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 5/08 K:lProj/Aoc&PresIClRC/DateDue.indd MECHANISMS OF R-TYPE CALCIUM CHANNEL REGULATION IN THE ENTERIC NERVOUS SYSTEM . By Vinogran Naidoo A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Neuroscience 2009 ABSTRACT MECHANISMS OF R-TYPE CALCIUM CHANNEL REGULATION IN THE ENTERIC NERVOUS SYSTEM By Vinogran Naidoo The enteric nervous system (ENS) regulates gastrointestinal function, and calcium channels are crucial to ENS function. There are distinct classes of calcium currents. L-type currents require a strong depolarization for activation, are long lasting, and are the main calcium currents recorded in muscle and endocrine cells. P/Q-type, N-type, and R-type currents, also require strong depolarizations for activation. They are expressed primarily in neurons, where they initiate neurotransmission at most fast synapses. T-type calcium currents are activated by weak depolarizations and found in neurons and cardiac myocytes. R-type calcium currents were previously recorded from guinea-pig myenteric neurons suggesting that R-type calcium channels may be important in the ENS. The aim of my dissertation was to determine their functional significance in the ENS. I investigated the role of R-type channels in the myenteric plexus at four different levels: 1) the cellular level by mapping 01E- subunit immunoreactivity in various regions of the GI tract; 2) intracellular recordings from myenteric neurons to determine whether R-type channels mediate slow synaptic excitation in intrinsic primary afferent (AH-type) neurons, and/or whether calcium entry via non-R-type channels mediate fast synaptic excitation in interneurons and motorneurons (S-type); 3) functional studies of R- type channel-mediated neuromuscular transmission because excitatory and inhibitory motor express a1E immunoreactivity; and 4) physiologically integrative level by analyzing the effects of R-type calcium channel antagonists on peristalsis. My studies revealed novel mechanisms of R-type calcium regulation in the ENS. 1) R-type channel immunoreactivity was found in intrinsic primary afferent nerve endings, in nerve terminals of ascending interneurons, and on excitatory and inhibitory motorneurons that supply the muscle layers. 2. The R- type calcium channel antagonist NiCl2 inhibited the calcium shoulder and the amplitude of the late afterhyperpolarization in AH neurons. NiCI2 also acted at nerve terminals of fully cholinergic ascending interneurons to inhibit fast ganglionic neurotransmission. 3) The R-type channel couples to nitric oxide synthesis and release from nerve terminals of inhibitory motorneurons. 4) NiClz applied to isolated ileal segments in vitro impaired peristalsis. Taken together, these results suggest an important role for R-type channels as mediators of neurotransmitter release or as transducers of other synaptic events in myenteric neurons. Success with a selective activator of chloride channels that relieves chronic idiopathic constipation, suggests that treatments with ion channel agonists or antagonists may help alleviate motility disorders. The R-type calcium channel may therefore be a potential drug target in combating motility disorders. ACKNOWLEDGEMENTS I wish to first offer my sincerest thanks to Dr. Jim Galligan. Dr. Galligan has been the best mentor I could have ever hoped for. Thank you Dr. Galligan for your immense support, kindness, patience, guidance, advice, dedication, and for keeping faith in my abilities. Thank you also for accepting me into your laboratory, and for supporting me financially during the past 3 years. I would have never got even close to this point in my career without your help, and I am proud to have trained under you. I wish to also thank my guidance committee for the time they took out of their busy schedules to serve on my committee. Thank you Drs. David Kreulen, Stephen Schneider, Ke Dong and Hongbing Wang for your helpful discussions and advice. I truly appreciate your efforts. I also need to thank Dr. Stephen Schneider for the synaptophysin antibody. Thank you also to the International Institute of Education (HE) and the Fulbright Scholarship Board for giving me this unique opportunity to study for my doctorate in the United States. You are a testament that dreams do come true. Thank you also to the Nelson Mandela Fellowship Fund and Michigan State University for additional financial support. To the Graduate School at MSU, and the Neuroscience Program, thank you as well for funding my research. I certainly cannot forget my friends and colleagues in both Galligan labs. Thank you Xiaochun, Hui Wang, Dima, Hong, Hua, and Emmy for your friendship, support and laughter, through the best and the toughest of times. It’s iv been a tremendous five years with you all, and I will never forget you. Eileen ROdriguez contributed to the Neuromuscular transmission studies presented in my dissertation, and I thank her for that great collaboration. To my fellow students in the Neuroscience Program, thank you for your collegiality and friendship. Last but not least, to my wonderful family back in South Africa. It’s been a long 5 years since I left home. You’ve been the sources of my inspiration and my guiding light, even though you all are thousands of miles away. Thank you for your support, love and understanding. I will be home soon. To my brother Seelendran, this dissertation is for you. TABLE OF CONTENTS LIST OF TABLES .............................................................................................. viii LIST OF FIGURES ............................................................................................. ix CHAPTER 1 General Introduction ..................................................................... 1 The Nervous System ........................................................................................ 2 Structure and function of a neuron ................................................................... 3 Neurotransmitters, neuropeptides and synaptic transmission .......................... 4 Classical neurotransmitters ........................................................................... 5 Neuropeptides ............................................................................................... 6 The Autonomic Nervous System ...................................................................... 7 Innervation of the gastrointestinal tract ............................................................. 8 Extrinsic innervation of the ENS .................................................................... 8 lntrinisic innervation: the ENS ..................................................................... 11 Electrophysiological characteristics of myenteric neurons .............................. 18 Neurochemical coding and synaptic transmission in the ENS ........................ 20 Slow synaptic transmission in the ENS ....................................................... 20 Fast synaptic transmission in the ENS ........................................................ 23 Neuromuscular transmission in the ENS ........................................................ 24 Small intestinal motility patterns ..................................................................... 28 Peristalsis .................................................................................................... 28 GI motility disorders ........................................................................................ 30 Voltage-gated calcium channels in the nervous system ................................. 33 Overview ..................................................................................................... 33 Calcium channels in the ENS ...................................................................... 41 R-type calcium channels in the ENS ........................................................... 42 Ion channels as targets for disease treatment in the ENS .............................. 43 CHAPTER 2 Hypothesis and specific aims ................................................... 45 CHAPTER 3 lmmunohistochemical localization of R-type calcium channels in the guinea pig enteric nervous system ...................................................... 49 Abstract .......................................................................................................... 50 Introduction ..................................................................................................... 52 Materials and methods ................................................................................... 56 Results ........................................................................................................... 65 Discussion ...................................................................................................... 73 Summary and conclusions ............................................................................. 84 vi CHAPTER 4 Contribution of R-type calcium channels to the function of intrinsic primary afferent neurons and ascending interneurons in the guinea pig ileum myenteric plexus ............................................................... 108 Abstract ........................................................................................................ 109 Introduction ................................................................................................... 1 10 Materials and methods ................................................................................. 113 Results ......................................................................................................... 1 17 Discussion .................................................................................................... 123 Summary and conclusions ........................................................................... 132 CHAPTER 5 R-type calcium channels couple to nitric oxide-mediated inhibitory neurotransmission in the guinea pig ileum ................................ 147 Abstract ........................................................................................................ 148 Introduction ................................................................................................... 1 50 Materials and methods ................................................................................. 154 Results ......................................................................................................... 158 Discussion .................................................................................................... 163 Summary and conclusions ........................................................................... 169 CHAPTER 6 Inhibition of R-type calcium channels impairs peristalsis in the guinea pig ileum ............................................................................................. 177 Abstract ........................................................................................................ 178 Introduction ................................................................................................... 179 Materials and methods ................................................................................. 181 Results ......................................................................................................... 185 Discussion .................................................................................................... 192 Summary and conclusions ........................................................................... 200 CHAPTER 7 General discussion and conclusions ..................................... 215 REFERENCES ................................................................................................. 235 vii LIST OF TABLES Chapter 1. Table 1.1. Descriptions of enteric neuronal morphologies as described by Dogiel ....................................................................................... 12 Chapter 3. Table 3.1. Primary antibodies used to study the distribution of R-type calcium channels ............................................................................. 85 viii LIST OF FIGURES Chapter 1. Figure. 1.1. The enteric nervous system (ENS) ......................................... 11 Figure. 1.2. Morphological characteristics of enteric neurons ........................ 13 Figure 1.3. Model depicting the membrane structure and transmembrane topology of a voltage-gated calcium channel ............................................. 35 Chapter 3. Figure 3.1. Western blot analyses of (11 E subunit expression ...................... 87 Figure 3.2. The a1E-subunit is not expressed in the submucosal plexus in the ileum and colon .................................................................................. 88 Figure 3.3. Co-localization of a1E- and SP-ir in nerve fibers in the LMMP in the colon, stomach and ileum ...................................................................... 90 Figure 3.4. Localization of a1E-ir with calbindin-ir and calretinin-ir in the LMMP in the colon and duodenum ....................................................................... 92 Figure 3.5. Localization of a1E and 5-HT-ir in the LMMP in the small intestine.93 Figure 3.6. Co-localization of a1E-ir (red fluorescence) and VlP-ir (green fluorescence) in the LMMP in the ileum and stomach .................................. 95 Figure 3.7. Confocal photomicrographs showing co-Iocalization of a1E-ir (red fluorescence) and NOS-ir (green) in the myenteric plexus ........................... 97 Figure 3.8. Mean number of neurons per ganglion .................................... 99 Figure 3.9. Co-Iocalization of a1E-ir (red fluorescence) with SP-, VIP- and N03- ir (green fluorescence) in nerve fibers supplying the circular muscle layer ...... 100 Figure 3.10. Co—localization of a1E-ir (red fluorescence) with TH-ir (green) in the myenteric plexus ................................................................................ 101 Figure 3.11. The density of neuronal cell bodies co-Iocalizing (11 E and TH....103 ix Figure 3.12. Distribution of a1E (red fluorescence) with calbindin, SP, calretinin, and VIP (green fluorescence) in dissociated myenteric neurons in the guinea pig ileum ............................................................................................... 104 Figure 3.13. Co-localization of a1E-ir with NOS, synaptophysin, SP with synaptophysin, and a1E-ir with pan-cadherin in the guinea pig ileal whole mount (WM) LMMP and myenteric neurons in culture ........................................ 106 Chapter 4. Figure 4.1. CdClz inhibits action potentials in AH neurons ........................ 133 Figure 4.2. CdCIz does not affect action potentials in S-type neurons ......... 134 Figure 4.3. NiCIz inhibits the action potential in AH neurons ..................... 135 Figure 4.4. Action potentials in S-type neurons are not sensitive to NiClz....136 Figure 4.5. NiClz, SNX-482 and w—CTX inhibit the action potential in AH neurons .................................................................................... 137 Figure 4.6. The late afterhyperpolarization (late AHP) in AH neurons is inhibited by NiCIz ......................................................................................... 138 Figure 4.7. Effect of CCICIz on sEPSPs in AH neurons ............................ 139 Figure 4.8. NiClz (50 pM) did not inhibit sEPSPs ................................... 140 Figure 4.9. Sensitivity of sEPSPs to w-conotoxin (w-CTX) (100 pM) ......... 141 Figure 4.10. Amplitude of fEPSPs after treatment with CdCI2 and during time controls ........................................................................................ 142 Figure 4.11. NICIz does not inhibit fEPSPs in cholinergic/purinergic S-type neurons ....................................................................................... 143 Figure 4.12. SNX-482 does not inhibit fEPSPs in cholinergic/purinergic S-type neurons ....................................................................................... 144 Figure 4.13. NiClz and w-CTX inhibit fEPSPs in purely cholinergic S-type neurons ....................................................................................... 145 Fig 4.14. Effect of SNX-482 and w-CTX on purely cholinergic fEPSPs ..... 146 Chapter 5. Figure 5.1. Representative traces showing NANC responses after transmural electrical field stimulation of the guinea pig LMMP in vitro ........................ 170 Figure 5.2. Concentration-response curves for the area under the curve (AUC) of relaxation responses caused by EF S .............................................. 171 Figure 5.3. LMMP contractions in the absence of histamine .................... 172 Figure 5.4. Effects of NiCIz, NLA, and apamin on LMMP contractions ....... 173 Figure 5.5. Effect of NiClz, and NiClz + NLA on inhibitory junction potentials (IJPs) evoked in guinea pig smooth muscle cells .................................. 176 Chapter 6. Fig 6.1. Peristalsis video imaging ....................................................... 202 Figure 6.2. EP50 pressure-response curve and representative traces ......... 203 Figure 6.3. Time controls showing stable peristaltic motor patterns ........... 205 Figure 6.4. Effect of w—conotoxin (w-CTX), MCI; and CdCIz on the AUC expressed as percent change ............................................................ 206 Figure 6.5. Effect of MCI; and NLA on (A) frequency, (3) AUC (C) fluid volume expelled .............................................................................................. 207 Figure 6.6. Effect of NLA and NiCl2 on (A) frequency, (B) AUC and (C) fluid volume expelled ............................................................................. 209 Figure 6.7. The peristaltic motor pattern changes with NiCl2, and NiCI2 + w-CTX treatment ......................................................................... 211 Figure 6.8. Effect of w-CTX alone, and w-CTX with NiClz, on peristalsis parameters ................................................................................... 21 3 Chapter 7. Figure 7.1. Proposed Model for Inhibitory Neuromuscular transmission...230 Figure 7.2. Schematic of the overall effect of NiClz on the peristaltic Reflex ......................................................................................... 232 xi LIST OF ABBREVIATIONS B-nicotinomide adenine dinucleotide phosphate Acetylcholine After hyperpolarization Area under the curve Autonomic nervous system Cadmium Chloride Calbindin Calretinin Central nervous system Chloride Choline acetyltransferase Endoplasmic reticulum Enteric nervous system Excitatory junction potential Fast excitatory post synaptic potential Functional Gastrointestinal disorder Gamma-aminobutyric acid Gastrointestinal Immunoreactivity Inhibitory junction potential Interstitial cells of Cajal xii B-NAD ACh AHP AUC ANS CdClz Calb Calret CNS Cl‘ ChAT ER ENS EJP fEPSP FGID GABA GI IJP ICC lntrinisic primary afferent neurons Irritable bowel syndrome Longitudinal muscle myenteric plexus Migrating myoelectric complex Neurokinin Neuropeptide Y Neurotransmitter Nickel Chloride Nicotinic acetylcholine receptors Nitric oxide synthase Nitric oxide Nitro-L-arginine Non adrenergic non cholinergic Norepinephrine Omega-agatoxin Omega-conotoxin Phosphate buffered saline Potassium pyridoxal-phosphate-6—azophenyl-2',4'-disulfonate Serotoninl5-Hydroxytryptamine Slow excitatory post synaptic potential Sodium Substance P xiii IPANs IBS LMMP MMC NK NPY NT NiCl2 nAChRs NOS NO NLA NANC NE w-ATX w-CTX PBS K“ (PPADS) 5-HT sEPSP Na+ SP Tertiary plexus Tetrodotoxin Tyrosine hydroxylase Vasoactive intestinal peptide Water xiv tp TI'X TH VIP H20 CHAPTER 1 GENERAL INTRODUCTION THE NERVOUS SYSTEM The nervous system is composed of the central nervous system (CNS) which includes the brain and spinal cord, and the peripheral nervous system which comprises peripheral nerves, receptors, effectors and ganglia that exist between the CNS and the periphery. In response to changes in the external environment or the internal milieu, the brain can detect, interpret and respond to these changes. This is achieved by the transduction of sensory information in the form of electrical impulses from the periphery to the CNS that are then transmitted via neurotransmitters to muscles, organs and glands. It is this capability of nerve cells to wire themselves into a network and create emergent outputs that are not simply predicted by the inputs, that is testament to the exquisite complexity of the nervous system. The British physiologist Sir Charles Sherrington referred to this elegance of neural control when he famously described the integrative function of the nervous system. That is, the ability of the nervous system to effect appropriate responses upon receiving different types of information is due to its functional interconnectivity and sensitivity. Towards this end, the manifestation of plasticity within neural circuits as a consequence of sensory inputs or motor outputs is an inherent property of our continually changing nervous system that allows us to adapt our behavior, acquire new tasks, remember past events and recognize objects. In this regard, the concept of how synaptic transmission shapes the connectivity and functions of neural circuits is crucial to understanding fundamental mechanisms of nervous system function. STRUCTURE AND FUNCTION OF A NEURON The great neuroanatomist Santiago Ramon y Cajal is most famous for being the primary proponent and developer of the Neuron Doctrine which states that “a neuron is an anatomically and functionally distinct cellular unit that arises through differentiation of a precursor neuroblast cell” (Guillery, 2005; Glickstein, 2006). The nerve cell, or neuron, is the structural and functional unit of the nervous system. Although several anatomical types of neurons exist, for example, pyramidal, granule, Purkinje and myenteric neurons, they all have several common features. Each neuron has a soma or cell body that contains a single nucleus which contains the genetic information for protein synthesis. The cell body is responsible for integration of incoming information. Many cellular organelles such as mitochondria, ribosomes, endoplasmic reticulum, the Golgi complex, Iysosomes, and cytoskeletal proteins, reside within neurons (Shepherd, 1994). Connected to the cell body are dendrites which collect afferent information and serve to increase the surface area via which neurons can communicate. The axon begins at the soma and extends to surrounding and distant nerve cells. Their function is neurotransmission through action potentials that encode signaling information. Neurons can be differentiated according to their configuration (multipolar or unipolar) and function (sensory, interneurons and motorneurons). Sensory neurons have typically unipolar processes, while interneurons and motorneurons have multipolar processes. The classical definition of a sensory neuron is a neuron that transmits information into the CNS, and therefore has an afferent modality. Efferent (motor) neurons transmit information to effector cells, such as muscle cells within the gastrointestinal (GI) tract. The function of interneurons is integration of information in the communication pathway between sensory and motorneurons. Neurons therefore are functionally distinct entities that have specific roles that depend on their intrinsic properties (Jessell & Kandel, 1993). NEUROTRANSMITTERS, NEUROPEPTIDES AND SYNAPTIC TRANSMISSION There has been an increasing and active interest in the various chemicals utilized by neurons. Advances in this area have led to an understanding of their structure, biochemical and pharmacological characteristics, functions, and how they serve as diffusible chemical messengers between neurons. Several benchmarks need to be met before a chemical is considered to be a neurotransmitter: 1) it must be synthesized and released from neurons, 2) it must act at the postsynaptic cell to produce specific changes in that cell, such as changes in input resistance, 3) its action should be blocked by known antagonists, and 4) there must be a mechanism through which its action is terminated (Deutch & Roth, 1999). Neurotransmitters are subdivided into the classical transmitters and neuropeptides. Classical neurotransmitters The classical neurotransmitters are small, low molecular weight chemicals that are produced in nerve terminals. Specific enzymes catalyze the biosynthesis of neurotransmitters before they are packaged into synaptic vesicles by a vesicular transporter. Acetylcholine, the biogenic amines (norepinephrine, epinephrine, dopamine, serotonin, histamine), some amino acids (glutamate and aspartate) and the ubiquitous CNS neurotransmitter gamma-aminobutyric acid (GABA) are all classical neurotransmitters (Amara & Arriza, 1993; Amara & Kuhar, 1993). Although neurotransmitter distribution is heterogeneous, specific neurotransmitters modulate activity in different regions within the CNS and peripheral nervous system. The basic stages in classical chemical neurotransmission are: a) synaptic vesicle (~ 50 nm in diameter) synthesis and docking of vesicle loaded with neurotransmitters at the active zone; b) action potential initiation at the axon hillock and invasion of the synaptic terminal by the arrival of the depolarizing action potential; 0) calcium entry though voltage-gated calcium channels; d) fusion of vesicles with the plasma membrane and release of neurotransmitter into the synaptic cleft; e) binding of neurotransmitters to specific receptors; f) transduction of the signal resulting in a postsynaptic response; 9) termination of neurotransmission by active reuptake of the transmitter at presynaptic nerve terminals or metabolic inactivation by degradative enzymes in the synaptic cleft (Jessell & Kandel, 1993; Kelly, 1993). Neuropeptides Neuropeptide synthesis is energetically more demanding than small-molecule transmitter synthesis. Whereas the small molecule neurotransmitters are synthesized in neuronal nerve terminals, peptide transmitters are synthesized in the neuron cell body. The messenger RNA (mRNA) coding for the neuropeptide is translated into a peptide precursor on ribosomes in the rough endoplasmic reticulum (ER) within the cell body. The pre- propeptide is then transported to the Golgi complex where it is packaged into large dense-core vesicles (~ 75-150 nm in diameter). These vesicles are then transported by fast axonal transport from the cell body to the nerve terminal. lmportantly, neuropeptides can undergo post-translational modifications including glycosylation, methylation and sulfation. This has implications for the stability of the mature peptide. Further, large-dense core vesicles do not always undergo exocytosis at the active zones, but can release neuropeptides at extra-synaptic regions. Nonetheless, calcium influx into the nerve terminal is a critical determinant of neuropeptide release. To this end, the synaptic machinery necessary for peptide transmitter release is highly sensitive to calcium, more than that of the small molecule transmitters. Peptide transmitters therefore do not require large increases in intracellular calcium concentrations for exocytosis (Jessell & Kandel, 1993; Kelly, 1993; Goda & Davis, 2003). THE AUTONOMIC NERVOUS SYSTEM The autonomic nervous system (ANS) is the division of the peripheral nervous system that is responsible for involuntary sensory and motor control of visceral organs. The function of autonomic nerves is regulated by networks of nuclei that are located in the brainstem, hypothalamus and forebrain (cerebral cortex, basal ganglia, Iimbic system, thalamus and olfactory bulbs and tracts). These brain regions generally do not impart conscious sensation from visceral reflexes (Langley, 1922). The ANS is responsible for the co-ordination of key survival functions, such as breathing, heart-rate, digestion, and urogenital regulation. The ANS has three divisions: sympathetic, parasympathetic, and enteric (Richerson, 2003). Communication between the CNS and the sympathetic and parasympathetic pathways occurs through preganglionic neurons that are located in nuclei in the brainstem and spinal cord, and postganglionic neurons found in peripheral ganglia that receive synaptic input from preganglionic neurons. The cell bodies of sympathetic preganglionic neurons are found in the interrnediolateral cell column at levels T1 through L3 in the thoracic and lumbar regions respectively of the spinal cord. Axons from these preganglionic cell bodies exit the spinal cord through the ventral root and enter the white myelinated rami. They then synapse within a sympathetic chain (paravertebral) ganglion, or may exit that paravertebral ganglion to synapse on postganglionic sympathetic neurons in prevertebral ganglia (celiac, superior mesenteric, and inferior mesenteric). Parasympathetic preganglionic neuronal cell bodies are found in the craniosacral division of the autonomic nervous system. Preganglionic axons travel in the vagus, oculomotor, facial and glossopharyngeal cranial nerves and synapse with postganglionic parasympathetic neurons located mostly within target organs. Both sympathetic and parasympathetic preganglionic neurons use acetylcholine (ACh) as the transmitter. While postganglionic parasympathetic neurons also release ACh on to final target cells, most postganglionic sympathetic neurons release primarily norepinephrine (NE) onto the viscera. The sympathetic and parasympathetic divisions of the ANS produce opposite effects on the viscera (Richerson, 2003) INNERVATION OF THE GASTROINTESTINAL TRACT Extrinsic innervation of the ENS Most enteric neurons are not directly innervated by the CNS (Galligan, 2002b). This salient feature of the GI tract implies that the gut must manifest this independence from CNS input via intact intrinsic neural networks. The small intestine and the proximal region of the large intestine are less dependent on sympathetic and parasympathetic innervation than the esophagus, stomach and distal colon (Goyal & Hirano, 1996). The ENS nonetheless does receive sympathetic, parasympathetic, extrinsic vagal and spinal neuronal input (Grundy et al., 2006). Parasympathetic innervation of the gut occurs via the vagus and pelvic nerves. The myenteric plexus receives the majority of vagal efferent input, and primary afferent neurons with cell bodies in nodose ganglia branch throughout the plexus forming synapses with enteric neurons (Kirchgessner & Gershon, 1989; Holst et al., 1997). Cell bodies of spinal afferent neurons reside in dorsal root ganglia and their nerve endings mainly innervate the large intestine. Spinal afferents relay nociceptive information about the GI tract to the CNS (Cervero, 1994). These extrinsic sensory nerves, taken together, are important determinants of the physiological status of the gut, and do so on a moment to moment basis. Parasympathetic innervation of specific regions in the GI tract could mean that activation of nicotinic acetylcholine receptors (nAChRs) by nicotine needs to be strong in order for the reflex pattern of emesis and defecation to occur. Sympathetic innervation of the gut wall occurs though postganglionic neurons located with prevertebral ganglia. Tyrosine hydroxylase (TH) is the rate-limiting enzyme for the synthesis of catecholamines including NE (Grima et al., 1985). Based on the presence of TH in sympathetic nerve endings, three subtypes of branching, varicose sympathetic axons exist that supply the ENS: 1) those that target blood vessels contain TH and neuropeptide Y (NPY); 2) those that provide input to the submucosa contain TH and somatostatin; and, 3) those supplying myenteric neurons contain only TH (Macrae et al., 1986). In fact, within the myenteric plexus, NE acting at az-adrenoreceptors blocks neurotransmitter release from ACh-containing myenteric neurons, leading to an inhibition of motility (Nishi & North, 1973b; Hirst & McKirdy, 1974; Frigo et al., 1984). It has been suggested that the greater innervation of specific regions of the ENS may be because of an override system that the CNS has put into place to be able to control local gut motility. Intrinsic innervation: the ENS The mammalian small intestine contains approximately 100 million neurons, about as many neurons as that found in the spinal cord (Gershon, 1997). The ENS (Fig. 1.1) consists of two ganglionated plexuses: the myenteric plexus which is found in between the longitudinal and circular smooth muscle, and which innervates the GI smooth muscle, and the submucosal plexus located betWeen the circular muscle and the mucosa, and which regulates neuroimmune interactions, local blood flow, secretion and absorption across the GI epithelium (Kunze & Furness, 1999). The myenteric and submucosal plexuses are connected by numerous fiber bundles that facilitate secretomotor and motility reflexes (Furness et al., 1990). Within each plexus, the ganglia are connected via interganglionic fiber tracts, and these two nerve plexuses therefore provide the basis for all neural reflex activity in the GI tract. These reflexes are mediated by three classes of enteric neurons: intrinsic primary afferent neurons (lPANs) lsensory neurons, interneurons and motor neurons. Between 1895 and 1899, the Russian histologist AS Dogiel conducted extensive anatomical experiments in which he traced and characterized the morphologies of human and guinea pig enteric neurons (Table 1). He found three distinct subtypes, now known as Dogiel types I, II and III. Dogiel type I neurons (Fig. 1.2) are flat, stellate cells with lamellar dendrites, are unipolar, and with a cell body diameter between 13 and 35 pm in length and 9 and 22 pm in width. Dogiel type II neurons are larger in size (from 22-47 pM) and have round/oval cell bodies (Brehmer et al., 1999). A prominent feature of IPANs is that they are 10 Myenteric lnterganglionic ganglion fiber tract Submucosal Circular ganglion muscle Longitudinal muscle Submucosal _///- artery 200 pm Figure. 1.1. The enteric nervous system (ENS). The mucosa with villi protrudes towards the intestinal lumen. There are two ganglionated plexuses: the myenteric plexus located between the longitudinal and circular muscle layers that controls motility, and the submucosal plexus located between the mucosa and the circular muscle. The submucosal plexus controls blood flow of arterioles across the epithelial cell layer, as well as secretion and absorption (Furness, 2006a) 11 Table 1.1. Descriptions of enteric neuronal morphologies as described by Dogiel (after Brehmer et al., 1999) Types and numbers of processes Dendrites Axonal course I 4-20 dendrites, 1 axon Branch and end within the ganglion of origin. Short, thick, flat, wih varicosities, lamellar. Through neighboring ganglia, with collaterals sporadically to the muculature ll 1-16 dendrites, 1 axon Leave ganglion of origin. Structural resemblance to axons. Ramify into long, thin, smoothly contoured branches. To other ganglia lll 2-10 dendrites, 1 axon Ramify and end within ganglion of origin. Through other ganglia 12 Dogiel type I Dogiel type II Figure. 1.2. Morphological characteristics of enteric neurons. (A) Dogiel type I neurons are flat with short lamellar dendrites and a single long axon. These neurons are the interneurons and motorneurons. (B) Dogiel type II neurons have smooth surfaces with several dendrites and multipolar axonal processes. These neurons are the IPANs (Wood, 2006a). multipolar. In the guinea pig small intestine, IPANs constitute 26% of the total number of myenteric neurons and 11% of submucosal neurons (Furness et al., 2004). Dogiel type III neurons are filamentous neurons with long branched processes (Lomax et al., 1999; Furness, 2006a). lPANs IPANS have Dogiel ll morphology (Kirchgessner & Gershon, 1988; Furness et al., 1998; Pan & Gershon, 2000). This has generated some controversy since lPANs do not actually convey any conscious sensations from the intestine and are therefore not really sensory afferent neurons (Wood, 2006a). However, within the context of the intestine, the definition of IPANs are: sensory neurons in the wall of the gut that transduce mechanical or chemical stimuli into electrical activity within .a local reflex arc (Furness ef al., 2004; Brookes & Costa, 2006). The term IPANs will be used during the course of my dissertation. Most IPANs synapse with other IPANs to form positive feedfonrvard loops. This means that excitation within these loops or networks can build rapidly so that simultaneous activation of entire neural networks around the intestinal circumference is elicited. These bursts of excitation then drive the entire reflex circuitry towards threshold leading to the subsequent activation of interneurons and motorneurons that supply GI smooth muscle. The result is ascending contraction or descending relaxation. The axons of IPANS project from the myenteric and submucosal plexuses to the mucosal layer (Furness et al., 2004). The mucosa also contains enterochromaffin (EC) cells that, when activated via 14 chemical or mechanical (Kunze et al., 1998) stimulation of the mucosa, release 5-hydroxytryptamine (5-HT) (Kirchgessner et al., 1992; Mawe et al., 2006) and the other sensory mediators such as ATP (Bertrand, 2003). This indirect activation of IPANs occurs through 5-HT acting at 5-HT3 (Bertrand et al., 2000) and 5-HT1p (Kirchgessner et al., 1992) receptors on myenteric and submucosal AH nerve terminals respectively, and ATP acting at P2X receptors (Furness et al., 2004) elicit antidromic action potentials that propagate into the soma of the IPAN. The IPAN then provides synaptic input to interneurons, which in turn activate motor neurons to produce an excitation of GI smooth muscle oral to and an inhibition anal to the stimulus site, along distinctly polarized neural pathways (Galligan, 2002b). However, it should be remembered that the mucosal processes of lPANs are also activated directly, by chemical stimuli applied to the mucosa (Kunze et al., 1995; Brookes & Costa, 2006). According to Brookes and Costa (2006), a number of subsets of IPANs exist: 1) submucosal IPANs that respond to mechanical stimulation of the mucosa; 2) myenteric IPANs that respond to chemical stimulation of the mucosa; and 3) myenteric lPANs that are sensitive to tension along the gut wall as they posses mechanosensitive ion channels (Kunze et al., 1998; Kunze et al., 2000). Both myenteric and submucosal lPANs are neurochemically coded by the calcium-binding protein, calbindin, and this is the best marker known for lPANs (Pompolo & Furness, 198 8). 15 Myenteric interneurons The plasticity of the enteric neural circuitry is reflected by the presence of interneurons that are organized in a chain-like fashion along polarized pathways. All interneurons have Dogiel type I morphology (Furness et al., 1993). In the guinea pig ileum, there are four classes of interneurons: a single class of ascending interneurons and three classes of descending interneurons (Brookes, 2001). lnterneurons communicate via fast synaptic transmission (Galligan, 2002b). Ascending interneurons comprise 5 % of all myenteric neurons and stain positive for choline acetyltransferase (ChAT)/Calretinin/Substance P (SP)/enkephalin (Brookes et al., 1997). Their primary neurotransmitter is acetylcholine (ACh). These neurons provide synaptic input to other ascending interneurons and to excitatory circular and longitudinal muscle motorneurons (Pompolo & Furness, 1993, 1995). Their functional importance is to allow rapid spread of excitation and for the effective co-ordination of motor responses of the gut smooth muscle. In the guinea pig ileum, the first type of descending interneuron is immunoreactive for ChAT and the peptide hormone somatostatin. This class (4 % of myenteric neurons) is believed to be involved in the migrating myoelectric complex (MMC) (Furness, 2000). The second class of descending interneurons (2%) includes those that participate in the secretomotor and motility reflex. These neurons are immunoreactive for ChAT and 5-HT (at 5-HT3 receptors) and use both ACh and 5-HT as their primary neurotransmitters. Because at least 90% of the body’s 5-HT is localized within EC cells, this may represent a significant 16 source of the 5-HT that cause fast synaptic excitation within this interneuronal pathway. The third group of descending interneurons (5%) are those involved in the local reflex pathway. These neurons are neurochemically coded with ChAT/nitric oxide synthase (NOS)/vasoactiVe intestinal peptide (VlP)/gastrin- releasing peptide (GRP)/NPY, and their principal transmitters are ACh and ATP (Furness, 2000; Brookes, 2001). Evidence indicates that ACh and ATP may be the important transmitters released from nerve terminals of these neurons (LePard et al., 1997). The heterogeneity of three classes of descending interneurons may be to allow diverse communication patterns to take place, from the spread of inhibition along long (~ 40 mm) distances of the gut, to co- ordinating the MMC as well as being crucial mediators of relaxation (Brookes & Costa, 2006). Myenteric motorneurons All enteric motorneurons have Dogiel Type I morphology. Excitatory and inhibitory circular and longitudinal muscle motorneurons produce GI smooth muscle. contraction and relaxation. In the guinea pig ileum, excitatory circular and longitudinal muscle motorneurons constitute 12 % and 25 % of the total number of myenteric neurons respectively. Excitatory circular muscle motorneurons have the neurochemical phenotype ChAT/SP/enkephalin, while excitatory longitudinal muscle motorneurons immunostain positive for ChAT/calretinin/SP (Brookes & Costa, 1990). ACh and SP are the main neurotransmitters released from these neurons. Inhibitory circular and longitudinal muscle motorneurons make up 16 % l7 and 2 % respectively of the neurons in the myenteric plexus. These neurons have the chemical phenotype NOSNIP/Pituitary adenylate cyclase-activating peptide (PACAP)/gamma-aminobutyric acid (GABA) (Costa et al., 1986a; Furness et al., 1987; Brookes & Costa, 1990). Circular muscle motorneurons are important components that are crucial to peristalsis. Longitudinal muscle motorneurons may be responsible for fine-tuning motility in a graded manner (Brookes & Costa, 2006). Substantial evidence suggests that motorneurons in each of the muscle layers are activated synchronously (Hennig et al., 1999; Spencer et al., 2002). ELECTROPHYSIOLOGICAL CHARACTERISTICS OF MYENTERIC NEURONS lPANs are electrophysiogically characterized as AH neurons (Nishi 8. North, 19733; Hirst et al., 1974). The defining features of AH neurons are: a resting membrane potential of approximately -70 mV, a prominent calcium shoulder on the repolarizing phase of the action potential, the long and late afterhyperpolarization (late AHP), and an absence of anodal-break excitation at the offset of a hyperpolarizing pulse (Furness et al., 2004; Wood, 2006a). The major component of the upstroke during the AH somal action potential (approximately 75-110 mV in amplitude) is a tetrodotoxin (Tl‘X)-sensitive sodium current, and a TTX-insensitive calcium current is most prominent on the falling phase of the action potential (North, 1973; Hirst et al., 1985; Rugiero et al., 2003). The longer-lasting calcium current is carried by N-type (a1B) calcium 18 channels and generates a calcium-shoulder during the falling phase of the action potential (Rugiero et al., 2002). R-type calcium channels may also contribute to this calcium conductance (Rugiero ef al., 2002). Calcium entry induces further calcium release by activating ryanodine receptors on the endoplasmic reticulum (Hillsley et al., 2000; Vogalis et al., 2001). The action potential is followed by a characteristic long-lasting (2-30 3) afterhyperpolarization (AHP) mediated by the amplified level of intracellular calcium that activates an intermediate-conductance (lK) calcium-dependent potassium channel found only within the cell bodies (Nishi & North, 1973a; Morita et al., 1982; Furness et al., 2003). Since the AHP limits the action-potential firing rate within the cell bodies of AH neurons, these neurons cannot fire trains of action potentials and therefore control the frequency of sensory information output to interneurons and motorneurons (Hirst et al., 1974). Furthermore, the hyperpolarization produced by the AHP triggers the activation of a non-selective cationic current (In) that increases the excitability of AH neurons (Galligan et al., 1990). If IPANs/AH-type neurons are sensory neurons, then S-type neurons are the interneurons and motorneurons. In S-type neurons, the AHP following the action potential is short, lasting 20-100 ms (Hirst et al., 1974). Action potentials in these neurons are generated by the voltage- dependent Na‘ and K‘ conductances in line with the classic Hodgkin and Huxley model. The resting membrane potentials of S neurons are modulated by an inwardly-rectifying potassium current (Kir) (Rugiero et al., 2002). 