It is the cache of ${baseHref}. It is a snapshot of the page. The current page could have changed in the meantime.
Tip: To quickly find your search term on this page, press Ctrl+F or ⌘-F (Mac) and use the find bar.

World Journal of Gastroenterology-Baishideng Publishing
World J Gastroenterol. 2011 May 21; 17 (19) : 2357-2364.
Published online 2011 May 21. doi: 10.3748/wjg.v17.i19.2357.
Targeting voltage-gated sodium channels for treatment for chronic visceral pain
Fei-Hu Qi, You-Lang Zhou and Guang-Yin Xu.
Fei-Hu Qi, You-Lang Zhou, Guang-Yin Xu, Jiangsu Key Laboratory of Neuroregeneration, Nantong University, Nantong 226001, Jiangsu Province, China
Guang-Yin Xu, Institute of Neuroscience, Key laboratory of Pain Research and Therapy, Soochow University, Suzhou 215123, Jiangsu Province, China
Guang-Yin Xu, Division of Gastroenterology, Department of Internal Medicine, University of Texas Medical Branch, Galveston, TX 77555-0655, United States
Author contributions: Qi FH and Zhou YL performed the experiments and wrote the paper; Xu GY edited the paper.
Correspondence to: Guang-Yin Xu, MD, PhD, Division of Gastroenterology and Hepatology, Department of Internal Medicine, University of Texas Medical Branch, Galveston, TX 77555-0655, United States. guangyin.xu@gmail.com
Telephone: +1-409-7473044  Fax: +1-409-7473084
Received December 14, 2010; Revised March 16, 2011; Accepted March 23, 2011;
Abstract
Voltage-gated sodium channels (VGSCs) play a fundamental role in controlling cellular excitability, and their abnormal activity is related to several pathological processes, including cardiac arrhythmias, epilepsy, neurodegenerative diseases, spasticity and chronic pain. In particular, chronic visceral pain, the central symptom of functional gastrointestinal disorders such as irritable bowel syndrome, is a serious clinical problem that affects a high percentage of the world population. In spite of intense research efforts and after the dedicated decade of pain control and research, there are not many options to treat chronic pain conditions. However, there is a wealth of evidence emerging to give hope that a more refined approach may be achievable. By using electronic databases, available data on structural and functional properties of VGSCs in chronic pain, particularly functional gastrointestinal hypersensitivity, were reviewed. We summarize the involvement and molecular bases of action of VGSCs in the pathophysiology of several organic and functional gastrointestinal disorders. We also describe the efficacy of VGSC blockers in the treatment of these neurological diseases, and outline future developments that may extend the therapeutic use of compounds that target VGSCs. Overall, clinical and experimental data indicate that isoform-specific blockers of these channels or targeting of their modulators may provide effective and novel approaches for visceral pain therapy.
Keywords: Voltage-gated sodium channel, Dorsal root ganglion, Visceral pain, Functional gastrointestinal disorders, Treatment
INTRODUCTION
Treating chronic pain remains a significant clinical challenge, and in particular, current treatments for visceral pain conditions are still based mainly on empirical rather than mechanistic evidence. Search for a consistently effective pharmacological agent to treat these pain conditions remain elusive. Functional gastrointestinal (GI) pain, one of the most prevalent forms of chronic visceral pain, is commonly interpreted as a consequence of hypersensitivity of GI nociceptive pathways, either of the peripheral sensory nociceptors in the gut or of the central neurons[ 1 , 2]. Human and animal studies have identified several channels as pivotal for enhanced signal transmission along the pain axis, including voltage-gated sodium channels (VGSCs) Na v1.3, Na v1.7, Na v1.8, and Na v1.9, with the latter three preferentially expressed in peripheral sensory neurons[ 3 - 5]. VGSCs are integral membrane glycoproteins that are essential for generation and conduction of electrical impulses in excitable cells. Many of the most common neurological disorders, such as epilepsy, migraine, neurodegenerative diseases, and chronic pain involve abnormalities of neuronal excitability[ 6 , 7]. There is a growing body of evidence that implicates abnormal expression and function of VGSCs in these disorders[ 8 - 12]. Although there is an incomplete understanding of these channels and/or primary site of action at which these compounds exert their effects and significant side effects are often encountered, there is enough evidence emerging to give hope that a more refined approach may be achievable[ 13]. This article does not aim to provide an exhaustive list of potential targets, but to focus on some of the most recent advances in the field of pain research and medicine and to highlight where therapeutic strategies for relieving visceral pain may emerge in the near future.
