I... ELISE": .- 315. . .1 5: 15.1% 51...: 3.. . ml. 9. 3:7...3-1 . C l a 5 ... In 1!.11 ‘. :3 l... . :5: ‘5 ‘ . xii}: u 25!.‘03 v El :0. .‘uflWF. 2. . “an. £15.11 £3.6an 4‘. if ~11 hi I— I! v.“ \K. t ‘ THESIQ LIBRARY 100% Michigan State University This is to certify that the dissertation entitled DYNAMICS OF FAST SYNAPTIC EXCITATION IN MYENTERIC NEURONS presented by JIANHUA JIM REN has been accepted towards fulfillment of the requirements for the Ph.D. degree in Neuroscience /3 Z M Ei‘rgpyésor‘gsguy 9% ”-5-— § - Date L/ MSU is an Affinnative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/ClRC/DaleDue.indd-p.1 DYNAMICS OF FAST SYNAPTIC EXCITATION IN MYENTERIC NEURONS By Jianhua Jim Ren A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR of PHILOSOPHY Neuroscience Program 2006 ABSTRACT DYNAMICS OF FAST SYNAPTIC EXCITATION IN MYENTERIC NEURONS By J ianhua Jim Ren The goal of this dissertation is to understand the dynamic properties of fast excitatory synaptic transmission in myenteric neurons in the small intestine. Three specific issues have been addressed: 1) identification of purinergic P2X receptor subtypes expressed in myenteric neurons; 2) characterization of fast excitatory synaptic transmission during bursts of synaptic activity and 3) the cellular and molecular mechanisms by which presynaptic 5-HT4 receptors mediate facilitation of fast synaptic transmission. Two types of electrophysiological recording methods were used: 1) intracellular microelectrode recordings from single myenteric neurons in acutely isolated longitudinal muscle-myenteric plexus (LMMP) preparations and 2) patch-clamp recordings from single myenteric neurons maintained in primary culture. Identification of P2X receptor subtypes was conducted in acutely isolated LMMP preparations in vitro from the small intestine of guinea pigs and mice which P2X2 or P2X; gene was knockout. My data suggest that P2X receptors were P2X; homomers in myenteric S neurons and P2X2/3 heteromers in AH neurons in mouse ileum. However, in guinea pig ileum, P2X receptors in myenteric S neurons contained P2X3 subunits. P2X receptors in AH neurons did not contain P2X; subunits. My data also showed that P2X; subunit-containing receptors in myenteric S neurons of guinea pig ileum coupled to Ca2+- activated 1Cr channels to mediate a hyperpolarization that followed agonist-induced depolarization. Fast excitatory postsynaptic potentials (fEPSPs) evoked in guinea pig LMMP preparations in vitro declined (rundown) during trains of stimulation. The rundown rate was frequency-dependent. Neither desensitization of nAChRs or P2X receptors nor receptor-mediated presynaptic inhibition contributed to rundown of fEPSP. Depletion of transmitter vesicles in the readily releasable pool in presynaptic terminals was proposed as a mechanism to account for synaptic rundown. 5-HT4 receptors are localized to presynaptic terminals in myenteric neurons. In my study, 5-HT4 agonists increased fast EPSP amplitude during trains of stimulation and facilitated recovery afier rundown. The effects of 5-HT4 receptor agonists on fEPSPs were mediated through activation of the adenylate cyclase-cAMP-PKA signaling pathway. Detector-patch studies in cultured myenteric neurons suggested that 5-HT4 agonists facilitated fast synaptic transmission by increasing neurotransmitter release probability from single varicosities. These results provide novel information on pre- and postsynaptic mechanisms regulating synaptic transmission in the enteric nervous system. Rundown of fEPSPs during bursts may be a presynaptic mechanism to limit the synapses from over-activity. After-hyperpolarization mediated by P2X receptors and facilitation mediated by 5-HT4 receptors may provide the cellular mechanisms by which neurotransmitters or drugs that can mimic neurotransmitters’ action can alter neuronal activity. In the View of practice, the differential expression of P2X subunits in myenteric neurons will help to select drug targets for development of novel treatment of GI motility disorders. Studies of the mechanisms of 5-HT4 receptor-mediated facilitation of synaptic transmission may provide insights into the neurophysiological basis of GI motility disorders. ACKNOWLEDGEMENTS First and foremost I would like to thank my mentor Dr. James J. Galligan for giving me the opportunity working in his laboratory. 1 have benefited for my scientific career and daily life so much from his patient teaching, his wise advice and the excellent environment he provides. These past few years working with him will be an invaluable experience for rest of my life. I would also like to thank my thesis advisory committee members for their helpful discussion, constructive suggestions and encouragement. They were: Drs. Steven Hedeimann, David Kreulen and Mary Rheuben. I dedicate this dissertation to my lovely wife, as she has always been a loyal audience for my work. iv TABLE OF CONTENTS LIST OF TABLES viii LIST OF FIGURES ix INTRODUCTION 1 The enteric nervous system 2 Overview 2 Anatomy of the GI tract and arrangement of the enteric plexuses ---- 4 Morphology and ultrastructure of enteric neurons 7 Electrophysiology of myenteric neurons 8 Functional classification of enteric neurons 15 Intrinsic neural circuits in the myenteric plexus l7 Extrinsic innervation 20 Submucosal neurons 21 Synaptic transmission 23 Overview 23 Slow synaptic transmission in enteric neurons 24 I Fast synaptic transmission and ligand-gated ion channels in myenteric neurons 25 Dynamics of neurotransmitter release 31 Recycle of synaptic vesicles 33 5-HT4 receptors in myenteric neurons 36 Research goal 38 Specific aims 39 METHODS 42 Animals 42 Longitudinal muscle myenteric plexus preparations 42 Myenteric neurons maintained in primary culture 43 Intracellular recording from single myenteric neurons 44 Patch-clamp recording from primary cultured myenteric neurons ------------ 45 Immunohistochemistry 46 Drugs and drug application 46 Statistics 49 RESULTS 50 Subtypes of P2X receptor in myenteric neurons in guinea pig and mouse small intestine 51 Part A. P2X receptor subtypes in myenteric neurons in mouse small intestine 51 Characteristics of myenteric neurons from P2X2+/+, P2X2'l', P2X3H+ and P2X;”' mice 51 P2X2 subunits knockout altered P2X-meidated responses in S neurons 59 Properties of P2X receptors were not affected by P2X3 subunits knockout in S neurons 66 or, B-mATP depolarizes AH neurons from P2X?+ and P2X2'l' mice - 66 ATP- and a, B-mATP- induced depolarization in AH neurons from szf” and P2X;"' mice 75 Part B. Pharmacological identification of p2X subtypes in guinea pig myenteric neurons 75 Depolarization induced by ATP and a,B-mATP in AH and S neurons in guinea pigs 75 Effects of TNP-ATP on fEPSPs and ATP-induced depolarization in S neurons in guinea pigs 78 P2X receptors couple to a calcium-dependent potassium conductance in guinea pig myenteric neurons 85 An afterhyperpolarization (AHP) is induced by activation of P2X receptors 85 The ionic basis of the afterhyperpolarization 88 Immunochemical characterization of S neurons with after- hyperpolarization 93 Fast excitatory synaptic transmission during burst of neuronal activity—98 Fast synaptic transmission declined in amplitude during trains of electrical stimulation 98 fEPSCs exhibit paired pulse depression 101 Recovery from synaptic rundown 101 The contribution of nAChRs and P2X receptors to rundown of fEPSPs 104 Presynaptic mechanisms 109 Presynaptic 5-HT4 receptor-mediated regulation of fast excitatory synaptic transmission 125 5-HT4 receptor agonist increases fEPSPs during trains of stimulation - 125 5-HT4 receptor agonist accelerates recovery time of fEPSP after rundown 125 Forskolin mimics 5-HT4 receptor agonist effects 130 H-89 abolishes the effects of 5-HT4 receptor agonist I30 5-HT4 receptor agonists increase release probability from single varicosities 13 5 DISCUSSION 143 P2X receptor subtypes expressed in myenteric neurons 144 Properties of murine myenteric neurons 145 vi P2X receptor-mediated responses in murine AH neurons ------------ 147 Fast synaptic transmission and P2X receptor-mediated responses in murine S neurons 148 P2X subtypes in myenteric neurons in guinea pig small intestine ---151 P2X receptor-mediated hyperpolarization in guinea pig myenteric neurons 153 Ionic basis of afterhypolarization 153 Functional significance of AHP in myenteric S neurons ------------- 157 Rundown of fEPSPs during bursts of neuronal activity 159 Rundown of fEPSPs 159 Receptor desensitization did not contribute to synaptic rundown --- 160 Rundown is not caused by presynaptic inhibition 161 Changes in presynaptic action potentials do not contribute synaptic rundown 163 Depletion of transmitter accounts for synaptic rundown ------------- 165 Rundown of fEPSPs during bursts of neuronal activity 168 Activation of 5-HT4 receptors attenuates synaptic run-down and accelerates recovery from run-down I71 5-HT4 receptor-mediate presynaptic regulation of synaptic transmission is through adenylate cyclase-cAMP-protein kinase A signaling pathway 173 Activation of 5-HT; receptors increases probability of transmitter release from single varicosities 174 SUMMARY AND SIGNIFICANCE 178 BIBLIOGRAPHY l 83 vii LIST OF TABLES Table 1. Drugs and ligands used in this dissertation 48 Table 2. The electrical properties of myenteric neurons in mouse small intestine -------- 53 Table 3. Effects of 5-HT4 receptor agonists on release probability 142 viii Figure 01. Figure 02. Figure 03. Figure 04. Figure 05. Figure 06. Figure 07. Figure 08. Figure 09. Figure 10. Figure l I. Figure 12. LIST OF FIGURES The general organization of the enteric nervous system 5 Electrophysiological recordings from single myenteric neurons in LMMP preparations and neurons maintained in primary culture 10 Myenteric AH and S neurons 12 Intrinsic neural circuitry in the myenteric plexus l8 The action potentials evoked in myenteric S neurons from wild type, P2X; and P2X3 knockout mice 55 The action potentials evoked in myenteric AH neurons from wild type, P2X2, P2X; knockout mice 57 Effects of PPADS on fEPSPs from murine myenteric S neurons in wild type and P2X; knockout tissues 60 Effects of mecamylamine on fEPSPs from murine myenteric S neurons in wild type and P2X; knockout tissues 62 Responses of murine myenteric S neurons in wild type and P2X; knockout 64 tissues to nicotine, ATP and or, B-mATP Effects of PPADS on fEPSPs from murine myenteric S neurons in wild type and P2X3 knockout tissues 67 Effects of mecamylamine on fEPSPs from murine myenteric S neurons in wild type and P2X; knockout tissues 69 Responses of murine myenteric S neurons in wild type and P2X3 knockout 71 tissues to ATP and or, B-mATP ix Figure 13. Figure 14. Figure 15. Figure 16. Figure 17. Figure 18. Figure 19. Figure 20. Figure 21. Figure 22. Figure 23. Figure 24. Figure 25. Figure 26. Figure 27. Responses of murine myenteric AH neurons in wild type and P2X; knockout tissues to or, B-mATP 73 Responses of murine myenteric AH neurons in wild type and P2X3 knockout tissues to ATP and or, B-mATP 76 Responses of myenteric S and AH neurons in guinea pig small intestine to ATP and or, B-mATP 79 Effects of TNP-ATP on ATP-induced depolarization in guinea pig S neurons 81 Effects of TNP-ATP on fEPSP in guinea pig S neurons 83 or, B-mATP induces an after-hyperpolarization following depolarization in guinea pig S neurons 86 a, B—mATP-induced after-hyperpolarization is Ca2+ sensitive 89 or, B-mATP—induced after—hyperpolarization is mediated by an increase in 1C conductance 9 1 Identification of Kc, subtypes mediating after-hyperpolarization ------------- 94 Identification of neurons that generate afterhyperpolarization 96 Trains of electrical nerve stimulation induce rundown of fEPSPs ------------- 99 Paired stimulation caused synaptic depression 102 Time-dependent recovery from rundown 105 Both cholinergic and purinergic component of synaptic transmission rundown 107 Postsynaptic desensitization does not contribute to synaptic rundown ------ 1 10 Figure 28 Figure 29. Figure 30. Figure 31. Figure 32. Figure 33. Figure 34. Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Figure 41. . Antagonists against presynaptic inhibitory receptors do not alter synaptic rundown Pertursis toxin does not affect synaptic rundown Action potentials do not fail during the train stimulation at 10 Hz ----------- Effects of IBTx on action potentials in S and AH neurons The effect of IBTx on synaptic rundown Prucarlopride increases cholinergic and purinergic fEPSP Renzapride increases fEPSPs elicited by train of stimulation Renzapride reduces recovery time of fEPSP from rundown F orskolin mimics the effects of renzapride H-89 abolishes the effects of renzapride 113 115 119 121 123 126 128 131 133 I36 Activation of presynaptic 5-HT4 receptors increases release probability from single varicosities 139 A proposed model for P2X-mediated afterhyperpolarization in S neurons --154 A proposed model for 5-HT4 receptor-mediated presynaptic regulations --- 169 A summary of neural circuitry in myenteric plexus with emphasis on the new findings in the thesis 180 Note: Images in figures 01, 02, 03, 04, 22, 39, 40 and 41 are presented in color. A black and white copy may not represent the original information. xi INTRODUCTION THE ENTERIC NERVOUS SYSTEM Overview The enteric nervous system (ENS), sympathetic and parasympathetic nervous systems compose the autonomic nervous system (ANS). The ENS is intrinsic to the gastrointestinal (GI) tract. The ENS is the collection of neurons and supporting cells located within the wall of GI tract, including the neurons within the pancreas and gall bladder (Fumess and Costa, 1987). The ENS is composed of two major interconnected ganglionic plexuses: the myenteric (Auerbach’s) and submucosal (Meissner’s) plexuses (Fumess and Costa, 1987). Auerbach and Meissner are the two neuroanatomists who first provided detailed descriptions of the arrangement of the neurons and nerve fibers of the plexuses that bear their names. The descriptions of the two groups of neurons as the myenteric and submucosal plexuses are more meaningful in terms of their anatomical arrangement and this terminology is currently used widely. The ENS is highly autonomous as it can function independent of central nervous system (CNS) control. Although it is innervated by efferent nerve fibers originating in neurons in sympathetic, parasympathetic ganglia and sensory afferent fibers from neurons in the nodose and dorsal root ganglia, the ENS can control most GI functions after these extrinsic nerves have been cut. The ENS is often referred as the “little brain” or “second brain” to denote its functional autonomy (Gershon, 1999). The ENS can operate autonomously because it contains all of the components required for most GI reflexes (Bertrand, 2003; Fumess et al., 2004). The original data supporting this conclusion can be tracked back to 1899 when Bayliss and Starling observed a pressure- evoked descending propulsive wave consisting of oral contraction and anal relaxation (Bayliss and Starling, 1899). They removed sympathetic and parasympathetic innervations of the intestine and found no interruption of the pressure-evoked reflex. They proposed the “ Law of the Intestine” which stated that “local stimulation of the gut produces excitation above and inhibition below the excited spot. These effects are dependent on the activity of the local nervous mechanisms.” (Bayliss and Starling, 1899). In 1917 Trendelenburg conducted a further study and observed the same type of reflex in guinea pig gut in vitro, which Trendelenburg called peristalsis (F umess and Costa, 1987). These pioneering studies demonstrated that the peristaltic reflex could occur in vitro when the bowel is obviously isolated from the CNS and sensory ganglia. They provided indirect evidence that the ENS contains all the elements and complete intrinsic neural circuits that underlie the coordinated movement of GI smooth muscle. Modern experimental techniques have facilitated steady progress in identifying these components and the neural circuits they form (Fumess, 2000; Fumess et al., 2004; Bertrand 2003. Also see below “classification of myenteric neurons and neural circuits” for details). Briefly, the individual components include sensory neurons that transduce chemical or mechanical stimulation to initiate a reflex; intemeurons that relay the inputs and mediate the reflex and; excitatory and inhibitory motor neurons that cause coordinated contraction and relaxation of smooth muscle to accomplish peristalsis (Bertrand, 2003; Fumess et al., 2004). So, when J. N. Langley formally defined the ANS, he cited anatomical and functional evidence that led him classify the ENS as separate from the sympathetic and parasympathetic divisions. In his comprehensive review (Langley, 1921), he pointed out that: “...we should expect the cells of A uberbach ’s and Meissner ’s plexuses to be on the course of the bulbar and sacral nerves, but as there was no clear proof of their central connection, and as their obvious histological characters diflered from those of any other peripheral nerve cell, I placed them in a class by themselves as the enteric nervous D system... ’ Anatomy of the GI tract and arrangement of the enteric plexuses The GI tract is a muscular tube extending from the mouth to the anus. In general, the GI tract includes the esophagus, stomach, duodenum, jejunum, ileum and large intestine. The duodenum, jejunum and ileum compose the small intestine. The GI tract has a similar laminar structure in these different segments and consists of two smooth muscle layers, mucosa and the enteric plexuses (Schofield, 1968; Fumess and Costa, 1987). There is an inner circular and an outer longitudinal smooth muscle layer. The anatomic arrangement for all these layers is, in order from serosal to lumen side, longitudinal smooth muscle, myenteric plexus, circular smooth muscle, submucosal plexus and mucosal plexus (Figure 1). This kind of arrangement provides an anatomical basis for the function of the two enteric plexuses fimction. The myenteric plexus controls motility as it exists between two smooth muscle layers. The submucosal plexus regulates secretory and absorptive function and it lies just below the mucosal layer. Figure l. The general organization of the enteric nervous system A schematic diagram adapted from Fumess and Costa (1987) illustrates the organization of the enteric nervous system as seen in a whole mount of small intestine. The lumenal mucosa is at the bottom while the serosal surface is at the top. The submucosal plexus is located between the circular muscle and the mucosal layer. The myenteric plexus is located between the longitudinal and circular smooth muscle layers. In guinea pig small intestine, the myenteric plexus is attached tightly to the longitudinal muscle. Therefore, the longitudinal muscle-myenteric plexus (LMMP) preparation is used to study myenteric neurons. Synaptic connections in the myenteric plexus that are present in vivo are preserved in the LMMP preparations. Figure l Interganglionic Myenteric Ganglion Fiber Tract Submucosal Ganglion Longitudinal Muscle Morphology and ultrastructure of enteric neurons The ENS contains more than 100 million neurons, which are as many as the number of neurons in the spinal cord. Alex Dogiel (1899) provided a complete and accurate description of the morphology of enteric neurons. He used methylene blue as histological stain to reveal neuronal morphology. He demonstrated three groups of enteric neurons based on soma shape, soma size and the number and distribution of neuronal processes (Fumess and Costar, 1987). The morphological classification scheme of enteric neurons is named after him. Dogiel type I cells are flat, with stellate or angular shapes. The size of cell bodies was 13- 35 pm in length and 9-22 pm in width. They had one axon (unipolar) and 4-20 or more short dendrites. Because many of these cells sent single axons to the muscle, Dogiel concluded these neurons were motor neurons. Dogiel type 11 cells were described as angular, star or spindle shaped with multiple long neural processes. The cells had smooth surfaces with long axes of 22-47 um and short axes of 13-22 pm. These cells projected multiple processes to the mucosa and adjoining ganglia and were recognized as sensory neurons. Dogiel type III cells were similar to type 1 cells. They had a uniopolar shape with 2—10 filamentous dendrites. Type III neurons were suggested to be motor neurons. Later studies have suggested there was an overlap between type I and III neurons in morphology therefore these two cell types were re- classified as type I neurons. Ultrastructural studies of enteric ganglia reveal important differences from other autonomic ganglia. Enteric ganglia do not have internal capillaries and connective tissue (Cook and Burnstock, 1976) and the enteric neurons in ganglia are packed tightly without satellite cell sheaths (Llewellyn-Smith et al., 1981). Electron microscopy has identified eight types of neurons within myenteric ganglia in guinea pig small