1,253
Views
0
CrossRef citations to date
0
Altmetric
Review

The influence of Nav1.9 channels on intestinal hyperpathia and dysmotility

, &
Article: 2212350 | Received 10 Mar 2023, Accepted 02 May 2023, Published online: 15 May 2023

ABSTRACT

The Nav1.9 channel is a voltage-gated sodium channel. It plays a vital role in the generation of pain and the formation of neuronal hyperexcitability after inflammation. It is highly expressed in small diameter neurons of dorsal root ganglions and Dogiel II neurons in enteric nervous system. The small diameter neurons in dorsal root ganglions are the primary sensory neurons of pain conduction. Nav1.9 channels also participate in regulating intestinal motility. Functional enhancements of Nav1.9 channels to a certain extent lead to hyperexcitability of small diameter dorsal root ganglion neurons. The hyperexcitability of the neurons can cause visceral hyperalgesia. Intestinofugal afferent neurons and intrinsic primary afferent neurons in enteric nervous system belong to Dogiel type II neurons. Their excitability can also be regulated by Nav1.9 channels. The hyperexcitability of intestinofugal afferent neurons abnormally activate entero-enteric inhibitory reflexes. The hyperexcitability of intrinsic primary afferent neurons disturb peristaltic waves by abnormally activating peristaltic reflexes. This review discusses the role of Nav1.9 channels in intestinal hyperpathia and dysmotility.

Introduction

The Nav1.9 channel is a type of voltage-gated sodium channel (VGSC), that can regulate the resting membrane potential and amplify subthreshold stimuli [Citation1]. The α subunit encoded by the SCN11A gene is the main functional subunit of Nav1.9 channels [Citation2]. Nav1.9 channels are highly expressed in membranes of small diameter afferent neurons in dorsal root ganglions (DRGs) and trigeminal ganglions, and Dogiel II neurons in the enteric nervous system (ENS) [Citation3,Citation4]. In the past, Nav1.9 channels were mainly considered to be involved in the formation of pain sensing [Citation5]. The small-diameter neurons of the DRG are the primary neurons involved in pain transduction. The pathogenic variations of the SCN11A gene lead to three genetic pain related disorders: familial episodic pain type III (FEP type III), congenital insensitive to pain (CIP), and genetic small fiber neuropathy (SFN) [Citation6,Citation7]. Some patients carrying the pathogenic variants in the SCN11A gene exhibit not only abnormal pain, but also gastrointestinal symptoms [Citation8–11]. Recently, Nav1.9 channels have also been found to participate in regulating colonic motility [Citation12]. Intestinal hyperpathia and dysmotility are vital physiopathologic mechanisms in certain functional gastrointestinal disorders (FGIDs), such as irritable bowel syndrome (IBS) [Citation13,Citation14]. This review addresses the role of Nav1.9 channels in intestinal hyperpathia and dysmotility. Its underlying mechanisms are explained.

Structure and function of Nav1.9 channels

The Nav1.9 channel consists of α and β subunits. The α subunit is the main functional subunit and forms ion-selective channels. The β subunit is the coregulatory subunit. A total of 10 sodium channels have been found in mammals, including 9 VGSCs. According to the α subunit, VGSCs are named Nav1.1–1.9 channels. The molecular weight of the α subunit is approximately 260 kDa, and there is only one α subunit per channel. The α subunit contains four homologous transmembrane domains (DI-DIV), which are linearly arranged in the same chain. Each domain contains six transmembrane α helical segments (S1-S6) [Citation15]. S1-S4 comprise the voltage sensing domain (VSD). The interlinked peptide chain between S5 and S6 is called the pore loop (P loop). The P loops of the four domains form the sodium ion selective filter. The S5-S6 and P loop form the pore domain. In the VSD, there is a positively charged amino acid residue (arginine or lysine) in S4 located on every 3 amino acid residues. Inactivation motifs (IFMs) containing isoleucine, phenylalanine, and methionine are found in the intracellular ring between the DIII and DIV domains. The IFM can block the channel and participate in inactivation of the channel when the membrane depolarizes continuously. There are drug and toxin binding sites on the α subunit [Citation16]. β subunits modulate the activation and inactivation kinetics of VGSCs, regulate channel density at the membrane, and may also have a role in regulating channel subcellular localization [Citation1,Citation17] (). The expression of Nav1.9 channels peaks in adolescence and declines significantly after the beginning of adulthood [Citation7].

Figure 1. Schematic diagram of the VGSC structure. VGSCs comprise 4 transmembrane domains (DI-IV). Each domain consists of 6 transmembrane segments (S1–6). S1–4 comprise the voltage sensing domain. There are positively charged amino acid residues in S4 on every 3 amino acid residues. The interlinked peptide chain between S5 and S6 is called the pore loop (P loop). The P loops of the four domains form the sodium ion selective filter. The S5-S6 and P loops form the pore domain. The intracellular inactivation motif between DIII and DIV can block the channel. There is a tetrodotoxin binding site on the channels. The 4 domains form a pore inserted through the cell membrane. D, domain; S, segment; TTX, tetrodotoxin; +, positive charge. This figure was modified from Coates [Citation3].

Figure 1. Schematic diagram of the VGSC structure. VGSCs comprise 4 transmembrane domains (DI-IV). Each domain consists of 6 transmembrane segments (S1–6). S1–4 comprise the voltage sensing domain. There are positively charged amino acid residues in S4 on every 3 amino acid residues. The interlinked peptide chain between S5 and S6 is called the pore loop (P loop). The P loops of the four domains form the sodium ion selective filter. The S5-S6 and P loops form the pore domain. The intracellular inactivation motif between DIII and DIV can block the channel. There is a tetrodotoxin binding site on the channels. The 4 domains form a pore inserted through the cell membrane. D, domain; S, segment; TTX, tetrodotoxin; +, positive charge. This figure was modified from Coates [Citation3].

The depolarization of the cell membrane potential induces the activation and opening of VGSCs. It selectively mediates transmembrane sodium influx, and is one of the molecular bases of cell action potential generation. The gating mechanisms of VGSCs can be summarized as a “two-gate three-state model” [Citation3]. VGSCs have an activation gate and an inactivation gate, and three states, including resting (standby), activation and inactivation states. When the cell membrane is at resting potentials, the activation gate of VGSCs closes, and the inactivation gate opens. VGSCs could respond to the electrical signal depolarization stimulus. When cell membrane depolarization reaches the activation threshold, the position of S4 could shift, leading to conformational changes in the α subunit. The activation and inactivation gates of VGSCs both open, temporarily allowing sodium ions to enter cells. Then, in the fast inactivation mode, the activation gate opens. However, the inactivation gate closes. The IFM blocks the channel area. Later, other ion channels are activated to generate a hyperpolarized membrane potential (approximately −80 mV), which causes S4 to return to its resting position. Both the activation and inactivation gates close, which is referred to as slow inactivation. In the inactive state, VGSCs cannot be activated again. With the repolarization of the membrane potential, the channel revives and returns to its resting state [Citation7] ().

Figure 2. The gating mechanisms of VGSCs. (a) the cell membrane is at the resting potential. The pore is closed, and the IFM is in the resting conformation. (b) the depolarization above the threshold potential makes S4 move outwards and the pore open. Sodium ions flow into the cell. (c) the changes in potential after pore opening cause the IFM to block the open pore. (d) Activity of other ion channels produces a hyperpolarized membrane potential, causing S4 to return to its resting position. In addition, the pore closed. This figure was modified from Bennett [Citation7].

