Enteric Nervous System Structure and Neurochemistry Related to Function and Neuropathology




Abstract


This chapter will discuss major recent advances that have been made in our understanding of how the different functional and neurochemical classes of enteric neurons participate in neurogenic motor activities in the lower gastrointestinal tract that underlie muscle contraction and propulsion of ingest content. Also, we will discuss the very recent advances that have used optogenetics successfully, for the first time this year to control the excitability of the enteric nervous system. Finally, we will reveal some of our own new data demonstrating the potency with which optogenetics can be used to successfully induce colonic contractions and control the timing of colonic propulsion and expulsion of fecal content from the large bowel.




Keywords

Peristalsis, Myenteric plexus, Colonic migrating motor complex, Optogenetics, Sensory neuron, Enteric nervous system, Hirschsprung’s disease, Chagas disease

 





Introduction


The gastrointestinal (GI) tract is the only internal organ with its own nervous system, known as the enteric nervous system (ENS), which is concealed entirely within the gut wall and can function fully independently of any neural inputs from the central nervous system (CNS). What is particularly unique about the ENS and the GI tract compared to all other organs is that it is capable of responding to sensory stimuli in vitro (without an involvement from any extrinsic afferents) and can generate complex neurogenic motor patterns, even when isolated from the animal. It is sometimes overlooked, but the ENS is itself an entire division of the autonomic nervous system and has developed all the neural circuits necessary to not only respond to sensory stimuli but also trigger polarized polysynaptic neural pathways that lead to muscle contraction oral to and relaxation anal to physiological stimuli. Activation of enteric neural circuits, following local stimulation in vitro, is sufficient to propel ingested contents along the isolated bowel.


The ENS has evolved not only in the GI tract of vertebrates but also in tiny invertebrates, like flies. For example, a prominent ENS has developed in the GI tract of the fly Drosophila melanogaster , and there is compelling evidence that the same neurotransmitters that are expressed in larger vertebrates have also evolved in the ENS of invertebrates. This is supported by the finding that choline acetyl transferase (ChAT) is expressed in enteric neurons from the midgut of Drosophila and peristalsis occurs in this region. In recent years, much has been learnt about the development of the ENS in grasshoppers and locusts and while there are important differences between the ENS of vertebrates and invertebrates, there are also some very clear similarities emerging in terms of neurogenesis, migration, and differentiation of enteric neurons.


The importance of the ENS in normal day-to-day propulsion of colonic content is perhaps best exemplified in genetic diseases like Hirschsprung’s disease, where the GI tract develops without any enteric ganglia in the terminal part of the colon and/or rectum. This region is referred to as the aganglionic region. But, interestingly, this aganglionic region still receives extrinsic spinal efferent and afferent nerve inputs. And, interestingly, within the aganglionic region interstitial cells of Cajal (ICC), which are intestinal pacemaker cells, still develop. However, by themselves, ICC and smooth muscle cells are unable to propel contents sufficiently, leading to improper transit. The outcome is that newborn mammals born with Hirschprung’s disease fail to develop the necessary propulsive neurogenic motor patterns, and typically megacolon develops. Without resection, mammals with large aganglionic segments of bowel typically die shortly after birth. While it is well acknowledged that inputs from the CNS are not required for the activation of enteric neural pathways and propulsion of intestinal content, there is extensive neural innervation and control of the ENS by central pathways arising from spinal cord.


The ENS composed of two discrete plexuses: one known as the myenteric plexus (originally called Auerbach’s plexus), which lies between the two major muscle layers (the circular and longitudinal muscle coats) of the GI tract. The major function of the myenteric plexus is to generate neurogenic contractions and relaxations of the smooth muscle layers that are vital for the coordinated timing and genesis of neurogenic motor patterns, necessary for propulsion of ingested contents. The other nerve plexus that comprises the ENS is known as the submucosal plexus (originally called Meissner’s plexus). This plexus is largely associated with secretomotor reflexes and absorption. In some parts of the GI tract, like the stomach, there is no submucosal plexus and motor reflexes and propulsion still occur. In these regions, it is thought that the myenteric plexus takes on the role of the submucosal plexus.


There are literally hundreds of thousands of nerve cell bodies within the myenteric and submucosal plexuses. Individual neurons within each plexus are localized to small ganglia that also contain variable numbers of glial cells. Each ganglion consists of variable numbers of different functional classes of enteric neurons that are distributed in not opographical order throughout each plexus. Submucosal ganglia are noticeably smaller than myenteric ganglia and contain significantly fewer numbers of neurons. Also, each submucosal ganglion is further spaced apart from one another when compared to ganglia in the myenteric plexus. Each ganglion, within the submucosal and myenteric plexus communicates to neighboring ganglia via synaptic connections, involving different neurotransmitters that generate fast and/or slow synaptic potentials. In the colon of mice, for example, there are ~ 40 rows of myenteric ganglia around the circumference of the colon and a similar number of rows of ganglia along every 6 mm of the length of colon (at resting length). This means that if one assumes each myenteric ganglion has on average 20 neurons, then for every 6 mm along the length of colon, there are ~ 32,000 myenteric neurons. If one assumes the density of myenteric ganglia and number of myenteric neurons is similar along the full length of colon (which is ~ 7 cm in length at rest), then this means there are ~ 370,000 myenteric neurons in the mouse colon.