BIOPHYSICAL AND MOLECULAR PROPERTIES OF VGSCS
VGSCs are integral membrane glycoproteins that consist of a central α-subunit of 260 kDa associated with one or more auxiliary β-subunits of about 35 kDa (Figure 1A)[ 5 , 14]. Nine α-subunit subtypes have been cloned and functionally expressed. These subtypes are designated Na v1.1-Na v1.9 for the proteins and SCN1A-SCN5A and SCN8A-SCN11A for the genes; SCN6A/SCN7A codifies the related protein Na x. The VGSC α-subunit, which forms the ion-conducting pore and the channel gate for activation and inactivation, contains four domains (D I-D IV), each with six α-helical transmembrane segments labeled S1-S6 (Figure 1A). Pore loops between S5 and S6 (highlighted in red) in each of the four domains form the selectivity filter of the channel. Each pore loop contributes a single amino acid (aspartate from D I, glutamate from D II, lysine from D III, and alanine from D IV), which together form a narrow ring that is mainly responsible for conferring Na + selectivity. The four S6 segments form the cytoplasmic end of the pore, which binds various types of therapeutically important pore-blocking compounds, including local anesthetics, and antiarrhythmic drugs. The S4 segments (highlighted in green) in each of the four domains contain regularly spaced, positively charged amino acid residues and serve as voltage-sensors, coupling membrane depolarization to channel activation. The intracellular loop between D III and D IV forms the fast-inactivation gate that occludes the cytoplasmic end of the pore when the channel inactivates. The C-terminal cytoplasmic domain is important for setting some of the properties of fast inactivation and contains binding sites for interacting proteins. In addition to fast inactivation, a distinct process called slow inactivation develops during prolonged depolarizing plateaus and during high frequency repetitive firing. The kinetics of onset and recovery of slow inactivation are about four orders of magnitude slower than those of fast inactivation. Slow inactivation does not depend on the fast inactivation gate formed by the intracellular loop between D III and D IV, but instead mainly involves rearrangements of the pore of the channel[ 15].
Figure 1
Figure 1
Figure 1
Schematic of a voltage-gated sodium channel polypeptide and Nav1.8 positive colon specific dorsal root ganglion neurons. A: The transmembrane organization includes four domains (D I-D IV) joined by three intracellular loops (L1-L3) and the intracellular N and C termini. Cylinders represent probable α-helical segments. P in circles, sites of demonstrated protein phosphorylation by protein kinase A (PKA) or PKC; green, S4 voltage sensors; h in which square, inactivation particle in the inactivation gate loop (modified, with permission, from Catterall[14]); B: Colon specific dorsal root ganglion neurons labeled by Dil (red); C: Nav1.8 positive neurons (green). Colon specific Nav1.8 positive neurons are stained yellow. Bar = 50 μm for both B and C.
VGSCS IN PRIMARY SENSORY NEURONS INNERVATING VISCERAL ORGANS
Primary sensory neurons, with their cell bodies located in the dorsal root ganglion (DRG), are the first link in the somatosensory pathway. They encode their messages in the form of a series of action potentials whose depolarizing upstroke is produced by sodium channels[ 3 - 5 , 16]. Most neuroscientists and neurologists are familiar with the textbook description of VGSC function (Figure 2). VGSCs contribute to the control of membrane excitability and underlie action potential generation. They are closed at resting membrane potentials characteristic of quiescent neurons. In response to membrane depolarization, they open within a few hundred microseconds (a process termed activation), resulting in an inward sodium ion (Na +) current, and then convert within a few milliseconds to a non-conducting inactivated state through a process called fast inactivation. Transient Na + influx through thousands of rapidly opening and inactivating VGSCs results in the familiar transient macroscopic Na + current detected in whole-cell voltage clamp studies[ 16]. This transient current gives rise to the depolarizing phase of the action potential in neurons, and eventually causes the sensation of pain (Figure 2).
Figure 2
Figure 2
Figure 2
A model showing the three states of the sodium channel. Tissue damage/inflammation leads to activation of the sodium channel from a closed state. Once activated, sodium ions entered through the channel and evoke the action potential of the cell membrane, eventually producing the sensation of pain. Once the channel becomes inactivated, it must go to a closed state before being activated again. Action potentials and inward currents were recorded from a colon specific dorsal root ganglion neuron from a healthy rat.
Distinct distribution
Expression of the different subtypes of VGSCs is tissue- and/or cell-specific. The main subtypes expressed in adult brain neurons are Na v1.6. Na v1.2 in unmyelinated axons and in myelinated axons early in development (before being replaced by Na v1.6), and Na v1.1 in the neuronal somata. In rodents, Na v1.3 is expressed primarily in embryonic neurons; however, expression in the human central nervous system remains comparatively high into adulthood. Na v1.4 is the main subtype in adult skeletal muscle, whereas Na v1.5 is expressed in cardiac muscle, and it is also expressed in some neurons. In DRG neurons, Na v1.6 is the major sodium channel isoform at the nodes of Ranvier[ 17], is present in unmyelinated fibers in the sciatic nerve[ 18 , 19] and is preferentially expressed by TrkC (tyrosine receptor kinase C) neurons. Na v1.7, Na v1.8, and Na v1.9 are found in peripheral primary sensory afferents. They are more abundant in C-fiber neurons[ 16 , 20]. Small diameter DRG neurons can be broadly classified into isolectin B4 (IB4)(+) and IB4(-) neurons. James Brock had reviewed sensory and biophysical properties differences between IB4(+) and IB4(-) neurons[ 21]. Our preliminary experiments showed that most colon specific DRG neurons are Na v1.8 positive (Figure 1B and C).
Developmentally regulated expression
Expression of the different subtypes is developmentally regulated in DRG neurons[ 22 - 24]. Na v1.3 is overexpressed at embryonic day E1, peaks at E17, is reduced by postnatal day P15, and is not detectable by P30. Na v1.6 is upregulated from E17 only in large neurons. Expression of both Na v1.8 and Na v1.9 increases with age, beginning at E15 and E17, respectively, and reaching adult levels by P7. Their distribution is restricted mainly to those subpopulations of primary sensory neurons in developing and adult DRGs that give rise to unmyelinated C-fibers (neurofilament 200 negative). Na v1.8 is expressed in a higher proportion of neuronal profiles than Na v1.9 at all stages during development, as in the adult[ 24]. It is worthy of note that nerve injury would evoke expression of some of these channels that are not present in the sensory nerve system in the adult, and may contribute to abnormal pain sensation[ 20 , 25 , 26].