Figure 2. The gating mechanisms of VGSCs. (a) the cell membrane is at the resting potential. The pore is closed, and the IFM is in the resting conformation. (b) the depolarization above the threshold potential makes S4 move outwards and the pore open. Sodium ions flow into the cell. (c) the changes in potential after pore opening cause the IFM to block the open pore. (d) Activity of other ion channels produces a hyperpolarized membrane potential, causing S4 to return to its resting position. In addition, the pore closed. This figure was modified from Bennett [Citation7].

The α subunit of the Nav1.9 channel is encoded by the SCN11A gene (OMIM: * 604385). At present, the Human Gene Mutation Database (HGMD) contains information regarding 32 pathogenic mutations of the SCN11A gene. Human Nav1.9 channels contain 1791 amino acid residues and are classified as tetrodotoxin-resistant (TTX-R) channels in pharmacology [Citation18]. TTX is a potent neurotoxin that blocks VGSCs. TTX binding to the α-subunit within the outer vestibule of the VGSC in a highly selective manner, blocking the entry of Na+ ions through the channel. VGSCs are classified on the basis of its sensitivity to TTX. TTX-R VGSCs are inhibited at micromolar concentrations of TTX. However, TTX-sensitive VGSCs are blocked by nanomolar concentrations. TTX-sensitive VGSCs include Nav1.1, Nav1.2, Nav1.3, Nav1.4, Nav1.6 and Nav1.7 channels. Nav1.5, Nav1.8 and Nav1.9 channels belong to TTX-R VGSCs [Citation19]. The electrophysiological characteristics of Nav1.9 channels are relatively unique. Its activation voltage is very low and is close to the resting membrane potential of the neuronal membrane (−70 mV ~ −40 mV). In addition, the inactivation of Nav1.9 channels is ultraslow, with steady-state activation and inactivation overlapping. Nav1.9 channels can open spontaneously and continuously at voltages within a specific overlap range, generating a large window current [Citation20]. Thus, Nav1.9 channels act as threshold channels to regulate resting membrane potentials by generating inward sodium currents. In addition, Nav1.9 channels respond to subthreshold stimulations [Citation20,Citation21], which reduces the discharge threshold of a single action potential and increases the number of repeated neuronal discharges [Citation22].

Heterologous expression of Nav1.9 channels was difficult previously, which limited the screening of Nav1.9 channels modulators [Citation23]. Recently HEK-293 cells and ND7/23 cells that heterologously express Nav1.9 channels were established [Citation24,Citation25]. They can facilitate identification of selective Nav1.9 modulators. There are spider venom-derived peptides specifically activating Nav1.9 channels [Citation26,Citation27]. The leech peptide HSTX-I exerts significant analgesic function by specifically inhibiting Nav1.8 and 1.9 channels [Citation28]. However, another study suggested that HSTX-I was not a useful lead for development of analgesic Nav1.8 and 1.9 modulators [Citation29]. The specific inhibitors of Nav1.9 channels need further study.

The role of Nav1.9 channels in intestinal hyperpathia

Nav1.9 channels and pain production

The process of pain generation is complex. The endings of nociceptive sensory neurons express some ligand-gated ion channels, G-protein-coupled receptors and tyrosine kinase receptors. Through these receptors, nociceptive neurons transmit signals regarding extreme temperature, low pH, and various chemical stimuli into electrical signals. VGSCs play a key role in the response to small depolarization potentials and action potential generation [Citation7] (). Then, neural impulses are transmitted to the dorsal horn of the spinal cord, which is the primary integration center for pain signal processing in the trunk and limbs. Neural impulses are gradually relayed to the cerebral cortex through the spinal cord. The cerebral cortex produces pain. Hyperalgesia indicates a decreased pain threshold for stimuli. Allodynia is a condition in which pain is triggered by stimuli that are not supposed to cause pain.

Figure 3. The generation of action potentials in nociceptor neurons. (a) These neurons transduce signals regarding extreme temperatures (Temp), low pH, and various chemical stimuli into electrical signals by some of the ligand-gated ion channels (blue), G-protein-coupled receptors (light blue), and tyrosine kinase receptors (green). Then, VGSCs, such as Nav1.7, Nav1.8 and Nav1.9 (red), have a key role in responding to small depolarizations and action-potential generation. The hyperpolarizing activated cation channel HCN2 acts as a so-called pacemaker, modulating ectopic activity after nerve injury. Voltage-gated potassium channels (yellow) are important breaks in excitability. This figure was cited from Bennett [Citation7]. (b) Representative action potential waveform recorded from a small-diameter nociceptive DRG neuron, excited by a depolarizing current step of 200 pA. The contributions of the Nav1.7, Nav1.8 and Nav1.9 channels to the generation of the action potential are indicated. This figure was modified from Bennett [Citation1].

Figure 3. The generation of action potentials in nociceptor neurons. (a) These neurons transduce signals regarding extreme temperatures (Temp), low pH, and various chemical stimuli into electrical signals by some of the ligand-gated ion channels (blue), G-protein-coupled receptors (light blue), and tyrosine kinase receptors (green). Then, VGSCs, such as Nav1.7, Nav1.8 and Nav1.9 (red), have a key role in responding to small depolarizations and action-potential generation. The hyperpolarizing activated cation channel HCN2 acts as a so-called pacemaker, modulating ectopic activity after nerve injury. Voltage-gated potassium channels (yellow) are important breaks in excitability. This figure was cited from Bennett [Citation7]. (b) Representative action potential waveform recorded from a small-diameter nociceptive DRG neuron, excited by a depolarizing current step of 200 pA. The contributions of the Nav1.7, Nav1.8 and Nav1.9 channels to the generation of the action potential are indicated. This figure was modified from Bennett [Citation1].

The DRG is an oval enlargement of the spinal nerve at the intervertebral foramen, which is also known as the sensory ganglia. The DRG is formed from the aggregated cell bodies of pseudomonopolar neurons in vertebrates. The free ends of the process of these cells are used as receptors, which are distributed in the skin, joints, muscles and internal organs. The central process of DRG neurons projects nerve impulses to the posterior horn of the spinal cord. The sensory neurons contained in the DRG are important signal integration and conduction centers in sensory pathways, such as those for pain, temperature and pressure sensations. Small-diameter neurons in the DRG are primary neurons involved in pain transduction [Citation30]. DRG neurons can be divided into large-, medium- and small-diameter neurons. Large-diameter neurons mainly emit Aα and Aβ fibers with coarse myelin. Medium-diameter neurons mainly emit Aδ fibers with fine myelin. Small neurons mainly emit C fibers without myelin [Citation1]. Nav1.9 channels are mainly expressed in small-diameter DRG neurons [Citation31,Citation32], but also in medium- and large-diameter neurons [Citation33]. Nav1.9 channels are distributed on the cell membranes of the cell bodies, central processes and peripheral processes of these primary afferent neurons [Citation5,Citation34,Citation35]. During the action potential production of small-diameter DRG neurons, the Nav1.9 channel is responsible for amplifying subthreshold stimuli. The Nav1.7 channel is responsible for amplifying subthreshold stimuli and is involved in the formation of action potential ramus. The Nav1.8 channel is the main VGSC responsible for the formation of the action potential rising phase [Citation1] ().

The Nav1.9 channel is involved in the production of inflammatory pain [Citation36]. First, inflammatory factors, such as bradykinin, histamine and prostaglandin E2, can directly or indirectly upregulate the activity of Nav1.9 channels [Citation37]. Inflammation reduces local cholesterol levels. The decrease in cholesterol levels can sensitize nociceptive neurons and promote hyperalgesia [Citation38]. In addition, the inflammatory response upregulates the expression of Nav1.9 channels in DRG sensory neurons [Citation39].