Morphological Classification of Enteric Neurons


The first morphological identification of enteric neurons was made in the late 1800s by Dogiel and La Villa. At first, Dogiel devised a particular nomenclature, which consisted of Type I, Type II, and Type III neurons. Later, he modified this classification to include Types IV-VII neurons. Nowadays, enteric neurons are most commonly classified morphologically into one of two major classes, either Dogiel Type I or Dogiel Type II. Neurons with Dogiel Type I morphology consist of relatively small cell bodies, with dendrites that do not project far from the cell soma and usually, a single long axon. A number of different functional classes of enteric neurons have morphologies consistent with Dogiel Type I neurons, including ascending and descending interneurons and inhibitory and excitatory motor neurons. A class of filamentous Dogiel Type I neurons has also been identified, which are also uniaxonal, but display extensive filamentous dendrites (see Fig. 14.1 ). Direct intracellular recordings from a population of filamentous interneurons in the distal colon have revealed that these neurons exhibit mechanosensory properties in response to maintained stretch.




Fig. 14.1


Complex dendritic morphology of an ascending interneuron in the guinea-pig distal colon. (A) A single uniaxonal S-neuron with long branching dendrites extending from its soma. (B) An expanded region of the three neurons injected in panel (A). (C) Ongoing action potentials and fast EPSPs in the recorded neuron.


In contrast, Type II neurons have a very distinctive shape and projection profile. They have large smooth cell bodies, often oval in shape, that consist of many long branching axons that ramify extensively within myenteric ganglia. In the guinea pig small intestine, some Type II neurons have long aboral projections. The functional role of these outputs is not clear, but they may serve as a multifunctional neuron (e.g., interneuron), in addition to potentially fulfilling the role as an intrinsic sensory neuron.





First Recordings From the ENS


The first known documented recordings from enteric neurons were made by Yokoyama in the mid-1960s. Although the morphological features and neurochemical coding of the neurons were unknown, Dr. Yokoyama had successfully recorded the electrical activity from neurons in the ENS, using extracellular electrodes in the isolated rabbit intestine. During his study, he was able to detect extracellular compound discharges of nerve action potentials, where it was stated that action potentials had a duration of about 20 ms. A few years later, Dr. Wood reported a more detailed description of the different patterns of neuronal firing within the myenteric plexus of the isolated cat intestine, where action potentials had a duration of 2–5 ms. In the rabbit, Yokoyama reported discharges of action potentials with an interval between bursts of 5 and 25 s and rhythmic segmental- type movements correlated with the nerve activity and not with the action potentials in the longitudinal muscle. In the cat, the average interval between bursts was 6.9 s. The study of Dr. Wood was highly significant because it was the first to demonstrate that enteric neurons could respond to mechanical stimuli with bursts of action potentials ( Fig. 14.2 ). He showed that both phasic- and tonic-type discharges of action potentials could occur and when mechanical stimulation was imposed on the myenteric ganglion by withdrawing the suction electrode, a tonic discharge of action potentials was elicited.




Fig. 14.2


Tonic and phasic mechanoreceptors in the myenteric plexus of isolated cat intestine. (A) Phasic mechanoreceptor that generated spontaneous action potentials increased its firing rate during application of mechanical stimuli. The top two records are a continuous recording. (B) In contrast, the bottom two recordings show a tonic discharge of action potentials that was elicited after removal of the suction electrode.

(From Fig. 5 in Wood JD. Electrical activity from single neurons in Auerbach’s plexus. Am J Physiol 1970; 219 (1):159–69, with permission.)


The first detailed intracellular recordings from myenteric neurons were made coincidentally by two independent laboratories, one in the northern hemisphere and the other in the southern hemisphere. Essentially, the two laboratories revealed very similar findings. That is, they identified two distinct populations of myenteric neurons based on electrical responses to current injection and the synaptic inputs they received in response to electrical stimulation of neighboring fiber tracts. Both studies agreed that one major type of cell fired tonically to current injection and received prominent fast synaptic inputs. These were called S-type neurons by or Type I neurons by, whereas the second population that rarely responded with any synaptic inputs to electrical stimulation, consistently generated a prolonged after hyperpolarization that ensured following single-action potentials. These were called Type II neurons by Dr. North’s laboratory or after-hyperpolarizing (AH) neurons in the laboratory of Hirst et al. Also, Dr. Hirst noted in their experiments that it was not possible “… to evoke any synaptic activity in AH cells by transmural stimulation of the plexus, and no spontaneous synaptic activity has ever been recorded from these cells.” They suggested that “…that AH neurones, in common with primary afferent neurones, may not be innervated.” At the same time, Nishi and North also made the comment “…that the cells with the slow after-hyperpolarization do not have synaptic inputs, are compatible with the hypothesis that they are in the afferent side of an intramural reflex…” While there was no doubt that neural reflexes could be elicited by local stimulation of isolated segments of bowel, it had not been demonstrated that these reflexes were generated by sensory neurons, whose cell bodies resided within the gut wall. It was conceivable, at that time, that axon collaterals of extrinsic spinal afferent axons, whose nerve endings resided within the gut wall, but whose cell bodies resided in dorsal root ganglia (DRG) could initiate neural reflexes in isolated segments of bowel.


