated prolactin alone on median eminence noradrenaline and dopamine in the rat. Endocrinology 110 (5): 1584-1591, 1982. 204 Mulder, A. H., Wardeh, G., Hogenboom, F ., and Frankhuyzen, A. L.: Selectivity of various opioid peptides towards delta-, kappa; and mu-opioid receptors mediating presynaptic inhibition of neurotransmitter release in the brain. Neuropeptides 14 (2): 99-104, 1989. Murai, I., Garris, P. A., and Ben-Jonathan, N.: Time-dependent increase in plasma prolactin after pituitary stalk section: role of posterior pituitary dopamine. Endocrinology 124 (5): 2343-9, 1989. Nemeroff, C. B.: Neurotensin: perchance an endogenous neuroleptic? Biol Psychiatry 15 (2): 283-302, 1980. Nemeroff, C. B.: The interaction of neurotensin with dopaminergic pathways in the central nervous system: basic neurobiology and implications for the pathogenesis and treatment of schizophrenia. Psychoneuroendocrinology 11 (1): 15-37, 1986. Nkitovitch-Winer, M. B.: Effect of hypopysial stalk transection on luteotropic hormone secretion in the rat. Endocrinology 77: 658-666, 1965. Nordstréim, A. L., Farde, L., Wiesel, F. A., Forslund, K., Pauli, S., Halldin, C., and Uppfeldt, G.: Central D2-dopamine receptor occupancy in relation to antipsychotic drug effects: a double-blind PET study of schizophrenic patients. Biol Psychiatry 33 (4): 227-235, 1993. Oades, R. D., and Halliday, G. M.: Ventral tegmental (A10) system: neurobiology. 1. Anatomy and connectivity. Brain Research 434 (2): 117-165, 1987. Ogren, S. O., and Fuxe, K.: D1- and D2-receptor antagonists induce catalepsy via different efferent striatal pathways [corrected] [published erratum appears in Neurosci Lett 1988 Jun 29;89(2):258]. Neurosci Lett 85 (3): 333-8, 1988. Oliver, C., Mica], R. S., and Porter, J. C.: Hypothalamic-pituitary vasculature: evidence for retrograde blood flow in the pituitary stalk. Endocrinology 101 (2): 598-604, 1977. Palkovits, M.: Isolated removal of hypothalamic or other brain nuclei of the rat. Brain Research 59: 4449-4450, 1973. Pan, J. T., Tian, Y., Lookingland, K. J., and Moore, K. E.: Neurotensin-induced activation of hypothalamic dopaminergic neurons is accompanied by a decrease in pituitary secretion of prolactin and alpha-melanocyte-stimulating hormone. Life Sciences 50 (25): 2011-2017, 1992. Pasinetti, G., Govoni, 8., di Giovine, S., Spano, P. F., and Trabucchi, M.: Dopamine enhances Met-enkephalin efflux from rat striatal slices. Brain Research 293 (2): 364- 367, 1984. Pasteels, J. L.: Sécrétion de prolactine par l'hypophyse en culture de tissus. C. R. Acad. Sci. (Paris) 253: 2140, 1961. Paterson, S. J., Robson, L. E., and Kosterlitz, H. W.: Classification of opioid receptors. Br Med Bull 39 (1): 31-6, 1983. Paxinos, G., and Watson, C.: The Rat Brain: in Stereotaxic Coordinates, Academic Press, New York, 1986. 205 Pehek, E. A., Warner, R. K., Bazzett, T. J., Bitran, D., Band, L. C., Eaton, R. C., and Hull, E. M.: Microinjection of cis-flupenthixol, a dopamine antagonist, into the medial preoptic area impairs sexual behavior of male rats. Brain Res 443 (1 -2): 70-6, 1988. Pickel, V. M., Chan, J., and Sesack, S. R.: Cellular basis for interactions between catecholaminergic afferents and neurons containing Leu-enkephalin-like irnmunoreactivity in rat caudate-putamen nuclei. Journal of Neuroscience Research 31 (2): 212-230, 1992. Popper, P., Abelson, L., and Micevych, P. E.: Differential regulation of alpha-calcitonin gene-related peptide and preprocholecystokinin messenger RNA expression in alpha- motoneurons: effects of testosterone and inactivity induced factors. Neuroscience 51 (1): 87-96, 1992. Popper, P., Ulibarri, C., and Micevych, P. E.: The role of target muscles in the expression of calcitonin gene- related peptide mRNA in the spinal nucleus of the bulbocavemosus. Brain Research Molecular Brain Research 13 (1 -2): 43-51, 1992. Porter, T. E., Grandy, D., Bunzow, J., Wiles, C. D., Civelli, O., and Frawley, L. 8.: Evidence that stimulatory dopamine receptors may be involved in the regulation of prolactin secretion. Endocrinology 134 (3): 1263-1268, 