19 NEUROCHEMICAL CODING AND SYNAPTIC TRANSMISSION IN THE ENS Functional classification of enteric neurons according to the neurochemical markers they express and their axonal projections is well established in the guinea pig (Brookes, 2001). As mentioned previously, ACh, SP, NO, VIP, 5-HT, the calcium-binding proteins calbindin-D28K (calbindin) and calretinin are all markers for specific subsets of enteric neurons (Costa et al., 2000; Brookes, 2001; Quinson et al., 2001). Moreover, there are two types of chemically-mediated synaptic mechanisms used by enteric neurons: fast and slow neurotransmission (Bornstein et al., 1994; Costa et al., 1996; Furness, 2000) Slow synaptic transmission in the ENS Long-lasting changes in neuronal excitability results from transmitter binding to a family of cell surface receptors called G-protein-coupled receptors. These are heterotrimeric seven-transmembrane domain receptors with on, B and y subunits. G-proteins employ two main systems to mediate their effects: the cyclic AMP (CAMP), and phosphoinositol systems. A third system, the arachidonic system, is also used by G-proteins, but for the purpose of my dissertation, the arachidonic system will not be discussed here. The cAMP system is activated when a ligand (eg. norepinephrine) binds to a receptor (eg. [3 adrenergic receptor) that is in turn linked to a Gs transducing protein. G0,s then stimulates adenylate cyclase which then catalyzes the conversion of ATP into the second messenger CAMP. CAMP then activates protein kinase A. The transducer 20 of the phosphoinositol pathway is G... Go,q activates the primary effector phospholipase C which in turn cleaves phosphoinositol phosphate (PIPz) to yield two second messengers, diacylglycerol (DAG) and inositol 1,4,5-triphosphate (lP3). IP3 acts at ligand-gated IP3 receptors on the endoplasmic reticulum (ER) to cause the release of calcium from these intracellular membrane stores. DAG remains in the membrane where it activates protein kinase C (PKC) (Siegelbaum et al., 2000). Slow synaptic excitation thus serves to modulate channel and intracellular enzymes to produce a remarkable array of responses, from changing the neural circuitry to altering synaptic transmission pathways. For example, ion channels such as K+ channels can be phosphorylated, leading to a change in their probability of opening. Once these channels are phosphorylated, they close, causing a long-lasting increase in excitability that can be experimentally detected as an increase in input resistance (Galligan, 1998; Galligan et al., 2000). This particular mechanism is a hallmark of slow synaptic excitation in AH neurons. Trains of high frequency electrical stimuli (5—20 Hz) elicit sEPSPs in AH neurons (Morita & North, 1985). Slow synaptic excitation occurs via G-protein-coupled receptors due to the long latency (> 50 ms) of the sEPSP (Galligan, 2002b). sEPSPs recorded from AH neurons are mediated by ACh and SP that are co- released from the same nerve endings (Galligan, 1999). ACh acting at M muscarinic receptors, however, is not the major contributor to the sEPSP in AH neurons because sEPSPs evoked in the majority of these neurons in the presence of muscarinic receptor antagonists are unaffected (North et al., 1985). SP acting at neurokinin-3 (NK3) receptors coupled to inositol phospholipid 21 hydrolysis and PKC is believed to be the principal mediator of slow excitatory neurotransmission in the ENS (Guard et al., 1988; Bertrand & Galligan, 1995). Forskolin, a direct activator of adenylate cyclase coupled-PKA, also mimics the sEPSP implying that the PKA pathway may be also be involved in mediating the sEPSP in IPANs (Bertrand & Galligan, 1995). NK3 receptors are located on IPANs and on interneurons that receive synaptic input from lPANs (Holzer & Holzer-Petsche, 2001). lPANs also receive slow synaptic input from other IPANs and therefore form self-reinforcing networks (Kunze & Furness, 1999). SP also acts at postsynaptic PKC-coupled NK1 receptors on excitatory and inhibitory enteric motor neurons (Holzer & Holzer-Petsche, 2001). lPANs can therefore directly mediate slow synaptic excitation of motor neurons (Holzer & Holzer- Petsche, 2001; Galligan, 2002b). The sEPSP is due to: a) suppression of the leak potassium conductance (Galligan, 1998), b) phosphorylation of calcium- dependent potassium channels causing them to close (Bertrand & Galligan, 1995), and 0) increase in chloride conductance in about 40 % of myenteric AH neurons (Bertrand & Galligan, 1994; Wood & Kirchgessner, 2004). Changes in input resistance assume a biphasic mode, where the first phase of the increase in input resistance is the closure of calcium-dependent lK channels, while increased CI' conductance is responsible for the second phase of decreased input resistance (Wood, 2006a). In order for the AH neuron to generate an action potential, the sEPSP has to reach threshold. Once this occurs, calcium then enters through N-type channels to trigger calcium-induced-calcium-release from the ER (Rugiero et al., 2002). Calcium then activates the IK channel, increasing 22 its probability of opening. At the same time, protein phosphatase 28 (calcineurin) is also activated by the released calcium and dephosphorylates the IK channel, thus permitting it to re—open (Vogalis et al., 2004). The IK channel is then free to mediate the late AHP that is the characteristic feature of AH neurons. What then is the functional importance of sEPSPs in the ENS? Firstly, during a sEPSP, excitatory inputs to that AH neuron will be potentiated. Second, intestinal motility requires a sustained neural response to cause prolonged excitation or inhibition in the effector systems (smooth muscle, secretory glands, blood vessels) (Bertrand & Galligan, 1995; Wood & Kirchgessner, 2004), and the electrical properties of AH neurons appear to adequately fit this description. Based on their electrophysiological properties, IPANS/sensory neurons are classified as “AH” neurons, and interneurons and motor neurons are defined as “S” neurons Fast synaptic transmission in the ENS A single focal stimulus applied to interganglionic connectives produces fEPSPs in S neurons (Clerc et al., 1999; Alex et al., 2002; Galligan, 2002b). fEPSPs recorded from S neurons are mediated by ACh acting at ligand-gated pentameric nAChRs and ATP acting at trimeric P2X receptors (Galligan et al., 2000). ACh is the principal excitatory neurotransmitter in the ENS (Nishi & North, 1973a; Zhou et al., 2002). nAChRs and P2X receptors are non-specific cation channels. Enteric nAChRs are also calcium-permeable (Trouslard et al., 1993). In 27% of S-type myenteric neurons with oral polarity, hexamethonium, a nAChR antagonist completely inhibited the fEPSP (LePard et al., 1997; LePard & 23 Galligan, 1999). This indicated that ACh was the only transmitter in the ascending pathway (Spencer et al., 2000). ATP is a co-transmitter with ACh in approximately 67% of S neurons in selective descending pathways (LePard & Galligan, 1999; Galligan et al., 2000). Further analysis revealed that 11% of 8- type neurons with aboral polarity were blocked by hexamethonium and 5-HT3, indicating that this subset of S-neuron uses both ACh and 5-HT for fast synaptic communication. In fact, in the small intestine, 5-HT3 receptors are found in the soma of AH and S-type neurons, where they mediate fastly desensitizing inward currents (Zhou & Galligan, 1999; Galligan, 2002a). Taken together, these data indicate that most myenteric neurons mediate fast excitatory synaptic transmission predominantly by ACh and ATP. Interestingly, nAChRs have been localized in the somatodendritic region (Torocsik et al., 1991) and nerve terminals (McGehee et al., 1995) of myenteric interneurons and motorneurons (Galligan, 1999, 2002a). According to Galligan (2002a), the paucity of recorded 5-HT3-mediated fEPSPs may be due to the small percentage of 5-HT-containing S-type interneurons. NEUROMUSCULAR TRANSMISSION IN THE ENS Autonomic nerves release neurotransmitters en passage from varicosities that occur at 5-15 pm intervals along axons (Blakeley & Cunnane, 1979; Blakeley et al., 1982). ACh, released from excitatory longitudinal and circular muscle motor neurons, acts at presynaptic M2 muscarinic and postjunctional M3 cholinergic receptors (Matsui et al., 2000; Eglen, 2001) and is 24 the major excitatory neurotransmitter responsible for contraction of GI smooth muscle (Paton, 1.955; Paton & Zar, 1968; Bogeski et al., 2005). Cholinergic excitatory junction potentials (EJPs) have also been recorded from guinea pig GI smooth muscle (Burnstock & Holman, 1963). SP, acting at NK1 receptors located on interstitial cells of Cajal (ICC) and on the smooth muscle (Sternini et al., 1995; Southwell & Furness, 2001), and at NKZ receptors on the muscle (Maggi, 2000) is an excitatory co-transmitter involved in these contractions (Holzer & Holzer- Petsche, 2001; Lecci et al., 2002). NK1 receptors are localized to the soma of NOS-containing inhibitory motor neurons (Portbury et al., 1996). Pharmacological evidence later showed that tachykinins excite NK1 receptors on inhibitory motor pathways within the ENS of the guinea-pig ileum, causing those inhibitory neurons to release NO, thus depressing motor activity (Lecci et al., 1999; Saban et al., 1999; Bian et al., 2000). nAChRs, however, are also present on the somatodendritic and nerve terminal regions (Wonnacott, 1997) of longitudinal and circular muscle motor neurons (Galligan, 1999; MacDermott et al., 1999; Schneider et al., 2000). Since excitatory longitudinal and circular muscle motor neurons contain both choline acetyltransferase (ChAT) and SP (Brookes, 2001), ACh and SP may be released from the same nerve endings (Galligan, 1999). Presynaptic nAChRs may therefore function as autoreceptors to facilitate SP release, causing non-cholinergic GI smooth muscle contractions (Galligan, 1999; MacDermott et al., 1999). There appears to be an oral-aboral increase in inhibitory innervation to GI smooth muscle. Burnstock and colleagues used sucrose-gap recordings with 25 transmural field stimulation to show that in the presence of the muscarinic receptor antagonist atropine, and guanethidine which blocks adrenergic neurotransmission, that 1 Hz stimuli pulses evoked transient hyperpolarizations or inhibitory junction potentials (IJPs) from smooth muscle from guinea pig taenia coli (Burnstock et al., 1963; Burnstock & Holman, 1963; Burnstock et al., 1964). TTX completely blocked the UPS. In guinea pig, murine and human, increasing the electrical stimuli of nerve fibers supplying the muscle layer causes UPS to become biphasic, consisting of an initial large-amplitude response, followed by a smaller-amplitude and slower second response (Wang et al., 2007). Relaxation of GI smooth muscle is mediated by non-adrenergic non-cholinergic (NANC) inhibitory enteric neurons (Furness, 2000). NO is the primary mediator of this relaxation (Li & Rand, 1990; Sanders 8 Ward, 1992) but ATP and VIP also contribute to inhibitory NANC responses (Makhlouf & Grider, 1993; Furness, 2000). The NO-synthesizing enzyme NOS, ATP and VIP co-Iocalize within inhibitory circular and longitudinal muscle motor neurons and may therefore act in concert, in parallel or in series to exert muscle relaxation (Costa et al., 1992; Grider & Jin, 1993; Brookes, 2001). NO is synthesized on demand, diffuses into smooth muscle cells and binds to soluble guanylate cyclase. Guanylate cyclase catalyzes the conversion of guanosine-5'-triphosphate (GTP) to cyclic guanosine- 3',5'-monophosphate (cGMP) (Lincoln et al., 2001). If intracellular calcium levels increase via calcium entry through voltage-gated L-type calcium channels or from intracellular calcium stores via activation of inositol triphosphate (IP3) or ryanodine receptors to cause excitation (contraction) of smooth muscle cells 26 (Sanders & Koh, 2006), then a decrease in intracellular calcium should produce inhibition (relaxation) of the muscle. Elevated cytosolic cGMP lowers intracellular calcium levels by activating a cGMP-dependent protein kinase (PKG) (Hofmann et al., 2006). PKG phosphorylates large-conductance calcium-dependent potassium (BK) channels (Robertson et al., 1993; Alioua et al., 1998), thus hyperpolarizing the cell, reducing calcium influx via L-type calcium channels and relaxing GI smooth muscle (Sausbier et al., 2000; Geiselhoringer et al., 2004). ATP, acts at P2Y1 receptors on GI smooth muscle to produce fast IJPs. In guinea pig circular muscle, the fast lJPs are completely inhibited by the selective P2Y1 receptor antagonist MR82179 (Boyer et al., 1998). A recent study in mouse colonic circular muscle showed that the purinergic co-enzyme B-nicotinamide adenine dinucleotide (B-NAD) may also be a non-adrenergic non-cholinergic (NANC) inhibitory neurotransmitter (Mutafova-Yambolieva et al., 2007). Hyperpolarization responses after picospritzing B-NAD onto muscle cells revealed that in the presence of B-NAD, MR32179 drastically inhibited those responses. Those data were confirmed with calcium-imaging studies performed in human embryonic kidney (HEK)-293 cells where calcium transients to B-NAD were completely blocked by MRS2179. Therefore B-NAD mediates it effects on GI smooth muscle via P2Y1 receptors (Mutafova-Yambolieva et al., 2007). Moreover, electrical field stimulation of murine colonic strips maintained in vitro showed that ATP and B-NAD are co-released, and that their release may be mediated predominantly by N-type calcium channels (Mutafova-Yambolieva et al., 2007; Smyth ef al., 2009). 27 VIP acts at Gs-coupled vasoactive—pituitary adenylate cyclase- activating peptide (VPAC) receptors on lCCs and GI smooth muscle cells to produce relaxation (F urness, 2000; Kurjak et al., 2001). VIP, via cyclic AMP, may activate a calcium-dependent potassium channel to elicit this response (Kawasaki et al., 1997; Bayguinov et al., 1999). SMALL INTESTINAL MOTILITY PATTERNS Peristalsis Peristalsis is a polarized motor reflex that is responsible for moving intestinal contents in an oral to anal direction (Trendelenburg, 2006; Gwynne & Bornstein, 2007). Phasic changes in intraluminal pressure are due to contractions and relaxations of GI smooth muscle (Huizinga et al., 2006). The GI tract produces three important motor patterns. The first is non-propulsive segmental contractions that mix and churn food; this enhances absorption of essential dietary nutrients. The second type of motor patterns are those that are fully propulsive contractions that propel food anally (Trendelenburg, 2006). This latter form of peristalsis was initially described by Bayliss and Starling (1899) and is referred to as the peristaltic reflex or “the law of the intestine”. Using small intestinal segments in vitro and in vivo, Bayliss and Starling proved that ascending excitation and descending inhibition occurs above and below a physiological stimulus presented within the lumen of the gut, one that mimicked a bolus of food (Bayliss & Starling, 1899). Those authors further showed that the circular and longitudinal muscle layers contract and relax synchronously to propel 28 food anally. The muscle layers also exhibits regular and rhythmic contractions called slow waves. Slow waves are depolarizing potentials that produce oscillations of the membrane potential of the smooth muscle, and these oscillations are close to threshold for calcium influx via L-type calcium channels (Sanders, 2008; Huizinga & Lammers, 2009). Slow waves are generated by the non-neuronal non-smooth muscle interstitial cells of Cajal, and are conducted to the adjacent smooth muscle cells via gap-junctions (Farrugia, 1999, 2008). Contraction and inhibition of the muscle occur through neurotransmitters released from enteric nerves that innervate the smooth muscle. The peristaltic reflex is mediated by both chemical stimulation of the mucosa (Bertrand et al., 2000) or distention (Hennig et al., 1999). Intestinal segments isolated from the guinea pig show consistent fully propagative peristaltic reflex patterns. This neuraIIy-mediated pattern is completely abolished in the presence of the voltage- gated sodium channel blocker tetrodotoxin (TTX), but myogenic rhythmic patterns persist (Huizinga et al., 1998). Experiments done with intestinal segments show that fluid-induced distention of the rat and mouse intestine in vitro does not evoke the classic Trendelenburg peristaltic reflex (Seerden et al., 2005). Instead, oscillatory contractions of GI smooth muscle cells in those animals have been frequently demonstrated and appear to be under low neural excitation, similar to the TTX-resistant contractions and relaxations observed in the guinea pig intestine (Seerden et al., 2005). Therefore, peristaltic motor patterns differ between species, and in rats and mice they are mainly due to the 29 high frequency of slow waves that manifest as the predominant motor patterns in those animals (Huizinga et al., 1998). The third electrical and motor pattern in the GI tract is the MMC that occurs during the interdigestive (fasting) state in mammals (Galligan et al., 1989). MMCs begin in the gastroduodenal region of the GI tract and slowly propagate towards the terminal ileum as a slowly-moving wave complex (Szurszewski, 1969). This wave is created by rapid and intense contractions that sweep along the gut. Humans and guinea pigs have approximately the same number of MMC cycles that occur every 70-90 minutes (Furness, 2006a). There are four known stages of the MMC: 1) a prolonged quiescent phase; 2) an irregular phase consisting of increasing action potential frequency coinciding with an increase in motilin synthesis and release from the duodenal mucosa; 3) a peak phase where the MMC propagates; and 4) a brief period of declining and irregular activity (Carlson et al., 1972). The functional significance of the MMC is the removal of undigested food, bacteria, and residual intestinal secretions (Huizinga & Lammers, 2009). GI MOTILITY DISORDERS The GI tract is a specialized hollow tube that consists of the oral cavity, pharynx, esophagus, stomach, small and large intestine, and anus. The function of those organs is to allow the sequential degradation of food by enzymes into smaller molecules during the transit of food. In this way, the body has access to vital minerals, nutrients, and water before they are expelled. This 30 goal is accomplished by the enteric nervous system (ENS), a collection of neurons embedded in the gut wall, that can function independently of input from the CNS (Gershon, 1981). Dysregulation of the ENS causes changes in motility, secretion and absorption, and leads to motility disorders (Sharkey & Kroese, 2001; Furness, 2006b). Gl motility disorders are important areas for the health of the United States and the rest of the world, and comprise about 40% of GI problems for which patients seek health care (Camilleri et al., 2005). They impose a heavy burden of illness, decreased quality of life, and decreased work productivity on those who are affected. Disorders of the ENS produce motor, secretory, and inflammatory and immunological disturbances of GI function. These disorders are believed to be associated with the degeneration of inflammatory cells or altered signaling of enteric neurons, and these changes are manifested as disturbances in GI transit or functional obstruction (Goyal & Hirano, 1996). Patients with FGle exhibit a greater altered motility response to psychological or physiological stressors when compared to control subjects (Drossman et al., 2002; Parkman et al., 2004). A major impact of GI disorders is its global social and economic burden. Irritable bowel syndrome (IBS), for example, has a prevalence rate of over 10% in known IBS cases and costs the United States economy at least $25 billion a year (Camilleri, 2001). IBS is probably the best known GI motility disorder. IBS is pathophysiologically complex, and a hallmark of this disorder is an alteration of the sensory innervation to the GI tract, causing visceral hypersensitivity in IBS patients. Hypersensitivity of IPANs has been implicated in patients with IBS, such that 31 their reduced threshold of activation would heighten motility reflexes (Galligan, 2004). There is no current cure for this condition. Patients with IBS have either increased or reduced gut motility, or alternating constipation and diarrhea in the same patient, and painful cramps (Drossman, 1999; Thompson et al., 1999). Used as an index of hypersensitivity, balloon bowel-distention experiments show that IBS patients possess visceral hypersensitivity (visceral hyperalgesia), suggesting that the neural sensory system in IBS patients may be dysregulated (Ritchie, 1973; Mertz et al., 1995; Bernstein et al., 1996). IBS is believed to have an immune component (Furness, 2006a). In small intestinal biopsies taken from IBS patients with elevated mast cell counts, degranulating mast cells were localized close to nerve terminals that supply the mucosa (Barbara et al., 2004). The resulting inflammatory insult to the GI tract increases excitability of enteric neurons (Sharkey & Kroese, 2001; Nurgali et al., 2007). The manifestation of plasticity within neural circuits as a consequence of sensory inputs or motor acts is therefore an inherent property of our continually changing central and peripheral nervous systems. Moreover, most HIV/AIDS patients (50-93%) report GI problems including dysphagia, diarrhea, nausea, vomiting, weight loss, abdominal pain, anorectal disease, GI bleeding and tumors during the course of their illness (Gazzard, 1988). . There is a positive correlation between a progressively compromised immune system and an increased prevalence of these GI symptoms (May et al., 1993) because the small intestine is a crucial site in immune defense due to its large surface area. Understanding mechanisms of 32 GI disorders is therefore imperative due to their pervasiveness across many disease-types. VOLTAGE-GATED CALCIUM CHANNELS IN THE NERVOUS SYSTEM Overview Voltage-gated calcium channels (Fig. 1.3) are found in virtually every excitable cell. Initial studies in the barnacle demonstrated properties of calcium channels: 1) intracellular calcium chelators favor excitability, 2) permeant divalent ions compete for entry into these channels, and 3) calcium channels are blocked by divalent metal ions including CdClz, NiCI2 and Co” (Hagiwara & Byerly, 1981b, 19813; Hille, 2001). They are physiologically crucial because they regulate a myriad of calcium-dependent intracellular events. Moreover, they possess the unique ability of translating electrical signals into chemical signals critical for neurotransmitter and hormone release, muscle contraction, gene expression, neurite outgrowth, synaptogenesis, synaptic plasticity and cell death (Berridge, 1998; Lee et al., 2002; Catterall et al., 2005). Structural, pharmacological and biophysical (voltage-dependence and kinetics) studies have described several distinct classes of neuronal voltage-gated calcium channels: a T-type referred to as a low-threshold current whose channel opens at more negative voltages (~ -64 mV; maximum activation ~-30 mV) with rapid inactivation (< 100 ms) (Perez-Reyes, 2003; Talavera & Nilius, 2006), and the high-threshold currents L-, N-, HQ and R-types whose channels open at more depolarized voltages (~ -30 mV; maximum activation ~ 0 mV) (Bean, 1989; Tsien 33 et al., 1991; Hoffman ef al., 1994) and generally inactivate slower (> 100 ms) (Yamakage & Namiki, 2002) (Table 2). Voltage-gated calcium channels are pentameric with six transmembrane or helices. These channels have five subunits: a1, 8, (126 and y. The pore-forming (between 85 and S6) subunit is a1, and is between 190-250 kDa in size with approximately 2000 amino acid residues. The a1 subunit also contains the voltage sensor (S4), channel gates, and drug/toxin binding sites (Isom et al., 1994; Hofmann et al., 1999; Ertel et al., 2000). Positive charges residing in the voltage sensor serves to activate the channel complex. The short intracellular loop between domains I and II has a GM and PKC phosphorylation site. Interaction of a1 with GM subunits reduces the rate of activation of the channel by reducing the open channel probability (Dolphin, 2003b). Phosphorylation of a1, however, increases the probability of calcium channel opening. A calcium-calmodulin site is positioned on the C- terminal tail. The large intracellular loop connecting domains II and III has the SNARE-binding region referred to as a synprint site. SNARES constitute the synaptic vesicle release machinery in neurons (Catterall, 2000). The synaptic membrane proteins syntaxin, synaptosome—associated protein of 25 kDa (SNAP25), and vesicle-associated membrane protein of 25 kDa (synaptobrevin) form the SNARE complex (Hofmann et al., 1999). The [3 subunit is hydrophilic and located intracellularly. It is mainly responsible for the proper insertion of the 0:1 subunit into the membrane. (3 subunits dramatically enhance the cell surface 34 Figure. 1.3. Model depicting the membrane structure and transmembrane topology of a voltage-gated calcium channel. Arrangement of the channel in the plasma membrane. The pore forming subunit is a1. The (125, B, and y subunits are the accessory subunits. Sites for channel glycosylation are indicated by “g”. Voltage-gated calcium channels are tetramers, each subunit contains six transmembrane (TM) helices. Note the conduction pore-loop (green) between TM 5 and TM 6, and the voltage-sensor S4 helices (Catterall & Few, 2008). 35 expression of a1 subunits and affect the kinetics and voltage-gating properties of the channel (Dolphin, 2003a). The a26 subunits are conjoined; the a2 subunit is found extracellularly and the y subunit is a transmembrane protein. 0:25 subunits also increase cell membrane expression of a1, but has less pronounced effects on channel kinetics (Davies et al., 2007). The y subunit has thus far not been intensively studied. These subunits, however, appear to have differential effects on a1 membrane expression, that is, y subunits may either have no effect on a1 expression, or may even reduce the level of expression (Catterall, 2000). LVA channel inactivation is similar to the voltage-dependent process of Na+ channels. However, HVA calcium channels show either slow (voltage-dependent) or fast (calcium-ion dependent) inactivation. Both serve to prevent the overloading the neuron with calcium as excessive calcium concentrations become cytotoxic. Calcium sensitive inactivation of HVA calcium channels appears to be calmodulin-mediated. HVA channels have calmodulin-binding domains on the intracellular C-terminal tail that bind calmodulin with extremely high affinity (Hofmann et al., 1999). Neurons therefore use these negative feedback mechanisms for survival. Synaptic transmission requires a dramatic elevation in intracellular calcium ions entering through voltage-gated calcium channels at concentrations of 50-100 uM (Matthews, 1996). These channels are believed to be the N- and P/Q-type calcium channels and densely associate with calcium microdomains in the nerve terminal (Hofmann et al., 1999). The calcium ions then interact with the synaptic machinery. While synaptotagmin is the putative calcium sensor, 36 there have been reports where calcium-mediated exocytosis still occurs in synaptotagmin-knockout mice (Shao et al., 1996; Sudhof & Rizo, 1996). T-type calcium channels T-type calcium channels produce a transient current with a tiny conductance with rapid activation and inactivation. These channels are strongly expressed in kidney and heart but have low expression in neural tissue (Perez- Reyes, 2003). T-type channels are very sensitive to high concentrations of MCI; (Vassort & Alvarez, 1994; Vassort et al., 2006). a1G, H, and l are the molecular counterparts of T-type calcium channels. L-type calcium channels L-type calcium channels have a large single-channel conductance and give rise to a long-lasting current. L-type channels are the principal channels in smooth, skeletal and cardiac muscle, where they mediate calcium entry into cells for excitation-contraction coupling in response to long or consistent depolarizations (Hille, 2001). L-type channels are highly sensitive to dihydropyridines (eg. nifedipine) and this is a distinctive property of these channels (Bean, 1989). 0:18, C, D and F are the molecular counterparts of L-type calcium channels (Bean, 1989). 37 N-type calcium channels N-type calcium channels were first localized in chick sensory neurons (Fox et al., 1987). Those authors found a calcium conductance that was dihydropyridine-insensitive and intermediate to that of T- and L-type currents. N- type (N for neuronal) channels are blocked by the peptide toxin w-conotoxin GVIA (w—CTX). w-CTX is a toxin isolated from the Pacific cone-shelled snail Conus geographus (McCleskey et al., 1987; Olivera et al., 1991). N-type calcium channels are important mediators of neurotransmitter release at central and peripheral synapses (Smith & Cunnane, 1996, 1997). a1B is the molecular counterpart of N-type channels (Bean, 1989). PIQ-type calcium channels Around the time that N-type calcium channels were being characterized with w-CTX, Rodolfo Llinas and colleagues (1991) were working on Purkinje cell function in the cerebellum. One of the tools at their disposal was a peptide toxin, w-agatoxin IVA (w-ATX) from the American funnel web spider Agelenopsis aperta. They observed a very slowly inactivating calcium current that was especially sensitive to w-ATX. They called this current the P (for Purkinje)-type calcium current (Hillman et al., 1991). However, four years later, a second w-ATX-sensitive calcium current was found in cerebellar granule cells that displayed rapid inactivation kinetics (Randall & Tsien, 1995). This was named the Q-type calcium current. The P/Q-currents are now combined because 38 they were shown to be homologous spliced variants (Bourinet et al., 1999). a1A is the molecular equivalent of P/Q-type channels (Bean, 1989). R-type calcium channels Soong and others (1993) initially classified the R-type (R for resistant, residual or remaining) calcium current as a LVA T-type channel. However, the R- type calcium current is really a HVA current (Schneider et al., 1994) that may activate at somewhat more negative voltages (~ -40 mV) and often inactivates more rapidly than the other high-threshold currents (Soong et al., 1993; Yamakage & Namiki, 2002). Unlike the other calcium currents, a complete understanding of the molecular properties of the R-type calcium channel remains elusive. The principal candidate for the ion-conducting subunit of R-type calcium channels is the a1E subunit, encoded by the Cat/2.3 gene (Zhang ef al., 1993). The a1E subunit was initially identified in the rabbit hippocampus, cerebral cortex and corpus striatum (Niidome et al., 1992). Using different splice-variants of the Cav2.3 gene, a1E was expressed in Xenopus oocytes and localized in adult rat brain (Soong et al., 1993). Recombinant a1E was subsequently expressed in HEK cells and Xenopus oocytes (Schneider et al., 1994). Soong et al. (1993) and Schneider et al. (1994) both confirmed a1E sensitivity to blockade by NiClz. It has further been shown that the toxin SNX-482 (Newcomb et al., 1998) and low concentrations of NiCl2 selectively block R-type channels (Sochivko et al., 2002). In situ hybridization and the reverse transcriptase-polymerase chain reaction both revealed extensive a1E mRNA expression in the adult mouse brainstem, 39 forebrain and cerebellum (Williams et al., 1994). lmmunohistochemical localization of a1E has been demonstrated in neurons within the rat caudate- putamen, thalamus, hypothalamus, amygdala, somata and axons of Purkinje cells (Yokoyama et al., 1995; Grabsch et al., 1999), cell bodies of CA1 pyramidal cells, mitral cells of the olfactory bulb (Yokoyama et al., 1995), human Purkinje and cerebellar granule cells (Yokoyama et al., 1995; Volsen et al., 1997), rat cardiac myocytes and human distal tubules (Weiergraber et al., 2000). R-type calcium channels comprise a significant portion of the calcium current recorded from murine dentate granule cells, hippocampal CA1 and cortical neurons (Sochivko et al., 2002). The study by Sochivko and others (2002) was conducted in transgenic mice where the a1 E gene was knocked out. Currents recorded from those CNS neurons showed that the R-type current was not eliminated but reduced. R-type calcium channel knockout-mice also show deficits in fear behavior (Lee et al., 2002), spatial memory as measured by the Morris water- maze (Kubota et al., 2001), and pain perception (Saegusa et al., 2000). The a1E- knockout mice in the study by Saegusa and others (2000) were not susceptible to inflammatory pain. R-channels may mediate synaptic communication by those neurons in descending anti-nociceptive pathways. Moreover, another study showed that a1E mutant mice are hyperglycemic (Matsuda et al., 2001). These data together highlight the physiological importance of the R-type channel in various organ systems. Thus Tottene and others (2000) used antisense oligonucleotides against the Cav2.3 gene. Those authors observed both sensitive and insensitive R-type channel conductances in response to SNX-482 in 40 cerebellar granular neurons maintained in culture. In studies with P/Q-type (01A) channel knockout mice, neurotransmitter release at the somatic neuromuscular junction occurred via R-type channels, suggesting that these channels are positioned close to presynaptic calcium sensors (Urbano et al., 2003). An interaction between R-type calcium channels and the synaptic vesicle protein synaptotagmin has been demonstrated previously (Wiser et al., 2002), further strengthening the importance of this channel in short-term plasticity and synaptic transmission. Calcium channels in the ENS Electrophysiological and immunohistochemical studies demonstrate the presence of all calcium channels, with the exception of T-type channels, in the ENS (Hirning et al., 1990; Kirchgessner & Liu, 1999; Smith et al., 2003). L- type calcium channels, however, are not major contributors to the synaptic excitation of enteric neurons (Reis et al., 2000; Reis et al., 2002). Further, the inhibition of L-type channels on GI smooth muscle reduces or abolishes spontaneous or evoked muscle movement (Horowitz et al., 1996; Farrugia, 1999). FIG type channels are expressed by myenteric neurons and contribute to neurotransmitter release from enteric nerves (Reis et al., 2000; Smith et al., 2003). However, weak immunoreactivity has often been associated with P/Q-type calcium channel (Kirchgessner & Liu, 1999). Nonetheless, data from guinea pig caecal submucosal (Cunningham ef al., 1998) and ileal myenteric neurons indicate that P/Q-type channels do contribute to neurotransmitter release (Reis et 41 al., 2000; Reis et al., 2002). N-type calcium channels are ubiquitously expressed in the guinea pig plexuses (Kirchgessner & Liu, 1999). In myenteric AH neurons, N-type channels were shown to be the main contributor to the calcium shoulder on the repolarization phase of the action potential (Furness et al., 1998; Vogalis et al., 2002). R-type calcium currents have also been recorded in guinea pig myenteric neuron maintained in culture (Bian et al., 2004). R-type calcium channels in the ENS R-type calcium channels are the principal subtype in the somatodendritic region of myenteric neurons maintained in primary culture (Rugiero et al., 2002; Bian et al., 2004), suggesting that R-type calcium channels may contribute to synaptic transmission and regulation of membrane excitability of myenteric neurons. Only two studies thus far have indicated functional roles for R-type calcium channels in the ENS (Bian et al., 2004; Bian & Galligan, 2007). Bian and others (2004) used whole-cell voltage clamp recordings to show that at least 50% of the calcium current present in myenteric neurons maintained in primary culture is the R-type current. After using a strong depolarization (0 mV) pulse to activate the R-type current, they applied N-, PlQ-, and L—type calcium channel antagonists sequentially to block the calcium currents. However, 50% of the remaining calcium current remained. This residual current was only blocked after application of MCI; (50 uM). This observation was confirmed when NiCl2 and SNX-482 (0.1 pM) mutually occluded the calcium current (Bian et al., 2004). That study suggested a somatodendritic role for R-type calcium channels in the ENS. A further role for R-type channels has been postulated based on a 42 selective coupling with az-adrenoreceptors (Bian & Galligan, 2007). Norepinephrine released from sympathetic postganglionic neurons inhibits GI motility (Paton & Vizi, 1969; Stebbing et al., 2001). Presynaptically-located a2- adrenoreceptors are activated by norepinephrine and inhibit neurotransmitter release by inhibiting calcium influx through voltage-gated calcium channels (Paton 8. Vizi, 1969; Stebbing et al., 2001). R-type calcium channels couple to inhibition of az-adrenoreceptors via a pertussis toxin-sensitive G-protein (Bian & Galligan, 2007). R-type channels may therefore be a target for norepinephrine released from sympathetic nerve terminals (Bian & Galligan, 2007). ION CHANNELS AS TARGETS FOR DISEASE TREATMENT IN THE ENS The concept of how alterations in synaptic transmission shape the connectivity and functions of enteric neural circuits has a key bearing on the consideration of normal physiological versus pathophysiological changes in the ENS. Voltage-gated calcium channels are critical mediators of synaptic transmission. Excessive calcium entry, however, is also cytotoxic. Calcium influx via calcium channels must therefore be precisely regulated because alterations in calcium channel function would produce widespread cellular changes. Perturbations in ion channel function affect diverse organ systems. In the neuromuscular system, ion channel diseases results from gain-of—function mutations to cause, amongst others, epilepsy, ataxia, myotonia and cardiac arrhythmias (Cooper & Jan, 1999). In other systems, cystic fibrosis and Bartter syndrome, for example, result from defects in chloride (CI‘) and potassium (K*) 43 channels that alter the transepithelial transport gradients of salt and water (Cooper & Jan, 1999). In the ENS, the type 2 Cl' channel (ClC-2) is found in enterocytes in the small intestine and colon (Lipecka et al., 2002). ClC-2 channels regulate intracellular volume and pH levels and activation of these channels on the enterocytes increase intestinal fluid secretion (Cuppoletti ef al., 2004). The highly selective ClC-2 agonist lubiprostone is now being used to successfully treat patients with chronic constipation (Cuppoletti et al., 2004). A recent electrophysiological study using whole-cell patch clamp recordings demonstrated that R-type calcium channels are the predominant calcium channel subtype expressed by myenteric neurons (Bian et al., 2004). This raised the questions of whether a) R-type calcium channels are localized selectively within subsets of myenteric neurons, b) the mechanisms by which these calcium channels contribute to synaptic transmission at neuro-neuronal synapses and neuromuscular junctions in the myenteric plexus and GI smooth muscle, respectively, and c) the physiological significance of the large R-type currents found by Bian et al. (2003). Voltage gated calcium channels are key pharmacological targets for treating pain, stroke, ‘epilepsy, migraine and hypertension (French & Zamponi, 2005). These observations support the concept that ion disturbances in ion channel function can contribute to the pathophysiology of some diseases, and that ion channels can also be the targets for therapeutically useful drugs. These studies will provide new insights into the contribution of R-type calcium channels to GI motility. 44 CHAPTER 2 HYPOTHESIS AND SPECIFIC AIMS 45 OVERALL HYPOTHESIS Gastrointestinal (GI) disorders decrease the quality of life for affected patients (Drossman et al., 1993; Stewart et al., 1999). These disorders are classified as either structural (organic) or functional, but patients may simultaneously suffer from both types. Inflammatory bowel disease (IBD) is an organic GI disorder with clearly defined injury and inflammation that affect motility (Stewart et al., 1999). When no pathological (anatomical, physiological or biochemical) etiology is apparent, the patient is diagnosed as having a functional GI disorder (FGID) (Longstreth ef al., 2006). There has been a surge in the diagnosis of FGle including IBS in recent years. The pathophysiological basis for FGIDs is unknown because FGID is due to an interaction of psychosocial factors and altered gut physiology via the brain-gut axis. This is pertinent as the GI tract has its own (enteric) nervous system (ENS) that allows the gut to function quasi-autonomously (Furness, 2000). Synaptic transmission requires voltage—gated calcium channels as mediators for a host of important biological events including neurotransmitter release and muscle contraction (Cooper & Jan, 1999; Lorenzon & Beam, 2000). R-type calcium currents can be recorded from guinea-pig myenteric neurons suggesting that pore-forming 01E calcium channel subunits are expressed in the ENS (Rugiero et al., 2002; Bian et al., 2004). At least 50% of the current in myenteric neurons is the R-type calcium current. This suggests an important role for R-type channels as mediators of neurotransmitter release or as transducers of other synaptic events in myenteric neurons, and implicates this channel as a potential drug target in combating motility disorders. 46 I therefore tested the overall hypothesis that R-type calcium channels are therefore functionally significant in controlling GI motility. This function was explored further through the following specific aims. Specific Aim 1. Myenteric neurons express R-type calcium currents (Bian et al., 2004), but the specific subset of myenteric neuron that expresses the R-type calcium channel is not known. These studies tested the hypothesis that R-type calcium channels are localized to AH neurons and/or S-type neurons. Specific Aim 2. Fast and slow synaptic neurotransmission may be selectively modulated by calcium channel subtypes. These studies tested the hypothesis that R-type or non-R-type calcium channels are respectively responsible for substance P (SP) release as the mediator of the slow excitatory postsynaptic potential (sEPSP) in AH neurons, and acetylcholine (ACh) release as the mediator of the fast EPSP in S neurons. Specific Aim 3. lmmunohistochemical data show that a1E is present in nerve endings of excitatory motor neurons [SP-immunoreactivity (ir)], and descending interneurons and inhibitory motor neurons [vasoactive intestinal peptide (VlP)/nitric oxide synthase (NOS)-ir]. These studies tested the hypothesis that R- type calcium channels mediate nerve-to-muscle transmission in the guinea pig isolated ileum. 47 Specific Aim 4. My goal was to use an integrative approach to examine the functional role of R-type channels in mediating the peristaltic reflex in the guinea pig ileum. I used a modified method of Trendelenburg (Billbring et al., 1958; Huizinga et al., 1998; Bian et al., 2003) to test the hypothesis that R-type calcium channels are necessary or are sufficient for propulsive peristalsis in vitro. 48 CHAPTER 3 IMMUNOHISTOCHEMICAL LOCALIZATION OF R-TYPE CALCIUM CHANNELS IN THE GUINEA PIG ENTERIC NERVOUS SYSTEM 49 ABSTRACT R-type calcium currents can be recorded from guinea pig myenteric neurons suggesting that a1E calcium channel subunits are expressed in the enteric nervous system. However, the neuronal subset(s) expressing a1E subunits are unknown. We used whole mounts from the stomach, intestine and colon of adult guinea pigs, and ileal myenteric neurons maintained in primary culture to map a1E-subunit expression. I also used antibodies against specific markers to determine if a1E was expressed selectively in neuronal subsets. a1E- immunoreactivity (ir) was found in colonic and gastric, but not intestinal myenteric cell bodies in the longitudinal muscle-myenteric plexus (LMMP). a1E-ir never co- localized with serotonin (a marker for descending interneurons). a1E-ir co— localized with substance P (SP) in myenteric nerve fibers and in the muscle layers throughout the gut. a1E-ir co-Iocalized with calretinin in cell bodies and nerve fibers in primary cultured neurons, in some varicosities in the LMMP, and with calbindin in cell bodies only in dissociated neurons. a1E-ir co-Iocalized with vasoactive intestinal peptide (VIP) and nitric oxide synthase (NOS) in myenteric nerve fibers and in muscle layers; only the colon contained a1E-VlP-ir neurons. a1E-NOS-ir neurons were found in the gastric myenteric plexus. a1E-ir co- localized with tyrosine hydroxylase in myenteric nerve fibers. a1E-ir was not detected in the submucosal plexus. I concluded that a1E-ir is expressed by cell bodies of myenteric intrinsic primary afferent neurons (calbindin-containing) and in ascending interneurons (calretinin-containing), inhibitory motorneurons (VIP/NOS-ir) and nerve endings of excitatory motorneurons (SP-ir). a1E-ir is 50 contained in sympathetic nerves supplying the myenteric but not the submucous plexus. R-type calcium channels could participate in synaptic transmission in the myenteric but not submucosal plexus. R-type calcium channels may also participate in neuromuscular transmission. 51 INTRODUCTION The enteric nervous system (ENS), a collection of neurons embedded in the gut wall, can function independently of input from the central nervous system (Gershon, 1981). The ENS consists of two ganglionated plexuses: the myenteric plexus which provides innervation to the gastrointestinal smooth muscle, and the submucosal plexus which regulates neuroimmune interactions, local blood flow, secretion and absorption across the gastrointestinal epithelium (Kunze & Furness, 1999). Changes in enteric neuronal function can therefore cause gastrointestinal motility disorders. Pharmacological and biophysical studies have described several distinct classes of neuronal voltage-activated calcium channels: the high-threshold currents L-, Ni, P/Q- and R-types, and a T-type referred to as a low-threshold current (Hoffman et al., 1994). The function of R-type calcium channels in the nervous system has not been firmly established. However, In studies using P/Q- type channel knockout mice, neurotransmitter release at the somatic neuromuscular junction occurred via R-type channels, suggesting that these channels are positioned close to sites of synaptic transmission (Urbano et al., 2003). In addition, an interaction between R-type calcium channels and the synaptic vesicle protein synaptotagmin has been demonstrated previously (Wiser ef al., 2002), further strengthening the importance of this channel in synaptic transmission. Neuronal calcium channels are composed of a pore-forming and voltage- sensing ail-subunit in association with accessory subunits which serve to 52 modulate the properties of the channel complex (Isom et al., 1994; Ertel et al., 2000). The pore forming subunit of R-type calcium channels is the a1E-subunit, encoded by the Cav2.3 gene. The a1E subunit was initially identified in the rabbit hippocampus, cerebral cortex and corpus striatum (Niidome et al., 1992). Recombinant a1E-subunits were expressed in human embryonic kidney (HEK) cells and Xenopus oocytes (Schneider et al., 1994) and these investigators established that R-type calcium currents could be blocked selectively by low concentrations (5 50 pM) of nickel chloride. It has further been shown that the spider toxin SNX-482 (0.1 pM) can also selectively block R-type calcium currents (Vajna et al., 1998; Tottene et al., 2000; Sochivko et al., 2002). In situ hybridization, reverse transcriptase-polymerase chain reaction and immunohistochemical studies have revealed that al E-subunits are expressed throughout the central nervous system of mice (Williams et al., 1994), rats (Yokoyama et al., 1995; Grabsch ef al., 1999) and humans (Yokoyama et al., 1995; Volsen et al., 1997). a1E-subunits have also been localized to non- neuronal tissues such as cardiac myocytes from rats and human distal tubules from the kidney (Weiergraber et al., 2000). a1E-subunits and R-type calcium channels are also expressed in sympathetic neurons (Murakami et al., 2007), sensory neurons (Fang et al., 2007) and somatic motor neurons (Urbano et al., 2003; Pardo et al., 2006) but localization of a1 E-subunits has not been studied in other parts of the peripheral nervous system. Electrophysiological and immunohistochemical studies demonstrate the presence of all calcium channels, with the exception of T-type channels, in the ENS (Hirning et al., 1990; 53 Kirchgessner & Liu, 1999; Smith et al., 2003). L-type calcium channels, however, are not major contributors to the synaptic excitation of enteric neurons (Reis et al., 2000; Reis et al., 2002). Furthermore, the inhibition of L-type channels on GI smooth muscle reduces or abolishes spontaneous or evoked muscle movement (Horowitz et al., 1996; Farrugia, 1999). PM type channels are expressed by I myenteric neurons and contribute to neurotransmitter release from enteric nerves (Reis et al., 2000; Smith et al., 2003). R-type calcium channels may also contribute to somatodendritic action potentials or neurotransmitter release from myenteric nerve terminals (Rugiero et al., 2002; Bian et al., 2004). However, those studies did not identify specific functional classes of myenteric neurons: intrinsic primary afferent neurons (IPANs) or interneurons and motorneurons that express the R-type channel. Electrophysiological studies have established that R-type calcium channels are the principal subtype in the somatodendritic region of guinea pig ileal myenteric neurons maintained in primary culture (Bian et al., 2004). These data suggest that R-type calcium channels may contribute to synaptic transmission and regulation of membrane excitability of myenteric neurons. Functional classification of enteric neurons according to the neurochemical markers they express and their axonal projections is well established in the guinea pig ‘gut (Brookes, 2001; Hansen, 2003). Substance P (SP), nitric oxide (NO), vasoactive intestinal polypeptide (VIP), 5-hydroxytryptamine (5-HT), the calcium-binding proteins calbindin and calretinin, and tyrosine hydroxylase (TH) are all markers for subsets of enteric neurons. 