It was not until 1995 that evidence emerged for the existence of a population of enteric neurons that had sensory properties. This was demonstrated when all synaptic transmissions were blocked in the ENS (using a tissue bathing solution of reduced extracellular calcium to block synaptic transmission), and it was demonstrated that under these conditions, exogenous serotonin (5-HT) puffed onto the mucosa elicited proximal process potentials in the cell somas of myenteric AH neurons. Later, a similar type response was demonstrated in the proximal colon, when electrical stimulation was applied to the mucosa and proximal process potentials were evoked in mucosally projecting Dogiel Type II neurons. Proximal process potentials occur in neuronal cell bodies when action potentials, which have been generated in the neurites, fail to elicit an action potential in the cell body. Unlike synaptic potentials that vary considerably in amplitude in the cell soma, proximal process potentials are of the same amplitude and do not increase in amplitude with an imposed membrane hyperpolarization (which is the opposite of fast EPSPs). The conclusion to these experiments was that the nerve endings of Dogiel Type II neurons (which project into the mucosa) can be directly excited by chemosensory stimuli, without relying on synaptic inputs from any other classes of neurons. This was taken as the first evidence that a population of sensory neurons exists in a peripheral organ. Based on these findings, the laboratory of Dr. Furness coined the terminology “Intrinsic Primary Afferent Neuron (or IPAN),” to represent myenteric Dogiel Type II (AH) neurons, which responded directly to mechanical and/or chemosensory stimuli, without synaptic transmission.


The notion that a population of sensory neurons exists within the ENS has not been fully accepted by all investigators. Dr. Wood has suggested that the terminology of IPAN “…is inaccurate because AH neurons are not sensory afferent neurons.” He suggests that justification for naming these neurons IPANs, based on data showing a paracrine activation of AH neurons by 5-HT is insufficient and suggests that “…AH neurons are interneurons that are synaptically connected for feed-forward excitation of neighboring AH neurons in driver circuits.” Indeed, morphologically, it is clear that AH neurons have extensive ramifications within myenteric ganglia onto other enteric neurons and some have long aboral projections for many millimeters. Since the original work of Nishi and North and Hirst et al. has been revealed that some AH neurons do receive fast synaptic inputs ( Fig. 14.3 ), which are mediated by nicotinic receptors.




Fig. 14.3


Spontaneous fast EPSPs in a myenteric AH neuron of the guinea-pig distal colon. (A) Spontaneous and evoked action potentials and periods of after hyperpolarization. Also, note spontaneous fast EPSPs can occur in this neuron of irregular amplitudes. (B) Two spontaneous after-hyperpolarizations and one evoked action potential and after- hyperpolarization. During this AH, some fast EPSPs are evident (see expanded period in panel (B)). (C) Spontaneous action potentials and AH periods with fast EPSPs. (D) An expanded period of the evoked action potentials in panel (C).


To test the hypothesis that a population of intrinsic sensory neurons exists within the wall of the GI tract, which could be capable of eliciting intrinsic neural reflexes, without any involvement from extrinsic afferents, segments of guinea pig intestine were extrinsically denervated (in vivo) about 9–11 days before the bowel was resected from the animal. The success of the denervation was performed histochemically to show degradation of extrinsic axons in the gut wall. When these preparations of intestine were removed from the animal, it was revealed that ascending and descending reflexes could still be readily elicited in the circular muscle, even though all extrinsic nerves had been histologically depleted. This data revealed that in the absence of all extrinsic afferent endings in the gut wall, sensory stimuli applied in vitro could still evoke ascending excitatory and descending inhibitory pathways to the smooth muscle layers.


The mechanisms underlying mechanosensory transduction in the ENS of the colon have some significant differences compared to the small intestine, even within the same species. For example, in guinea-pig distal colon maintained circumferential stretch evokes an ongoing discharge of ascending excitatory and descending inhibitory nerve pathways, which fire simultaneously in the circular ( Fig. 14.4 ) and longitudinal muscle layers. This stretch-activated motor pattern did not require extrinsic capsaicin-sensitive afferents and was unaffected by muscle paralysis with an L-type Ca 2 + channel blocker. Clearly, in the distal colon, a population of sensory neurons must not require muscle tone or tension for the mechanotransduction of stretch-activated firing of ascending excitatory and descending inhibitory pathways in the ENS. In contrast to the distal colon, in the guinea-pig ileum, paralysis of muscle tone or tension has a major effect to reduce mechanosensitivity of intrinsic sensory neurons. The mechanisms underlying the discrepancy of mechanotransduction between the two regions are not understood.




Fig. 14.4


Simultaneous intracellular recordings from two circular muscle cells show simultaneous activation of ascending excitatory and descending inhibitory neuronal pathways. Note, the similarity in amplitudes between oral EJPs and anal IJPs despite the two electrodes was separated by 20 mm. The mechanotransduction process underlying this stretch-activated pathways is resistant to changes in muscle tension or tone, since these recordings were made in the presence of nifedipine, to block L-type Ca 2 + channels.