intestine based on the size and the intracellular unltrastructure (Cook and Bumstock, 1976). However, the submucosal ganglionic neurons do not differ from each other in their ultrastructural appearance as they all have broad and thin dendrites, axo-somatic and axo-dendritic synapses and similar perikarya of 12-30 pm (Wilson et al., 1981; Komuro et al., 1982). Ultrastructural studies have also revealed that there are multiple types of synaptic vesicles that can be differentiated based on size and vesicle content. These observations have led to the suggestion that enteric neurons use multiple transmitters and some nerve terminals contain and release more than one type of transmitter (Fumess and Costa, 1987; Llewellyn-Smith et al., 1989). Myenteric neurons are directly contacted by varicosities. However the pre-and postsynaptic specializations are not different from synapses in other autonomic ganglia or in the CNS (Llewellyn-Smith et al., 1989). Electrophysiology of myenteric neurons Electrophysiological studies of the ENS began with extracellular recordings from the myenteric plexus (Wood, 1970). Later, intracellular microelectrode recordings were made from the myenteric plexus (Hirst et a1 1974; Nishi and North 1973) and submucosal plexus (Hirst and Mckirdy 1975). Patch-clamp recording techniques have also been used to record from cultured myenteric neurons (Zhou and Galligan, 1996) and submucosal neurons (Barajas-lopez et al., 1998) and in acutely isolated LMMP preparations (Rugiero et al., 2002). Studies in LMMP preparations provide information that represents Physiological conditions in vivo while studies done in cultured neurons allow investigations of electrophysiological properties in more detail than is possible using the intact plexus preparation. Figure 2 contains schematic diagrams showing the setups for these two types of electrophysiological recordings. These studies have characterized the electrophysiological properties of myenteric neurons and the ion channel basis of these properties (Galligan et al., 1989; Galligan et al., 1990; Galligan, 20023 and b; Rugiero et al., 2002; Fumess and Sanger, 2002; Bertrand, 2003). Most intracellular electrophysiological studies of myenteric neurons have been done in preparations from the guinea pigs. The reason to choose the guinea pigs as animal model for GI studies is because guinea pig gut is readily dissected, reasonably long and mechanically quiescent when studied in vitro. There have also been studies done in mouse myenteric plexus (F urukawa et al., 1986). Gene manipulation in mice provides an opportunity to study the electrophysiological properties of specific proteins (Ren et al., 2003; Bian et al., 2003). Most neuronal properties are similar between guinea pigs and mice. The following description is mainly drawn from guinea pigs unless noted. Two distinct types of myenteric neuron have been identified based on their electrophysiological characteristics. S neurons and AH neurons were described by Hirst et al (1974); while Nishi and North (1973) found neurons with the same properties but called these type I and type II neurons (Figure 3). The nomenclature of Nishi and North is not longer used because of its similarity to the Dogiel classification of cell morphology. In addition, Hirst’s nomenclature was based on the properties of the neurons and this is more meaningful. Neurons responding to single focal stimulation of interganglionic nerve fibers with a fast excitatory postsynaptic potential (fEPSP) were termed S neurons; the fEPSP is the predominant mechanism of synaptic excitation in the GI tract. Action Figure 2. Electrophysiological recording from single myenteric neurons A. Intracellular electrophysiological recordings are made from single myenteric neurons in LMMP preparations. Myenteric ganglia can be visualized using a microscope with differential interference contrast optics. A sharp microelectrode filled with 2 M KCl and with a tip resistance of ~100 M ohm is used to impale single neurons. Synaptic activity can be evoked by electrical stimulation of interganglionic nerve bundles. Agonists can be applied locally by pressure ejection from a drug pipette positioned close to the impaled neuron. B. Patch-clamp recordings are made from single myenteric neurons maintained in primary culture. The picture shows the cultured myenteric neurons visualized using a microscope with differential interference contrast optics. Cultured myenteric neurons grow processes that form synapses with other myenteric neurons. Synaptic responses can be elicited by stimulating presynaptic single nerve fibers. Agonists can be applied locally by gravity-driven flow from a drug pipette positioned close to the target neuron. Figure 2 A STIMULATING ELECTRODE Drug pipette var—Tr —— I—, -- RECORDING ’ _, - r, , ELECTRODE Stimulating Electrode Recording Electrode Figure 3. Myenteric AH and S neurons A. Photomicrographs of an AH neuron morphology and its electrophysiological properties. AH neurons are bipolar with several dendrites and a smooth cell body. Action potentials in AH neurons have a long-lasting (1-10 5) hyperpolarization pointed by the arrow (lower left). Trains of stimulation (as arrow shows) evoke slow EPSPs from AH neurons (lower right). B. S neurons have one axon and multiple short dendrites. Single electrical stimuli (as arrow shows) elicit fEPSPs from S neurons (lower right). Action potentials in S neurons have a short afterhyperpolarization as the arrow shows (lower left). For morphological studies, microelectrodes filled with 4% Neurobiotin were used to record from these neurons. Neurobiotin was localized with a Cy3 conjugated avidin. 12 Figure 3 A All neuron / Afterhyperpolarization —2 s f 2 s St' I t. Action potential 1mu a Ion sEPSP B S neuron #w _ "T v "‘ '1 — 2 S / 10 ms Afierhyperpolarization Stimulation Action potential fEPSP potentials recorded from the soma of S neurons were mediated solely by voltage-gated NaJr channels and they can be blocked by tetrodotoxin (TTX). The action potential in S neurons is followed by an afterhyperpolarizaion of ~20 ms duration (Hirst et al 1974; Bomstein et al., 1994). Most S neurons fire action potentials continuously when depolarized by a current pulse injected through the recording microelectrode (Bomstein et al., 1994). Inward rectifier potassium current (Ikir) modulates the resting membrane potential of S neurons (Galligan et al., 1989; Rugiero et al 2002). As myenteric S neurons receive mainly excitatory synaptic inputs, the membrane potential is unlikely to go to values more negative than 13.. (~ -80 mV) where Kir channels operate more actively and potassium ions leave the cell. Under physiological conditions, IKir hyperpolarizes the neurons to maintain resting membrane potential. AH neurons are so named because the action potentials generated in soma are followed by long-lasting (1-10 5) afierhyperpolarization (AHP) (Figure 3) (Nishi and North, 1973; Hirst et al., 1974; Fumess et a1, 1998). AH neurons cannot fire a train of action potentials because the AHP limits the firing rate. The AHP is mediated by activation of an intermediate conductance calcium-dependent potassium channel (Imp) (Morita et al., 1982; Hirst et al., 1985; Galligan et al., 1989; Vogalis et al., 2002). The action potential of AH neurons has two components: 1) a TTX-sensitive sodium current and, 2) a calcium current. Calcium action potentials are mediated by N-type calcium channels as N-type calcium channel blocker m—conotoxin reduces the duration of action potentials and the amplitude of the afterhyperpolarization (F umess et al., 1998; Vogalis et al., 2001; Rugiero et al., 2002). In addition, immunoreactivity for the alB (N-type) l4 calcium channel subunits is localized to AH neurons, whereas only weak alA (P/Q-type calcium channel subunit) immunoreactivity was observed (Kirchgessner and Liu, 1999). R type calcium channels are only expressed in the nerve fibers (Naidoo, personal communication) therefore it does not seem to contribute to action potential but to contribute to regulation of synaptic transmission. AH neurons contain calbindin immunoreactivity, which is a neurochemical marker to differ AH neurons from S neurons. A cationic current activated by hyperpolarization (1..) appears to be one major conductance in AH neurons (Galligan et al., 1990; Rugiero et al., 2002). Activation of 1;, was suggested to increase excitability in AH neurons by keeping the membrane potential of AH neurons less hyperpolarized in response to activation of [AHP and Ikir. Some AH neurons receive fast synaptic inputs but slow synaptic transmission is the primary mechanism of synaptic excitation in AH neurons (Galligan et al., 2000; Fumess et al., 2004) Functional classification of myenteric neurons Classification of myenteric neurons has accomplished using morphological, electrophysiological or neurochemical properties (Fumess, 2000; Brookes 2001). A combined classification scheme that includes function would be more significant in a way that can facilitate understanding the functional significance of the myenteric neurons with different properties. This revised scheme requires correlating the electrophysiology, the neurochemistry and the morphology of enteric neurons with their functions. This will help to uncover the mechanisms underlying neural regulation of GI function. 15 There has been significant progress in defining the types of enteric neurons in the context of their roles in enteric neural circuitry (Bertrand, 2003; Fumess et al., 2004). Correlation of electrophysiology, morphology and function has been accomplished by injecting marker dyes into neurons that had been recorded from and subsequently examining cell shape and neurochemical content (Hodgkiss and Lees, 1983; Bomstein et al., 1984; Iyer et al., 1988; Messenger et al., 1994; Clerc et al., 1998; Lomax and Fumess, 2000; Tamura et al., 2001; Nurgali et al., 2003). Based on these kinds of studies it has been established that the myenteric plexus contains three major types of neurons. They are motor neurons, intemeurons and intrinsic primary afferent neurons (IPANs). IPANs are so termed is to differentiate them from the extrinsic sensory neurons whose cell bodies are in dorsal root and nodose ganglia, which have afferent connections with the ENS (Fumess et al., 2004). Motor and intemeurons have Dogiel type I morphology and IPANs have Dogiel type II morphology (Figure 3). Five subtypes of motomeurons, one subtype of ascending and four subtypes of descending intemeurons have been identified (Fumess, 2000; Brookes, 2001; Fumess and Sanger, 2002). There are also a small number of intestinofugal neurons (<1% in number) that send afferent fibers to prevertebral ganglion and these connections are part of intestinofugal reflex pathways (Fumess, 2002; 2003). Each type of neuron has neurochemical coding and electrophysiological characteristics (Fumess, 2000; Fumess and Sanger, 2002; Bertrand, 2003). For example, inhibitory longitudinal motor neurons express nitric oxide synthase (NOS), vasoactive intestinal peptide (VIP) and GABA while excitatory longitudinal motor neurons express choline acetyltransferase, calretinin and substance P. IPANs are 16 the only neuron type that expresses calbindin. Electrophysiologically, motor- and intemeurons are S neurons while IPANs are AH neurons. Intrinsic neural circuits in the myenteric plexus Studies of neuron types have provided the information that helps to understand ' how myenteric neurons are organized to form a functional network that underlies the peristaltic reflex (Figure 4). Because the peristaltic reflex is based on distention-evoked contraction oral to the point of distention and relaxation anal to the distention point, the organization of the intrinsic neural circuitry must be polarized accordingly. The accumulated data suggest that IPANs are excited indirectly by intraluminal chemical or mechanical stimulation through sensory mediators released from enterochromaffin (EC) cells in mucosa. The identified mediators released from EC cells are 5-HT and ATP (Bertrand, 2003). 5-HT acts at 5-HT3 receptors and ATP acts at P2X receptors on nerve terminals to activate myenteric IPANs antidromically (Fumess et al., 2004). S-HTlp localized to terminals of submucosal IPANs may also mediate sensory information (Gershon, 2004). IPANs then activate intemeurons via cholinergic and purinergic synaptic transmission in the descending pathway and cholinergic synaptic transmission in the ascending pathway (Galligan, 2002b). Interneurons then activate motomeurons, which excite smooth muscle orally and inhibit smooth muscle anally to generate the peristaltic reflex. It is very clear that fast synaptic transmission between these neurons plays an important role in process sensory information (see “fast synaptic transmission” for details). However, the mechanisms by which fast synaptic transmission may encode l7 Figure 4. Intrinsic neural circuitry in myenteric plexus This diagram a propose model of the enteric neural circuit underlying the peristaltic reflex. Intrinsic primary afferent neurons (IPANs) are AH neurons while inter- and motomeurons are S neurons. Intraluminal mechanical or chemical stimulation of the mucosa releases 5-HT from enterochromaffin cells. S-HT acting at 5-HT3 receptors localized to nerve terminals of myenteric IPANs excites the IPAN antidromically. Activation of IPANs releases transmitters to activate intemeurons in orally and anally directed synaptic pathways. Subsequent activation of intemeurons stimulates inhibitory or excitatory motor neurons to relax or contract smooth muscle, respectively. IPAN: intrinsic primary afferent neuron; LM: longitudinal muscle; MP: myenteric plexus; SM: smooth muscle; SMP: submucosal plexus; 18 Figure 4 sensory motomeuron intemeuron neuron intemeuron motomeuron has: _L MP éACh ”AC CM SMP Mucosa Enterochromaffin cell Mechanical or chemical stimuli and regulate sensory information causing responses to a variety of physiological stimulation are still unknown. Extrinsic innervation Myenteric and submucosal plexuses also receive extrinsic inputs through parasympathetic and sympathetic nerves. The sympathetic preganglionic cell bodies are in thoracic (T) and lumbar (L) segments of the spinal cord. Their axons end on the postganglionic neurons in celiac ganglia, superior and inferior mesenteric ganglia. Sympathetic postganglionic noradrenergic fibers innervate the myenteric plexus, the submucous plexus, and the mucosa, but have little input to the muscle layers (F urness and Costa, 1987; King and Szurszewski, 1989; Llewellyn-smith et al., 1993). The majority of postganglionic sympathetic neurons innervating the ENS are found in celiac-mesenteric ganglia. They project via splanchnic nerves to the stomach and the small intestine. Some sympathetic innervation originates from cervical chain ganglia and projects to the upper small intestine. The parasympathetic innervation of the GI tract originates in the cervical and sacral portions of the spinal cord and terminates in the esophagus, stomach, pancreas, sigmoid colon, rectum and anus (Kirchgessner and Gershon, 1989). Only few vagal efferent fibers innervate the small intestine. Sympathetic and parasympathetic innervation regulates GI function. Sympathetic postganglionic neurons release norepinephrine (NE) to inhibit enteric excitatory cholinergic neurons through presynaptic mechanism (Hirst and McKirdy, 1974) by acting at a2-adrenergic receptors (Frigo et al., 1984). Activation 20 of sympathetic nervous system results in decreased motility of GI smooth muscle. Parasympathetic nervous system provides only excitatory innervation to GI tract. Stimulation of the parasympathetic nervous systems results in the release of ACh, which causes an increase in GI activity. There is also afferent innervation that conveys sensory information from the GI tract to the CNS where GI reflex is coordinated and integrated with conscious behavioral responses. The extrinsic sensory innervation originates from enteric plexus and follows two routes to the CNS. The cell bodies of vagal afferent neurons are located in nodose ganglia while the cell bodies of spinal afferent neurons are located in dorsal root ganglia. Pseudo-unipolar sensory neurons in nodose ganglia send their central processes to the nucleus tractus solitarius (NTS) of the medulla in brain stem and their peripheral processes to the GI tract via vagal nerve (Berthoud et al., 2004). Vagal afferent is considered to convey the sensory information from the GI tract that generates the consciousness, for example satiety (Grundy and Scratcherd, 1997). Spinal afferents enters spinal cord to make synapses with the second order neurons that convey GI nociceptive information to the CNS. There is a differential innervation of these afferents throughout the GI tract. Vagal afferents end largely in the upper GI tract including the esophagus and stomach (Powley and Phillips, 2002). For spinal afferents, pelvic afferents are limited in the lower bowel. In contrast, splanchnic afferents innervate the whole GI tract. Submucosal neurons The submucosal plexus is a second important component of the ENS. It underlies the regulation of intestinal secretory reflexes (Cooke, 1999). Submucosal neurons receive 21 their intrinsic innervation mainly from myenteric neurons and partly from other submucosal neurons (Bomstein et al., 1987). They also receive major extrinsic innvervation from noradrenergic neurons in the prevertebal sympathetic ganglia (Mihara et al., 1997). In submucosal plexus, four distinct types of submucosal neurons have been identified based on neurochemical properties (Mihara et al., 1997; Fumess, 2000; Brookers, 2001). Some submucosal neurons have Dogiel type II morphology. The other submucosal neurons have filamentous morphology (Brookes, 2001). Non-cholinergic secretomotor/vasodilator neurons are immunoreactive for dynophin (DYN), galanin (GAL) and VIP. These neurons account for 45% of submucosal neurons and they use VIP as their secretomotor neurotransmitter. Cholinergic secretmotor neurons account for 30% and they are immunoreactive for ChAT, calcitonin gene-related peptide (CGRP) and neuropeptide Y (NPY). Submucosal IPANs (10%) that have immunoreactivity of ChAT, DYN and substance P (SP) and they use Ach as transmitter. Cholinergic intemeurons comprise 15% of submucosal neurons and they are immunoreactive for ChAT and DYN. Electrophysiologically, submucosal neurons can generate three types of synaptic potentials in response to focal nerve stimulation (Mihara, 1993). Fast EPSPs can be recorded from motor and intemeurons afier single electrical stimuli. Fast EPSPs are mediated by cholinergic inputs from neurons in both myenteric and submucosal plexuses. Slow EPSPs are evoked by trains of stimulation and slow EPSPs are likely mediated by SP, VIP and 5-HT (Mihara et al., 1997). Slow IPSPs can be recoded from non- cholinergic secretmotor neurons: slow IPSPs are mediated by norepinephrine released from extrinsic sympathetic nerves. Norepinephrine acts at a2-adrinergic receptors (Surprenant and North, 1987). 22 SYNAPTIC TRANSMISSION Overview Neurons communicate with each other through synaptic transmission, which occurs at a specialized neuronal structure, the synapse, a milestone-like concept coined by C. Shenington. Enteric neurons use chemical synaptic transmission to mediate neural signals. Synaptic transmission at chemical synapses is complex in structure, fiinction and modulation mechanisms (Eccles, 1982; Walmsley et al., 1998). Neurotransmitters released from presynaptic nerve terminals act at receptors located at postsynaptic neurons (Ruff, 2003). Transmitter release occurs from nerve terminals after an influx of calcium produced by an action potential, triggers the excytosis of neurotransmitter vesicles clustered at the active zone (Smith and Augustine, 1988; Zucker, 1993; Martin, 2002). Some transmitters are recovered by uptake by specific transporters and some will be degraded in the synaptic cleft by enzymes (McMahan et al., 1978; Amara and Arriza, 1993). Most chemical synaptic transmission is unidirectional. However, the retrograde synaptic transmission may occur in some tissues (Jessel and Kandel, 1993; Levenes et al., 2001; Diana et al., 2002). Synaptic modulation can occur by targeting presynaptic nerve terminals or/and postsynaptic neurons. There are two broad mechanisms of chemical synaptic transmission based on synaptic delay between a presynaptic action potential and the onset of the postsynaptic change in membrane potential caused by the release neurotransmitter: slow and fast synaptic transmission. 