Expression in glial cells
Na-G, a distinct type of “glial” Na channel expressed at high levels in Schwann cells in vivo[ 27 , 28], was thought to play a role in regulation of cytoplasmic Na + homoeostasis. However, recent studies show that action potential-like events have been recorded in glial precursor cells, astrocytes and a subset of oligodendrocyte precursor cells. More recently, Frieboes reported de novo expression of Na v1.8 in endoneurial Schwann cells following injury[ 10]. This is surprising since Na v1.8 expression is thought to be restricted to sensory neurons. No similar studies have been reported under pathophysiological conditions in visceral organs. Nevertheless, these data indicate a sodium channel in glial cells is linked to the development of neuropathic pain. Further investigation of the underlying molecular basis of VGSCs in glial cells and neural-glial interaction could yield promising targets for treatment of chronic pain.
VGSCS AND CHRONIC VISCERAL PAIN
There are relatively few published studies on roles of sodium channels in visceral pain and a uniform theme has not emerged. Saito et al[ 29] demonstrated a mutation in the cardiac sodium channel gene SCN5A in interstitial cells of Cajal (ICC) in irritable bowel syndrome (IBS) patients, suggesting SCN5A as a candidate gene in the pathophysiology of IBS. Additionally, Verne et al[ 30] showed there was a reduction in abdominal pain in IBS patients receiving intrarectal lidocaine, indicating a possible normalization of abnormal sodium channels in IBS patients. Yiangou et al[ 31] carried out an immunohistochemistry study on biopsies from patients with idiopathic rectal hypersensitivity which showed that the nerve fibers immunoreactive to the sodium channel Na v1.7 were significantly increased in this group compared with controls. Collectively, these clinical data suggest that voltage-gated sodium channels appear to play an important role in visceral pain[ 5 , 9].
Since no pharmacological tools to separate sodium currents are available, it is hard to relate the channel current to its molecular subtype. However, tetrodotoxin (TTX) is the most used drug to separate the TTX-S (sensitive) and TTX-R (resistant) sodium channel currents. Whole-cell patch-clamp recording of fast blue-labeled bladder afferent neurons demonstrated that the majority (70%) of bladder neurons that were small in size had capsaicin-sensitive TTX-R sodium channels[ 32 , 33]. More than 80% of gastric DRG neurons displayed TTX-R action potentials[ 34]. These data demonstrated that visceral sensory neurons expressed the TTX-R sodium channels that are thought to contribute to pain processing.
Sensitization of VGSCs
Several animal models have been used to determine the role for VGSCs under visceral pain conditions. These included the trinitrobenzenesulphonic acid (TNBS)-induced colitis, painful chronic pancreatitis, and acetic acid-induced gastric ulcers. TNBS-induced colitis significantly enhanced neuronal excitability of colon specific DRG neurons, which is associated with a significant potentiation of slow inactivating TTX-R Na + current density in mice[ 35] and in guinea pigs[ 36]. Changes in expression of Na v1.7, 1.8 and 1.9 protein and mRNA during TNBS colitis were measured using Western blotting and quantitative polymerase chain reaction. On day 7 of colitis, there was a 3-fold increase in Na v1.8 protein in ganglia from T 9-13, but Na v1.7 and 1.9 levels remained unchanged. Surprisingly, there was no corresponding change in the Na v1.8 α-subunit mRNA levels[ 37], indicating that the TNBS specifically evoked an increase in the numbers of Na v1.8 channels, which contributed to an increased current. In a rat model of gastric hyperalgesia, Gebhart and colleagues reported that ingestion of iodoacetamide increased the peak TTX-R sodium current which was associated with a left-shifted voltage-dependent activation in DRG neurons[ 38]. In a rat model of painful chronic pancreatitis, however, no change in sodium peak current density was observed in pancreas-specific DRG neurons (Xu et al[ 39], unpublished data) although there was an alteration in voltage-gated potassium channel function in the same model. Although the data from different models are complicated and somehow controversial, they point out the fact that changes in function and expression of VGSCs are organ or/and disease specific.
As mentioned previously, no pharmacological tools to separate TTX-R currents are available. However, researchers have employed antisense oligodeoxynucleotide or Na v1.8 and Na v1.9 null mice to evaluate their current roles in animal models of visceral hypersensitivity. Inhibition of expression of Na v1.8 (PN3/SNS) sodium channels by an antisense oligodeoxynucleotide injected intrathecally suppressed bladder hyperactivity and c-fos expression in the spinal cord induced by chemical irritation of the urinary bladder in rats[ 32]. In Nippostrongylus brasiliensis infection-induced transient jejunitis, hyperexcitability was absent in Na v1.8 -/- mice[ 40]. Further experiments showed that Na v1.8-null mice displayed normal nociceptive behavior provoked by acute noxious stimulation of abdominal viscera. Na v1.8-null mice also showed weak pain and no referred hyperalgesia to intracolonic capsaicin and mustard oil[ 41]. In a model of cyclophosphamide-induced bladder cystitis, Na v1.8 null mice showed normal responses. These data suggest that Na v1.8 plays an essential role in mediating spontaneous activity in sensitized nociceptors. However, the role for Na v1.9 is inconsistent. Hillsley et al[ 40] reported that Na v1.9 –/– mice maintained hyperexcitability in Nippostrongylus brasiliensis infection-induced transient jejunitis. On the other hand, Martinez reported that R-848 (the toll-like receptor 7 activator)-induced hypersensitivity was blunt in Na v1.9-null mice although their normal pain responses were similar to those of wild-type mice[ 42]. Hence, they suggested that Na v1.9 channels do not significantly contribute to normal visceral pain responses to acute colonic mechanical stimulation but may be important for the development of inflammation-related visceral hyperalgesic responses. This idea was further supported by a recent report that peripheral nerve recordings from pelvic afferents in Na v1.9-null mice revealed a lack of sensitization to intravesicularly applied prostaglandin (PG) E2[ 43]. Again, this points out that the role for Na v1.9 may be organ and/or disease specific.