54 The predominance of R-type calcium currents in myenteric neurons suggests an important role for these calcium channels in enteric neurotransmission (Bian et al., 2004). However, since the specific classes of myenteric neurons have not been identified, it remains to be determined as to whether R-type channels are localized to subsets of these neurons. Furthermore, there have been no studies of the cellular distribution of 01E calcium channel subunits in specific regions of the gut. The aim of the present investigation was to therefore localize a1E subunits with neurochemical markers for functional classes of neurons in the ENS of the guinea pig. 55 MATERIALS AND METHODS Tissue collection All animal use procedures were approved by the Institutional Animal Care and Use Committee at Michigan State University. Adult male Hartley strain guinea pigs (n = 4; 300-400 g, Bioport, Lansing, MI) were deeply anesthetized with isoflurane (Abbott Laboratories, Chicago, IL), stunned by a blow to the back of the head, and then killed by severing the carotid arteries. Segments of the ileum, duodenum, gastric corpus, proximal and distal colon were dissected out and placed in 0.01 M phosphate-buffered saline (PBS; pH 7.2). Protein isolation and Western blotting The specificity and sensitivity of the a1E antibody were verified by Western blot. Longitudinal muscle-myenteric plexus (LMMP), submucosal and circular muscle preparations from the guinea pig ileum were isolated on ice and placed directly into liquid nitrogen. The rat brain and heart, and guinea pig heart served as positive controls for (11E expression. Tissues were ground to a powder in liquid nitrogen, and ice-cold homogenization buffer (125 mmol/L Tris (pH 6.8), 4% SDS, 20% glycerol, 0.5 mmol/L phenylmethylsulfonyl fluoride, 1 mmol/L orthovanadate, 10 pg/ml aprotinin, 10 pg/ml leupeptin] added. Homogenates were vortexed, sonicated briefly and transferred to a plastic centrifuge tube and spun at 4 °C to pellet debris. The supernatant was separated from the pellet and analyzed for protein concentration (BCA protein kit, Sigma- Aldrich, St. Louis, MO). Total protein (50 pg) was boiled for 5 minutes with 56 standard 4:1 sample buffer. Proteins were separated on 1 mm-thick, 10% SDS polyacrylamide gels using a Mini Bio-Rad® lll apparatus and all samples were loaded into the gel in parallel. After transfer, the nitrocellulose membranes were blocked at 4 °C overnight in 5% milk with 0.025% NaN3 and incubated with the 01E primary antibody (Table 1). Membranes were then rinsed three times in tris- buffered saline (TBS) with Tween (0.1%) with a final rinse in TBS and incubated with the appropriate horseradish peroxidase-linked secondary antibody (1:2000; Cell Signaling Technology, Beverly, MA, USA) for 1 hour at 4 °C with continuous shaking. ECL® reagents (Amersham Life Sciences, Arlington Heights, IL, USA) were used to visualize bands. Finally, gels were stained with Gel Code Blue (Pierce, Rockford, IL, USA) to verify protein loading. Whole-mount preparations Tissues were flushed of content, cut open along the mesenteric border, stretched out and pinned tautly onto a silicone elastomer-Iined (Sylgard, Dow Corning, Midland, MI) petri-dish, and fixed in 10% neutral-buffered formalin (Sigma-Aldrich) overnight at 4 0C. Forrnalin-fixed tissues were cleared using 3 washes with dimethyl sulfoxide (J T Baker, Phillipsburg, NJ) at 10 minute intervals. Tissues were then washed in 0.01 M PBS (3 X 10 min intervals) at room temperature and stored in 0.01 M PBS at 4 0C. Using fine scissors and forceps, the submucosal plexus, circular muscle and LMMP layers were sequentially peeled away and trimmed into ~8 mm2 57 pieces. Tissues were placed into separate wells of a 24-well tissue culture plate (Corning Glass Works, Corning, NY). Each well contained 0.5 ml of 0.01 M PBS. Primary cultures of myenteric neurons Newborn guinea pigs (~36 h in age, ~70 g in weight) were sacrificed by severing the major neck blood vessels after deep isoflurane anesthesia. The ileum was removed and placed in cold oxygenated (95% Oz; 5% 002) sterile Krebs’ solution of the following composition (mM): NaCI, 117; KCI, 4.7; CaClz, 2.5; MgClz, 1.2; NaHCOa 25; NaHPO4, 1.2; glucose, 11. The LMMP was removed from the entire length of the ileum and cut into 5 mm segments. The dissected LMMP was divided into four aliquots and placed into 1 ml Krebs’ solution containing 1600 U of trypsin (Sigma-Aldrich) for 30 min at 37 °C. The tissues were then triturated 30 times and centrifuged at 3000 g for 5 min. The supernatant was discarded and the pellet resuspended in 1 ml Krebs’ solution containing 2000 U of crab hepatopancreas collagenase (Calbiochem, San Diego, CA) for 30 min at 37 °C. The suspension was again triturated and centrifuged for 5 min at 3000 g. The pellet was then resuspended in Eagle’s minimum essential medium containing 10% fetal calf serum, gentamicin (10 ug/ml), penicillin (100 U/ml), and streptomycin (50 pg/ml) (Sigma-Aldrich). Cells were plated on coated poly-L- lysine glass coverslips and grown in an incubator in a 5% CO; atmosphere for up to 3 weeks. After 2 days in culture, cytosine arabinoside (10%) was added to the culture medium to inhibit smooth muscle and fibroblast proliferation. 58 lmmunohistochemical procedures All primary antibodies (see Table 1) were diluted in 0.01 M PBS/1% Triton X-100. For myenteric plexus preparations, the a1E primary antibody was incubated in combination with anti-calbindin, anti-calretinin, anti-5-HT, anti-SP, anti-VIP, anti-NOS and anti-TH primary antibodies. With the exception of anti-5- HT and anti-TH, the above antibodies were also applied to circular muscle preparations from each of the five regions. Primary antibody incubations were performed at 4 °C overnight in a humidified chamber. The tissues were then washed in 0.01 M PBS (3 X 10 min), and incubated with the appropriate secondary antibodies coupled to indocarbocyanine (CY3) or fluorescein isothiocyanate (FITC) for 90 min at room temperature. For the purposes of consistency during co-localization experiments, the ME primary antibody was always localized with a CY3-conjugated secondary antibody while all other primary antibodies were localized with a secondary antibody conjugated to F ITC. Tissue sections were subsequently washed in 0.01 M PBS (3 X 10 min) and mounted in buffered glycerol for confocal or conventional fluorescence microscopy. All photomicrographs are confocal images unless otherwise indicated. Sequential dual-immunolocalization of a1E and nNOS. Since the a1E and NOS antibodies were raised in rabbit, an established protocol was used to study co-localization of these antigens (Franzusoff et al., 1991; Negoescu et al., 1994). Tissue sections were first pretreated with 10% normal goat serum (Sigma- 59 Aldrich) (diluted in 0.01 M PBS/ 1% Triton X-100) for 1 h at room temperature to block non-specific binding sites. The sections were then incubated with the NOS primary antibody diluted 1:200 in 0.01 M PBS/ 1% Triton X-100, in a humidified chamber at 4 °C overnight. Sections were then washed with 0.01 PBS (3 X 10 min). To prevent non-specific binding of the secondary antibody, the sections were then incubated with 0.01 M PBS/ 1% Triton X-100 containing 10% normal goat serum for 1 h at room temperature. Incubation with a goat anti-rabbit secondary antibody conjugated to FITC was performed for 2 h at room temperature. The tissues were then washed with 0.01 PBS (3 X 10 min). Non- immune rabbit serum (Pierce Biotechnology, Rockford, IL) was then applied to the tissue sections for 1 h at room temperature to saturate binding sites on the first secondary antibody (goat anti-rabbit-FITC conjugate), thus preventing the second primary antibody (a1 E) binding. Monovalent goat anti-rabbit unconjugated fragment antigen binding (Fab) antibodies (Jackson lmmunoresearch Laboratories, West Grove, PA) (1:50 diluted in 1% Triton X- 100) were then added to the tissue sections and incubated in a humidified chamber at 4 °C overnight. These Fab antibody fragments bind to free antigen- binding regions on the first primary antibody and to sites on the non-immune rabbit-IgG molecules, thus obstructing the binding of the second secondary antibody. The sections were then washed in 0.01 PBS (3 X 10 min) and blocked with 0.01 M PBS/ 1% Triton X-100/10% normal goat serum for 1 h at room temperature. Tissue sections were then incubated with the second primary antibody (alE) diluted 1:400 in 0.01 M PBS/1% Triton X-100 in a humidified 60 chamber at 4 °C overnight, and subsequently washed in 0.01 PBS (3 X 10 min). The sections were again blocked with 0.01 M PBS/1% Triton X-100/ 10% normal goat serum for 1 h at room temperature. A goat anti-rabbit-CY3 conjugate was then applied to the tissue sections (2 h at room temperature). The tissues were rinsed in 0.01 PBS (3 X 10 min) and mounted with buffered glycerol for confocal and epifluorescence microscopy. To eliminate potential sources of signal cross-talk (Brelje et al., 1993; Schneider & Lopez, 2002) and to validate the specificity of immunostaining, three concurrent method control experiments were performed. These method controls included the omission of either one or the other primary antibody, and subsequent incubation with the corresponding secondary antibodies. The first control involved the omission of the first primary antibody, the exclusion of the second primary antibody served as the second control, and in the third control both primary antibodies were included. Antibody specificity for anti-aIE was determined using preadsorption tests on whole mount sections of the guinea pig stomach and intestine, and no a1E-ir was observed (see Fig. 3.30 inset). Confocal and epifluorescence microscopy. Tissue sections were examined using a Leica TCS-SLr laser scanning confocal system (Leica Microsystems Heidelberg, Mannheim, Germany) attached to a Leica Instruments DMLFSA microscope. Multicolor immunofluorescence analyses were then performed using a single line argon ion laser, (488 nm), which was used to excite FITC while a helium-neon laser (543 nm) excited CY3. Images were then 61 obtained using a 63 x 1.32 HCX PL APO oil immersion objective (Leica). The methods of Majlof and Forsgren (1993) and Schneider and Lopez (2002) were followed to facilitate accurate confocal imaging (Majlof & Forsgren, 1993; Schneider & Lopez, 2002). Sequential scan settings were used to obtain images at each wavelength individually. A xyz series range for a particular observer- detennined area was defined, and a z-series of optical sections through the area of interest and orthogonal to the optical axis was obtained. The number of sections taken was calculated automatically by doubling the optimized step size. Paired red and green images from the different focal planes were then acquired with Leica confocal software (version 2.61) and merged to produce either an image montage or a single composite image. In the photomicrographs shown in this study, a1E-ir is always presented as red staining, with the alternate neurochemical marker presented as green staining. Yellow immunolabeling in a merged image represents antigen co-Iocalization. Conventional epifluorescence microscopy was used to indicate the extent of immunolabeling and to provide a histological orientation of labeled structures within each of the layers examined. Utilizing rhodamine and FITC filter sets, images corresponding to higher-resolution confocal images were viewed with a 40x numerical aperture objective coupled to a Nikon Eclipse TE2000-U inverted microscope (Nikon, Kanagawa, Japan). Images were then captured using MetaMorph 6.2r2 (Universal Imaging Corp, Downingtown, PA). 62 Colchicine treatment in vitro. To enhance a1E staining within small intestinal myenteric cell bodies, LMMP sections were dissected and pinned tautly onto a silicone elastomer-lined petri dish. The preparations were then incubated in sterile cell culture media comprising Eagle’s minimum essential medium supplemented with fetal bovine serum (10%), gentamicin (10 jig/ml), penicillin (100 U/ml) and streptomycin (50 pg/ml) and colchicine (80 pM) (Sigma-Aldrich) for 24 h at 37 °C (Furness et al., 1989). The tissues were then washed with 0.01 M PBS (3 X 10 min intervals) and then fixed in formalin overnight at 4 °C in preparation for immunohistochemistry. Data analysis Cell counts of co-localized labeling in neurons. Co-Iocalized a1E/nNOS-ir cell bodies in one-hundred myenteric ganglia were counted in each of nineteen plexus preparations from four guinea pigs. These data were then exported to GraphPad Prism (GraphPad, San Diego, USA) for statistical analysis. A paired t-test, or a one-way ANOVA followed by a Tukey-Kramer post- hoc test was used to determine statistical significance. P < 0.05 was considered significant. The data are presented as mean :I: S.E.M. Quantification of co-Iocalization. Quantification of co-localization in merged images was assessed by a two-tiered approach: 1) line profiles, indicating pixel intensities, were drawn between structures deemed to have signal overlap (Zeiss Pascal LSM 5, Carl Zeiss lnc., Thornwood, NY); 2) object- 63 based analyses involving segmentation and thresholding techniques that separate and then classify varicosities as mathematically-defined individual objects (Silver & Stryker, 2000) were performed using Image Pro-Plus 4.5.1 (Media Cybernetics, Silver Spring, MD). A 7X7 pixel matrix high-gauss spatial filter (single-pass, strength 1) was applied to facilitate optimum signal-to-noise ratios. Figures were then imported into Adobe Photoshop CS 8.0 (Adobe Systems, Mountain View, CA) for further processing. Care was taken to ensure that ‘when the overall sharpness, brightness and contrast of images were adjusted, that such manipulations did not alter the information content of the original images. Images in this thesis/dissertation are presented in color. 64 RESULTS Western blot studies Western blot analysis (Figure. 3.1) demonstrated the presence of the 01E subunit (~170 kDa) in the myenteric plexus of the guinea pig ileum. The guinea pig heart and myenteric plexus showed multiple low molecular weight bands. The strong expression of the 170 kDa a1E band in the circular muscle and its absence from the submucosa is supported by the immunohistochemical studies described below. Summary of a1E-ir in the whole mount LMMP throughout the GI tract a1E—ir was observed in cell bodies and axons of gastric neurons within myenteric ganglia and interconnecting nerve bundles. or1 E-positive nerve fibers were abundant and diffusely distributed throughout the ganglia and interconnective fibers. This distribution pattern of a1E-ir was similar in all GI regions. In the small intestine (ileum and duodenum), no a1E-ir was detected in myenteric cell bodies. In the colon (proximal and distal), however, a1E-ir was present in cell bodies and axonal processes of some myenteric neurons. In all GI regions, strong a1E-ir was observed in nerve fibers supplying the circular smooth muscle. Longitudinal muscle nerve fibers in the tertiary plexus also exhibited a1E-ir, but this labeling was not as intense as that visualized in the circular muscle. 65 Summary of a1E-ir in ileal myenteric neurons maintained in primary culture Since R-type calcium channel currents were recorded from dissociated ileal myenteric neurons in primary culture, and because no a1E-ir was observed in myenteric cell bodies from the whole-mount LMMP in the guinea pig ileum, I investigated a1 E-ir in myenteric neurons in primary culture. While some neurons showed labeling throughout their cell bodies and axons, a1 E-ir within the cell bodies was predominantly intracellular. Co-Iabeling studies a1E is not found in the submucosal plexus aIE-ir was completely absent in the submucosal plexus in all regions of the gut. For example, I attempted to localize a1E-ir in submucosal ganglia of the ileum and proximal colon (Figure. 3.2) and no labeling was detected in any of these ganglia (Figure. 2A1 and C). Submucosal ganglia were revealed using an antibody raised against SP and numerous SP-ir nerve fibers were found in submucosal ganglia (Figure. 3.2A2). SP-ir but not a1E-ir was found in extrinsic sensory nerves supplying submucosal arterioles (Figure. 3.28). TH is the rate limiting enzyme in norepinephrine synthesis and TH-ir is a marker for sympathetic nerves supplying the gastrointestinal tract. TH-ir revealed that sympathetic nerves supply submucosal ganglia and arterioles but these nerve fibers did not express a1E-ir (Figure. 3.20, D, E). 66 Co-Iocalization of a1E-ir and SP-ir in the myenteric plexus SP-ir and a1E-ir were co-Iocalized in varicose nerve fibers in myenteric ganglia throughout the gut, but this was not a complete overlap. This result is illustrated with data obtained in the proximal colon (Figure. 3.3A, B) and gastric corpus (Figure. 3.3C1, CZ). Nerve cell bodies containing a1E-ir were also detected in myenteric ganglia of the proximal colon but SP-ir was not detected in these neurons (Figure. 3.38). a1E-ir and SP-ir were also co-localized in nerve fibers in myenteric ganglia (Figure. 3.3C1, CZ and D) and in varicose nerve fibers supplying the longitudinal muscle layer (Figure. 3.3E) and circular muscle layers (not shown). In the duodenum, co-Iocalization of a1E-ir and SP-ir occurred in nerve fibers in ganglia, in internodal strands and in the tertiary plexus supplying the longitudinal muscle layer. The intensity of alE-ir labeling in the duodenal myenteric plexus was greater than that observed in the ileum, particularly in the interganglionic fiber tracts. I also identified some a1E-SP-ir in myenteric neurons in culture (Figure. 3.12C2, D2, and C3, D3). a1E-ir with calbindin and calretinin The calcium binding protein, calbindin, is the strongest marker for IPANs in the myenteric plexus (Song et al., 1991; Costa et al., 1996), while a second calcium binding protein, calretinin, is a marker for orally projecting interneurons and longitudinal muscle excitatory motorneurons in the myenteric plexus (Brookes, 2001). In whole mount LMMP sections, calbindin-ir was detected in myenteric nerve cell bodies in the stomach, ileum, duodenum and in the colon. 67 Representative images of calbindin-ir in nerve cell bodies are provided for the proximal colon (Figure. 3.4A, B). Although a1E-ir was readily detected in nerve fibers and nerve cell bodies in the stomach and colon, I never detected co- localization of a1E-ir and calbindin-ir in nerve cell bodies in the whole mount LMMP. These data are illustrated by results obtained in the proximal colon (Figure. 3.4A, B). a1E-calbindin co-localization was, however, present in myenteric neurons in culture (Figure. 3.1201, D1). Calretinin-ir was detected in nerve fibers in myenteric ganglia throughout the gut and co—Iocalized with a1E is some varicosities. These data are illustrated by results obtained in duodenal myenteric ganglia (Figure. 3.4C1, C2). Calretinin-ir was also detected in myenteric neurons of the distal colon but a1E-ir was not co-localized in calretinin- ir neurons (Figure. 3.4D1, DZ). In culture, myenteric neurons showed some co- expression of ME and calretinin in cell bodies (Figure. 3.12C4, D4) but a1E- calretinin-ir labeling was intense in axons (Figure. 3.12C5, D5). a1E-ir and 5-HT-ir are not co-expressed in neurons 5-HT is contained in a subset of descending interneurons in the myenteric plexus and I used 5-HT as a marker for this class of neuron (Costa & Furness, 1982; Costa et al., 1982). a1E and 5-HT did not co-Iocalize in the myenteric plexus in any gut tissue examined or in dissociated neurons. a1E-ir was detected in varicose nerve fibers which formed distinct ring-shaped structures surrounding non-a1E-ir cell bodies in the ileum (Figure. 3.5A1). Varicose nerve fibers in internodal strands also showed intense and extensive a1E-ir. There 68 were relatively few 5—HT-ir nerve fibers but 5-HT-ir in these individual nerve fibers was intense (Figure. 3.5A2). Close analysis revealed intense a1E-ir and 5-HT-ir labeling of varicose nerve fibers within myenteric ganglia in the duodenum and 5- HT labeling was more prominent in the duodenum than in the ileum. It is possible that myenteric neurons in the duodenum contain more 5-HT than other regions (Meyer & Brinck, 1999). However, 5-HT-ir and a1E-ir were contained in separate populations of nerve fibers (Figure. 3.581, 82). In the corpus, strong a1E-ir was present in nerve fibers in myenteric ganglia and in interganglionic connectives. In those areas, nerve fibers displaying 5-HT-ir were more abundant than in the small intestine as reported previously (Mawe et al., 1989). a1E-ir was detected in cell bodies and axonal processes in myenteric ganglia and internodal strands of the proximal and distal colon but a1E-5—HT co-localization was never observed (not shown). Co-Iocalization of a1E- and VlP-ir VlP-ir is localized to cell bodies and nerve fibers in myenteric ganglia, to nerve fibers in the tertiary plexus which supplies the longitudinal muscle and to nerve fibers supplying the circular muscle layer (Costa & F urness, 1983; Brookes et al., 1991b; Brookes et al., 1992). I found VIP-ir nerve fibers in ileal myenteric ganglia and in the tertiary plexus and that arrangement of a1E-ir was very similar to that of VIP-ir in these regions (Figure. 3.6A1, A2). Closer examination of ileal myenteric ganglia revealed evidence of some co-localization of VlP-ir and a1E-ir in varicose nerve fibers (Figure. 3.681, 82). Similar overlap of VlP-ir and a1E-ir 69 in varicose nerve fibers was also detected in the tertiary plexus supplying ileal longitudinal muscle (Figure. 3.6C, D). Varicose nerve fibers in myenteric ganglia in the gastric corpus also displayed overlapping (11 E- and VlP-ir (Figure. 3.6E, F). Co-Iocalization of a1E-ir with NOS-ir l detected NOS-ir in myenteric neurons in the stomach, ileum and proximal colon (Figure. 3.7A1, 81, C and E1). However, a1E-ir labeling in neuronal cell bodies was only detected in the stomach (Figure. 3.7A2, 82) and colon (Figure. 3.7E2). In the gastric myenteric plexus I found 3.9 :I: 0.60 a1E- NOS-ir neurons per ganglion, based on counts done in 100 ganglia each from gastric myenteric plexus from four guinea pigs. There was an anal-oral gradient in the numbers of a1E-NOS-ir neurons in the colon as there were 5.7 :l: 0.5 neurons per ganglion in the proximal colon while there were 4.7 i: 0.6 a1-NOS-ir neurons in the distal colon (P < 0.05) (Figure. 3.8). NOS-ir neurons were prominent in ileal myenteric ganglia, but a1E-ir neurons were never detected in these ganglia (Figure. 3.7C). I detected co- localization of a1 E-ir and NOS-ir in nerve fibers in the ileal myenteric plexus (7D1, 702). However, there was not complete overlap of a1E and NOS labeling as l detected nerve fibers which were labeled by only one or the other antibody. Therefore, a1 E is a marker for a subpopulation of NOS-containing nerve fibers. I did not detect a1E-ir in myenteric nerve cell bodies in the ileum although I did detect varicose nerve fibers that contained a1E-ir and NOS-ir. It is possible 70 that a1E-ir does not accumulate in myenteric nerve cell bodies in this region of the gut. I used a colchicine-incubation protocol in an effort to increase a1 E-ir in ileal myenteric neurons. Colchicine disrupts axonal transport and should cause an accumulation of proteins in nerve cell bodies (Costa et al., 1996). However, a1E-ir was not detected in myenteric neurons even in small intestinal preparations that had been incubated with colchicine (not shown). Intense co- immunolabeling for a1 E and NOS was, however, observed in the cell bodies and axons of dissociated ileal neurons (Figure. 3.13A1-A3). Identification of excitatory and inhibitory muscle motorneurons l next asked whether a1E-ir is co-expressed with SP-ir, VlP-ir and NOS-ir in nerve fibers supplying the muscle layers. In all regions a1E- and SP-ir co- localized in many nerve fibers supplying the longitudinal and circular muscle layers. This result is illustrated by data from nerve fibers supplying ileal circular muscle (Figure. 3.9A1, A2). SP-ir in nerve fibers in the muscle layers is a marker for excitatory motor innervation (Brookes et al., 1991b; Brookes et al., 1992; Brookes et al., 1998). I therefore reasoned that the a1E—ir that did not contain SP-ir was present in inhibitory nerve fibers. VIP-ir and NOS-ir were used as markers for inhibitory nerve fibers (Costa and Furness, 1983; Brookes, 1993) and it was found that a1E-VlP-ir and a1E-NOS-ir were co-localized in nerve fibers supplying the muscle layers throughout the gut. This result is illustrated with data obtained in the circular muscle of the stomach (Figure. 3.98) and proximal colon (Figure. 3.9C). 71 Co-Iocalization of a1E-ir and TH-ir As discussed above TH-ir is a marker for sympathetic nerve fibers supplying the gut. Sympathetic nerves inhibit gut motility by inhibiting myenteric neuronal activity and myenteric ganglia are innervated by sympathetic nerve fibers. I next determined if the sympathetic nerve supply of the myenteric plexus expressed a1E-ir. I found that nerve fibers in the myenteric plexus that contained TH-ir also contained a1E-ir (Figure. 3.10A, B). In the myenteric plexus brightly fluorescent rings of a1E-TH-ir varicosities circled non-a1 E-ir cell bodies. lntensely labeled ME and TH varicosities were also found in ganglionic fibers in the myenteric plexus and internodal strands of the duodenum (Figure. 3.10C). The network of ganglia and interconnecting nerve fibers in the duodenum consistently demonstrated intense a1E-ir and TH-ir. In the stomach, a1E- and TH-ir were localized to varicose nerve fibers in the myenteric ganglia and interganglionic fiber tracts. In the gastric myenteric plexus, nerve fibers co- expressing a1E-ir and TH-ir were less dense than in the corresponding areas of the other gut regions. In the colon (Figure. 3.100), a1 E-ir and TH-ir were co- Iocalized in nerve cell bodies and nerve fibers in myenteric ganglia. The same pattern of immunohistochemical labeling applied to the distal colon, where both the a1E and TH antibodies labeled nerve cell bodies and varicose nerve fibers. There were more a1E-TH-ir neurons per myenteric ganglion in the proximal colon (2.1 1 0.06) compared to the distal colon (1.4 1 0.02, P < 0.05) based on counts in 100 ganglia in preparations from four guinea pigs (Figure. 3.11). 72 Furthermore, an interesting observation was that although not all a1E-cell bodies were TH-ir, all TH-ir neurons were a1E-ir. Co-Iocalization of a1E-synaptophysin-ir and SP-synaptophysin-ir Synaptophysin is an integral synaptic vesicle membrane protein believed to be involved in neurotransmitter release after phosphorylation by tyrosine kinases (Evans & Cousin, 2005). Interestingly, in the guinea pig ileum, a1E and synaptophysin did not co-localize in nerve fibers in myenteric ganglia or in the tertiary plexus (Figure. 3.13A2-C2; A3-C3). Very rare co-expression of these two proteins, however, was found in some varicosities. Prominent SP-synaptophysin- ir was observed in varicosities in myenteric ganglia and longitudinal muscle in the tertiary plexus (Figure. 3.13 A4-C4). DISCUSSION Identification of the chemical coding of functional classes of enteric neurons has been of great help in understanding how the ENS functions. The results reported in this study identified subtypes of neurons that express R-type calcium channels. This information will assist in identifying the contribution of R- type calcium channels to the control of gastrointestinal function. a1E is not expressed by submucosal neurons Submucosal neurons control water and electrolyte secretion by the gastrointestinal epithelium and they also control local blood flow by providing 73 vasodilator input to submucosal arterioles. Submucosal arteriolar diameter is also controlled by sympathetic nerves that supply these blood vessels and by extrinsic SP-calcitonin—gene related peptide (CGRP) containing sensory nerves (Vanner & Macnaughton, 2004). SP and CGRP are both vasodilator substances (Uddman et al., 1986). Neither the extrinsic nerve supply of submucosal ganglia nor arterioles (sympathetic and sensory) nor intrinsic submucosal neurons expressed a1E-ir. Although I was unableto localize or1E-ir in the submucosal plexus, l was able to confirm, using TH immunohistochemistry, the presence of periarteriolar sympathetic nerve fibers and sympathetic nerve fibers terminating in submucosal ganglia. Similarly, I was able to demonstrate the presence of SP-ir nerve fibers in submucosal ganglia and around submucosal arterioles. These data suggest that drug treatments which either block or facilitate the function of R-type calcium channels would have little effect on secretomotor activity or blood flow in the submucosa. a1E-ir is localized to cell bodies of myenteric plexus lPANs in the small intestine Intestinal IPANS contain calbindin and SP and they are the first neurons activated in reflex pathways responding to chemical and mechanical stimuli of the mucosa (Furness et al., 1999; Furness et al., 2004). In the small intestine, IPANS constitute approximately 26% of myenteric neurons and these same neurons have axons that project to other myenteric ganglia (Brookes et al., 1995). In the small intestine, I could not detect a1E-ir nerve cell bodies, even in 74 tissues that had been incubated with colchicine in an effort to enhance cell body content of a1 E-ir. Our calbindin antibody did not label nerve fibers although I was able to detect prominent a1E-ir in nerve fibers in the myenteric plexus. I could, however, assess the overlap of a1E with calbindin in small intestinal myenteric neurons maintained in culture. In these neurons, alE-calbindin-ir was co- Iocalized mainly within the cell soma. Very little co-staining was observed in axonal processes. This, in agreement with the study by Rugiero et al. (2002), suggests that R-type calcium channels may contribute to the tetrodotoxin- insensitive calcium current during the somal action potential in lPANs. SP-ir is contained in IPANS which send processes to myenteric ganglia and SP-ir is also contained in ascending interneurons (Brookes, 2001). Ascending interneurons in the guinea pig myenteric plexus constitute approximately 5% of the total number of neurons present (Brookes, 2001). Motorneurons will not have processes in myenteric ganglia so they will not contribute to SP-ir in the plexus (Costa et al., 2000). I found substantial co- localization of a1E- and SP-ir in myenteric ganglia. Calretinin, however, is also a marker for ascending interneurons in the myenteric plexus (Brookes et al., 1991a; Brookes, 2001) and I used an antibody against this protein to localize nerve fibers in the myenteric plexus. Within the 5% of ascending interneurons, both SP and calretinin share the chemical coding in this type of neuron. In doubled-stained preparations, 1 detected a1 E-ir and calretinin-ir nerve fibers, and found that a small population of myenteric nerve fibers expressed both a1E-ir and calretinin. Therefore, a1E-ir may be present in the nerve endings of this 75 subset of calretinin-containing ascending interneurons. Based on these data, I conclude that the a1E-ir nerve fibers in the small intestinal myenteric plexus arise from SP-containing and calretinin-containing excitatory interneurons. l was able to detect a1E-ir and calbindin-ir neurons in the gastric myenteric plexus but I never detected co-localization of these two antigens in the same neuron. I did not determine their co-expression in gastric or colonic myenteric neurons in culture. Similar to my data in the small intestine, I found that, in the gastric corpus, a1E-ir and SP-ir were co-Iocalized in nerve fibers in myenteric ganglia and in the muscle layers. SP-ir and calbindin are localized to a subset of gastric myenteric neurons but there are also SP-ir neurons which do not contain calbindin or calretinin (Furness et al., 1988; Reiche et al., 1999; Reiche & Schemann, 1999; Schemann et al., 2001). The SP-ir but calbindin/calretinin negative neurons are either a separate population of ascending interneurons or circular muscle excitatory motorneurons (Michel et al., 2000; Schemann et al., 2001). My data in the ileum show that while a1 E-ir co- localized with calbindin mainly in cell bodies in culture, a1E-ir was also present in calretinin-positive nerve fibers in the whole mount LMMP and in myenteric neurons in culture. It would therefore seem likely that aIE-ir nerve fibers in gastric myenteric ganglia are probably from ascending interneurons. I also show that alE-ir and SP-ir co-Iocalized in nerve fibers supplying the muscle layers in the stomach and these must arise from excitatory motorneurons. Therefore, R- type calcium channels may contribute to synaptic transmission in myenteric ganglia and to excitatory neuromuscular transmission in the gastric corpus. 76 In the colon, IPANS contain calbindin and SP-ir (Messenger & Furness, 1990; Lomax et al., 1999; Lomax & Furness, 2000) and I found numerous nerve fibers containing both a1E-ir and SP-ir suggesting that nerve fibers might arise from colonic IPANS. However, unlike in the small intestine I was able to detect both a1E-ir and calbindin-ir neuronal cell bodies in the colon myenteric plexus but these proteins were expressed in separate neurons. Therefore, the nerve fibers containing a1E-ir and SP-ir are unlikely to come from calbindin-containing IPANS in the colon. There is a second set of SP-ir neurons that contain choline acetyltransferase but not calbindin in the myenteric plexus of guinea pig colon; these are ascending excitatory motorneurons supplying the circular muscle (Lomax and Furness, 2000). The SP-orl E-ir nerve fibers l detected are likely to arise from this subgroup of neurons. These data suggest that R-type calcium channels would contribute to excitatory neurotransmission to the circular muscle layer in guinea pig colon. a1E does not co-Iocalize with 5-HT in the myenteric plexus 5-HT is contained in some myenteric neurons and by enterochromaffin (EC) cells (Blilbring & Gershon, 1967). The importance of 5-HT in mucosal sensory transduction and as a transmitter in the regulation of secretory and motor functions in the gastrointestinal tract is well documented. In this regard, 5- HT released by EC cells may activate extrinsic and intrinsic sensory, motor and secretomotor neurons, and effector cells such as enterocytes and smooth muscle 77 cells (Tonini, 2005). 5-HT is also a mediator of slow synaptic excitation in some myenteric neurons (Monro et al., 2005). In the small intestine and stomach, there are ChAT-5-HT-containing descending interneurons (Clerc et al., 1998; Meedeniya et al., 1998; Lomax & Furness, 2000; Reiche et al., 2000). In the stomach 5-HT-ir neurons also contain calbindin while 5-HT-ir descending interneurons in the distal colon are similar to the small intestine in that they contain calbindin (Lomax & Furness, 2000). I found that a1E-ir and 5-HT-ir were present in separate populations of nerve fibers in myenteric ganglia and interganglionic strands in all regions of the gut examined. Therefore, I conclude that R-type calcium channels are unlikely to contribute to 5-HT release from nerve terminals in the myenteric plexus. a1E is expressed by enteric inhibitory neurons NO and VIP are released from inhibitory motor neurons supplying the muscle layers in the gut (Murthy et al., 1995; Xue et al., 2000). I found that alE-ir co-Iocalized with VlP-ir and NOS-ir in cell bodies in myenteric ganglia, internodal connecting fibers and in nerve fibers supplying the longitudinal and circular muscle layers. a1 E-ir and NOS-ir were co-localized in myenteric cell bodies and varicosities in the stomach and colon, and to varicose nerve fibers in myenteric ganglia, internodal strands and the muscle layers. In the ileum, VIP and NOS are present in descending interneurons, and inhibitory motor neurons. VIP but not NOS is also co-expressed with ChAT in some descending interneurons (Brookes, 2001). In the duodenum, however, NOS-ir was not 78 detected in VIP-containing descending interneurons, but motorneurons supplying the muscle layers in the intestine contain these neurotransmitters (Clerc et al., 1998). Taken together, the localization of a1E with VIP and NOS in myenteric cell bodies and nerve fibers, and in dissociated myenteric neurons, has a number of implications: 1) a1E-VIP/NOS-ir varicose nerve fibers in the myenteric plexus of the small intestine and stomach arise from descending interneurons; 2) a1E- VIP-ir cell bodies in the colon and a1E-NOS ir neurons in the stomach and colon, are either descending interneurons or inhibitory motor neurons projecting to the circular and longitudinal muscle layers; 3) a1E-VIP-NOS labeling in nerve fibers supplying the muscle layers suggests that a1 E is present in the nerve endings of inhibitory motorneurons. These results therefore suggest that R-type channels are likely involved in inhibitory neurotransmission to the muscle layers and in synaptic transmission in descending pathways in the myenteric plexus. a1E-ir is expressed by sympathetic nerves supplying the myenteric plexus TH is a marker for sympathetic nerve fibers and most TH-ir in the gut is extrinsic in origin. The exception is the proximal colon where amine- containing myenteric neurons have been observed (Costa et al., 1971; Furness & Costa, 1971). In the present study I found that: 1) TH-ir was not present in myenteric neurons in the small intestine or stomach, but TH-ir co-localized with a1E-ir in nerve fibers in myenteric ganglia and interganglionic connectives; 2) the proximal colon exhibited a substantially greater number of TH-ir neurons than the distal colon, 3) both regions of the colon displayed a1E-ir/TH-ir nerve terminals. 79 Three separate subgroups of sympathetic neurons project to the gut. These are TH/NPY neurons that supply the intestinal blood vessels (Furness et al., 1983), TH/somatostatin neurons that target the submucosal ganglia and mucosa (Costa & Furness, 1984) and TH-containing nerve fibers that contain another neurochemical marker and supply the myenteric plexus. My data suggest that a1E is a marker of sympathetic neurons that target the myenteric plexus and these sympathetic neurons have the neurochemical code of TH/a1E. A recent study showed that a2-adrenoreceptors couple to and inhibit the R-type calcium channel by mediating presynaptic inhibition of slow and fast synaptic transmission in the ENS (Bian & Galligan, 2007). Therefore, R-type calcium channels may be important for regulation of sympathetic neural control of gut motility. a1 E-ir does not co-Iocalize with synaptophysin The absence of co-localization of a1E and synaptophysin in myenteric ganglia suggests a number of possibilities: 1) since synaptophysin constitutes only 7% of the total synaptic vesicle membrane protein (Evans & Cousin, 2005), alE may only be present on synaptophysin-negative varicosities; 2) the SNARE proteins syntaxin and SNAP-25, but not synaptobrevin, interact with N-type calcium channels by binding to the intracellular loop between domains II and III on the (11 subunit (Sheng et al., 1994; Sheng et al., 1996; Catterall & Few, 2008). Furthermore, N- and PIQ type calcium channels co-localize with syntaxin-1 (Catterall & Few, 2008). Putative a1E-ir varicosities may therefore contain a 80 neuronal vesicle population comprising of other synaptic vesicle membrane proteins such as synapsin, synaptotagmin (Wiser et al., 2002), or perhaps even an as yet unidentified vesicle membrane protein; 3) synaptophysin may not be responsible for regulating neurotransmitter release from R-type calcium channel- containing nerve terminals. This is supported by evidence in CA1 hippocampal pyramidal neurons where the amplitude of excitatory postsynaptic currents, probability of vesicle release and overall synaptic transmission were unimpeded in mice that were synaptophysin-deficient (McMahon et al., 1996). Those authors suggest that the function of synaptophysin may be a redundant one or that it may have a more subtle function not necessary for regulating exocytosis. This may indeed explain the surprising observation in my study that 0:1E and synaptophysin do not co-localize. It is unlikely that the intense and diffuse 0:1E-ir fibers in the myenteric plexus represent non-functional synapses or are completely non-varicose fibers, since I had observed definitive SP-synaptophysin punctate co-Iabeling in the LMMP (Figure. 3.13C4, D4). It would therefore be feasible in a future study to investigate the co-Iocalization of 0:1E-ir with other synaptic vesicle membrane proteins, or presynaptic plasma membrane—anchored proteins such as syntaxin or neurexin. 0:1 E-ir does not co-localize with pan-cadherin I did not observe any co-Iocalization of 0:1 E with the cell membrane marker pan- cadherin in neurons in primary culture or in the whole mount LMMP (Figure. 3.13A5-CS, A6-C6). Four auxiliary calcium channel subunits, B, 0:28, and y are 81 essential for the function of the 0:1 protein (Hoffman et al., 1994) and all are expressed in neural tissue (Castellano et al., 1993; Stephens et al., 1997). The [3 subunit is located intracellularly and binds to the interaction domain on the 0:1 subunit, while the 0:2 subunit is an extracellular protein. The 0:2 subunit is anchored to 0:1 via the integral membrane protein 8. The y subunit is also an integral membrane protein, but anchored directly to 0:1 (Catterall & Few, 2008). The 8 subunit enhances expression of the 0:1 protein and peak calcium current density at the cell membrane, presumably by trafficking a large number of functional calcium channels to the surface (Brice ef al., 1997; Hofmann et al., 1999; Bichet et al., 2000; Altier et al., 2002). Interestingly, the expression of specific isoforrns of the [3 subunit varies during postnatal development (Vance at al., 1998). In that study, 84 protein expression increased ten-fold, concomitant with rat postnatal (PO-adult) development. 0:28 subunits contribute moderately to cell membrane expression of 0:1 subunits, but y subunits have no effect on 0:1 membrane expression (Catterall & Few, 2008). In fact, 0:28 subunits strongly facilitate 0:1E expression and function by increasing the rate of voltage- dependent R-type calcium channel activation and inactivation (Qin et al., 1998). Together, co-expression of 0:28 and 8 calcium channel subunits with 0:1 increases surface expression of 0:1 to approximately 5 and 35% respectively (Williams et al., 1992; Klugbauer et al., 2003; Bernstein & Jones, 2007). Evidence suggests that the 8 subunit unlocks a signal motif on the HI 82 intracellular loop on the 0:1 subunit, thus preventing retention of the 0:1 protein in the endoplasmic reticulum (ER) (Bichet et al., 2000). The absence of co—localization of 0:1E with pan-cadherin therefore suggests that: 1) since the 0:1E subunit has a number of glycosylation and phosphorylation sites (eg. by protein kinase C on the I-ll intracellular loop), 0:1E may have been post-translationally modified, preventing it’s folding in the rough ER and subsequent association with the accessory subunits for targeting to the cell membrane; 2) the 8 subunit may not have access to the putative ER- retention signal and thus cannot direct the movement of the 0:1E subunit from this intracellular compartment to the plasma membrane; 3) the cytoplasmic 0:1E staining observed could represent immature forms of the pore-forming subunit. The 8 subunit would then function to restrict 0:1E trafficking to the membrane to prevent cytotoxic insults such as uncontrolled calcium entry; 4) alternatively, 0:1E may be a mature protein, but the 8 subunit has been phosphorylated by protein kinases A and C, or undergone palmitoylation (Dolphin, 2003a). Palmitoylation regulates neurotransmission pre and postsynaptically (eI-Husseini Ael & Bredt, 2002). This post-translational processing of the B (or even 0:28) protein would again mean ineffective trafficking of 0:1E to the plasma membrane. 0:1E would thus be localized mainly intracellularly; and 5) the cytoplasmic 0:1E staining in myenteric cell bodies in culture suggest that this may be a purely developmental occurrence. Incremental [34 expression occurs from P0 to adult in the rat brain (Vance at al., 1998), and it is possible that in newborn guinea pigs that 8 subunits 83 may not be in their mature conformations, and therefore unable to facilitate efficient membrane targeting of 0:1 E. SUMMARY AND CONCLUSIONS The 0:1E subunit of the R-type calcium channel is expressed by nerve fibers throughout the myenteric but not submucosal plexus in the guinea pig gastrointestinal tract. Based on an established neurochemical code for subclasses of myenteric neurons and myenteric nerve fibers in the guinea pig, I conclude that 0:1E-ir is found in the nerve terminals of sympathetic nerves supplying the myenteric plexus, ascending interneurons, inhibitory motorneurons supplying the muscle layers and the cell bodies of myenteric IPANS in the small intestine (Table 2). These observations suggest that it may be possible to selectively inhibit or facilitate neurotransmitter release from specific subsets of nerve endings using drugs which act to alter the activity of R-type calcium channels. 84 Table 3.1. Primary antibodies used to study the distribution of R-type calcium channels . Dilutio “PSI . . . Antibody n Lot No specres and Supplier’s Characterization References Source Western blot of membrane Rabbit extracts from rat brain (ACC-006 ideflgies péotggig :Bfl 99 Sagggga H, , ’ a an ~ a. ; 01E 1'300 AN'OZ £2232: Absence of staining after Fisher TE, l s r a el) ' preincubation with control 2000 ' antigen in rat cerebellar tissue. Mouse, 8 .- D ”3353’ 'T-Z‘”"°”"’lli"i K. . a in in - , lneyepi eiawoece aseos 28K 1200 113K486” (fig: extracts labels a single CD, 1994) Aldrich St protein of 28 kDa. Louis, MO) Goat Western blot shows a single Jacobowitz 50700582 (A31550 P grate? Of 319 ki’a'n l viiM Emil. . . . . . ’ rea sorp ion 0 ra co ica ins y , ca"et'"'" 120° 8 8333?; tissue with antibody and 1991; CA) ’ control antigen shows Winsky et al, absence of immunolabeling. 