What is clear is that in the guinea-pig distal colon, the stretch-activated repetitive firing of ascending excitatory and descending inhibitory neural pathways were resistant to muscle paralysis, suggesting that a fundamentally different process was required for sensory neuronal transduction than was identified for the muscle tension-sensitive activation of myenteric AH neurons in the small intestine. Interestingly, in these paralyzed preparations of distal colon, direct recordings from myenteric AH neurons revealed this class of neuron was electrically silent at the same time as these stretch-activated pathways were active ( Fig. 14.5 ). However, in these stretched segments of guinea-pig distal colon, a population of myenteric neurons was recorded from the guinea-pig colon, which were activated by circumferential stretch, even in the presence of a low Ca 2 + solution (to prevent all synaptic transmission). These neurons were identified as interneurons with extensive dendritic morphology and with S-class electrophysiology, which received prominent fast synaptic inputs and fired tonically to current injection, but also generated proximal process potentials. This finding provided support for the idea that a population of interneurons may serve a dual function, not only to transmit synaptic traffic along the length of bowel but also to respond directly to mechanical stimuli. Neuronal imaging data from Dr. Schemann’s laboratory further support the idea that a population of intrinsic sensory neurons could be multifunctional neurons. Their work identified rapidly adapting mechanosensitive enteric neurons which were called (RAMEN). In the guinea-pig small intestine, these neurons comprised 25% of all neurons in the myenteric plexus. It was found that RAMEN were activated with phasic action potential discharge pattern to deformation of the ganglion, and this response was suggested to be reproducible and increased as the applied force increased. Interestingly, of all RAMEN tested 72% were cholinergic, 22% were nitrergic, and 44% were identified to express NeuN and calbindin, two reliable markers of IPANs. Mechanosensitivity was observed in 31% and 47% of retrogradely traced interneurons and motor neurons, respectively. Because it was found that a substantial population of myenteric neurons was interneurons or motor neurons, which still responded to von Frey hair compression, or intraganglionic volume injection, the team suggested that many neurons in the ENS are multifunctional mechanosensitive neurons (MEN). These investigators have now identified MENs not only in the guinea-pig ileum but also mouse ileum and colon and human intestine.




Fig. 14.5


Simultaneous intracellular recording from a myenteric AH neuron and circular muscle cell in isolated preparations of guinea-pig distal colon maintained under circumferential stretch. (A) The preparation in which the longitudinal muscle was removed and myenteric plexus faces uppermost position. An independent microelectrode was impaled into an AH neuron and a second electrode into a circular muscle cell. (B) A stretch-activated discharge of IJPs in the circular muscle at the same time as the AH neuron was silent. Only when current was injected into the neuron did it fire a brief burst of action potentials and a prolonged after-hyperpolarization. (C) The large multipolar Dogiel Type II neuron from which recordings in B were made. (D) An expanded period of the evoked action potentials in panel (B).





Motor Neurons in the ENS


Unlike skeletal muscle that has only an excitatory innervation, smooth muscle of the GI tract is innervated by both excitatory and inhibitory motor neurons. There have been some important advances in our understanding of enteric motor neurons. One of the important advances was the discovery that enteric motor neurons to the longitudinal and circular muscle are indeed innervated by separate and distinct populations of motor neurons. This was first suspected when differences in neurotransmission were recorded during simultaneous intracellular recordings from the two muscle layers our understanding of the motor neuronal populations that innervate the circular and longitudinal muscle layers. Then, with the development of DiI retrograde tracing from the two distinct muscle layers, it was confirmed that indeed different populations of motor neurons innervate the two muscle layers of the bowel, with different projections and neurochemical characteristics. Longitudinal muscle motor neurons ramify in no preferential orientation around the circumference of the bowel, whereas circular muscle motor neurons have clear circumferential projections, and a single motor neuron to the circular muscle can typically innervate about one-third of the circumference of the bowel. In human colon, the projections of circular muscle motor neurons are very similar to those in laboratory animals. That is, the excitatory circular muscle motor neurons project locally, or orally for short distances (typically < 2 mm) and contain ChAT, while inhibitory circular muscle motor neurons typically project short distances anally and contain the enzyme nitric oxide synthase (NOS) and often vasoactive intestinal peptide (VIP).


In response to maintained circumferential stretch of the bowel, studies have now shown that common interneurons synapse onto both distinct populations of inhibitory and excitatory motor neurons. This was demonstrated when stretch-evoked discharges of inhibitory junction potentials (IJPs) were found to be linearly related in amplitude, even though no electrotonic coupling existed between the two layers. This was supported by the finding that when IJPs occurred simultaneously in both the circular and longitudinal muscles, application of a nitric oxide synthesis inhibitor blocked the slow IJP in the longitudinal muscle, but the prominent fast apamin-sensitive IJP in the circular layer was not detected passively propagating in longitudinal cells.


A most interesting recent advance in our understanding of enteric motor neurons is that a population of NOS immunoreactive Dogiel Type I neurons responded directly to mechanical stimuli. This population likely includes inhibitory motor neurons, which can therefore be presumed to be potentially “multifunctional neurons.” Indeed, it had been suspected for some time that a population of enteric motor neurons may exhibit mechanosensitive properties, and hence may be capable of releasing neurotransmitter when all known excitatory synaptic inputs to these neurons had been blocked. This notion had emerged as an example, in the guinea-pig distal colon where acetylcholine appeared to be able to be released onto the smooth muscle in stretched colonic preparations, even in the presence of hexamethonium when all synaptic transmission in the ENS had been prevented. This was realized when atropine potently reduced circular muscle tone of the colon, even though these same reparations had nicotinic transmission blocked.