Both Nav1.7 and Nav1.9 are involved in the summation of subthreshold stimuli and determining the threshold for action potential generation. Pathogenic mutations in Nav1.7 channels cause several pain related disorders in human, including congenital insensitivity to pain, inherited erythromelalgia, paroxysmal extreme pain disorder and inherited small-fiber neuropathy [Citation1]. Patients with mutations in Nav1.7 channels do not manifest the same intestinal complications that are observed in patients with mutations in Nav1.9 channels. Nav1.7 is mainly abundantly expressed in vagal sensory neurons and DRG sensory neurons [Citation1]. Therefore, mutant Nav1.7 channels have limited effect on ENS. These patients are not affected with obvious intestinal dysmotility. But Nav1.7 channels can influence gut sensation. For example, paroxysmal extreme pain disorder is originally known as familial rectal pain [Citation40].

Pain-related genetic diseases associated with SCN11A gene variation

Pathogenic SCN11A variants can lead to three different genetic diseases: FEP type III, genetic SFN and CIP [Citation5]. FEP type III is also called familial episodic limb pain, which exhibits clinical features of paroxysmal limb pain. It is a rare autosomal dominant genetic disease. Its incidence in the population has not been reported. Only individual cases have been reported at present [Citation8,Citation9,Citation41–44]. Gain-of-function SCN11A mutations cause FEP type III. Mutant Nav1.9 channels in patients with FEP are more likely to open, which results in the hyperexcitability of small-diameter DRG neurons. Then, the hyperexcitability of the neurons ultimately leads to hyperpathia [Citation42].

SFN is also known as painful small fiber neuropathy. Its main clinical manifestations are pain, numbness, burning and other paresthesia. SFN is usually accompanied by autonomic nervous dysfunction, such as hyperhidrosis, and gastrointestinal motility disorders. It can be caused by metabolic diseases, drug, autoimmune, infection, genetics and other causes, which affect the fibers of the peripheral sensory nerves. Gain-of-function mutations in the SCN11A gene lead to partial genetic SFN [Citation45–49]. This mechanism can partially explain the pathogenesis of SFN [Citation50,Citation51].

The main clinical characteristics of CIP caused by the SCN11A gene are insensitivity to pain, self-injurious behaviors, and slow wound healing. The condition may be accompanied by itching, Charcot arthropathy, muscle weakness and abnormal intestinal motility. It is a rare genetic disease, and most studies are case reports [Citation10,Citation11,Citation52–54]. The obvious gain-of-function SCN11A mutations could mediate significantly more sodium influx and directly reach the inactivation potentials of Nav1.7 and Nav1.8 channels. Nav1.7 and Nav1.8 channels cannot mediate sodium influx, which hinders the formation of action potentials, and ultimately leads to insensitivity to pain [Citation55].

Some patients carrying SCN11A mutations showed intestinal dysmotility [Citation8–11]. For example, a patient with CIP carrying the L396P variant had constipation [Citation10]. The FEP patient with the R222S variant exhibited diarrhea [Citation41]. In addition, the CIP patient with the L811P variant has alternate symptoms of diarrhea and constipation [Citation11].

Alterations in the function of Nav1.9 channels and intestinal hyperpathia

Scn11a gene knockout (Scn11a−/−) mice showed no significant changes in visceral nociceptive sensitivity under the noninflammatory state. However, Scn11a−/− mice exhibited decreased visceral nociceptive hyperreactivity induced by inflammatory mediators. In electrophysiological tests, the responses of intestinal afferent nerves to pain-causing inflammatory mediators (adenosine triphosphate (ATP) and prostaglandin E2) were significantly lower than those in the wild-type (WT) group [Citation56]. In addition, Scn11a−/− and WT mice were tested for the sensitivity to colorectal pain induced by mechanical distension stimuli using the abdominal withdrawal reflex (AWR) test. There was no significant difference between the two groups. Then, after the mice were subjected to enema with an inflammatory inducer, WT mice showed higher levels of colorectal pain sensitivity than the placebo group. However, Scn11a−/− mice responded similarly to the placebo treatment. This finding suggested that Nav1.9 channels were involved in inflammation-related acute colorectal hyperalgesia [Citation57]. Such afferent fibers, which are normally insensitive to stimuli but are activated in an inflammatory state, are known as silent or “dormant” pain receptors [Citation58].

In contrast, Scn11aR222S/R222S mice carried the gain-of-function R222S variant in the Scn11a gene, which was homologous to the R222S variant in FEP patients. They showed not only somatic hyperalgesia, but also visceral hyperalgesia. In the AWR test, the Scn11aR222S/R222S mice presented with increased pain sensitivity to mechanical stimulation. In addition, the acute inflammatory visceral pain thresholds of Scn11aR222S/R222S mice also decreased significantly [Citation59]. Some patients with pathogenic SCN11A gene variants also have abdominal pain symptoms [Citation8,Citation9].

The potential mechanisms underlying the effects of Nav1.9 channels on visceral hyperalgesia

Extrinsic afferent neurons distributed in the gastrointestinal tract include spinal afferent nerves and vagal afferent nerves, both of which innervate the whole gastrointestinal tract. In the upper digestive tract, the main extrinsic afferent nerves come from vagus nerves. In the lower digestive tract, the distribution of vagus nerves gradually decreases, and the number of spinal afferent nerves increase. The vagus nerves are not involved in sensory formation. The main external afferents distributed in the colon are spinal afferents, which have cell bodies that are located in the DRG [Citation60]. Lumbosacral and vertebral DRGs project fibers to the distal colon and bladder along with the pelvic nerves [Citation61,Citation62]. Except for the distal colon, spinal afferent nerves in other parts of the intestinal tract mainly travel along the splanchnic nerves and merge into the thoracic and lumbar DRGs [Citation63].

The small-diameter neurons in the DRG are primary neurons of peripheral pain transduction, which mainly belong to type C nerve fibers [Citation64]. They could be divided into non-peptidergic neurons that could be labeled by IB4 and peptidergic neurons that expressed calcitonin gene related peptide (CGRP). The Nav1.9 channel is mainly expressed in nonpeptidergic DRG neurons (IB4+) [Citation35]. In addition, 14% of sensory neurons distributed in the colon are IB4+ neurons [Citation65]. Single-cell RNA sequencing of DRG neurons revealed that the Nav1.9 channel is expressed in peptidergic neurons, non-peptidergic neurons, myelinated neurons and nonmyelinated neurons [Citation66]. A study showed that Nav1.9 channels are expressed in half of the nerve fibers projected to the intestine by DRG [Citation56]. Gain-of-function changes in Nav1.9 channels to some extent lead to hyperexcitability of small-diameter DRG neurons. Under the circumstance, some subthreshold stimulations can generate nerve impulses in these neurons. Small-diameter DRG neurons are primary neurons for pain signal processing in the trunk and limbs. Therefore, the alterations of Nav1.9 channels could furtherly cause hyperpathia, including visceral hyperalgesia. Small-diameter DRG neurons in Scn11aR222S/R222S mice, for instance, showed hyperexcitability. These mice manifested visceral hyperpathia [Citation59].

The role of Nav1.9 channels in intestinal dysmotility

The neural regulation of intestinal motility

The ENS is relatively independent of the central nervous system (CNS), also known as the gut brain or the second brain. The ENS has complete internal neural reflex circuits. It can maintain intestinal function, even if the connections with the CNS and autonomic nervous system (ANS) are completely cut off [Citation67]. The ENS comprises the myenteric plexus and the submucosal plexus. The myenteric plexus is mainly responsible for the regulation of intestinal movements. The submucosal plexus is involved in the secretion of water and electrolytes, as well as the neural regulation of intestinal blood flow [Citation68]. The ENS plays a leading role in the neuroregulation of intestinal motility. For example, in one experiment, the exogenous nerve fibers of isolated intestinal segments in guinea pigs were completely cut off. Then, intestinal segments continued to carry out complex neuro-driven propulsive movements in the organ bath by relying on the ENS [Citation69].