1989 Mouse Absence of staining after (A816007, incubation with antibody and 5-HT 1:200 522888 Abcam, antigen in cryopreserved izaafibggfi)? Cambridge, ventral mesencephalic ' MA) neurons- Rat, clone (84150103505 Complete crossreactivity for Substance P 1,200 240907 AbD ' Substance P inbrat brain Cue1llg7gt al, issues y gigs: radioimmunoassay NC) Vasoactive Sheep Pre-incubation with VIP Zhu et al intestinal 1:2000 25030911 (A81581, antibody and antigen 1997 ’ peptide (VIP) Millipore) abolishes immunostaining 85 Table 3.1 (cont’d) Neuronal nitric oxide synthase (nNOS) t200 Tyrosine hydroxylase (TH) 1:200 Pan-cadherin 1:200 Synaptophys 1,200 in ' LV144962 9 21071256 598458 32 K4862 Sheep (A81 529, Invitrogen); Rabbit (Dr. 8 Mayer, U of Graz, Austria) Mouse, clone TOHA1 .1 (MAB358, Millipore) Mouse, clone CH- 19 (A86528, Abcam) Mouse, clone SVP- 38 ($5768, Sigma) 86 Preadsorption of NOS antibody with antigen abolishes staining in rat frontoparietal cortex Recognizes a~60 kDa TH protein by Western blot in rat cerebral cortex. Recognizes a protein ~135 kDa by Western blot in mouse hippocampal Iysates. lmmunoblot shows a 38 kDa synaptophysin-positive band in rat brain extracts. Cauli B et al, 2004 Rolletschek A et al, 2001; Semenenko, F. et al. 1986 Bae GU et al, 2008; Grimm M0 at al, 2008 Eshkind and Leube,1995 250K0- 160KB — ‘4 105K0— Brain Heart Heart LMMP CM SM k J k J Y W Rat Guinea Pig Figure 3.1. Western blot analyses of 0:1E subunit expression in rat brain and heart, guinea pig heart, longitudinal muscle-myenteric plexus (LMMP), small intestinal circular muscle layer (CM) and submucosa (SM). Note the multiple spliced variants detected in. the myenteric plexus, and a prominent band near 170 kDa, the same size as that present in the rat brain and heart and guinea pig head. 87 A1 Ileum S.PlexusP 0:1E A2 Ileum S. Plexus SP 8 Ileum S Plexus 0:1E + SP C Colon S.PlexusP 0:1E+TH D Colon S. Plexus 0:1E+TH E Colon S Plexus 0:1E+TH Figure. 3.2. The a1E-subunit is not expressed in the submucosal plexus in the ileum and colon. A1. SP-ir (green fluorescence) and absence of a1E-ir in a submucosal ganglion (g) in the ileum. A2. SP-ir (arrow) is present in submucosal ganglia. B. A merged image showing that arterioles (bv) in the submucosa were SP-ir (arrowhead) but not a1E-ir. C. A merged image showing TH-ir nerve fibers in a colonic submucosal ganglion do not contain 0:1E-ir. D. A merged image shown TH-ir (arrowhead) in periarterial nerve fibers. E. A merged image showing a submucosal ganglion adjacent to an arteriole; periarterial and intraganglionic nerve fibers contain TH-ir but not 0:1E-ir (arrow and arrowhead, respectively). Scale bars = 200 am (A1-A2); 50 pm (BC); 20 pm (D) and 30 um (E). 88 Figure. 3.3. Co-Iocalization of a1E- and SP-ir in nerve fibers in the LMMP in the colon, stomach and ileum. A. Merged (yellow labeling) low-magnification epifluorescence photomicrograph illustrating o1E-ir (red fluorescence) and SP-ir (green) co-Iocalization in varicose nerve fibers (arrow) with no co-labeling in nerve cell bodies (ch) in the proximal colon. The arrowhead indicates a1E-SP co- localization in nerve fibers supplying the longitudinal muscle. Inset show a1E-ir in a myenteric ganglion and in the tertiary plexus (tp) supplying the longitudinal muscle. B. Confocal image showing a1E-ir and SP-ir co-localization (yellow) in varicose nerve fibers in the proximal colon (arrow). 0:1E-IR nerve cell bodies are also visible (arrowhead). C1. Co—Iocalization of 0:1 E and SP-ir in varicose nerve fibers in the myenteric plexus in the stomach. Arrowheads denote examples of varicosities containing both antigens. The dotted line A-B corresponds to the line profile analysis shown in C2. D. Confocal image showing a1E-SP-ir co- localization in nerve fibers in an ileal myenteric ganglion. Arrowhead indicates a varicosity containing both antigens but fibers containing either 0:1E-fr or SP-ir are also visible. The inset is a preadsorption control where preincubation of the 0:1E antibody with the antigenic peptide (1pg antibody/1 pg peptide) abolished immunolabeling in the ileum LMMP. E. Nerve fibers of the tertiary plexus supplying the longitudinal muscle. Some nerve fibers (arrowhead in D) contained 0:1E- but not SP-ir. Scale bars = 200 pm (A); 100 pm (8); 18 pm (C1); 40 pm (D); 20 um (E)- 89 ALLERQXI L COLON a1F + SP 8 PROXIMAL COLON \ C1 STOMACH C2 STOMACH . ' 300' _ 0:1E 250- _ SP N O O l Intensity 8 ‘ ‘ (34%ka . é . .8ng88. Distance (:lM) E ILEAL TP 90 Figure. 3.4. Localization of a1E-ir with calbindin-ir and calretinin-ir in the LMMP in the colon and duodenum. A. 0:1E-ir (red fluorescence) is contained in nerve fibers and nerve cell bodies (arrowhead) while calbindin (Calb; green fluorescence) is contained only in nerve cell bodies (arrows) in the proximal colon. B. Higher magnification image showing that 0:1E-ir (arrowhead) and calbindin-ir (arrow) are contained in different neurons in the myenteric plexus of the proximal colon. C1, C2. Images showing that a1E-ir (red) and calretinin (Calr; green) co-Iocalize in some varicosities in the duodenal myenteric plexus (CZ arrowhead). D1, DZ. 01E and calretinin also co-Iocalize in some varicosities (arrowhead in DZ) but not in cell bodies in the LMMP (arrow in DZ) in the distal colon. Scale bars = 100 pm (A); 50 pm (8); 50 pm (CD). 91 A Colon 0:1E +Calb. B Colon 0:1E + Calb. C1 Duodenum 0:1E B Duodenum 0:1E+Calr.. Colon 0:1E+Calr.. 92 Figure. 3.5. Localization of 01E and 5-HT-ir in the LMMP in the small intestine. A1, A2. Low-magnification epifluorescencephotomicrographs showing a1E-ir (red fluorescence) and 5-HT-ir (green) (arrowheads) are not co-localized in nerve fibers in myenteric ganglia (g) and internodal strands (is) in the ileum. Note the varicose clusters of a1E-ir that surround the neuron in (A). BI. Confocal image showing 0:1E-ir (arrowhead) and 5-HT-ir (arrow) in separate nerve fibers in the duodenum. The 5-HT-ir nerve fibers can be seen coursing through the myenteric ganglia independent of a1E-ir fibers. Similar results were obtained in the stomach ileum and colon. 82. The dotted line A-B indicates the pixel profile analysis from 81. This confirms that 0:1E-ir and 5-HT-ir are localized in different nerve fibers in the myenteric plexus. Scale bars = 100 pm (A-B); 14 pm (C). 93 Bl Duodenum 0:1E+5-HT 94 Figure. 3.6. Co-Iocalization of a1E-ir (red fluorescence) and VIP-ir (green fluorescence) in the LMMP in the ileum and stomach. A1, A2. Low magnification epifluorescence images demonstrating 01E and VIP-ir in varicosities (arrows) in myenteric ganglia and internodal strands in the guinea pig ileum. Note the prominent a1E-VIP positive nerve fibers in the tertiary plexus supplying the longitudinal muscle layer (arrowheads). B1. Confocal image of co- Iocalized (yellow labeling) o1E-VIP-ir varicosities in the ileum. 82. Dotted line A- B is the line along which the pixel intensity was determined. Analysis of pixel intensities shows co-localization of 0:1E-ir and VlP-ir in this nerve fiber. C, D. Low-magnification epifluorescence photomicrograph illustrating a1E-VIP-ir in nerve fibers supplying the longitudinal muscle in the tertiary plexus (arrowheads). E, F. Epifluorescence (E) and confocal (F) images showing a1E-ir and VlP-ir co- localization (arrowheads) in varicose nerve fibers in a myenteric ganglion in the stomach. Scale bars = 200 pm (A1,AZ); 55 pm (8); 20 um (D-F). 95 A1 ILEUM 01E A2 ILEUM B1 ILEUM 01E+V|P C ILEALTP a1E+V|P 'D ILEALTP E STOMACH 01E + VIP F STOMACH 01 E + VIP 96 Figure. 3.7. Confocal photomicrographs showing co-localization of 0:1E-ir (red fluorescence) and NOS-ir (green) in the myenteric plexus. A1. NOS-ir in neurons in the gastric corpus. A2. the same ganglion contains 0:1E-ir neurons. A3. Merged (yellow) image of A1 and AZ showing overlap of 0:1E-ir and NOS-ir in gastric myenteric neurons. 81. High magnification images showing NOS-ir (arrowhead) localizes to myenteric neuronal cell bodies in the gastric corpus. BZ. 0:1E-ir (arrowhead) is present in the same neurons. B3. Merged image from panels 81 and 82 shows co-localization of 0:1E-ir and NOS-ir in nerve cell bodies (arrowhead). Varicosities containing 0:1E-ir without NOS-ir are also visible in these images (arrow). C. Neurons (arrowheads) containing NOS-ir in an ileal myenteric ganglion. D1. 0:1E-ir and NOS-ir are co-localized in pericellular varicose nerve fibers (arrows). DZ. Pixel intensity analysis through varicosities in D1. E1-E3: Co-localization of 0:1E-ir and NOS-ir in neurons in the myenteric plexus of the proximal colon. The inset in D3 illustrates a string of 0:1E-NOS-ir varicosities (arrowhead) in a myenteric ganglion from the proximal colon. Scale bars = 200 pm (AI-A3); 50 pm (8-E3); inset in D3 = 10 pm). 97 A1 81 C E1 STOMACH NOS A2 STOMACH NOS B2 ILEUM P. COLON NOS + 01E D1 « E2 STOMACH 01 E STOMACH NOS + O1E STOMACH 01E STOMACH NOS + (11 E ILEUM NOS + 01E P. COLON (11E E3 P. COLON NOS + G1E 98 0:1E-NOS-ir neurons/ganglion Stomach Proximal colon Distal colon Figure. 3.8. Mean number of neurons per ganglion obtained from 100 myenteric ganglia, 19 plexus tissue sections and 4 guinea-pigs. The proximal colon displayed the highest number of 0:1E-nNOS-ir neurons (5.7 1: 0.5), followed by the distal colon (4.7 :l: 0.6) and then the corpus (3.9 :t 0.6). There was a significant difference in the number of 0:1E-nNOS-ir neurons between the stomach and proximal colon. 99 A1 Ileum 0:1E+SP B Stomach 0:1E+VIP C Ileum 0:1E+NOS Figure. 3.9. Co-localization of 0:1E-ir (red fluorescence) with SP-, VIP- and NOS-ir (green fluorescence) in nerve fibers supplying the circular muscle layer. A1. 0:1E-ir co-localized (yellow) extensively with SP-ir in the circular muscle layer of the ileum. Inset shows a low magnification image of a1E-ir in the nerve plexus in the circular muscle layer. A2. Line A-B indicates a line profile analysis through the varicosities indicated by the arrowhead. B. 0:1E-ir nerve fibers in the circular muscle co-expressed VlP-ir in the stomach. C. 0:1E-ir and NOS-ir are co-localized in nerve fibers in the circular muscle layer in the proximal . colon. Scale bars = 20 pm. 100 Figure. 3.10. Co-Iocalization of a1E-ir (red fluorescence) with TH-ir (green) in the myenteric plexus. A. Merged (yellow) epifluorescence image showing co-localization of 0:1E-ir and TH-ir in varicose nerve fibers (arrowheads) in a myenteric ganglion in the ileum. B. Merged confocal image showing co- localization of 0:1E-ir and TH-ir in varicose nerve fibers in a myenteric ganglion in the ileum. Insets show individual images use to produce the main picture. C. 0:1E-ir and TH-ir co-localized in nerve fibers in ganglia (arrows) in the duodenum. Insets show individual images of 0:1E-ir and TH-ir. D. An 0:1E-TH-ir nerve cell body (arrowhead) and axon (arrow) in a ganglion in the proximal colon. Insets show un-merged images of 0:1E-ir and TH-ir. Scale bars = 200 pm (A); 20 pm (8); 16 pm (C) and 20 pm (D). 101 A Ileum 0:1E+TH B Ileum 0:1E+TH 0:1E u ' I" C Duodenum 0:1E+TH i _,Q t \ 102 0:1E-TH-ir neurons/ganglion Proximal colon Distal colon Figure. 3.11. The density of neuronal cell bodies co-localizing 0:1E and TH was higher in the proximal (4.2 :I: 0.6) than in the distal colon (2.8 :t 0.5), and this difference was significant at P <0.05. 103 Figure. 3.12. Distribution of 0:1E (red fluorescence) with calbindin, SP, calretinin, and VIP (green fluorescence) in dissociated myenteric neurons in the guinea pig ileum. A1. 0:1E-ir (red) in neuronal cell bodies was cytoplasmic in all myenteric neurons observed. B1. Calbindin-ir was observed only in some neuronal cell bodies. C1, D1. Merged image (A1, 81) showing co- Iocalization (yellow) of 0:1E- and calbindin in a neuronal cell body. AZ. 0:1E-ir strongly localized within the cell soma and to lesser extent in axons. BZ. Intense SP-ir was only detected around the periphery of myenteric cell bodies, and in axonal processes. C2, D2. This cell appears to have some co-localization (yellow, C2) at the periphery of the cell body. A3-D3. A cluster of three myenteric neurons, two of which are 0:1E-ir. The third unlabeled neuron is positive for neither 0:1E-ir nor SP. SP-ir was only present in neuronal processes. A4. Myenteric neurons immunoreactive for 0:1E. 84. Two calretinin-containing neurons show intense immunoreactivity in cell bodies. C4, D4. Calretinin-ir almost masks the effect of 0:1E-ir labeling in this image, but these two neurochemical markers do co-Iocalize within the cell bodies. A5-05. Prominent 0:1E-calretinin-ir in myenteric nerve fibers was observed, as evidenced by pixel intensity analysis (05). A6. Cytoplasm and nerve fiber immunostained for 0:1E- ir. 86. VIP-ir was only seen in the nerve fibers of myenteric neurons in culture. No VIP immunostaining within the cell soma was visualized. C6-DS. Although some co-localization for 0:1 E and VIP was present at the periphery of this 0:1E-ir cell, there are two distinct neurons present, one 0:1E-ir, the other VIP-ir. No or very little co-Iocalization is seen along the length of the axons. 104 is: Calbindin wCZ Merge ‘B4 Calretinin Merge (tp) C4 Merge A (11 E fibers B5 Calretinin fibers Figure. 3.13. Co-Iocalization of o1E-ir with NOS, synaptophysin, SP with synaptophysin, and 0:1E-ir with pan-cadherin in the guinea pig ileal whole mount (WM) LMMP and myenteric neurons in culture. A1-D1. Confocal images and line pixel intensity analysis demonstrating 01E and NOS-ir co- localization (yellow) in cell bodies and axons in myenteric neurons in culture. AZ- DZ. Confocal image showing an absence of co-localization between 01E and synaptophysin-ir. Insets in AZ and 82 are low-magnification epifluorescence photomicrographs of the LMMP showing intense but restricted ganglionic (g) synaptophysin-ir; and a1E-ir throughout the LMMP respectively. A3-D3. (11 E and synaptophysin do not co-localize in the tertiary plexus, but rare co-expression (arrow in C3) was sometimes observed. A4-D4. Synaptophysin and SP strongly co-Iocalized to varicosities in the myenteric plexus. A5-D5. In myenteric neurons maintained in culture, 01E did not co-Iocalize with the membrane marker pan- cadherin. A6-D6. In the LMMP, and with the rare exception of a few oIE-pan- cadherin-ir varicosities, co-Iabeling for 01E and pan-cadherin was not consistently detectable. 106 C1 Merge f 0 4 Q. A3 or1E (tp) C3 Merge (tp) A5 Pan-cadherin A2 Synaptophysin 9 B3 Synaptophysin (tp) A4 Synaptophysin 9 r 6 Pan-cadherin 107 82 01E B6 01E CHAPTER 4 CONTRIBUTION OF R-TYPE CALCIUM CHANNELS TO THE FUNCTION OF INTRINSIC PRIMARY AFFERENT NEURONS AND ASCENDING INTERNEURONS IN THE GUINEA PIG ILEUM MYENTERIC PLEXUS 108 ABSTRACT R-type calcium channels are expressed by subsets of myenteric neurons in the guinea pig ileum. In the present study, I used intracellular electrophysiological techniques preparation to determine the function of R—channels in myenteric neurons in the acutely isolated longitudinal muscle-myenteric plexus. I used the non-selective calcium channel antagonist, CdCIz (100 uM), the R-type calcium channel blockers NiCI2 (50 pM) and SNX-482 (0.1 0M), and the N-type calcium channel blocker w-conotoxin GVIA (CTX 0.1 uM). I studied the effects of these drugs on action potentials and fast and slow excitatory postsynaptic potentials (fEPSPs, sEPSPs) in S and AH neurons. CTX, but not NiClz, blocked sEPSPs in AH neurons. NICI2 (n = 14) and SNX-48Z (n = 6) inhibited purely cholinergic, but not mixed cholinergic purinergic, fEPSPs in S neurons. MCI; and SNX-482 reduced the duration and amplitude of action potentials in AH but not S neurons. NiCI2 also reduced the amplitude of the late afterhyperpolarization (n = 4). These data show that R-type calcium channels do not contribute to calcium entry into nerve terminals releasing mediators of sEPSPs recorded from AH neurons but they do contribute to release of acetylcholine as the mediator of fEPSPs in a subset of S-type neurons. These data indicate that R-type calcium channels may be target for drugs that could selectively modulate activity of AH neurons or could alter fast synaptic excitation is subsets of pathways in the myenteric plexus. 109 INTRODUCTION The enteric nervous system (ENS) is a division of the autonomic nervous system dedicated to local control of gastrointestinal (GI) function. The ENS exerts this control via intrinsic primary afferent neurons (IPANS), interneurons and motor neurons that reside in enteric ganglia within the gut wall. The myenteric plexus is located between the circular and longitudinal smooth muscle layers and controls GI motility. The submucosal plexus, found between the mucosa and circular muscle, facilitates absorptive and secretory responses (Costa et al., 2000; Brookes, 2001). Together, these neural networks produce a myriad of coordinated GI responses (Wood, 2006a). Efficient functioning of the GI tract therefore relies on the integrative behavior of the enteric neurocircuitry. Two types of myenteric neurons have been distinguished according to their electrophysiological properties: AH- and S-type neurons (Hirst el‘ al., 1974). AH-type neurons are the IPANs, and S-type neurons are interneurons and motorneurons (Furness et al., 1998). AH neurons have a prominent shoulder on the repolarizing phase of the action potential due to inward calcium currents through one or more voltage-gated calcium channels. Calcium influx generates a long-lasting (1-20 s) late afterhyperpolarization mediated by activation of intermediate conductance calcium dependent potassium (lK) channels (Furness et al., 2004). Trains of electrical stimuli applied to interganglionic nerve fibers evoke slow excitatory postsynaptic potentials (sEPSPs) that last seconds to minutes in AH neurons. A prominent mediator of the sEPSP is substance P (SP) acting at G-protein-coupled neurokinin-3 receptors (Galligan et al., 2000; Furness 110 et al., 2004; Wood 8: Kirchgessner, 2004). The action potential in S-type neurons is blocked by the voltage-activated sodium channel antagonist tetrodotoxin ('I'I'X) (Hirst et al., 1974). Single stimuli applied to interganglionic nerve fibers produce a fast EPSP (fEPSP) in S-type neurons. ACh, acting at nicotinic cholinergic receptors (nAChR), is the major excitatory neurotransmitter in S neurons. However, non-cholinergic fEPSPs have also been observed and these fast synaptic responses are mediated by ATP acting at P2X purinergic receptors (LePard et al., 1997; LePard 8: Galligan, 1999; Galligan, 2002a). Myenteric neurons express all high-voltage-activated calcium channels (L, N, PIQ, and R-type), but not low-voltage-activated T-type calcium channels (North 8: Tokimasa, 1987; Kirchgessner & Liu, 1999; Reis et al., 2000). Neurotransmitter release, contraction and secretory functions in the ENS are thus regulated by high-voltage-activated calcium channels (Smith et al., 2003). The dihydropyridine-sensitive L-type calcium channel is found mainly on GI smooth muscle cells, and is not an active participant in synaptic excitation within the plexus (Marino et al., 1993). L—, N—, P/Q- and R-type channels are present mainly on neurons and open at more depolarized voltages (~ -30 mV; maximum activation ~ 0 mV) and generally inactivate slower (> 100 ms) (Bean, 1989; Hoffman et al., 1994; Yamakage 8. Namiki, 2002). The R (resistant) current may activate at somewhat more negative voltages (~ -40 mV) and often inactivates more rapidly than the other high-threshold currents (Fox et al., 1987; Lorenzon & Foehring, 1995; Foehring et al., 2000). HVA channels subtypes are blocked by different drugs. N- and PIQ-type calcium channels are blocked by w-conotoxin 111 GVIA (w-CTX) (Barajas-Lopez et al., 1996; Chen & Kirchgessner, 2002) and 00- agatoxin IVA (w-ATX) (Tsien et al., 1988; Catterall, 2000; Bian et al., 2004) respectively. Low concentrations of NiClz (30-50 uM) and the spider toxin SNX- 482 (0.1-1 uM) block R-type calcium channels (Randall 8 Tsien, 1995; Newcomb et al., 1998; Wang et al., 1999; Tottene et al., 2000). R-type calcium channels have been implicated in neurotransmitter release at rat hippocampal mossy and associative-commissural fiber synapses (Gasparini et al., 2001), oxytocin release from rat neurohypophysial terminals (Wang et al., 1999), and excitation-contraction coupling (Wu et al., 1998; Albillos et al., 2000). While a role for R-type calcium channels in the central nervous system (CNS) has been established, the function of the R-type calcium channel in the ENS is not known. Bian et al. (2004) showed that 50 % of the calcium current in guinea-pig ileal myenteric neurons maintained in primary culture is carried by R-type channels. The R-type channel was also shown to be coupled to the az-adrenoreceptor in guinea-pig myenteric neurons, implying that R-type channels may be targets for sympathetic nerve-derived norepinephrine to regulate synaptic transmission (Bian & Galligan, 2007). The co-Iocalization of the R-type channel with SP, calretinin, vasoactive intestine peptide (VIP) and nitric oxide synthase (NOS) in varicose nerve fibers and cell bodies of calbindin- positive myenteric neurons in the guinea-pig ileum (Naidoo et al., 2006) strongly suggests that R-type calcium channels may mediate calcium entry into specific subsets of myenteric neurons. 112 The objective of my study was to determine the contribution of R—type calcium channels to synaptic transmission in the acutely-isolated longitudinal muscle-myenteric plexus (LMMP) preparations of the adult guinea pig ileum. The native synaptic connections present in the guinea pig intestine in vivo are preserved in this in vitro preparation. I therefore investigated whether calcium entry through R-type calcium channels contributes to synaptic transmission or action potentials in guinea pig ileum myenteric neurons. MATERIALS AND METHODS Tissue collection and preparation All animal use protocols were approved by the Institutional Animal Care and Use Committee at Michigan State University. Adult male Hartley guinea pigs (Z50-300 g, Bioport, Lansing, MI) were anesthetized with isoflurane (Abbott Laboratories, Chicago, IL), stunned by a blow to the back of the head and exsanguinated. A segment of the ileum approximately 10 cm proximal to the ileo- cecal junction was quickly removed and placed in pre-warmed (37 °C) oxygenated (95% 02; 5% C02) Krebs’ solution of the following composition (in mM): 117 NaCl, 4.7 KCI, 2.5 CaClz, 1.2 MgClz, 1.2 NaHzPO4, 25 NaHCOg, and 11 glucose. The Krebs’ solution contained scopolamine (1 pM) and nifedipine (1 pM) to block muscarinic receptors and L-type calcium channels, respectively. A 5 mm2 section of the ileum was cut open along the mesenteric border and pinned flat (Fine Science Tools, Foster City, CA) in a silicone elastomer-Iined (Sylgard, Dow Corning, Midland, MI) petri dish containing Krebs’ solution. The submucosa 113 and circular muscle were peeled away and the longitudinal muscle-myenteric plexus (LMMP) was transferred to a smaller silicone elastomer-lined perfusion chamber. The chamber was mounted on a stage of an Olympus microscope (Tokyo, Japan). The LMMP was pinned tautly and superfused continuously with warm (37 °C) Krebs’ solution at a flow rate of 4 ml/min. The preparation was equilibrated for 30 min before commencing intracellular recordings. Intracellular electrophysiological recordings Myenteric neurons in the acutely isolated LMMP were impaled with a 2 M KCI-containing glass microelectrode (Frederick-Haer, Brunswick, ME) with a tip resistance of 80-120 Mn. Membrane potential was recorded using an Axoclamp 2A amplifier (Axon Instruments, Foster City, USA) in bridge mode. Synaptic potentials were evoked focally with a monopolar glass electrode (tip diameter 60 pM) filled with Krebs’ solution (World Precision Instruments, Sarasota, Florida) by stimulating interganglionic nerve fiber tracts. The resting membrane potential (RMP) of the impaled neuron was allowed to stabilize for 10 minutes without applying intracellular holding current. Myenteric (AH and S-type) neurons were then classified electrophysiologically according to published criteria (Hirst et al., 1974; Morita & North, 1985; Wood, 1996). Short high frequency trains of stimulation (20 Hz, 500 ms), provided by a pulse generator (Master 8, A.M.P.I., Jerusalem, Israel) and a constant current stimulation unit (Grass Technologies, West Warwick, RI) were used to elicit slow excitatory postsynaptic potentials (sEPSPs). Single stimuli (0.5 ms duration, 0.1 Hz stimulus rate, 4-9 mA, 10 s 114 intervals) were used to evoke fast excitatory postsynaptic potentials (fEPSPs). When recording fEPSPs, the membrane potential was hyperpolarized to -90 mV to prevent fEPSPs from reaching action potential threshold. Signals were recorded using an analog to digital converter (Digidata 1322A, Molecular Devices Corp., Sunnyvale, CA) and Axoscope and Clampfit 10 software (Molecular Devices), and a desktop computer. Amplified signals were sampled at 2 kHz, filtered at 1 kHz and a digital average of 6 fEPSP sweeps was then obtained. The amplitudes of synaptic responses were measured under control conditions and after various drug treatments. Action potentials (APs) and late afterhypolarizations (late AHPs) from AH neurons, and APs from S-type neurons, were elicited by injection of brief intracellular depolarizing current pulses (2 ms duration, 0.1 Hz, 0.02-1 nA, 6V, at 10 s intervals) through the recording electrode (De Laet et al., 2002; Nurgali et al., 2007). AH neurons were identified by the prominent calcium shoulder on the repolarization phase of the action potential (Hirst et al., 1985). Myenteric AH and S-type APs and late AHPs were recorded at the resting membrane potentials of those neurons. I measured the amplitude, the half-maximum width, and the area under the curve (AUC) of the evoked APs and late AHPs. Half-maximum width was defined as the time at which the postsynaptic response is at half of its maximum value (Schneider, 2008). Neurons with RMPs less than -40 mV were excluded from the study. 115 Drug application Local drug application was accomplished using a quartz micropipette (30 -— 40 pm tip diameter) placed within 50 to 150 pm of the impaled neuron. Drugs were superfused near the impaled neuron and solution flow was gated using a VC-8 Valve Controller application system (Warner Instruments, Hamden, CT). The impaled neuron was exposed to drug treatments for at least 6 minutes before responses were recorded. Drugs Nifedipine, scopolamine, pyridoxaI-phosphate-6-azophenyl-2',4'- disulfonate (PPADS), NICIz, and mecamylamine hydrochloride were obtained from Sigma-Aldrich. w-conotoxin GVIA (CTX), SNX-482 and tetrodotoxin (TTX) were purchased from Alomone labs (Jerusalem, Israel). All drugs were diluted in deionized water except for nifedipine which was dissolved in 95% ethanol. Statistical analyses Data are presented as mean values :I: S.E.M for n neurons. Concentration- response data were fitted using non-linear regression and the Hill equation (GraphPad Prism 4, GraphPad, San Diego). Statistical differences between groups were analyzed with a t-test or one-way ANOVA with Tukey-Kramer post— hoc test comparisons. P < 0.05 was the criterion for determining statistical significance. 116 RESULTS CdCIz inhibits action potentials in AH but not S neurons To establish whether blockade of calcium channels would inhibit the calcium shoulder on the descending phase of the AH action potential, I applied the non-selective calcium channel antagonist CdCIz (100 pM) (Jun et al., 2004) to the perfusion bath. In all 5 neurons tested, the calcium shoulder was blocked by CdCl2 and this shortened the action potential half-width (Figure. 4.1A,B); CdCIz did not change action potential amplitude in AH neurons (Figure. 4.1A,C). After washout of CdClz, the voltage-gated sodium channel antagonist TTX (0.3 pM) was applied. TTX reduced action potential amplitude but not half-width (Figure. 4.1A—C). In all 5 neurons examined, CdClz did not affect the amplitude or half- maximum width of the action potential (Figure. 4.2). However, the action potential in S neurons was blocked by TTX (Figure. 4.2). R-type calcium channel blockers inhibit action potentials AH- but not S- neurons l next determined the effect of the selective R-type calcium channel blocker NlClz (50 pM) (Tottene et al., 2000; Gasparini et al., 2001) on the AH and S action potential. NiClz reduced the AH neuron action potential half width by 44% and the peak amplitude by 9% (n=12 P < 0.05; Figure. 4.3). However, NiCl2 (50 pM) did not change the action potential in S neurons (n=1Z, Figure. 4.4). These data provide pharmacological evidence that R-type calcium channels contribute to the action potential in AH but not S-type neurons. 117 Previous studies indicate that the calcium hump on the descending phase of the myenteric AH action potential is due to calcium entry through primarily N- type calcium channels (Franklin et al., 1992; Furness et al., 1998; Kang et al., 2003). Strong immunohistochemical evidence points to the presence of N-type channels in calbindin-containing myenteric IPANS (Kirchgessner & Liu, 1999). On this basis, a significant portion of the calcium current is likely to be the N-type current. Analysis of the AH action potential in 3 neurons demonstrated that NiCl2 (50 pM) and w-CTX (0.1 pM) did not affect the amplitude of the action potential (Fig. 4.5A,B). However, NICIz and w-CTX produced additive reductions in the action potential half width in AH neurons. At low concentrations, SNX-48Z (0.1 pM) is a specific antagonist of R-type calcium channels (Newcomb et al., 1998; Gasparini et al., 2001; Bian et al., 2004). l next determined whether SNX-482 inhibited action potentials in AH neurons. SNX-482 did not affect the mean amplitude of the AH action potential but it did reduce the action potential half-width (Figure. 4.SD,E). The subsequent addition of w-CTX, produced a further inhibition of action potential half width (Figure. 4.5D) and also reduced the peak action potential amplitude. These data indicate that : 1) NiClz and SNX-48Z produce an almost identical effect on the action potential in AH neurons; and 2) N- and R-type calcium channels are the predominant contributors to the action potential in myenteric AH neurons. Since CdClz and NiClz did not alter the action potential of S-neurons, I therefore did not test the effects of SNX-48Z or w-CTX on the action potential of those neurons. 118 NiCI2 inhibits the slow AHP in AH neurons As NiClz reduced the calcium shoulder on the action potential in AH neurons then the amplitude of the slow AHP should correspondingly decrease as the slow AHP is mediated by a calcium activated potassium channel. As anticipated, NICIz (50 pM) reduced the peak AHP amplitude by 42 % (Figure. 4.6A,B). NiClz also shortened the time course of the slow AHP as reflected by the reduction in the area under the curve (Figure. 4.6A,C) which was reduced by 46%. R-type calcium channels do not contribute to slow synaptic excitation of AH neurons. If R-type calcium channels contribute to the somal AH action potential, the next question I asked was whether R-type calcium channels contribute to release of the mediators of slow excitatory synaptic potentials (sEPSPs) in AH neurons. Enteric neurons use two means of communication, fast and slow chemical neurotransmission (Galligan, 2002b). Because neurotransmitter release from myenteric nerve terminals is regulated by voltage-gated calcium channels (Smith et al., 2003), I first ensured that blockade of all calcium channels was feasible. The non-selective calcium channel blocker, CdClz (100 uM), was used to block sEPSPs evoked by short trains (20 Hz 500 ms) of focal electrical stimulation applied to of interganglionic nerves strands, In all 5 AH neurons tested, CdClz (100 uM) largely blocked the sEPSP (Figure. 4.7A). The mean amplitude of the control sEPSP was 21.13 :I: 0.94 mV, while the mean amplitude in the presence 119 of CdClz was 1.2 :I: 0.3 mV. A complete concentration-response curve revealed a concentration—dependent calcium channel inhibition by CdCl2 (1 — 100 uM) (n = 5; i050 ¥—- 12 pM) (Figure. 4.73). The decline in sEPSP amplitude caused by CdClz was not related to a time-dependent decline in synaptic transmission because time controls revealed that sEPSP amplitude was stable overthe time course of the pharmacology experiments (Figure. 4.70). The effect of MCI; (50 pM) on sEPSP was studied in 16 AH neurons and it was found that NiCIz did not change sEPSP amplitude in any of these neurons (Figure. 4.8A,B). R-type calcium channels therefore do not contribute to slow synaptic excitation in AH myenteric neurons. To determine the voltage-gated channel subtype responsible for the sEPSP, I tested w—CTX (0.1 uM). In all 4 neurons tested with NiCIz resistant sEPSPs, w-CTX largely blocked the sEPSP (Figure. 4.9A,8). These results suggest that predominantly N-type calcium channels mediate sEPSPs in myenteric IPANS. R-type calcium channels do not contribute to cholinergic/purinergic fEPSPs in S-type neurons CdCI2 (1—100 pM) blocked fEPSPs recorded from S neurons in a concentration-dependent manner (Figure. 4.10A,B; n=5; IC50=10 pM). To demonstrate that this inhibition was due to calcium channel inhibition and not to a time dependent decline in synaptic transmission, time-control experiments were done where fEPSPs were evoked in drug-free conditions in 5 min intervals for 30 120 min. The amplitude of fEPSPs (n = 7; Figure. 4.10C) did not significantly change during this time period. Previous studies revealed the presence of three phannacologically distinct subsets of myenteric S-type neurons based on the mediators of fEPSPs in these cells. There are cholinergic fEPSPs (25% ‘of neurons), mixed cholinergic/purinergic fEPSPs (67 % of neurons), and cholinergic/serotonergic fEPSPs (11 % of neurons) (Galligan & Bertrand, 1994; LePard 8: Galligan, 1999; Zhou & Galligan, 1999). I therefore next asked whether R-type calcium channels contribute to fast synaptic transmission in one of more of these classes of S-type neurons. To answer these questions, I used NiCl2 (50 pM), the nAChR antagonist mecamylamine (10 pM), and the P2X receptor antagonist PPADS (10 pM) (Galligan, 20023; Ren et al., 2003) in an effort to block fEPSPs. Mecamylamine is a more potent nAChR antagonist than hexamethonium in myenteric neurons (Zhou et al., 2002). In a subset of 15 neurons tested, NiC|2 did not affect fEPSP amplitude (Figure. 4.11 A,B). Subsequent application of mecamylamine significantly reduced the fEPSP amplitude by 28 % (Figure. 4.11A,B; P < 0.05). Subsequent addition of PPADS further reduced the fEPSP amplitude by 77 % (Figure. 4.11A,B; P<0.001) compared to control. I then attempted to mimic the effect of NiCl2 (50 uM) with SNX-48Z (0.1 uM) on the fEPSP in 3 S-type neurons with mixed cholinergic/purinergic fEPSPs. SNX-48Z did not change fEPSP amplitude in these neurons but subsequent application of w-CTX (0.1 uM) reduced the fEPSP amplitude by 76 % (Figure. 4.12 A,8; P<0.05). These results indicate that calcium entry via primarily N-type, but not R- 121 type calcium channels mediate fEPSPs in cholinergic/purinergic S-type myenteric neurons. R-type calcium channels contribute to cholinergic fEPSPs in S-type neurons l next tested the hypothesis that the R-type channel is present at nerve endings using only acetylcholine as the fast synaptic transmitter. In 14 S-type neurons NiClz (50 0M) reduced the fEPSP by 23 % (Figure. 4.13A,B; P < 0.05). NICI2 was washed out and in these cells mecamylamine (10 uM) inhibited the fEPSP by 95 % (Figure. 4.13A,8; P <0.01). In the next experiments (n = 6), | first ensured that mecamylamine completely blocked the fEPSP. Then NiClz (50 pM) reduced the fEPSP by 45 % (Figure. 4.13C, P < 0.001) followed by w-CTX (0.1 pM) which inhibited the fEPSP amplitude by 42 % compared to NiClz alone (P < 0.05). I finally determined whether SNX—482 (0.1 pM) would again mimic the effect of MCI; (50 pM). In 4 neurons examined, mecamylamine completely inhibited the fEPSP (not shown). After washout of mecamylamine SNX 482 was applied and it reduced the fEPSP by 40 % (Figure. 4.14A,B; P<0.01). Addition of w-CTX (0.1 uM), then further reduced the fEPSP amplitude by 69 % (Figure. 4.14A,B; P < 0.01). 122 DISCUSSION The ability of the ENS to effect appropriate responses upon receiving different types of information is due to its functional interconnectivity and sensitivity (Furness & Costa, 1987). The manifestation of plasticity within enteric neural circuits as a consequence of distention or chemical stimulation of the mucosa, is an inherent property of the adaptive and integrative behavior of the ENS. The neurochemistry and synaptic connections of enteric neurons shape the function of the enteric neural circuitry. The synaptic connections focus attention on voltage-gated gated calcium channels in synaptic transmission in the ENS. Voltage-gated calcium channels regulate the functions of IPANS, interneurons and motorneurons, and tailor reflex circuits to motility and secretory patterns (Holzer et al., 1997). Electrophysiological studies demonstrated that ~50 % of the calcium current in ileal guinea pig myenteric neurons maintained in primary culture is the R-type calcium current (Bian et al.,2004). That study, however, did not identify the specific subsets of myenteric neurons from which the R—currents were recorded. lmmunohistochemical data later established the presence of R- type calcium channels in neurons and varicosities in the myenteric plexus in the adult guinea pig ileum (Naidoo et al., 2006). The present study provides the first evidence that R-type calcium channels contribute to the action potential in AH- type neurons, and to synaptic transmission in a subset of S-type neurons in the guinea pig LMMP. 123 The AH neuron action potential is N- and R-type calcium channel- dependent To reconcile the above descriptions of R-type calcium channels in the myenteric plexus, I first looked at the kinetics of action potentials from AH and S- type neurons. NiCl2 (50 pM) and SNX-482 (0.1 pM) did not affect the action potential amplitude or half-maximum width in S-type neurons. However, in AH neurons, both MCI; and SNX-482 significantly attenuated the mean peak amplitude and the half-maximum width in comparison with the control action potential. w-CTX (0.1 pM) largely inhibited the remainder of the calcium shoulder on the repolarization phase of the AH action potential. The calcium shoulder is due in part to activation of a TTX-insensitive N-type channel activated calcium conductance in myenteric AH neurons (Vogalis et al., 2001; Rugiero et al., 2002). Moreover, strong N-type, but weak PIQ-type calcium channel immunoreactivity was found in IPANs (Kirchgessner & Liu, 1999). PIQ-type calcium channels (Rugiero et al., 2002) and L-type channels do not appear to contribute to the calcium shoulder in guinea pig myenteric neurons (North & Tokimasa, 1987; Kunze et al., 1994). Interestingly, Rugiero et al. (2002) found a residual calcium current after blockade of N-, L- and PIQ-type channels, and postulated a contribution by the R-type channel to the somal action potential in AH neurons. My results demonstrate that when the R-type and N-type calcium channels are blocked in tandem, the duration of the AH action potential is reduced by approximately 95 %. F urtherrnore, when I examined the effect of R-type channels on the late AHP, I found that NiCIz inhibited the late AHP amplitude by 42 %. 124 Taken together, my results confirm that primarily N-type calcium channels (~ 54 %) regulate calcium influx during action potential in AH-neurons. However, I show for the first time that R-type calcium channels (~ 40 %) also contribute significantly to calcium entry during action potential firing in the somatodendritic region of lPANs. R-type calcium channels have been localized in neurons throughout the CNS (Gasparini et al., 2001; Lee et al., 2002; Dietrich et al., 2003; Weiergraber et al., 2008). In mouse hippocampal CA1 pyramidal neurons (Kavalali et al., 1997; Isomura et al., 2002; Yasuda et al., 2003) and amygdala neurons (Lee et al., 2002), R-type calcium channels participate in dendritic action potential generation. Inhibition of calcium channels in amygdala neurons heightened excitability and reduced the AHP amplitude in those neurons (Schiess et al., 1993; Schiess et al., 1999). My data indicate that in the ENS, R-type calcium channels may contribute to the AH neuron action potential by regulating somatic and dendritic excitability in lPANs. This is supported by previous evidence that R- type channels require strong depolarizations and that R-type currents have fast activation kinetics in rat cortical (Lorenzon & Foehring, 1995; Foehring et al., 2000) and guinea pig myenteric neurons (Bian et al., 2004). The late AHP maintains the resting conductance of AH neurons via voltage-insensitive IK channels that limit the firing rate of AH neurons to one or two action potentials in response to a depolarizing intracellular current pulse (Galligan et al., 2000; Furness et al., 2004). When R-type calcium channels on the somatodendritic region of lPANs were inhibited, the late AHP was correspondingly attenuated. R- 125 type calcium channels are therefore important contributors to the regulation of myenteric AH neuron excitability. R-type calcium channels do not contribute to slow synaptic excitation in AH-type neurons Slow synaptic excitation is mediated by G-protein coupled receptors that produce long-lasting slow-synaptic responses with a long latency. SP acting at neurokinin-3 (NK3) and ACh acting at M muscarinic receptors, are the principal mediators of slow excitatory neurotransmission in the ENS (Guard et al., 1988; Costa et al., 2000; Galligan, 2002b). A novel finding from the present study points to N-type rather than R-type calcium channels as the chief mediators of sEPSPs in AH myenteric neurons. This conclusion is based on the observation of an 89 % inhibition of the evoked sEPSP amplitude by w-CTX. N-type calcium channels predominate on autonomic nerve terminals (Mochida et al., 1996) while PIQ-channels are more prevalent on nerve terminals in the CNS (Takahashi & Momiyama, 1993). Those channel subtypes contribute to synaptic transmission in those regions (Hirning et al., 1988; Miller et al., 1988; Dolphin, 2003b). My data suggest that in the ENS, trains of stimulation could cause calcium entry through N-type channels to release SP. SP would then act at the postsynaptic NK3 receptor through protein kinases A (Morita & North, 1985; Bertrand 8: Galligan, 1995) or C (Guard et al., 1988; Bertrand & Galligan, 1995) to produce the sEPSP. 126 R-type calcium channels contribute to fast synaptic excitation in a subset of S-type neurons Since R-type calcium channels are not involved in slow synaptic excitation, l next investigated their role in fast synaptic transmission. In the ENS, fast synaptic transmission in the ENS is accomplished via ligand-gated ion channels where acetylcholine (ACh) acting at nAChRs, ATP at P2X receptors and 5-HT at 5-HT3 receptors are the predominant mechanisms of fast excitatory neurotransmission (Galligan, 2002a). My results demonstrate that nerve terminals releasing only acetylcholine as a fast synaptic transmitter in S-type neurons are sensitive to NiClz and SNX-482. Further, R-type channels are not present on nerve terminals releasing both acetylcholine and ATP as fast synaptic transmitters as amplitude of the fEPSP does not change in the presence of either NiClz or SNX-482. In both groups, w-CTX inhibited the fEPSP amplitude, suggesting that N- and R-type channels are intimately involved in modulation of neurotransmitter release from nerve terminals of fully cholinergic myenteric S- neurons. I believe that R-type calcium channels are localized to nerve terminals of fully cholinergic S-type ascending interneurons. Ascending interneurons use ACh acting at nAChRs and constitute 5 % of the total number of myenteric neurons in the guinea pig LMMP (Brookes, 2001). Descending interneurons use ACh acting at nAChRs, and ATP acting at P2X and PZY receptors. I concluded that R-type calcium channels are present on the nerve endings of ascending interneurons because those neurons were completely sensitive to mecamylamine (Galligan & 127 Bertrand, 1994; LePard et al., 1997; LePard & Galligan, 1999). In a second subset of S-type interneurons, PPADS (10 pM) blocked the ATP-mediated component of the mixed cholinergic/purinergic fEPSP (LePard et al., 1997). In fact, the fully cholinergic NiClz—sensitive S-neurons observed in my study may be calretinin-coded (Naidoo et al. to be submitted). LePard and Galligan (1999) surgically interrupted all nerve fibers to study the ascending and descending projections of myenteric interneurons. They found that ascending interneurons use only ACh for neurotransmission, whereas descending interneurons use both ACh and ATP. The involvement of R-type calcium channels in fast synaptic excitation in the CNS is conflicting (Gasparini ef al., 2001; Dietrich et al., 2003). Dietrich et al. (2003) showed that presynaptic R-type calcium channels do not participate in fast neurotransmitter release at the mossy fiber-CA1 synapse in adult mice hippocampal neurons. However, separate evidence from neonatal rat hippocampal neurons showed that R-type channels contribute to fast synaptic transmission, and may be important for regulating neurotransmitter release and modulation of synaptic plasticity (Gasparini et al., 2001). These contrasting observations may be due to species and/or developmental differences. Nonetheless, my data corroborate those of Gasparini and others (2001). At the mossy and associative-commissural fiber fast synapses, low concentrations of NICIz (50 uM) and SNX-482 (0.3-1 uM) inhibited the amplitude of excitatory postsynaptic currents (EPSCs) equally (Gasparini et al., 2001). The addition of SNX-482 followed by NiClz did not further inhibit the EPSC amplitude. However, 128 an antagonist of N- and PIQ-type calcium channels (to-CTX-MVIIC) significantly reduced the EPSC amplitude, indicating a role for R-type calcium channels in fast synaptic transmission. MCI; and SNX-482 are also mutually occlusive in myenteric neurons maintained in culture (Bian et al., 2004). In the present study, NiCIz (50 uM) significantly inhibited the fEPSP by 23 % (n = 14) and by 45 % (n = 6). This discrepancy could be because of low R-type channel expression in some of those (23 %-containing) S-neurons from which fEPSPs were evoked. SNX-482 attenuated the mean fEPSP amplitude by 40 %. Taken together, these results suggest that MCI,» and SNX-482 act specifically on the same pathway to inhibit fast synaptic transmission in S-type neurons. Whether NiCIz and SNX-482 are acting pre- or postsynaptically to inhibit the fEPSP remains to be determined as it is possible that R-type calcium channels also actively participate in synaptic integration. A synaptic interaction protein site (synprint) site important for interaction with the vesicular release machinery (synaptobrevin, syntaxin and SNAP-25; the SNARE complex) was believed to be unique to N- and PIQ-type but not R-type channels (Maximov et al., 1999). N- and PIQ-type channels physically and tightly couple to syntaxin 1A and SNAP-25 located on the nerve terminal membrane (Jarvis & Zamponi, 2001). R-type channels, however, may functionally interact with the SNARE complex at a putative unidentified interaction site (Bergsman & Tsien, 2000; Wiser et al., 2002; Cohen et al., 2003; Cohen & Atlas, 2004). This Dre-association of the R-type channel with synaptic proteins suggests these Calcium channels may be responsible for rapid synaptic vesicle exocytosis. My 129 data suggest that R- and N-type calcium channels spatially co-operate to facilitate efficient neurotransmission in fully cholinergic myenteric S neurons. In fact, this co-operativity between R- and N-type channels has been observed at the mouse endplate (Urbano et al., 2003). R-type channels may be also located closer to the calcium sensor than N-type channels (Urbano et al., 2003), possibly to synaptotagmin for fast synchronous neurotransmitter release (Catterall & Few, 2008). Therefore, in myenteric S-neurons, a possible mechanism for synaptic transmission via R-type channels could be: 1) a train of action potentials depolarizes the (NiCIz-sensitive) S-neuron nerve terminal allowing calcium to enter though R-type channels, 2) calcium activates calcium/calmodulin- dependent (CAM) protein kinase II that then phosphorylates the cytoskeletal tethering protein synapsin on the SP-containing vesicle, 3) the vesicle is targeted to the docking site, 4) the low affinity calcium sensor synaptotagmin binds to syntaxin once calcium enters the cell, and exocytosis occurs. Taken together, R- type calcium channels may therefore contribute to fast synaptic excitation in the myenteric plexus of the guinea pig ileum. The question then emerges, why are R-type calcium channel localized to ascending interneurons and specifically to a fully cholinergic subset of S-type neurons in the ENS? Does the ENS, by utilizing several calcium channel subtypes, employ a compensatory mechanism to allow efficient synaptic transmission? First, the topographical organization of N- and R-type calcium channels on nerve terminals may be important to short- and long-term synaptic plasticity for the fine-tuning of information encoding. Calcium channels are 130 modulated by G87 subunits released from Gi/Go proteins (Dolphin, 2003b). R- type calcium channels may be coupled to the inhibitory az-adrenoreceptor (Bian & Galligan, 2007). Furthermore, the SNARE complex enhances efficient exocytosis if calcium channels are close to the active zone. These properties allow synaptic transmission to be precisely regulated (Catterall & Few, 2008). The presence of both N- and R-calcium channel subtypes may serve to enhance synaptic transmission for the effective coordination of gut effector system (GI smooth muscle, secretory glands and blood vessels) responses. Second, the guinea pig ileum contains only a single class of ascending interneurons, and these neurons provide synaptic input to other ascending interneurons and to excitatory circular and longitudinal muscle motorneurons (Brookes, 2001). It remains possible that the R-channel is present in ascending interneurons to ensure that transmission proceeds unimpeded. R-type calcium channels, located on purely cholinergic S-neurons could, by co-operating with N- type calcium channels, serve as “booster stations” for ACh release as ACh is the predominant mediator of fast synaptic excitation in the ENS. This is supported by data that R-type channels are more closely associated with the synaptic release machinery than N-type channels (Urbano el‘ al., 2003). R-type channels may not mediate fEPSPs through mixed cholinergic/purinergic S-neurons, presumably because the release of two transmitters, ACh and ATP, could increase the overall probability of successful neurotransmission. These neurons would rely on calcium entry via only N-type calcium channels for exocytosis. The functional significance of R-type channels in S-neurons may thus be one of redundancy. 