23 Slow synaptic transmission in myenteric plexus Slow synaptic transmission occurs through activation of G-protein coupled receptors (GPCRs) by neurotransmitters. GPCR activation alters the membrane potential of postsynaptic neurons through activation or inhibition of a number of second messengers leading to close or open of ion channels. GPCRS are a family of membrane proteins with seven transmembrane domains (Ji et al., 1998). The third intracellular loop is the G-protein binding site (Wess, 1997). G-proteins are trimers composed of a, B and 7 subunits (Hamm, 1998). In general, two mechanisms mediate extracellular signals by activation of GPCRs. The first one is a signaling pathway through activated G protein a subunit resulting in an increase or decrease in the activity of a target enzyme (typically a protein kinase) via a second messenger (Gilman, 1987). The protein kinase then modulates the target effectors such as ion channels and gene expression to alter neuronal properties (Hille, 1994; Hamm, 1998). In addition, the G protein [3 and 7 subunit complex can directly act at the ion channels as membrane de-limited factors to modulate ion channel kinetics (Wickman et al., 1994; Huang et al., 1995; Clapham and Neer, 1997). As there are multiple-step processes following activation of GPCRs including agonist binding, the subsequent activation of signaling pathways and the activation of target ion channels, GPCRs mediate slow synaptic transmission with an onset of 50-100 ms and effects on neuronal excitability lasting seconds to minutes. Slow synaptic transmission is an important signaling mechanism in the ENS (Surprenant and North, 1988; Bertrand and Galligan, 1994; Bertrand and Galligan, 1995; Pan et al, 1997; Alex G et al., 2001; Galligan, 2002 a, b). In the ENS, examples of GPCRs include NK—3 receptors activated by substance P (SP) (Holzer and Holzer-Petsche, 2001), 5-HT.p receptors activated by 5- 24 HT (Pan et al., 1997) and P2Y receptors (Bomstein et al., 2002) activated by ATP. Slow synaptic responses in the ENS are due to a concurrent decrease in potassium conductance and increase in chloride conductance (Bertrand and Galligan, 1994). Slow synaptic transmission occurs predominately in AH neurons (Galligan, 20023). Fast synaptic transmission and ligand-gated ion channels in myentric neurons Ligand-gated ion channel-mediated fast synaptic transmission is the second broad class of synaptic transmission in the ENS (Galligan, 2002a). Ligand-gated ion channels are membrane proteins composed of multiple subunits. The ligand binding sites and the ion channels are localized within the same protein complex (Lester, 1992). There is no second messenger needed to convey the primary signals via ligand-gated channels (Uings and Farrow, 2000; Breitinger, 2001). Therefore, the synaptic delay is less than 1 ms and the duration of transmitter release and response lasts less than 100 ms (Panas et al., 2000). Several ligand-gated ion channels mediate fEPSPs in myenteric neurons. The major receptors include nicotinic cholinergic receptors (nAChRs), purinergic P2X receptors and 5-HT3 receptors. These receptors have been localized to S neurons and they receive synaptic inputs from AH neurons and other S neurons to mediate reflex activation (Galligan, 2002a). Nicotinic acetylcholine receptors (M C hRs) Acetylcholine (ACh), acting at nAChRs, is the predominant contributor to fast synaptic excitatory neurotransmission in the ENS (Nishi and North, 1973; Hirst et al., 1974; Galligan and Bertrand, 1994; Zhou et al, 2002). In myenteric plexus, application of 25 nAChR antagonists inhibits fEPSPs, at least partly, in 92% of S neurons. In submucosal S neurons fEPSPs are exclusively mediated by ACh acting at nAChRs (Hirst and Mckirdy, 1975; Suprenant, 1984). In the myenteric plexus, 25% of S neurons generate fEPSPs that are completely blocked by nAChR antagonists (Galligan et al., 2000); fEPSPs in the remaining neurons have non-cholinergic components (Galligan and Bertrand, 1994; Galligan et al., 2000). Cholinergic fEPSPs play a primary role in ascending reflex pathway in guinea pigs (Johnson et al., 1996; Spencer et al., 2000). Cholinergic fast synaptic transmission also mediates signals in descending reflex pathways in the rat colon and guinea pig ileum (Johnson et al., 1996; Bian et al., 2003). Neuronal nAChRs are heteromeric pentamers that are composed of different combinations of a and B subunits; each subunit four transmembrane domains. The electrophysiological and pharmacological properties of nAChRs are determined by the specific subunit composition (Albuquerque et al., 1996; Jones et al., 1999; Skok, 2002). Fast EPSPs and ACh-induced responses in myenteric neurons have a reversal potential of 0 mV (Galligan and Bertrand, 1994), which suggests that nAChRs are permeable to cations. Data from immunohistochemical and pharmacological studies done in myenteric neurons maintained in primary culture suggest that S neurons do not express ()7 subunit- containing nAChRs in the cell body but most S neurons express nAChRs that contain (1301584 subunits (Kirchgessner and Liu, 1998; Zhou et al., 2002). Cholinergic fEPSPs can also be recorded from AH neurons (Fumess et al., 1998; Galligan et al., 2000). But the amplitude of fEPSP in AH neurons is smaller than that in S neurons as there may be a lower density of nAChRs expressed in AH neurons (Barajas-Lopez et al., 2001; Galligan, 2002a). Patch-clamp studies performed in cultured myenteric neurons indicate that 26 nAChR-mediated inward currents have a reversal potential of 0 mV and the current- voltage relationship is inward rectified. These results confirm that nAChRs are non- selective cation channels and also suggest that nAChRs operate more actively in depolarizing the neurons to increase excitability. nAChRs desensitize fast within 0.3 second by 80% in response to steady-state application of ACh. Single channel currents have linear relationship with the potentials, which suggests that inward rectification of whole cell currents is due to a decrease in nAChR open probability at positive membrane potential (Zhou et al., 2002; Brown et al, 2003). In addition to nAChRs localized to somatodendritic region of S neurons, evidence indicates that nAChRs may also be localized to nerve terminals. These nAChRs mediate release of tachykinin peptides in myenteric ganglia and at the motor neuron- circular/longitudinal muscle junction (Galligan, 1999; Schneider and Galligan, 2000; Schneider et al., 2000). The mechanisms and the physiological functions of presynaptic nAChRs remain firrther studied. Purinergic P2X receptors The non-cholinergic component of fEPSPs recorded from 67% of myenteric S neurons is sensitive to the P2X receptor antagonists PPADS or suramin (Galligan and Bertrand, 1994; Galligan 2002a, b). This indicates that ATP acting at P2X receptors is a co-mediator with ACh acting at nAChRs in mediating fast synaptic excitation in myenteric neurons. But the purinergic fEPSPs are not evenly distributed in the whole gastrointestinal tract. Recordings from different segments of the gut indicate that purinergic fEPSPs predominate in ileum compared with that in duodenum, jejunum, 27 proximal and distal colon (LePard et al., 1997). There is also a preferred oral-anal polarity of purinergic projections in myenteric plexus (LePard and Galligan, 1999). In guinea pig ileum, by stroking the mucosal villi and recording synaptic responses, it has been concluded that P2X receptors mediate fast synaptic transmission between intemeurons in descending excitatory reflex pathway (Spencer et al., 2000; Monro et al., 2002). P2X receptors are also shown to mediate synaptic transmission from intemeurons to inhibitory motor neurons in guinea pig ileum (Bian et al., 2000). P2X receptors may also mediate activation in AH neurons and contribute to initiation of peristalsis (Bertrand and Bomstein, 2002). In this study, ATP applied to the cell body evoked a large depolarization in most AH neurons. The depolarization was blocked by PPADS. Application of ATP to the mucosal terminals of AH neurons can trigger action potentials that propagate back to the soma. The action potentials were also abolished by PPADS. Although the pharmacology of P2X receptors mediating these responses need to be further studied, P2X receptors obviously play physiological function in AH neurons mediating sensory information. However, the localization of ATP sources remains unclear. Study from vagal neurons has suggested that ATP may be co-transmitter with histamine in mast cells (Kreis et al., 1998). ATP is also localized in the vesicles in EC cells (Tamir and Gershon, 1990). In vivo, ATP released from these sources may serve as stimulator to P2X receptors. P2X receptors are ligand-gated non-selective cation channels. But the molecular structure is different from nAChRs as P2X receptors may be trimers consisting of combinations of one or more of seven different subunits (Khakh et al., 2001; North, 2002; Galligan, 2002a, b). Each subunit has two membrane-spanning domains (Khakh et 28 al., 2001). Studies of heterogeneous expression of P2X receptor subunits have generated a series of data for the properties of different P2X subunit combinations (North, 2002). P2X receptors can be composed of the same or different subunits to form homomers or heteromers, respectively. All P2X subunits can form homomers except the P2X6, which only combines with other subunits, whereas all P2X subunits can form heteromers except the P2X7, which can only form homomers. The specific composition of the P2X receptor determines the unique pharmacological and physiological properties. Pharmacological data have indicated there are or,B-mATP-sensitive P2X subtypes (P2X1 or P2X3) as well as P2X; subtype in myenteric S neurons of guinea pig intestine (Zhou and Galligan, 1996; LePard et al., 1997). Immunohistochemical studies have shown that P2X2 and P2X3 subunits are expressed in myenteric neurons (Castelucci et al., 2002; Poole et al., 2002). However, lack of selective drugs has impaired a full pharmacological identification of P2X receptor subtypes in myenteric neurons. Electrophysiological characterization of P2X receptors made in cultured myenteric neurons has shown similarities and differences to nAChRs. ATP-induced whole-cell currents show inward rectification, which is due to a decrease in the open probability of single channels at more positive membrane potentials as single channel current-voltage relationship is linear (Zhou and Galligan, 1996). The reversal potential of 0 mV indicates that P2X receptors are permeable to cations. A major difference between myenteric P2X receptors and nAChRs is that ATP induces inward currents that desensitize by 80% in 7 seconds and ACh currents desensitize by 80% in less than 1 3 (Zhou and Galligan, 1996; Zhou et al., 2002). This difference may contribute to their distinct fimctions in mediating synaptic signals in enteric neurons. 29 5-HT3 receptors 5-HT3 receptors are ligand-gated cation channels (Derkach et al., 1989; Fletcher and Barnes, 1998). These receptors belong to a same superfamily with nAChRs in overall structure. 5-HT3 receptors are pentamers with each subunit having four transmembrane domains. In the GI tract, 5-HT3 receptors are found in myenteric sensory nerve endings localized to the mucosal layer where they initiate the motor reflex by responding to 5-HT released from EC cells when mucosal villi are stimulated (Foxx-Orenstein et al., 1996; Bertrand et al., 2000; Galligan, 2002 a, b). In addition to triggering the sensory signal, 5- HT3 receptors also mediate fast synaptic transmission in some S neurons (Zhou and Galligan, 1999). These 5-HT3 receptors are localized to the cell body of myenteric S and AH neurons in small intestine where they mediate rapidly developing and desensitizing depolarization (Zhou and Galligan, 1999; Zhai et al., 1999). Fast EPSPs from about 11% S neurons have components that are not cholinergic or purinergic (Galligan et al., 2000). These fEPSPs are inhibited by ondansetron, a 5-HT3 receptor antagonist. S-HT3 receptor- mediated fEPSPs are involved in both descending and ascending GI reflex pathway (Monro et al., 2002; Yuan et al., 1994). The electrophysiological and pharmacological properties of 5-HT3 receptors have been studied in myenteric neurons maintained in primary culture (Zhou and Galligan, 1999; Zhai et al., 1999). More than 80% neurons generate inward currents in response to 5-HT. These currents desensitize with a double exponential time course of 1.1 and 6.9 s. 5-HT3 receptor-mediated currents do not rectify at positive membrane potentials as the nAChRs and P2X receptors do. The current-voltage relationship for 5-HT-induced single 30 channel currents is linear, which suggests that change of potentials does not alter single channel open probability. Dynamics of neurotransmitter release A number of chemicals including ATP, ACh, biogenic amines and amino acids have been identified as neurotransmitters in the ENS. A neurotransmitter has to meet the following criteria: 1) it is synthesized in the neuron; 2) it is present in the presynaptic terminal and exerts a defined action on the effectors; 3) it mimics the synaptic response when applied exogenously and 4) there is a mechanism to stop its action (Kandel et al., 2000). In the ENS, identified neurotransmitters mediating fEPSP are ACh, ATP, 5-HT, glutamate (Galligan, 2002a, b). Neurotransmitter release is triggered by calcium influx produced by an action potential leading to exocytosis of neurotransmitter-filled vesicles in quanta units. The exocytosis of transmitter vesicles is via the formation of a complex involving many specific proteins (Hanson et al., 1997). The released vesicles can be recycled by endocytosis, which is one mechanism for generating new synaptic vesicles (Schweizer et al., 1995; Murthy and Stevens, 1998; Sudhof, 2004). Although many vesicles are contained in a synaptic terminal, the vesicles are organized into firnctionally different pools (Rizzoli and Betz, 2005); 1) the readily releasable pool, 2) the recycling pool and 3) the reserve pool. A single action potential triggers release of a small percentage (~5%) of vesicles to release (Tong and Jahr, 1994; Stevens and Wang, 1995; Zucker and Regehr, 2002). These vesicles are from the readily releasable pool (RRP) (Stevens and Tsujimoto, 1995; Stevens and Wesseling, 1999; Zucker and Regehr, 2002). Ultrastructural studies have shown that, the vesicles in the 31 RRP are clocked at active zone sites where calcium channels are localized (Schikorski and Stevens, 1997). The RP can be depleted rapidly by a short train of electrical stimulation at high frequency (Elmqvist and Quastel, 1965, Richards et al., 2003). Most transmitter vesicles (typically 80 - 90%) are contained in the reserve pool, which cannot be released until they move into the active zone to refill RRP, a process called replenishment. Release from the reserve pool occurs only in response to intense stimulation. The recycling pool contains vesicles that are dynamically released and reused under different conditions of neuronal activity (Murthy and Stevens, 1999; Rizzoli and Betz, 2005). Vesicles in the reserve pool can also be released and recycled during periods of prolonged and intense synaptic activity (Murthy and Stevens, 1999). Styryl dyes such as FM1-43 are used to monitor synaptic vesicle cycling and significant progress has been made to elucidate these processes (Cochilla et al., 1999; Sudhof, 2000). Depletion of transmitter vesicles in the RRP is replenished slowly, which is an explanation for rundown of synaptic transmission (Sudhof, 2000; Zucker and Regehr, 2002). Replenishment of the RRP is also an explanation for synaptic recovery, which is a calcium-dependent process (Dittrnan and Regehr, 1998; Wang and Kaczmarek, 1998). Synaptic vesicle trafficking between these pools is dependent on a number of proteins (Chi et al., 2003). One of most important proteins is synapsin. Synapsin docks vesicles to the cytoskeleton and is a target protein for phosphorylation by protein kinases (Sudhof et al., 1989; Chi et al., 2003). Synapsin phosphorylation dissociates vesicles from the cytoskeleton and mobilizes them for trafficking from the reserve pool to the RRP (Turner et al., 1999). Activation of 5-HT4 receptors can stimulate protein kinase A (PKA) in the ENS (see below). Therefore it is a reasonable hypothesis that such receptors can mediate 32 the regulation of neurotransmitter trafficking to affect synaptic activity. In the ENS, the dynamic processes associated with synaptic transmission during trains of stimulation have not been studied. Recycling of synaptic vesicles Action potentials induce neurotransmitter release by initiating the processes that lead to fusion of synaptic vesicles with the presynaptic membrane and exocytosis of neurotransmitter. Afier exocytosis, synaptic vesicles undergo a recycling process including endocytosis and refilling with neurotransmitters so that they are prepared for another cycle of transmitter release in response to the next action potential. This process is an important mechanism that allows synaptic terminals to maintain an active supply of transmitter-containing vesicles. The pioneering studies of synaptic vesicle recycling were conducted at the frog neuromuscular junction (NMJ) using electrophysiology and electron microscopy (Heuser and Reese, 1973; Ceccarelli et al., 1973). In these studies, NMJ preparations were stimulated in presence of horseradish peroxidase (HRP) or dextran. Exocytosis Was assessed by recording the responses of the muscle cells to ACh released from nerve terminals. Endocytosis was measured as the amount of HRP or dextran containing in synaptic vesicles after tissue fixation. The membrane-attached objects with the markers were considered as the endocytosed vesicles. Based on these studies, two models were proposed for synaptic vesicle recycling. Using stimulation at high fi'equency, Heuser and Reese (1973) suggested “during stimulation the intracellular compartments of this synapse change shape and take up extracellular protein in a manner which indicates that 33 synaptic vesicle membrane added to the surface during exocytosis is retrieved by coated vesicles and recycled into new synaptic vesicles by way of intermediate cisternae This model became known as the “slow endocytotic pathway”. Ceccarelli and colleagues, however, proposed a different pathway for synaptic vesicle recycling. They used stimulation at low frequency and found (a), that synaptic vesicles fuse with, and re-form from, the membrane of the nerve terminal during and after stimulation and (b), that the re-formed vesicles can store and release transmitter (Ceccarelli et al., 1973). This suggestion later leads to a “fast endocytotic pathway” model. Later, extensive studies have been conducted to probe the mechanisms of synaptic vesicle endocytosis. Processes of endocytosis can be directly monitored by measuring capacitance change (Sun et al., 2002; Sun and Wu, 2001) or by optical recording (Gandhi and Stevens, 2003). Recycling can be assessed by monitoring the uptake of fluorescent tracers (Wu and Betz, 1998; Cochilla et al., 1999; Murthy and Stevens, 1999). Based on these studies, three endocytotic pathways have been proposed (Sudhof, 2004): 1) “kiss and stay”, 2) “kiss and run”, and 3) clathrin-mediated recycling. The “kiss and stay” model proposes that synaptic vesicles recycle directly afier closure of the vesicle membrane-plasma membrane fusion pore formed during exocytosis. In this model, vesicles are refilled with neurotransmitters while the vesicles remain docked to the active zone. Vesicle recycling and refilling do not require the formation of a clathrin coat around the vesicle. The “kiss and run” model also proposes that vesicles endocytose without a clathrin-coated intermediate. The difference from the “kiss and stay” model is that the vesicles do not stay in the active zone (Ceccarelli et al., 1973; Richards et al., 34 2000). After endocytosis, recycled vesicles mix into the reserve pool. "Kiss and stay” and “kiss and run” pathways are mechanisms for fast recycling of synaptic vesicles (Pyle et al., 2000; Sun et al., 2002; Rizzoli et al., 2003). The third pathway, clathrin-mediated endocytosis, is a slow recycling process. Vesicle recycling in this model requires formation of a clathrin-scaffold that organizes several adaptor and accessory proteins (Royle and Lagnado, 2003). The electron-dense coats around vesicles are formed by clathrin (Heuser and Reese, 1973; Cremona and De Camilli, 1997). Clathrin-mediated endocytosis and recycling may occur directly or via endosomes. Both slow and fast endocytosises require GTP hydrolysis (Jockusch et al., 2005; Yamashita et al., 2005). Rapid recycling including “kiss and stay” and “kiss and run” is a primary pathway for synaptic vesicle retrieval during low frequencies of stimulation (Jockusch et al., 2005). The slow recycling pathway is triggered during high frequency stimulation (Sun et al., 2002; Jockusch et al., 2005). In studies done on synapses in the calyx of Held, capacitance measurements revealed that endocytosis is completed with time constant of 56 ms after a single vesicle exocytotic event. This value increases to 115 ms during stimulation at frequencies < 2 Hz and the time constant increases to 2.3 s and 8.3 s after 10 stimuli at 20 or 333 Hz, respectively (Sun et al., 2002). In retinal bipolar cells, using capacitance measurements of membrane retrieval indicated that clathrin-dependent endocytosis has a slow phase of membrane retrieval with time constant of 10 3 while clathrin-independent pathway only takes 1 s (Jockusch et al., 2005). Although the data suggest that the clathrin-dependent slow recycling pathway is activated selectively upon high—frequency stiumulation and clathrin-independent fast pathway is stimulated by low-frequency stimulation, the mechanisms responsible for 35 these differences remain unclear. The proteins involved in these two recycling pathways have not been fully characterized. S-HT4 receptors in myenteric neurons 5-HT4 receptors belong to the superfamily of GPCRs (Barnes and Sharp, 1999). 