Modulation of VGSCs
There is growing evidence to support the hypothesis that inflammatory mediators are involved in the modulation of VGSCs function and expression under pathophysiological states (Figure 3). Among these are well-studied nerve growth factor (NGF) and PGE2. Blockade of NGF and tyrosine kinase receptors inhibits peripheral mechanical sensitivity accompanying cystitis in rats[ 44]. NGF continuously infused into the intrathecal space at the L6-S1 level of the spinal cord for 1 or 2 wk using osmotic pumps (0.5 μL/h) significantly lowered the threshold for spike activation. In addition, the number of TTX-R action potentials during 600 ms depolarizing pulses was significantly and time-dependently increased[ 33]. Another inflammatory messenger PGE2 induced a rapid (< 15 s) increase in TTX-R I Na that was associated with a hyperpolarizing shift in the conductance-voltage curve (3.4 ± 0.7 mV), an increase in the rate of inactivation (4.21 ± 0.7 ms at 0 mV), and no change in steady-state availability. These results do suggest that modulation of TTX-R I Na by NGF and PGE2 in colonic afferents is an underlying mechanism of hyperalgesia associated with inflammation of the colon, and that this current constitutes a novel target for therapeutic relief of visceral inflammatory pain[ 45 - 47]. Other modulators include, but are not limited to, aquaporins (AQP-1), interleukin (IL)-1β and gaseous modulators/transmitters. AQP-1 may directly interact with Na v1.8 channels in DRG neurons[ 48], while IL-1β may act via a p38 mitogen-activated protein kinase pathway[ 49]. Hydrogen sulfide (H 2S), as an endogenous novel third gaseous modulator/transmitter, might directly or indirectly modulate the function of VGSCs since it enhanced excitability of primary sensory neurons, and its endogenous generating enzyme was upregulated under visceral pain conditions[ 50].
Figure 3
Figure 3
Figure 3
Possible mechanisms underlying the modulation of voltage-gated sodium channel activity. Inflammatory mediators such as nerve growth factor (NGF), prostaglandin (PG) E2, interleukin (IL)-1β may sensitize sodium channels by different intracellular signal transduction pathways such as protein kinase A (PKA), protein kinase C (PKC), mitogen-activated protein kinase and calmodulin (CaM). In addition, a recent study showed that aquaporin (AQP)-1 directly interacted with the sodium channel, thus contributing to the perception of pain[48]. TTX: Tetrodotoxin; LIF: leukemia inhibitory factor; Trk: Tyrosine receptor kinase; TNF: Tumor necrosis factor; TRAF: Tumor necrosis factor receptor-associated factor; TRADD: TNF receptor-associated death domain protein.
The regulation of VGSCs by intracellular protein kinases is another important mechanism for maintenance of persistent behavioral hypernociception. Direct phosphorylation of Na v1.7 by pERK1/2 and modulation of Na v1.8 by p38 regulated the gating properties of these channels[ 49 , 51]. In addition, functional regulation of the Na v1.8 by PKA and PKCepsilon in the primary sensory neuron is important for the development of the peripheral pro-nociceptive state induced by repetitive inflammatory stimuli and for the maintenance of behavioral persistent hypernociception[ 52]. Since the Na v1.8 C-terminus carries a conserved calmodulin-binding isoleucine-glutamine motif, it is reasonable to speculate that these two proteins can interact in vivo. Indeed, calmodulin demonstrated a regulation of Na v1.8 currents[ 53]. Unfortunately, our understanding of regulation of VGSCs in visceral pain conditions lags behind our knowledge of mechanisms of somatic pain. Future experiments are warranted to determine whether and how Na v1.8 channels are modulated by inflammatory and stress mediators in the peripheral nerve system.
MANAGEMENT OF VISCERAL PAIN WITH VGSC BLOCKERS
Treating chronic pain remains a significant clinical challenge, and in particular current treatment options for visceral pain are very limited and marginally effective. Na + channel blockade is one of the most powerful and best-proven analgesic principles, beginning with the use of local anesthetics for sensory blockade and then with the discovery that Na v-blocking anticonvulsants also have benefit as pain therapy clinically[ 5 , 13 , 16 , 54 - 57]. Based on their mechanism of action, sodium channel blockers are divided into the following categories: (1) Extracellular blockers, which bind to and occlude the extracellular pore of the channel, e.g. TTX and saxitoxin; (2) Intracellular blockers, which block from the intracellular side of the channel, including local anesthetics, class I antiarrhythmic agents, and some anticonvulsants; and (3) Unknown mechanisms of inhibition of the sodium current in ventricular cells of guinea pigs by e.g. caffeine[ 58], or A-803467, a specific blocker of the Na v1.8 channel (SCN10A)[ 59].