131 SUMMARY AND CONCLUSIONS R-type calcium channels in the guinea pig ileum make a substantial contribution to the action potential in AH neurons. The calcium inflection on the downstroke of the AH action potential is mainly mediated by both N- and R-type calcium channels. R-type channel blockers significantly inhibited the fEPSP in ascending interneurons that release only ACh. N-type channels also contribute to fast synaptic excitation in ascending pathways. I did not detect R-type channel mediation of cholinergic/purinergic S-neurons. However, N-type channels were the primary contributors to the fEPSPs in those neurons. R-type calcium channels may therefore be important to control of cholinergic synaptic excitation. This could have profound implications for the overall function of the ENS as ACh predominates fast synaptic transmission and is one of the most important neurotransmitters in ENS neurons. 132 AH neurons Control CdCIZ I190 PM) TTX (0-3 PM) i I J ’ AA / IL.” / fl b- *_T —\..-_ W, ..-.._ __ _-..-..- ‘—-\ Figure. 4.1. CdCIZ inhibits action potentials in AH neurons. (A) A 2 ms intracellular depolarizing current pulse was used to evoke an action potential. The calcium shoulder (arrow in Control trace) on the repolarizing phase of the action potential was inhibited by CdCIZ. Tetrodotoxin (TTX, 0.3 pM) blocked the Na+ sensitive component during the depolarizing phase of the action potential to produce a smaller amplitude and slower calcium action potential. (8) Mean data 1: S.E.M (n = 5) showing that CdCIZ, but not TTX, reduced action potential half width (***, P< 0.01 vs Control and 'I'I'X). (C) TTX but not CdCIZ reduced action potential amplitude C“, P < 0.05 vs Control and CdCIZ). 133 S neurons Control CdCIZ (100 Ml) TTX (0.3 MA) I) . W l M: K L’M__ i \ 2...— —/ _'/—‘\ _f—'L fl Figure. 4.2. CdCIZ does not affect action potentials in S-type neurons. (A) CdClZ (100 pM) did not affect the action potential in S-type neurons while 'I‘I’X (0.3 uM), completely blocked the action potential in the same neurons. (8) Mean data :I: S.E.M (n = 5) from experiments similar to A. 134 AH neurons Control NiCIZ (50 pM) Washout .0. AN W” L... /J\L_,_ N... .fla—v-m -2.“ Mfl .._.. Figure. 4.3. NiC12 inhibits the action potential in AH neurons. Representative traces showing attenuation of the calcium shoulder (arrow in Control trace) on the repolarizing phase of the action potential in the presence of NiClZ. Mean data showing that NiCIZ shortened the action potential half-width (n = 12; *** P < 0.001) in a reversible manner. NiClZ also reversibly reduced the action potential amplitude * P < 0.05). Data are mean :I: S.E.M. n is the number of neurons from which AH action potentials were elicited. 135 S neurons Control NiCIZ (50 uM) Washout WI WI N k...___.._... ‘ L22... ' LOW“. .f L _, a” Figure. 4.4. Action potentials in S-type neurons are not sensitive to NiCIZ. Representative traces showing that the amplitude and half-maximum width of the S-type action potential are unchanged by NiCIZ. Mean data showing that NiClZ (50 pM) did not change action potential half-width or peak amplitude. Data are mean :I: S.E.M. n = 8 neurons. 136 AH neurons Control NiCIZ (50 till) crx (0.1 pill) fl) 3 /ll _ \ LL 1%.. A — ——.~.-_—-Aa Figure. 4.5. NiCIZ, SNX-482 and w-CTX inhibit the action potential in AH neurons. Representative traces showing that NiCIZ (50 uM) reduced the calcium shoulder. The residual shoulder was then completely blocked by w-CTX . (0.1 uM). Mean data showing that the action potential half-width after NiCIZ- treatment was smaller compared to control (n = 3; ** = P < 0.01). w-CTX further reduced the half-width (*** = P < 0.001 vs. control). NiCIZ and w-CTX did not affect the peak AH action potential amplitude (n = 3; P > 0.05). SNX-482 inhibited the action potential width at half-amplitude compared to control (n = 3; * = P < 0.05). w-CTX then further attenuated the half-maximal width (n = 3; ** = P < 0.01 vs. control). The peak amplitude was inhibited by w-CTX but not SNX-482. Data are mean :I: S.E.M. n is the number of neurons. 137 AH neurons Control NiCIZ (50 pM) Washout Figure. 4.6. The late afterhyperpolarization (late AHP) in AH neurons is inhibited by NiCIZ. Representative traces showing that the amplitude and half- maximum width of the late AHP are inhibited after application of NiClZ (50 uM). Mean data showing that the amplitude of the late AHP is reduced after addition of MCI; (50 uM) treatment compared to control (n = 4; * = P < 0.05). The late AHP width at half-amplitude was unchanged in the presence of NiClZ (n = 4; P > 0.05). NiCIZ significantly decreased the area under the curve (AUC) of the late AHP compared to the AUC of the control late AHP (n = 4; * = P < 0.05). Data are mean t S.E.M. n represents the number of neurons. 138 sEPSPs in AH neurons Control CdCI2 (100 M) Figure. 4.7. Effect of CdCIZ on sEPSPs in AH neurons. A train of electrical pulses (20 Hz, 500 ms) applied to an interganglionic connective produces a sEPSP in an AH neuron (NS = nerve stimulation). The sEPSP is inhibited by CdCIZ (100 pM). Concentration-response curve for sEPSP inhibition by CdCI2. Curves are non-linear fits of the Hill equation. The peak amplitude of the sEPSP was stable over time (P > 0.05; one-way ANOVA). Data points are mean :I: S.E.M. n = 5 neurons. 139 sEPSPs in AH neurons Control Figure. 4.8. NiCIZ (50 pM) did not inhibit sEPSPs. a sEPSP in an AH neuron is not inhibited by NiCl2. 140 sEPSPs in AH neurons Control %- NS NiCIZ (50 0M) Hm /I - - - — - - - 0 NS I CTX(0.1 ltM) 4!! NS Figure. 4.9. Sensitivity of sEPSPs to w-conotoxin (w-CTX) (100 uM). sEPSPs in AH neurons are not inhibited by NiCIZ, but are blocked by w-CTX. Data are mean 1: S.E.M similar to that shown in A. * = P < 0.05 (NiCIZ versus 00- CTX); ** = P < 0.01 (Control versus w-CTX). n = 4 neurons. 141 fEPSPs in S neurons Control CdCI2 (100 pM) Wash Figure. 4.10. Amplitude of fEPSPs after treatment with CdCI2 and during time controls. CdCI2 (100 pM) completely inhibited the fEPSP. Concentration- response curve for inhibition of the fEPSP by CdCI2 (n = 5). The peak amplitude of the fEPSP was stable over time (n = 7; P > 0.05). Data are mean :I: S.E.M. n is the number of neurons from which fEPSPs were recorded. 142 Cholinergic/purinergic fEPSPs in S neurons Control Figure. 4.11. NiCl2 does not inhibit fEPSPs in cholinergic/purinergic S-type neurons. Representative traces showing that in a subpopulation of S-type neurons, NiCl2 (50 pM) did not inhibit the fEPSP. Mecamylamine (10 pM) reduced the fEPSP. Subsequent application of mecamylamine and PPADS (10 pM) then largely inhibited the fEPSP. Mean data indicating that the amplitude of the fEPSP under control conditions is not different from the fEPSP amplitude after NiCl2 treatment (n = 15; P > 0.05). Application of mecamylamine (10 pM) reduced the fEPSP amplitude (n = 15; * = P < 0.05). Mecamylamine and PPADS (10 pM) then inhibited the fEPSP (*** = P < 0.001). Data are mean :l: S.E.M. n = 15 neurons. 143 Cholinergic/purinergic fEPSPs in S neurons Control \-——--—-- Figure. 4.12. SNX-482 does not inhibit fEPSPs in cholinergic/purinergic 8- type neurons. Representative traces showing that SNX-482 (0.1 uM) has no effect on the fEPSP amplitude. Application of w-CTX (0.1 uM) blocked the fEPSP. This population of S-type neurons is sensitive to inhibition by a combination of mecamylamine (10 pM) and PPADS (10 uM). Mecamylamine (10 pM) reduced the fEPSP amplitude (data not shown). Mecamylamine (10 pM) and PPADS (10 pM) then largely inhibited the fEPSP. Application of w-CTX (0.1 pM) blocked the fEPSP (n = 3; * = P < 0.05). Data are mean :I: S.E.M. n = 3 neurons. 144 Cholinergic fEPSPs in S neurons ‘ Control Mecamylamine (10 pl“) ~o- WMMO'-W n.“ ~- Wash Figure. 4.13. NiCl2 and w-CTX inhibit fEPSPs in purely cholinergic S-type neurons. Representative traces showing that in a sub-population of S-type neurons, NiC|2 (50 uM) reduced the fEPSP. Mecamylamine (10 pM) then completely blocked the fEPSP. Mean data indicating that the amplitude of the fEPSP after NiCl2 treatment is significantly smaller compared to control (* = P < 0.05). Mecamylamine 10 uM) reduced the fEPSP amplitude (*** = P < 0.001), and this was smaller compared to NiC|2 treatment (*** = P < 0.001). n = 14 neurons. (C) w-CTX (0.1 pM) further inhibited the fEPSP (trace not shown). 145 :\ Control |I \ ‘l \~ ----.,.' ....................... 'CTX (0.1 pM) Figure. 4.14. Effect of SNX-482 and w-CTX on purely cholinergic fEPSPs. Representative traces showing that SNX-482 (0.1 pM) reduces the fEPSP amplitude in purely cholinergic S—type neurons. w-CTX (0.1 pM) subsequently inhibited the fEPSP. Mean data showing that SNX-482 inhibited the amplitude of the fEPSP compared to control (n = 3; *** = P < 0.001). Mecamylamine (10 pM) then blocked the fEPSP (data not shown). After washout of mecamylamine, w- CTX (0.1 pM) blocked the fEPSP (n = 3; *** = P < 0.001). Data are mean 1: S.E.M. n = 4 neurons. 146 CHAPTER 5 R-TYPE CALCIUM CHANNELS COUPLE TO NITRIC OXIDE- MEDIATED INHIBITORY NEUROTRANSMISSION IN THE GUINEA PIG ILEUM 147 ABSTRACT Previous immunohistochemical studies showed that R-type calcium channels and nitric oxide synthase (NOS) co-localize in nerve terminals supplying the muscle layers in the guinea pig ileum. It is not known if calcium entry through R- type channels couples to NOS activation and/or ATP release. The present study determined the contribution of R-type calcium channel activation to release of NO and ATP from myenteric motorneurons supplying the muscle layers of the guinea pig ileum. I measured the effects of the NOS inhibitor, nitro L-arginine (NLA) and the R-type calcium channel blocker NiClz on neurogenic relaxations and contractions of longitudinal muscle-myenteric plexus muscle strips maintained in vitro. Neurogenic relaxations under histamine tone and contractions in non- histamine-contracted tissues were evoked by trains (20 Hz, 0.3 ms) of transmural electrical stimulation. I also used intracellular recordings from smooth muscle cells to investigate the contribution of R-type channels to release of ATP/B-NAD as the mediators of the inhibitory junction potential (IJP). Tetrodotoxin (TTX), and the non-selective calcium channel antagonist CdCI2 blocked electrically-evoked relaxations confirming that they were neurogenic. NLA inhibited neurogenic relaxations in a concentration-dependent (10-100 pM) manner; the maximum inhibition was 60%. NiClz produced a similar inhibition of neurogenic relaxations and contractions. Co-application of NiCIz and NLA did not produce a greater inhibition of neurogenic relaxations or contractions than either blocker alone. lJPs were blocked by TTX, CdCI2 and apamin. NiClz or NLA applied alone or together did not alter IJP amplitude or time course. R-type calcium channels may couple 148 to NOS activation and NO-mediated muscle relaxation in the guinea pig ileum. R- channels do not couple to release of the mediator of apamin-sensitive IJPs. ATP and NO released from the same inhibitory nerves may be regulated differently, or NO and other inhibitory neurotransmitters are released from separate populations of nerve fibers. 149 INTRODUCTION The enteric nervous system (ENS) controls gastrointestinal (GI) function (Furness et al., 1998). The ENS is composed of two nerve plexuses: the submucosal plexus which controls the absorption of nutrients, secretion, and local blood flow across the epithelium, and the myenteric plexus which has a pivotal role in the control of GI motility (GI) (Bornstein et al., 1994; Lomax & Furness, 2000). Non-adrenergic non-cholinergic (NANC) nerves mediate the majority of inhibitory responses in the ENS (Sanders & Ward, 1992). Stimulation of NANC inhibitory enteric neurons relaxes GI smooth muscle (Gao et al., 2006; Burnstock, 2008) and causes inhibitory junction potentials (lJPs) (Burnstock et al., 1964). Pharmacological and electrophysiological studies revealed that inhibitory neurotransmitters released from NANC nerves use varying mechanisms to cause smooth muscle relaxation (He & Goyal, 1993; Smits & Lefebvre, 1996; De Man et al., 2003). The most prominent inhibitory neurotransmitters in the ENS are ATP (Burnstock et al., 1997), vasoactive intestinal peptide (VIP) (Fahrenkrug et al., 1979; Rekik et al., 1996), carbon monoxide (Kwon et al., 2001), nitric oxide (NO) (Crist et al., 1992; Sanders & Ward, 1992; Mashimo et al., 1996) and possibly B-nicotinamide dinucleotide ([3- NAD) (Mutafova-Yambolieva et al., 2007). B-NAD/ATP, VIP and NO may all be released from the same motor nerve varicosity (Burnstock, 2008). ATP and or B-NAD, released from enteric motorneurons, is an important inhibitory enteric neurotransmitter that acts at Gq-protein-coupled P2Y1 receptors on Gl circular smooth muscle (Wood, 2006b). P2Y1 receptors mediate 150 the fast IJP (Gallego et al., 2006). MRSZ179 is a P2Y1 receptor antagonist (Boyer et al., 1998) that blocks the fast IJP in myenteric neurons (Gao et al., 2006; Wang et al., 2007). Apamin (0.1 pM) is a blocker of small-conductance calcium—dependent K‘ channels (SK) and selectively inhibits the fast IJP (Mackenzie & Burnstock, 1980). NO is an important signaling molecule in the central nervous system (CNS) and the ENS. NO can function as a retrograde (Holscher, 1997) transmitter, or as a conventional anterograde transmitter where it is dependent on the calcium-dependent activation of the neuronal isoforrn of nitric oxide synthase (nNOS) (Ohashi et al., 2007). In the ENS, nNOS is the dominant isoforrn of NOS (Ward et al., 1992) and NO is an important inhibitory neurotransmitter (Sanders & Ward, 1992). NO relaxes GI smooth muscle (Katsoulis et al., 1996; Mashimo et al., 1996; Kim et al., 1999). The slow IJP is NO-mediated (He & Goyal, 1993; Stark et al., 1993). Regulation of nitrergic neurotransmission however, is still not well understood. The identification of the NO regulatory mechanism is complicated by the fact that NO is a highly diffusible gas that is neither pre-formed nor stored in secretory vesicles (Bredt et al., 1990). Therefore its mechanism of regulation must be different from that which is used to control the release of vesicle-stored neurotransmitters. Regulation of readily releasable vesicles is consequently an important index of neurotransmitter release after each stimulation event (Rao et al., 2008). NO production and release may be regulated by a catalytically active pool of nNOS via PDZ binding 151 domains that allow nNOS to anchor to the plasma membrane at the active zone (Rao et al., 2008). Neurotransmitter release requires calcium influx through voltage-gated calcium channels (Catterall, 2000). R-type calcium channels are high-threshold channels and require strong depolarization for activation (Foehring et al., 2000) and ENS (Bian et al., 2004; Bian & Galligan, 2007). R-type voltage-gated calcium channels are expressed in myenteric neurons and together with the N-, PlQ-, and L-type channels, account for the total calcium current recorded from those neurons (Bian et al., 2004). The functions of R-type channels in myenteric neurons are not yet clear. In myenteric neurons maintained in primary culture, the R-type calcium channels may contribute to the somatodendritic action potential or to action potential-induced calcium entry required for neurotransmitter release from myenteric nerve terminals (Bian et al., 2004). Low concentrations of MCI; (50 (M) and SNX-482 (0.1 pM) block R-type calcium channels (Randall & Tsien, 1995; Newcomb et al., 1998; Wang et al., 1999; Tottene et al., 2000). w—conotoxin GVIA (w-CTX) (0.1 uM) and w-agatoxin IVA (0.1 pM) block N-type (Barajas-Lopez et al., 1996; Chen & Kirchgessner, 2002) and PIQ-type channels (Tsien et al., 1988; Catterall, 2000; Bian et al., 2004) respectively. Previous data has shown that the R-type calcium channel and nNOS co- localize in myenteric neurons, and in nerve fibers supplying the smooth muscle in the guinea pig ileum (Naidoo et al., 2006). The present study investigated whether myenteric neurons mediate descending inhibition via R-type channels, 152 and whether R-type calcium channels and nNOS are spatially coupled on nerve endings of enteric motorneurons to cause GI smooth muscle relaxation. 153 MATERIALS AND METHODS Ethical approval All animal use protocols were reviewed and approved by the Institutional Animal Care and Use Committee at Michigan State University. Longitudinal muscle-myenteric plexus (LMMP) preparation Adult male Hartley-strain guinea pigs weighing 250 g (Bioport, Lansing, Michigan; animals) were anesthetized under isofiurane inhalation (Abbott Laboratories, Chicago, IL), stunned and exsanguinated. The ileum was removed and placed in pre-warmed oxygenated (95% 002, 5% C02) Krebs’ solution of the following composition (mM): 117 NaCI, 4.7 KCI, 2.5 CaCIz, 1.2 M9021, 1.2 NaHzPO4, 25 NaHCOa, and 11 glucose. The Krebs’ solution contained scopolamine (1 pM) to block muscarinic cholinergic receptors. A glass cannula was inserted into a 2 cm segment of the ileum and the longitudinal muscle attached with the myenteric plexus (LMMP) was then teased off with a cotton swab soaked in Krebs’ solution. The LMMP preparation was mounted with silk Iigatures between a platinum foil electrode and a stationary hook connected to an isometric force transducer (Grass Instruments, FT03C, Quincy, MA, USA) and placed in a 20 ml jacketed organ bath containing Krebs’ solution at 37 °C. A resting tension of 1 g was applied to each of four 2 cm segments of LMMP preparations. Tissues were allowed to equilibrate for 60 minutes, during which the Krebs’ solution was replaced in 15 minute intervals. 154 Transmural electrical field stimulation (EFS) To study inhibitory mechanisms in the LMMP, each tissue was first pre- contracted with histamine (1 pM) to induce a sustained baseline resting tension (Osthaus and Galligan, 1991). Histamine acts directly at H1 receptors on the longitudinal muscle in the guinea pig ileum to cause an increase in muscle tone (lzzo et al., 1998). Non-adrenergic non-cholinergic (NANC)-mediated relaxations were induced by brief trains of transmural electrical stimuli (20 Hz, 0.3 ms pulse duration, 0.25 Hz train rate, 50 V) (Osthaus & Galligan, 1992). To determine the action of the antagonist on the inhibitory responses of the ileal longitudinal muscle, non-cumulative concentration response curves were created with a 15 minute interval between successive doses. I also investigated EFS-evoked frequency-dependent contractions in the non pre-contracted LMMP (20 Hz, 0.3 ms, 50 V). To determine the action of the antagonist on the excitatory response of the ileal longitudinal muscle, non-cumulative experiments were conducted and the effects of co-application of antagonists on the excitatory responses were studied. Transmural electrical stimuli in both set of experiments were provided by a Grass S48 stimulator and mechanical activity of the LMMP was recorded with Labscribe (iWorx, Dover, NH, USA) and a personal computer. Tissue preparation for intracellular electrophysiological recordings A 1.5 cm section of the ileum was cut open along the mesenteric border and pinned flat (Fine Science Tools, Foster City, CA) in a silicone elastomer-Iined (Sylgard, Dow Corning, Midland, MI) petri-dish containing Krebs’ solution. The 155 Krebs’ solution contained scopolamine (1 pM) and nifedipine (1 uM) to block muscarinic receptors and L-type calcium channels on smooth muscle cells, respectively. The submucosa was removed to expose the circular muscle- longitudinal muscle-myenteric plexus (CM-LMMP). A 5 mm2 section of the CM- LMMP was then transferred to a smaller silicone elastomer-lined recording chamber (2 ml) with the smooth muscle layer facing up. The chamber was mounted on a stage of a microscope (Olympus, Tokyo, Japan). The CM-LMMP was pinned tautly and superfused with oxygenated Krebs’ solution at a flow rate of 4 mllmin at 37 °C. The preparation was equilibrated for 40 min before commencing intracellular recordings. Intracellular electrophysiological recordings Single smooth muscle cells in the CM-LMMP were impaled with glass microelectrodes filled with 2 M KCI (tip resistance 80-120 Mn). Membrane potential was recorded using an Axoclamp 2A amplifier (Axon Instruments, Foster City, USA) in bridge mode. IJPs were evoked focally with a monopolar glass electrode (tip diameter 60 pM) (World Precision Instruments, Sarasota, Florida) filled with Krebs’ solution by electrically stimulating myenteric nerve fibers supplying the smooth muscle. Single pulses (1 Hz, 4 ms). Short frequency trains of stimulation (5 Hz and 10 Hz, 1 s, 50 V), provided by a pulse generator (Master 8, A.M.P.I., Jerusalem, Israel) and a constant current stimulation unit (Grass Technologies, West Warwick, RI) were used to evoke lJPs. The resting membrane potential (RMP) of the smooth muscle cell was allowed to stabilize for 156 10 minutes without applying intracellular holding current. Cells with RMPs less than -40 mV were not included in the analysis. Signals were recorded using an A/D converter (Digidata 1322A, Molecular Devices Corp., Sunnyvale, CA), Axoscope and Clampfit 10 software (Molecular Devices), and a desktop computer. Amplified signals were sampled at 2 kHz and filtered at 1 kHz. I measured the area under the curve (AUC) as this is an integration of both IJP amplitude and duration. This allowed us to study the changes in responses induced by drugs. AUCs were measured under control conditions and after various drug treatments. Drugs All chemicals were obtained from Sigma-Aldrich (St Louis, MO), except for w-CTX, w-ATX, and tetrodotoxin (TTX) (Alomone Labs, Jerusalem, Israel). Drugs were dissolved in deionized water with the exception of nitro-L-arginine (NLA), which was dissolved in hydrochloric acid (0.1 N) and nifedipine which was dissolved in 95% ethanol. Data Analysis Data are presented as the mean 1: S.E.M. For the EFS study, n refers to the number of animals from which tissues were obtained. For the intracellular IJP recordings, n is the number of smooth muscle cells. Concentration-response data were fitted using non-linear regression and the Hill equation (GraphPad Prism, San Diego). A one-way ANOVA, two-way ANOVA or Student’s t-test for paired or 157 unpaired data were used when applicable. P < 0.05 was considered statistically significant. RESULTS Electrically induced relaxations of the histamine-contracted LMMP Transmural stimulation of histamine (1 pM)-contracted LMMP preparations (n = 4) produced a triphasic response. This response was composed of a fast contraction (Figure. 5.1Aa, B1), followed by a relaxation (Fig. 1Ab, B1), and a rebound contraction (Figure. 5.1Ac, B1). I first tested whether these responses were neurogenic. The voltage-dependent sodium channel blocker TTX (0.3 pM) completely inhibited the relaxation and rebound contraction of the longitudinal smooth muscle but not the initial fast contraction (Figure. 5.1B2). The CdCI2 blocked EFS-induced contractions and relaxations (Figure. 5183, 2A). Apamin (0.1 uM) inhibited the EFS-evoked relaxation (Figure. 5.1B4) and revealing a large-amplitude slow contraction. R-type calcium channel blockers inhibit neurogenic relaxation of the LMMP Previous data showed that R-type calcium channels and NOS co-localize in the LMMP (Naidoo et al., 2006). The R-type calcium channel blocker, NiClz (0.1-100 uM), caused a concentration-dependent inhibition of EFS—evoked relaxation with a maximum 41 % reduction in the control response (P < 0.0001) (Figure. 5.2B). NLA (0.1-100 uM) also reduced the peak relaxation amplitude in a concentration—dependent manner with a maximum inhibition of 53 % (P< 158 0.0001). Co—application of NiClz with NLA (0.1-100 pM) inhibited the relaxation amplitude by 48 "/0. Comparison of these dose response curves revealed that NiCI2 did not significantly affect the dose response curve for NLA, and the converse was also true (P > 0.05). During these initial sets of experiments, I observed that NiClz and NLA inhibited the EFS-evoked relaxation and also produced a concentration dependent increase in the amplitude of the rebound contraction. I next studied the effect of MCI,» and NLA on the amplitude of this NANC contraction in the absence of histamine induced tone. Transmural stimulation (20 Hz, 0.3 ms) of the non-pre-contracted LMMP produced a biphasic contraction response. The biphasic contraction response at the beginning of the stimulation contained an early fast contraction followed by a late slow contraction (n=10, Figure. 5.3A1, B1). There was no evidence of a time-related change in the LMMP slow contraction (Figure. 5.3A2). I then tested the effect of MCI; (50 pM), NLA (100 uM) and apamin (0.1 pM) on this slow contraction. In these experiments I only measured the AUC (g.s) of the slow contraction because drug treatment had a pronounced effect on the amplitude and duration of the response. NiCI2 increased the AUC of the contraction by 147 % (Figure. 5.3B1, 382, 4A; P < 0.05). NLA in the presence of NiCIz increased the AUC by a further 13 % and this was not statistically different from NiCIz alone (Figure. 5.383; Figure. 5.4A P > 0.05). Subsequent addition of apamin with NiCl2 and NLA and increased the AUC by 53 % and this was statistically greater from the effect of NiCl2 by itself (Figure. 5.334; P < 0.01), and from the effect of MCI; in the presence of NLA (P < 0.05 159 Figure. 5.4A). TTX (0.3 pM) blocked the slow but not fast contraction while CdCI2 (100 pM) blocked both responses was applied to all LMMP preparations, and subsequently blocked the slow contraction (Figure. 5.3B5 and Figure. 5.6). In second set of experiments preparation where NLA was applied first, NLA increased the AUC of the contraction by 118 "/0 (Figure. 5.43, P < 0.01). NLA in the presence of NICIz did not produce any additional effect on the contraction (P > 0.05) when compared to NLA by itself. Co-application of NLA with NICI2 and apamin increased the AUC of the contraction by 37 % which was greater than the effect of MCI; and NLA together (P < 0.05). In the third set of experiments, apamin increased the AUC by 95 % (Figure. 5.40, P < 0.01). NLA in the presence of apamin increased the AUC by an additional 42 % (P < 0.01) while subsequent addition of MCI; did not produce any additional increase over that caused by apamin and NLA (Figure. 5.40). lJPs are NiCl2 insensitive l next investigated IJPs caused by focal stimulation of myenteric plexus interganglionic fiber tracts. Resting membrane potentials (RMPs) of impaled smooth muscle cells under control conditions was -51 1: 1 mV and did not change in the presence of drugs. Figures 5.5A1-A3 show control IJPs evoked in response to 1, 5 and 10 stimuli. I chose these frequencies because increasing electrical stimulus frequency causes the differential and additional release of enteric neurotransmitters from the same nerve endings (Bywater & Taylor, 1986; Crist et al., 1991; Undi et al., 2006; Benko et al., 2007). The amplitude and 160 duration of the IJP increased with increasing number of stimuli and NiClz (50 pM) did not affect lJPs (Figure. 5.5B1-B3). In the presence of NiCIZ, NLA (100 uM) also did not change the properties of the UPS (Figure. 5.5C1-03, H). In these cells, and regardless of the order of application, neither NLA nor NiClz by themselves affected lJPs (Figure. 5.5I, n = 4; 5 Hz and 10 Hz frequencies not shown). The selective purinergic P2Y1 receptor antagonist MRS2179 (10 pM) blocked the UPS at 1 Hz, 5 Hz and 10 Hz stimulation (Figure. 5.501-D3). TTX (0.3 uM) (Figure. 5.5 E) also blocked IJPs. CdCI2 (1-100 pM), produced a leftward-shift in the stimulus-response curve (Figure. 5.5F, I) and caused a concentration-dependent inhibition of the UPS (n = 4) at all stimulus frequencies. I therefore used 100 uM CdCI2 in subsequent experiments. While MCI; and NLA had no effect on the UPS they were blocked by CdCI2 (Figure. 5.50, H). Finally, application of apamin (0.1 pM) (Figure. 5.5G) also inhibited the IJP responses. Together, these data suggest that R-type calcium channels couple to NO synthesis and release but not ATP/B-NAD release as mediator(s) of the IJP. NiClz-insensitive lJPs are mediated by N- and PIQ-type calcium channels. In order to provide support for selective coupling of R-type calcium channels to NO release, and because CdCI2 inhibited the UPS, I used, together with NiCIz, w-CTX and w-ATX to inhibit N- and PIQ-type channels respectively. NiCl2 did not affect the UP (Figure. 5.6B1-B3, E) compared to control (Figure. 5.6A1-A3). Additive application of w-CTX (0.1 uM) in the presence of NiClz 161 reduced IJPs evoked at all frequencies (Figure. 5.601-03). Addition of w-ATX (0.1 pM) then further reduced the AUC (Figure. 5.601-D3). The single stimuli IJP was almost completely blocked, while at 5 and 10 Hz a residual IJP persisted. Apamin (0.1 pM) blocked the residual lJPs. Concentration—response curves with w—CTX (0.001-0.1 pM) and w-ATX (0.001-0.1 pM) revealed almost complete inhibition of UPS (Figure. 5.5I). These data indicate that N- and PIQ-type calcium channels mediate release of ATP/B-NAD as the mediators of the IJP. 162 DISCUSSION This study investigated the contribution of R-type calcium channels to inhibitory neurotransmission in the ENS of the guinea pig ileum. Enteric inhibitory nerves release NO and ATP/B-NAD to relax GI smooth muscle. ATP/B-NAD act at P2Y1 receptors on smooth muscle to cause apamin-sensitive inhibitory junction potentials (UPS) and relaxation. NO increases intracellular cGMP to cause muscle relaxation. I used transmural EFS in the LMMP and conventional intracellular recordings from smooth muscle cells in the ileum to show that R-type calcium channels couple to NO to mediate relaxation of GI smooth muscle. l have further demonstrated that N- and PIQ-type calcium channels control ATP/B- NAD release. R-type calcium channels couple to NOS release in the guinea pig LMMP I first investigated nerve-mediated relaxations of the guinea pig LMMP using transmural EFS in the presence of scopolamine. My first experiments used histamine which caused a sustained contraction of the longitudinal muscle. EFS then evoked consistent nerve-mediated relaxations. Three responses were observed after EFS: a fast contraction, a relaxation and then a rebound contraction. The relaxation response and the rebound contraction were nerve mediated as they were blocked by TTX. The initial fast contraction, however, was TTX-resistant. There are a couple of reasons for the TTX-resistant contraction. First, calcium entry through voltage-gated calcium channels probably does not cause this contraction since blockade of those channels still produces a TTX- 163 resistant contraction (Kristufek et al., 1999). Nicotinic ACh receptors (nAChRs) may be localized to nerve terminals of excitatory longitudinal muscle myenteric motorneurons (Galligan, 1999). nAChRs are non-specific ligand-gated cation channels that have a greater permeability to calcium than sodium and potassium ions (Vemino et al., 1992). In the presence of TTX, EFS would cause a depolarization of the nerve terminal and activation of presynaptic nAChRs to allow calcium entry via the nAChR, leading to neurotransmitter release (Sacaan et al., 1996; Lena & Changeux, 1997; Schneider et al., 2000). Second, it is possible that high-frequency (5 Hz, and 10 Hz) nerve stimulation in myenteric neurons still causes additional calcium entry through activation of voltage-gated calcium channels, different to that observed in sympathetic neurons (Kristufek et al., 1999). Calcium influx through voltage-gated calcium channels could directly cause neurotransmitter release, or indirectly by inducing calcium release from intracellular calcium stores within the smooth endoplasmic reticulum (Smith et al., 2003). The TTX-resistant contraction may therefore be a calcium-mediated contraction. Concentration-response curves with NiCI2 and NLA showed that whether these drugs were co-applied, or whether they were applied alone, they produced an equivalent effect on neurogenic relaxation. Osthaus and Galligan (1992) observed that LMMP tissues exhibiting biphasic nerve-mediated response of the histamine-contracted LMMP showed: a fast relaxation followed by a rebound non-cholinergic contraction, or a fast relaxation followed by a slower and smaller amplitude relaxation. My data, however, show a fast contraction, followed by a 164 slower and large amplitude relaxation, and a slow rebound contraction of large amplitude. These varying responses suggest that the interplay between excitatory and inhibitory nerves is complex and that neurotransmitters released from enteric nerves can produce several motor responses. I therefore next determined the contribution of R-type calcium channels to facilitation of LMMP contractions as a measure of simultaneous release of inhibitory neurotransmitters. MCI; and NLA, applied alone, increased the contraction and co-application of NICIz and NLA did not produce an additive effect. The addition of apamin however, significantly increased the contraction. Apamin applied alone increased the contraction and subsequent addition of NLA or NiClz further increased the contraction. These data suggest that R-type calcium channels couple to NO synthesis and release, but R-type channels are not coupled to ATP/B—NAD release. R-channels do not couple to ATP release as a mediator of apamin-sensitive lJPs I next characterized the relationship between R-type calcium channels and inhibitory neurotransmission by recording lJPs evoked from smooth muscle cells. NICIz and NLA added independently, and together, did not affect the UPS. However, apamin completely blocked IJPs evoked at all frequencies. Furthermore, MRSZ179 also completely inhibited the UPS. These responses were neurogenic as they were blocked by 'ITX. ATP, NO and VIP are released from the same enteric nerve endings on to GI smooth muscle, and UPS with 165 different kinetics have been observed (Mackenzie & Burnstock, 1980; Benko ‘et al., 2007; Burnstock, 2008). ATP produces a fast IJP, NO produces a slower IJP (Hirst et al., 2004), and VIP elicits very slow and tonic relaxations in the guinea pig ileal circular muscle (He & Goyal, 1993; Burnstock, 2008). In guinea pig jejunum and ileum, purinergic lJPs are either purely purinergic (26 %), partially purinergic (70 %) or non-purinergic (4 %) (Wang et al., 2007). In that study, apamin (0.5 uM) and MRSZ179 (2 uM) completely blocked the IJP. Wang and others (2007) used a single pulse (2 ms duration) whereas I used three different stimuli at 1 Hz, 5 Hz and at 10 Hz. In all cells tested, apamin and MRS 2179 completely blocked the IJP. This suggests that all UPS I had evoked were ATP/BNAD responses that are not mediated by R-type calcium channels. These NiClz-insensitive IJPs were largely blocked with w-CTX and w-ATX indicating that the IJP is ATP-mediated. At 10 Hz, a biphasic IJP was observed. This cannot be VIP or NO-mediated since apamin completely inhibited all evoked lJPs. ATP acting at P2Y1 receptors on smooth muscle causes the fast IJP (Crist et al., 1992; Gallego et al., 2006), thus corroborating my data. However, my results further indicate that the slower and broader duration IJPs (at 5 Hz and 10 Hz) are also be due to ATP/BNAD acting at P2Y1 receptors. Functional significance of R-type calcium channels in neuromuscular transmission Taken together, it is possible that R-type calcium channels may couple to NO release to mediate inhibition of GI smooth muscle. R-type channels and NOS 166 co-localize in nerve fibers that supply the smooth muscle layer. ATP/B-NAD released from the same nerve fibers would act at P2Y1 via N- and PIQ-type calcium channels. There is presently no marker for ATP localization. Therefore, whether R-type channels co-localize with ATP and NOS on the same or different myenteric motor nerve terminals remains to be determined. R-type channels may therefore be present on separate nerve fibers to regulate NO release from motor nerve varicosities, or they may be present on the same NOS-ATP containing fibers but couple to NOS for the differential regulation of NO release on to GI smooth muscle. Accumulating evidence indicates that all three major myenteric inhibitory neurotransmitters (NO, ATP and VIP) are released from the same motor nerve terminals (Costa et al., 1986b; Crist et al., 1992; Furness et al., 1995). It is therefore likely that R-type calcium channels are present with NO and ATP on the same nerve endings. Recent data from the mouse gut suggest a model in which enteric motor nerve terminals contain catalytically active forms of NOS (nNOSa) that are anchored via PDZ-binding domains (Lai & Lp, 2003; Kim & Sheng, 2004) to the plasma membrane (Rao et al., 2008). My data therefore suggest that NOS may be docked at the nerve terminal, close to R-type calcium channel release sites. A proposed mechanism would be as follows: when the nerve terminal is depolarized by invading action potentials, calcium, entering through R-type calcium channels, binds to calmodulin to activate NOS that converts L-arginine into NO. NO then diffuses across the neuroeffector junction, binds to soluble guanylate cyclase, activates 167 protein kinase G (PKG) which then phosphorylates myosin light chain kinase to cause relaxation of the muscle (Lecci et al., 2002).. R-type calcium channels play important roles throughout the nervous system. For example, at hippocampal CA3-CA1 synapses, R-type channels contribute to presynaptic calcium entry (Wu & Saggau, 1995). R-type channels are also involved in excitation-contraction coupling in the brainstem (Wu et al., 1998) and in chromaffin cells (Albillos et al., 2000). R-type calcium channels also contribute to fast synaptic transmission at hippocampal mossy fiber and associative-commissural synapses (Gasparini et al., 2001). In the ENS, R-type channels may be important to control of descending inhibition between myenteric motorneurons and effector tissues. NO and ATP are the most prominent inhibitory neurotransmitters in the ENS (Sanders & Ward, 1992; Lecci et al., 2002; Furness, 20063). It is possible that enteric motorneurons, by using multiple calcium channel subtypes, which in turn are coupled to important inhibitory neurotransmitters, ensure that transmission to the smooth muscle is achieved. The functional significance of multiple inhibitory neurotransmitters could therefore be a compensatory one, that is, deficiencies in either NOS/ATPNIP and/or a calcium channel subtype would be compensated, at least in part, by another inhibitory neurotransmitter or calcium channel (Mashimo et al., 1996; Mashimo et al., 2000). The same reasoning that multiple transmitters serve as protective mechanisms to ensure continual gut function can also be ascribed to the excitatory transmitters ACh and SP. This is due to the paucity of CNS innervation 168 to the ENS (Galligan, 2002b), and implies that the GI tract needs to be able to function in the absence of CNS input. SUMMARY AND CONCLUSIONS The present study demonstrated the involvement of R-type calcium channels in neurotransmission to the GI muscle layer. My data support a model in which R-type calcium channels couple to NOS to mediate NO release from myenteric motor axon varicosities to the smooth muscle. This mediation occurs via an apamin-insensitive mechanism. Therefore R-type channels do not couple to ATP/BNAD release, but rather to NO as a mediator of relaxation responses, demonstrated by EFS of LMMP preparations, and recordings of UPS from smooth muscle cells. All IJPs observed in this study were completely inhibited by apamin but not by MCI; and NLA, thus providing indirect evidence of a spatial relationship and coupling between R-type calcium channels and NO release. The cross-talk between R-type channels and NOS would need to be further evaluated as to whether there is a physical coupling between the channel and the enzyme, or whether R-type channels are linked to NOS through intracellular signaling mechanisms. This study helps us understand how the inhibitory signaling pathway is involved in neuromuscular control of motility via R-type channels. This may be important as the drugs that target R-type calcium channels would help to facilitate or inhibit neurotransmitter release form nerve endings of myenteric motor neurons. 169 Resting tension (Histamine, 1 pM) B1 Control relaxations B2 TI’X (0.3 pM) B3 CdCI2 (100 uM) B4 Apamin(0.1uM) W M Fig. 5.1. Representative traces showing NANC responses after transmural electrical field stimulation of the guinea pig LMMP in vitro. Relaxations were induced in the presence of histamine (1 pM) and scopolamine (1 pM). After nerve stimulation (NS), a fast contraction (a) was followed by a relaxation (b), and then a rebound contraction (c) (A, B1). While LMMP relaxations were abolished by TTX (0.3 (M), a fast contraction persisted. The entire response was, however, completely sensitive to CdCI2, a non-selective antagonist of voltage—gated calcium channels. Apamin (0.1 pM) blocked the relaxation. 170 A B d) .E 5 6°' n =10 - NiCI2 % 125‘ n =4 % so« I NLA g 100. .5 40.1 A NIC12 + NLA a .s g 75‘ %304 é 5* g zo- 92 25- «I g a 10‘ g o ' I E, c I I I I I OS 45‘ '5 '5 " °\o -8 -7 -a -5 .4 .3 [CdCI21(M) [Antagonist] (M) Figure. 5.2. Concentration-response curves for the area under the curve (AUC) of relaxation responses caused by EFS. CdCI2 caused a successive inhibition of the AUC (A). (B) When NLA was added first, it produced a concentration-dependent inhibition expressed as a percent of the relaxation response against the histamine contraction (P < 0.0001). The action of NLA was not statistically different in the presence of NiCI2 (P > 0.05). NiCI2, added first, also caused a concentration-dependent inhibition relaxation response (P < 0.0001). NiCI2 + NLA together had no further effect on the relaxation (P > 0.05). Data are mean 1: S.E.M. n is the number of tissues from which responses were obtained. 171 A. ** 20- ":10 ** *** 15- T 0.5- AUC (gs) slow contraction 3 l 0.0- - Control - NiC|2 (50 pM) - NiCI2 + NLA(1OO prl) I: NiC|2 + NLA+ Apamin (0.1 pM) * * B c 1.5- n = 10 l—‘——l .9 *** E T *-‘ ** *** c - 8 1.0 E: (I) 71? 0.5- 92 O D < 0.0- - Control - NLA(100pM) I NLA+ NiC|2 (50 pM) I::I NLA+ NiC|2 + Apamin (0.1 pM) Figure. 5.3. Effects of NiCI2, NLA, and apamin on LMMP contractions. Co- application of apamin + NLA + NiC|2 did not cause a significant increase compared to apamin + NLA (P > 0.05). 172 Figure. 5.4. Effect of NiCI2, and NiC|2 + NLA on inhibitory junction potentials (IJPs) evoked in guinea pig smooth muscle cells. NiCI2 (B1-B3, H), and NiCI2 + NLA (C1-C3) did not have any effect on evoked lJPs at all frequency levels of nerve stimulation (NS) compared to control (A1-A3). MRS 2179 (D1-D3), TTX (E) and apamin (G) completely inhibited the contractions. CdCI2 (F) reduced the LMMP contraction but did not block it. Concentration- response curves (I) for NiCI2 showed no effect on the UPS, whereas CdCI2 largely inhibited contraction responses. Data are mean :t S.E.M. from n muscle cells. 173 A1 Control (1 Hz) NiC|2 (1 HZ) ”If“ C1 NiC|2+NLA (1 Hz) ”'1 A2 Control (5 Hz) 82 NiC|2 (5 Hz) I‘M M C2 NiC|2+NLA (5 Hz) 174 B3 A3 Control (10 Hz) '1' 'l-II . \ C3 a. I l' 1" I/ b.311- Innull III I; 1ll NiC|2 (10 Hz) ) Figure. 5.4 (cont’d) D1 MRSZ179 (1 Hz) DZ MR82179 (5 Hz) DB MRSZ179 (10 Hz) E TTX (10 Hz) F CdCIZ (10 Hz) G ADamII) (10 Hz) _‘ __ II M "W w ”'Y' "' 17““ MW H 1 Hz, 5 Hz and 10 Hz nerve stimdation I 1 HZ nerve stimulation I Control (n = 5) I NiCI2 (n = 4) - w-c‘rx) (n = 4) I NiC|2 (n = 5) v co-ATX) (n = 4 :1 CdC|2 (n = 4) 30000 I NiCI2 + NLA (n = 5) 125 > g 100 E. g 75 5 < 50 3 33 < 25 I I J—-9-'8-7-'6-5-‘4-3 1 5 10 -25 Number Of stimuli (HZ) [Calcium Antagonist] (M) 175 1 Hz, 5 Hz and 10 Hz nerve stimulation - Control (n = 6) ll NiC|2 (n = 6) 250001 CI NiCI2 + CTX (n = 5) m - NiCI2 +CTX+ATX(n=5) r-——'I zoooo- .... ** E >' 15000-1 E. 0 10000- D < E O k Number of stimuli (Hz) Figure. 