5-HT4 receptors couple to the Gs subtype of G-protein activating adenylate cyclase increasing cAMP leading to activation of PKA. Several 5-HT4 receptor subtypes have been identified as products of alternative splicing in different species (Medhurst et al., 2001). All splice variants of the 5-HT4 receptors possess similar N-terrninal sequences up to residue 359, but differ in the length of the C-terminal segment. This segment is an important determinant of constitutive activity and receptor regulation (Blondel et al., 1998). One of the differences is the desensitization rate in which the shorter variants (h5- HT4a and r5-HT4s) desensitize slowly whereas the longer variants (h5-HT4b and r5-HT4L) desensitize more rapidly (Claeysen et al., 1998). The small intestine expresses predominately 5—HT4b receptors (Blondel et al., 1998). Radioligand binding and functional studies in the intestine suggest that 5-HT4 receptors are present on smooth muscle cells, excitatory cholinergic motor neurons, and terminals of IPANs. More specifically, 5-HT4 receptors are located on the terminals of calcitonin gene-related peptide (CGRP)-containing IPANs. Electrophysiological study indicated 5-HT4 receptor agonists did not alter resting membrane potential of myenteric neurons or and slow synaptic transmission when intracellular recordings were made in cell body (Pan and Galligan, 1995). This result suggested that 5-HT4 receptors were not localized to soma. 36 Electron microscopic evidence also indicates that 5-HT4 receptors are localized to synaptic terminals (Liu et al., 2005). 5-HT acting at 5-HT4 receptors triggers initiation of the peristaltic reflex (Craig et al., 1990; Craig and Clarke, 1991; Foxx-Orenstein et al., 1996; Grider et al., 1996; Grider et al., 1998; Jin et al., 1999). 5-HT4 receptors modulate fast synaptic transmission (Kibinger and wolf, 1992; Pan and Galligan, 1994), which underlies 5-HT4 receptor agonist-induced increases in GI motility (Grider et al., 1998). Further studies show that 5- HT4 receptors couple to facilitation of fast synaptic transmission by enhancing both ACh and ATP release (Pan and Galligan. 1994; LePard et al., 2004). It has been also shown that forskolin mimics 5-HT4 receptor-mediated facilitation of fast synaptic transmission and PKA selective inhibitors blocked 5-HT4-mediated synaptic facilitation (Galligan et al., 2003). This 5-HT4 receptor-mediated facilitation is used as the neurophysiological basis for the prokinetic actions of clinical drugs such as 5-HT4 receptor agonists, cisapride and tegaserod, which are used to treat G1 motility disorders. However, the underlying molecular and cellular mechanisms remain unknown. 37 RESEARCH GOAL The goals of the present study were to investigate fast synaptic transmission during trains of nerve stimulation and to identify mechanisms of synaptic modulation and plasticity. Purinergic P2X receptors in postsynaptic neurons mediate fEPSPs and sensory transduction in the ENS. The subunit compositions that determine P2X receptor properties affect synaptic transmission and sensory transduction. Identification of the P2X receptor subtype expressed by enteric neurons will help to understand the functional significance of fast excitatory synaptic transmission. Previous studies on fast synaptic transmission have investigated fEPSP induced by a single stimulation. However, fEPSP triggered by physiological stimulation occurs in bursts, which represent the situation in vivo. Therefore it is important to investigate fEPSPs elicited by trains of action potentials to understand the physiological functions of fEPSPs. Presynaptic 5-hydroxytryptamine (5-HT4) receptors mediate facilitation of fast synaptic transmission in myenteric neurons. This dynamic change mediated by 5-HT4 receptors is very important as agonists of 5-HT4 receptors are used as prokinetic drugs to treat certain GI disorders such as irritable bowel syndrome (138) and gastroesophageal reflux disease (GERD). Identification of the mechanism underlying 5-HT4 receptor- mediated facilitation will be useful in understanding the pathology and drug therapy of these GI motility disorders. 38 Fast synaptic transmission is an important mechanism for enteric neurons to perform their physiological functions. Alteration of fast synaptic transmission by abnormal physiological condition can develop into GI disorders. GI disorders are common problems throughout the world. Although they are rarely life threatening, they negatively impact quality of life (Fumess and Sanger, 2002). Therefore study of dynamic change of fast synaptic transmission will, in one way help to understand the neural circuits underlying the ENS regulation of GI physiological functions; in the other way provide information for GI pathology and GI disorder drug therapy. Specific Aims 1. Identification of P2X receptor subtypes in myenteric neurons in small intestine. This study was done in tissues from mice and guinea pigs. Although immunohistochemical localization of P2X subunits in guinea pig myenteric neurons has been done (Castelucci et al., 2002; Poole et al., 2002; VanNassauw et a1. 2002), the subtypes of P2X receptors mediating fEPSPs in these tissues are not fully identified. Selective agonists and antagonists were used to evaluate eletrophysiological responses to identify P2X receptor subtypes in guinea pig myenteric neurons. In addition to using pharmacological methods, I also took advantage of P2X2 and P2X; gene knockout mice to investigate the contribution of these two subunits to electrophysiological properties of myenteric neurons in mice. 2. Fast synaptic transmission elicited by trains of electrical stimulation. 39 This study was conducted in myenteric neurons of guinea pig small intestine in vitro. Intracellular electrophysiological recording techniques including single electrode voltage-clamp (SEVC) were used to characterize fast synaptic transmission during trains of electrical stimulation in LMMP preparations in vitro. Fast synaptic transmission elicited by single electrical nerve stimulation has been studied extensively. However, fast excitatory postsynaptic potentials (fEPSPs) occur in bursts in response to physiological stimulation. These studies have characterized fast synaptic transmission during trains of stimulation and examined the contributions of ATP and ACh to fEPSPs during a short train of electrical stimulation. 3. Presynaptic S-HT4 receptor-mediated modulation of fast synaptic transmission. 5-HT4 receptors mediate enhancement of single fEPSPs in myenteric neurons. 5- HT4 receptors may also regulate the dynamic change of synaptic transmission evoked by trains of stimulation. These studies have tested the hypothesis that activation of 5-HT4 receptors reduced the decline in amplitude (rundown) of fEPSPs during trains of stimuli and facilitates the rate of recovery from synaptic run-down. The regulation is mediated by 5-HT agonists activating adenylate cyclase-cyclic 3’, 5’ adenosine monophosphate (cAMP)-PKA signaling pathway. 5-HT4 receptors mediate facilitation of fEPSPs by increasing transmitter release but the cellular mechanism remains unclear. These studies have examined the hypothesis that activation of 5-HT4 receptors enhances the release probability from single release sites (varicosities). Two experimental methods were used in these studies. Intracellular electrophysiological recordings were made from LMMP 40 preparations from guinea pig tissue in vitro. Patch-clamp recordings were made in guinea pig myenteric neurons maintained in primary culture. 41 METHODS Animals All the procedures for handling animals were used in this study were approved by the All University Committee on Animal Use and Care at Michigan State University. Intact myenteric plexuses were dissected acutely from adult Guinea pigs (Hartley, male, 250 — 350 g) and Mice. Guinea pigs were available from Bioport Company (Lansing, MI). Mice were obtained from Roche Bioscience (Palo Alto, CA). P2X; or P2X3 gene knockout (P2X2'l' or P2X3'l') mice were generated as described (Cockayne et al. 2000). All mice used in this study have the genetic background 1290la X C57BL/6 (Harlan), and were derived from homozygous F2 crosses. The comparison analysis was performed between knockout mouse and its wild type from same family. Genotype confirmation of all animals was carried out by southern blot analysis as previously described (Cockayne et al. 2000). Longitudinal muscle myenteric plexus (LMMP) preparation Guinea pigs were euthanized by halothane inhalation and by severing the major neck blood vessels. Mice were anesthetized with halothane and then killed by cervical dislocation. The small intestine was removed from mice or guinea pigs and placed in oxygenated (95% Oz and 5% C02) Krebs’ solution of the following composition (millimolar): NaCl, 117; KCI, 4.7; CaClz, 2.5; MgClz, 1.2; NaHC03, 25; NaHzPO4, 1.2; 42 glucose, 11. The Krebs’ solution also contained nifedipine (1 uM) for blocking L-type calcium channels and scopolamine (1 pM) for blocking muscarinic cholinergic receptors to inhibit longitudinal muscle contractions. These drugs do not affect normal myenteric neuronal properties. A 1.5 cm segment of intestine was cut open along the mesenteric border and pinned out flat with the mucosal surface up in a petri dish lined with silastic elastomer. A longitudinal muscle-myenteric plexus (LMMP) preparation was made by peeling away the mucosal, submucosal, and circular muscle layers using fine forceps and scissors. A 5 mm2 piece of longitudinal muscle myenteric plexus was transferred to a small (2 ml volume) recording chamber lined with silastic elastomer. The preparation was stretched lightly and pinned to the chamber bottom using stainless steel pins (50 pm diameter). The preparation was superfused with 37°C oxygenated Krebs’ solution at a flow rate of 4 ml/min. Individual myenteric ganglia were visualized at 200x magnification using an inverted microscope (Olympus CK-2) with differential interference contrast optics. Myenteric neurons maintained in primary culture Tissue for primary myenteric cell culture was obtained from newborn guinea pigs (~36 h in age; ~70 g in weight). Newborn guinea pigs were sacrificed by severing the major neck blood vessels and spinal cord after deep halothane anesthesia. The small intestine was placed in cold (4°C) sterile-filtered Krebs' solution of the following composition: ll7mM NaCl, 4.7mM KCI, 2.5 mM CaClz, 1.2 mM MgC12, 1.2 mM NaHzPO4, 25 mM NaHCO3, and 11 mM glucose. The longitudinal muscle myenteric plexus was removed from the entire length of small intestine and cut into 5-mm-long 43 pieces. The dissected tissues were divided into four aliquots and placed in 1 ml of Krebs' solution containing 1600 U of trypsin (Sigma Chemical Co., St. Louis, M0) for 25 to 30 min at 37°C. After trypsin incubation, the tissues were triturated 30 times and then centrifuged at 900g for 5 min with a bench-top centrifuge. The supernatant was discarded, and the pellet was then be resuspended in sterile Krebs' solution and incubated (25-30 min, 37°C) in Krebs' solution containing 2000 U crab hepatopancreas collagenase (Calbiochem-Novabiochem, Corp., La Jolla, CA). The suspension was triturated and then centrifuged for 5 min. The pellet was suspended in Eagle's minimum essential medium containing 10% fetal calf serum, gentamicin (10 pg/ml), penicillin (100 U/ml), and streptomycin (50 ug/ml) (all from Sigma). Cells were plated on plastic dishes coated with poly-L-lysine and maintained in an incubator at 37°C in an atmosphere of 5% C02 for up to 2weeks. After 2 days in culture, 10 pM cytosine arabinoside was added to the minimum essential medium to limit smooth muscle and fibroblast proliferation, and the medium was changed twice weekly thereafter. Intracellular electrophysiological recording from myenteric neurons Individual myenteric ganglia can be visualized at 200x magnification using an inverted microscope (Olympus CK-2) with differential interference contrast optics. Intracellular recordings were obtained from single neurons using glass microelectrodes filled with 2 M KCl and a tip resistance of 80-120 MO. An amplifier with an active bridge circuit (Axoclamp 2A, Axon Instruments, Foster City, CA) was used to record membrane potential. In most experiments, the membrane potential was hyperpolarized using constant DC current to avoid action potentials when evoking fEPSPs. The 44 amplified signals then were filtered at 1 kHz using a four-pole, low-pass Bessel filter (Warner Instruments, Hamden, CT), and digitized at 2 kHz using a Digitdata 1200 analog/digital converter (Axon Instruments). Data were acquired and stored using Axotape 2.02 or Axoscope 8.2 software (Axon Instruments). All fEPSPs were elicited using a glass pipette (tip diameter 40-60 uM) filled with Krebs’ solution as a focal stimulating electrode. The stimulating electrode was positioned closely over an interganglionic nerve strand. Nerve fibers were stimulated electrically using single stimuli of 0.5 msec of duration at a rate of 0.1 Hz. A digital average of eight individual fEPSPs was used as a measurement of fEPSP amplitude in the absence or presence of drug treatments. Patch-clamp recording from primary cultured neurons Whole-cell and outside-out patch-clamp recordings were obtained via standard methods. Recordings were carried out at room temperature with patch pipettes with tip resistances of 3 to 5 M9 for whole-cell and 5 to 10 M9 for single-channel currents in - outside-out patches; seal resistances were >5 G9. The tips of pipettes used for single- channel recordings were coated with Sylgard (Dow Corning, Midland, MI). The pipette solution contained the following: 160 mM CsCl, 2 mM MgC12, 1 mM EGTA, 10 mM HEPES, lmM ATP, and 0.25 mM GTP, the pH and osmolarity were adjusted to 7.4 (with CsOH) and 315 mosmong (with CsCl), respectively. All recordings were made with an Axopatch 200A amplifier. Data were acquired with pClamp 8.0 software. Currents were sampled at 2 kHz and were filtered at 1 kHz (4-pole Bessel filter, Warner Instruments, Hamden, CT) and stored on a computer hard drive for offline analysis. 45 Immunohistochemistry After electrophysiological recordings with Neurobiotin-filled microelectrodes, the myenteric plexus-longitudinal muscle preparation was fixed overnight at 4 °C in Zamboni’s fixative (2 % (v/v) formaldehyde and 0.2 % (v/v) picric acid in 0.1 M sodium phosphate buffer, pH 7.0). The fixative was removed using three washes of dimethyl sulfoxide at 10 min intervals. Tissues were then washed three times with phosphate- buffered saline (PBS) (0.01 M; pH 7.2) at 10 min intervals. Subsequently the preparation was incubated overnight with a primary antibody against nitric oxide synthase (NOS) at room temperature. After primary antibody incubation, tissues were washed three times at 10 min intervals with PBS. Tissues were then incubated (1.5 h at 23 °C) with goat anti- rabbitt IgG (1:40 dilution in PBS; Jackson Immunoresearch Laboratories, West Grove, PA, USA) conjugated to fluorescein isothiocyanate (FITC) to recognize NOS immunoreactivity and Texas Red to recognize the neurobiotin-filled impaled neurons. Tissues were washed three times with PBS and mounted in buffered glycerol for fluorescence microscopy. Images were obtained using Nikon, Eclipse TE2000-U (Tokyo, Japan). Drugs and drug application All drugs used in the studies are listed in Table 1. In general, four different methods were used for drug application. 46 1. Superfusion: Antagonists were applied by superfusion in a known concentration by addition to the superfusing Krebs’ solution. Antagonists were applied for 6-20 minutes prior to measuring the amplitude of evoked responses. 2. Pressure ejection: Local application of agonists was accomplished by ejection from the tip of a micropipette (~20 pm tip diameter) placed within 150 pm of the impaled neuron. Agonists were applied using short pulses of nitrogen gas (3 to 35 ms, 10 psi) using a Picospritzer II (General Valve, Fairfield, NJ). 3. Ionophoresis: ACh or ATP was applied by ionophoresis from an electrode (10 MO 0.05) in tissues from P2X2+l+ and P2X2'/' mice, respectively; 2.0 :l: 0.2 and 2.4 :k 0.1 ms (P > 0.05) in P2X?” and P2X3‘/' mice, respectively. The peak amplitudes of the action potentials slow afterhyperpolarization were 5.9 i 1.1 and 5.8 :l: 0.7 mV in P2X2+/+ and P2X;"' mice (P > 0.05), respectively. These numbers were 5.2 i 0.7mV and 5.4 i 0.6mV in (P > 0.05) in P2X)“+ and P2X3'l' mice, respectively. The duration of the 52 Table 2. Electrical properties of myenteric neurons in mouse small intestine A. Resting membrane potential (RMP) and input resistance (Rim) of AH and S neurons in P2X3 +/+ and P2X3 "' mice. AH neurons S neurons P2X3 *’* P2X3 "' P2X; “i P2X; "' (n=5) (n=7) (n=22) (n=22) Mean 632t4 633:2 62i2 59i2 RMP (mV) Range 50 — 71 57 -— 73 49 — 73 42 ~ 72 Mean 172i35 153i20 121:1:14 127i10 Rinput (M0) Range 93 — 285 79 — 212 57 — 277 65 — 204 53 B. Resting membrane potential (RMP) and input resistance (Rum) of AH and S neurons in P2X2 +/+ and P2X; 4' mice. AH neurons S neurons 122x2 *’* P2X; "' P2X2 *’* P2X2 "' (n=3) (n=7) (n=1 7) (n=2 1) Mean 69i5 62i4 63:1:2 58i2 RMP (mV) Range 61 -— 80 48 — 81 50 — 74 45 — 80 Mean l36:t35 132i21 122i13 112i9 Rinput (M9) Range 96 - 207 85 — 221 65 — 241 54 — 167 54 Figure 5. Action potentials in myenteric S neurons in mouse small intestine A. Action potentials recorded from myenteric S neurons in tissues from P2X; knockout mice (P2X2‘l') and wild type mice (P2X2+/+). S neurons have action potentials followed by short duration afterhyperpolarizations. TTX blocks the action potential in S neurons. B. Action potentials recorded from S neurons in tissues from a P2X3’/' and P2X3‘L/+ mice. These action potentials also have a short afterhyperpolarization (as arrows point) and they are blocked by TTX. 55 Figure 5 A P2X2 W sz2 -/- Control L 10 ms TTX (0.3 uM) B Control TTX (0.3 pM) 20 mV 0.15 nA 1 5ms P2X H/«Jk/ j - ALL. 3L. -- T. _. xv w , __ K 5ms I 20 mV 0.5 ms P2X3"’ {v A __ w_ _ .—e K 56 Figure 6. Action potentials in myenteric AH neurons in mouse small intestine A. Action potentials recorded from AH neurons in P2X2'/' knockout and P2X2’L/+ tissues are similar. They have long-lasting afterhyperpolarization (as arrows show). B. Similarly, there is no difference between action potentials recorded from AH neurons in P2X3'I’ and P2X3+/+ tissues and they have long-lasting hyperpolarization (arrows). In addition, action potential recorded from AH neurons is not blocked by TTX. 57 Figure 6 A szz H+ A, \ 20 mV PZXZ J- I A/ E 5— ms B Control TTX (0.31M _I—1 _J—_‘l 20 mV 5—150. 15 nA 5__ ms P2)(3 +l+ r wr- .--w :fi‘r‘w‘r v'\ 20 mV I P2X3 --/ A afterhyperpolarization was 6.4 :t 0.9 s in AH neurons from P2X2+l+ mice and 5.7 :i: 0.4 s (P > 0.05) in AH neurons from P2X2"/' mice. In P2X3H+ and P2X3'/' mice, these numbers were 6.3 i 1.33 and 6.5 i 0.7 s (P> 0.05), respectively. P2X; subunit knockout altered P2X receptor-mediated responses in S neurons Fast EPSPs recorded from S neurons in tissues of P2X2H+ and P2X2'I‘ mice were not different in the peak amplitude or time course (Table 2, Figure 7). This result suggests that P2X2 subunits may not contribute to the P2X receptors mediating fEPSPs in S neurons. However, in tissues from P2X2+Pr mice, nearly all fEPSPs were reduced in amplitude by the P2 receptor antagonist pyridoxalphosphate- 6-azophenyl-2,, 4,- disulphonic acid (PPADS, 10 uM) (Figure 7). All fEPSPs were blocked by the combined application of PPADS and the nicotinic acetylcholine receptor antagonist, mecamylamine (10 uM). However, PPADS had no effect on fEPSPs recorded from S neurons in tissues from P2X2'l' mice (Figure 7). Mecamylamine reduced the amplitude of fEPSPs recorded 4'” tissues, but blocked fEPSPs recorded from S neurons in tissues from S neurons in P2X; from P2X2'l’ mice (Figure 8). The amplitude of fEPSP in presence of mecamylamine recorded from S neurons in P2X2+l+ tissues was larger than that in S neurons of P2X;"' tissues (Figure 8). Puff application of nicotine caused a depolarization of S neurons from P2X2'l' mice and this response was similar in amplitude to that recorded from S neurons in P2X2+/+ tissues (Figure 9). Focal application of ATP, but not or,B-mATP depolarized S neurons in tissues from P2X?” mice (Figure 9). Neither ATP nor or,B-mATP caused a response in S neurons from P2X2’/' mice (Figure 9). 