Among the neuronal Na + channel subtypes, Na v1.8 is believed to be of importance for certain visceral pain states, and Na v1.8-preferring channel blockers should be able to relieve pain without causing severe effects (due to the restricted expression of this channel type). The compounds are lidocaine, mexiletine, benzocaine, and ambroxol, which are clinically used to treat pain after local or systemic administration. Of these compounds, ambroxol has been reported to effectively suppress pain symptoms in animal models of chronic, neuropathic and inflammatory pain[ 60]. The analgesic effects of ambroxol by either systemic administration to animals, or by topical application in humans can be explained by ambroxol-induced blockade of ion channels in peripheral neurons. However, this has not yet been used to treat visceral pain. Pregabalin effectively inhibits TNBS-induced chronic colonic allodynia in the rat[ 61]. Pregabalin increased distension sensory thresholds to normal levels in IBS patients with rectal hypersensitivity[ 62]. Although downregulation of the α 2δ calcium channel subunit may be one of the mechanisms underlying the analgesic effect of pregablin[ 63], the precise mechanism of pregabalin action is not fully understood. Nevertheless, these findings support the idea that pharmacological substances that modify the function of VGSCs would be potentially useful in treating pain conditions[ 11]. Further experiments are needed to provide more evidence for physicians to treat visceral pain with these pharmacological substances.
Numerous natural toxins have evolved to target sodium channels, either by blocking current through the pore or by modifying channel gating. Among well-studied toxins, the peptide conotoxins from marine snail venoms and δ-atracotoxin and scorpion venom toxins (e.g. birtoxin) constitute another promising source of such modulators. These peptide toxins are of considerable interest not only as probes for investigating the functioning of ion channels and receptors but also as potential therapeutics for neurological disorders, including neuropathic pain and epilepsy[ 64 , 65]. There are three classes of conopeptides that modulate VGSCs: the pore-blocking μ-conotoxins, the δ-conotoxins which delay or inhibit VGSC inactivation, and the μO-conotoxins which inhibit VGSC Na + conductance independent of the TTX binding site. Some of these toxins have been evaluated in animal studies and in preliminary clinical studies, but others are highly novel and may develop into potential drugs for the treatment of pain states by engineering a selective µ-conopeptide that has high affinity and efficacy.
It is noted that until now none of the known neurotoxins or small molecule VGSC inhibitors/modulators is highly selective for a specific VGSC subtype. However, the recent discovery of a genetic link in inherited pain syndromes has advanced our understanding of the contribution of sodium channels to pain in humans and given us hope that it is possible to generate subunit specific blockers. “Gain-of-function” mutations in SCN9A, the gene which encodes Na v1.7, have been linked to two human-inherited pain syndromes, inherited erythromelalgia and paroxysmal extreme pain disorder, while “loss-of-function” mutations in SCN9A have been linked to complete insensitivity to pain[ 12 , 66]. Inhibition of expression of Na v1.8 (PN3/SNS) sodium channels by an antisense oligodeoxynucleotide[ 32] or by Na v1.8 knockdown[ 40 , 41] would be useful tools for treatment of chronic visceral pain in the future.
CONCLUSION
An emerging theme that unifies many supposedly diverse functional GI disorders is altered neuronal excitability, caused by abnormal expression and function of membrane ion channels in organ-specific primary sensory neurons. VGSCs as the main determinants of intrinsic neuronal excitability are particularly appealing targets for pharmacological intervention. Other ion channels such as vanilloid receptor, ATP receptors or potassium channels, which are certainly involved in pain processing, have not been discussed in this review. Nevertheless, the advent of new experimental medicine techniques presents us with an opportunity to test the effectiveness of novel medicines in great detail. The ability to perform hypothesis-driven research with tool compounds in targeted patients with demonstrable visceral hypersensitivity is a stimulating prospect. As VGSCs are expressed in the glial cells, future studies should be directed toward understanding whether glia-neuron interaction is involved in the peripheral sensitization of nociceptors under visceral pain conditions. Since Na + channel subtypes are expressed in a regionally and temporally specific pattern and current VGSC blockers offer little discrimination between various VGSC subtypes, the development of selective blockers could increase their clinical usefulness. For example, a drug that selectively inhibits the Na v1.8 VGSC, which appears to be crucially and specifically involved in visceral nociception, would presumably act as a novel and powerful analgesic, with few side effects in most indications.
Footnotes
Supported by an NIH grant AT005158
Peer reviewer: Dan L Dumitrascu, Professor, President, Romanian Society of Neurogastroenterology, 2nd Medical Department University of Medicine and Pharmacy Iuliu Hatieganu Cluj, Romania
S- Editor Tian L L- Editor Cant MR E- Editor Ma WH
References
1.
Mayer EA, Bradesi S, Chang L, Spiegel BM, Bueller JA, Naliboff BD. Functional GI disorders: from animal models to drug development. Gut. 2008; 57:384-404. [PubMed]
2.
Xu GY, Shenoy M, Winston JH, Mittal S, Pasricha PJ. P2X receptor-mediated visceral hyperalgesia in a rat model of chronic visceral hypersensitivity. Gut. 2008; 57:1230-1237. [PubMed]
3.
Caffrey JM, Eng DL, Black JA, Waxman SG, Kocsis JD. Three types of sodium channels in adult rat dorsal root ganglion neurons. Brain Res. 1992; 592:283-297. [PubMed]
4.