5.5. Effect of NiCI2, and NiCI2 + w-CTX and NiC|2 + w-CTX + w-ATX on lJPs. NiCI2 did not have any effect on evoked lJPs at all frequency levels of stimulation compared to control. Additive application of w-CTX reduced the AUC. Subsequent application of w-ATX largely blocked LMMP contractions at the 1 Hz frequency, and significantly reduced but did not block the contractions at the higher 5H2 and 10 Hz frequencies. This combined effect of NiC|2, w-CTX + w- ATX was similar to the result obtained after CdCI2 application. Data are mean :t S.E.M. from n muscle cells. 176 CHAPTER 6 INHIBITION OF R-TYPE CALCIUM CHANNELS IMPAIRS PERISTALSIS IN THE GUINEA PIG ILEUM 177 ABSTRACT R-type calcium channels are expressed by nitric oxide synthase (NOS) containing inhibitory motorneurons and some interneurons in the guinea ileum myenteric plexus. The goals of this study were to determine if inhibition of R-type calcium channels alters fluid induced peristalsis in the guinea pig ileum in vitro. Tetrodotoxin (0.3 pM), and the non-specific calcium channel blocker cadmium chloride (CdCI2, 1-100 pM) blocked peristalsis confirming that the reflex was neurogenic. NiCI2 (50 (M, which blocks selectively R-type calcium channels), reduced peristaltic contractions and efficiency (measured as fluid expelled). Additional treatment with the NOS inhibitor, NLA (100 (M) further inhibited peristalsis. NLA applied alone increased contraction amplitude, frequency and efficiency and additional application of NiCI2 inhibited peristalsis. w-conotoxin GVIA (0.1 “M, a concentration which selectively inhibits N-type calcium channels), partially inhibited peristalsis, subsequent addition of MCI; produced a further inhibitory effect. These data indicate that R-type calcium channels couple to NOS to regulate neurotransmission at neuroeffector junctions in the peristaltic reflex pathway. Selective inhibition of NO release from inhibitory motorneurons by NLA enhances peristalsis. While NiCI2 also inhibits NO release from motor neurons, it has additional sites of action which cause inhibition of peristalsis. NiCI2 and w-conotoxin GVIA produce additive inhibition of peristalsis indicating that R-and N-type calcium channels are localized to different components of the reflex circuitry mediating peristalsis. 178 INTRODUCTION Control of gastrointestinal (GI) motility is accomplished in part by interactions between the central nervous system (CNS), the enteric nervous system (ENS), myogenic and hormonal mechanisms (Grundy et al., 2006). The myenteric plexus is the division of the ENS that is primarily responsible for controlling GI motility. The peristaltic reflex is a fundamental motility reflex that propels content along the length of the gut (Gwynne & Bornstein, 2007). Peristalsis is a reflex that is initiated by distention of the intestine and it relies on the sequential activation of intrinsic primary afferent neurons (IPANs), interneurons and motorneurons in the myenteric plexus to produce oral muscle contraction and aboral muscle relaxation relative to the stimulus (Hennig et al., 1999; Huizinga et al., 2006). Acetylcholine and Substance P, released from excitatory longitudinal and circular muscle motorneurons, are the major excitatory neurotransmitters responsible for contraction of GI smooth muscle and the emptying phase of peristalsis (Paton & Zar, 1968; Bogeski et al., 2005). Relaxation of GI smooth muscle is mediated by non-adrenergic non-cholinergic (NANC) inhibitory enteric neurons and NO is a mediator of this relaxation (Sanders & Ward, 1992; Furness, 2000). Contraction of GI smooth muscle is dependent in part on activation of L- type calcium channels (Horowitz et al., 1996; Farrugia, 1999; Sanders, 2008). Although L-type calcium channels are also expressed in the ENS, they do not contribute to synaptic transmission (North & Tokimasa, 1987; Franklin & Willard, 1993; Reis et al., 2000). N, P/Q and R-type calcium channels are also expressed 179 in the ENS (Hirning et al., 1990; Kirchgessner & Liu, 1999; Smith et al., 2003). N-type channels contribute to the calcium shoulder during the repolarizing phase on the AH action potential and to synaptic transmission by (Franklin & Willard, 1993; Rugiero et al., 2002). N-type calcium channels are blocked by w- conotoxin GVIA (McCleskey et al., 1987). PIQ-type channels also contribute to neurotransmitter release in the ENS (Reis et al., 2000). w-Agatoxin IVA is a specific antagonist of PIQ channels (Mintz et al., 1992). R-type calcium currents have been recorded from guinea pig myenteric neurons maintained in primary cell culture (Bian et al., 2004). R-type calcium channels also contribute to calcium currents in duodenal myenteric neurons in the intact myenteric plexus (Rugiero et al., 2002; Furness, 2006a). Immunoreactivity (ir) to the pore-forming subunit (a1E) of the R-type calcium channel has been localized to subsets of myenteric neurons (Naidoo et al., 2006). a1E-ir was co-localized with nitric oxide synthase (NOS) in nerve fibers supplying the muscle layers and NO contributes to non-cholinergic non- adrenergic GI smooth muscle relaxation (Lecci et al., 2002). NO-mediated neurotransmission in particularly important for the peristaltic reflex (Ciccocioppo et al., 1994; Waterman & Costa, 1994; Holzer et al., 1997). However, the relationship between R-type calcium channel function, NO-mediated neuromuscular transmission and the peristaltic has not been studied. | used a modified Trendelenburg system (Billbring et al., 1958; Huizinga et al., 1998; Bian et al., 2003) to determine whether antagonism of R-type calcium channels affects fluid-induced peristalsis in the adult guinea pig ileum in vitro. 180 MATERIALS AND METHODS Tissue collection and preparation. All animal use protocols were approved by the Institutional Animal Care and Use Committee at Michigan State University. Adult male Hartley guinea pigs (250-300 g, Bioport, Lansing, MI) were anesthetized via isoflurane (Abbott Laboratories, Chicago, IL) inhalation, stunned by a blow to the back of the head and killed by severing the carotid arteries. A segment of the ileum approximately 10 cm proximal to the ileo-cecal junction was quickly removed and placed in pre-warmed (37 °C) oxygenated (95 % 02; 5 % C02) Krebs’ solution of the following composition (in mM): 117 NaCl, 4.7 KCI, 2.5 CaClg, 1.2 MgClz, 1.2 NaHzPO4, 25 NaHCOa, and 11 glucose. Assessment of peristalsis in vitro. Assessment of peristalsis in vitro was accomplished using a modified Trendelenburg preparation. A segment (6.5 cm) of the ileum was transferred to and pinned (Fine Science Tools, Foster City, CA) in a silicone elastomer-lined (Sylgard, Dow Corning, Midland, MI) petri dish containing warmed Krebs’ solution. Fire-polished ends of two glass cannulas (Radnoti Glass Technology, Monrovia, CA) were inserted into the oral and aboral ends of the ileum respectively, and secured firmly with 4-0 silk suture (Harvard Apparatus, Holliston, MA). The ileal segment was mounted horizontally in a plexiglass organ bath (volume 40 ml) containing Krebs’ solution and equilibrated at 37 °C for 45 min. The Krebs’ solution was replaced in 15 minute intervals during the equilibration period. Fluid-induced distention was achieved by raising the level of a Krebs’-containing reservoir connected by a plastic filling tube to the 181 glass tube inserted into the oral end of the ileum. The filling tube was connected to a pressure transducer (TRN 050; Kent Scientific Corp., Torrington, CT) via a T-connector, allowing intraluminal pressure increases to be monitored as the reservoir height was increased. Signals from the pressure transducer were digitized and acquired using an analog to digital converter (Minidigi 1A, Molecular Devices Inc., Foster City, CA), a DC strain gage amplifier (Grass Technologies, West Warwick, RI), and Axoscope v10 software (Molecular Devices Inc). The glass tube at the aboral end of the ileum was fitted with a one-way value to prevent backflow of Krebs’ solution. This tube drained into a graduate cylinder for measurement of the volume of fluid expelled during each episode (see below). Experimental protocol. After the 45 minute equilibration period, intraluminal pressure was first increased from a basal to a threshold (T) level. T was defined as the minimum pressure (mmHg) required for the initiation of peristalsis. Peristalsis was defined as a contraction that propagated from the oral to the anal end of the segment. Typically, these contractions occurred in regular intervals throughout the distention period. The pressure was elevated for 1 minute for each episode and then returned to baseline where the tissue was left to recover for 5 minutes. A range of increasing intraluminal pressures, in 0.5 mmHg intervals, from T to T+5 mmHg was used. This allowed us to calculate the half-maximal effective intraluminal pressure (EPso) for time control and antagonist concentration-response studies. Time control experiments were performed over a period of 50 min, and the Krebs’ solution was replaced every 15 min. I measured the amplitude of the peak contraction (mmHg). frequency of 182 contractions (waves/ 3), and the area under the curve (AUC; mmHg.ms) (Clampfit; Molecular Devices) during a 60 s recording period. The volume of fluid expelled (ml) at the end of each distention period was also measured. Antagonist concentration-response curves. Concentration-response curves were obtained by using the T+EP50 (mmHg) value at increasing antagonist concentrations for each segment of ileum. The non-selective calcium channel antagonist CdCI2 (1-100 pM), and the R— and N-type calcium channel antagonists NiCI2 (1—50 pM) and w-conotoxin GVIA (cu-CTX) (0.001-1 (M), respectively, were added to the bath. The tissue was incubated in successively higher antagonist concentrations for 10 min prior to testing its effect on peristalsis. Pressure-response curves. Pressure-response curves were obtained under control drug-free conditions and at specific antagonist concentrations using increasing intraluminal pressure values from T to T+5. The aim of the pressure- response curve experiments was to investigated the effects of NiC|2 (50 (M), the nitric oxide synthase inhibitor nitro-L-arginine (NLA; 100 pM), w-CTX, the sodium channel blocker tetrodotoxin (TTX; 0.3 pM), the small-conductance potassium channel antagonist apamin (0.1 pM) on the peristaltic reflex. All chemicals were purchased from Sigma-Aldrich (St Louis, MO), except for w-CTX and TTX (Alomone Labs., Jerusalem, Israel). Video imaging and spatio-temporal mapping. The construction of spatio-temporal maps was performed as previously described (de Backer et al., 2008). Briefly, a camera (Logitech, Fremont, CA), placed directly above the 183 organ bath, was used to capture peristaltic motor movements of the ileum over a 60 s recording period (25 frames/s, 640 x 480 pixel resolution). The video file was imported into lmageJ using a plugin for QuickTime v7 (www.gpple.com) (Figure 6.1A). A threshold procedure was applied to all video frames and edge- detection was used to define the gut periphery. Images were converted into binary format (Figure. 6.1B) and a region of interest was selected. Next, two further lmageJ plugins were used: Amplitude Profiler, which determines the change in intestinal diameter (mm); and Spatio-temporal Mapping software, which facilitates 3-dimensional (diameter distance and time) rendering of intestinal motility. I measured the minimum and maximum absolute diameters (mm) and the overall mean diameter (mm) of the ileum for a 60 s duration (de Backer et al., 2008). Statistics. Data are presented as mean values 1 S.E.M for n animals from which ileal segments were obtained. Concentration-response data were fitted using non-linear regression and the Hill equation (GraphPad Prism 4, GraphPad, San Diego). Statistical analysis was performed using either Student's t test for paired data, or a one-way ANOVA with Tukey—Kramer multiple comparison as a post-hoc test when more than two groups were compared. The effect of drugs on measured parameter responses at various intraluminal pressures between groups was assessed using a two-way repeated-measures ANOVA; post hoc multiple comparisons were corrected using Bonferroni's procedure. P < 0.05 was the criterion for determining statistical significance. 184 RESULTS Determining the EPso. To determine the EPso, I used, from a basal level, 0.5 mmHg step intervals in intraluminal pressure, ranging from T to T+5. I then calculated the EP50 by measuring the active area of the AUC which will incorporate frequency and amplitude of peristaltic contraction during each episode (Figure. 6.2A, B, C) during this range. The EP50 was 1.9 :I: 0.03 mmHg (n = 8). Using the EPso, l verified the stability of the reflex over a period of 50 minutes with peristaltic episodes at 5 minute intervals. There was no decline in the AUC or frequency of contraction during this period (Figure. 6.3A, B). Effects of calcium channel blockers on peristalsis. I also used the EPso to investigate the effect of CdCI2 on peristalsis. CdCI2 (1-100 pM) inhibited the peak amplitude of peristaltic contractions and their frequency. These changes are indicated by the concentration-dependent decrease in the AUC (Figure. 6.4) caused by CdCI2. Previous studies suggest that the N-type calcium channel may be the predominant calcium channel in the ENS (Takahashi et al., 1992; Smith et al., 1999). I therefore first investigated whether w-CTX could inhibit the peristaltic motor pattern. w-CTX (1—100 nM) produced a concentration-dependent inhibition of the peak contraction amplitude of 39.31 % (Figure. 6.4). Low concentrations of NiC|2 (50 uM) selectively blocks R-type calcium channels in the nervous system (Randall & Tsien, 1995; Tottene et al., 2000). NiCI2 (1-50 pM) also caused a concentration-dependent reduction in the peak amplitude, reducing it by 19.95 % (Figure. 6.4). The peristaltic reflex was impaired but not completely blocked by 185 either of NiC|2 or w-CTX. TTX (0.3 uM) completely inhibited the peristaltic reflex, thus confirming that this observed pattern of small intestinal motility was neurogenic through activation of myenteric neurons (data not shown). NiCI2 and NLA inhibit the peristaltic reflex Previous immunohistochemical data show that the pore-forming subunit of the R-type calcium channel, a1 E, strongly co-localizes with NOS in cell bodies and varicose nerve fibers in the myenteric plexus, and in nerve fibers innervating the circular muscle (Naidoo et al., 2006). I therefore hypothesized that R-type calcium channels may couple to NOS activation and NO-mediated muscle relaxation in the guinea pig ileum. To explore the role of a possible interaction between the R-type calcium channel and NO, I performed two sets of the following experiments where either NiCI2 (50 nM) or NLA (100 uM) was applied to the bath in combination with apamin (0.1 pM): 1) NiCI2 alone; NiCI2 + NLA; NiC|2 + NLA + apamin; and 2) NLA alone; NLA + NiCI2; NLA + NiC|2 + apamin. Peristaltic frequency. In experiments in which NiCI2 was added first, I detected a significant decline in the peristaltic frequency compared to the control response (Figure. 6.5, 6.6A). The frequency of contractions after co-application of NiCI2 and NLA compared to NiC|2-treatment alone was unchanged. However, when NLA was applied first, I detected a significant increase in the number of propagating peristaltic contractions at intraluminal pressures 1.9 to 2.9 mmHg 186 (Figure. 6.7, 6.8A). After co-application of NLA and NiCI2, the peristaltic frequency decreased below control values at 1.9 to 2.9 mmHg. AUC. l detected a significant decrease in the AUC at 1.9 to 2.9 mmHg when NiC|2 was added first (Figure. 6.5, 6.63). Subsequent application of NiCI2 and NLA did not change the AUC. In the presence of NLA first, however, the AUC increased. When NLA and NiCI2 were subsequently applied together, the AUC was reduced at 1.9 to 2.9 mmHg (Figure. 6.7, 6.8B). Volume expelled. The ileum emptied less when NiC|2 was added first, deviating from the control response at 1.9 to 2.9 mmHg (Figure. 6.6C). The addition of MCI; and NLA did not further decrease the volume of expelled fluid. When the ileum was treated with NLA first, the volume expelled increased over the control response at 2.4 and 2.9 mmHg. Subsequent application of NiC|2 with NLA significantly reduced the overall volume of fluid expelled at 1.4 to 2.9 mmHg (Figure. 6.8C). Mean Diameter. In the presence of NiCI2 and then NiCI2 with NLA, the mean diameter increased significantly at intraluminal pressures 1.9 to 2.9 mmHg (Figure. 6.6D). Spatio-temporal maps. Throughout this study, peristalsis, when evoked under drug-free conditions, was consistently robust with fully propagating motor 187 patterns (Figures. 6.5A1, A2; 7A1, A2; 9A1, A2, 11A1, A2). My interpretation of these motor patterns relied on the concurrent analysis of simultaneously- recorded ejection pressure waves (Figure. 6.5A1). In Figure. 6.5A1, the duration of the emptying phase of peristalsis was approximately 2.5 s. Figure. 6.5A2 is a spatio-temporal map depicting the motor patterns as cycles by characterizing the propagation of peristaltic waves in time and space. Examination of a control peristaltic reflex cycle (from the midpoint of one trough to the midpoint of the next trough; Figure. 6.5A2 line a-b ) on the map shows that the time taken for one cycle is approximately 5-6 s. This 5-6 s duration represents the emptying phase duration (see Figure 6.5A1). The spatio-temporal map therefore indicates that two fully propagating motor reflexes, each taking 2.5-3 s, traversed the ileal segment within a total time period of 5-6 s. It should therefore be noted that one cycle, as represented on my map, is a superimposition of multiple propagating motor patterns that summate to produce that cycle. When NiC|2 was applied first, the time taken for one cycle to occur steadily increased (Figure. 6.5B1, B2) until the propulsive motor reflex was replaced by regular small-amplitude oscillatory contractions. NiC|2 and NLA together, largely suppressed, but did not completely inhibit the reflex (Figure. 6.5C1, C2). NiCI2 and w-CTX inhibit the peristaltic reflex Because NiCI2 did not completely inhibit the reflex (see above), I investigated the contribution of other calcium channel subtypes to peristalsis. This would help us determine the extent of R-type calcium channel involvement 188 in modulation of the peristaltic reflex. I chose to study the interaction between NiCI2 (50 pM) and w-CTX (100 nM). I used an identical protocol to my NLA experiments to monitor the peristaltic motor pattern, that is: 1) NiCI2 first; then NiC|2 + w-CTX; and 2) w-CTX first; then w-CTX + NiC|2. Peristaltic frequency. When the ileum was treated initially with NiCI2 alone, I observed a decrease in the frequency of peristaltic waves, and threshold pressure was unaffected. The wave frequency only decreased significantly when NiCI2 and w-CTX were co-applied (Figure. 6.9, 6.10A). However, when w-CTX was applied first, the peristaltic reflex was impaired. In all tissues examined, peristalsis did not commence until at least 15 s after the desired level of pressure was attained; the initiation of the reflex appeared to be particularly affected. Co- application of w-CTX and NiCI2 then mostly suppressed the propulsive wave at all levels of intraluminal pressures tested (Figure. 6.11, 6.12A). AUC. Examination of the AUC revealed a similar sensitivity of the ileum to treatment with NiCI2 and w-CTX. Initial application of NiC|2 did not produce a significant decrease in the AUC, and the AUC was only reduced after incubation with both NiCI2 and w-CTX (1.4 — 2.9 mmHg) (Figure. 6.9, 6.10B). I further found that when w-CTX was added to the bath first, and that when the ileal segment was further incubated with both calcium channel antagonists, that this had a similar diminishing effect on the AUC (Figure. 6.11, 6.12B, also compare Figure. 6.90 with Figure. 6.110). 189 Volume expelled. In the presence of NiCI2, the volume of fluid expelled was not different from control values (Figure. 6.10C). When the ileum was incubated with both NiC|2 and w—CTX, the volume emptied declined significantly (0.9 — 2.9 mmHg). Reversibly, the application of w-CTX first, significantly reduced the fluid volume expelled (1.9 — 2.9 mmHg). Co-application of MCI; and w-CTX in both sets of experiments then caused a steep decline in fluid expulsion volume at 0.9 — 2.9 mmHg (Figure. 6.12C). Mean diameter. The presence of both NiCI2 + w-CTX (not shown), and w—CTX + NiC|2 caused a significant increase in the mean diameter of the ileum in both experiments (Figure. 6.120). This increase was particularly greater at higher intraluminal pressures (1.9 - 2.9 mmHg). Spatio-temporal maps. Analysis of spatio-temporal maps showed consistent regular peristaltic reflexes under drug-free conditions (Figure. 6.9A2). NiC|2 caused a progressive reduction in the strength of contraction and amplitude of each peristaltic wave. NiCI2 also doubled the time taken for each motor pattern to fully propagate from 2.5 s to approximately 4.5 s (Figure. 6.9B2), and this duration increased throughout the recording session. NiC|2 together with w-CTX completely abolished the motor pattern (Figure. 6.902). Peristaltic activity, however, visualized as propagating oscillations uniform in amplitude, was still 190 observed. Reverse peristalsis was also found, indicating that in the presence of NiC|2 and w-CTX, motility is not always aborally-directed. When I applied w-CTX first, threshold pressure was not altered. Propagating oscillations were again present, but only for approximately 30 s from the onset of fluid-distention (Figure. 6.11B1, B2). Moreover, these oscillatory contractions were not uniform in size and/or time. The regular motor pattern then commenced and appeared to have, at least in part, recovered from the initial w- CTX insult. Incubation of the ileum with both w—CTX and NiCI2 completely inhibited facilitation of the distension-induced motor reflex (Figure. 6.11C1, C2). 191 DISCUSSION The mammalian small intestine contains approximately 100 million neurons, about as many neurons as that found in the spinal cord (Gershon, 1997). The functions of these neurons are controlled primarily by the ENS. The ENS is therefore a critical regulator of GI function by coordinating motility (peristalsis), absorption and secretory responses, thus indicating that altered enteric neuronal signaling can therefore cause GI motility disorders. In this study I therefore sought to understand the physiological role of the voltage-gated R- type calcium channels in regulating peristalsis. The major findings of this study are that the R-type calcium channels are coupled to NO release, and that this channel, together with the N-type calcium channel are the major contributors to fluid distention-induced peristalsis in vitro. Specifically, the evoked peristalsis reflex is impaired by the association of the R-channel and NO, and by the R- and N-type calcium channel subtypes. These findings have important implications for furthering our understanding of peristalsis as they provide a mechanism by which ion-channels can be targeted to control GI motility. Voltage-gated calcium channels mediate neurotransmitter release and are therefore important regulators of synaptic transmission. I first tested for a role of calcium channels in peristalsis in the guinea pig ileum in the presence of CdCI2. CdCI2 inhibited the peristaltic reflex by blocking calcium channels at all synapses, consequently increasing the intraluminal threshold pressure and the amplitude of the peristaltic wave. Peristaltic motility however, was not abolished, but still present as small amplitude, shallow, and regular oscillatory contractions. These 192 oscillations could be due to the intrinsic myogenic mechanisms of GI smooth muscle cells via adjacent interstitial cells of Cajal (ICC) (Huizinga et al., 2006; Sanders, 2008). ICCs generate the slow wave that passively transmits to the smooth muscle cell that ultimately, in conjunction with neural input, determine the frequency and strength of contractions (Farrugia, 2008). I therefore concluded that calcium channels are necessary for the release of neurotransmitters from enteric motor axons to generate peristalsis. I observed identical results when 'l'l'X was applied to the ileal segment; the reflex was inhibited, confirming neural mediation. Inhibition of the peristaltic reflex by NiCI2 and NLA. R-type calcium channels represent a viable control point for neurotransmission if the ENS (Bian et al., 2004; Bian & Galligan, 2007), and a1E subunits are localized in cell bodies of IPANS and in nerve endings of excitatory (SP-containing) and inhibitory (NOS and VIP-containing) motorneurons (Naidoo et al., 2006). These channels may therefore mediate synaptic excitation in the ascending or descending reflex pathway. In my first set of experiments, I therefore investigated whether R-type calcium channels contribute to the peristaltic reflex. I used NLA to further investigate a selective coupling of the R- channel to NO inhibitory neurotransmission. NiCI2 inhibited, but did not block peristalsis. Subsequent application of NLA with NiCI2 did not produce any additional changes in peristalsis suggesting that NLA and NiC|2 act on the same mechanism. Interestingly, when NLA was applied first the peak amplitude, 193 l'll‘fl A: I. 1- -— - - frequency, AUC and volume expelled all increased, particularly at higher threshold pressures. Subsequent application of NiCI2 with NLA inhibited peristalsis. Using the peristaltic frequency as a point of reference, this then suggests at least two possible scenarios: 1) When NLA was applied first, the frequency of peristaltic contractions significantly increased. NLA inhibits NOS to increase the propagating peristaltic contraction frequency (Ciccocioppo et al., 1994). However, when NiCI2 was subsequently applied in the presence of NLA, the contraction frequency decreased. Since NiCI2 (50 (M) inhibits calcium entry though the R-type channel, if the R-channel and NOS are selectively coupled, then calcium would not be able to activate NOS. However, it also appeared that only when the R-channel was blocked, that the ability of NLA to inhibit NOS was impaired because the resultant frequency declined. The source of this decay points to a physical interaction between the R-channel and NOS, and suggests that when the R-type channel is blocked, NOS cannot be activated and therefore NOS cannot be blocked by NLA. The fact that many peristaltic waves are still propagating suggests that NiCI2 may be inducing a “braking” system on NLA-induced NOS inhibition that would othenrvise produce a hyperexcitable ileum. This effect may require synaptic release of multiple inhibitory enteric neurotransmitters from motor axons on to the smooth muscle. Recent evidence indicates that enteric motor nerve terminals in the mouse may contain catalytically active. forms of NOS (nNIOSa) that are anchored via PDZ binding domains to the plasma membrane (Stricker et al., 1997; Rao et al., 2008). NOS may therefore be docked at the 194 nerve terminal, close to R-type calcium channel release sites. My data suggest however that NO is probably not the transmitter being released because a physical association of the R-type channel to NOS requires that calcium enters through the R-type channel to facilitate NO release. This implies that the lower peristaltic frequency may be due to mediation by another inhibitory neurotransmitter via another calcium channel subtype. After NiC|2 and NLA treatment, apamin completely inhibited the peristaltic reflex. A likely candidate for this neurotransmitter is therefore ATP/BNAD as a mediator of apamin-sensitive inhibitory neurotransmission. ATP and NO are released from the same nerve terminals (Costa et al., 1986b; Crist et al., 1992) and constitute the major contributors to inhibitory neurotransmission in the ENS (Burnstock, 2008). Moreover, NO and ATP are reported to exert independent effects on GI motility in the rat ileum and porcine esophagus (Benko et al., 2006; Farre et al., 2006). When NLA and apamin were co-applied to the rabbit distal colon (Ciccocioppo et al., 1994) and guinea pig small intestine (Waterman & Costa, 1994; Holzer et al., 1997), the gut became hyperexcitable but with non-propulsive segmental motor patterns and very little or no emptying. However, in my study, I had added apamin after NiCI2 + NLNNLA + NiCI2 and in both cases, apamin did not affect the peristaltic reflex. The resulting motor pattern was similar to the effect by TTX on the peristaltic reflex. An important test would have been to add apamin first before either NiCI2 or NLA. The emptying phase of peristalsis involves the activation of excitatory and inhibitory motor neurons (Waterman & Costa, 1994). In order for efficient propagative peristalsis to occur, there must therefore be 195 functional synchrony between activation of these motorneuronal inputs to the smooth muscle. 2) To confirm this coupling, I did the reverse experiment. When NiCI2 was added first, the frequency of contractions was reduced, suggesting the possible release of NO. Interestingly, when NiC|2 with NLA were co-applied, the frequency remained the same. In the presence of NiCI2, NLA was unable to inhibit NOS and could not produce an increase in the contraction frequency. This verified my conclusion from the previous experiment that entry via the R-type calcium channel may be coupled to the activation of NOS. Inhibiting the R-type channel possibly prevents inhibition of NOS. If the R-channel is inhibited, and selectively coupled to NOS, then the low peristaltic frequency observed is probably mediated by another inhibitory neurotransmitter. I observed that apamin again completely blocked peristalsis. Blocking the R-channel may alternatively also prevent the release of excitatory neurotransmitters. My data however argue that ATP release from inhibitory motor axons may be responsible for the reduced peristaltic frequency rather than low excitatory neurotransmission. Norepinephrine, released from sympathetic nerve endings, act at az-adrenoreceptors to inhibit intestinal motility (Surprenant & North, 1988; Stebbing et al., 2001). A selective coupling between the R-type calcium channel and az-adrenoreceptors on myenteric neuronal cell bodies and nerve terminals has recently been proposed (Bian & Galligan, 2007). az-adrenoreceptors have been detected on cell bodies of myenteric neurons but the exact phenotype of myenteric neuron expressing these receptors is not yet known (Bian & Galligan, 196 2007). The R-type channel may be present on the cell bodies of lPANs where it modulates the slow excitatory postsynaptic potential and somal action potential. That study showed that agonist-activated az-adrenoreceptors inhibit calcium entry through R-type channels, because in the presence of NiCI2, inhibitory control by the az-adrenoreceptor over the R-type channel was removed. In my study, when NLA was applied first, it is possible that the R-type channel was unable to prevent inhibition of NOS during the peristaltic reflex due to the release of norepinephrine from sympathetic nerve endings. Norepinephrine, acting at pre-junctional az-adrenoreceptors on myenteric motor axons, would then inhibit calcium influx by decreasing the probability of opening of the R-type channel, or activating presynaptic K” channels (Hein, 2006). This principle would also hold when NiC|2 was applied first. NiCI2-block of the R-type channel presumably removes inhibition by the az-adrenoreceptor to cause either a conformational change in the calcium channel protein, or a change in the R-channel-NOS signaling cascade that renders NOS inaccessible to inhibition by NLA. Inhibition of the peristaltic reflex by NiC|2 and w-CTX In the present study, I demonstrate that the peristaltic reflex in the guinea pig ileum is mostly mediated by R- and N-type calcium channels. This conclusion is based on the observation that co-application of MCI; and the N-type calcium channel antagonist w-CTX completed inhibited the peristaltic reflex. My concentration-response curve for NiCI2 and w-CTX shows that the additive inhibitory effect of these two calcium channel antagonists is approximately 60 %. 197 When NiCI2, was added first, the number of propagating peristaltic waves, peak peristaltic amplitude, area under the curve, and volume of fluid expelled all significantly decreased compared to control responses. These parameters appeared to be particularly affected at higher threshold pressures. The reflex was impaired but was not blocked. Application of w—CTX then completely inhibited the peristaltic reflex. When w—CTX was applied first, non-propulsive segmentation patterns were observed, and propagating contractions only began after at least 25 5. By itself, w-CTX did not change the threshold pressure for initiation of peristalsis. This suggests that N-type calcium channels mediate neurotransmission between interneurons and motorneurons, and between motorneurons and the smooth muscle. However, in the presence of w-CTX and NiC|2, the reflex was abolished, signifying inhibition of neurotransmission at all synapses. I still observed intrinsic myogenic activity, indicating that peristalsis still exists, albeit with a low level of neural excitation. To date, only two studies have reported on the functional significance of R-type calcium channels in the GI tract (Bian et al., 2004; Bian & Galligan, 2007). Most studies on R-type channel function have been done in the CNS and the R-channel has been localized to nerve endings at many excitatory synapses (Wu et al., 1998; Tottene et al., 2000; Gasparini et al., 2001; Dietrich et al., 2003). In wild-type mice, PIQ-type calcium channels are the major calcium channel subtype found on nerve terminals in the calyx of Held (Wu et al., 1999), cerebellum (Matsushita et al., 2002) and hippocampus (Qian & Noebels, 2000). Furthermore, PIQ-type channels primarily mediate acetylcholine release at mammalian endplates (Urbano et al., 2003; 198 Pagani et al., 2004). Urbano et al. (2003), using PIQ-type calcium channel knockout mice showed that: 1) R- and N-type calcium channels are upregulated, and 2) R-type channels were less numerous than N-type channels, but are located closer to presynaptic calcium sensors at the active zone. In the ENS, PIQ-type calcium channels display weak immunoreactivity weak in the soma and nerve terminals of myenteric'neurons, whereas N-type calcium channel-ir was diffuse and intense (Kirchgessner & Liu, 1999). Further, PIQ-type calcium channels contribute only 20 % to the total calcium channel current (Bian et al., 2004). In the present study, my data suggest that the major contributors to the peristaltic reflex are the N-and R-type calcium channels. Previous immunohistochemical and electrophysiological data indicate that a1E co- localizes with NOS, VIP and SP in enteric motor neurons, and that R-type channels are present in a specific population of purely-cholinergic calretinin- containing ascending interneurons (Naidoo and Galligan, unpublished data). The present data indicate that NiCI2, in the presence of fluid-induced radial distention, clearly induces a shift in motor activity of the ileum, from neurally-mediated fully propagating reflex patterns to segmental contractions associated with non- propulsive churning movements of the fluid content. | speculate that this shift may be due to a block of neurotransmission at specific synapses along the peristaltic reflex pathway. At these synapses, the concentration of R-type channels may be high. Ascending interneurons constitute 5 % of the total number of myenteric neurons in the ileum (Furness, 2006a) and 25 % of these are fully cholinergic (Galligan et al., 2000). At purely-cholinergic ascending excitatory 199 synapses, R-type channels would mediate all cholinergic neurotransmission from those calretinin-containing interneurons. In fact, Gasparini et al. (2001) showed that 50 (M NiC|2 reduced fast glutamatergic transmission in the CNS by at least 51-59 %. Because R-channels are also present on SP- and NOSNIP-containing enteric motorneurons, they may control excitatory and inhibitory neurotransmission to the muscle. By inhibiting calcium-entry with NiCI2 at all these synapses, the reflex circuitry is dysregulated. R-type channels make an important contribution to the peristaltic reflex, but in its absence, this reflex can still continue, as I have observed under conditions of fluid-evoked distention. Inhibiting its function, however, disrupts motility. The data suggest that at least two calcium channels are necessary for the peristaltic reflex. Perhaps for the peristaltic reflex to occur, two calcium channels, from a combination of N-type with the R-type calcium channel are necessary for this facilitation. This may be part of a fail-safe mechanism that has built into the reflex circuitry to ensure the efficacy of the propagating peristaltic motor pattern in the guinea pig. SUMMARY AND CONCLUSIONS Inhibition of R-type calcium channels in an in vitro model of fluid-induced peristalsis suppresses but does not completely inhibit the peristaltic reflex. The R-type channel is coupled to NOS and when both are inhibited, the peristaltic reflex is abolished. The R-type and N-type calcium channels function in tandem to facilitate the peristaltic reflex. Further understanding of the interaction between the various voltage-gated channel subtypes holds promise for improving our 200 understanding of GI motility, and the development of therapeutic strategies against motility diseases. 201 Figure 6.1. Peristalsis video imaging. A. A single frame from a peristalsis movie recording of the ileum imported into Image J. B. All movie frames were then converted into binary images. Image thresholding and edge detection delineated the ileum periphery. By selecting regions of the ileum and using intestinal motility software, I was able to detect changes in gut diameter. These changes were pooled and built into 3- dimensional spatio-temporal maps with an Image J-plugin linked to the graph-plotting program Gnuplot. 202 Figure 6.2. EP5o pressure-response curve and representative traces. (A) Pressure-response curve in drug-free Krebs’ solution for calculating the magnitude of the area under the curve (AUC) against successive increases in intraluminal pressure. The half maximal effective pressure (EP50) was 1.9 mmHg. Data points are mean :I: S.E.M. n represents the number of animals. Curves are non-linear fits of the Hill equation. (B) The reservoir was raised from a baseline pressure where no peristalsis was observed, to a threshold pressure (T, arrow) that produced a peristaltic reflex. (C) The area under the curve (AUC) was calculated for every pressure interval from T to T+5. These intervals were in 0.5 mmHg increments after the T value was known. Note that only the active area of the AUC was calculated. The arrow indicates the desired pressure level, and the dotted red line (at the arrow point) is the baseline for calculation of the AUC. The amplitude and frequency of peristaltic waves, in addition to the volume expelled, were also calculated. 203 A 125- Control (n = 8) t 3 ‘1 i 5 i 45‘ Intraluninal pressure (““01 3 Control Pressure Trace (Threshold, T) Baseline pressure C Control Pressure trace: T+5 I Frequency % T+5 Amplitude l 1' .......AU. # 1 L Baseline 10 5 re ................................................................................................................ 204 A 1000001 Control (n = 5) 95000( 90000J 850004 800001 AUC (mmHg.ms) 75000 0 10 20 30 40 50 60 Minutes after equilbration 0'19' Control (n = 5) 0.18“ 0.17- M 0.16‘ o. 1 5 T U I V I I 0 10 20 30 40 50 60 Minutes after equilbration Peristaltic frequency (waves! 8) Figure. 6.3. Time controls showing stable peristaltic motor patterns after measuring (A) the AUC and (B) the peak peristaltic frequency in 5 min resting intervals for a period of 50 min after tissue equilibration. Peristalsis was evoked and recorded for 60 s. The ileal segment showed no evidence of fatigue during the 50 min experimental period. 205 1501 I OJ‘CTX (n = 4) - NiC|2 (n = 8) CI CdCIZ (n = 5) d; 100‘ ,\ U D < 50‘ G T I fl I -10 -8 -6 -4 -2 [Calcium antagonist] (M) Figure. 6.4. Effect of w-conotoxin (cu-CTX), NiC|2 and CdClz on the AUC expressed as percent change. w-CTX (1-100 nM) and NiC|2 (1-50 pM) decreased the AUC by 39% and 20% respectively. CdCI2 (1-100 uM) caused a concentration-dependent inhibition of the AUC, and blocked the AUC completely at 100 pM. Data are mean :I: S.E.M. n indicates the number of animals, 206 Figure. 6.5. Effect of NiCI2 and NLA on (A) frequency, (B) AUC, and (C) fluid volume expelled. These four parameters all decreased in response to NiCI2-treatment alone. Subsequent application of NLA did not produce a shift in any of the parameters. This was particularly evident at higher intraluminal pressures: frequency and AUC 1.9 — 2.9 mmHg; volume expelled 1.4 — 2.9 mmHg, and mean diameter 1.9 - 2.9 mmHg, compared to control (two-way ANOVA; P < 0.05). Data are mean :I: S.E.M. *P < 0.05; **P < 0.01; *** P < 0.001. 207 Peristaltic frequency (waves! 3) 40°00 - Control AUC AUC (mmHg.ms) .§§ Volume expelled (mi) 30000 - NiCI2 + NLA 5 0.25 - Control Frequency 0.4 0.9 1.4 1 .9 2.4 2.9 Intraluminal Pressure (mmHg) - NICI2 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) Volume expelled (ml) (per 60 seconds) - NiC|2+ NLA n=5 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) 208 Figure 6.6. Effect of NLA and NiCI2 on (A) frequency, (B) AUC and (C) fluid volume expelled (refer to Figure. 6.7). Addition of NLA first, increased all the above parameters of peristalsis. Subsequent application of NiCI2 reduced these parameters. A. The peak NLA contraction was significantly different from NLA + NiCI2 but not from control. B. The frequency of peristalsis after NLA treatment was higher compared to NLA + NiC|2 treatment, particularly at intraluminal pressures 1.4 — 2.9 mmHg. NLA, and NLA + NiC|2, were significantly different from control at pressures 1.9 - 2.9 mmHg. C. The AUC after NiCI2 and NLA treatment was greater than either control or NiC|2 alone, at 1.9 — 2.9 mmHg. Data are mean :I: S.E.M. 209 AUC (mmHg.ms) i? To a: i E. > o r: o = a' o It Io E 8 .2 h a: 0. 50000 - Control 0.10 0.05 0.00 Frequency '2' 0.20 - Control “M ** I NLA H - NLA+ NiC|2 0.15 ":4 0.4 0.9 1.4 1 .9 2.4 2.9 lntraluminal Pressure (mmHg) AUC Volume expelled (ml) 0.4 0.9 1 .4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) Volume expelled (ml) (per 60 seconds) 8 - Control ** I NLA m If" 6 I NLA+NICI2 n=4 4 2 0 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) 210 Figure 6.7. The peristaltic motor pattern changes with NiCI2, and NiC|2 + w-CTX treatment (refer to Figure. 6.9). A. The frequency after NiC|2 application was reduced (1.4 - 2.9 mmHg vs. control). Addition of w-CTX completely inhibited the reflex wave frequency (0.9 - 2.9 mmHg vs. control; (1.9 - 2.9 mmHg vs. control). B. The AUC was inhibited after NiCI2 treatment (1.9 -- 2.9 mmHg vs. control). Subsequent application of w-CTX then completely blocked the AUC (0.9 — 2.9 mmHg vs. control; 2.4 - 2.9 mmHg vs. NiCI2). C. Volume of fluid expelled decreased significantly compared to control at intraluminal pressures 2.4 and 2.9 mmHg. Following NiC|2 application, the volume ejected was significantly reduced (0.9 - 2.9 mmHg vs. control; 2.9 mmHg vs. NiCI2). D. Mean intestinal diameter and. Under control conditions, the ileum demonstrated a small mean diameter. After w-CTX treatment, the diameter increased. Following w-CTX and NiC|2 treatment, the diameter significantly increased (1.9 - 2.9 mmHg vs. control; 1.9 -— 2.9 mmHg vs. w-CTX). Data are mean :I: S.E.M. 211 Peristalfic frequency (waves! s) AUC (mmHg.ms) Frequency 0-20‘ - Control - N'CIZ (“5‘ .- NICI2+ CTX . * A n - 5 H a i it 0.10' * 0.05- ** *** *** *** I ** 0.00d 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) AUC 500001 I Control I NiC|2 .I W I NiCI2 + w-crx sooood n = 4 . ** * H :7 200W * rooom * ** *** *** m 0.1 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) Volume expelled (ml) (per 60 seconds) a 12' - Control 5 10- I NiCI2 3 8. I NiCI2 + m-crx ** 3 n = 4 ft 64 o .l 0 4 S l 3 2 > 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) 212 Figure. 6.8. Effect of w-CTX alone, and w-CTX with NiC|2, on peristalsis parameters. A. After w-CTX treatment, the frequency significantly decreased compared to control (0.9 - 2.9 mmHg). Addition of NiC|2 then inhibited the frequency (0.4 - 2.9 mmHg vs. control; 1.4 - 2.9 mmHg vs. w-CTX + NiC|2). w- CTX together with NiC|2 blocked propagating reflex contractions, but not small oscillating contractions B. The AUC decreased significantly after w-CTX treatment compared to control at intraluminal pressures 1.4 and 2.9 mmHg. The AUC was then significantly reduced after application of NiC|2 (0.9 - 2.9 mmHg vs. control; 1.9 - 2.9 mmHg vs. w-CTX). C. Volume of fluid expelled declined significantly in the presence of w-CTX (1.9 - 2.9 mmHg vs. control) and after w- CTX + NiC|2 treatment (1.9 - 2.9 mmHg vs. w-CTX). D. Under control conditions, the ileum demonstrated a small mean diameter. After w-CTX treatment, the diameter increased. Following w-CTX and NiCI2 treatment, the diameter significantly increased (1.9 - 2.9 mmHg vs. control; 1.9 - 2.9 mmHg vs. w-CTX). Data are mean :I: S.E.M for n tissues. 213 AUC (mmHg.ms) Peristaltic frequency (waved s) Frequency I Control CI w-crx I (o-CTX + NiCI2 n = 4 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) AUC I Control 80°00] :3 max I -CTX + NiCI2 m *** soooot 0’ ,2, H "' n = 4 it 40000“ ** 20000‘ W ** *** | . 