59 Figure 7. PPADS inhibits fEPSPs recorded from S neurons in tissues from P2X?” but not in tissues from P2X2'/' mice. A. Representative fEPSPs recorded from S neurons in P2X2”+ (left) and P2X2'/’ (right) tissues in the absence or presence of PPADS (10 pM). The fEPSPs recorded from neurons in tissues from P2X2+H mice were inhibited by PPADS (left) while fEPSPs recorded from neurons in P2X2’l‘ tissues were unaffected by PPADS (right). B. Pooled data show that the amplitudes of fEPSPs recorded from S neurons in P2X2+H (n = 5) and P2X2’l' (n = 6) tissues were similar. PPADS did not alter fEPSP amplitude recorded from P2X2'l' S neurons but it inhibited the fEPSP in S neurons from P2X2+/+ tissues. Data are mean :i: SEM. “*” Significantly different from P2X;+le control (p<0.05). 60 Figure 7 P2X2+’+ P2X2'" / Control PPADS Control PPADS '20 mV 20 ms B [:3 Control 30 - - PPADS (10 pM) A 20 - > E. 8% o. 10 r k! 0 P2X *’* P2X "' 61 Figure 8. Mecamylamine reduced fEPSPs in S neurons from tissues of P2X2+H mice but completely blocked fEPSPs in S neurons from P2X2’/' mice. A. Representative fEPSPs recorded from S neurons in P2X2”+ (left) and P2X2'/’ (right) tissues. The fEPSP in P2Xz+l+ neuron is only partly inhibited by mecamylamine (10 uM), but the fEPSP is blocked by mecamylamine in a neuron from a P2X2'l' mouse. +/+ B. Pooled data show that mecamylamine reduced fEPSP amplitude in P2X; neurons (n = 5). However, mecamylamine blocked fEPSPs in P2X2‘l’ neurons (n = 9). “*” Significantly different from P2X?” or P2X2'I' control amplitude (P < 0.05); “#” +/+ Significantly different from the amplitude of the P2X2 fEPSP in the presence of mecamylamine (P < 0.05). Data are mean i SEM. 62 Figure 8 A P2x,+/+ P2X2‘" Control Control / , flamine Mecamylamine ,— '20 mV ‘ 20 ms B 30 ' 1:: Control - Mecamylamine (10 pM) A ,__T__ I > 20 .. .5, 0. if 10 L15 # * 0 L P2X,” P2X "' 2 63 Figure 9. Responses of murine myenteric S neurons in PZXZW and P2X2'l‘ tissues to nicotine, ATP and a, B—mATP A. Nicotine induced similar amplitude depolarizations of S neurons from PZXZH+ and P2X2'l' mice. ATP depolarized S neurons from P2X2+l+ but not P2X2'I' mice, while or, B-mATP did not depolarize any S neurons. Drugs were applied (at the arrows) by pressure ejection from a pipette positioned near the impaled neurons. The concentration of each drug in the pipette was 1 mM. B. Mean data from experiments similar to those shown in A. There was no difference in the amplitude of the nicotine-induced depolarization recorded from S neurons in tissues from P2X?+ (n = 15) and P2X2'l' (n = 12)mice. “*” The ATP-induced depolarization was significantly smaller in P2X2”' neurons compared to those in recorded from P2X2H+ neurons. or, B-mATP did not elicit a response in S neurons from either type of mouse. 64 Figure 9 Nicotine P2X2+/+ H TWW P2X2"' ATP N 00 O O .3 O Depolarization (mV) O 65 or,B-mATP Properties of P2X receptors were not affected by P2X; subunit knockout in S neurons Single stimuli applied to interganglionic nerve strands evoked fEPSPs in S neurons from tissues of P2X3+I+ and P2X3‘l‘ mice (Figure 10, 11). The mean amplitude of fEPSPs recorded from neurons in P2X3+l+ and P2X3'/' tissues was not significantly different (Figure 10, 11). The pharmacology of fEPSPs between P2X3+I+ and P2X3’l’ tissues was similar. All fEPSPs were reduced, but not completely blocked, by the nicotinic receptor antagonist mecamylamine (10 uM) (Figure 10). Amplitudes of fEPSPs in presence of mecamylamine were not different in P2X3H+ and P2X3’l' mice. Similarly, fEPSPs in tissues from P2X3+l+ and P2X3'/' mice were both partly inhibited by PPADS (Figure 11). Also, fEPSP amplitude in S neurons from P2X3+I+ and P2X3'l' mice were not different in presence of mecamylamine. ATP caused a depolarization that was similar in amplitude in S neurons from P2X3+l+ and P2X3'/' mice (Figure 12). However, or, B-mATP did not cause depolarization in S neurons in tissues from either type of mice (Figure 12). or, p-mATP depolarizes AH neurons from szf’+ and szz"' mice Focal application of afi-mATP (lmM), an agonist selective for P2X. and WK; containing P2X receptors caused a rapidly developing depolarization that was associated with a decrease in membrane input resistance in AH neurons from P2X2+l+ and P2X2'l' mice (Figure 13). The or,B-mATP-induced response was mimicked by ATP applied to the same neurons and was blocked by PPADS. Amplitudes of or, B-mATP-induced 66 Figure 10. Effects of mecamylamine on fEPSPs from murine myenteric S neurons in P2X3‘i’+ and P2X;"' tissues. A. Representative recordings show mecamylamine (10 uM) reduced the amplitudes of fEPSPs in neurons in both P2X3H+ and P2X3'l’ tissues. B. Pooled data from experiments shown in A. There was no significant difference in the amplitude of fEPSPs recorded from S neurons from P2X3+l+ and P2X3'/' mice. Mecamylamine reduced the amplitude of the fEPSP in all S neurons but it did not completely block the fEPSP. * Significantly different from fEPSP amplitude recorded in drug free solution. 67 Figure 10 i>2x,+/+ Control / Mecamytamine (10 uM) 20 ms B 30 ' - Control E Mecamylamine (10 pM) fEPSP (mV) 9*. {5 <75 LL O P mm” mx+ 68 Figure 11. Effects of PPADS on fEPSPs from murine myenteric S neurons in P2X3”+ and P2X3'l' tissues. A. Representative fEPSPs recorded from S neurons in tissues from P2X3,+/+ and P2X3"' mice are inhibited but not blocked by PPADS. B. Pooled data from experiment illustrated in A. PPADS produces a significant inhibition of fEPSPs but it does not block the synaptic response. There was no difference in the amplitude of the fEPSPs recorded from S neurons in P2X3+H and P2X3”' tissues. 69 Figure 11 P2X,“+ szy- __I 10 mV __I 20 ms - Control 25_ n=5 [:1 PPADS (10 nM) 204 fEPSP (mV) P2X,+/+ P2X3'/' 70 Figure 12. Responses of murine myenteric S neurons in P2X3m and P2X3‘/' tissues to ATP and a, B—mATP A. Representative recordings Show that ATP, but not or, B-mATP induced depolarization in S neurons. B. Pooled data indicate there were no differences in the amplitude of ATP- induced depolarizations in S neurons from P2X?” and P2X3’l‘ mice. or, B-mATP did not induce depolarization in any tested S neurons. 71 Figure 12 . I l . .‘l.‘.,. '1 1‘ III I ‘20 mV 055 on... n... n. _ATP (2 mM) 1 c: a, B-mATP 03 O A n=12 Depolarization (mV) 8 n: P2X3+I+ P2X3-l- 72 Figure 13. Responses of murine myenteric AH neurons in P2X?” and P2X2'" tissues to or, B-mATP A. Representative recordings of agonist-induced responses from AH neurons in P2X2+H and P2X2'/' mouse tissues. or, B-mATP was applied by pressure ejection (at the arrows) fi'om a pipette positioned near the impaled neuron. The concentration of or, B- mATP in the pipette was 1 mM. B. Mean data show that the amplitude of the or, B-mATP-induced depolarization was similar in tissues from P2X2+l+ (n = 3) and P2X2'/' (n = 4) mice. 73 Figure 13 A a,B-mATP P2X2+/+ IW (x,B-mATP P2X2'" "IM— 20 mV 1—' _I10 mV S 0.5 s B 25 ’S E, C .9 ’5 .N <3 0 O. (D D szz+l+ sz2 4- 74 depolarizations were similar in tissues from P2X2H+ and P2X2'/" mice (Figure 13). ATP— and or, B-mATP— induced depolarization in AH neurons from P2X3J"+ and P2X3'l' mice Local application of ATP (2 mM) from a pipette positioned near the neurons caused a depolarization that was similar in amplitude in neurons from P2X3H+ and P2X3’l' mice (Figure 14). The ATP-induced response was blocked by PPADS and was associated with a decrease in membrane input resistance (Figure 14). PPADS blocks some P2X receptors and P2Y. receptors for ATP so responses by ATP in murine neurons could be mediated by either P2X or P2Y. receptors. To address this possibility, I tested the effect of or, B-mATP, an agonist of P2X. and P2X3 subunit-containing P2X receptors. It was +/+ found that or, B-mATP caused depolarization in AH neurons in preparations from P2X3 but not P2X3'l' mice (Figure 14). B. Pharmacological Identification Of P2X Receptor Subtypes In Guinea Pig Myenteric Neurons Depolarization induced by ATP and or,B-mATP in AH and S neurons in guinea pigs Intracellular electrophysiological recordings were obtained from 112 myenteric S type neurons from guinea pig ileum. 85% of the neurons (95 cells) were depolarized by local application of a, B-mATP, an ATP structure analog identified as an agonist selective 75 Figure 14. Responses of murine myenteric AH neurons in P2X?” and P2X3’" tissues to ATP and or, B—mATP A. Representative responses induced by ATP and or, B-mATP. ATP depolarized AH neurons from both types of mice. or, B-mATP evoked depolarization in AH neurons in tissues from P2X3‘L/+ but not P2X3'/' mice. B. Pooled data from experiments illustrated in A. The amplitude of ATP-induced depolarizations was similar in AH neurons from P2X3H+ and P2X3'l' mice whilst or, [3- mATP induced depolarization in AH neurons from P2X3+H but not P2X3'/’ mice. 76 Figure 14 A ATP mp-mATP V V > E P2X3+/+ 8 g I . , 1 s ”XS/W B -ATP (2 mM) 25_ E: (X, B'mATP (2 mM) 1 _ § 204 "-5 ”=3 "=6 e . l C .9 15‘ ‘c'u' . -E' g 10. 0 d 8 5 c: . o. P2X3+’+ P2X,-/- 77 for P2X. and P2X3 subunits containing P2X receptors. The rest 15% of the S neurons (17 cells) were insensitive to ATP or a, B-mATP. They were cholinergic neurons and fast synaptic excitation was mediated solely by ACh acting at nAChRs as mecamylamine (10 uM) completely blocked fEPSPs recorded from these neurons. The amplitudes of ATP and a, B-mATP-induced depolarization were 21.5 mV and 22.5 mV (P>0.05) (Figure 15), respectively, which were blocked by PPADS (10 pM). A total of 28 AH neurons was studied; 17 neurons were tested for ATP sensitivity. 65% of these neurons were depolarized by ATP. All 28 AH type neurons were tested for a, B-mATP sensitivity and only 17% of them were depolarized by local application of a, B-mATP. The amplitudes of depolarization caused by ATP and a, B- mATP were 14.7 mV and 11 mV, respectively (Figure 15). Effects of TNP-ATP on fEPSPs and ATP-induced depolarization in S neurons TNP-ATP (10 pM), an ATP analog, is an antagonist selective for P2X], P2X; or P2X2/3 receptors. It was found that TNP-ATP reversibly reduced depolarizations induced by ATP (Figure 16). Superfusion of TNP-ATP also reversibly reduced amplitude of fEPSPs recorded from S type neurons (Figure 17). These data also suggest that P2X receptors of S neurons contain P2X., P2X3 or P2X2/3 subunits. 78 Figure 15. Depolarization induced by ATP (lmM) and or, B—mATP (1 mM) in S and AH myenteric neurons in guinea pig small intestine A. Representative recordings of depolarizations induced by or,B-mATP and ATP in AH neurons. The arrowheads refer to pressure injection of or,B-mATP or ATP. Only 17% of the tested AH neurons have similar responses to these two agonists. B. Representative recording traces show depolarization induced by ATP and (1,13- mATP in S myenteric neurons. The depolarizations were blocked by PPADS (10 uM). 85% S neurons have similar responses to these two agonists. C. Pooled data indicated the amplitudes of or, B—mATP and ATP-induced depolarization in S neurons are similar, which are bigger than these in AH neurons. Data are the mean :t SEM obtained from 5 AH neurons and 12 S neurons. “*” indicates a significant difference from responses in AH neurons. 79 Figure 15 A o, B-mATP ATP |20mV ,~ ~ 1 s PPADS (10 M) Ezra, B-mATP (1 mM) - ATP (1mM) Depolarization (mV) AH Neurons S Neurons 80 Figure 16. Effect of TN P-ATP on depolarizations induced by ATP in S neurons A. Representative traces show that TNP-ATP reduced amplitude of ATP-induced depolarizations. The effect of TNP-ATP was reversible. Arrowheads indicate pressure ejection of ATP. B. Data are mean i SEM from experiments (n=7) similar to that shown in A. “*” indicates a significant difference from control and wash groups. 81 Figure 16 A ATP Control Depolarization (mV) 00 O N O _x O J 10 mV ATP ATP 1 S V TNP-ATP (10 pM) Wash Control TNP—ATP Wash 82 Figure 17. Effect of TNP-ATP on fEPSPs in S neurons A. Representative recording traces indicate TNP-ATP reduced the amplitude of fEPSPs. The effect of TNP-ATP on fEPSP amplitude was reversible. B. Data are mean i SEM from experiments (n=4) similar to that shown in A. “*” indicates a significant difference from control and wash groups. 83 Figure 17 A Control TNP-ATP (10 |JM) Wash _J 10mV 10ms B 25 { fEPSP (mV) Control TNP-ATP Wash 84 P2X RECEPTORS COUPLE TO CALCIUM-DEPENDENT POTASSIUM CONDUCTANCE IN GUINEA PIG MYENTERIC NEURONS An afterhyperpolarization is induced by activation of P2X receptors Activation of P2X receptors by a, B-mATP induced a depolarization in myenteric S neurons of guinea pig small intestine. This resulted from cations flowing through P2X receptors as input resistance (Rinpm) was reduced during depolarization. However, it was observed that, in 52 of 95 (55%) S type neurons, (1, B-mATP-induced depolarizations were followed by a hyperpolarization (Figure 18). PPADS abolished both the depolarization and hyperpolarization. The amplitudes of depolarization and hyperpolarization were positively correlated (Figure 18). It was also observed that membrane conductance was increased during hyperpolarization, which suggested that opening of ion channels mediates the after-hyperpolarization. To eliminate the possibility that a, B-mATP-elicited hyperpolarization was due to change of membrane potential, single electrode voltage clamp (SEVC) recordings from similar preparations were made to confirm a, li-mATP-induced depolarization and hyperpolarization. In SEVC recording, inward and outward currents were recorded in S neurons in response to a, B-mATP, which underlie the depolarization and hyperpolarization, respectively (Figure 18). Similarly, the amplitudes of inward and outward currents were also positively correlated and they were blocked by PPADS. 85 Figure 18. or, B—mATP—induced a biphasic response in S neurons. A. Representative recording showing that or, B-mATP induced a depolarization followed by a hyperpolarization (resting membrane potential = —50 mV). Downward deflections in the trace are voltage responses caused by hyperpolarizing current pulses. A decrease in the amplitude of these responses during the depolarization and hyperpolarization indicates an increase in membrane conductance. Arrowheads indicate pressure ejection of a,B-mATP. B. Current recording in SEVC mode shows a similar result as in A. or, B-mATP induced an inward current followed by an outward current at a holding potential of —50 mV. The downward deflections are currents caused by voltage steps. An increase in the amplitude means an increase in membrane conductance. Arrowheads indicate pressure ejection of or, B-mATP. C. There is a positive correlation between the amplitude of the hyperpolarization and the depolarization. 86 Figure 18 A B |20mV 23 N I 0.2nA 25 C 30- E 25- e v _ e e. 8 20+ 0 U ,,/’ "a » r w" .2. ‘15- /./0.0 . .9 . . e)” . 0.0 O 10- // e D. x L. I!!!“ O. . O 8. 5- ° 0 e e f . 0 x l L o 510.15‘20‘25‘30 Depolarization (mV) 87 The ionic basis of the afterhyperpolarization One mechanism for increased conductance in mediating hyperpolarization could be an efflux of K+ ions. It has been reported that intermediate Cazl-activited K+ channels mediate the action potential afterhyperpolarization in AH neurons (Vogalis et al., 2002; Fumess et al., 2004; Neylon et al., 2004). In myenteric S neurons, a depolarization induced by nicotine is followed by a hyperpolarization, which is also due to activation of calcium-activated potassium channels (Tokimasa et al., 1983). a, B-mATP-induced hyperpolarizations in S neurons might also be mediated by Cay-activated K+ channels. Calcium dependency of the a, B-mATP after-hyperpolarization was examined by changing extracellular calcium concentration from normal level of 2.5 mM to 0.25 mM. Lowering extracellular [Ca2+] at 0.25 mM almost abolished fEPSP and reduced (1, B- mATP-induced depolarization from 20.6 mV to 17.8 mV (Figure 19). As a consequence, it reversibly and dramatically reduced the hyperpolarization from 14.6 mV to 1.8 mV (Figure 19). The subsequent experiment was conducted to examine whether the u,B-mATP afterhyperpolarization or outward current was carried by KL SEVC recordings were made from LMMP preparations. The inward and outward currents were induced by a, B- mATP. The relationships between holding potential and inward and outward currents were linear (Figure 20). The reversal potentials of inward and outward currents were 3.3 mV and —79 mV, respectively. These data indicate that the inward current underlying depolarization was carried by non-selective cations through P2X receptors as the reversal 88 Figure 19. a, B—mATP-induced afterhyperpolarization is Ca2+ sensitive A. Low extracellular Ca2 I (0.25 mM) abolished fEPSPs. Restoring extracellular Ca2+ to 2.5 mM restored fEPSP amplitude. Ca2+ at 0.25 mM reduced the depolarization induced by or, B-mATP. The after-hyperpolarization was also reduced. Restoring Ca2+ to 2.5 mM restored the depolarization and hyperpolarization. B. Pooled data (mean i “*9, SEM) obtained from similar experiments (n=7) to that shown in A. indicates a significant difference from 2.5 mM Ca2+ conditions. 89 Figure 19 A 2.5 mM Ca2+ 0.25 mM Ca2+ 2.5 mM Ca2+ 10 mV r 20 ms J20 mV " N 2 s B - Normal Krebs (2.5 mM Ca2+) 25 - Low Ca Krebs (0.25 mM Ca2+) 9 E o '0 3 '6. E <1: * Depolarization Hyperpolarization 90 Firl lull \'Ol 10 Figure 20. 0., B—mATP—induced afterhyperpolarization is K+ dependent. A. Representative recordings show inward and outward currents at different holding potentials. B. Current voltage relationship of the inward current. C. Current voltage relationship for outward current induced by or, B-mATP in S-neurons. Current voltage relationships were linear. The reversal potential for the inward current was 3 mV. The reversal potential for the outward current was -79 mV. Data are mean 3: SEM obtained from 7 neurons. 91 Figure 20 92 A B “'B’mATP Membrane potential (mV) -100 -80 -60 -40 -20 0 -30 mV 1,. " 'f-0.1 , O S --o.2 a . 5’. 1.03 3 --0.4 J -70 mV -90 mV . 0.08 9 - 0.04 3 E. ’5 . n . . l . 1 x n . 1 000 3 mimbrane potential (mV) ‘ -0.04 potential is close to 0 mV. The outward current mediating the afterhyperpolarization was carried by KL because the reversal potential is approximately same as K“ equilibrium potential of -80 mV. 1 then began to identify the Cay-activated K+ channel (KCa) subtypes that mediated the afterhyperpolarization or outward currents in myenteric S neurons. Three toxins for major KCa subtypes were used (Vergara et al., 1998). Iberiotoxin (IBTx) at concentration of 0.1 uM selectively blocks Kc, with large conductance (BK). Apamin at 0.1 uM selectively blocks Kca with small conductance (SK). Clotrimazole at 10 uM selectively blocks Kca with intermediate conductance (1K). Application of these antagonists had no effects on AHP induced by a, fi-mATP in S neurons (Figure 21). Immunochemical characterization of S neurons with after-hyperpolarization P2X3 receptors are expressed primarily in the inhibitory motor neurons ((LePard et al., 1997; Poole et al. 2002; Van Nassauw et al. 2002). As shown above, not in all S type neurons but in an approximately half of them, P2X3 receptors coupled to a Kc, that mediates the after-hyperpolarization. One hypothesis could be this subset S neurons may utilize the P2X; receptors to accomplish their unique function of those neurons. Here I tested the idea that these S neurons were descending inhibitory motor neurons. The descending inhibitory motor neurons can be labeled by an antibody against NOS. The results indicated that the S neurons with a, B-mATP-induced hyperpolarization are either NOS-ir positive or NOS-ir negative. There was no significant association between this subset of S type neurons and descending inhibitory motor neurons (Figure 22). 93 Figure 21. Lack of effect of KC, channel blockers on the a,B—mATP-induced afterhyperpolarization. A. IBTx, apamin and Clotrimazole, selective blockers of big, small and intermediate conductance calcium-activated potassium channels, respectively, had no effect on the afterhyperpolarization induced by or, B-mATP. Data are mean i SEM obtained from 6, 4 and 7 neurons in IBTx, apamin and Clotrimazole, respectively. B. Clotrimazole blocked action potential afterhyperpolarization in AH neurons in the same preparations from which S neurons were studied. These data confirmed the effectiveness of application of Clotrimazole. 94 Figure 21 A 20 — A I:IContro| E ' _Treatment v 16 - r: .9 ' s 12- 5 . o e 8 — a) Q . E‘ o 4- a: < lberoritoxin Clotrimazole Apamin 0.1pM 1O uM 0.1pM B |20mV 25 control “‘1 - ._... Clotrimazole fl wash 95 Figure 22. Identification of the neurochemical phenotype of S neurons that generate the a, B-mATP induced afterhyperpolarization There was no association between S neurons generating afterhyperpolarization and NOS-ir neurons. Green fluorescence indicates NOS-ir while red fluorescence reveals neurobiotin used to mark neurons from which electrophysiological recordings were obtained. A neuron generating the afterhyperpolarization contained NOS-ir (A) while C shows an example of a neuron with an afterhyperpolarization but that did not contain NOS-ir. A neuron that did not show an afterhyperpolarization contained NOS-ir neuron (B) while D shows an example of a neuron without NOS-ir neuron (D). 