Neurons in the ENS are mainly divided into Dogiel type I and Dogiel type II neurons in morphology. The cell bodies of Dogiel type I neurons project multiple short dendrites and a single elongated axon. Dogiel type II neurons have smooth cell bodies and project multiple axons, but no dendrites or only a few dendrites [Citation70]. In addition, ENS neurons are divided into sensory neurons, interneurons and motor neurons in terms of functional characteristics [Citation71], and AH-type neurons and S-type neurons in terms of electrophysiological characteristics [Citation72]. AH-type neurons are characterized by the larger action potentials with an infection on the falling phase and long afterhyperpolarizing potential (AHP;>2 s) that follows. S-type neurons are typified by the brief action potentials without slow AHPs, and they present fast excitatory postsynaptic potentials (EPSPs) [Citation68]. The intrinsic sensory neurons/intrinsic primary afferent neurons (IPANs) of the ENS generally have the characteristics of Dogiel II neurons in morphology, and belong to AH neurons in electrophysiological characteristics [Citation67,Citation73,Citation74]. However, some Dogiel type I neurons (intermediate neurons) can function as mechanosensory neurons [Citation75,Citation76]. Nav1.9 channels are mainly expressed in AH-type/Dogiel type II neurons in the ENS [Citation4,Citation77,Citation78]. Nav1.9 channels flow at subthreshold voltages, generating tonic firing. They remain active during the falling phase, opposes action potential repolarization [Citation4].

The gastrointestinal tract displays a wide range of complex motor patterns. Two of the major neurogenic contractile behaviors are segmentation contractions [Citation79,Citation80] and peristalsis [Citation81–83]. Peristaltic reflexes are the basis of most intestinal propulsive movements [Citation84]. However, the identities of the neural circuits participating in other motility patterns remain elusive. How the gut transitions between different motility patterns requires further investigation [Citation68].

Dogiel type II neurons play an important role in intestinal peristaltic reflexes [Citation60]. In peristaltic reflexes, AH neurons/IPANs are activated by mechanical pulling on the intestinal wall and/or chemical stimulation of intestinal contents. After generating nerve impulses, IPANs transmit excitatory signals to both ascending and descending interneurons by releasing substance P (SP), acetylcholine (Ach), glutamate and calcitonin gene-related peptide (CGRP). Then, ascending interneurons release SP and Ach to the excitatory motor neurons on the oral side. The excitatory motor neurons release SP and Ach at the neuromuscular junction, causing smooth muscle contraction. At the same time, the descending interneurons release 5-HT and Ach to the inhibitory motor neurons in the anal side. Then, the inhibitory motor neurons release nitric oxide (NO), vasoactive intestinal polypeptide (VIP) and ATP to the neuromuscular junction, causing the smooth muscle to relax [Citation68,Citation85].

The extrinsic neuromodulation of intestinal motility mainly depends on the ANS. The main neurotransmitter released by sympathetic nerves at synapses is noradrenaline (NE). NE relaxes the smooth muscles of the intestine and inhibits peristalsis. Ach is the main neurotransmitter released by the parasympathetic nerve at synapses. Ach can make intestinal smooth muscle contract and promote intestinal motility [Citation86]. Intestinofugal afferent neurons (IFANs) are a special type of afferent neuron in the ENS. In other words, there are four types of intestinal afferent neurons: IPANs, IFANs, primary afferent neurons with cell bodies located in the DRG and primary afferent neurons with cell bodies located in the vagal ganglion [Citation87]. IFANs belonging to Dogiel II neurons can project impulses to sympathetic neurons in the prevertebral ganglion (PVG), which in turn innervate the intestine [Citation87,Citation88]. IFANs and sympathetic PVG neurons form the basis of entero-enteric inhibitory reflexes, which do not require CNS involvement [Citation89,Citation90]. When mechanical stimulation of intestinal dilation activates the IFANs, the information is transmitted upward to sympathetic motor neurons in the PVG [Citation91,Citation92], and these sympathetic postganglial nerve fibers project into the intestine and inhibit intestinal motility [Citation60]. For example, when a section of the colon of guinea pigs is stretched and dilated, the external sympathetic reflex pathway can inhibit peristalsis of adjacent isolated colon segments in the organ bath [Citation93].

There are several subtypes of VGSCs expressed in the human ENS, including Nav1.1, Nav1.2, Nav1.3, Nav1.5, Nav1.6, Nav1.7 and Nav1.9 channels. The Nav1.5 and Nav1.9 channels are expressed in enteric neurons of mice [Citation63]. Nav1.9 channels are expressed in the intermuscular plexus and submucosal plexus [Citation77,Citation94]. The Nav1.9 channel is mainly distributed in Dogiel II neurons, while the Nav1.5 channel is highly expressed in Dogiel I and Dogiel II neurons in mice. Nav1.9 channels are responsible for amplifying subliminal stimuli and regulating the firing threshold of action potentials in Dogiel II neurons in the ENS, while Nav1.5 channels open to mediate sodium influx and form the ascending branch of action potentials in mice [Citation4,Citation78]. In other words, the Nav1.9 channel is expressed in the ENS, and in small-diameter neurons of the DRG that are associated with the ENS and CNS. Therefore, the functional and structural changes in Nav1.9 channels have a structural basis for affecting intestinal function.

Alterations in the function of Nav1.9 channels and intestinal dysmotility

Loss-of-function alterations in Nav1.9 channels promote intestinal motility in mice, whereas gain-of-function changes in Nav1.9 channels inhibit intestinal motility in mice. In one study, the migrating motor complex in the isolated colon of Scn11a−/− mice was examined. The whole colon was taken and fixed in an organ bath filled with Kirschner’s fluid. The mechanical activity of the circumferential muscles in the proximal, middle and distal parts of the WT and Scn11a−/− mice was recorded. The mean frequency and area under the contraction curve in the Scn11a−/− mice were significantly higher than those in WT mice [Citation12]. This finding suggested that loss-of-function alterations in Nav1.9 channels promote intestinal motility.

However, a CIP patient with a gain-of-function L811P mutation was observed to exhibit reduced intestinal motility during laparoscopic surgery [Citation11]. We summarized the functional effects of SCN11A variants on electronical properties, pain sensing, and potential roles in intestinal hyperpathia and dysmotility ().

Table 1. Summarization of SCN11A variants.

Scn11a+/L799P mice carried the mutation that is homologous to the L811P mutation in humans. The intestinal segment of Scn11a+/L799P mice showed a slightly decreased spontaneous peristalsis frequency, but there was no statistically significant difference from that observed in the WT mice [Citation95]. The carbon powder propelling test was used to investigate the difference in intestinal motility of Scn11aR222S/R222S mice in vivo. The distances of the carbon powder traveled in the Scn11aR222S/R222S mice were significantly shorter than those in WT mice. These results suggested that intestinal peristalsis was slower in Scn11aR222S/R222S mice. The mechanical activity of small intestine segments was evaluated. The frequency of contractions in intestinal segments from Scn11aR222S/R222S mice was lower than that in WT mice [Citation59]. Therefore, the gain-of-function changes in Nav1.9 channels inhibit intestinal motility in mice.

The potential mechanisms underlying the effects of Nav1.9 channels on intestinal dysmotility

Visceral sensory afferent neurons overstimulate motor neurons in the PVG, which can affect local motion of the gastrointestinal tract [Citation14]. Scn11aR222S/R222S mice showed a slight increase in intestinal NE concentrations in tissues [Citation59] (). The gain-of-function changes in Nav1.9 channels may lead to hyperexcitable IFAN afferent fibers, which could abnormally activate entero-enteric inhibitory reflexes. The sympathetic motor neurons in the PVG are then stimulated. These sympathetic efferent nerve fibers can then release the neurotransmitter norepinephrine, and inhibit intestinal motility.