Interneurons


Traditionally, enteric interneurons were thought of as neurons that simply receive synaptic inputs from either other interneurons or sensory neurons, then relay synaptic outputs (excitatory or inhibitory) to either other interneurons, or motor neurons in ascending or descending pathways along the bowel. This notion has been revised in recent years. Nowadays, a picture has emerged that a population of interneurons are multifunctional and can likely serve roles as both a sensory neuron and interneuron. As mentioned, in the guinea-pig distal colon maintained circumferential stretch evokes sensory responses in a population of myenteric interneurons. These interneurons have extensive dendritic processes that ramify throughout the ganglion of origin ( Fig. 14.1 ). Based on immunohistochemical staining techniques, in the guinea-pig distal colon, there are four major neurochemical classes of descending interneuron and three major classes of ascending interneuron. In the guinea-pig ileum, there is reported to be only one major class of ascending interneuron, from which detailed electrophysiological characteristics have been obtained. No mechanosensory properties were identified in ascending interneurons of the guinea-pig ileum. However, they do receive prominent fast nicotinic synaptic inputs predominantly from other ascending pathways. In contrast to only one class of ascending interneuron in the guinea-pig ileum there are four major classes of descending interneuron in this region, based on neurochemical classification. In the human colon, there are two major types of ascending interneuron. One type which comprises 90% of all ascending interneurons contains ChAT, while the remainder contain neither NOS or ChAT. There are three major classes of descending interneuron in the human colon, the greatest population contains NOS, but interestingly lack ChAT. 5-Hydroxytryptamine-like immunoreactivity has been identified in a small population of enteric neurons of human foetuses, but no evidence has been revealed that 5-HT is a neurotransmitter in human ENS, since all synaptic inputs are blocked by hexamethonium.


An important advance in our understanding of the control of excitability of specific classes of enteric neurons was made when it was discovered recently that calcium/calmodulin- dependent protein kinase II was found to be colocalized in myenteric neurons of guinea-pig, human, rat, and mouse intestine. It was revealed that CaMKII expression was localized in myenteric neurons expressing calretinin and NOS and noncholinergic secretomotor neurons of the submucosalplexus. When CaMKII activity was blocked with the selective inhibitor KN-62, it caused a suppression of slow excitatory postsynaptic potentials and prevented tetrodotoxin-induced increases in contractility, suggesting that tonic neurogenic inhibition in the colon is dependent upon CaMKII-signaling pathways. Interestingly, CaMKII signaling acted to suppress purinergic slow synaptic potentials that have been shown to be mediated by the P2Y1 receptor on submucosal neurons. Fast synaptic transmission in the ENS involves P2X receptors, and P2X7 are highly expressed on neurons expressing calbindin, calretinin, neuropeptide Y, substance P, and NOS.





Submucosal Neurons


There have been some important advances in our understanding of submucosal neurons in recent years. These neurons are well known to play a major role in electrolyte and water secretion. In a recent study, it was demonstrated that the properties of submucosal neurons ad their control of epithelial ion transport changed along the length of the colon. Also, the size of submucosal ganglia decreased, at the same time as the proportion of cholinergic neurons actually increased along the length of colon. In mice, a pronounced difference between the submucosal plexus compared to other species is the overt lack of intrinsic sensory neurons. The functional ramifications of this are unclear. What is clear is that the colonic migrating motor complex (CMMC), which is a major colonic motor pattern in the isolated whole mouse, can be reliably generated without the submucosal plexus; and mucosa and only requires an intact myenteric plexus and smooth muscle layers.


There have been some important advances in our understanding of synaptic transmission in the submucosal plexus. A study showed that neurally released nitric oxide acts to increase the amplitude of inhibitory postsynaptic potentials (IPSPs) in submucosal neurons, but likely acts to depress the amplitude of slow excitatory postsynaptic potentials (slow EPSPs).


Recent studies have investigated the effects of cholera toxin (CT) on excitability of submucosal neurons. While it is well known that the hypersecretion induced by cholera toxin requires activation of secretomotor neurons, the classes of neurons affected by CT are incompletely understood. In isolated segments of jejunum from guinea pigs, a recent study showed perhaps surprisingly, that submucosal neurons were unaffected, but myenteric neurons showed hyperexcitability. This increase in excitability of myenteric AH neurons was found to be dependent upon NK3 receptors, but not ligand-gated 5-HT3 receptors.