1994. Pugsley, T. A., Davis, M. D., Akunne, H. C., Mackenzie, R. G., Shih, Y. H., Damsma, G., Wikstrom, H., Whetzel, S. Z., Georgie, L. M., Cooke, L. W., Demattos, S. B., Corbin, A. E., Glase, S. A., Wise, L. D., Dijkstra, D., and Heffiier, T. G.: Neurochemical and functional characterization of the preferentially selective dopamine D3 agonist PD 128907. Journal of Pharmacology and Experimental Therapeutics 275 (3): 1355-1366, 1995. Richtand, N. M., Kelsoe, J. R, Sega], D. S., and Kuczenski, R.: Regional quantification of D1, D2, and D3 dopamine receptor mRNA in rat brain using a ribonuclease protection assay. Brain Research Molecular Brain Research 33 (1): 97-103, 1995. Roth, R. H., and Elsworth,J. D.: Biochemical pharmacology of mide dopamine neurons. In Psychopharmacology: The Fourth Generation of Progress, ed. by F. E. Bloom and D. J. Kupfer, pp. 227-243, Raven Press, New York, 1996. Roth, R. H., Murrin, L. C., and Walters, J. R.: Central dopaminergic neurons: effects of alterations in impulse flow on the accumulation of dihydroxyphenylacetic acid. EuropeanJournal of Pharmacology 36 (1): 163-171, 1976. Roufogalis, B. D., Thornton, M., and Wade, D. N.: Specificity of the dopamine sensitive adenylate cyclase for antipsychotic antagonists. Life Sci 19 (6): 927-34, 1976. Saller, C. F ., and Salama, A. I.: D-1 and D-2 dopamine receptor blockade: interactive effects in vitro and in vivo. Journal of Pharmacology and Experimental Therapeutics 236 (3): 714-720, 1986. Saller, C. F ., and Salama, A. 1.: D-1 dopamine receptor stimulation elevates plasma prolactin levels. European Journal of Pharmacology 122 (1): 139-142, 1986. Sanghera, M. K., Anselmo Franci, J., and Mch, S. M.: Effect of medial zona incerta lesions on the ovulatory surge of gonadotrophins and prolactin in the rat. Neuroendocrinology 54 (5): 433-438, 1991. 206 Schumann, H. J.: Nachweis von oxytyramin (dopamin) in sympathischen nerven und ganglien. Arch Exp. Path. Pharmak. 227: 556-573, 1956. Schwartz, J. C.: Multiple dopamine receptors: functional implications. Clin Neuropharmacol 15 (Suppl 1 Pt A): 1A-2A, 1992. Seabrook, G. R., Patel, 8., Hutson, P. H., Ragan, C. I., and Bristow, L. J.: Functional relevance of dopamine receptors [letter]. Trends in Pharmacological Sciences 16 (12): 406-408, 1995. Seeman, P., Chan-Wong, M., Tedesco, J., and Wong, K.: Brain receptors for antipsychotic drugs and dopamine: direct binding assays. Proc Natl Acad Sci U S A 72 (11): 4376-80, 1975. Seeman, P., Lee, T., Chan-Wong, M., and Wong, K.: Antipsychotic drug doses and neuroleptic / dopamine receptors. Nature 261 (5562): 717-9, 1976. Seeman, P., and Ulpian, C.: Dopamine D1 and D2 receptor selectivities of agonists and antagonists. Adv Exp Med Bi01235: 55-63, 1988. Selmanoff, M.: The lateral and medial median eminence: distn'bution of dopamine, norepinephrine, and luteinizing hormone-releasing hormone and the effect of prolactin on catecholamine turnover. Endocrinology 108 (5): 1716-22, 1981. Setler, P. E., Sarau, H. M., Zirkle, C. L., and Saunders, H. L.: The central effects of a novel dopamine agonist. European Journal of Pharmacology 50 (4): 419-430, 1978. Shaar, C. J., and Clemens, J. A.: The role of catecholamines in the release of anterior pituitary prolactin in vitro. Endocrinology 95 (5): 1202-1212, 1974. Shin, S. H.: Dopamine-induced inhibition of prolactin release from cultured adenohypophysial cells: Spare receptors for dopamine. Life Sciences 22: 67-74, 1978. Sibley, D. R., Leff, S. E., and Creese, 1.: Interactions of novel dopaminergic ligands with D-1 and D-2 dopamine receptors. Life Sciences 31 (7): 637-645, 1982. Smith, M. W., Bienkowski, M. J., Darlington, W. H., Dinh, D. M., Harris, D. W., Huff, R. M., Lajiness, M. E., Lawson, C. F., McCall, R. B., McGuire, J. C., Moon, M. W., and Schlachter, S. K.: Monoaminergic receptor profiles of imidazoquinolinones: Agonists potentially useful for the treatment of Parkinson's Disease. Society for Neuroscience Abstracts 21: 853, 1996. Spanagel, R., Herz, A., and Shippenberg, T. S.: The effects of opioid peptides on dopamine