Figure 4. The underlying mechanisms for the hyperexcitability in Nav1.9 channels that causes intestinal dysmotility. (a) the influence on entero-enteric inhibitory reflexes. Gain-of-function alteration in Nav1.9 channels could lead to hyperexcitability of IFANs. The excess stimulation by IFANs might cause sympathetic neurons in PVG releasing more NE. NE could dampen intestinal peristalsis. LM: longitudinal smooth muscle, MP: myenteric plexus, CM: circular smooth muscle, SMP: submucosal plexus, PVG: prevertebral ganglion, IFANs: intestinofugal afferent neurons, NE: noradrenaline. (b) Alteration in peristalsis reflexes. Gain-of-function change in Nav1.9 channels cause hyperexcitability of IPANs. IPANs: intrinsic primary afferent neurons. (c) the abnormal peristalsis reflexes triggered by hyperexcitability IPANs may destroy intestinal peristalsis waves. Overactive firing in IPANs might lead to overlapping paradoxical signals. The mixed signals may disrupt normal pressure gradients in peristalsis. This figure was modified from Zhao [Citation59]. .

Figure 4. The underlying mechanisms for the hyperexcitability in Nav1.9 channels that causes intestinal dysmotility. (a) the influence on entero-enteric inhibitory reflexes. Gain-of-function alteration in Nav1.9 channels could lead to hyperexcitability of IFANs. The excess stimulation by IFANs might cause sympathetic neurons in PVG releasing more NE. NE could dampen intestinal peristalsis. LM: longitudinal smooth muscle, MP: myenteric plexus, CM: circular smooth muscle, SMP: submucosal plexus, PVG: prevertebral ganglion, IFANs: intestinofugal afferent neurons, NE: noradrenaline. (b) Alteration in peristalsis reflexes. Gain-of-function change in Nav1.9 channels cause hyperexcitability of IPANs. IPANs: intrinsic primary afferent neurons. (c) the abnormal peristalsis reflexes triggered by hyperexcitability IPANs may destroy intestinal peristalsis waves. Overactive firing in IPANs might lead to overlapping paradoxical signals. The mixed signals may disrupt normal pressure gradients in peristalsis. This figure was modified from Zhao [Citation59]. .

In regions of inflammation, IPANs are spontaneously active and synaptic activities are enhanced. These changes result in the overlap of upward excitatory signals and downward inhibitory signals in the inflammatory region. In addition, inhibitory neuromuscular transmission is decreased. When peristaltic waves arrive at the region, the peristalsis is disrupted by the mixed signals and the suppressed neuromuscular transmission [Citation96]. Scn11aR222S/R222S mice showed dampened intestinal motility. IPANs in Scn11aR222S/R222S mice might exhibit hyperexcitability [Citation59] (). They might be activated by subthreshold stimulations. If abnormal signals triggered by overactive IPANs were contradictory to normal signals in the same area, mixed signals would appear. The mixed signals might disrupt normal pressure gradients during peristalsis (). In addition, the persistent TTX-R Na+ currents mediated by the Nav1.9 channels may be involved in the regulation of neuronal excitability in vagal afferent neurons. The activation of visceral afferent fibers may induce autonomic changes and alterations in colonic tone (e.g. vagally mediated gastrocolonic motor response). Therefore, Nav1.9 channels may affect intestinal motility by autonomic changes. However, the regulation process of gastrointestinal motility is very complex, and the effect of Nav1.9 channels on intestinal motility and its specific mechanisms still need to be further studied. Performing RNA-seq sequencing on intestinal tissue in both Scn11a variants knock-in and wild-type mice, for instance, is helpful to investigate potential pathways for this complex network. In order to analyze the electrical activities, it is necessary to isolate enteric neurons from the myenteric plexus of Scn11a variants knock-in mice. Then using whole-cell patch clamp technique to check whether IFANs and IPANs manifesting hyperexcitability. Besides, intestinal myoelectrical activities in Scn11a variants knock-in mice are worthwhile to research.

Outlook on the relationship between Nav1.9 channels and IBS

Intestinal hyperalgesia and dysmotility play an important role in the occurrence and development of functional bowel diseases, especially IBS. IBS is a functional bowel disorder characterized by abdominal pain, distention, or abdominal discomfort. It is often accompanied by changes in bowel habits [frequency and/or traits]. Routine clinical examinations typically show no biochemical or morphological abnormalities that could explain these symptoms. The pathophysiological mechanism of IBS has not been fully elucidated, but visceral hypersensitivity (VH) is considered to be the core pathogenic mechanism underlying IBS. VH refers to the enhanced sensitivity of visceral tissues and organs to stimuli, which mainly manifest as hyperalgesia and dysalgesia [Citation97]. Currently, the internationally recognized diagnostic criteria for IBS, the Rome IV criteria, are recurrent abdominal pain with a frequency of more than 1 day/week in the last 3 months as a necessary condition for the diagnosis of IBS [Citation98]. VH occurs in 33–90% of patients with IBS [Citation99]. Moreover, studies have found that visceral and somatosensory thresholds in patients with IBS are both significantly lower than those in healthy controls [Citation100]. Gastrointestinal dysmotility is an important pathophysiological mechanism underlying IBS, and is mainly manifested in intestinal dysmotility [Citation101,Citation102]. In patients with constipation-predominant IBS, the amplitude and frequency of the intestinal migrating motor complex (MMC) are reduced [Citation103], and the intestinal transmission time is also prolonged [Citation101].

IBS following acute gastrointestinal infection is known as postinfectious irritable bowel syndrome (PI-IBS) [Citation104]. IBS was previously considered to be a functional gastrointestinal disease without morphological changes and biochemical abnormalities. However, it is currently believed that chronic low-grade inflammation is involved in the pathogenesis of PI-IBS [Citation105,Citation106]. It is well known that the Nav1.9 channel is normally “dormant” or “silent,” but is involved in inflammatory pain. The expression levels and activities of Nav1.9 channels are upregulated after acute inflammation [Citation37,Citation39]. In addition to acute inflammation, the Nav1.9 channel is involved in hyperalgesia to mechanical and thermal stimuli in mice under subacute and chronic inflammatory states [Citation107]. Therefore, whether the Nav1.9 channel is involved in the pathogenesis of IBS deserves further study.

Conclusion

Nav1.9 channels play a role in intestinal hyperpathia and dysmotility. The gain-of-function alterations in Nav1.9 channels cause the hyperexcitability of small diameter neurons in the DRG and Dogiel II neurons in the ENS. The hyperexcitability of small -diameter DRG neurons projecting to the intestine leads to intestinal hyperpathia. In addition, the increasing excitability of Dogiel II neurons in the ENS might affect intestinal peristaltic reflexes and entero-enteric inhibitory reflexes, which could cause intestinal dysmotility. VH and gastrointestinal dysmotility are vital pathophysiological mechanisms of IBS. Therefore, functional changes in Nav1.9 channels may be involved in the pathogenesis of IBS, but further studies are needed.

Author contributions

Conceptualization, Chenyu Zhao; software, Chenyu Zhao; writing – original draft preparation, Chenyu Zhao; writing – review and editing, Xi Zhou and Xiaoliu Shi. All authors have read and agreed to the published version of the manuscript.

Disclosure statement

No potential conflict of interest was reported by the authors.

Data availability statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Additional information

Funding

The author(s) reported there is no funding associated with the work featured in this article.