Djouhri L, Newton R, Levinson SR, Berry CM, Carruthers B, Lawson SN. Sensory and electrophysiological properties of guinea-pig sensory neurones expressing Nav 1.7 (PN1) Na+ channel alpha subunit protein. J Physiol. 2003; 546:565-576. [PubMed]
5.
Dib-Hajj SD, Binshtok AM, Cummins TR, Jarvis MF, Samad T, Zimmermann K. Voltage-gated sodium channels in pain states: role in pathophysiology and targets for treatment. Brain Res Rev. 2009; 60:65-83. [PubMed]  [DOI]
6.
Wada A. Roles of voltage-dependent sodium channels in neuronal development, pain, and neurodegeneration. J Pharmacol Sci. 2006; 102:253-268. [PubMed]
7.
Chahine M, Chatelier A, Babich O, Krupp JJ. Voltage-gated sodium channels in neurological disorders. CNS Neurol Disord Drug Targets. 2008; 7:144-158. [PubMed]
8.
Cregg R, Momin A, Rugiero F, Wood JN, Zhao J. Pain channelopathies. J Physiol. 2010; 588:1897-1904. [PubMed]
9.
Dib-Hajj SD, Cummins TR, Black JA, Waxman SG. Sodium channels in normal and pathological pain. Annu Rev Neurosci. 2010; 33:325-347. [PubMed]
10.
Frieboes LR, Palispis WA, Gupta R. Nerve compression activates selective nociceptive pathways and upregulates peripheral sodium channel expression in Schwann cells. J Orthop Res. 2010; 28:753-761. [PubMed]
11.
Leo S, D’Hooge R, Meert T. Exploring the role of nociceptor-specific sodium channels in pain transmission using Nav1.8 and Nav1.9 knockout mice. Behav Brain Res. 2010; 208:149-157. [PubMed]
12.
Reimann F, Cox JJ, Belfer I, Diatchenko L, Zaykin DV, McHale DP, Drenth JP, Dai F, Wheeler J, Sanders F. Pain perception is altered by a nucleotide polymorphism in SCN9A. Proc Natl Acad Sci USA. 2010; 107:5148-5153. [PubMed]
13.
England S. Voltage-gated sodium channels: the search for subtype-selective analgesics. Expert Opin Investig Drugs. 2008; 17:1849-1864. [PubMed]
14.
Catterall WA. From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron. 2000; 26:13-25. [PubMed]
15.
Ji YH, Liu T. The study of sodium channels involved in pain responses using specific modulators. Shengli Xuebao. 2008; 60:628-634. [PubMed]
16.
Momin A, Wood JN. Sensory neuron voltage-gated sodium channels as analgesic drug targets. Curr Opin Neurobiol. 2008; 18:383-388. [PubMed]
17.
Caldwell JH, Schaller KL, Lasher RS, Peles E, Levinson SR. Sodium channel Na(v)1.6 is localized at nodes of ranvier, dendrites, and synapses. Proc Natl Acad Sci USA. 2000; 97:5616-5620. [PubMed]  [DOI]
18.
Craner MJ, Hains BC, Lo AC, Black JA, Waxman SG. Co-localization of sodium channel Nav1.6 and the sodium-calcium exchanger at sites of axonal injury in the spinal cord in EAE. Brain. 2004; 127:294-303. [PubMed]  [DOI]
19.
Black JA, Renganathan M, Waxman SG. Sodium channel Na(v)1.6 is expressed along nonmyelinated axons and it contributes to conduction. Brain Res Mol Brain Res. 2002; 105:19-28. [PubMed]
20.
Fukuoka T, Kobayashi K, Yamanaka H, Obata K, Dai Y, Noguchi K. Comparative study of the distribution of the alpha-subunits of voltage-gated sodium channels in normal and axotomized rat dorsal root ganglion neurons. J Comp Neurol. 2008; 510:188-206. [PubMed]
21.
Brock J. Sea anemone ‘sting’ isolates IB4-negative sensory neurones. J Physiol. 2010; 588:1143. [PubMed]  [DOI]
22.
Waxman SG, Kocsis JD, Black JA. Type III sodium channel mRNA is expressed in embryonic but not adult spinal sensory neurons, and is reexpressed following axotomy. J Neurophysiol. 1994; 72:466-470. [PubMed]
23.
Felts PA, Yokoyama S, Dib-Hajj S, Black JA, Waxman SG. Sodium channel alpha-subunit mRNAs I, II, III, NaG, Na6 and hNE (PN1): different expression patterns in developing rat nervous system. Brain Res Mol Brain Res. 1997; 45:71-82. [PubMed]
24.
Benn SC, Costigan M, Tate S, Fitzgerald M, Woolf CJ. Developmental expression of the TTX-resistant voltage-gated sodium channels Nav1.8 (SNS) and Nav1.9 (SNS2) in primary sensory neurons. J Neurosci. 2001; 21:6077-6085. [PubMed]
25.
Lindia JA, Köhler MG, Martin WJ, Abbadie C. Relationship between sodium channel Nav1.3 expression and neuropathic pain behavior in rats. Pain. 2005; 117:145-153. [PubMed]
26.
Kerr NC, Holmes FE, Wynick D. Novel mRNA isoforms of the sodium channels Na(v)1.2, Na(v)1.3 and Na(v)1.7 encode predicted two-domain, truncated proteins. Neuroscience. 2008; 155:797-808. [PubMed]
27.