0‘ ' 0.4 0 9 1 4 1 9 2.4 2.9 lntraluminal Pressure (mmHg) Volume expelled (ml) (per 60 seconds) i 12' I Control " 10' C] (D'CTX 3 8) I (o-CTX + NiCI2 *** 3 .4 o E 4‘ 2 2' o > 0- 0.4 0.9 1.4 1.9 2.4 2.9 lntraluminal Pressure (mmHg) 214 CHAPTER 7 GENERAL DISCUSSION AND CONCLUSIONS 215 Gastrointestinal motility disorders pose a severe reduction on the quality of life for affected individuals. Despite a plethora of studies aimed at understanding the pathophysiology of GI disorders, there are still no cures. The reason probably has to do not only with the complexities of the disorders themselves, but also with the overall impact on the ENS. For example, in patients with IBS and Crohn’s disease, the ENS is a specific target of inflammatory cells including lymphocytes and mast cells that alter the enteric neural circuitry. After tissue inflammation, signals generated by inflammatory cells induce changes in extrinsic sensory nerve endings, lPANs, interneurons, and motorneuronal signaling to the smooth muscle (Kirkup et al., 2001). The effect of impaired or altered synaptic communication means that the ENS has to attempt to overcome inflammatory insults by deconvoluting new sensory signals and combining them within the enteric circuitry to produce an adaptive response. However, in IBD and Crohn’s disease patients, drastic therapeutic strategies such as corticosteroid administration, or as a last resort, surgical intervention, are used to help patients with moderate to severe forms of those diseases. A reason for the lack of cures for GI disorders also probably has to do with the high sensitivity and semi- autonomous nature of the intrinsic reflex circuitry of the ENS. By high sensitivity I mean, for example, that even a minor alteration in the signaling pathway of the ascending excitatory pathway may have overall consequences on muscle contraction. Similarly, dysregulation of the descending inhibitory pathway would produce an altered result on muscle relaxation. IPANs are connected in feed- forward circuits, are the first neurons activated by distention or mucosal 216 stimulation, and act to gate excitation in the myenteric and submucosal circuitry. The integrity of the nerve endings of lPANs is therefore important for continual gut function. Furthermore, IPANs, interneurons and motorneurons are connected at multiple points along the GI tract, implying that a response to a given stimulus can occur at any given point in the gut wall. The semi-autonomous nature of the GI tract is advantageous and pragmatic in terms of having its own “little brain”, in that transmission over long distances (from the CNS to the ENS) and the packaging of 100 million enteric neurons into the CNS has been avoided through evolution. However, this may also be to the detriment of the ENS. That is, because of the paucity of connections between the CNS and the ENS, the ENS is only able to respond to dysregulatory mechanisms mostly by itself. Therefore, when the GI reflex circuitry is affected by an inflammatory “soup”, the outcome is severe and can be debilitating. Psycho-social factors are also believed to be involved in GI motility disorders, but conclusive evidence has yet to be presented. The search for therapeutic strategies to combat motility disorders, although challenging, is ongoing, and the studies presented in my dissertation contribute to that search. Because of the immense complexity that GI disorders present, I believe that a good strategy to target ENS dysfunction is to study the mechanisms of the ENS reflex circuitry. After all, that is the region where many GI disorders and diseases produce their overall effects. Calcium channels are the critical conduits of synaptic transmission. They transduce electrical signals into chemical ones, and therefore regulate communication between excitable cells. They are also involved in contraction of 217 smooth and skeletal muscle, secretion, and gene expression. Calcium channels have also been established as key components in the ENS. Seminal experiments by Bian and Galligan (2004) demonstrated that at least 50% of the calcium current in myenteric neurons maintained in primary culture is R-type calcium channel-mediated. In light of that evidence, my dissertation addressed the functional importance of the R-type channel in the ENS. The guinea pig has been an invaluable and major contributor in furthering our understanding of ENS function. Other species, however, such as rats and mice have also been used to study the ENS. Much of our present understanding of the ENS is based on structural, morphological, neurochemical and physiological studies done in guinea pigs. The ileum is also the most studied and best characterized GI region. Study of local ENS control can also best be studied in this region because ileal myenteric neurons are sparsely innervated by vagal (parasympathetic) preganglionic cholinergic neurons and sympathetic postganglionic noradrenergic neurons located in celiac ganglia (Goyal & Hirano, 1996). CNS contribution to motility control in the ileum is therefore limited, implying that motility is predominantly locally regulated. The guinea pig was therefore my model system of choice for these studies. A major finding in my study was that a1E-ir is not present in the submucosal plexus, and therefore does not contribute to absorptive or secretory functions in the guinea pig GI tract. However, extensive a1E-ir in varicose nerve fibers is present in the myenteric plexus in all regions of the GI tract. The 218 predominance of R-type calcium currents in myenteric neurons suggests an important role for this class of calcium channels in enteric neurotransmission and motility. Interestingly, there was no expression of a1E within cell bodies in the guinea pig small intestine. Cell bodies of myenteric neurons in the stomach and colon, however, do express R-type channel immunoreactivity, with the colon demonstrating a greater expression of a1E-ir. Overall, a1E staining in the colon was greater compared to the stomach and small intestine. The reasons for the intense a1E labeling in the colon LMMP could be related to the function of the colon. The colon absorbs water and electrolytes from the intestinal content and converts it into fecal matter. Furthermore, migrating myoelectric complex (MMC) patterns are absent in the colon, but slow waves are present. The MMC is an electromechanical pattern that occurs in the fasting state. It is responsible for the propulsion of residual undigested food and sloughed enterocytes from the stomach through the small intestine. It therefore serves a “housekeeping” function. The musculature in the proximal colon exhibits two basic motor patterns: segmental (haustral) contractions, and neurogenically—mediated mass peristalsis, while the distal colon only displays the former (Furness, 2006a). The function of the R-type channel, present on cell bodies and on nerve terminals, could therefore be to augment peristalsis in the colon. This could occur by mediating calcium entry during somatodendritic action potentials, or neurotransmitter release. I did not find a1E-calbindin-ir colocalization in the colon LMMP. However, I did observe strong co-localization of a1E and SP, a1E and calretinin (nerve fibers) and a1E with NOS and VIP. This suggested that a1E 219 was present in: 1) excitatory and inhibitory circular and longitudinal muscle motorneurons, 2) ascending interneurons, and, 3) lPANs. The prevailing dogma is that calbindin is the best known marker for myenteric lPANs. However, according to Brookes and Costa (2006), at least two types of myenteric lPANs exist: 1) an IPAN majority that respond to chemical stimulation of the mucosa; and 2) an IPAN minority that are sensitive to tension along the gut wall (Kunze et al., 1998; Kunze et al., 2000). In fact, a third subset of IPANs may also exist, those that are sensitive to both tension and chemical stimuli. Those data suggest that calbindin may not necessarily fit the gold-standard for confirming IPANs, and my data support this theory. SP is a marker for both IPANs and excitatory longitudinal muscle neurons. SP may therefore be present in IPAN populations that are calbindin-negative. The SP antibody I had used never labeled cell bodies, while calbindin never labeled nerve terminals. However, because SP and a1E co-localize in the myenteric LMMP, the R-type channel may therefore be present in nerve terminals of colonic IPANs that are calbindin-negative. They may also be present on nerve endings of excitatory motorneurons. There are two ways to answer these possibilities. One is to record action potentials and sEPSPs from AH neurons in LMMP preparations in the colon, in the presence of NiCI2 (50 nM) or SNX-482 (0.1 nM). The second method would be to perform neuron cell culture to determine whether a1E is expressed in colonic myenteric neurons. R-type channel staining in the small intestine deviated from the stomach and the colon. The abundance of axonal varicosities in the intestinal 220 whole mount LMMP suggests that the function of the R-type channel may be more towards that of synaptic transmission. Intracellular recordings of action potentials from single AH showed that NiCI2 did not reduce the amplitude of the AH action potential, but did inhibit the half-width by almost 50 %. Addition of w-CTX then completely blocked the calcium shoulder. The late AHP was also significantly reduced in the presence of NiCI2. Rugiero and others (2002) showed a major contribution of N-type channels to the action potential, and also implicated the R-type channel as a contributor to the total inward calcium current. In my study, I have shown the N- and R-type calcium channels contribute equally to the somatodendritic action potential in AH neurons. NiCI2 had no effect of on action potentials in S-type neurons. The effect of NiC|2 on fEPSPs was interesting. I initially observed that some fEPSPs were significantly inhibited by NiC|2, whereas others were completely unaffected. There is a single group of ascending interneurons but 3 groups of descending interneurons in the guinea pig ileum (Brookes, 2001). The ascending interneurons comprise just 5 % of the total number of myenteric neurons in the myenteric plexus and are immunoreactive for ChAT and calretinin. Those neurons use ACh as their neurotransmitter. It should be noted that some excitatory longitudinal muscle motorneurons (25 %) are also calretinin-ir. In the guinea pig LMMP, the axons of excitatory longitudinal muscle motorneurons are confined to the tertiary plexus (Brookes & Costa, 2006), are short, and project in this plexus close to their cell bodies (Lomax et al., 1999; F urness, 2000; Brookes & Costa, 2006). In the LMMP preparation, there is therefore a greater chance of 221 recording from interneurons (ascending or descending) rather than longitudinal (excitatory or inhibitory) muscle motorneurons as l was stimulating interganglionic fiber tracts. Furthermore, motorneurons do not synapse with other enteric neurons (LePard & Galligan, 1999). Together, descending interneurons account for 11% of the total number of myenteric neurons in the guinea pig LMMP and use either ACh and ATP, or ACh and 5-HT for neurotransmission (LePard et al., 1997; LePard & Galligan, 1999; Galligan, 2002b). A third group of descending interneurons is associated with the MMC and has Dogiel type III (filamentous) morphology. The majority of the interneurons (67%) belongs to the mixed cholinergic/purinergic group, while 25% are completely purinergic (Galligan, 2002b). Components of the fEPSP can therefore be isolated using mecamylamine (nAChR antagonist), PPADS (P2X receptor antagonist) or ondansetron (5-HT3-receptor antagonist). My data shows that NiCI2 and SNX—482 selectively act at the nerve endings of purely cholinergic nerve fibers to inhibit R- type calcium channels as the fEPSP amplitude was reduced. Neurons that use both ACh and ATP for fast transmission were unaffected by NiCI2 and SNX-482 treatment. This result means that R-type calcium channels are localized to calretinin interneurons, and this is supported by my findings that a1E and calretinin co-Iocalize in some varicosities in the LMMP. The paucity of co- localization between a1E and calretinin supports this conclusion as the R-type channel is present in just 5 "/0 of purely cholinergic ascending interneurons. In both subtypes (cholinergic/purinergic), w-CTX largely inhibited the fEPSP following R-channel blockade. Therefore, mainly R- and N-type channels 222 contribute to fast synaptic in purely cholinergic S-type interneurons, whereas only N-type calcium mostly mediate fast transmission neurons at mixed cholinergic/purinergic nerve endings. The ideal way to determine whether NiCI2 is acting pre-or postsynaptically to inhibit calcium entry through R-type calcium channels would be to “puff” onto impaled neurons an agonist of the mediator of the sEPSP or fEPSP in the presence of NiC|2. Previous data (DeVries M, unpublished) showed that NiCI2 acts presynapticallly to inhibit synaptic trasmission. My present data further indicate that CdCI2 and NiC|2 do act presynaptically. This conclusion is based on: 1) CdCI2 did not affect the neuronal membrane potential; 2) NiCI2 did not affect the sEPSP; 3) NiC|2 acts at nerve terminals in fully cholinergic S-type ascending interneurons; 4) NiC|2 does not affect fEPSPs in descending interneurons. Taken together, these data rule out a general non-specific effect on the postsynaptic cell. Perhaps the reason why the R-type channel is present on nerve endings of purely cholinergic ascending interneurons is based on redundancy. That is, the ENS has included a mechanism using two calcium channels that act as fail-safes in instances of, for example, a toxin—attack that acts to block chemical neurotransmission. Having two calcium channels at a single synapse would therefore help the ENS circumvent noxious insults. Because the ENS has a single population of S-type interneurons, the R-type channel may consequently be very important in this regard. ACh is an important enteric neurotransmitter, and the importance of purely cholinergic interneurons may be to spread excitation to other cell types, 223 including excitatory motorneurons, to allow the coordination of motor responses over long distances in the GI tract. R-type calcium channels also mediate inhibitory neurotransmission to the muscle layers, suggesting a functional coupling between the R-type calcium channel and NOS synthesis and NO release (Figure. 7.1). The R-channel does not appear to be linked to ATP release from nerve terminals. It is possible that the R-type channel, localized to myenteric nerve endings is coupled to NO synthesis and release, physically or via an intracellular signaling mechanism, to cause relaxation of GI smooth muscle through an ATP (and possibly B-NAD)- insensitive mechanism. This mechanism would be as follows: intracellular nNOSa would be anchored at the nerve terminal membrane, and activated by the influx of calcium. NO would be synthesized by intracellular NOS, and NO would then passively diffuse to affect the postsynaptic cell. These data therefore have important implications for the spatial organization of calcium-channel- mediated exocytosis. The effect of NiC|2 on R-type calcium channel function in the whole ileum is important for the understanding of possible therapeutic strategies against GI motility disorders, particularly in light of ion channels as potential drug targets (Lipecka et al., 2002; Cuppoletti et al., 2004). For example, the peptide Ziconotide is a synthetic analogue of w-CTX used to treat chronic pain (Miljanich, 2004). Similarly, the spider toxin peptide inhibitor guangxitoxin-1 targets the delayed rectifying potassium channel Kv2.1 and enhances glucose-dependent 224 insulin secretion. This has important implications for the treatment of type II diabetes (Herrington et al., 2006; Kaczorowski et al., 2008). However, the challenges in translating ion-channel drugs from bench to bedside is difficult due to, amongst others, species differences, and just how efficacious such drugs would be in humans (Kaczorowski et al., 2008). NiC|2 by itself impairs the peristalsis reflex. Peristalsis can still occur, but mixing and churning movements of the intestine were observed 30 seconds after fluid-induced peristalsis. Where exactly is NiC|2 acting? NiC|2: 1) reduces the calcium shoulder and the late AHP amplitude of the AH action potential; 2) acts at nerve terminals of purely cholinergic interneurons; 3) reduces inhibitory neurotransmission by coupling to NO to the muscle layers; 4) may also act at other unknown sites. By disrupting synaptic communication in a number of pathways, such as the ascending excitatory pathway, and at inhibitory neuromuscular junctions, the function of the reflex circuitry is compromised. N- and PIQ-type calcium channels appear to be localized near R-type channels, such that calcium channel co-operativity between those major calcium channel subtypes may be important for facilitation of enteric reflexes (Pietrobon, 2002, 2005). My data indicate the presence of R-type calcium channels within specific populations of myenteric neurons. R-type channels may therefore have an important function as a relay point for information signaling within those neurons. lmpairing its function alters the peristaltic reflex. A schematic of a potential coupling between voltage-gated R-type calcium channels (VGCC) and NOS on myenteric inhibitory motor nerve 225 terminals is shown in Figure. 7.2. In Figure. 7.2A, under drug-free conditions, fluid-distention induces a propagative peristaltic reflex. At an inhibitory neuroeffector synapse, NOS is anchored within the active zone to the plasma membrane. This synapse is also under inhibitory control by sympathetic nerves that release norepinephrine (NE). NE, acting at presynaptic GilGo-coupled a2- adrenoreceptors (azAR) will inhibit release of neurotransmitters from the myenteric inhibitory motor nerve fiber. If the azAR and R-type channel are coupled perhaps by diffusible second messengers, then the probability of the R- channel being open will be reduced, and calcium influx will be blocked. If the R- channel and NOS are coupled, then NO cannot be released if the R-channel is under the control of the azAR. If the R-channel and NOS are not coupled, then NO can be released via calcium entry through another calcium channel, possibly the N-type calcium channel. NO can then diffuse across the neuroeffector junction into the smooth muscle cell, bind to guanylate cyclase, to cause muscle relaxation. When NLA is added first (Figure. 7.28), and distention evoked, the frequency of contractions increases. NLA inhibits NOS leading to increased smooth muscle contraction. NE will again act at the azAR and because it is coupled to the R-channel, calcium entry through this channel will be then blocked by azAR activation. If the R-channel and NOS are coupled, then in the presence of NLA alone, NOS cannot be under the control of the R-channel, because the channel is still inhibited by the azAR. NLA is now free to exert its inhibitory actions on NOS, leading to increased contractions. 226 When NiC|2 is added in the presence of NLA (Figure. 7.2C), and peristalsis induced, the frequency of contractions decreased. NE is still released from the sympathetic nerve terminal and activates the azAR. The azAR, however, cannot inhibit the R-type channel because NiCI2 (50 pM) removes this inhibition (Bian & Galligan, 2007). Calcium entry therefore cannot occur. Because the contraction frequency is reduced (by NLA and NiCI2), this suggests that NLA cannot be inhibiting NOS. Speculatively, this may be because of an intracellular coupling pathway or a physical interaction between the R-type channel and NOS. Reduced calcium entry via the R-type channel would lead to less calmodulin binding, resulting in reduced NOS activation. NO therefore cannot be synthesized via the R-channel at this synapse. The inhibitory response must be due to release of another neurotransmitter via another calcium channel. Apamin completely inhibits the peristaltic reflex, and my data also show that the R- and N-type channels act in tandem to block the reflex. Therefore, N-type (or even PlQ-) channel-mediated ATP neurotransmission must account for the low frequency of the propagating peristaltic reflex. This would also help explain the sustained inhibition in the reverse experiment when NiCI2 was added first, followed by NICI2 and NLA. Tissue damage associated with inflammation, a cellular response to injury, evokes the production of a myriad of chemical mediators from specific cell types such as mast cells and lymphocytes at the site of injury (Levine 8. Reichling, 1999). The initial events in tissue inflammation, provoked by chemical, thermal, 227 or mechanical stimuli, involve the synthesis and release of such pro-inflammatory mediators that comprise an “inflammatory soup”. These compounds include, amongst others, protons, histamine, peptides (bradykinin, BK), lipids (prostaglandins), neurotransmitters (ATP and serotonin) and nerve growth factor (NGF) (Samad et al., 2002). Specifically, these mediators, via the activation of intracellular signal transduction pathways alter the sensitivity of voltage-gated ion channels to produce peripheral sensitization. Sensitization may also be triggered by inflammatory mediators from nociceptive neurons themselves (neurogenic inflammation) (Julius & Basbaum, 2001). Chronic inflammatory GI diseases include ulcerative colitis and Crohn’s disease and can be debilitating to affected patients. In Crohn’s disease, inflammatory mediators produce long-term changes that affect the mucosal and submucosal layers, and also neuromuscular transmission. In ulcerative colitis, the mucosal and submucosal are affected. Both those conditions produce rapid intestinal transit accompanied by abdominal cramping (Koch et al., 1988). Animals models using the nematode worm T. spiralis or the chemical trinitro benzene sulfonate have successfully reproduced inflammation in the GI tract. Inflammation depolarizes the resting membrane potentials of GI smooth muscle cells and increases their excitability. The excitability of enteric neurons is also increased with increased input resistance (Linden et al., 2003). Under conditions of inflammation, calcium influx through L- type calcium channels is suppressed, which in turn inhibits the calcium- calmodulin-dependent phosphorylation of the myosin laight chains, thus preventing contraction. The function of L-type calcium channels is also affected 228 by nuclear factor KB, a ubiquitous transcription factor responsible for immune and inflammatory reactions. Interestingly, the release of calcium from intracellular calcium stores are unaffected (Shi and Sarna, 2000; Shi and Sarna, 2006). According to Shi and Sama (2006), both neurotransmitter storage in, and release from enteric nerve terminals, are altered. Taken together, those results indicate that voltage-gated calcium channels are affected during inflammation, and that it is therefore possible to target them pharmacologically as viable control points. In my study, the R-type channel is clearly involved in intestinal motility, and it would be useful to study the function of those channels in the inflamed intestine. Inflammation increases GI transit, and I would therefore expect NiCI2 to inhibit the peristaltic response. The studies reported in this dissertation show that R-type channels are integral components of the ENS. The design of therapeutic strategies using NiCI2 is challenging due to its toxic nature. However, success has been achieved using toxins that block N-type calcium channels. Perhaps using SNX-482 may represent a feasible way of targeting these channels. Overall, the motility of the ileum in the presence of NiC|2 is impaired. This suggests that R-type channel antagonists may be useful as anti-diarrheal drugs, so that the lives of patients who suffer with GI diseases and disorders can be improved. 229 Figure. 7.1. Proposed Model for Inhibitory Neuromuscular transmission. (A1) R-type calcium channels may be present on the same nerve terminals as N- and PIQ-type calcium channels. R—channels couple to NO release, whereas N- and PIQ-type channels couple to ATP and/or B-NAD release. Both mechanisms produce relaxation of the smooth muscle. (A2) R-type calcium channels may be present on different nerve terminals compared to N- and PIQ-type calcium channels. 230 A1. RWCa" Charm N and \ _, PIQ channel R",.,,,;’,,’,,, SMOOTH MUSCLE N 8. P/Q‘ channel R-type I calcium - channel NOS 231 Figure. 7.2. Schematic of the overall effect of NiCI2 on the peristaltic reflex. Fluid induced distention in the presence of NiCI2 impairs peristalsis but its effects may extend to other regions of the GI as well. The response of the ileum under control conditions (A) after NLA (applied first) (B), and NLA + NiCI2 (C) treatments are shown. 232 \ Myenteric inhibitory motor nerve varicosity A1 . Control ...........,<_3o/Gi NE , . NOS NOS l ' N-type VDCC @ R-type VGCC NO NO Sympathetic I nerve terminal + cGMP guanylate cyclase I PKG Relaxation ‘/ GI smooth muscle A2. NLA first Myenteric inhibitory motor nerve varicosity ..........,C_5i/Go NLA a2AR NOS I NOS NOS Gig X N-type VDCC ® R-type vocc NO NO ' GI smooth muscle Sympathetic I nerve terminal GTP » cGMP guanylate cyclase I PKG Relaxation ‘/ 233 Figure. 7.2 cont’d A3. NLA + NiC|2 Myenteric inhibitory motor nerve varicosity NLA x X / 0LMR........§_5i/Go NE ’ ’,// \ N08 N08 nos I I I I I I I I I I I I I I I I I I I J / X N-type vocc 5 @ R-type VDCC NO .................. . NiC|2 P2Y1 receptor Sympathetic NO nerve terminal I + cGMP guanylate cyclase I PKG Relaxation / GI smooth muscle 234 REFERENCES Albillos A, Neher E & Moser T. (2000). R-Type Ca2+ channels are coupled to the rapid component of secretion in mouse adrenal slice chromaffin cells. J Neurosci 20, 8323-8330. Alex G, Clerc N, Kunze WA & Furness JB. (2002). Responses of myenteric S neurones to low frequency stimulation of their synaptic inputs. Neuroscience 110, 361-373. Alioua A, Tanaka Y, Wallner M, Hofmann F, Ruth P, Meera P & Toro L. (1998). The large conductance, voltage-dependent, and calcium-sensitive K+ channel, Hslo, is a target of cGMP-dependent protein kinase phosphorylation in vivo. J Biol Chem 273, 32950-32956. Altier C, Dubel SJ, Barrere C, Jarvis SE, Stotz SC, Spaetgens RL, Scott JD, Cornet V, De Waard M, Zamponi GW, Nargeot J & Bourinet E. (2002). Trafficking of L-type calcium channels mediated by the postsynaptic scaffolding protein AKAP79. J Biol Chem 277, 33598-33603. Amara SG & Arriza JL. (1993). Neurotransmitter transporters: three distinct gene families. Curr Opin Neumbiol 3, 337-344. Amara SG & Kuhar MJ. (1993). Neurotransmitter transporters: recent progress. Annu Rev Neurosci 16, 73-93. Andrews PL, Grundy D & Scratcherd T. (1980). Reflex excitation of antral motility induced by gastric distension in the ferret. J Physiol 298, 79-84. Barajas-Lopez C, Peres AL & Espinosa-Luna R. (1996). Cellular mechanisms underlying adenosine actions on cholinergic transmission in enteric neurons. Am J Physiol 271, C264-275. Barbara G, Stanghellini V, De Giorgio R, Cremon C, Cottrell GS, Santini D, Pasquinelli G, Morselli-Labate AM, Grady EF, Bunnett NW, Collins SM & Corinaldesi R. (2004). Activated mast cells in proximity to colonic nerves correlate with abdominal pain in irritable bowel syndrome. Gastroenterology 126, 693-702. ‘ Bayguinov O, Keef KD, Hagen B & Sanders KM. (1999). Parallel pathways mediate inhibitory effects of vasoactive intestinal polypeptide and nitric oxide in canine fundus. Br J Pharmacol 126, 1543-1552. Bayliss WM & Starling EH. (1899). The movements and innervation of the small intestine. J Physi0124, 99-143. 235 Bean BP. (1989). Classes of calcium channels in vertebrate cells. Annu Rev Physiol 51, 367-384. Benko R, Undi S, Wolf M, Magyar K, Tovolgyi Z, Rumbus Z & Bartho L. (2006). P(2) purinoceptors account for the non-nitrergic NANC relaxation in the rat ileum. Naunyn Schmiedebergs Arch Pharrnacol 373, 319-324. Benko R, Undi S, Wolf M, Vereczkei A, lllenyi L, Kassai M, Cseke L, Kelemen D, Horvath OP, Antal A, Magyar K & Bartho L. (2007). P2 purinoceptor antagonists inhibit the non-adrenergic, non-cholinergic relaxation of the human colon in vitro. Neuroscience 147, 146-152. Bergsman JB & Tsien RW. (2000). Syntaxin modulation of calcium channels in cortical synaptosomes as revealed by botulinum toxin C1. J Neurosci 20, 4368-4378. Bernstein CN, Niazi N, Robert M, Mertz H, Kodner A, Munakata J, Naliboff B & Mayer EA. (1996). Rectal afferent function in patients with inflammatory and functional intestinal disorders. Pain 66, 151-161. Bernstein GM & Jones OT. (2007). Kinetics of internalization and degradation of N-type voltage-gated calcium channels: role of the alpha2/delta subunit. Cell Calcium 41 , 27-40. Berridge MJ. (1998). Neuronal calcium signaling. Neuron 21, 13-26. Bertrand PP. (2003). ATP and sensory transduction in the enteric nervous system. Neuroscientist 9, 243-260. Bertrand PP & Galligan JJ. (1994). Contribution of chloride conductance increase to slow EPSC and tachykinin current in guinea-pig myenteric neurones. J Physiol 481 (Pt 1), 47-60. Bertrand PP & Galligan JJ. (1995). Signal-transduction pathways causing slow synaptic excitation in guinea pig myenteric AH neurons. Am J Physiol 269, G71 0-720. Bertrand PP, Kunze WA, Furness JB & Bornstein JC. (2000). The terminals of myenteric intrinsic primary afferent neurons of the guinea-pig ileum are excited by 5-hydroxytryptamine acting at 5-hydroxytryptamine-3 receptors. Neuroscience 101, 459-469. Bian X & Galligan JJ. (2007). Alpha2-adrenoceptors couple to inhibition of R-type calcium currents in myenteric neurons. NeurogastroenterolMotil19, 845- 855. 236 Bian X, Ren J, DeVries M, Schnegelsberg B, Cockayne DA, Ford AP & Galligan JJ. (2003). Peristalsis is impaired in the small intestine of mice lacking the P2X3 subunit. J Physiol 551, 309-322. Bian X, Zhou X & Galligan JJ. (2004). R-type calcium channels in myenteric neurons of guinea pig small intestine. Am J Physiol Gastrointest Liver Physiol 287, G134-142. Bian XC, Bertrand PP, Furness JB & Bornstein JC. (2000). Evidence for functional NK1-tachykinin receptors on motor neurones supplying the circular muscle of guinea-pig small and large intestine. Neurogastroenterol Motil 1 2, 307-31 5. Bichet D, Cornet V, Geib S, Carlier E, Volsen S, Hoshi T, Mori Y & De Waard M. (2000). The l-Il loop of the Ca2+ channel alpha1 subunit contains an endoplasmic reticulum retention signal antagonized by the beta subunit. Neuron 25, 177-190. Blakeley AG & Cunnane TC. (1979). The packeted release of transmitter from the sympathetic nerves of the guinea-pig vas deferens: an electrophysiological study. J Physiol 296, 85-96. Blakeley AG, Cunnane TC & Petersen SA. (1982). Local regulation of transmitter release from rodent sympathetic nerve terminals? J Physiol 325, 93-109. Bogeski G, Shafton AD, Kitchener PD, Ferens DM & Furness JB. (2005). A quantitative approach to recording peristaltic activity from segments of rat small intestine in vivo. NeurogastroenterolMotil17, 262-272. Bornstein JC, Furness JB & Kunze WA. (1994). Electrophysiological characterization of myenteric neurons: how do classification schemes relate? J Auton Nerv Syst 48, 1-15. Bourinet E, Soong TW, Sutton K, Slaymaker S, Mathews E, Monteil A, Zamponi GW, Nargeot J & Snutch TP. (1999). Splicing of alpha 1A subunit gene generates phenotypic variants of P- and Q-type calcium channels. Nat Neurosci 2, 407-415. Boyer JL, Mohanram A, Camaioni E, Jacobson KA & Harden TK. (1998). Competitive and selective antagonism of P2Y1 receptors by N6-methyl 2'- deoxyadenosine 3',5'-bisphosphate. BrJ Pharmacol 124, 1-3. Bredt DS, Hwang PM & Snyder SH. (1990). Localization of nitric oxide synthase indicating a neural role for nitric oxide. Nature 347, 768-770. 237 Brehmer A, Schrodl F & Neuhuber W. (1999). Morphological classifications of enteric neurons-1OO years after Dogiel. Anat Embryol (Bell) 200, 125- 135. Brelje TC, Wessendorf MW & Sorenson RL. (1993). Multicolor laser scanning confocal immunofluorescence microscopy: practical application and limitations. Methods Cell Biol 38, 97-181. Brice NL, Berrow NS, Campbell V, Page KM, Brickley K, Tedder l & Dolphin AC. (1997). Importance of the different beta subunits in the membrane expression of the alpha1A and alpha2 calcium channel subunits: studies using a depolarization-sensitive alpha1A antibody. EurJ Neurosci 9, 749- 759. Brookes SJ. (2001). Classes of enteric nerve cells in the guinea-pig small intestine. Anat Rec 262, 58-70. Brookes SJ & Costa M. (1990). Identification of enteric motor neurones which innervate the circular muscle of the guinea pig small intestine. Neurosci Lett 118, 227-230. Brookes SJ, Hennig G & Schemann M. (1998). Identification of motor neurons to the circular muscle of the guinea pig gastric corpus. J Comp Neurol 397, 268-280. Brookes SJ, Meedeniya AC, Jobling P & Costa M. (1997). Orally projecting interneurones in the guinea-pig small intestine. J Physiol 505 (Pt 2), 473- 491. Brookes SJ, Song ZM, Ramsay GA & Costa M. (1995). Long aboral projections of Dogiel type II, AH neurons within the myenteric plexus of the guinea pig small intestine. J Neurosci 15, 4013-4022. Brookes SJ, Song ZM, Steele PA & Costa M. (1992). Identification of motor neurons to the longitudinal muscle of the guinea pig ileum. Gastroenterology 103, 961 -973. Brookes SJ, Steele PA & Costa M. (1991a). Calretinin immunoreactivity in cholinergic motor neurones, interneurones and vasomotor neurones in the guinea-pig small intestine. Cell Tissue Res 263, 471-481. Brookes SJ, Steele PA & Costa M. (1991b). Identification and immunohistochemistry of cholinergic and non-cholinergic circular muscle motor neurons in the guinea-pig small intestine. Neuroscience 42, 863- 878. 238 :m ’1. Brookes SJH & Costa M. (2006). Functional Histoanatomy of the Enteric Nervous System. In Physiology of the Gastrointestinal Tract, ed. Johnson RJ, Barret KE, Merchant JL, Ghishan F K, Said HM & Wood JD, pp. 577-602. Elsevier Academic Press, Burlington, MA. Bfilbring E, Crema A & Saxby OB. (1958). A method for recording peristalsis in isolated intestine. British Journal of Pharmacology 13, 440-443. BiJlbring E & Gershon MD. (1967). 5-hydroxytryptamine participation in the vagal inhibitory innervation of the stomach. J Physio/192, 823-846. Burnstock G. (2008). The journey to establish purinergic signalling in the gut. Neurogastroenterol Motil 20 Suppl 1 , 8-19. Burnstock G, Campbell G, Bennett M & Holman ME. (1963). Inhibition Of The Smooth Muscle On The Taenia Coli. Nature 200, 581-582. Burnstock G, Campbell G, Bennett M & Holman ME. (1964). Innervation Of The Guinea-Pig Taenia Coll: Are There Intrinsic Inhibitory Nerves Which Are Distinct From Sympathetic Nerves? Int J Neuropharmacol 3, 163-166. Burnstock G, Campbell G, Satchell D & Smythe A. (1997). Evidence that adenosine triphosphate or a related nucleotide is the transmitter substance released by non-adrenergic inhibitory nerves in the gut. 1970. EN Phannacol120, 337-357; discussion 334-336. Burnstock G & Holman ME. (1963). Smooth muscle: autonomic nerve transmission. Annu Rev Physiol 25, 61-90. Bywater RA & Taylor GS. (1986). Non-cholinergic excitatory and inhibitory junction potentials in the circular smooth muscle of the guinea-pig ileum. J Physiol 374, 1 53-164. Camilleri M. (2001). Management of the irritable bowel syndrome. Gastroenterology 120, 652-668. Camilleri M, Dubois D, Coulie B, Jones M, Kahrilas PJ, Rentz AM, Sonnenberg A, Stanghellini V, Stewart WF, Tack J, Talley NJ, Whitehead W & Revicki DA. (2005). Prevalence and socioeconomic impact of upper gastrointestinal disorders in the United States: results of the US Upper Gastrointestinal Study. Clin Gastroenterol Hepatol 3, 543-552. Carlson GM, Bedi BS & Code CF. (1972). Mechanism of propagation of intestinal interdigestive myoelectric complex. Am J Physiol 222, 1027-1030. 239 Castellano A, Wei X, Birnbaumer L & Perez-Reyes E. (1993). Cloning and expression of a neuronal calcium channel beta subunit. J Biol Chem 268, 12359-12366. Catterall WA. (2000). Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol 16, 521-555. Catterall WA & Few AP. (2008). Calcium channel regulation and presynaptic plasticity. Neuron 59, 882-901. Catterall WA, Perez-Reyes E, Snutch TP & Striessnig J. (2005). International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev 57, 411- 425. Cervero F. (1994). Sensory innervation of the viscera: peripheral basis of visceral pain. Physiol Rev 74, 95-138. Chen RS. (2006). Calcium channel gamma-6 subunit: in search of its functions and mechanisms of actions. In Neuroscience, pp. 7. University of Illinois at Urbana-Champaign, Illinois. Chen WP & Kirchgessner AL. (2002). Activation of group II mGlu receptors inhibits voltage—gated Ca2+ currents in myenteric neurons. Am J Physiol Gastrointest Liver Physiol 283, G1282-1289. Ciccocioppo R, Onori L, Messori E, Candura SM, Coccini T & Tonini M. (1994). Role of nitric oxide-dependent and -independent mechanisms in peristalsis and accommodation in the rabbit distal colon. J Phannacol Exp Ther 270, 929-937. Clerc N, Furness JB, Kunze WA, Thomas EA & Bertrand PP. (1999). Long-term effects of synaptic activation at low frequency on excitability of myenteric AH neurons. Neuroscience 90, 279-289. Clerc N, F urness JB, Li ZS, Bornstein JC & Kunze WA. (1998). Morphological and immunohistochemical identification of neurons and their targets in the guinea-pig duodenum. Neuroscience 86, 679-694. Cohen R & Atlas D. (2004). R-type voltage-gated Ca(2+) channel interacts with synaptic proteins and recruits synaptotagmin to the plasma membrane of Xenopus oocytes. Neuroscience 128, 831-841. Cohen R, Elferink LA & Atlas D. (2003). The CZA domain of synaptotagmin alters the kinetics of voltage-gated Ca2+ channels Ca(v)1.2 (Lo-type) and Ca(v)2.3 (R-type). J Biol Chem 278, 9258-9266. 240 Cooper EC & Jan LY. (1999). Ion channel genes and human neurological disease: recent progress, prospects, and challenges. Proc Natl Acad Sci U S A 96, 4759-4766. Costa M, Brookes SJ & Hennig GW. (2000). Anatomy and physiology of the enteric nervous system. Gut 47 Suppl 4, iv15-19; discussion iv26. Costa M, Brookes SJ, Steele PA, Gibbins I, Burcher E & Kandiah CJ. (1996). Neurochemical classification of myenteric neurons in the guinea-pig ileum. Neuroscience 75, 949-967. Costa M 8 Furness JB. (1982). Neuronal peptides in the intestine. Br Med Bull 38, 247-252. Costa M & Furness JB. (1983). The origins, pathways and terminations of neurons with VIP-like immunoreactivity in the guinea-pig small intestine. Neuroscience 8, 665-676. Costa M & Furness JB. (1984). Somatostatin is present in a subpopulation of noradrenergic nerve fibres supplying the intestine. Neuroscience 13, 911- 919. Costa M, Furness JB, Cuello AC, Verhofstad AA, Steinbusch HW & Elde RP. (1982). Neurons with 5—hydroxytryptamine-like immunoreactivity in the enteric nervous system: their visualization and reactions to drug treatment. Neuroscience 7, 351-363. Costa M, Furness JB & Gabella G. (1971). Catecholamine containing nerve cells in the mammalian myenteric plexus. Histochemie 25, 103-106. Costa M, Furness JB & Gibbins IL. (1986a). Chemical coding of enteric neurons. Prog Brain Res 68, 217-239. Costa M, Furness JB & Humphreys CM. (1986b). Apamin distinguishes two types of relaxation mediated by enteric nerves in the guinea-pig gastrointestinal tract. Naunyn Schmiedebergs Arch Phannacol 332, 79-88. Costa M, Furness JB, Pompolo S, Brookes SJ, Bornstein JC, Bredt DS & Snyder SH. (1992). Projections and chemical coding of neurons with immunoreactivity for nitric oxide synthase in the guinea-pig small intestine. Neurosci Lett 148, 121-125. Crist JR, He XD & Goyal RK. (1991). The nature of noncholinergic membrane potential responses to transmural stimulation in guinea pig ileum. Gastroenterology 1 00, 1006-1 015. 241 Crist JR, He XD & Goyal RK. (1992). Both ATP and the peptide VIP are inhibitory neurotransmitters in guinea-pig ileum circular muscle. J Physiol 447, 119- 131. Cunningham SM, Mihara S & Higashi H. (1998). Presynaptic calcium channels mediating synaptic transmission in submucosal neurones of the guinea- pig caecum. J Physiol 509 (Pt 2), 425-435. Cuppoletti J, Malinowska DH, Tewari KP, Li QJ, Sherry AM, Patchen ML & Ueno R. (2004). SPI-0211 activates T84 cell chloride transport and recombinant human ClC-2 chloride currents. Am J Physiol Cell Physiol 287, C1173- 1 183. Davies A, Hendrich J, Van Minh AT, Wratten J, Douglas L & Dolphin AC. (2007). Functional biology of the alpha(2)delta subunits of voltage-gated calcium channels. Trends Pharmacol Sci 28, 220-228. de Backer O, Blanckaert B, Leybaert L & Lefebvre RA. (2008). A novel method for the evaluation of intestinal transit and contractility in mice using fluorescence imaging and spatiotemporal motility mapping. Neurogastroenterol Motil 20, 700-707. De Laet A, Adriaensen D, Van Bogaert PP, Scheuermann DW & Timmermans JP. (2002). Immunohistochemical localization of voltage-activated calcium channels in the rat oesophagus. Neurogastroenterol Motil 14, 173-181. De Man JG, De Winter BY, Seerden TC, De Schepper HU, Herman AG & Pelckmans PA. (2003). Functional evidence that ATP or a related purine is an inhibitory NANC neurotransmitter in the mouse jejunum: study on the identity of P2X and P2Y purinoceptors involved. BrJ Phannacol140, 1 108-1 1 16. Deutch AY & Roth RH. (1999). Neurotransmitters. In Fundamental Neuroscience, ed. Zigmond M, Bloom FE, Landis SC, Roberts JL & Squire LR, pp. 193- 234. Academic Press, San Diego, California. Dietrich D, Kirschstein T, Kukley M, Pereverzev A, von der Brelie C, Schneider T & Beck H. (2003). Functional specialization of presynaptic Cav2.3 Ca2+ channels. Neuron 39, 483-496. Dolphin AC. (2003a). Beta subunits of voltage-gated calcium channels. J Bioenerg Biomembr 35, 599-620. Dolphin AC. (2003b). G protein modulation of voltage-gated calcium channels. Pharmacol Rev 55, 607-627. 242 Drossman DA. (1999). The functional gastrointestinal disorders and the Rome ll process. Gut 45 Suppl 2, ll1-5. Drossman DA, Camilleri M, Mayer EA & Whitehead WE. (2002). AGA technical review on irritable bowel syndrome. Gastroenterology 123, 2108-2131. Drossman DA, Li Z, Andruzzi E, Temple RD, Talley NJ, Thompson WG, Whitehead WE, Janssens J, Punch-Jensen P, Corazziari E & et al. (1993). US. householder survey of functional gastrointestinal disorders. Prevalence, sociodemography, and health impact. Dig Dis Sci 38, 1569- 1 580. Eglen RM. (2001). Muscarinic receptors and gastrointestinal tract smooth muscle function. Life Sci 68, 2573-2578. el-Husseini Ael D & Bredt DS. (2002). Protein palmitoylation: a regulator of neuronal development and function. Nat Rev Neurosci 3, 791 -802. Ertel EA, Campbell KP, Harpold MM, Hofmann F, Mori Y, Perez-Reyes E, Schwartz A, Snutch TP, Tanabe T, Birnbaumer L, Tsien RW & Catterall WA. (2000). Nomenclature of voltage-gated calcium channels. Neuron 25, 533-535. Evans GJ & Cousin MA. (2005). Tyrosine phosphorylation of synaptophysin in synaptic vesicle recycling. Biochem Soc Trans 33, 1350-1353. Fahrenkrug J, Schaffalitzky de Muckadell OB, Holst JJ & Jensen SL. (1979). Vasoactive intestinal polypeptide in vagally mediated pancreatic secretion of fluid and HCO3. Am J Physiol 237, E535-540. Fang 2, Park CK, Li HY, Kim HY, Park SH, Jung SJ, Kim JS, Monteil A, Oh SB & Miller RJ. (2007). Molecular basis of Ca(v)2.3 calcium channels in rat nociceptive neurons. J Biol Chem 282, 4757-4764. Farre R, Auli M, Lecea B, Martinez E & Clave P. (2006). Phannacologic characterization of intrinsic mechanisms controlling tone and relaxation of porcine lower esophageal sphincter. J Pharmacol Exp Ther 316, 1238- 1248. Farrugia G. (1999). Ionic conductances in gastrointestinal smooth muscles and interstitial cells of Cajal. Annu Rev Physiol 61, 45-84. Farrugia G. (2008). Interstitial cells of Cajal in health and disease. Neurogastroenterol Motil 20 Suppl 1, 54-63. 243 Foehring RC, Mermelstein PG, Song WJ, Ulrich S & Surrneier DJ. (2000). Unique properties of R-type calcium currents in neocortical and neostriatal neurons. J Neurophysiol 84, 2225-2236. Fox AP, Nowycky MC & Tsien RW. (1987). Kinetic and pharmacological properties distinguishing three types of calcium currents in chick sensory neurones. J Physiol 394, 149-172. Franklin JL, Fickbohm DJ & Willard AL. (1992). Long-tenn regulation of neuronal calcium currents by prolonged changes of membrane potential. J Neurosci 12, 1726-1735. Franklin JL & Willard AL. (1993). Voltage-dependent sodium and calcium currents of rat myenteric neurons in cell culture. J Neurophysi0169, 1264- 1275. Franzusoff A, Redding K, Crosby J, Fuller RS & Schekman R. (1991). Localization of components involved in protein transport and processing through the yeast Golgi apparatus. J Cell Biol 112, 27-37. French RJ & Zamponi GW. (2005). Voltage-gated sodium and calcium channels in nerve, muscle, and heart. IEEE Trans Nanobioscience 4, 58-69. Frigo GM, Lecchini S, Marcoli M, Tonini M, D'Angelo L 8: Crema A. (1984). Changes in sensitivity to the inhibitory effects of adrenergic agonists on intestinal motor activity after chronic sympathetic denervation. Naunyn Schmiedebergs Arch Pharmacol 325, 145-152. Furness JB. (2000). Types of neurons in the enteric nervous system. J Auton Nerv Syst 81, 87-96. ' Furness JB. (2006a). The Enten'c Nervous System. Blackwell Publishing, Melbourne, Victoria, Australia. Furness JB. (2006b). Novel gut afferents: Intrinsic afferent neurons and intestinofugal neurons. Auton Neurosci 1 25, 81 -85. Furness JB, Alex G, Clark MJ & Lal W. (2003). Morphologies and projections of defined classes of neurons in the submucosa of the guinea-pig small intestine. Anat Rec A Discov Mol Cell Evol Biol 272, 475-483. Furness JB, Bornstein JC, Pompolo S, Young HM, Kunze WA, Yuan SY & Kelly H. (1993). The nerve circuits for motility control in the gastrointestinal tract. J Smooth Muscle Res 29, 143-144. 244 Furness JB & Costa M. (1971). Morphology and distribution of intrinsic adrenergic neurones in the proximal colon of the guinea-pig. Z Zellforsch Mikrosk Anat 120, 346-363. Furness JB & Costa M. (1987). The enteric nervous system. Churchill Livingstone, Adelaide. Furness JB, Costa M & Eckenstein F. (1983). Neurones localized with antibodies against choline acetyltransferase in the enteric nervous system. Neurosci Lett 40, 105-109. F urness JB, Costa M, Rokaeus A, McDonald TJ & Brooks B. (1987). Galanin- immunoreactive neurons in the guinea-pig small intestine: their projections and relationships to other enteric neurons. Cell Tissue Res 250, 607-615. Furness JB, Jones C, Nurgali K & Clerc N. (2004). Intrinsic primary afferent neurons and nerve circuits within the intestine. Prog Neurobiol 72, 143- 164. Furness JB, Keast JR, Pompolo S, Bornstein JC, Costa M, Emson PC & Lawson DE. (1988). lmmunohistochemical evidence for the presence of calcium- binding proteins in enteric neurons. Cell Tissue Res 252, 79-87. Furness JB, Kunze WA, Bertrand PP, Clerc N & Bornstein JC. (1998). Intrinsic primary afferent neurons of the intestine. Prog Neurobiol 54, 1-18. F urness JB, Kunze WA & Clerc N. (1999). Nutrient tasting and signaling mechanisms in the gut. II. The intestine as a sensory organ: neural, endocrine, and immune responses. Am J Physiol 277, G922-928. F urness JB, Lloyd KC, Sternini C & Walsh JH. (1990). Projections of substance P, vasoactive intestinal peptide and tyrosine hydroxylase immunoreactive nerve fibres in the canine intestine, with special reference to the innervation of the circular muscle. Arch Histol Cytol 53, 129-140. F urness JB, Pompolo S, Murphy R 8. Giraud A. (1989). Projections of neurons with neuromedin U-like immunoreactivity in the small intestine of the guinea-pig. Cell Tissue Res 257, 415-422. Furness JB, Young HM, Pompolo S, Bornstein JC, Kunze WA & McConalogue K. (1995). Plurichemical transmission and chemical coding of neurons in the digestive tract. Gastroenterology108, 554-563. Gallego D, Hernandez P, Clave P & Jimenez M. (2006). P2Y1 receptors mediate inhibitory purinergic neuromuscular transmission in the human colon. Am J Physiol Gastrointest Liver Physiol 291, G584-594. 