References

  • Bennett DL, Clark AJ, Huang J, et al. The role of voltage-gated sodium channels in pain signaling. Physiol Rev. 2019 Apr 1;99(2):1079–15.
  • Hockley JR, Winchester WJ, Bulmer DC. The voltage-gated sodium channel NaV 1.9 in visceral pain. Neurogastroenterol Motil. 2016 Mar;28(3):316–326.
  • Coates MD, Vrana KE, Ruiz-Velasco V. The influence of voltage-gated sodium channels on human gastrointestinal nociception. Neurogastroenterol Motil. 2019 Feb;31(2):e13460.
  • Osorio N, Korogod S, Delmas P. Specialized functions of Nav1.5 and Nav1.9 channels in electrogenesis of myenteric neurons in intact mouse ganglia. J Neurosci. 2014 Apr 9;34(15):5233–5244.
  • Dib-Hajj SD, Black JA, Waxman SG. NaV1.9: a sodium channel linked to human pain. Nat Rev Neurosci. 2015 Sep;16(9):511–519.
  • Yang L, Li L, Tang H, et al. Alcohol-aggravated episodic pain in humans with SCN11A mutation and ALDH2 polymorphism. Pain. 2020 Jul;161(7):1470–1482.
  • Bennett DL, Woods CG. Painful and painless channelopathies. Lancet Neurol. 2014 Jun;13(6):587–599.
  • Kabata R, Okuda H, Noguchi A, et al. Familial episodic limb pain in kindreds with novel Nav1.9 mutations. PLoS ONE. 2018;13(12):e0208516. DOI:10.1371/journal.pone.0208516
  • Huang J, Estacion M, Zhao P, et al. A novel gain-of-function Nav1.9 mutation in a child with episodic pain. Front Neurosci. 2019;13:918.
  • King MK, Leipold E, Goehringer JM, et al. Pain insensitivity: distal S6-segment mutations in Na(V)1.9 emerge as critical hotspot. Neurogenetics. 2017 Jul;18(3):179–181.
  • Leipold E, Liebmann L, Korenke GC, et al. A de novo gain-of-function mutation in SCN11A causes loss of pain perception. Nature Genet. 2013 Nov;45(11):1399–1404.
  • Copel C, Clerc N, Osorio N, et al. The Nav1.9 channel regulates colonic motility in mice. Front Neurosci. 2013;7:58.
  • Saha L. Irritable bowel syndrome: pathogenesis, diagnosis, treatment, and evidence-based medicine. World J Gastroenterol. 2014 Jun 14;20(22):6759–6773.
  • Boeckxstaens G, Camilleri M, Sifrim D, et al. Fundamentals of neurogastroenterology: physiology/motility – sensation. Gastroenterology. 2016 Feb 18;150(6):1292–1304.e2.
  • Catterall WA. Forty years of sodium channels: structure, function, pharmacology, and epilepsy. Neurochem Res. 2017 Sep;42(9):2495–2504.
  • Catterall WA, Goldin AL, Waxman SG. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol Rev. 2005 Dec;57(4):397–409.
  • Catterall WA, Lenaeus MJ, Gamal El-Din TM. Structure and pharmacology of voltage-gated sodium and calcium channels. Annual review of pharmacology and toxicology. Annu Rev Pharmacol Toxicol. 2020 Jan 6;60(1):133–154.
  • Zeisel A, Hochgerner H, Lönnerberg P, et al. Molecular architecture of the mouse nervous system. Cell. 2018 Aug 9;174(4):999–1014.e22.
  • González-Cano R, Ruiz-Cantero MC, Santos-Caballero M, et al. Tetrodotoxin, a potential drug for neuropathic and cancer pain relief? Toxins (Basel). 2021 Jul 12;13(7). doi :10.3390/toxins13070483.
  • Cummins TR, Dib-Hajj SD, Black JA, et al. A novel persistent tetrodotoxin-resistant sodium current in SNS-null and wild-type small primary sensory neurons. J Neurosci. 1999 Dec 15;19(24):Rc43.
  • Herzog RI, Cummins TR, Waxman SG. Persistent TTX-resistant Na+ current affects resting potential and response to depolarization in simulated spinal sensory neurons. J Neurophysiol. 2001 Sep;86(3):1351–1364.
  • Baker MD, Chandra SY, Ding Y, et al. GTP-induced tetrodotoxin-resistant Na+ current regulates excitability in mouse and rat small diameter sensory neurones. Journal of Physiology. 2003 Apr 15;548(Pt 2):373–382.
  • Wu Y, Ma H, Zhang F, et al. Selective voltage-gated sodium channel peptide toxins from animal venom: pharmacological probes and analgesic drug development. ACS Chem Neurosci. 2018 Feb 21;9(2):187–197.
  • Lin Z, Santos S, Padilla K, et al. Biophysical and pharmacological characterization of Nav1.9 voltage dependent sodium channels stably expressed in HEK-293 cells. PLoS ONE. 2016;11(8):e0161450. DOI:10.1371/journal.pone.0161450
  • Vanoye CG, Kunic JD, Ehring GR, et al. Mechanism of sodium channel NaV1.9 potentiation by G-protein signaling. J General Physiol. 2013 Feb;141(2):193–202.
  • Peng S, Chen M, Xiao Z, et al. A novel spider toxin inhibits fast inactivation of the Na(v)1.9 channel by binding to domain iii and domain iv voltage sensors. Front Pharmacol. 2021;12:778534.
  • Zhou X, Ma T, Yang L, et al. Spider venom-derived peptide induces hyperalgesia in Na(v)1.7 knockout mice by activating Na(v)1.9 channels. Nat Commun. 2020 May 8;11(1):2293.
  • Wang G, Long C, Liu W, et al. Novel sodium channel inhibitor from leeches. Front Pharmacol. 2018;9:186.
  • McMahon KL, Tay B, Deuis JR, et al. Pharmacological activity and NMR solution structure of the leech peptide HSTX-I. Biochem Pharmacol. 2020 Nov;181:114082.
  • Berta T, Qadri Y, Tan PH, et al. Targeting dorsal root ganglia and primary sensory neurons for the treatment of chronic pain. Expert Opin Ther Targets. 2017 Jul;21(7):695–703.
  • Dib-Hajj SD, Tyrrell L, Cummins TR, et al. Two tetrodotoxin-resistant sodium channels in human dorsal root ganglion neurons. FEBS Lett. 1999 Nov 26;462(1–2):117–120.
  • Benn SC, Costigan M, Tate S, et al. Developmental expression of the TTX-resistant voltage-gated sodium channels Nav1.8 (SNS) and Nav1.9 (SNS2) in primary sensory neurons. J Neurosci. 2001 Aug 15;21(16):6077–6085.
  • Fang X, Djouhri L, Black JA, et al. The presence and role of the tetrodotoxin-resistant sodium channel Na(v)1.9 (NaN) in nociceptive primary afferent neurons. J Neurosci. 2002 Sep 1;22(17):7425–7433.
  • Dib-Hajj SD, Tyrrell L, Escayg A, et al. Coding sequence, genomic organization, and conserved chromosomal localization of the mouse gene Scn11a encoding the sodium channel NaN. Genomics. 1999 Aug 1;59(3):309–318.
  • Dib-Hajj S, Black JA, Cummins TR, et al. NaN/Nav1.9: a sodium channel with unique properties. Trends Neurosci. 2002 May;25(5):253–259.
  • Priest BT, Murphy BA, Lindia JA, et al. Contribution of the tetrodotoxin-resistant voltage-gated sodium channel NaV1.9 to sensory transmission and nociceptive behavior. Proc Natl Acad Sci USA. 2005 Jun 28;102(26):9382–9387.
  • Maingret F, Coste B, Padilla F, et al. Inflammatory mediators increase Nav1.9 current and excitability in nociceptors through a coincident detection mechanism. J General Physiol. 2008 Mar;131(3):211–225.