Spinal Afferents


Spinal afferent neurons are of particular interest because they play the major role in the transduction and transmission of noxious and innocuous stimuli from visceral organs, like the GI tract. These afferents are also of interest in the ENS because there is evidence that extrinsic spinal afferents may modulate both the excitability of the ENS and peristalsis. While the location of the cell bodies of spinal afferents is well known to lie within DRG, the location of the nerve endings of spinal afferents that actually transduce sensory stimuli into action potentials has only recently been identified in the GI tract. This has arisen because of the development of a new anterograde tracing technique to selectively label only spinal afferent axons, and not all the other axons that run alongside spinal afferents as they enter the visceral organ. The spinal afferent endings that innervate the distal colon and rectum are extraordinarily complex. A single spinal afferent axon that enters the colorectum can give rise to well over 200 varicose axons, in at least three different anatomical layers, for example, the submucosa, circular muscle, and myenteric ganglia. A total of 14 different classes of spinal afferent ending were identified in the mouse distal colorectum, almost all of which were immunoreactive to CGRP. In contrast, in the stomach, there were only eight distinct types of spinal afferent endings, the majority of which innervated the myenteric ganglia and circular muscle. There is a very high degree of colocalization between the transient receptor potential vaniloid 1 (TRPV1) receptor/ion channel and CGRP in spinal afferent endings in the esophagus, stomach, and colon. However, there is no clear evidence for TRPV1 expression in enteric neuronal cell bodies. On the other hand, studies have shown that there is clear expression for the canonical TRPC family of ion channels in enteric neurons, specifically TRPC1 in myenteric cholinergic, calretinin, and nitrergic neurons and cholinergic and noncholinergic neurons of the submucosal plexus.





Intestinofugal Neurons


Intestinofugal neurons are enteric neurons, whose cell bodies lie in the ENS, but have axonal projections that exit the GI tract, via the mesenteric nerves and make synaptic connections with sympathetic prevertebral ganglia. The first realization of the existence of intestinofugal neurons was made when the isolated guinea-pig colon was distended while intracellular electrophysiological recordings were made from neurons in the inferior mesenteric ganglia (IMG) that remained in neural continuity with the colon. In this colon-IMG preparation, a discharge of fast EPSPs was evoked in IMG neurons in response to colonic distension. This afferent response to the IMG was transmitted via the lumbar colonic nerves (LCN), since lesions to the LCN abolished all distension-evoked responses. When nicotinic receptor antagonists were applied selectively to the colon (but not IMG), the discharge of fast EPSPs evoked in the IMG following colonic distension were substantially reduced (but not blocked). Similar results were obtained when the nicotinic antagonist hexamethonium was applied to the colon of guinea pigs and recordings made from the coeliac ganglia, or the superior mesenteric ganglia (SMG) of the mouse. This consistent finding between species showed that intestinofugal neurons were predominantly second-order neurons, whose activation was largely due to synaptic inputs arising from other mechanosensitive enteric neurons. The fact that hexamethonium applied selectively to the colon did not block all distension-evoked activation of intestinofugal neurons suggests one of two possibilities. First, that a population of mechanosensitive enteric neurons could synaptically activate intestinofugal neurons via nonnicotinic distension-sensitive pathways within the ENS. Or, alternatively, there could be a population of intestinofugal neurons that are primary afferent neurons with cell bodies in the ENS and which respond directly to mechanical stimuli. Bywater further tested this notion specifically. He investigated whether nonnicotinic pathways in the ENS could activate intestinofugal neurons. To do this, he replaced the normal Krebs solution to the colon only, with a solution containing low Ca 2 + to block all synaptic transmission in the ENS. He found that when all synaptic transmission was blocked in the ENS, colonic distension still evoked some fast EPSPs in the IMG. Although no definitive recordings were made directly from intestinofugal neurons, Bywater concluded logically that some intestinofugal neurons must project directly from the colon to the IMG which can be activated directly, that is, without any synaptic inputs following mechanical stimulation.


Recent studies have now demonstrated that intestinofugal neurons in the guinea-pig distal colon are indeed directly mechanosensitive and can respond to mechanical probing by Von Frey hairs. Hibberd et al. recently showed that during recordings from intestinofugal neurons, increases in muscle tension without any change in the muscle length did not alter sensory firing. This led to these workers to the conclusion that distension-evoked firing in intestinofugal neurons did not require muscle tone or tension, consistent with the same conclusion proposed earlier by Miller and Szurszewski in the mouse colon. Interestingly, in the guinea-pig colon, it was found that circumferential shortening of muscle length due to contraction that was maintained under a constant load reduced sensory firing. However, when intramural muscle tension was increased, without any accompanying changes in muscle length did not affect afferent firing. It was concluded that intestinofugal neurons provide an important afferent output from the large intestine that is dependent upon changes in circumferential length, essentially independent of changes in muscle tension. Interestingly, it was found that contraction of the colon that occurred either spontaneously, or when induced by stretch was actually associated with increased firing immediately preceding contractions. This is slightly different from findings of Miller and Szurszewski using the isolated mouse colon-SMG preparation. They found that fast EPSP input into SMG neurons increased during filling (distension) of the colon, but was always reduced during contraction. These investigators provided convincing evidence that, at least in mice, intestinofugal neurons monitor changes in intracolonic volume. Miller and Szurszewski also clearly showed that activation of intestinofugal neurons was dependent upon changes in circumferential length, not changes in longitudinal length of the colon. This was a most convincing experiment, where in the same preparation no fast EPSP was evoked in the SMG to longitudinal stretch, but extensive fast EPSPs were evoked in the same SMG neurons to circumferential stretch. It is worth noting that activation of intestinofugal afferents to the distal colon is very different from the activation of spinal afferents to the distal colon. For example, in the distal colon, spinal afferents are potently activated by contraction of the colon and stretch of the colorectum in both circumferential and longitudinal axis. Whereas intestinofugal neurons are only activated by circumferential stretch.