Gautron S, Dos Santos G, Pinto-Henrique D, Koulakoff A, Gros F, Berwald-Netter Y. The glial voltage-gated sodium channel: cell- and tissue-specific mRNA expression. Proc Natl Acad Sci USA. 1992; 89:7272-7276. [PubMed]
28.
Felts PA, Black JA, Dib-Hajj SD, Waxman SG. NaG: a sodium channel-like mRNA shared by Schwann cells and other neural crest derivatives. Glia. 1997; 21:269-276. [PubMed]
29.
Saito YA, Strege PR, Tester DJ, Locke GR 3rd, Talley NJ, Bernard CE, Rae JL, Makielski JC, Ackerman MJ, Farrugia G. Sodium channel mutation in irritable bowel syndrome: evidence for an ion channelopathy. Am J Physiol Gastrointest Liver Physiol. 2009; 296:G211-G218. [PubMed]  [DOI]
30.
Verne GN, Sen A, Price DD. Intrarectal lidocaine is an effective treatment for abdominal pain associated with diarrhea-predominant irritable bowel syndrome. J Pain. 2005; 6:493-496. [PubMed]  [DOI]
31.
Yiangou Y, Facer P, Chessell IP, Bountra C, Chan C, Fertleman C, Smith V, Anand P. Voltage-gated ion channel Nav1.7 innervation in patients with idiopathic rectal hypersensitivity and paroxysmal extreme pain disorder (familial rectal pain). Neurosci Lett. 2007; 427:77-82. [PubMed]  [DOI]
32.
Yoshimura N, Seki S, Novakovic SD, Tzoumaka E, Erickson VL, Erickson KA, Chancellor MB, de Groat WC. The involvement of the tetrodotoxin-resistant sodium channel Na(v)1.8 (PN3/SNS) in a rat model of visceral pain. J Neurosci. 2001; 21:8690-8696. [PubMed]
33.
Yoshimura N, Bennett NE, Hayashi Y, Ogawa T, Nishizawa O, Chancellor MB, de Groat WC, Seki S. Bladder overactivity and hyperexcitability of bladder afferent neurons after intrathecal delivery of nerve growth factor in rats. J Neurosci. 2006; 26:10847-10855. [PubMed]
34.
Dang K, Bielefeldt K, Gebhart GF. Differential responses of bladder lumbosacral and thoracolumbar dorsal root ganglion neurons to purinergic agonists, protons, and capsaicin. J Neurosci. 2005; 25:3973-3984. [PubMed]
35.
Beyak MJ, Ramji N, Krol KM, Kawaja MD, Vanner SJ. 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; 287:G845-G855. [PubMed]
36.
Stewart T, Beyak MJ, Vanner S. Ileitis modulates potassium and sodium currents in guinea pig dorsal root ganglia sensory neurons. J Physiol. 2003; 552:797-807. [PubMed]
37.
King DE, Macleod RJ, Vanner SJ. Trinitrobenzenesulphonic acid colitis alters Na 1.8 channel expression in mouse dorsal root ganglia neurons. Neurogastroenterol Motil. 2009; 21:880-e64. [PubMed]
38.
Gebhart GF, Bielefeldt K, Ozaki N. Gastric hyperalgesia and changes in voltage gated sodium channel function in the rat. Gut. 2002; 51 Suppl 1:i15-i18. [PubMed]
39.
Xu GY, Winston JH, Shenoy M, Yin H, Pasricha PJ. Enhanced excitability and suppression of A-type K+ current of pancreas-specific afferent neurons in a rat model of chronic pancreatitis. Am J Physiol Gastrointest Liver Physiol. 2006; 291:G424-G431. [PubMed]
40.
Hillsley K, Lin JH, Stanisz A, Grundy D, Aerssens J, Peeters PJ, Moechars D, Coulie B, Stead RH. Dissecting the role of sodium currents in visceral sensory neurons in a model of chronic hyperexcitability using Nav1.8 and Nav1.9 null mice. J Physiol. 2006; 576:257-267. [PubMed]
41.
Laird JM, Souslova V, Wood JN, Cervero F. Deficits in visceral pain and referred hyperalgesia in Nav1.8 (SNS/PN3)-null mice. J Neurosci. 2002; 22:8352-8356. [PubMed]
42.
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; 12:934-944. [PubMed]
43.
Ritter AM, Martin WJ, Thorneloe KS. The voltage-gated sodium channel Nav1.9 is required for inflammation-based urinary bladder dysfunction. Neurosci Lett. 2009; 452:28-32. [PubMed]
44.
Guerios SD, Wang ZY, Boldon K, Bushman W, Bjorling DE. Blockade of NGF and trk receptors inhibits increased peripheral mechanical sensitivity accompanying cystitis in rats. Am J Physiol Regul Integr Comp Physiol. 2008; 295:R111-R122. [PubMed]
45.
Gold MS, Zhang L, Wrigley DL, Traub RJ. Prostaglandin E(2) modulates TTX-R I(Na) in rat colonic sensory neurons. J Neurophysiol. 2002; 88:1512-1522. [PubMed]
46.
Ma K, Zhou QH, Chen J, Du DP, Ji Y, Jiang W. TTX-R Na+ current-reduction by celecoxib correlates with changes in PGE(2) and CGRP within rat DRG neurons during acute incisional pain. Brain Res. 2008; 1209:57-64. [PubMed]
47.