  • Amsalem M, Poilbout C, Ferracci G, et al. Membrane cholesterol depletion as a trigger of Nav1.9 channel-mediated inflammatory pain. Embo J. 2018 Apr 13;37(8). doi :10.15252/embj.201797349.
  • Yu YQ, Zhao F, Guan SM, et al. Antisense-mediated knockdown of Na(V)1.8, but not Na(V)1.9, generates inhibitory effects on complete Freund’s adjuvant-induced inflammatory pain in rat. PLoS ONE. 2011 May 10;6(5):e19865.
  • Fertleman CR, Ferrie CD, Aicardi J, et al. Paroxysmal extreme pain disorder (previously familial rectal pain syndrome). Neurology. 2007 Aug 7;69(6):586–595.
  • Okuda H, Noguchi A, Kobayashi H, et al. Infantile Pain episodes associated with novel Nav1.9 mutations in familial episodic pain syndrome in japanese families. PLoS ONE. 2016;11(5):e0154827. DOI:10.1371/journal.pone.0154827
  • Zhang XY, Wen J, Yang W, et al. Gain-of-function mutations in SCN11A cause familial episodic pain. Am J Hum Genet. 2013 Nov 7;93(5):957–966.
  • Ginanneschi F, Rubegni A, Moro F, et al. SCN11A variant as possible pain generator in sensory axonal neuropathy. Neurol Sci. 2019 Jun;40(6):1295–1297.
  • Sambuughin N, Ren M, Capacchione JF, et al. Multifactorial origin of exertional rhabdomyolysis, recurrent hematuria, and episodic pain in a service member with sickle cell trait. Case Rep Genet. 2018;2018:6898546.
  • Huang J, Han C, Estacion M, et al. Gain-of-function mutations in sodium channel Na(v)1.9 in painful neuropathy. Brain. 2014 Jun;137(Pt 6):1627–1642.
  • Han C, Yang Y, de Greef BT, et al. The domain II S4-S5 linker in Nav1.9: a missense mutation enhances activation, impairs fast inactivation, and produces human painful neuropathy. Neuromolecular Med. 2015 Jun;17(2):158–169.
  • Kleggetveit IP, Schmidt R, Namer B, et al. Pathological nociceptors in two patients with erythromelalgia-like symptoms and rare genetic Nav 1.9 variants. Brain Behav. 2016 Oct;6(10):e00528.
  • Eijkenboom I, Sopacua M, Hoeijmakers JGJ, et al. Yield of peripheral sodium channels gene screening in pure small fibre neuropathy. J Neurol Neurosurg Psychiatry. 2019 Mar;90(3):342–352.
  • Wadhawan S, Pant S, Golhar R, et al. Na(v) channel variants in patients with painful and nonpainful peripheral neuropathy. Neurol Genet. 2017 Dec;3(6):e207.
  • Zhou L. Small fiber neuropathy. Semin Neurol. 2019 Oct;39(5):570–577.
  • Cazzato D, Lauria G. Small fibre neuropathy. Curr Opin Neurol. 2017 Oct;30(5):490–499.
  • Schon KR, Parker APJ, Woods CG. Congenital Insensitivity to Pain Overview. In: Adam M, Ardinger H, and Pagon R, et al., editors. GeneReviews(®). Seattle (WA): University of Washington, Seattle, Copyright © 1993-2023.2018; NBK481553 .
  • Poojary S, Jaiswal S, Shah KS, et al. Sisters with no pain, no tears: a report of a new variant of hereditary sensory and autonomic neuropathy (Type IX) caused by a novel SCN11A mutation. Indian J Dermatol. 2020 Jul-Aug;65(4):299–303.
  • Phatarakijnirund V, Mumm S, McAlister WH, et al. Congenital insensitivity to pain: fracturing without apparent skeletal pathobiology caused by an autosomal dominant, second mutation in SCN11A encoding voltage-gated sodium channel 1.9. Bone. 2016 Mar;84:289–298.
  • Huang J, Vanoye CG, Cutts A, et al. Sodium channel NaV1.9 mutations associated with insensitivity to pain dampen neuronal excitability. J Clin Investig. 2017 Jun 30;127(7):2805–2814.
  • Hockley JR, Boundouki G, Cibert-Goton V, et al. Multiple roles for NaV1.9 in the activation of visceral afferents by noxious inflammatory, mechanical, and human disease-derived stimuli. Pain. 2014 Oct;155(10):1962–1975.
  • Martinez V, Melgar S. Lack of colonic-inflammation-induced acute visceral hypersensitivity to colorectal distension in Na(v)1.9 knockout mice. Eur J Pain. 2008 Oct;12(7):934–944.
  • Vanner S, Greenwood-Van Meerveld B, Mawe G, et al. Fundamentals of Neurogastroenterology: basic Science. Gastroenterology. 2016 Feb 18;150(6):1280–1291.
  • Zhao C, Jin J, Hu H, et al. The gain-of-function R222S variant in Scn11a contributes to visceral hyperalgesia and intestinal dysmotility in Scn11aR222S/R222S Mice. Front Neurol. 2022;13:856459.
  • Spencer NJ, Hu H. Enteric nervous system: sensory transduction, neural circuits and gastrointestinal motility. Nat Rev Gastroenterol Hepatol. 2020 Jun;17(6):338–351.
  • Spencer NJ, Kyloh MA, Travis L, et al. Sensory nerve endings arising from single spinal afferent neurons that innervate both circular muscle and myenteric ganglia in mouse colon: colon-brain axis. Cell Tissue Res. 2020 Jul;381(1):25–34.
  • Herrity AN, Rau KK, Petruska JC, et al. Identification of bladder and colon afferents in the nodose ganglia of male rats. J Comp Neurol. 2014 Nov 1;522(16):3667–3682.
  • Erickson A, Deiteren A, Harrington AM, et al. Voltage-gated sodium channels: (Na(v))igating the field to determine their contribution to visceral nociception. Journal of Physiology. 2018 Mar 1;596(5):785–807.
  • Brierley SM, Hughes PA, Harrington AM, et al. Identifying the ion channels responsible for signaling gastro-intestinal based pain. pharmaceuticals. (Basel, Switzerland). 2010Aug26; 3(9): 2768–2798. 10.3390/ph3092768
  • Beyak MJ, Ramji N, Krol KM, et al. Two TTX-resistant Na+ currents in mouse colonic dorsal root ganglia neurons and their role in colitis-induced hyperexcitability. Am J Physiol Gastrointest Liver Physiol. 2004 Oct;287(4):G845–55.
  • Usoskin D, Furlan A, Islam S, et al. Unbiased classification of sensory neuron types by large-scale single-cell RNA sequencing. Nat Neurosci. 2015 Jan;18(1):145–153.
  • Li JH, Duan R, Li L, et al. Unique characteristics of “the second brain” - the enteric nervous system. Sheng Li Xue Bao. 2020 Jun 25;72(3):382–390.
  • Fung C, Vanden Berghe P. Functional circuits and signal processing in the enteric nervous system. Cellular and molecular life sciences: cMLS. Nov. 2020;77(22):4505–4522. DOI:10.1007/s00018-020-03543-6
  • Costa M, Furness JB. The peristaltic reflex: an analysis of the nerve pathways and their pharmacology. Naunyn-Schmiedeberg’s Arch Pharmacol. 1976 Jul;294(1):47–60.
  • Brehmer A, Schrodl F, Neuhuber W. Morphological classifications of enteric neurons–100 years after Dogiel. Anatomy and Embryology. 1999 Aug;200(2):125–135.
  • Costa M, Spencer NJ, Brookes SJH. The role of enteric inhibitory neurons in intestinal motility. Autonomic neuroscience: basic & clinical. Auton Neurosci. 2021 Nov;235:102854.