release in the nucleus accumbens: an in vivo microdialysis study. Journal of Neurochemistry 55 (5): 1734-1740, 1990. Spano, P. F ., Govoni, S., and Trabucchi, M.: Studies on the pharmacological properties of dopamine receptors in various areas of the central nervous system. Advances in Biochemistry and Psychopharmacology 19:155-65: 155-165, 1978. Swazey, J. P.: CPZ Enters Psychiatry: France, 1951-1952. In Chlorpromazine in Psychiatry, pp. 111-141, The MIT Press, Cambridge, 1974a. Swazey, J. P.: Henri Laborit: The nature and management of shock. In Chlorpromazine in Psychiatry, pp. 61-81, The MIT Press, Cambridge, 1974b. 207 Takahara, J., Arimura, A., and Schally, A. V.: Suppression of prolactin release by a purified porcine PIF preparation and catecholamines infused into a rat hypophysial portal vessel. Endocrinology 95 (2): 462-465, 1974. Talwalker, P. K., Ratner, A., and Meites, J.: In vitro inhibition of pituitary prolactin synthesis and release by hypothalmic extract. American Journal of Physiology 205 (2): 213-218, 1963. Tice, M. A., Hashemi, T., Taylor, L. A., Duffy, R. A., and McQuade, R. D.: Characterization of the binding of SCH 39166 to the five cloned dopamine receptor subtypes. Pharmacology, Biochemistry, and Behavior 49 (3): 567-571, 1994. Toney, T. W., Lookingland, K. J., and Moore, K. E.: Role of testosterone in the regulation of tuberoinfundibular dopaminergic neurons in the male rat. Neuroendocrinology 54 (1): 23-29, 1991. Turpen, C., and Dunn, J. D.: The effect of surgical isolation or ablation of the medial basal hypothalamus on serum prolactin levels in male rats. Neuroendocrinology 20 (3): 224- 234, 1976. Uhl, G. R., Kuhar, M. J., and Snyder, S. H.: Neurotensin: immunohistochemical localization in rat central nervous system. Proc Natl Acad Sci U S A 74 (9): 4059-63, 1977. van den Po], A. N., Herbst, R. S., and Powell, J. F.: Tyrosine hydroxylase-immunoreactive neurons of the hypothalamus: a light and electron microscopic study. Neuroscience 13 (4): 1117-56, 1984. Van Maanen, J. H., and Smelik, P. G.: Induction of pseudopregnancy in rats following local depletion of monoamines in the median eminence of the hypothalamus. Neuroendocrinology 3 (3): 177-186, 1968. Verbalis, J. G., Stricker, E. M., Robinson, A. G., and Hoffman, G. B: Journal of Neuroendocrinology 3: 205-213, 1991 . Wagner, C. K., Eaton, M. J., Moore, K. E., and Lookingland, K. J. Efferent projections from the region of the medial zona incerta containing A13 dopaminergic neurons: a PHA-L anterograde tract- tracing study in the rat. Brain Research 677 (2): 229-237, 1995. Wagstaff, J. D., Bush, L. G., Gibb, J. W., and Hanson, G. R.: Endogenous neurotensin antagonizes methamphetamine-enhanced dopaminergic activity. Brain Research 665 (2): 237-244, 1994. Wagstaff, J. D., Gibb, J. W., and Hanson, G. R.: Dopamine D2-receptors regulate neurotensin release from nucleus accumbens and striatum as measured by in vivo microdialysis. Brain Research 721 (1 -2): 196-203, 1996. Walsh, R. J., Posner, B. I., Kopriwa, B. M., and Brawer, J. R.: Prolactin binding sites in the rat brain. Science 201 (4360): 1041 -3, 1978. Wamsley, J., Hunt, M. E., McQuade, R. D., Alburges, M. E., and Wamsley, J. K.: [3H]SCH39166, a D1 dopamine receptor antagonist: Binding characteristics and localization. Experimental Neurology 111 (2): 145-151, 1991. 208 Welsch, C. W., Squiers, M. D., Cassel], B, Chen, C. L., and Meites, J. Median eminence lesions and serum prolactin: influence of ovariectomy and ergocomine. American Journal of Physiology 221: 1714-1717, 1971. Young, W. S., Bonner, T. 1., and Brann, M. R.: Mesencephalic dopamine neurons regulate the expression of neuropeptide mRNAs in the rat forebrain. Proceedings of the National Academy of Science USA 83 (24): 9827-9831, 1986. Zhou, Q. Y., Grandy, D. K, Thambi, L., Kushner,J. A., Van Tol, H. H., Cone, R., Pribnow, D., Salon, J., Bunzow, J. R., and Civelli, O.: Cloning and expression of human and rat D1 dopamine receptors. Nature 347 (6288): 76-80, 1990. 209 HIGRN STATE UNIV. lllsllljll llllLl ollzll Ill lellllgllllllllllllll