Taken together, in recent years, direct evidence has been presented to support the notion that at least a population of intestinofugal neurons in the ENS are directly mechanosensory and these neurons are predominantly S-type neurons with Dogiel Type I morphology. It is still not fully clear what the physiological function of intestinofugal neurons is in the GI tract. It is possible that in response to colorectal filling or distension, the activation of sympathetic efferent neurons in prevertebral ganglia by intestinofugal neurons would release noradrenaline into the ENS and directly onto smooth muscle layers to facilitate accommodation of the colon and rectum for the advancing fecal mass.





Enteric Neurons in the Stomach


The organization of the ENS in the stomach is different from the small and large intestines. In the stomach, there is no submucosal plexus, or very sporadic submucosal ganglia and it is thought that the myenteric plexus fulfils the role of the submucosal plexus in this region. Of particular interest is that myenteric neurons with electrophysiological characteristics of AH neurons are absent from the gastric corpus and uncommon in the gastric antrum. Excitation of the gastric antrum and corpus is largely controlled through the vagus nerves, which provides a prominent excitatory synaptic input into the majority of gastric neurons. There is evidence that local distension of the antrum activates intrinsic reflexes and elicits IJPs in the neighboring smooth muscle layers. Although not confirmed, it is possible that these local distension-evoked reflexes in the stomach are triggered by a population of intrinsic sensory neurons which are not Dogiel Type II neurons. As mentioned above, there is now gathering evidence that a population of enteric neurons with S-type electrophysiology and morphologies of Dogiel Type I neurons can also respond directly to mechanical stimulation. And, these neurons may elicit distension-evoked junction potentials in the smooth muscle cells of the stomach.





Role of 5-HT in the ENS and GI Motility


In recent years, there has been major unexpected revisions in our understanding of the role of endogenous 5-HTin the control of GI motility. It is perhaps surprisingly in this day and age, but in terms of GI motility, 5-HT is still one of the most enigmatic substances endogenously synthesized in the gut wall. There is no doubt that 5-HT is synthesized in very high quantities in the mucosa and to a much smaller extent in the ENS (about 1% of neurons). There is also no doubt that exogenous 5-HT (and many other exogenous substances) potently stimulates the ENS and GI motility. However, many exogenous substances can stimulate the ENS and may not even be synthesized in the gut wall. Also, there is a major difference between the effects of exogenous application of 5-HT, compared with endogenously synthesized 5-HT. In recent years, there has been some startling new data that have led to a complete change in the way we view the role of endogenous 5-HT in the generation of neurogenic motor patterns in the GI tract.


Since the early 1950s, there had been a gathering and compelling case that endogenous 5-HT played an important role in the generation and control of propulsive neurogenic motor patterns, like peristalsis and migrating motor complexes in the small intestine and CMMCs in the large intestine. However, with the advent of highly refined genetic techniques and dynamic amperometric recordings to accurately monitor release of 5-HT in real time, the evidence that endogenous 5-HT plays an important role in the generation of GI-motor patterns, at least in the healthy bowel, has never been more unconvincing.


Originally, it was first proposed in the mid-1950s that endogenous 5-HT played a role in the generation of peristalsis after Dr. Büllbring’s laboratory at Oxford University discovered that endogenous 5-HT could be released from the mucosa at a similar time that peristalsis occurred. The fact that there was a similarity in timing between the release of 5-HT from the mucosa and the onset of neurogenic peristalsis led these investigators to propose that release of endogenous 5-HT from the mucosa was causally related to the generation of peristalsis. Dr. Büllbring and Dr. Lin suggested that “…the threshold of intraluminal pressure required to elicit peristaltic contractions was controlled by the amount of 5-HT synthesized locally.” They further suggested that “…that 5-HT, formed and stored locally in the mucous membrane, is released in proportion to the rise in intraluminal pressure and that it sensitizes pressure receptors which are situated in the mucosa.” At the time, these ideas must have seemed revolutionary. And, respect should be given to these early scientists who, in the early 1950s and even up to the 1990s, had comparatively poor technology (relative to now) to help them address the complex question of the role of 5-HT in GI motility.


In support of the early hypothesis that 5-HT was a key mediator of peristalsis, it had clearly been demonstrated that exogenous exposure of 5-HT to the isolated bowel could potently stimulate peristalsis. Then, in the 1980s and 1990s, different laboratories, including our own work, showed that selective antagonists of 5-HT receptors could potently abolish or inhibit neurogenic motor patterns, such as peristalsis and CMMCs in the large bowel.