  • Hirst GD, Holman ME, Spence I. Two types of neurones in the myenteric plexus of duodenum in the guinea-pig. Journal of Physiology. 1974 Jan;236(2):303–326.
  • Bertrand PP, Kunze WA, Bornstein JC, et al. Electrical mapping of the projections of intrinsic primary afferent neurones to the mucosa of the guinea-pig small intestine. Neurogastroenterol Motil. 1998 Dec;10(6):533–541.
  • Bertrand PP, Thomas EA. Multiple levels of sensory integration in the intrinsic sensory neurons of the enteric nervous system. Clin Exp Pharmacol. 2004 Nov;31(11):745–755.
  • Spencer NJ, Smith TK. Mechanosensory S-neurons rather than AH-neurons appear to generate a rhythmic motor pattern in guinea-pig distal colon. Journal of Physiology. 2004 Jul 15;558(Pt 2):577–596.
  • Mazzuoli G, Schemann M. Multifunctional rapidly adapting mechanosensitive enteric neurons (RAMEN) in the myenteric plexus of the guinea pig ileum. Journal of Physiology. 2009 Oct 1;587(Pt 19):4681–4694.
  • O’Donnell AM, Coyle D, Puri P. Decreased Nav1.9 channel expression in Hirschsprung’s disease. J Pediatr Surg. 2016 Sep;51(9):1458–1461.
  • Copel C, Osorio N, Crest M, et al. Activation of neurokinin 3 receptor increases Na(v)1.9 current in enteric neurons. Journal of Physiology. 2009 Apr 1;587(Pt 7):1461–1479.
  • Gwynne RM, Bornstein JC. Mechanisms underlying nutrient-induced segmentation in isolated guinea pig small intestine. Am J Physiol Gastrointest Liver Physiol. 2007 Apr;292(4):G1162–72.
  • Gwynne RM, Thomas EA, Goh SM, et al. Segmentation induced by intraluminal fatty acid in isolated guinea-pig duodenum and jejunum. Journal of Physiology. 2004 Apr 15;556(Pt 2):557–569.
  • Costa M, Dodds KN, Wiklendt L, et al. Neurogenic and myogenic motor activity in the colon of the guinea pig, mouse, rabbit, and rat. Am J Physiol Gastrointest Liver Physiol. 2013 Nov 15;305(10):G749–59.
  • Spencer NJ, Nicholas SJ, Robinson L, et al. Mechanisms underlying distension-evoked peristalsis in guinea pig distal colon: is there a role for enterochromaffin cells? Am J Physiol Gastrointest Liver Physiol. 2011 Sep;301(3):G519–27.
  • Spencer NJ, Dinning PG, Brookes SJ, et al. Insights into the mechanisms underlying colonic motor patterns. Journal of Physiology. 2016 Aug 1;594(15):4099–4116.
  • Bornstein JC, Costa M, Grider JR. Enteric motor and interneuronal circuits controlling motility. Neurogastroenterol Motil. 2004 Apr;16(Suppl 1):34–38.
  • Hansen MB. The enteric nervous system II: gastrointestinal functions. Pharmacol Toxicol. 2003 Jun;92(6):249–257.
  • Wehrwein EA, Orer HS, Barman SM. Overview of the anatomy, physiology, and pharmacology of the autonomic nervous system. Compr Physiol. 2016 Jun 13;6(3):1239–1278.
  • Furness JB. Novel gut afferents: intrinsic afferent neurons and intestinofugal neurons. Autonomic neuroscience: basic & clinical. 2006 Apr 30;125(1–2):81–85. DOI: 10.1016/j.autneu.2006.01.007
  • Linden DR. Enhanced excitability of guinea pig inferior mesenteric ganglion neurons during and following recovery from chemical colitis. Am J Physiol Gastrointest Liver Physiol. 2012 Nov 1;303(9):G1067–75.
  • Linden DR. Colitis is associated with a loss of intestinofugal neurons. Am J Physiol Gastrointest Liver Physiol. 2012 Nov 15;303(10):G1096–104.
  • Furness JB. Intestinofugal neurons and sympathetic reflexes that bypass the central nervous system. J Comp Neurol. 2003 Jan 13;455(3):281–284.
  • Bywater RA. Activity following colonic distension in enteric sensory fibres projecting to the inferior mesenteric ganglion in the guinea pig. J Auton Nerv Syst. 1994 Jan-Feb;46(1–2):19–26.
  • Miller SM, Szurszewski JH. Colonic mechanosensory afferent input to neurons in the mouse superior mesenteric ganglion. Am J Physiol. 1997 Feb;272(2 Pt 1):G357–66.
  • Kreulen DL, Szurszewski JH. Reflex pathways in the abdominal prevertebral ganglia: evidence for a colo-colonic inhibitory reflex. Journal of Physiology. 1979 Oct;295:21–32.
  • Hetz S, Acikgoez A, Moll C, et al. Age-related gene expression analysis in enteric ganglia of human colon after laser microdissection. Front Aging Neurosci. 2014;6:276.
  • Ebbinghaus M, Tuchscherr L, Segond von Banchet G, et al. Gain-of-function mutation in SCN11A causes itch and affects neurogenic inflammation and muscle function in Scn11a+/L799P mice. PLoS ONE. 2020;15(8):e0237101. DOI:10.1371/journal.pone.0237101
  • Mawe GM. Colitis-induced neuroplasticity disrupts motility in the inflamed and post-inflamed colon. J Clin Investig. 2015 Mar 2;125(3):949–955.
  • Farzaei MH, Bahramsoltani R, Abdollahi M, et al. The role of visceral hypersensitivity in irritable bowel syndrome: pharmacological targets and novel treatments. J Neurogastroenterol Motil. 2016 Oct 30;22(4):558–574.
  • Mearin F, Lacy BE, Chang L, et al. Bowel Disorders. Gastroenterology. 2016 Feb 18;150:1393-1407.
  • Gwee KA, Lu CL, Ghoshal UC. Epidemiology of irritable bowel syndrome in Asia: something old, something new, something borrowed. J Gastroenterol Hepatol. 2009 Oct;24(10):1601–1607.
  • Liu L, Liu BN, Chen S, et al. Visceral and somatic hypersensitivity, autonomic cardiovascular dysfunction and low-grade inflammation in a subset of irritable bowel syndrome patients. J Zhejiang Univ Sci B. 2014 Oct;15(10):907–914.
  • DuPont AW, Jiang ZD, Harold SA, et al. Motility abnormalities in irritable bowel syndrome. Digestion. 2014;89(2):119–123. DOI:10.1159/000356314
  • Park JH, Baek YH, Park DI, et al. Analysis of rectal dynamic and static compliances in patients with irritable bowel syndrome. Int J Colorectal Dis. 2008 Jul;23(7):659–664.
  • Fassov J, Lundby L, Worsøe J, et al. A randomised, controlled study of small intestinal motility in patients treated with sacral nerve stimulation for irritable bowel syndrome. BMC Gastroenterol. 2014 Jun 25;14:111.
  • Berumen A, Edwinson AL, Grover M. Post-infection Irritable Bowel Syndrome. Gastroenterol Clin North Am. 2021 Jun;50(2):445–461.
  • Barbara G, Grover M, Bercik P, et al. Rome foundation working team report on post-infection irritable bowel syndrome. Gastroenterology. 2019 Jan;156(1):46–58.e7.
  • Sinagra E, Pompei G, Tomasello G, et al. Inflammation in irritable bowel syndrome: myth or new treatment target? World J Gastroenterol. 2016 Feb 21;22(7):2242–2255.
  • Lolignier S, Amsalem M, Maingret F, et al. Nav1.9 channel contributes to mechanical and heat pain hypersensitivity induced by subacute and chronic inflammation. PLoS ONE. 2011;6(8):e23083. DOI:10.1371/journal.pone.0023083