96 Figure 22 S neuron with AHP S neuron without AHP 97 FAST EXCITATORY SYNAPTIC TRANSMISSION DURING BURSTS OF NEURONAL ACTIVITY Fast synaptic transmission declined in amplitude during trains of electrical stimulation First fEPSPs were recorded when a train of electrical stimulation was applied on presynaptic nerve fibers at different frequencies. Fast EPSPs showed a decline in their amplitudes during test stimulus train at 5, 10 and 20 Hz. However, trains of stimulation could also elicit slow excitatory post-synaptic potentials (sEPSPs) in S and AH neurons in the ENS (Grafe et al., 1980; Morita and North, 1985; Wood and Kirchgessner, 2004). The depolarization occurring during sEPSPs would reduce the electrochemical driving force for Na+ and Ca2+ influx through nAChRs and P2X receptors thereby reducing fEPSP amplitude during a stimulus train. Therefore, I used single electrode voltage clamp (SEVC) method to hold the membrane potential at -60 mV in order to record fEPSC amplitude at a constant membrane potential. Holding the membrane potential constant would eliminate changes in driving force as a cause for alterations in the amplitude of fast synaptic responses during trains of stimulation. The SEVC experiments showed the similar results. At 0.5 Hz, the fEPSC declined in amplitude by 50% after the first stimulus but was maintained throughout the remainder of the stimulus train (Figure 23). However, at stimulus frequencies of 5, 10 and 20 Hz, the amplitude of fEPSCs declined to zero and the rate of rundown was frequency-dependent and was fitted by a single exponential decay (Figure 23). The time constants (r) for rundown of the fEPSCs at 5, 10 98 Figure 23. Trains of electrical nerve stimulation cause synaptic rundown. A. Representative recordings of fEPSCs during trains of stimulation at 5, 10 and 20 Hz. B. Pooled data show that the rate of rundown in fEPSC amplitude was frequency- dependent and was fitted by a monoexponential decay (solid lines). Data are the mean i: SEM. obtained from 21 neurons at 5 Hz, 32 neurons at 10 Hz and 19 neurons at 20 Hz. 99 Figure 23 A 5 Hz 200 ms 10 Hz 100 ms 20 Hz 50 ms lnA B 1.2 - c ”O :3 'a 0.9 - E (U (I) 0 ‘3 06 e ' ' U m b ._u_ 2 0.3L 5 I 5H2 Z . 20 H2 _ 10 Hz 0.0 l l 1 l 0.0 0.2 0.4 0.6 0.8 1.0 Time (s) 100 and 20 Hz were 0.35 i 0.05, 0.2 i 0.02 and 0.1 i 0.02 s, respectively. These data indicate that fast excitatory synaptic responses rundown in amplitude during trains of stimulation and that synaptic rundown is not caused by changes in driving force that might occur during sEPSP. fEPSCs exhibit paired pulse depression Paired pulse facilitation (PPF) is a property of many synapses in the peripheral and central nervous systems. I tested for the presence of PPF at synapses in the myenteric plexus by measuring the ratio of the amplitude of fEPSCs elicited by pairs of stimuli applied at intervals of 200, 100 and 50 ms (5, 10 and 20 Hz frequencies, respectively) (Figure 24). The second fEPSC in the pair was always smaller than the first indicating that synapses in myenteric plexus exhibit paired pulse depression (PPD) rather than PPF (Figure 24). The second fEPSC evoked at 10 and 20 Hz exhibited less depression than that evoked at 5 Hz (Figure 24). Recovery from synaptic rundown Synaptic recovery is another important indicator of synaptic properties. It reflects synaptic vesicle trafficking alter depletion. 1 next investigated the time course of recovery from synaptic rundown following a 10 Hz train of stimulation. In these experiments, a train of 20 stimuli at 10 Hz caused fEPSPs to decline in amplitude to 0 mV. Recovery from rundown was tested by waiting for periods of 1 to 12 s before a single stimulation was used to elicit a fEPSP (Vtest); this amplitude was expressed as a 101 Figure 24. Paired stimuli cause synaptic depression A. The second fEPSC in a pair evoked by stimuli at 5 (200 ms interval), 10 (100 ms interval) and 20 (50 ms interval) Hz was smaller in amplitude than the first fEPSC. B. Mean data from experiments similar to that shown in A. Data are the ratio of the second fEPSC vs. the first fEPSC (Isecond/Ifirst) in a stimulus pair. * Indicates that synaptic depression at 10 (n=32) and 20 (n=l9) Hz is less than that occurring at 5 Hz (n=21). Data are mean i SEM. 102 Figure 24 A I 2001115 1 100 ms 50 ms lnA B 1.0- 0.8- second first / 10 Hz 20 Hz 5H2 103 fraction of the amplitude of the first EPSP in the preceding stimulus train (Vfirst) (Figure 25). Fully recovery from synaptic rundown occurred in less than 12 s and the recovery time course was fitted by a monoexponential function; the recovery time constant was 7 i 2 s (Figure 25). The contribution of nAChRs and P2X receptors to rundown of fEPSPs As ACh, acting at nAChRs, and ATP, acting at P2X receptors, contribute to most fEPSPs recorded from myenteric neurons in the guinea-pig ileum, 1 next examined if there was selective rundown of one of these fEPSP components during a lO-Hz train of stimulation. The P2X-mediated component of the fEPSP was isolated by using mecamylamine (10 uM), to block nAChRs. Under these conditions, it was found that the remaining fEPSP declined in amplitude with a time course (1’ :03 i 0.07 5, n=6) that was identical to that occurring in the absence of mecamylamine (r =0.3 i 0.07 s, n=6) (Figure 26). The cholinergic component of the fEPSP was isolated by using PPADS (10 uM) to block P2X receptors. The cholinergic fEPSP ran down with a time course (1 =0.4 i 0.08 5, n=6) similar to that measured under control conditions and in the presence of mecamylamine (Figure 26). More than 90% of the fEPSP was abolished by co- application of mecamylamine and PPADS (Figure 26) indicating that the rundown was a mixed cholinergic—purinergic fEPSP. ACh and ATP ionophoresis were utilized to investigate if fEPSP rundown was due to postsynaptic nAChR and/or P2X receptor desensitization. ACh and ATP 104 Figure 25. Time-dependent recovery of fEPSPs from synaptic rundown A. Diagram of the protocol used to measure the time course of recovery from synaptic rundown. A conditioning train of 20 stimuli at 10 Hz was used to cause fEPSP rundown. A single test stimulus applied at various intervals after the conditioning train was used to elicit a fEPSP to determine the extent of recovery. At is the time interval between the last stimulus in the conditioning train and the test stimulus. Recovery is expressed as the ratio of the amplitude of the fEPSP caused by the test stimulus (V103,) vs. the amplitude of the fEPSP caused by the first stimulus (Vma) in the conditioning train. B. The time course of recovery was fit by a single exponential function with a I =7 2t 2 3. Data points are the mean i SEM. of observations made from 14 neurons. 105 Figure 25 A Sfl rst 8test I -- - l fig” ................ é-m— -— ----—-— $56 > 10 Hz train of 20 stimuli At=1, 2, 3 s... Vtest/V 1 st .0 .o .-* «h (I) N .0 o 0'2'4 é'é'1'0f1'2 Time(S) 106 Figure 26. Mecamylamine or PPADS does not alter fEPSP rundown A. The amplitude of fEPSPs was reduced by mecamylamine or PPADS (each at 10 uM) but the time course of rundown was not changed by the antagonists. The rate of synaptic rundown was fitted by a monoexponential function with s values of 0.3 2’: 0.07, 0.3 i 0.07 and 0.4 d: 0.08 3 under control conditions and in the presence of mecamylamine and PPADS, respectively. Each point is the mean i SEM. (n=6). B. Co- application of mecamylamine or PPADS inhibited the fEPSP by more than 90%. Data are mean :1: SEM (n=6). 107 Figure 26 fEPSPs (mV) 20- I control A15" . PPADS E L A mecamylamine :10- 35 t 0.. la! 5” 0 l J r . 4 . 1 1.. l . 0.0 0.2 0.4 0.6 0.8 1.0 Time (s) 20 - 15 - ' II Control 10 _ o PPADS+MecamyIamine 5 _ 0 108 Ant-‘1! - ionophoresis caused depolarizations that were similar in amplitude and time course to fEPSPs recorded from the same neurons, which suggests the same group of receptors were activated by ACh/ATP and by synaptic stimulation (Figure 27). Then ACh and ATP were applied as trains (10 Hz) of ionophoretic pulses and fEPSPs were evoked in the same neurons using a 10 Hz stimulus train. As same as described above, fEPSPs declined in amplitude during the stimulus train. However, in the same neurons, neither ACh nor ATP responses declined in amplitude during a train of agonist application (Figure 27). These data indicate that synaptic rundown is not due to postsynaptic receptor desensitization. Presynaptic mechanisms Presynaptic inhibitory receptors. Nerve terminals in the myenteric plexus express a number of receptors that mediate presynaptic inhibition of neurotransmitter release. These receptors include muscarinic M2 muscarinic receptors, 5-HT1A receptors, opioid receptors, (12 adrenergic receptors and adenosine A1 receptors. It is possible that during trains of stimulation, myenteric nerves release the endogenous agonists for one or more of these receptors and that presynaptic inhibition of ACh and/or ATP is responsible for fEPSP rundown. To test this hypothesis, fEPSP rundown was measured before and after application of antagonists to each of the receptors listed above. The exception being an antagonist for the M2 muscarinic receptor as scopolamine (1 uM), which blocks all muscarinic receptor subtypes, was present in the Krebs’ solution in all studies (see Methods and Materials). I used naloxone (10 pM) to block opioid receptors, NAN-190 109 Figure 27. Postsynaptic desensitization does not contribute to synaptic rundown A. ACh applied by ion0phoresis caused a depolarization that mimicked the amplitude and time course of the fEPSP. During a 20 Hz stimulation train, the fEPSP ran down while responses caused by a 20 Hz train of ionophoretic pulses of ACh did not. Data are mean i SEM (n=4). B. ATP applied by ionophoresis caused a depolarization that mimicked the fEPSP. Responses caused by a 20 Hz train of ionophoretic pulses of ATP did not run down while fEPSPs caused by a 20 Hz stimulus train did. Data are Mean 4: SEM (n=4). 110 Figure 27 A fEPSP ACh ionophoresis k '10 mV '7 20 ms 12 _ ACh ionophoresis § ' ’ ' 9 ' i i i i i §0.8- i ’ "=5 93 i fEPSP g 0.4- i i i E i s i i ‘r 0. 0 0.0 0.102 03 0.4 0.5 Time(s) fEPSP ATP ionophoresis Bk k 5mV 20 ms 1.2r . _ 3 ATP IonophoreSIs c I 8.0.8- 3”” ii (D “.3 g 0.4. i fEPSP "=5 E l I I g i ci’ 0.0 W 0.0 0.1 0.2 0.3 0.4 0.5 Time (s) 111 (0.5 mM) to block S-HTIA receptors, idazoxan (10 pM) to block (12 adrenergic receptors, and 8-cyclopentyl-theophylline (CPT, 10 uM) to block A1 adenosine receptors. None of the antagonists tested altered fEPSP rundown during a IO-Hz stimulus train (Figure 28). It is possible that antagonists tested above have not blocked the receptor(s) responsible for presynaptic inhibition leading to synaptic rundown. In order to comprehensively eliminate presynaptic inhibition as a mechanism of synaptic rundown, I attempted to block the intracellular signaling pathway coupled to presynaptic inhibitory receptors. Most inhibitory presynaptic receptors are G-protein coupled receptors, which link to the pertussis toxin (PTX) sensitive Gi/Go subsets of G-proteins (Zamponi, 2001). So PTX was used to block the signaling pathway to confirm the effects of these receptors on synaptic rundown. Blockade of the inhibitory postsynaptic potential (IPSP) from submucosal S neurons can be an indicator of the effectiveness of the PTX-treatment protocol. Previous work has shown that the IPSP is mediated by norepinephrine acting a2-adrenergic receptors which couple via a PTX-sensitive G-protein to activation of a potassium channel (Surprenant and North, 1988). I showed that IPSPs were blocked in PTX pretreated submucosal preparation while fEPSPs recorded from the same cells were unchanged (Figure 29). The same protocol was used to treat myenteric plexus preparations with PTX. In these tissues, it was found that fEPSPs still exhibited rundown that was similar to that observed under control conditions (Figure 29). Action potentials. Another possible reason that results in synaptic rundown is that action potentials are not able to follow stimulus train. In order to test the hypothesis that 112 Figure 28. Antagonists of presynaptic inhibitory receptors do not alter synaptic rundown Each figure shows the mean i SEM. of successive fEPSPs normalized to the first fEPSP in the train before and after drug treatment. There were no differences between the control and drug treatment values. Data are mean 2‘: SEM (n = 4—5). 113 Figure 28 l Adenosine A1 receptors 1.0- I . I Control 0 CPT (10 pM) 0.5- i I O. ‘ I a. i 3 °- 00 ’ Ell . . . . . v 0.0 0.2 0.4 0.6 0.8 1.0 8 T' .5 Ime(s) E E . 2 l (12 adrenergic receptors 1.0- I I Control i . ldazoxan(10 pM) 0.5- g g l i i g 0.0 - . - . . . . :J—l— 0.0 0.2 0.4 0.6 0.8 1.0 Time (s) 114 l Opioid receptors 1.0- I i I Control I Naloxone (10 pM) 0.5- i i a i . 2 . i 0.0 . . - r- . . . L4— 0.0 0.2 0.4 0.6 0.8 1.0 Time (s) l 5-HT1 A receptors 1.0« I g I Control 3 E . NAN-190(0.5 pM) 0.5- i E E i i i 0.0 0.2 0.4 0.6 0.8 1.0 Time(s) Figure 29. Pertussis toxin (PTX) does not affect fEPSP rundown A. PTX pretreatment abolished the IPSP but not the fEPSP in submucosal neurons. Downward deflections in the IPSP traces are voltage responses caused by hyperpolarizing current pulses. A decrease in the amplitude of these responses during the IPSP indicates a decrease in membrane resistance. B. Successive fEPSPs recorded from myenteric neurons with and without PTX pretreatment. There were no differences between values obtained in treated and untreated tissues at any time point. Data are mean :l: SEM (n=8). 115 Figure 29 A Control -60 mV PTX Treatment l/\ l I III] Control (n=6) 20 . - PTXITreatment (n=3) 3‘ 15 - E l 8 10 = l E O. E 5- < h 0 '-— _— fEPSP IPSP B > e 8 1.0 _ 100 ms 0. a) a- p 1:! g 0.5 - .‘E \ + (a!) i Ma 0.0 - . - - - . .‘9 0.4 0.6 0.8 1.0 Time (S) 116 axonal action potential failure is not a mechanism of fEPSP rundown, we made recordings from neurons in which the focal stimulating electrode was positioned on the axon of the neuron from which the fEPSP was recorded. This allowed us to make simultaneous recordings of axonally propagated action potentials and fEPSPs during a 10 Hz stimulus train. These experiments were done in 10 neurons and it was found that antidromic action potentials were maintained throughout the stimulus train while fEPSPs declined in amplitude during the same stimulus train (Figure 30). BK channels. The activity of large conductance calcium-activated potassium (BK) channels in nerve terminals can modulate neurotransmitter release (Robitaille and Charlton, 1992). Calcium entering the nerve terminal during the action potential opens BK channels. The Opening of BK channels shortens action potential duration and hyperpolarizes the nerve terminal; both effects can reduce transmitter release. I used iberiotoxin (IBTx 100 nM), a selective antagonist of BK channels (Wanner et al., 1999) to determine if activation of these channels contributed to fEPSP rundown. The effect of IBTx on somal action potential duration was studied in 12 S neurons and 5 AH neurons. The duration of somal action potentials triggered by an intracellular current pulse was broadened by IBTx in 8 of 12 S neurons; in these cells, the action potential duration at half amplitude increased from 1.3 i: 0.1 to 2.0 5: 0.2 ms (P<0.05) (Figure 31). IBTx did not change action potential duration in AH neurons; the duration at half amplitude was 2.0 i 0.08 and 2.1 i- 0.07 ms (P>0.05) before and after IBTx treatment, respectively (Figure 31). lBTx did not change the resting membrane potential or input resistance in any neuron tested. IBTx did not significantly change the amplitude of the first fEPSP in a 117 I IO-Hz train of stimulation but in 8 of 12 neurons IBTx attenuated PPD. In the absence of IBTx, the amplitude of the second fEPSP in a train was 66 i 7% of the first fEPSP while in the presence of IBTx this value increased to 92 i 6% (P<0.05) (Figure 32). Although IBTx reduced PPD, fEPSP rundown was not altered as the ratio of amplitudes of the first and 10th fEPSP was similar in the absence and presence of IBTx (Figure 32). Similarly, IBTx did not change the time constant of fEPSP rundown. The I value for rundown at 10 Hz before IBTx was 0.4 i 0.1 s and in the presence of IBTx, this value was 0.5 i 0.1 s (P>0.05) (Figure 32). 118 Figure 30. Action potentials do not fail during the train stimulation at 10 Hz A. Recordings of antidromic action potentials and fEPSPs elicited by a 10 Hz train of stimulation. The fEPSP declined in amplitude while the antidromic action potential followed each stimulus in the train. B. Examples of the antidromic action potential and fEPSPs were shown on an expanded time scale. 119 Figure 30 fEPSP lst 2nd 10th 20th 5 ms 120 Figure 31. Effects of IBTx on action potentials in S and AH neurons A. Action potentials triggered by an intracellular depolarizing current pulse were broadened by IBTx in S but not AH neurons. B. Pooled data from experiments similar to that shown in (A) indicate that the action potential duration at half maximum amplitude is lengthened by lBTx in S (n=8) but not AH neurons (n=5). * Indicates a significant difference from control (P<0.05). Data are mean :t SEM. 121 . i Figure 3 1 A -50 mV _J2o mV 5ms -50 mV B I: Control 2.5 — IBTx(100nM) 2.0 75 E, 1 5 a) 1.0 I: 0.5 0.0 S neurons AH neurons 122 Figure 32. Effects of IBTx on fEPSPs during a 10 Hz stimulus train A. Representative fEPSPs before and after IBTx treatment. B. The second fEPSP in the train of fEPSPs is not depressed after IBTx treatment but the 10th fEPSP decreased to the same level occurring in the absence of lBTx. Data are the ratio of the second fEPSP vs. the first fEPSP (Vsccond/Vfirst) in a stimulus pair and the ratio of the 10th fEPSP vs. the first fEPSP in the stimulus train. * Indicates significantly different from control (P< 0.05). (C) Frequency-dependent rundown of fEPSPs before and after lBTx treatment. lBTx did not alter the time course of fEPSP rundown. Data are mean :1: SEM (n=8). 123 Figure 32 Mu. mMNMLJmmv 0.1s .. NM Hwa B 12 30. n- 25: E 08 :1 Control A20_ 9! ' -IBTx(1OO nM) E . '8 :15~ g 310’ m0.4- LU . E .. . é 5- 0‘ ‘ i i ' i 2 16 2nd fEPSP 10tthPSP 0-0 0.4 0.8 1. . Time (s) 124 PRESYNAPTIC 5-HT4 RECEPTOR-MEDIATED REGULATION OF FAST SYNAPTIC TRANSMISSION 5—HT4 receptor agonist increases fEPSPs during trains of stimulation First, I tested the effect of S-HT4 receptor agonist on cholinergic and purinergic synaptic transmission induced by single electrical stimulation. Renzapride at 0.1 uM increased both cholinergic and purinergic fEPSPs when presence of purinergic blocker, PPADS and cholinergic blocker, mecamylamine, respectively (Figure 33). Then the effect of renzapride on the fEPSPs elicited by trains of stimulation was examined. Under control condition, fEPSPs declined to zero at about the tenth stimulation during a 20 pulses conditioning stimulation at 10 Hz. The first fEPSP amplitude was 17 i 2 mV and fEPSP amplitude at the 20th stimulus was 0.7 :1: 0.3 mV (n=9). In the presence of renzapride, the amplitude of the first fEPSP was 28 i 3 mV and this declined to a steady state level at the 20 stimulus of 6 i 2 mV (n=9). However, the time constants of decline were similar. They were 0.6 i 0.1 s and 0.6 i 0.1 s before and afier renzapride (Figure 34). S—HT4 receptor agonist accelerates recovery after rundown The effect of renzapride on time of recovery from synaptic rundown following a 10 Hz train of stimulation was also investigated. As described above, in these experiments, a train of 20 stimuli at 10 Hz caused fEPSPs to decline in amplitude to 0 mV at control condition. Recovery from rundown was tested by waiting for periods of 1 125 Figure 33. Prucalopride increases non—cholinergic and cholinergic fEPSPs recorded from myenteric neurons A. Representative fEPSP recorded in the absence of drug (control) and in the presence of mecamylamine (10 uM) to isolate the non-cholinergic component of the fEPSP. Subsequent addition of prucalopride (0.1 uM) caused an increase in the amplitude of the non-cholinergic fEPSP. B. The P2 receptor antagonist, PPADS (10 11M), was used to isolate the cholinergic component of the fEPSP in the same neuron. Subsequent addition of prucalopride increased the amplitude of the cholinergic component of the fEPSP. 126 Figure 33 A i 10 mV AN“ 20 ms ll . Control Mecamylamine Mecamylamine + Prucalopride 25- ’>‘ E, n. a) at H! Control Mecamylamine Mecamylamine + Prucalopride B i10mV 20 ms Control PPADS PPADS + Prucalopride Control PPADS PPADS + Prucalopride 127 Figure 34. Renzapride increases fEPSPs elicited by trains of stimulation A. Representative recordings of fEPSPs during trains of stimulation. Fast EPSPs declined to zero at the tenth stimulation (arrow) under control conditions. Renzapride increased the amplitude of all fEPSPs during the stimulus train. The fEPSPs did not decline to zero at the last stimulation (arrow). B. Pooled data show that renzapride increased fEPSPs (lower left) but did not alter the time course of synaptic rundown (lower right). Data were fit by a mono-exponential funtion. 