Further, in the 1990s, the notion that the release of 5-HT from the mucosa was important for peristalsis in the colon was promulgated. This was proposed in isolated guinea pig colon where it was suggested that “…the presence of fecal pellets triggers the release of 5-HT, which acts via both 5-HT3 and 5-HT4 receptors to regulate propulsive activity in guinea pig colon.” Also, far more recently, it was stated in the mouse colon that “removing the mucosa appeared to abolish spontaneous CMMCs, suggesting that the mucosa is normally critical for their generation” and “the trigger for the CMMC appears to be spontaneous or evoked (i.e., a fecal pellet) release of 5-HT from enterochromaffin ‘EC’ cells to stimulate AH neurons.” Amazingly, this brave conclusion was accepted for publication without any recordings of 5-HT release from the mucosa, nor any recordings from any neurons. Others rationalized that if endogenous 5-HT release from the mucosa (and the presence of the mucosa) was critical for peristalsis or CMMCs to occur, then these motor patterns should cease to occur when the mucosa is removed and all mucosal 5-HT release is prevented. It was discovered that if the entire mucosa and submucosal plexus was removed from colon of guinea pigs, peristalsis, and the propulsion of fecal pellets still occurred along the colon, even though there was no more 5-HT release ( Fig. 14.6 ). Indeed, it had been demonstrated some time earlier from studies in the small bowel that removal of the mucosa did not block peristalsis, in the very preparation that Dr. Büllbring had used to make conclusions in the 1950s.




Fig. 14.6


Depletion of neuronal 5-HT with reserpine does not prevent peristalsis in the isolated guinea-pig distal colon. (A) Peristalsis and propulsion of a natural fecal pellet. (B) Immunohistochemical staining for 5-HT shows varicose axons in a myenteric ganglion and internodal strands. (C) After reserpine treatment to deplete neuronal 5-HT and removal of the mucosa and submucosal plexus, peristalsis still reliably occurred. (D) Confirmation that reserpine had depleted neuronal 5-HT.


Similarly, in mice, the hypothesis that the mucosa and release of 5-HT from the mucosa was “critical” for CMMCs to occur became inconceivable when it was discovered that removal mucosa of the mucosa from the colon did not block CMMCs in mouse colon. And, not long after it was hypothesized that release of 5-HT from the mucosa was critical for CMMC generation, it was shown by the same authors that CMMCs still occur in mutant mice lacking tryptophan hydroxylase 1, the enzyme that synthesizes mucosal 5-HT. At last, there was finally a consensus that mucosal 5-HT and the presence of the mucosa were not required for CMMC generation or propagation.


It was important to reconcile the original findings of Dr. Büllbring’s laboratory as to why 5-HT was released every time peristalsis occurred. This has now been explained by elegant real time recordings of 5-HT release by Dr. Bertrand, who developed amperometry to monitor dynamic release of 5-HT from the mucosa of the guinea pig small intestine, during peristalsis. He demonstrated clear results that the release of endogenous 5-HT from the mucosa was a consequence of peristalsis and was not the underlying cause of peristalsis. Dr. Bertrand showed that EC cells that synthesize and release 5-HT were highly sensitive to mechanical deformation. He revealed that muscle contraction, as occurs during peristalsis, was actually the cause of the release of 5-HT from the mucosa, rather than release of 5-HT being a catalytic step necessary for the initiation of peristalsis, as originally hypothesized. The same conclusions were found in the colon. When the first direct real time recordings of 5-HT release from the colonic mucosa were made it was also shown that, similar to the guinea pig small intestine, endogenous 5-HT could be released in a cyclical fashion, at the same time as CMMC generation. However, importantly, when the mucosa was carefully removed, it abolished all release of 5-HT, but CMMCs generation persisted. It is now apparent that if the mucosa is sharply dissected too aggressively from the myenteric plexus it destroys the delicate and intricate neural pathways required for CMMC generation. This would readily explain the erroneous conclusion that CMMCs are blocked by removal of the mucosa. Taken together, these recent findings have led to a major revision in our understanding of the functional role of endogenous 5-HT in the generation of propulsive motor patterns in the lower GI tract. It is important to now move forward and identify key mechanosensory transduction processes in the myenteric plexus that underlie distension-evoked peristalsis and CMMCs.


Our original assumption that 5-HT was likely to be a major player in the control of colonic motility, at least in mice, turned out to be incorrect because we recently discovered that when the same 5-HT3 antagonists were applied to isolated colon preparations that were completely depleted of endogenous 5-HT, the 5-HT3 antagonists still blocked CMMCs. In retrospect, this was thoroughly unsurprising because 5-HT is not a neurotransmitter in the ENS of mouse colon, as all fast synaptic inputs are nicotinic in origin.


It is important to recognize that while endogenous 5-HT is not a requirement for peristalsis nor CMMC generation or propagation to occur in healthy control bowel in vitro or in vivo ( Fig. 14.7 ), there is significant data that serotonergic signaling, especially with respect to 5-HT reuptake (via SERT) transporter; and genetic ablation of SERT exacerbates intestinal inflammation. Therefore, there is sound reason to believe that in human patients with inflammatory bowel disease (IBD) changes in SERT could contribute to abnormal GI transit. Therefore, serotonergic signaling in IBD could be a major mechanism underlying the pathogenesis of colorectal and intestinal inflammation. However, it needs to be confirmed that changes in SERT expression and 5-HT signaling are not a consequence of intestinal inflammation leading to changes in GI motility, rather than an actual underlying cause of the changes in motility.


Apr 21, 2019 | Posted by in ABDOMINAL MEDICINE | Comments Off on Enteric Nervous System Structure and Neurochemistry Related to Function and Neuropathology

Full access? Get Clinical Tree

Get Clinical Tree app for offline access