Maingret F, Coste B, Padilla F, Clerc N, Crest M, Korogod SM, Delmas P. Inflammatory mediators increase Nav1.9 current and excitability in nociceptors through a coincident detection mechanism. J Gen Physiol. 2008; 131:211-225. [PubMed]
48.
Zhang H, Verkman AS. Aquaporin-1 tunes pain perception by interaction with Na(v)1.8 Na+ channels in dorsal root ganglion neurons. J Biol Chem. 2010; 285:5896-5906. [PubMed]
49.
Binshtok AM, Wang H, Zimmermann K, Amaya F, Vardeh D, Shi L, Brenner GJ, Ji RR, Bean BP, Woolf CJ. Nociceptors are interleukin-1beta sensors. J Neurosci. 2008; 28:14062-14073. [PubMed]
50.
Xu GY, Winston JH, Shenoy M, Zhou S, Chen JD, Pasricha PJ. The endogenous hydrogen sulfide producing enzyme cystathionine-beta synthase contributes to visceral hypersensitivity in a rat model of irritable bowel syndrome. Mol Pain. 2009; 5:44. [PubMed]
51.
Stamboulian S, Choi JS, Ahn HS, Chang YW, Tyrrell L, Black JA, Waxman SG, Dib-Hajj SD. ERK1/2 mitogen-activated protein kinase phosphorylates sodium channel Na(v)1.7 and alters its gating properties. J Neurosci. 2010; 30:1637-1647. [PubMed]
52.
Villarreal CF, Sachs D, Funez MI, Parada CA, de Queiroz Cunha F, Ferreira SH. The peripheral pro-nociceptive state induced by repetitive inflammatory stimuli involves continuous activation of protein kinase A and protein kinase C epsilon and its Na(V)1.8 sodium channel functional regulation in the primary sensory neuron. Biochem Pharmacol. 2009; 77:867-877. [PubMed]
53.
Choi JS, Hudmon A, Waxman SG, Dib-Hajj SD. Calmodulin regulates current density and frequency-dependent inhibition of sodium channel Nav1.8 in DRG neurons. J Neurophysiol. 2006; 96:97-108. [PubMed]
54.
Matulenko MA, Scanio MJ, Kort ME. Voltage-gated sodium channel blockers for the treatment of chronic pain. Curr Top Med Chem. 2009; 9:362-376. [PubMed]
55.
Bhattacharya A, Wickenden AD, Chaplan SR. Sodium channel blockers for the treatment of neuropathic pain. Neurotherapeutics. 2009; 6:663-678. [PubMed]
56.
Clare JJ. Targeting voltage-gated sodium channels for pain therapy. Expert Opin Investig Drugs. 2010; 19:45-62. [PubMed]
57.
Zuliani V, Rivara M, Fantini M, Costantino G. Sodium channel blockers for neuropathic pain. Expert Opin Ther Pat. 2010; 20:755-779. [PubMed]
58.
Habuchi Y, Tanaka H, Furukawa T, Tsujimura Y. Caffeine-induced block of Na+ current in guinea pig single ventricular cells. Am J Physiol. 1991; 261:H1855-H1863. [PubMed]
59.
Jarvis MF, Honore P, Shieh CC, Chapman M, Joshi S, Zhang XF, Kort M, Carroll W, Marron B, Atkinson R. A-803467, a potent and selective Nav1.8 sodium channel blocker, attenuates neuropathic and inflammatory pain in the rat. Proc Natl Acad Sci USA. 2007; 104:8520-8525. [PubMed]  [DOI]
60.
Gaida W, Klinder K, Arndt K, Weiser T. Ambroxol, a Nav1.8-preferring Na(+) channel blocker, effectively suppresses pain symptoms in animal models of chronic, neuropathic and inflammatory pain. Neuropharmacology. 2005; 49:1220-1227. [PubMed]  [DOI]
61.
Diop L, Raymond F, Fargeau H, Petoux F, Chovet M, Doherty AM. Pregabalin (CI-1008) inhibits the trinitrobenzene sulfonic acid-induced chronic colonic allodynia in the rat. J Pharmacol Exp Ther. 2002; 302:1013-1022. [PubMed]
62.
Houghton LA, Fell C, Whorwell PJ, Jones I, Sudworth DP, Gale JD. Effect of a second-generation alpha2delta ligand (pregabalin) on visceral sensation in hypersensitive patients with irritable bowel syndrome. Gut. 2007; 56:1218-1225. [PubMed]  [DOI]
63.
Camilleri M. alpha2delta ligand: a new, smart pill for visceral pain in patients with hypersensitive irritable bowel syndrome?. Gut. 2007; 56:1337-1338. [PubMed]  [DOI]
64.
Ekberg J, Craik DJ, Adams DJ. Conotoxin modulation of voltage-gated sodium channels. Int J Biochem Cell Biol. 2008; 40:2363-1368. [PubMed]
65.
Zhang MM, Han TS, Olivera BM, Bulaj G, Yoshikami D. Mu-conotoxin KIIIA derivatives with divergent affinities versus efficacies in blocking voltage-gated sodium channels. Biochemistry. 2010; 49:4804-4812. [PubMed]
66.
Cox JJ, Reimann F, Nicholas AK, Thornton G, Roberts E, Springell K, Karbani G, Jafri H, Mannan J, Raashid Y. An SCN9A channelopathy causes congenital inability to experience pain. Nature. 2006; 444:894-988. [PubMed]