128 Figure 34 fEPSP (mV) N 03 O O .3 O O Control 4 LL11 I 1 F l Renzapride ‘23:! 20 mv 1.0» I value (8) 0.0 0.5 ‘ 1.0 7 1.5 2.0 ' Com. Renzapride 129 [E A....' ‘52. 3. to 12 s before a single stimulus was used to elicit a fEPSP (Vmso; this amplitude was expressed as a fraction of the amplitude of the first EPSP in the preceding stimulus train (Vfim) (Figure 35). Renzapride accelerated fEPSP recovery rate from rundown. The recovery time constants as calculated by exponential fittings were 7 i 2 3 under control condition whereas in the presence of renzapride the recovery time constant was reduced to 1.6 i 0.2 s (P < 0.05). F orskolin mimics 5-HT4 receptor agonist effects 5-HT4 receptors couple to Gs-protein. Activation of S-HT4 receptors stimulate a cascade including activiting adenylate cyclase, cAMP and PKA. To investigate if S-HT4 receptor agonist increases fEPSP through stimulating adenylate cyclase, forskolin, an adenylate cyclase stimulator, was superfused into the Krebs’ solution. It was found that forskolin mimicked the effects of S-HT4 receptor agonists. Forskolin increased fEPSPs during 10 Hz train of electrical stimulation (Figure 36). Fast EPSPs still declined with forskolin treatment however the decline curves were parallel before and after forskolin treatment (Figure 36). As 5-HT4 receptor agonists did to fEPSP recovery, forskolin also reduced fEPSP recovery time from rundown (Figure 36). Since S-HT4 receptors couple to the Gs-protein and forskolin mimics 5-HT4 receptor agonists, effects of 5-HT4 receptors on rundown and recovery must be through activation of adenylate cyclase. H-89 abolishes the effects of 5-HT4 receptor agonist The function of adenylate cyclase is to catalyze coversion of ATP to cyclic AMP (CAMP) which activates PKA. PKA will then phosphorylate a number of targets 130 Figure 35. Renzapride accelerates recovery from fEPSP rundown A. Renzapride accelerated recovery rate after fEPSP rundown. V was the first first fEPSP in the train and Vtest was the fEPSP evoked by a test stimulus at various time intervals after the stimulus train. The recovery time course was fitted by monoexponential function. B. The recovery time constant was reduced by renzapride. 131 Figure 35 A Renzapride .0 o: .O on Vtest/V1 st P o o z 4 e £3 11072 Time (8) n=14 1: of recovery (Sec) Control Renzapride 132 Figure 36. Forskolin (FSK) mimics the effects of renzapride A. F SK increased the fEPSP amplitude during a stimulus train. The fEPSPs did not decline to zero at the last stimulation in presence of FSK. B. FSK accelerated recovery rate after fEPSP rundown. 133 > 30 — 25 — g 20 i . Control g 15 _ o FSK % ’ i % ‘6 _ ésiiiiiéiiii o~. . . . . °i‘?*ao~. 0.0 0.5 1.0 1.5 2.0 Time (8) B 1.2 - 0.9 - $0.6 - 277) >~ 0_3 . Control . FSK 0.0 ~ Time (S) 134 such as proteins and ion channels. H-89, a PKA blocker, was used to examined if activation of PKA is needed for S-HT4 receptor agonist effects. As shown in Figure 37, renzapride increased fEPSPs and accelerated the recovery rate. Co—application of H-89 and rezapride abolished the renzapride effects (Figure 37). This result suggested that 5- HT4 receptor-mediated facilitation of synaptic transmission depended on its downstream signaling molecule PKA. S-HT4 receptor agonists increase release probability from single varicosities The above studies have shown that activation of S-HT4 receptors facilitated synaptic transmission through activating adenylate cyclase and PKA. But how PKA increases fEPSPs was not known. Did it increase release probability from single release site? Those experiments were conducted in acutely isolated LMMP preparations in vitro. In these studies, synaptic transmission recorded was triggered by electrical stimulation on a bundle of nerve fibers. However, in the ENS, transmitters are released from varicosities along single nerve fibers. The mechanism that S-HT4 receptor agonists increase transmitter release in single varicosities cannot be obtained from such studies. To address this question, the “sniffer” patch (detector patch) technique was employed in cultured myenteric neurons. Outside-out patch clamp recordings were made from myenteric neurons maintained in primary culture. ACh and ATP-induced inward currents were recoded to make sure there are nAChRs and/or P2X receptors in the outside—out patches. Then this 135 Figure 37. H-89 abolishes the effects of renzapride on rundown and recovery rate after fEPSP rundown. A. Renzapride increased the amplitude of all fEPSPs during stimulus train (10 Hz). Co-application of renzapride and H-89 did not alter fEPSP amplitude. B. Renzapride accelerated recovery rate afier fEPSP rundown. The acceleration was abolished by addition of H-89. 136 Figure 37 > N .b fEPSP (mV) KS 5 05 1.2 0.9 rst test 0.3 0.0 fl T I 1 Control Renzapride Renzapride + H89 0. n v D / 1 . Control _ o Renzapride . Renzapride + H89 0 2 4 6 8 1 0 Time (s) 137 patch was used as a detector to measure transmitter released from single varicosities. The transmitter release was triggered by stimulating a single nerve fiber. The currents flow through the detector patch represented the activity of a single release site. Twenty electrical stimuli were given in an interval of eight seconds. Because transmitter release from a given varicosity is intermittent, only some of twenty stimuli can trigger transmitter release to induce currents that can be recorded in the detector patch. Under control conditions, 27% stimulation elicited transmitter release (Figure 38; Table 3). Renzapride increased the ratio to 52% (Figure 38; Table 3). Similarly, another 5—HT4 receptor agonist teagasrod increased release probability from 27% to 49% (Figure 38; Table 3). This result suggests that activation of 5-HT.; receptors increases neurotransmitter release probability and this effect leads to facilitation of synaptic transmission. I38 Figure 38. Activation of presynaptic 5-HT.. receptors increases release probability from single varicosities Representative recordings of single channel events evoked by electrical stimulation of a single nerve fiber (as arrows show) before and after S-HT4 receptor agonists renzapride (Figure 38a) or tegaserod (Figure 38b). Renzapride or tegaserod increased single channel open probability. 139 Figure 38b Tegaserod (0.15 nM) \ Control \W W \Wmm W W \W W \W 20 ms 141 Table 3. Effects of S-HT4 receptor agonists on release probability. Treatment of renzapride (A) or tegaserod (B) increased transmitter release probability evoked by electrical stimulation (see text for details). A. Failure events Successful events Control 73 27 Renzapride 48 52* n=5, P < 0.001 B. Failure events Successful events Control 58 22 Tegaserod 41 39* n=4; P < 0.005 142 DISCUSSION 143 P2X RECEPTOR SUBTYPES EXPRESSED IN MYENTERIC NEURONS In this section, identification of P2X receptor subtypes of myenteric neurons was accomplished in the small intestine of guinea pigs and mice. Guinea pigs have been used as the animal model for studying G1 physiology and pharmacology broadly because the gut is readily dissected, reasonably long and mechanically quiescent when studied in vitro. Immunohistochemical localization of P2X receptor subtypes has been done in whole mount preparations of guinea pig small intestine (Pools et al., 2002; Castelucci et al., 2002). Pharmacological properties of P2X receptors in myenteric neurons have also been characterized in cultured cells (Zhou and Galligan, 1996; Khakh et al., 2000). However, no data are available about the function of P2X receptors in acutely isolated LMMP preparations. It is important to identify P2X receptor subtypes so that a complete characterization of synaptic mechanisms in neural circuits can be accomplished. I attempted to characterize pharmacological properties of P2X receptors in myenteric neurons in guinea pig tissues using limited selective agents. As a supplement to these pharmacological studies, I also investigated P2X receptor subtypes in tissues from mice in which P2X; and P2X; genes have been deleted. The advantage of gene deletion would produce precise information. My data indicate that, in guinea pig small intestine myenteric S neurons express P2X; receptors whereas AH neurons express P2X; receptors. In mice small intestine, myenteric S neurons express P2X; homomers whereas AH neurons express P2X2/3 heteromers. The differences may be due to selective expression of P2X subunits by 144 different species. But I cannot exclude the possibility that lack of specific P2X receptor ligands limited the conclusion made in guinea pigs. Properties of murine myenteric neurons Murine myenteric neurons have similar electrophysiological properties to guinea pigs. In the mouse ileum, AH neurons had a broad action potential that was only partly reduced by TTX. The action potential has a prominent calcium component that accounts for the long duration and action potential shoulder in AH neurons from guinea-pig ileum and colon (Fumess et al. 1998; Nuragli et al. 2004), rat colon (Browning and Lees, 1996), pig ileum (Cornelissen et al. 2001), mouse colon (Furakawa et al. 1986) and human colon (Brookes et al. 1987). Therefore, the calcium shoulder is a general property of AH neurons in the gastrointestinal tract of a number of species. Although not tested in this study, the long duration and TTX-resistance of mouse ileal myenteric AH neurons suggest that there is also a calcium component to the action potential. The action potential in mouse ileal myenteric AH neurons was followed by an afterhyperpolarization that lasted more than 2 s. This property is also similar to that in AH neurons in the myenteric plexus of the guinea-pig ileum and colon (Hirst et al. 1974; Fumess et a1. 1998;) rat colon (Browning & Lees, 1996), pig ileum (Cornelissen et al. 2001), mouse colon (Furakawa et al. 1986) and human colon (Brookes et al. 1987). In guinea pig and pig intestine, fEPSPs can be evoked in some AH neurons (Liu et al. 1997; Comelissen et al. 2001) while sEPSPs can be evoked in nearly all AH neurons in the guinea-pig gastrointestinal tract (Fumess et al. 1998). I found that single electrical stimuli did not elicit synaptic responses in mouse ileal AH neurons. As I recorded from only a small 145 number of AH neurons, it is possible that I did not sample those neurons receiving fast synaptic input or that I did not stimulate the interganglionic connectives providing the fast synaptic inputs to the neurons that I did sample. However, I was able to elicit sEPSPs in all AH neurons studied. Slow synaptic excitation is the predominant synaptic mechanism in guinea pig AH neurons (Kunze et al. 1993). Taken together, these data indicate that the properties of AH neurons in the mouse ileal myenteric plexus are similar to those identified in the gastrointestinal tract of other species. S type enteric neurons have an action potential that is completely blocked by TTX and all S neurons exhibit fEPSPs following single stimuli applied to the interganglionic connectives in the myenteric plexus (Hirst et al. 1974; Bomstein et al. 1994; Galligan et al. 2000). In the mouse myenteric plexus, I identified neurons that received fast excitatory synaptic inputs and the action potential in these neurons was completely blocked by TTX. Therefore, I concluded that these were S neurons. I also found that the fEPSP in S neurons was only partly blocked by mecamylamine, a nicotinic cholinergic receptor antagonist, and that fEPSPs were also inhibited by PPADS. These data indicate that both acetylcholine and ATP are fast synaptic transmitters in the mouse small intestine as they are in the guinea-pig intestine (LePard et al. 1997; Johnson et al. 1999) but there may be additional fast neurotransmitters including 5-HT (Galligan et al. 2000). Murine myenteric AH neurons were encountered infrequently using a random impalement strategy with intracellular microelectrodes. This observation is similar to that reported previously in a study of myenteric neurons in the mouse colon (Furukawa et al. 1986). If AH neurons are sensory neurons in the mouse intestine, as they are in the guinea pig intestine (Fumess et al. 1998), then normal intestinal reflex behaviors may require 146 only a small number of sensory neurons in the mouse. Alternatively, non-AH neurons may serve as sensory neurons in the mouse intestine. A previous study, for example, has shown S neurons in guinea-pig distal colon serve as mechanosensory neurons that are sensitive to stretch (Spencer and Smith, 2004). P2X receptor-mediated responses in AH neurons ATP caused a rapidly developing depolarization in all murine myenteric AH neurons. This depolarization was associated with a decrease in input resistance and was blocked by PPADS. These data are consistent with this response being mediated by either a P2X or P2Y. receptor. a, B-mATP is an agonist at homomeric P2X], homomeric P2X3 or heteromeric P2X2/3 receptors, but it does not activate P2Y receptors (North and Surprenant, 2000). I found that ATP caused a depolarization of AH neurons in P2X3+H and in P2X3’/' tissues, while or, B-mATP depolarized AH neurons in P2X3+l+ but not in P2X3’l’ tissues. These data indicate that P2X3 subunits contribute to the P2X receptor expressed by murine AH neurons. However, sensory neurons and some sympathetic neurons also express P2X2/3 heteromeric receptors (Lewis et a1. 1995; Zhong et al. 2000), and the loss of response to on, B-mATP in P2X3’l' mice could result from the loss of P2X2/3 heteromeric receptors in murine AH neurons. Other subunits must also be present in order for an ATP-induced depolarization to persist in AH neurons from P2X3'I' mice. P2X; subunits remaining in P2X3’/' mice could account for this ATP sensitivity. a, B- mATP caused similar amplitude depolarizations in AH neurons from PZXZH+ and P2X2’/' mice. As a, B-mATP activates P2X receptors containing P2X. or P2X3 subunits (North and Surprenant, 2000), it is possible that AH neurons express P2X receptors composed of 147 P2Xl or/and P2X3 subunits. However, my work has also shown that a, B-mATP does not depolarize AH neurons in the myenteric plexus of P2X;"' mice while ATP depolarized the same neurons. Therefore, the P2X receptors in murine AH neurons are likely to be P2X2/3 heteromers. My suggestion that the P2X; subunits are expressed in AH neurons is also supported by recent immunohistochemical studies in guinea-pig gastrointestinal tissues. These studies showed that P2X; subunits are localized to neurons that also contain the calcium binding protein, calbindin (Castelucci et al. 2002). Calbindin is a protein marker for most AH neurons in the guinea—pig GI tract (Fumess et al. 1998). However, it has also been shown that P2X; subunits are not found in calbindin-containing neurons in the guinea-pig GI tract but that these subunits are localized to calretinin- and NOS-containing cells (Nassauw et al. 2002; Poole et al. 2002). Calretinin and NOS are contained in S type neurons and are markers for intemeurons and motoneurons (Brookes, 2001). It is possible that receptor expression and subunit composition of P2X receptors in subsets of neurons are different in the mouse and guinea pig small intestine. A detailed study of the neurochemical content and receptor expression of subsets of murine enteric neurons is needed to address this issue. P2X receptors in AH neurons do not mediate fEPSPs as fEPSPs are not recorded from AH neurons. It may be due to no purinergic inputs to AH neurons. The functions of AH neuron somal P2X receptors remain to be addressed. Fast synaptic transmission and P2X receptor-mediated responses in S neurons 148 I hypothesized that a P2X receptor contributes to the fEPSP recorded from S neurons in the mouse small intestine. This hypothesis was based on the finding that fEPSPs were only partly inhibited by mecamylamine, indicating that a neurotransmitter in addition to acetylcholine mediates the fEPSP. As fEPSPs were also reduced by PPADS, the additional neurotransmitter in the mouse myenteric plexus is likely to be ATP. It is unlikely that P2X3 receptor subunits contribute to the P2X receptor expressed by S neurons as the properties of fEPSPs in tissues from P2X3H+ and P2X3'l' mice were +/+ -/- - and P2X3 mice were identical. Furthermore, S neurons in tissues from P2X3 depolarized by ATP but not by a, B-mATP. a, B-mATP insensitivity suggests that the P2X receptor in S neurons is a P2X; homomeric receptor (Zhou and Galligan, 1996). The fEPSPs recorded from S neurons in tissues from P2X?+ mice were partly inhibited by the nicotinic cholinergic receptor antagonist mecamylamine and by the P2 receptor antagonist PPADS. These results indicate that ACh and ATP mediate fast synaptic responses in the murine myenteric plexus as also occurs in the guinea pig intestine (Galligan and Bertand, 1994; LePard et al. 1997; Johnson et al. 1999) and colon (Nurgali et al. 2003). However, the fEPSP in P2X2'/' mice was blocked by mecamylamine while PPADS had no effect on fEPSPs in P2X2'/‘ tissues. These data indicate that P2X2 subunits are a critical component of the P2X receptor expressed by S neurons. Although P2X2 subunits mediate the purinergic component of fEPSPs, the average fEPSP amplitudes recorded from S neurons in P2X2'/' and P2X2‘i/+ tissues were not different. This result may be related to the findings from previous studies that showed a functional interaction between P2X and nicotinic acetylcholine receptors (Nakazawa, 1994; Zhou and Galligan, 1998; Searl et al. 1998; Barajas-Lopez et al. 1998; Khakh et al. 2000). In these studies it 149 was found that simultaneous activation of nicotinic and P2X receptors produces a response that is smaller in amplitude than the predicted sum of responses caused by individual activation of each receptor. It was concluded that there is a functional link between P2X and nicotinic receptors that results in cross-inhibition of responses mediated by these receptors (Nakazawa, 1994; Zhou and Galligan, 1998; Searl et a1. 1998; Barajas-Lopez et al. 1998; Khakh et a1. 2001). Deletion of the P2X; subunit gene would remove cross-inhibition and synaptic responses mediated by nicotinic receptors in P2X2'l' mice would be larger in amplitude than those occurring in neurons co-expressing the P2X; subunit. Lack of P2X2- nicotinic receptor cross-inhibition could account for the maintained fEPSP amplitude in P2X2'/' tissues. S neurons in tissues from PZXZH+ mice were depolarized by ATP but not by or, B- mATP. This result suggests that S neurons express P2X; homomeric receptors as 01,13- mATP does not activate P2X; homomers, but does activate P2X3 homomeric and P2X2/3 heteromeric receptors (Lewis et al. 1995; North and Surprenant, 2000). In addition, ATP failed to elicit a depolarization in S neurons from P2X2'l' mice. Based on these results, I conclude that the P2X receptor mediating fEPSPs in murine S neurons is a P2X; homomeric receptor. In the guinea pig intestine, P2X; subunit immunoreactivity has been localized to NOS-containing neurons (Castelucci et al. 2002). NOS—containing neurons are inhibitory motomeurons and descending intemeurons, and both of these classes of neuron would have S-type electrophysiological properties (Brookes, 2001). If the chemical coding of murine myenteric neurons is similar to that in the guinea pig intestine, these data would indicate that inhibitory motomeurons (Johnson et al. 1999) and some 150 descending intemeurons (LePard & Galligan, 1999; Bian et al. 2000; Monro et al. 2002) receive fast excitatory synaptic input mediated in part by P2X2 homomeric receptors. P2X subtypes in myenteric neurons in guinea pig small intestine Extensive immunohistochemical studies have attempted to identify P2X subunit compositions in guinea pig myenteric neurons (Poole et al., 2002; Castelucci et al., 2002). In these studies, it has been shown that P2X; subunits are localized to inhibitory motor neurons and AH neurons in myenteric plexus. P2X; subunits are localized to inhibitory motor neurons and S neurons. They are not found in AH neurons. In inhibitory motor neurons, which are NOS immunoreactive, P2X receptors are P2X2/3 heteromers. The pharmacological studies in other paper have found there are or, B-mATP-sensitive P2X receptor subtypes (P2X. or P2X3) contributing fEPSP in S neurons of guinea pig intestine (LePard et al. 1997). However patch-clamp recordings from myenteric neurons maintained in primary culture indicate that only in a small subset (10 %) of myenteric neurons, or, B-mATP caused a rapidly developing and desensitizing inward current, suggesting that P2X receptors contain P2X] or P2X3 subunits (Zhou & Galligan, 1996). In present study, all S neurons that were sensitive to ATP were depolarized by a, [3- mATP but only 17% AH neurons were sensitive to a, B-mATP. A selective antagonist for P2X] or P2X3 receptors, TNP-ATP reduced ATP-induced depolarization as well as reducing fEPSP amplitude in S neurons. These results support immunohistochemical identification and studies in acute isolated LMMPs in guinea pigs. However, these results are not consistent with guinea pig myenteric neurons maintained in primary culture and murine myenteric neurons. Most of neurons in primary culture do not desensitize to a, B- 151 mATP may be due to the alteration of P2X3 receptors in cultural condition. The difference between myenteric neurons in guinea pigs and mice may be due to the differentiation expression of P2X receptor between these two species; or myenteric S type neurons in mice and in guinea pigs may play different functions. These data in the present study suggest that, in guinea pig ileum, P2X receptors in myenteric S neurons contained P2X3 subunits while P2X receptors in AH neurons were composed of few P2X3 subunits. 