245 Galligan JJ. (1998). Mechanisms of excitatory synaptic transmission in the enteric nervous system. Tokai J Exp Clin Med 23, 129-136. Galligan JJ. (1999). Nerve terminal nicotinic cholinergic receptors on excitatory motoneurons in the myenteric plexus of guinea pig intestine. J Pharmacol Exp Ther 291, 92-98. Galligan JJ. (2002a). Ligand-gated ion channels in the enteric nervous system. Neurogastroenterol Motil 14, 611-623. Galligan JJ. (2002b). Pharmacology of synaptic transmission in the enteric nervous system. Curr Opin Pharmacol 2, 623-629. Galligan JJ. (2004). Enteric P2X receptors as potential targets for drug treatment of the irritable bowel syndrome. BrJ Pharmacol141, 1294-1302. Galligan JJ 8 Bertrand PP. (1994). ATP mediates fast synaptic potentials in enteric neurons. J Neurosci14, 7563-7571. Galligan JJ, Furness JB 8 Costa M. (1989). Migration of the myoelectric complex after interruption of the myenteric plexus: intestinal transaction and regeneration of enteric nerves in the guinea pig. Gastroenterology 97, 1 135-1 146. Galligan JJ, LePard KJ, Schneider DA 8 Zhou X. (2000). Multiple mechanisms of fast excitatory synaptic transmission in the enteric nervous system. J Auton Nerv Syst 81, 97-103. Galligan JJ, Tatsumi H, Shen KZ, Surprenant A 8 North RA. (1990). Cation current activated by hyperpolarization (IH) in guinea pig enteric neurons. Am J Physiol 259, G966-972. 630 N, Hu HZ, Zhu MX, Fang X, Liu S, Gao C & Wood JD. (2006). The PZY purinergic receptor expressed by enteric neurones in guinea-pig intestine. Neurogastroenterol Motil 18, 316-323. Gasparini S, Kasyanov AM, Pietrobon D, Voronin LL 8 Cherubini E. (2001). Presynaptic R-type calcium channels contribute to fast excitatory synaptic transmission in the rat hippocampus. J Neurosci 21, 8715-8721. Gazzard BG. (1988). HIV disease and the gastroenterologist. Gut 29, 1497-1505. Geiselhoringer A, Gaisa M, Hofmann F 8 Schlossmann J. (2004). Distribution of IRAG and cGKl-isofonns in murine tissues. FEBS Lett 575, 19-22. 246 Gershon MD. (1981 ). The enteric nervous system. Annual Review of Neuroscience 4, 227-272. Gershon MD. (1997). Genes and lineages in the formation of the enteric nervous system. Curr Opin Neurobiol 7, 101-109. Glickstein M. (2006). Golgi and Cajal: The neuron doctrine and the 100th anniversary of the 1906 Nobel Prize. Curr Biol 16, R147-151. Goda Y 8 Davis GW. (2003). Mechanisms of synapse assembly and disassembly. Neuron 40, 243-264. Goyal RK 8 Hirano l. (1996). The enteric nervous system. N Engl J Med 334, 1106-1115. Grabsch H, Pereverzev A, Weiergraber M, Schramm M, Henry M, Vajna R, Beattie RE, Volsen SG, Klockner U, Hescheler J 8 Schneider T. (1999). lmmunohistochemical detection of alpha1 E voltage-gated Ca(2+) channel isoforms in cerebellum, INS-1 cells, and neuroendocrine cells of the digestive system. The Journal of Histochemistry 8 Cytochemistry 47, 981 - 993. Grider JR 8 Jin JG. (1993). Vasoactive intestinal peptide release and L-citrulline production from isolated ganglia of the myenteric plexus: evidence for regulation of vasoactive intestinal peptide release by nitric oxide. Neuroscience 54, 521-526. Grima B, Lamouroux A, Blanot F, Biguet NF 8 Mallet J. (1985). Complete coding sequence of rat tyrosine hydroxylase mRNA. Proc Natl Acad Sci U S A 82, 617-621. Grundy D, Al-Chaer ED, Aziz Q, Collins SM, Ke M, Tache Y 8 Wood JD. (2006). Fundamentals of neurogastroenterology: basic science. Gastroenterology 130, 1391-1411. Guard S, Watling KJ 8 Watson SP. (1988). Neurokinin3-receptors are linked to inositol phospholipid hydrolysis in the guinea-pig ileum longitudinal muscle-myenteric plexus preparation. BrJ Pharmacol 94, 148-154. Guillery RW. (2005). Observations of synaptic structures: origins of the neuron doctrine and its current status. Philos Trans R Soc Lond B Biol Sci 360, 1281 -1 307. Gwynne RM 8 Bornstein JC. (2007). Mechanisms underlying nutrient-induced segmentation in isolated guinea pig small intestine. Am J Physiol Gastrointest Liver Physiol 292, G1 162-1 172. 247 Hagiwara S 8 Byerly L. (1981a). Calcium channel. Annu Rev Neurosci 4, 69-125. Hagiwara S 8 Byerly L. (1981b). Membrane biophysics of calcium currents. Fed Proc 40, 2220-2225. Hansen MB. (2003). The enteric nervous system III: a target for pharmacological treatment. Pharmacol Toxicol 93, 1-13. He XD 8 Goyal RK. (1993). Nitric oxide involvement in the peptide VIP- associated inhibitory junction potential in the guinea-pig ileum. J Physiol 461, 485-499. Hein L. (2006). Adrenoceptors and signal transduction in neurons. Cell Tissue Res 326, 541-551. Hennig GW, Costa M, Chen BN 8 Brookes SJ. (1999). Quantitative analysis of peristalsis in the guinea-pig small intestine using spatio-temporal maps. J Physiol 517 (Pt 2), 575—590. Herrington J, Zhou YP, Bugianesi RM, Dulski PM, Feng Y, Warren VA, Smith MM, Kohler MG, Garsky VM, Sanchez M, Wagner M, Raphaelli K, Banerjee P, Ahaghotu C, Wunderler D, Priest BT, Mehl JT, Garcia ML, McManus OB, Kaczorowski GJ 8 Slaughter RS. (2006). Blockers of the delayed-rectifier potassium current in pancreatic beta-cells enhance glucose-dependent insulin secretion. Diabetes 55, 1034-1042. Hille B. (2001). Voltage-gated calcium channels. Sinauer Associates, Sunderland, MA. Hillman D, Chen S, Aung TT, Cherksey B, Sugimori M 8 Llinas RR. (1991). Localization of P-type calcium channels in the central nervous system. Proc Natl Acad Sci U S A 88, 7076-7080. Hillsley K, Kenyon JL 8 Smith TK. (2000). Ryanodine-sensitive stores regulate the excitability of AH neurons in the myenteric plexus of guinea-pig ileum. J Neurophysiol 84, 2777-2785. Hirning LD, Fox AP, McCleskey EW, Olivera BM, Thayer SA, Miller RJ 8 Tsien RW. (1 988). Dominant role of N-type Ca2+ channels in evoked release of norepinephrine from sympathetic neurons. Science 239, 57-61. Hirning LD, Fox AP 8 Miller RJ. (1990). Inhibition of calcium currents in cultured myenteric neurons by neuropeptide Y: evidence for direct receptor/channel coupling. Brain Res 532, 120-130. 248 Hirst GD, Bywater RA, Teramoto N 8 Edwards FR. (2004). An analysis of inhibitory junction potentials in the guinea-pig proximal colon. J Physiol 558, 841-855. Hirst GD, Johnson SM 8 van Helden DF. (1985). The calcium current in a myenteric neurone of the guinea-pig ileum. J Physiol 361, 297-314. Hirst GD 8 McKirdy HC. (1974). Presynaptic inhibition at mammalian peripheral synapse? Nature 250, 430-431. Hirst GDS, Holman ME 8 Spence I. (1974). Two types of neurones in the myenteric plexus of duodenum in the guinea-pig. Journal of Physiology 236, 303-326. Hoffman F, Biel M 8 Flockerzi V. (1994). Molecular Basis for Ca2+ Channel Diversity. Annual Review of Neuroscience 17, 399-418. Hofmann F, Fell R, Kleppisch T 8 Schlossmann J. (2006). Function of cGMP- dependent protein kinases as revealed by gene deletion. Physiol Rev 86, 1-23. Hofmann F, Lacinova L 8 Klugbauer N. (1999). Voltage-dependent calcium channels: from structure to function. Rev Physiol Biochem Pharmacol 139, 33-87. Holscher C. (1997). Nitric oxide, the enigmatic neuronal messenger: its role in synaptic plasticity. Trends Neurosci 20, 298—303. Holst MC, Kelly JB 8 Powley TL. (1997). Vagal preganglionic projections to the enteric nervous system characterized with Phaseolus vulgaris- leucoagglutinin. J Comp Neurol 381, 81-100. Holzer P 8 Holzer-Petsche U. (2001). Tachykinin receptors in the gut: physiological and pathological implications. Curr Opin Phannacol1, 583- 590. Holzer P, Lippe IT, Tabrizi AL, Lenard L, Jr. 8 Bartho L. (1997). Dual excitatory and inhibitory effect of nitric oxide on peristalsis in the guinea pig intestine. J Pharmacol Exp Ther 280, 154-161. Horowitz A, Menice CB, Laporte R 8 Morgan KG. (1996). Mechanisms of smooth muscle contraction. Physiol Rev 76, 967-1003. Huizinga JD, Ambrous K 8 Der-Silaphet T. (1998). Co-operation between neural and myogenic mechanisms in the control of distension-induced peristalsis in the mouse small intestine. J Physiol 506 (Pt 3), 843-856. 249 Huizinga JD 8 Lammers WJ. (2009). Gut peristalsis is governed by a multitude of cooperating mechanisms. Am J Physiol Gastrointest Liver Physiol 296, G1-8. Huizinga JD, McKay CM 8 White EJ. (2006). The many facets of intestinal peristalsis. Am J Physiol Gastrointest Liver Physiol 290, G1347-1349; author reply 61348-1349. lsom LL, De Jongh KS 8 Catterall WA. (1994). Auxiliary subunits of voltage- gated ion channels. Neuron 12, 1183-1194. Isomura Y, Fujiwara-Tsukamoto Y, Imanishi M, Nambu A 8 Takada M. (2002). Distance-dependent Ni(2+)-sensitivity of synaptic plasticity in apical dendrites of hippocampal CA1 pyramidal cells. J Neurophysiol 87, 1169- 1174. lzzo AA, Costa M, Mascolo N 8 Capasso F. (1998). The role of histamine H1, H2 and H3 receptors on enteric ascending synaptic transmission in the guinea pig ileum. J Pharmacol Exp Ther 287, 952-957. Jarvis SE 8 Zamponi GW. (2001). Interactions between presynaptic Ca2+ channels, cytoplasmic messengers and proteins of the synaptic vesicle release complex. Trends Pharmacol Sci 22, 519-525. Jessell TM 8 Kandel ER. (1993). Synaptic transmission: a bidirectional and self- modifiable form of cell-cell communication. Cell 72 Suppl, 1-30. Jun JY, Choi S, Yeum CH, Chang W, You HJ, Park CK, Kim MY, Kong ID, Kim MJ, Lee KP, So I 8 Kim KW. (2004). Substance P induces inward current and regulates pacemaker currents through tachykinin NK1 receptor in cultured interstitial cells of Cajal of murine small intestine. EurJ Pharmacol 495, 35-42. Kaczorowski GJ, McManus OB, Priest BT 8 Garcia ML. (2008). Ion channels as drug targets: the next GPCRs. J Gen Physiol131, 399-405. Kang SH, Vanden Berghe P 8 Smith TK. (2003). Ca2+-activated Cl- current in cultured myenteric neurons from murine proximal colon. Am J Physiol Cell Physiol 284, C839-847. Katsoulis S, Schmidt WE, Schwarzhoff R, Folsch UR, Jin JG, Grider JR 8 Makhlouf GM. (1996). Inhibitory transmission in guinea pig stomach mediated by distinct receptors for pituitary adenylate cyclase-activating peptide. J Pharmacol Exp Ther 278, 199-204. 250 Kavalali ET, Zhuo M, Bito H 8 Tsien RW. (1997). Dendritic Ca2+ channels characterized by recordings from isolated hippocampal dendritic segments. Neuron 18, 651-663. Kawasaki J, Kobayashi S, Miyagi Y, Nishimura J, Fujishima M 8 Kanaide H. (1997). The mechanisms of the relaxation induced by vasoactive intestinal peptide in the porcine coronary artery. BrJ Pharmacol121 , 977-985. Kelly RB. (1993). Storage and release of neurotransmitters. Cell 72 Suppl, 43- 53. Kim CD, Goyal RK 8 Mashimo H. (1999). Neuronal NOS provides nitrergic inhibitory neurotransmitter in mouse lower esophageal sphincter. Am J Physiol 277, 6280-284. Kim E 8 Sheng M. (2004). PDZ domain proteins of synapses. Nat Rev Neurosci 5, 771-781. Kirchgessner AL 8 Gershon MD. (1988). Projections of submucosal neurons to the myenteric plexus of the guinea pig intestine: in vitro tracing of microcircuits by retrograde and anterograde transport. J Comp Neurol 277, 487-498. Kirchgessner AL 8 Gershon MD. (1989). Identification of vagal efferent fibers and putative target neurons in the enteric nervous system of the rat. J Comp Neurol 285, 38-53. Kirchgessner AL 8 Liu MT. (1999). Differential localization of Ca2+ channel alpha1 subunits in the enteric nervous system: presence of alpha1 B channel-like immunoreactivity in intrinsic primary afferent neurons. J Comp Neurol 409, 85-104. Kirchgessner AL, Tamir H 8 Gershon MD. (1992). Identification and stimulation by serotonin of intrinsic sensory neurons of the submucosal plexus of the guinea pig gut: activity-induced expression of Fos immunoreactivity. J Neurosci12, 235-248. Kirkup AJ, Brunsden AM 8 Grundy D. (2001). Receptors and transmission in the brain-gut axis: potential for novel therapies. l. Receptors on visceral afferents. Am J Physiol Gastrointest Liver Physiol 280, G787-794. Klugbauer N, Marais E 8 Hofmann F. (2003). Calcium channel alpha2delta subunits: differential expression, function, and drug binding. J Bioenerg Biomembr 35, 639-647. 251 ”4.5-2 "Q‘j'nu ' ’ 'V"Vf;' -o: ' j 'l.‘, ;;~ : , I .r i . ,;.,..o. ”no ., ..,.| . fln.ufig#-;W-I‘I{d3%IM-&Inflflm1.1,“. u .I:f.-;,5r.l‘...-l-r.g:' .}.' I.I - "3' .r. a. . if“. '-,',.,..n ‘ "s ,' n. 'z-fr; “,4,“ "'1'r‘i'o "ml. at Kristufek D, Stocker E, Boehm S 8 Huck S. (1999). Somatic and prejunctional nicotinic receptors in cultured rat sympathetic neurones show different agonist profiles. J Physiol 516 (Pt 3), 739-756. Kubota M, Murakoshi T, Saegusa H, Kazuno A, Zong S, Hu Q, Noda T 8 Tanabe T. (2001). Intact LTP and fear memory but impaired spatial memory in mice lacking Ca(v)2.3 (alpha(lE)) channel. Biochem Biophys Res Commun 282, 242-248. Kunze WA, Bornstein JC 8 Furness JB. (1995). Identification of sensory nerve cells in a peripheral organ (the intestine) of a mammal. Neuroscience 66, 1-4. Kunze WA, Bornstein JC, Furness JB, Hendriks R 8 Stephenson DS. (1994). Charybdotoxin and iberiotoxin but not apamin abolish the slow after- hyperpolarization in myenteric plexus neurons. Pflugers Arch 428, 300- 306. Kunze WA, Clerc N, Furness JB 8 Gola M. (2000). The soma and neurites of primary afferent neurons in the guinea-pig intestine respond differentially to deformation. J Physi01526 Pt 2, 375-385. Kunze WA, F urness JB, Bertrand PP 8 Bornstein JC. (1998). Intracellular recording from myenteric neurons of the guinea-pig ileum that respond to stretch. J Physiol 506 (Pt 3), 827-842. Kunze WAA 8 Furness JB. (1999). The enteric nervous system and regulation of intestinal motility. Annual Review of Physiology 61, 1 17-142. Kurjak M, F ritsch R, Saur D, Schusdziarra V 8 Allescher HD. (2001). Functional coupling between nitric oxide synthesis and VIP release within enteric nerve terminals of the rat: involvement of protein kinase G and phosphodiesterase 5. J Physiol 534, 827-836. Kwon S, Chung S, Ahn D, Yeon D 8 Nam T. (2001). Mechanism of carbon monoxide-induced relaxation in the guinea pig ileal smooth muscle. J Vet Med Sci 63, 389-393. Lai KO 8 Lp NY. (2003). Postsynaptic signaling of new players at the neuromuscular junction. J Neurocytol 32, 727-741. Langley JN. (1 922). The nerve fibre constitution of peripheral nerves and of nerve roots. J Physiol 56, 382-396 381. Lecci A, De Giorgio R, Bartho L, Sternini C, Tramontana M, Corinaldesi R, Giuliani S 8 Maggi CA. (1999). Tachykinin NK(1)receptor-mediated 252 inhibitory responses in the guinea-pig small intestine. Neuropeptides 33, 91-97. Lecci A, Santicioli P 8 Maggi CA. (2002). Pharmacology of transmission to gastrointestinal muscle. Curr Opin Pharmacol 2, 630-641. Lee SC, Choi S, Lee T, Kim HL, Chin H 8 Shin HS. (2002). Molecular basis of R- type calcium channels in central amygdala neurons of the mouse. Proc Natl Acad Sci U S A 99, 3276-3281. Lena C 8 Changeux JP. (1997). Role of Ca2+ ions in nicotinic facilitation of GABA release in mouse thalamus. J Neurosci 17, 576-585. LePard KJ 8 Galligan JJ. (1999). Analysis of fast synaptic pathways in myenteric plexus of guinea pig ileum. Am J Physiol 276, G529—538. LePard KJ, Messori E 8 Galligan JJ. (1997). Purinergic fast excitatory postsynaptic potentials in myenteric neurons of guinea pig: distribution and pharmacology. Gastroenterology 1 1 3, 1522-1 534. Li CG 8 Rand MJ. (1990). Nitric oxide and vasoactive intestinal polypeptide mediate non-adrenergic, non-cholinergic inhibitory transmission to smooth muscle of the rat gastric fundus. Eur J Pharmacol 191, 303-309. Lincoln TM, Dey N 8 Sellak H. (2001). Invited review: cGMP-dependent protein kinase signaling mechanisms in smooth muscle: from the regulation of tone to gene expression. J Appl Physiol 91, 1421-1430. Lipecka J, Bali M, Thomas A, Fanen P, Edelman A 8 F ritsch J. (2002). Distribution of ClC-2 chloride channel in rat and human epithelial tissues. Am J Physiol Cell Physiol 282, C805-816. Lomax AE 8 F urness JB. (2000). Neurochemical classification of enteric neurons in the guinea-pig distal colon. Cell Tissue Res 302, 59-72. Lomax AE, Sharkey KA, Bertrand PP, Low AM, Bornstein JC 8 Furness JB. (1999). Correlation of morphology, electrophysiology and chemistry of neurons in the myenteric plexus of the guinea-pig distal colon. J Auton Nerv Syst 76, 45-61. Longstreth GF, Thompson WG, Chey WD, Houghton LA, Mearin F 8 Spiller RC. (2006). Functional bowel disorders. Gastroenterology 130, 1480-1491. Lorenzon NM 8 Beam KG. (2000). Calcium channelopathies. Kidney Int 57, 794- 802. 253 Lorenzon NM 8 Foehring RC. (1995). Characterization of pharmacologically identified voltage-gated calcium channel currents in acutely isolated rat neocortical neurons. I. Adult neurons. J Neurophysiol 73, 1430-1442. MacDermott AB, Role LW 8 Siegelbaum SA. (1999). Presynaptic ionotropic receptors and the control of transmitter release. Annu Rev Neurosci 22, 443-485. Mackenzie l 8 Burnstock G. (1980). Evidence against vasoactive intestinal polypeptide being the non-adrenergic, non-cholinergic inhibitory transmitter released from nerves supplying the smooth muscle of the guinea-pig taenia coli. EurJ Pharmacol 67, 255-264. Macrae IM, Furness JB 8 Costa M. (1986). Distribution of subgroups of noradrenaline neurons in the coeliac ganglion of the guinea-pig. Cell Tissue Res 244, 173-180. Maggi CA. (2000). Principles of tachykininergic co-transmission in the peripheral and enteric nervous system. Regul Pept 93, 53-64. Majlof L 8 Forsgren PO. (1993). Confocal microscopy: important considerations for accurate imaging. Methods Cell Biol 38, 79-95. Makhlouf GM 8 Grider JR. (1993). Nonadrenergic Noncholinergic Inhibitory Transmitters of the Gut. News in Physiological Sciences 8, 195-199. Marino F, Marcoli M, De Ponti F, Lecchini S, Castelletti CM 8 Frigo GM. (1993). Inhibition of endogenous acetylcholine release by blockade of voltage- dependent calcium channels in enteric neurons of the guinea-pig colon. J Phann Pharmacol 45, 449-452. Mashimo H, He XD, Huang PL, Fishman MC 8 Goyal RK. (1996). Neuronal constitutive nitric oxide synthase is involved in murine enteric inhibitory neurotransmission. J Clin Invest 98, 8-13. Mashimo H, Kjellin A 8 Goyal RK. (2000). Gastric stasis in neuronal nitric oxide synthase-deficient knockout mice. Gastroenterology 119, 766-773. Matsuda Y, Saegusa H, Zong S, Noda T 8 Tanabe T. (2001). Mice lacking Ca(v)2.3 (alpha1 E) calcium channel exhibit hyperglycemia. Biochem Biophys Res Commun 289, 791-795. Matsui M, Motomura D, Karasawa H, Fujikawa T, Jiang J, Komiya Y, Takahashi S 8 Taketo MM. (2000). Multiple functional defects in peripheral autonomic organs in mice lacking muscarinic acetylcholine receptor gene for the M3 subtype. Proc Natl Acad Sci U S A 97, 9579-9584. 254 Matsushlta K, Wakamori M, Rhyu IJ, Arii T, Oda S, Mori Y 8 Imoto K. (2002). Bidirectional alterations in cerebellar synaptic transmission of tottering and rolling Ca2+ channel mutant mice. J Neurosci 22, 4388-4398. Matthews G. (1996). Neurotransmitter release. Annu Rev Neurosci19, 219-233. Mawe GM, Coates MD 8 Moses PL. (2006). Review article: intestinal serotonin signalling in irritable bowel syndrome. Aliment Pharmacol Ther 23, 1067- 1076. Mawe GM, Schemann M, Wood JD 8 Gershon MD. (1989). Immunocytochemical analysis of potential neurotransmitters present in the myenteric plexus and muscular layers of the corpus of the guinea pig stomach. Anat Rec 224, 431-442. Maximov A, Sudhof TC 8 Bezprozvanny I. (1999). Association of neuronal calcium channels with modular adaptor proteins. J Biol Chem 274, 24453- 24456. May GR, Gill MJ, Church DL 8 Sutherland LR. (1993). Gastrointestinal symptoms in ambulatory HIV-infected patients. Dig Dis Sci 38, 1388-1394. McCleskey EW, Fox AP, Feldman DH, Cruz LJ, Olivera BM, Tsien RW 8 Yoshikami D. (1987). Omega-conotoxin: direct and persistent blockade of specific types of calcium channels in neurons but not muscle. Proc Natl Acad Sci U S A 84, 4327-4331. McGehee DS, Heath MJ, Gelber S, Devay P 8 Role LW. (1995). Nicotine enhancement of fast excitatory synaptic transmission in CNS by presynaptic receptors. Science 269, 1692-1696. McMahon HT, Bolshakov W, Janz R, Hammer RE, Siegelbaum SA 8 Sudhof TC. (1996). Synaptophysin, a major synaptic vesicle protein, is not essential for neurotransmitter release. Proc Natl Acad Sci U S A 93, 4760- 4764. Meedeniya AC, Brookes SJ, Hennig GW 8 Costa M. (1998). The projections of 5-hydroxytryptamine-accumulating neurones in the myenteric plexus of the small intestine of the guinea-pig. Cell Tissue Res 291, 375-384. Mertz H, Naliboff B, Munakata J, Niazi N 8 Mayer EA. (1995). Altered rectal perception is a biological marker of patients with irritable bowel syndrome. Gastroenterology 109, 40-52. 255 Messenger JP 8 F urness JB. (1990). Projections of chemically-specified neurons in the guinea-pig colon. Arch Histol Cytol 53, 467-495. Meyer T 8 Brinck U. (1999). Differential distribution of serotonin and tryptophan hydroxylase in the human gastrointestinal tract. Digestion 60, 63-68. Michel K, Reiche D 8 Schemann M. (2000). Projections and neurochemical coding of motor neurones to the circular and longitudinal muscle of the guinea pig gastric corpus. Pflugers Arch 440, 393-408. Miljanich GP. (2004). Ziconotide: neuronal calcium channel blocker for treating severe chronic pain. Curr Med Chem 1 1 , 3029-3040. Miller RJ, Ewald DA, Fox AP, Hirning LD, McCleskey EW, Perney TM, Sturek M, Thayer SA, Tsien RW 8 Walker MW. (1988). The effect of calcium channel antagonists on peripheral neurones. Ann N YAcad Sci 522, 269- 277. Mintz IM, Adams ME 8 Bean BP. (1992). P-type calcium channels in rat central and peripheral neurons. Neuron 9, 85-95. Mochida S, Sheng ZH, Baker C, Kobayashi H 8 Catterall WA. (1996). Inhibition of neurotransmission by peptides containing the synaptic protein interaction site of N-type Ca2+ channels. Neuron 17, 781-788. Monro RL, Bornstein JC 8 Bertrand PP. (2005). Slow excitatory post-synaptic potentials in myenteric AH neurons of the guinea-pig ileum are reduced by the 5-hydroxytryptamine7 receptor antagonist SB 269970. Neuroscience 134, 975-986. Morita K 8 North RA. (1985). Significance of slow synaptic potentials for transmission of excitation in guinea-pig myenteric plexus. Neuroscience 14, 661 -672. Morita K, North RA 8 Tokimasa T. (1982). The calcium-activated potassium conductance in guinea-pig myenteric neurones. J Physiol 329, 341-354. Murakami M, Ohba T, Wu TW, Fujisawa S, Suzuki T, Takahashi Y, Takahashi E, Watanabe H, Miyoshi l, Ono K, Sasano H, Ito H 8 lijima T. (2007). Modified sympathetic regulation in N-type calcium channel null-mouse. Biochem Biophys Res Commun 354, 1016-1020. Murthy KS, Grider JR, Jin JG 8 Makhlouf GM. (1995). Interplay of VIP and nitric oxide in the regulation of neuromuscular activity in the gut. Arch Int Pharmacodyn Ther 329, 27-38. 256 Mutafova-Yambolieva VN, Hwang SJ, Hao X, Chen H, Zhu MX, Wood JD, Ward SM 8 Sanders KM. (2007). Beta-nicotinamide adenine dinucleotide is an inhibitory neurotransmitter in visceral smooth muscle. Proc Natl Acad Sci U S A 104, 16359-16364. Naidoo V, Bian X 8 Galligan JJ. (2006). Distribution of R-type calcium channels in the enteric nervous system of the guinea-pig. Gastroenterology 130 (Suppl. 2), A382. Negoescu A, Labat-Moleur F, Lorimier P, Lamarcq L, Guillermet C, Chambaz E 8 Brambilla E. (1994). F(ab) secondary antibodies: a general method for double immunolabeling with primary antisera from the same species. Efficiency control by chemiluminescence. J Histochem Cytochem 42, 433- 437. Newcomb R, Szoke B, Palma A, Wang G, Chen X, Hopkins W, Cong R, Miller J, Urge L, Tarczy-Hornoch K, Loo JA, Dooley DJ, Nadasdi L, Tsien RW, Lemos J 8 Miljanich G. (1998). Selective peptide antagonist of the class E calcium channel from the venom of the tarantula Hysterocrates gigas. Biochemistry 37, 15353-15362. Niidome T, Kim MS, Friedrich T 8 Mori Y. (1992). Molecular cloning and characterization of a novel calcium channel from rabbit brain. FEBS Letters 308, 7-13. Nishi S 8 North RA. (1973a). Intracellular recording from the myenteric plexus of the guinea-pig ileum. Joumal of Physiology 231, 471 -491. Nishi S 8 North RA. (1973b). Presynaptic action of noradrenaline in the myenteric plexus. J Physiol 231, 29P-30P. North RA. (1973). The calcium-dependent slow after-hyperpolarization in myenteric neurones with tetrodotoxin-resistant action potentials. British Journal of Pharmacology 49, 709-71 1 . North RA, Slack BE 8 Surprenant A. (1985). Muscarinic M1 and M2 receptors mediate depolarization and presynaptic inhibition in guinea-pig enteric nervous system. J Physiol 368, 435-452. North RA 8 Tokimasa T. (1987). Persistent calcium-sensitive potassium current and the resting properties of guinea-pig myenteric neurones. J Physiol 386, 333-353. Nurgali K, Nguyen TV, Matsuyama H, Thacker M, Robbins HL 8 Furness JB. (2007). Phenotypic changes of morphologically identified guinea-pig 257 myenteric neurons following intestinal inflammation. J Physiol 583, 593- 609. Ohashi K, Yamazaki T, Kitamura S, Ohta S, Izumi S 8 Kominami S. (2007). Allosteric inhibition of rat neuronal nitric-oxide synthase caused by interference with the binding of calmodulin to the enzyme. Biochim Biophys Acta 1770, 231-240. Olivera BM, Imperial JS, Cruz LJ, Bindokas VP, Venema VJ 8 Adams ME. (1991). Calcium channel-targeted polypeptide toxins. Ann N YAcad Sci 635, 114-122. Osthaus LE 8 Galligan JJ. (1992). Antagonists of nitric oxide synthesis inhibit nerve-mediated relaxations of longitudinal muscle in guinea pig ileum. J Pharmacol Exp Ther 260, 140-145. Pagani R, Song M, McEnery M, Qin N, Tsien RW, Toro L, Stefani E 8 Uchitel OD. (2004). Differential expression of alpha 1 and beta subunits of voltage dependent Ca2+ channel at the neuromuscular junction of normal and FIG Ca2+ channel knockout mouse. Neuroscience 123, 75-85. Pan H 8 Gershon MD. (2000). Activation of intrinsic afferent pathways in submucosal ganglia of the guinea pig small intestine. J Neurosci 20, 3295- 3309. Pardo NE, Hajela RK 8 Atchison WD. (2006). Acetylcholine release at neuromuscular junctions of adult tottering mice is controlled by N-(cav2.2) and R-type (cav2.3) but not L-type (cav1.2) Ca2+ channels. J Pharmacol Exp Ther 319, 1009-1020. Parkman HP, Hasler WL 8 Fisher RS. (2004). American Gastroenterological Association technical review on the diagnosis and treatment of gastroparesis. Gastroenterology 127, 1592-1622. Paton WD. (1955). The response of the guineapig ileum to electrical stimulation by coaxial electrodes. J Physiol 127, 40-41 P. Paton WD 8 Vizi ES. (1969). The inhibitory action of noradrenaline and adrenaline on acetylcholine output by guinea-pig ileum longitudinal muscle strip. Br J Pharmacol 35, 10-28. Paton WD 8 Zar MA. (1968). The origin of acetylcholine released from guinea- pig intestine and longitudinal muscle strips. J Physio/194, 13-33. Perez-Reyes E. (2003). Molecular physiology of Iow-voltage-activated t-type calcium channels. Physiol Rev 83, 1 17-161. 258 Pietrobon D. (2002). Calcium channels and channelopathies of the central nervous system. Mol Neurobiol 25, 31-50. Pietrobon D. (2005). Function and dysfunction of synaptic calcium channels: insights from mouse models. Curr Opin Neurobiol15, 257-265. Pompolo S 8 F urness JB. (1988). Ultrastructure and synaptic relationships of calbindin-reactive, Dogiel type II neurons, in myenteric ganglia of guinea- pig small intestine. J Neurocytol17, 771-782. Pompolo S 8 Furness JB. (1993). Origins of synaptic inputs to calretinin immunoreactive neurons in the guinea-pig small intestine. J Neurocytol 22, 531-546. Pompolo S 8 Furness JB. (1995). Sources of inputs to longitudinal muscle motor neurons and ascending interneurons in the guinea-pig small intestine. Cell Tissue Res 280, 549-560. Portbury AL, Furness JB, Young HM, Southwell BR 8 Vigna SR. (1996). Localisation of NK1 receptor immunoreactivity to neurons and interstitial cells of the guinea-pig gastrointestinal tract. J Comp Neurol 367, 342-351. Qian J 8 Noebels JL. (2000). Presynaptic Ca(2+) influx at a mouse central synapse with Ca(2+) channel subunit mutations. J Neurosci 20, 163-170. Qin N, Olcese R, Stefani E 8 Birnbaumer L. (1998). Modulation of human neuronal alpha 1E-type calcium channel by alpha 2 delta-subunit. Am J Physiol 274, C1 324-1331 . Quinson N, Robbins HL, Clark MJ 8 Furness JB. (2001). Calbindin immunoreactivity of enteric neurons in the guinea-pig ileum. Cell Tissue Res 305, 3-9. Randall A 8 Tsien RW. (1995). Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J Neurosci15, 2995-3012. Rao YM, Chaudhury A 8 Goyal RK. (2008). Active and inactive pools of nNOS in the nerve terminals in mouse gut: implications for nitrergic neurotransmission. Am J Physiol Gastrointest Liver Physiol 294, G627- 634. Reiche D, Michel K, Pfannkuche H 8 Schemann M. (2000). Projections and neurochemistry of interneurones in the myenteric plexus of the guinea-pig gastric corpus. Neurosci Lett 295, 109-112. 259 Reiche D, Pfannkuche H, Michel K, Hoppe S 8 Schemann M. (1999). lmmunohistochemical evidence for the presence of calbindin containing neurones in the myenteric plexus of the guinea-pig stomach. Neurosci Lett 270, 71-74. Reiche D 8 Schemann M. (1999). Mucosa of the guinea pig gastric corpus is innervated by myenteric neurones with specific neurochemical coding and projection preferences. J Comp Neurol 410, 489-502. Reis HJ, Biscaro FV, Gomez MV 8 Romano-Silva MA. (2002). Depolarization- evoked GABA release from myenteric plexus is partially coupled to L-, N-, and PIQ-type calcium channels. Cell Mol Neurobiol 22, 805-81 1. Reis HJ, Massensini AR, Prado MA, Gomez RS, Gomez MV 8 Romano-Silva MA. (2000). Calcium channels coupled to depolarization-evoked glutamate release in the myenteric plexus of guinea-pig ileum. Neuroscience 101, 237-242. Rekik M, Delvaux M, Tack I, Frexinos J 8 Bueno L. (1996). VIP-induced relaxation of guinea-pig intestinal smooth muscle cells: sequential involvement of cyclic AMP and nitric oxide. BrJ Pharmacol 118, 477-484. Ren J, Bian X, DeVries M, Schnegelsberg B, Cockayne DA, Ford AP 8 Galligan JJ. (2003). P2X2 subunits contribute to fast synaptic excitation in myenteric neurons of the mouse small intestine. J Physiol 552, 809-821. Richerson GB. (2003). The Autonomic Nervous System. In Medical Physiology, pp. 378-398. Saunders Publishing, Philadelphia. Ritchie J. (1973). Pain from distension of the pelvic colon by inflating a balloon in the irritable colon syndrome. Gut 14, 125-132. Robertson BE, Schubert R, Hescheler J 8 Nelson MT. (1993). cGMP-dependent protein kinase activates Ca-activated K channels in cerebral artery smooth muscle cells. Am J Physiol 265, C299-303. Rugiero F, Gola M, Kunze WA, Reynaud JC, Furness JB 8 Clerc N. (2002). Analysis of whole-cell currents by patch clamp of guinea-pig myenteric neurones in intact ganglia. J Physiol 538, 447-463. Rugiero F, Mistry M, Sage D, Black JA, Waxman SG, Crest M, Clerc N, Delmas P 8 Gola M. (2003). Selective expression of a persistent tetrodotoxin- resistant Na+ current and NaV1.9 subunit in myenteric sensory neurons. J Neurosci 23, 2715-2725. 260 Saban R, Nguyen N, Saban MR, Gerard NP 8 Pasricha PJ. (1999). Nerve- mediated motility of ileal segments isolated from NK(1) receptor knockout mice. Am J Physiol 277, G1173-1179. Sacaan Al, Menzaghi F, Dunlop JL, Correa LD, Whelan KT 8 Lloyd GK. (1996). Epibatidine: a nicotinic acetylcholine receptor agonist releases monOaminergic neurotransmitters: in vitro and in vivo evidence in rats. J Pharmacol Exp Ther 276, 509-515. Saegusa H, Kurihara T, Zong S, Minowa O, Kazuno A, Han W, Matsuda Y, Yamanaka H, Osanai M, Noda T 8 Tanabe T. (2000). Altered pain responses in mice lacking alpha 1E subunit of the voltage-dependent Ca2+ channel. Proc Natl Acad Sci U S A 97, 6132-6137. Sanders KM. (2008). Regulation of smooth muscle excitation and contraction. Neurogastroenterol Motil 20 Suppl 1, 39-53. Sanders KM 8 Koh SD. (2006). Two-pore-domain potassium channels in smooth muscles: new components of myogenic regulation. J Physiol 570, 37-43. Sanders KM 8 Ward SM. (1992). Nitric oxide as a mediator of nonadrenergic noncholinergic neurotransmission. Am J Physiol 262, G379-392. Sausbier M, Schubert R, Voigt V, Hirneiss C, Pfeifer A, Korth M, Kleppisch T, Ruth P 8 Hofmann F. (2000). Mechanisms of NO/cGMP-dependent vasorelaxation. Circ Res 87, 825-830. Schemann M, Reiche D 8 Michel K. (2001). Enteric pathways in the stomach. Anat Rec 262, 47-57. Schiess MC, Asprodini EK, Rainnie DG 8 Shinnick-Gallagher P. (1993). The central nucleus of the rat amygdala: in vitro intracellular recordings. Brain Res 604, 283-297. Schiess MC, Callahan PM 8 Zheng H. (1999). Characterization of the electrophysiological and morphological properties of rat central amygdala neurons in vitro. J Neurosci Res 58, 663-673. Schneider DA, Perrone M 8 Galligan JJ. (2000). Nicotinic acetylcholine receptors at sites of neurotransmitter release to the guinea pig intestinal circular muscle. J Pharmacol Exp Ther 294, 363-369. Schneider SP. (2008). Local circuit connections between hamster laminae Ill and IV dorsal horn neurons. J Neurophysiol 99, 1306-1318. 261 Schneider SP 8 Lopez M. (2002). Immunocytochemical localization of glutamic acid decarboxylase in physiologically identified interneurons of hamster spinal laminae Ill-V. Neuroscience 115, 627-636. Schneider T, Wei X, Olcese R, Costantin JL, Neely A, Palade P, Perez-Reyes E, Qin N, Zhou J, Crawford GD 8 et al. (1994). Molecular analysis and functional expression of the human type E neuronal Ca2+ channel alpha 1 subunit. Receptors Channels 2, 255-270. Seerden TC, Lammers WJ, De Winter BY, De Man JG 8 Pelckmans PA. (2005). Spatiotemporal electrical and motility mapping of distension-induced propagating oscillations in the murine small intestine. Am J Physiol Gastrointest Liver Physiol 289, G1043-1051 . Shao X, Davletov BA, Sutton RB, Sudhof TC 8 Rizo J. (1996). Bipartite Ca2+- binding motif in C2 domains of synaptotagmin and protein kinase C. Science 273, 248-251. Sharkey KA 8 Kroese AB. (2001). Consequences of intestinal inflammation on the enteric nervous system: neuronal activation induced by inflammatory mediators. Anat Rec 262, 79-90. Sheng ZH, Rettig J, Cook T 8 Catterall WA. (1996). Calcium-dependent interaction of N-type calcium channels with the synaptic core complex. Nature 379, 451-454. Sheng ZH, Rettig J, Takahashi M 8 Catterall WA. (1994). Identification of a syntaxin-binding site on N-type calcium channels. Neuron 13, 1303-1313. Shepherd GM. (1994). Neurobiology. Oxford University Press, New York. Siegelbaum SA, Schwartz JH 8 Kandel ER. (2000). Modulation of synaptic transmission: second messengers. In Principles of Neural Science, ed. Kandel ER, Schwartz JH 8 Jessell TM, pp. 229-252. McGraw-Hill, New York. Silver MA 8 Stryker MP. (2000). A method for measuring colocalization of presynaptic markers with anatomically labeled axons using double label immunofluorescence and confocal microscopy. J Neurosci Methods 94, 205-215. Smith AB 8 Cunnane TC. (1996). Omega-conotoxin GVIA-resistant neurotransmitter release in postganglionic sympathetic nerve terminals. Neuroscience 70, 817-824. 262 Smith AB 8 Cunnane TC. (1997). Multiple calcium channels control neurotransmitter release from rat postganglionic sympathetic nerve terminals. J Physiol 499 (Pt 2), 341-349. Smith TK, Burke EP 8 Shuttleworth CW. (1999). Topographical and electrophysiological characteristics of highly excitable S neurones in the myenteric plexus of the guinea-pig ileum. J Physiol 517 (Pt 3), 817-830. Smith TK, Kang SH 8 Vanden Berghe P. (2003). Calcium channels in enteric neurons. Curr Opin Pharmacol 3, 588-593. Smits GJ 8 Lefebvre RA. (1996). ATP and nitric oxide: inhibitory NANC neurotransmitters in the longitudinal muscle-myenteric plexus preparation of the rat ileum. Br J Pharmacol 118, 695-703. Smyth LM, Yamboliev IA 8 Mutafova-Yambolieva VN. (2009). N-type and PIQ- type calcium channels regulate differentially the release of noradrenaline, ATP and beta-NAD in blood vessels. Neurophannacology 56, 368-378. Sochivko D, Pereverzev A, Smyth N, Gissel C, Schneider T 8 Beck H. (2002). The Ca(V)2.3 Ca(2+) channel subunit contributes to R-type Ca(2+) currents in murine hippocampal and neocortical neurones. J Physiol 542, 699-710. Song ZM, Brookes SJ 8 Costa M. (1991). Identification of myenteric neurons which project to the mucosa of the guinea-pig small intestine. Neurosci Lett129, 294-298. Soong TW, Stea A, Hodson CD, Dubel SJ, Vincent SR 8 Snutch TP. (1993). Structure and functional expression of a member of the low voltage- activated calcium channel family. Science 260. Southwell BR 8 Furness JB. (2001). lmmunohistochemical demonstration of the NK(1) tachykinin receptor on muscle and epithelia in guinea pig intestine. Gastroenterology 120, 1140-1 151. Spencer NJ, Hennig GW 8 Smith TK. (2002). A rhythmic motor pattern activated by circumferential stretch in guinea-pig distal colon. J Physiol 545, 629- 648. Spencer NJ, Walsh M 8 Smith TK. (2000). Purinergic and cholinergic neuro- neuronal transmission underlying reflexes activated by mucosal stimulation in the isolated guinea-pig ileum. J Physiol 522 Pt 2, 321-331. 263 Stark ME, Bauer AJ, Sarr MG 8 Szurszewski JH. (1993). Nitric oxide mediates inhibitory nerve input in human and canine jejunum. Gastroenterology 104, 398-409. Stebbing M, Johnson P, Vremec M 8 Bornstein J. (2001). Role of alpha(2)- adrenoceptors in the sympathetic inhibition of motility reflexes of guinea- pig ileum. J Physiol 534, 465-478. Stephens GJ, Page KM, Burley JR, Berrow NS 8 Dolphin AC. (1997). Functional expression of rat brain cloned alpha1 E calcium channels in COS-7 cells. Pflugers Arch 433, 523-532. Sternini C, Su D, Gamp PD 8 Bunnett NW. (1995). Cellular sites of expression of the neurokinin-1 receptor in the rat gastrointestinal tract. J Comp Neuml 358, 531-540. Stewart WF, Liberrnan JN, Sandler RS, Woods MS, Stemhagen A, Chee E, Lipton RB 8 Farup CE. (1999). Epidemiology of constipation (EPOC) study in the United States: relation of clinical subtypes to sociodemographic features. Am J Gastroenterol 94, 3530-3540. Stricker NL, Christopherson KS, Yi BA, Schatz PJ, Raab RW, Dawes G, Bassett DE, Jr., Bredt DS 8 Li M. (1997). PDZ domain of neuronal nitric oxide synthase recognizes novel C-terrninal peptide sequences. Nat Biotechnol 15, 336-342. Sudhof TC 8 Rizo J. (1996). Synaptotagmins: C2-domain proteins that regulate membrane traffic. Neuron 17, 379—388. Surprenant A 8 North RA. (1988). Mechanism of synaptic inhibition by noradrenaline acting at alpha 2-adrenoceptors. Proc R Soc Lond B Biol Sci 234, 85-114. Szurszewski JH. (1969). A migrating electric complex of canine small intestine. Am J Physiol 217, 1757-1763. Takahashi T 8 Momiyama A. (1993). Different types of calcium channels mediate central synaptic transmission. Nature 366, 156-158. Takahashi T, Tsunoda Y, Lu Y, Wiley J 8 Owyang C. (1992). Nicotinic receptor- evoked release of acetylcholine and somatostatin in the myenteric plexus is coupled to calcium influx via N-type calcium channels. J Pharmacol Exp Ther 263, 1-5. Talavera K 8 Nilius B. (2006). Biophysics and structure-function relationship of T- type Ca2+ channels. Cell Calcium 40, 97-114. 264 Thompson WG, Longstreth GF, Drossman DA, Heaton KW, Irvine EJ 8 Muller- Lissner SA. (1999). Functional bowel disorders and functional abdominal pain. Gut 45 Suppl 2, ll43-47. Tonini M. (2005). 5-Hydroxytryptamine effects in the gut: the 3, 4, and 7 receptors. Neurogastroenterol Motil 17, 637-642. Torocsik A, Oberfrank F, Sershen H, Lajtha A, Nemesy K 8 Vizi ES. (1991). Characterization of somatodendritic neuronal nicotinic receptors located on the myenteric plexus. EurJ Pharmacol 202, 297-302. Tottene A, Volsen S 8 Pietrobon D. (2000). alpha(IE) subunits form the pore of three cerebellar R-type calcium channels with different pharmacological and permeation properties. J Neurosci 20, 171-178. Trendelenburg P. (2006). Physiological and pharmacological investigations of small intestinal peristalsis. Translation of the article "Physiologische und pharmakologische Versuche uber die Dunndarmperistaltik", Arch. Exp. Pathol. Pharmakol. 81, 55-129, 1917. Naunyn Schmiedebergs Arch Pharmacol 373, 101-133. Trouslard J, Mirsky R, Jessen KR, Burnstock G 8 Brown DA. (1993). Intracellular calcium changes associated with cholinergic nicotinic receptor activation in cultured myenteric plexus neurones. Brain Res 624, 103-108. Tsien RW, Ellinor PT 8 Home WA. (1991). Molecular diversity of voltage- dependent Ca2+ channels. Trends Pharmacol Sci 12, 349-354. Tsien RW, Lipscombe D, Madison DV, Bley KR 8 Fox AP. (1988). Multiple types of neuronal calcium channels and their selective modulation. Trends Neurosci 11, 431-438. Uddman R, Edvinsson L, Ekblad E, Hakanson R 8 Sundler F. (1986). Calcitonin gene-related peptide (CGRP): perivascular distribution and vasodilatory effects. RegulPept15, 1-23. Undi S, Benko R, Wolf M, lllenyi L, Horvath OP, Antal A, Csontos Z, Vereczkei A 8 Bartho L. (2006). Purinergic nerves mediate the non-nitrergic relaxation of the human ileum in response to electrical field stimulation. Brain Res Bull 71, 242-244. Urbano FJ, Piedras-Renteria ES, Jun K, Shin HS, Uchitel OD 8 Tsien RW. (2003). Altered properties of quantal neurotransmitter release at endplates of mice lacking PIQ-type Ca2+ channels. Proc Natl Acad Sci U S A 100, 3491-3496. 265 Vajna R, Schramm M, Pereverzev A, Arnhold S, Grabsch H, Klockner U, Perez- Reyes E, Hescheler J 8 Schneider T. (1998). New isoforrn of the neuronal Ca2+ channel alpha1 E subunit in islets of Langerhans and kidney-- distribution of voltage-gated CaZ+ channel alpha1 subunits in cell lines and tissues. Eur J Biochem 257, 274-285. Vance CL, Begg CM, Lee WL, Haase H, Copeland TD 8 McEnery MW. (1998). Differential expression and association of calcium channel alpha1 B and beta subunits during rat brain ontogeny. J Biol Chem 273, 14495-14502. Vanner S 8 Macnaughton WK. (2004). Submucosal secretomotor and vasodilator reflexes. NeurogastroenterolMoti/16 Suppl 1, 39-43. Vassort G 8 Alvarez J. (1994). Cardiac T-type calcium current: pharmacology and roles in cardiac tissues. J Cardiovasc Electrophysiol 5, 376-393. Vassort G, Talavera K 8 Alvarez JL. (2006). Role of T-type Ca2+ channels in the heart. Cell Calcium 40, 205-220. Vernino S, Amador M, Luetje CW, Patrick J 8 Dani JA. (1992). Calcium modulation and high calcium permeability of neuronal nicotinic acetylcholine receptors. Neuron 8, 127-134. Vogalis F, Furness JB 8 Kunze WA. (2001). Afterhyperpolarization current in myenteric neurons of the guinea pig duodenum. J Neurophysiol 85, 1941- 1951. Vogalis F, Harvey JR 8 F umess JB. (2004). Suppression of a slow post-spike afterhyperpolarization by calcineurin inhibitors. Eur J Neurosci 19, 2650- 2658. Vogalis F, Harvey JR, Lohman RJ 8 Furness JB. (2002). Action potential afterdepolarization mediated by a Ca2+-activated cation conductance in myenteric AH neurons. Neuroscience 115, 375-393. Volsen SG, Day NC, McCormack AL, Smith W, Craig PJ, Beattie RE, Smith D, lnce PG, Shaw PJ, Ellis SB, Mayne N, Burnett JP, Gillespie A 8 Harpold MM. (1997). The expression of voltage-dependent calcium channel beta subunits in human cerebellum. Neuroscience 80, 161-174. Wang G, Dayanithi G, Newcomb R 8 Lemos JR. (1999). An R-type Ca(2+) current in neurohypophysial terminals preferentially regulates oxytocin secretion. J Neurosci 19, 9235-9241. 266 Wang GD, Wang XY, Hu HZ, Liu S, Gao N, Fang X, Xia Y 8 Wood JD. (2007). Inhibitory neuromuscular transmission mediated by the P2Y1 purinergic receptor in guinea pig small intestine. Am J Physiol Gastrointest Liver Physiol 292, G1483-1489. Ward SM, Xue C, Shuttleworth CW, Bredt DS, Snyder SH 8 Sanders KM. (1992). NADPH diaphorase and nitric oxide synthase colocalization in enteric neurons of canine proximal colon. Am J Physiol 263, G277-284. Waterman SA 8 Costa M. (1994). The role of enteric inhibitory motoneurons in peristalsis in the isolated guinea-pig small intestine. J Physiol 477 (Pt 3), 459-468. Waterman SA, Tonini M 8 Costa M. (1994). The role of ascending excitatory and descending inhibitory pathways in peristalsis in the isolated guinea-pig small intestine. J Physiol 481 (Pt 1), 223-232. Weiergraber M, Henry M, Ho MS, Struck H, Hescheler J 8 Schneider T. (2008). Altered thalamocortical rhythmicity in Ca(v)2.3-deficient mice. Mol Cell Neurosci 39, 605-618. Weiergraber M, Pereverzev A, Vajna R, Henry M, Schramm M, Nastainczyk W, Grabsch H 8 Schneider T. (2000). Immunodetection of alpha1 E voltage- gated Ca(2+) channel in chromogranin-positive muscle cells of rat heart, and in distal tubules of human kidney. J Histochem Cytochem 48, 807- 819. Williams ME, Brust PF, Feldman DH, Patthi S, Simerson S, Maroufi A, McCue AF, Velicelebi G, Ellis SB 8 Harpold MM. (1992). Structure and functional expression of an omega-conotoxin-sensitive human N-type calcium channel. Science 257, 389-395. Williams ME, Marubio LM, Deal CR, Hans M, Brust PF, Philipson LH, Miller RJ, Johnson EC, Harpold MM 8 Ellis SB. (1994). Structure and functional characterization of neuronal alpha 1E calcium channel subtypes. J Biol Chem 269, 22347-22357. Wiser 0, Cohen R 8 Atlas D. (2002). Ionic dependence of Ca2+ channel modulation by syntaxin 1A. Proc Natl Acad Sci U S A 99, 3968-3973. Wonnacott S. (1997). Presynaptic nicotinic ACh receptors. Trends Neurosci 20, 92-98. Wood JD. (1996). Cellular neurophysiology of enteric neurons. In Physiology of the Gastrointestinal Tract, 4th edn, ed. Johnson RJ, Barret KE, Merchant 267 JL, Ghishan F K, Said HM 8 Wood JD, pp. 629-663. Elsevier Academic Press, Burlington, MA. Wood JD. (2006a). Cellular neurophysiology of enteric neurons. In Physiology of the Gastrointestinal Tract, 4th edn, ed. Johnson RJ, Barret KE, Merchant JL, Ghishan FK, Said HM 8 Wood JD, pp. 629-663. Elsevier Academic Press, Burlington, MA. Wood JD. (2006b). The enteric purinergic P2Y1 receptor. Curr Opin Pharmacol 6, 564-570. Wood JD 8 Kirchgessner A. (2004). Slow excitatory metabotropic signal transmission in the enteric nervous system. NeurogastroenterolMoti/16 Suppl 1, 71-80. Wu LG, Borst JG 8 Sakmann B. (1998). R-type Ca2+ currents evoke transmitter release at a rat central synapse. Proc Natl Acad Sci U S A 95, 4720-4725. Wu LG 8 Saggau P. (1995). Block of multiple presynaptic calcium channel types by omega-conotoxin-MVIIC at hippocampal CA3 to CA1 synapses. J Neurophysiol 73, 1 965-1 972. Wu LG, Westenbroek RE, Borst JG, Catterall WA 8 Sakmann B. (1999). Calcium channel types with distinct presynaptic localization couple differentially to transmitter release in single calyx-type synapses. J Neurosci 19, 726-736. Xue L, Farrugia-G, Miller SM, Ferris CD, Snyder SH 8 Szurszewski JH. (2000). Carbon monoxide and nitric oxide as coneurotransmitters in the enteric nervous system: evidence from genomic deletion of biosynthetic enzymes. Proc Natl Acad Sci U S A 97, 1851-1855. Yamakage M 8 Namiki A. (2002). Calcium channelsnbasic aspects of their structure, function and gene encoding; anesthetic action on the channels-- a review. Can J Anaesth 49, 151-164. Yasuda R, Sabatini BL 8 Svoboda K. (2003). Plasticity of calcium channels in dendritic spines. Nat Neurosci 6, 948-955. Yokoyama CT, Westenbroek RE, Hell JW, Soong TW, Snutch TP 8 Catterall WA. (1995). Biochemical properties and subcellular distribution of the neuronal class E calcium channel alpha 1 subunit. Joumal of Neuroscience 1 5, 641 9-6432. Zhang JF, Randall AD, Ellinor PT, Horne WA, Sather WA, Tanabe T, Schwarz TL 8 Tsien RW. (1993). Distinctive pharmacology and kinetics of cloned 268 neuronal Ca2+ channels and their possible counterparts in mammalian CNS neurons. Neuropharmacology 32, 1075-1088. Zhou X 8 Galligan JJ. (1999). Synaptic activation and properties of 5- hydroxytryptamine(3) receptors in myenteric neurons of guinea pig intestine. J Pharmacol Exp Ther 290, 803-810. Zhou X, Ren J, Brown E, Schneider D, CarabalIo-Lopez Y 8 Galligan JJ. (2002). Pharmacological properties of nicotinic acetylcholine receptors expressed by guinea pig small intestinal myenteric neurons. J Pharmacol Exp Ther 302, 889-897. 269