152 P2X RECEPTOR-MEDIATED AFTERHYPERPOLARIZATION IN GUINEA PIG MYENTERIC S N EURON S Ionic basis of afterhypolarization An important finding in the present study was that the depolarization induced by a, B-mATP was followed by a hyperpolarization in 55 percent of tested S neurons. The hyperpolarization was mediated by activation of potassium conductance because the reversal potential of its underlying outward current was equal to K+ equilibrium potential. The K+ conductance was gated by calcium because reduction of extracellular Ca2+ concentration greatly reduced the hyperpolarization. The influx of Ca2+ that activates K+ conductance was through P2X receptors as PPADS blocked depolarization therefore abolished hyperpolarization consequentially. The proposed model for P2X receptor- mediated hyperpolarization is shown as Figure 39. It has been known that P2X receptors have a predominant Ca2+ permeability that equals to or greater than those of the 100 times more Na+ (Bumashev, 1998). We also noticed that the amplitudes of hyperpolarization and depolarization were positively correlated although the other cations are involved. It may still be arguable that Ca2+ could move into the neurons through voltage-gated channels. It is unlikely the case in the present study since we recorded the KCa-mediated outward currents in single electrode voltage-clamp recordings. The influx of Ca2+ can have broad physiological effects by acting at various downstream targets, which are Caz+-dependent K+ channels (K5,) in the present study. 153 Figure 39. A proposed model for P2X-mediated afterhyperpolarization in S neurons P2X3 subunit-containing receptors and calcium-dependent potassium channels are expressed in myenteric S neurons in guinea pig small intestine. Specifically, they are located closely within a micro-domain (shown in dashed box). Activation of P2X3 receptors by 0L,B-mATP induces calcium influx. The calcium ions entry activates KCa that allows potassium flowing out of the cell, which results in afterhyperpolarization. 154 Figure 39 P2X3 Receptor 155 I attempted to identify the subtypes of K0, channels that mediate hyperpolarization in S neurons. Large conductance KCa (BK) and small conductance KCa (SK) have been described in the nervous system (Vergara et al., 1998; Faber and Sah, 2003; Stocker, 2004). Intermediate conductance Keg (1K) has been identified earlier in non—neural peripheral tissues, particularly smooth muscle tissues (V ergara et al., 1998). Recently [K is also documented in the enteric nervous system (Fumess et al., 2004). Further studies have shown that IK is expressed in myenteric AH neurons and mediates action potential aflerhyperpolarization in AH neurons (Vogalis et al., 2002; Neylon et al., 2004). 1K is activated when Ca2+ flows into the neurons during action potential firing, which mediates the long lasting action potential aflerhyperpolarization (Hirst et al., 1974; Fumess et al., 1998; Vogalis et al., 2001). The selective toxins iberiori toxin (IBTx), apamin and Clotrimazole against BK, SK and 1K were utilized to identify Kca subtype in S neurons (Vergara et al., 1998). lBTx broadened the duration of action potential by blocking BK in S neurons (Figure 31). Clotrimazole blocked action potential afterhyperpolarization in AH neurons (Figure 21B), which indicated the effectiveness of Clotrimazole on blocking lK. However none of these toxins affected P2X receptor- mediated afterhyperpolarization. This result suggests that KCa in S neurons does not belong to any of these three major types. Present study provides evidence that KCa channels are expressed in S neurons. KCa should mediate an afterhyperpolarization in S neurons as it does in AH neurons. However action potentials are not followed by long lasting hyperpolarization in S neurons. There is no action potential afterhyperpolarization in S type neurons because there is no Ca2+ influx during the action potential (Hirst et al., 1974), and not because there are no KCa 156 channels in S neurons. Ca2+ moves into the neurons during fEPSPs and a hyperpolarization should be observed after flSPSP. However a single purinergic fEPSP evoked by electrical stimulation on presynaptic nerve fibers was not followed by an after- hyperpolarization. One explanation would be that the amount of Ca2+ influx through P2X receptors during fEPSP is not sufficient to activate Kca channels. In addition, fast EPSPs are mediated by P2X3 receptors in most of myenteric S neurons (see discussion above). Ca2+ permeability of P2X receptor is dependent on different subtypes (Egan and Khakh, 2004). P2X; receptors have relative low permeability comparing with the other P2X subtypes. The other possibility would be that the P2X3 receptors expressed at the synaptic area do not couple to KCa channels. For example, immunohistochemistry staining indicates that only 20% Dogiel type 1 neurons express IK channel immunoreactivity and it is only seen in cell soma (Neylon et al., 2004). Functional significance of AHP in myenteric S neurons Another interesting observation was that only approximate half of tested S neurons displayed the agonist-induced afler-hyperpolarization. I proposed that ligand- gated hyperpolarization might be associated with a specific subset of S neurons that have the same function. S type myenteric neurons are motor or intemeurons (Fumess, 2000). This is based on the studies using combined electrophysiological and immunohistochemical methods (Fumess, 2000). Purinergic synaptic transmission contributes to the descending and ascending excitatory nervous reflex pathway as well as inhibitory descending pathway (Spencer et al., 2000; Fumess and Sanger, 2002). P2X receptors coupled to KCa may be a mechanism by which the neuronal firing rate is 157 controlled. 1 tested if this group of neurons belongs to a specific functional category of purinergic S neurons. I examined the co-localization of these neurons with NOS neurons, which are inhibitory motor neurons. However, my results did not show any association of NOS-ir neurons with the neurons that displayed a,B-mATP-induced ligand-gated hyperpolarization. A previous study has documented that a depolarization induced by activation of nicotinic AChRs is followed by a hyperpolarization in myenteric S neurons (Tokimasa et al., 1982). The underlying mechanism is similar, which is due to calcium entry and subsequent opening of Kc... Nicotinic AChRs and P2X receptors are both ligand-gated cation channels, which are permeable to non-selective cations including Ca”. They are the two primary ligand-gated ion channels in myenteric S neurons and have similar functions in mediating fast synaptic transmission. The similar hyperpolarization they induced may have similar physiological function. 158 RUNDOWN OF fEPSPS DURING BURSTS OF NEURONAL ACTIVITY In this section, fast synaptic transmission during trains of electrical stimulation was investigated. I found that fEPSPs declined during stimulus train and there was no contribution of desensitization of nAChRs and P2X receptors. Action potential failure, presynaptic inhibitory receptors or BK channels did not account for synaptic rundown. I proposed that synaptic rundown is due to depletion of vesicles in RP. Rundown of fEPSPs In isolated segments of guinea-pig ileum, it has been shown that mechanical stimulation of mucosal villi generates short bursts of fEPSPs that occur in the frequency range of 15—40 Hz (Bomstein et al., 1991). In similar experiments, it was shown that if the mechanical stimulus is applied at intervals of less than 10 s, the amplitude of fEPSPs declined during repetitive stimulation (Smith et al., 1992). Chemical stimulation (low pH or S-HT application) of the mucosa also elicits short bursts of fEPSPs (average of 14 fEPSPs per burst) occurring at frequencies between 5 and 20 Hz (Bertrand et al., 1997). I studied the dynamics of fast synaptic excitation during short trains of electrical stimulation of presynaptic nerve fibers. The frequencies of the stimulus trains that I used mimicked the frequencies of fEPSPs evoked by the physiological stimuli described above. In most neurons, these trains of electrical stimulation elicited fEPSPs that declined in amplitude in a frequency dependent manner. The conclusions are based on data obtained in S neurons, as these cells received fast excitatory synaptic input, and the action potentials were not followed a long-lasting (>1 5) action potential afterhyperpolarization. 159 My data differ from those obtained in studies of fast synaptic input to AH type neurons in the pig intestine where it was found that fEPSPs did not rundown when evoked by trains of stimulation at frequencies up to 20 Hz (Cornelissen et al., 2001). Similarly, in the rat colon, fast synaptic input to most myenteric S neurons is maintained during lO-Hz trains of stimulation (Browning and Lees, 1996). These observations suggest that there are species or tissue differences in the mechanisms regulating the availability and release of stores of fast synaptic transmitters in the myenteric plexus. Receptor desensitization did not contribute to synaptic rundown ACh acting at nAChRs and ATP acting at P2X receptors mediate most fEPSPs in myenteric S neurons in the guinea pig ileum (Galligan et al., 2000). Previous work showed that nAChRs expressed by guinea-pig myenteric neurons desensitize faster than P2X receptors expressed by the same cells (Zhou and Galligan, 1998; Brown and Galligan, 2003). Based on this difference in desensitization rate, I proposed that nAChRs and P2X receptors would make a differential contribution to the fEPSPs during trains of stimulation. My hypothesis was that the nicotinic component would desensitize early in a train of stimulation while P2X receptors would maintain synaptic transmission as the stimulus trained continued. However, the data indicate that this does not happen, as selective blockade of nAChRs or P2X receptors did not alter the rate of fEPSP rundown during a 10 Hz train of stimulation. Furthermore, when ACh and ATP were applied by trains of ionophoretic pulses to mimic fEPSPs during same stimulus train, the ionophoretic depolarizations did not decline in amplitude. These data indicate that little desensitization occurs during the brief occupation of nAChRs or P2X receptors by either 160 synaptically released or ionophoretically applied agonist. It was also found that mecamylamine or PPADS applied individually each reduced fEPSPs by about 35% while coapplication of these two blockers reduced the fEPSP by 90%. This result is consistent with the previous finding that P2X receptors and nAChRs are linked in a mutually inhibitory manner in guinea-pig ileum myenteric neurons (Zhou and Galligan, 1998; Khakh et al., 2000). That is simultaneous activation of the two receptors produces a response that is smaller than the predicted sum of responses caused by individual activation of the two receptors. Blocking either the nAChR or the P2X receptor removes cross-inhibition allowing a larger amplitude response at the unblocked receptor. Rundown is not caused by presynaptic inhibition Presynaptic inhibition is a predominant mechanism for modulating synaptic transmission in the ENS. a2—Adrenoceptors are localized to cholinergic neurons and these receptors mediate inhibition of acetylcholine release onto myenteric S neurons (Stebbing et al., 2001; Scheibner et al., 2002). Opioid receptors are expressed on cell bodies and processes of myenteric neurons and activation of opioid receptors inhibits the electrically evoked release of acetylcholine (Stemini, 2001). Muscarinic M2 receptors also mediate presynaptic inhibition of fast synaptic transmission in the myenteric plexus (North et al., 1985). Adenosine A1 (Christofi and Wood, 1993) and S-HTIA (Pan and Galligan, 1994) receptors also couple to presynaptic inhibition of fast synaptic excitation in the ENS. Acetylcholine, norepinephrine, enkephalins and 5-HT-containing nerve fibers are present throughout the myenteric plexus (Costa and Fumess, 1984; Costa et al., 1996; Brookes, 2001) and trains of nerve stimulation can release all of the substances. 161 Adenosine is a breakdown product of neurally released ATP (Westfall et al., 2002) and it will accumulate in the myenteric plexus during trains of nerve stimulation. Therefore, fast synaptic rundown could be due to presynaptic inhibition caused by the accumulation of endogenous agonists for presynaptic inhibitory receptors. However, the data show that fast synaptic rundown persists in the presence of muscarinic, a2 adrenergic, opioid, Al adenosine and 5-HT1A receptor antagonists. Based on these data, I conclude that presynaptic inhibition mediated by one or more of these receptors are not responsible for fEPSP rundown. However, although I tested the contribution of activation of several presynaptic receptors to fEPSP rundown, there could be many other inhibitory presynaptic receptors localized to myenteric nerve terminals. Inhibitory presynaptic receptors in the ENS belong to the family of G-protein coupled receptors which link to PTX-sensitive mechanisms and inhibition of neurotransmitter release (Zamponi, 2001). In order to eliminate presynaptic inhibition mediated by G-protein coupled receptors, I studied fast synaptic rundown in myenteric plexus preparations that had been treated with PTX to inactivate the Gi/Go class of G-proteins. Fast synaptic rundown was identical in control and in PTX-treated tissues indicating that presynaptic inhibition does not account for fast synaptic rundown. Although PTX-treatment did not alter synaptic rundown, the treatment protocol was effective in blocking the target G-proteins. This conclusion is based on the data showing that the a2-adrenoceptormediated IPSP recorded from submucosal S neurons was blocked in preparations that had been treated using our PTX incubation protocol. IPSPs in submucosal neurons are mediated by a2-adrenoceptors that couple directly to a K+ channel via a PTX-sensitive G-protein (Surprenant and North, 1988) I62 Changes in presynaptic action potentials do not contribute synaptic rundown Frequency-dependent decreases in nerve terminal Ca2+ influx could result in synaptic rundown. Decreased nerve terminal Ca2+ entry could be due to axonal action potential failure during high frequency nerve stimulation (Cunnane and Stjame, 1984). I made recordings from neurons in which fEPSPs and antidromically propagated action potentials were recorded simultaneously. These experiments showed that action potentials followed each stimulus in a 10 Hz train while fEPSPs declined in amplitude. These data indicate that action potential propagation into the nerve terminal is maintained throughout a train of stimuli and action potential failure does not contribute to fEPSP rundown. Frequency-dependent shortening of action potential duration could also contribute to synaptic rundown. This could occur either through Ca2+ channel inactivation or though a Cay-dependent activation of K+ channels which would accelerate action potential repolarization (Franciolini et al., 2001). Inactivation of Ca2+ channels in myenteric neurons occurs in the time range of 0.1— l s (Bian et al., 2004). As an individual action potential duration is 2 ms or less, it is unlikely that Ca2+ channel inactivation makes a significant contribution to fEPSP rundown in myenteric neurons. However, in the absence of direct data about calcium channel dynamics in the nerve terminal, I cannot categorically rule out a contribution of Ca2+ channel inactivation to fEPSP rundown. BK channels are a class of Ca2+-activated K+ channels expressed in nerve terminals and these channels function to limit nerve terminal action potential duration and Ca2+ entry (Robitaille and Charlton, 1992). IBTx, a BK channel blocker (Wanner et al., 1999), did not alter the time course of fEPSP rundown indicating that BK channel-dependent shortening of action potential duration do not contribute to fEPSP 163 rundown during trains of stimulation. However, I did find the IBTx inhibited PPD in a subset of neurons as the decline in amplitude of the second fEPSP in the presence of lBTx was less than that occurring in the absence of the toxin. This result suggests that BK channel activation contributes to PPD during the first two fEPSPs in a train of stimulation but that other processes mediate fEPSP rundown as the train duration lengthens. IBTx increased action potential duration in a subset of S neurons. These data indicate that BK channels regulate action potential duration in some S neurons and that Ca2+ must enter these neurons during a single action potential. Direct measurements have shown that short trains of action potentials can cause measurable increases in intracellular Ca2+ in some myenteric S neurons; N-type Ca2+ channels mediate this response (Shuttleworth and Smith, 1999). Although single action potentials did not evoke a measurable Ca2+ increase, changes caused by a single spike may be under the detection limits of the assay or there may be local increases in Ca2+ that were not detected. These local increases in Ca2+ could be restricted to sights near the BK channel as previous studies in hippocampal neurons have shown that there is a close spatial and functional relationship between N-type Ca2+ channels and BK channels (Marrion and Tavalin, 1998). A similar relationship may exist in some myenteric S neurons. This relationship is specific for some S neurons as IBTx did not alter the action potential in AH neurons. This result is consistent with previously published work showing that IBTx does not alter action potential duration in AH neurons (Kunze et al., 1994). 164 Depletion of transmitter accounts for synaptic rundown When nerve fibers were stimulated at a lower frequency (0.5 Hz), fEPSP amplitude declined by 50% after the first stimulus but was maintained for the remainder of the stimulus train. Higher frequencies of stimulation resulted in complete rundown of the fEPSP. There are two models that could explain these data. Firstly, there may be two separate neurotransmitter pools in myenteric nerve terminals. A readily releasable pool (RRP) is available for exocytosis during the first action potential in a train and at low stimulation frequencies; a reserve pool will partly replenish the RRP prior to subsequent stimuli (Rosenmund and Stevens, 1996; Pyle et al., 2000; Sudhof, 2000; von Gersdorff and Borst, 2002; Zucker and Regehr, 2002). The release probability of the RRP is very high as a single action potential can deplete this pool. In addition, we never observed PPF whose occurrence is most prominent at low probability synapses (Thomson, 2000; Zucker and Regehr, 2002). At stimulation frequencies >5 Hz, fEPSPs decline in amplitude because transmitter release rate exceeds the RRP refilling rate (Stevens and Tsujimoto, 1995; Rosenmund and Stevens, 1996; Dobrunz and Stevens, 1997; Lin et al., 2001). In myenteric nerve terminals, the refilling rate of the RRP is relatively slow (7 s) and the slow filling rate will limit the duration and frequency at which bursts of fEPSPs can occur. The mechanism by which vesicles are replenished after short-term depletion may vary among synapses with different recovery time constants. For example, the t for recovery is 5 s at synapses between hippocampal CA3 and CA1 pyramidal neurons (Debanne et al., 1996), 3 s at synapses between cerebellar granule and Purkinje cells, 4 s at the calyx of Held in the brain stem (von Gersdorff et al., 1997), and 2.8 s at synapses in the rat superior cervical ganglion (Lin et al., 2001). The recovery rate depends on 165 trafficking between vesicle pools and the mechanisms of vesicle movement at different synapses may be specialized to meet the particular demands of those synapses. Vesicle mobilization from a reserve pool can be a principal mechanism for synaptic recovery after rundown in the myenteric plexus. Mobilization from the reserve pool is relatively slow (Pyle et al., 2000) and this would account for synaptic rundown during trains of stimulation >5 Hz. A second model that would account for our data involves vesicle recycling (Sudhof, 2004). In this scheme, the RRP is released by the first action potential in a train. At low frequencies of stimulation (0.5 Hz), some of these vesicles recycle and are then available for release prior to the next action potential. At stimulation frequencies >5 Hz, the release rate exceeds the recycling rate and the RRP becomes depleted. Rapid vesicle recycling (1: