Open access peer-reviewed chapter

Targeting Acid-Sensing Ion Channels by Peptide Toxins

Written By

Chengchong Li, Yuhua Wang and Xiang-Ping Chu

Submitted: 27 October 2016 Reviewed: 31 August 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.70764

From the Edited Volume

Neurotoxins

Edited by J. Eric McDuffie

Chapter metrics overview

1,158 Chapter Downloads

View Full Metrics

Abstract

Acid-sensing ion channels (ASICs) are proton-gated ion channels that are highly expressed in the nervous system and play important roles in physiological and pathological conditions. They are also expressed in non-neuronal tissues with different functions. The ASICs rapidly respond to a reduction in extracellular pH with an inward current that is quickly inactivated despite the continuous presence of protons. Recently, protons have been identified as neurotransmitters in the brain. Until now, six different isoforms (ASIC1a, 1b, 2a, 2b, 3 and 4) in rodents have been discovered and they can be assembled into homotrimers or heterotrimers to form an ion channel. Peptide toxins targeting ASICs have been found from the venoms of spider Psalmotoxin-1 (PcTx1), sea anemones (APETx2 and PhcrTx1) and snakes (MitTx and mambalgins). They reveal different pharmacological properties and are selective blockers of ASICs, except for MitTx, which is a potent activator of ASICs. In this mini review, the structure, pharmacology and effects of peptide toxins on ASICs will be introduced and their therapeutic potentials for neurological and psychological diseases will be discussed.

Keywords

  • acid-sensing ion channels
  • peptide neurotoxins
  • pain
  • stroke
  • depression
  • neuron

1. Introduction

With great interests in venom toxins, scientists are extremely involved and enthusiastic about this area of research, as applications of these venoms for drugs could bring about a greater understanding of human diseases, potentially changing and advancing human healthcare [61, 65]. Venoms of species like spiders, sea anemones and snakes have been found to target ion channels with highly therapeutic potentials as drug candidates [17, 38]. To explore structure-function, gating mechanisms and tissue localization of many ion channels, animal venom toxins were important pharmacological tools in the ion channel field [28]. Certain peptides even lead to clinical development and venom-based drugs, such as ziconotide, which is an inhibitor of neuronal voltage-gated calcium channels isolated from Conus magus, designed for patients with intra-ctable pain who fail to respond to other drugs [57, 66].

Recently, protons have been identified as neurotransmitters in the brain [26]. One of the candidate targets for proton sensing is called “acid-sensing ion channels” (ASICs). Three decades ago, the proton-activated inward currents were discovered and recorded in neurons isolated from rat spinal ganglia and from the ganglion of trigeminal nerve by the pioneer Krishtal and Pidoplichko [48, 49]. Twenty years ago, Waldmann et al. first cloned the ASICs [80]. ASICs are widely expressed in the nervous system with high density [1, 62, 80]. Molecular cloning of ASICs has identified four genes (ACCN1–4) encoding at least six ASIC subunits in rodents (ASIC1a, ASIC1b, ASIC2a, ASIC2b, ASIC3 and ASIC4) [35, 84]. Structurally, each ASIC subunit consists of 500–560 amino acids with a simple topology: two transmembrane domains, large extracellular loop (370 amino acids) and short intracellular N- and C-terminals (35–90 amino acids). The structure of ASICs is different from traditional ligand-gated G-protein couple receptor, which has seven transmembrane domains. ASICs can form functional ion channels structurally appearing as trimeric complexes of these subunits [44], which form both homomeric and heteromeric channels with different electrophysiological and pharmacological properties [3, 11, 37, 45, 51, 67, 71]. Among all the ASIC subunits, the ASIC2b and ASIC4 subunits do not form functional homomeric proton-gated ion channels by themselves, but they can associate with other ASIC subunits to reveal new pharmacological properties on the heteromeric channels [21, 51, 67].

ASICs are mainly expressed in the central and peripheral nervous systems, chiefly found in neurons [80, 82]. In central nervous system, ASICs contributed to several physiological and pathological conditions, such as learning and memory, fear conditioning, pain, chemoreception, ischemia, seizures, drug addiction and neuroinflammation, where extracellular acidification occurs [5, 9, 12, 82, 83, 84]. More importantly, ASICs are involved in synaptic physiology and are neurotransmitter receptors critical for amygdala-dependent learning and memory [26]. In peripheral sensory neurons such as dorsal root ganglia (DRG), ASIC1, 2 and 3 are found. During pathological condition such as inflammation, tumors or wounds, peripheral tissue acidosis associated with pain occurs. ASICs are of particular interest because they are profoundly sensitive to moderate acidifications [18]. They are more sensitive than transient receptor potential vanilloid 1 (TRPV1), another ion channel activated by protons, capsaicin and heat in nociceptive neurons. ASICs can produce sustained depolarizing currents upon prolonged tissue acidification compatible with the detection of non-adapting pain [18]. ASIC currents and/or transcripts have also been found in glia, smooth muscle cells, lung epithelial cells, immune cells, urothelial cells, adipose cells, joint cells and osteoclasts, indicating that ASICs likely play a role in non-neuronal cells as well [18, 32, 35, 50, 59, 70, 86]. The review regarding the effects of peptide toxins on ASICs has also been discussed by previous publications [4, 5, 9, 10, 12, 17].

Advertisement

2. Targeting ASICs by peptide toxins

2.1. Psalmotoxin-1 (PcTx1)

Among all the peptide toxins, PcTx1 is the first peptide discovered for the ASICs. PcTx1 was identified from venom of the South American tarantula Psalmopoeus cambridgei [30, 31]. It is a potent and selective inhibitor for both homomeric ASIC1a and heteromeric ASIC1a/2b channels [31, 67]. Structurally, this toxin has 40 amino acids crosslinked by three disulfide bridges [31]. Pharmacologically, the IC50 of PcTx1 is 0.9 nM for homomeric ASIC1a channels [30, 31] and 2.6 nM for heteromeric ASIC1a/2b channels [67] in Xenopus oocytes expressed homomeric ASIC1a or heteromeric ASIC1a/2b channels. In our previous studies, PcTx1 at a concentration of 10 nM significantly inhibits ASIC currents in majority of cultured striatal and cortical neurons, respectively [45, 89]. At concentrations that effectively inhibit the homomeric ASIC1a current, it has no effect on the currents mediated by other configurations of ASICs such as heteromeric ASIC1a/2a channels [31] or known voltage-gated Na+, K+, Ca2+ channels as well as several ligand-gated ion channels [89]. Unlike amiloride, which is a blocker of epithelial sodium channel and directly blocks the ASICs, PcTx1 acts as a gating modifier [9, 35]. PcTx1 shifts the channel from its resting state toward the inactivated state by increasing its apparent affinity for protons [5, 17].

Purified PcTx1 or venom toxin was the first peptide used to explore the function of ASICs in neurological, psychological and other diseases [5]. Our previous studies have shown that PcTx1 reveals neuroprotective effects on mouse cultured cortical neurons subjected to extracellular acidosis as well as oxygen and glucose deprivation [88, 89]. In a rodent experimental stroke model (middle cerebral artery occlusion), central injection of venom toxin or PcTx1 significantly reduces the infarct volume by 60% and the protection by PcTx1 treatment lasts 1 week [60, 89]. Consistent with our findings, similar effect by application of PcTx1 was also found in a model of traumatic spinal cord injury in rats [39]. Venom toxin also shows certain protection in a mouse model of multiple sclerosis associated with axonal degeneration [33] as well as in the mouse MPTP model of Parkinson’s disease [2]. Moreover, PcTx1 decreases the acidosis-mediated cell death in cultured retinal ganglion cells [74]. Collectively, all the results support that PcTx1 might be a potential therapeutic agent for neurological disease [9, 12, 81, 87, 88].

ASIC1a is highly expressed in the amygdala, a brain region critical for fear, arousal and emotions [82, 84, 85]. Central injection of venom toxin reduces mouse innate fearing [14, 16], mouse depression-related behavior [15] and stress-induced elevation in core body temperature of mice [29]. The mechanisms of fear reduction, antidepressant and anxiolytic effects by PcTx1 are likely mediated by inhibition of ASIC1a-containing channels in the amygdala.

PcTx1 has also been used to study pain modulation in rodents [54]. Treatment by PcTx1 was shown to induce a potent analgesic effect in acute pain, inflammatory and neuropathic pain models in mice [54].

ASICs are involved in the central chemoreception [40, 71, 72]. Central injection of PcTx1 in the lateral hypothalamus (LH), nucleus of the solitary tract (NTS) and rostral ventrolateral medulla (RVLM) inhibits the acid-induced stimulating effect on respiration [40, 71, 72]. Thus, ASICs in the LH, NTS and RVLM contribute to central regulation of respiration.

ASICs are also expressed in non-neuronal tissue, including but not limited to smooth muscle cells (VSMC) from arteries, where they might play a role in mechanotransduction of the myogenic response and VSMC migration [25]. ASIC currents recorded in acutely dissociated mice cerebral artery smooth muscle cells are potentiated by PcTx1 in majority of the cells [13]. PcTx1 also reduces store-operated calcium entry in VSMCs in rat pulmonary arteries. By using PcTx1, ASIC1a-containing channels are involved in the vascular mechanotransduction.

PcTx1 itself cannot cross the blood-brain barrier. Therefore, the critical importance is how to deliver the PcTx1 to its correlated damaged specific brain region and to search a small molecule with similar effect as PcTx1 [9].

2.2. APETx2

The peptide toxin APETx2 was isolated from sea anemones (Anthopleura elegantissima) and is a selective inhibitor for ASIC3 and ASIC3-containing channels [22]. Structurally, APETx2 contains 42 amino acids crosslinked by three disulfide bonds, a compact disulfide-bonded core with a four-stranded beta-sheet. APETx2 possesses the disulfide-rich all-beta structural family of peptide toxins usually seen in animal venoms. Pharmacologically, APETx2 inhibits both homomeric ASIC3 channels and heteromeric ASIC3-containing channels in heterologous expression systems as well as primary cultures of sensory neurons in rodents. It inhibits the transient component of ASIC3 currents with an IC50 of 63 nM, without affecting sustained component of ASIC3 currents [22]. However, the affinity of this particular ASIC3 inhibitor is reduced when ASIC3 is associated with other ASIC subunits [22]. For instance, the IC50 for ASIC3/ASIC2b is about 117 nM, whereas the IC50 for ASIC3/ASIC1a is around 2 μM [22]. By acting at this external side, APETx2 directly inhibits the ASIC3 channel, and it does not modify the channel unitary conductance [5].

ASIC3 and ASIC3-containing channels are widely expressed in peripheral sensory neurons and play a critical role in pain modulation [8]. During chronic inflammation, the expression level of ASIC3 was upregulated in rat sensory neurons [52, 53, 77], which might be critical for the sensitization of cutaneous nociceptors during inflammation. Consistent with these findings, a reduction in pH in the skin of human volunteers was involved in non-adapting pain [73], and this cutaneous acid-induced pain is largely mediated by ASIC channels, because it is inhibited by amiloride [46, 56, 76]. Additionally, the non-amiloride ASIC blocker, A-317567 exhibits distinct in vitro and in vivo activities over amiloride [27]. Furthermore, by using APETx2, ASIC3 was identified as a sensor of cutaneous acidic pain and postoperative pain and as an integrator of molecular signals released during inflammation in rat, where it is involved in primary thermal hyperalgesia [18, 19, 20]. In correlation with this result, local peripheral application of APETx2 was found to attenuate mechanical hypersensitivity in a cutaneous inflammatory pain rat model [47].

ASIC3 is mainly expressed in small muscle afferents in rat [19, 58] and in more than 30% of sensory neurons innervating the knee joint in mouse [42]. The expression level of ASIC3 in sensory neurons is enhanced in models of muscle inflammation [79] and acute arthritis [42] in mice. The application of APETx2, in comparison with ASIC3 knockout and knockdown mice, revealed a critical role for ASIC3 in the generation of secondary mechanical hyperalgesia associated with central sensitization achieved in a mouse model of non-inflammatory muscular pain triggered by repeated acid injections into the muscle [63, 68] and in a mouse model of joint inflammation [41]. Consistent with these findings, peripheral application of APETx2 was also found to decrease mechanical hypersensitivity in a non-inflammatory muscular pain in rat [47]. Furthermore, ASIC3 is also involved in the development of primary cutaneous mechanical hyperalgesia induced by muscle inflammation [69, 78]. In a rat model of osteoarthritis, continuous intra-articular injections of APETx2 reduced pain-related behavior and secondary mechanical hyperalgesia [43]. An increase in ASIC3 expression was also seen in afferent sensory neurons of the knee joint [43].

APETx2 significantly reduces the exercise pressor reflex mediated by contracting skeletal muscle in rodents [36, 55, 75]. This is supported by the expression of ASIC3 in muscle metaboreceptors [58]. By using ASIC3 knockout mice, researchers have found minor changes in normal cutaneous mechanical sensitivity [8, 63], whereas other studies did not reveal a significant contribution to mechanosensory function [24]. By using selective inhibitor of ASIC3, ASIC3 has been shown to be a neuronal sensor for the skin vasodilation response to direct pressure in both humans and rodents and for skin protection against pressure ulcers in mice [34]. Thus, APETx2 reduces local vascular tone control through blockade of ASIC3 or ASIC3-containing channels.

2.3. Mambalgins

The two peptides of mambalgins (mambalgin-1 and mambalgin-2) were recently found from the venom of the snake Dendroaspis polylepis polylepis [23]. Structurally, these two toxins contain 57 amino acids and include eight cysteines linked by four disulfide bridges. Pharmacologically, mambalgins inhibit ASIC-like currents in cultured neurons of hippocampus and spinal cord. Furthermore, mambalgins inhibit homomeric ASIC1a, 1b, heteromeric ASIC1a/2a, 1a/2b and 1a/1b channels with IC50 between 50 and 200 nM. Functionally, mambalgins reveal analgesic effects in vivo in models of acute and inflammatory pain through either inhibition of ASIC1a and ASIC1a/2a channels in central nervous system or inhibition of ASIC1b channels in peripheral nervous system [5, 23]. Interestingly, the central analgesic effect of mambalgins revealed strong effect similar to morphine but produces less unwanted side effects [4, 23]. Further studies are needed to explore the cellular and molecular mechanisms responsible for such pain pathways, but brain ASICs appear as promising therapeutic targets for novel analgesic drugs [5]. It is also interesting to know whether mambalgins have other effects in brain besides pain modulation [9].

2.4. PhcrTx1

PhcrTx1 represents a newly discovered peptide, which was isolated from the sea anemones Pseudacris crucifer [64]. Structurally, it contains 32 amino acid residues. This peptide reveals an inhibitor cystine knot scaffold, which has been found in other venomous organisms, such as spider, scorpions and cone snails. Pharmacologically, PhcrTx1 inhibits peak ASIC currents in DRG neurons of rats with an IC50 of 0.1 μM. It does not affect the sustained component of the ASIC current or its desensitization rate. Furthermore, the toxin shows its effect in a closed state of the ASICs rather than an open state. PhcrTx1 also inhibits voltage-gated K+, but not voltage-gated Na+, currents in rat DRG neurons with an IC50 of 3.4 and 3.5 μM for peak and steady-state component, respectively. However, PhcrTx1 inhibits voltage-gated K+ currents in DRG neurons, but with significantly lower potency and efficacy than its ability for inhibition on ASIC currents. Thus, PhcrTx1 represents the frontrunner of a novel structural group of sea anemone toxin that acts on both ASICs and Kv channels with high and low potency, respectively [64]. It is interesting to know whether PhcrTx1 plays any functional role in ASICs.

2.5. MitTx

In 2011, MitTx was discovered from the venom of the Texas coral snake Micrurus tener tener [6]. Structurally, peptide MixTx contains two subunits (MitTx-α and MitTx-β) with a β-bungarotoxin-like structure. The MitTx-α subunit has a 60 amino-acid Kunitz-type peptide and the MitTx-β subunit consists of a 120 amino-acid phospholipase A2-like protein. They can associate with each other in a 1:1 ratio (Kd: 12 nM), but this interaction is non-covalent, unlike the β-bungarotoxins that are linked by an interchain disulfide bond. Pharmacologically, MitTx, unlike other inhibitory toxins for ASICs, strongly activates several homomeric and heteromeric ASICs [6, 7]. MitTx produces long-lasting profound effects on homomeric rodent ASIC1a and ASIC1b currents (EC50: 9 and 23 nM, respectively) and a much lower effect on ASIC3 current (EC50: 830 nM). During physiological pH condition (e.g. pH 7.4), MitTx reveals subtle effects on ASIC2a current, but potently enhances the ASIC current by shifting its activation curve toward less acidic pH. The effects of MitTx on sensory ganglion neurons from ASIC1a knockout mice were disappeared. Collectively, the data further suggest that effects of MixTx depend on ASIC1a-containing channels [6].

MitTx triggers a strong ASIC current in cultured sensory neurons in wild-type mice; these currents are lost in neurons from ASIC1a-knockout, but not from ASIC3-knockout mice. Consistent with this idea, injection of MitTx in the mice hindpaw displays a strong pain-related behavior (licking response). This effect is reduced in ASIC1a knockout mice but persists in ASIC3 knockout mice, suggesting the contribution of peripheral ASIC1a-containing channels in cutaneous pain [6]. It is needed to explore why MitTx produces lost-lasting effects in physiological concentration of pH on ASICs.

Advertisement

3. Conclusion

PcTx1 was the first peptide toxin found to block homomeric ASIC1a and heteromeric ASIC1a/2b channels. APETx2 was the second ASIC-targeting peptide discovered, and it inhibits ASIC3 channels. MitTx was discovered in 2011 and is a strong activator of ASICs during physiological conditions. Mambalgins have strong inhibition on ASIC1 channels. Another sea anemone peptide PhcrTx1 inhibits ASIC currents in DRG neurons. These peptide toxins have been very important to better understand the structure-function relationships of ASICs and their implication in physiological and pathological processes [5, 17]. ASIC-targeting peptides isolated from animal venoms that selectively block this class of channels are therefore not only instrumental as pharmacological tools to explore their function but also represent molecules of great potential therapeutic value [5]. ASIC channels appear therefore as targets for drug development in a variety of pathophysiological conditions [9].

Advertisement

Acknowledgments

We thank the support from Qiqihar Medical University (QY2016ZD-01 to X.P.C) and University of Missouri Research Board (X.P.C).

References

  1. 1. Alvarez de la Rosa D, Krueger SR, Kolar A, Shao D, Fitzsimonds RM, Canessa CM.Distribution, subcellular localization and ontogeny of ASIC1 in the mammalian central nervous system. The Journal of Physiology. 2003;546:77-87. DOI: 10.1113/jphysiol.2002.030692
  2. 2. Arias RL, Sung ML, Vasylyev D, Zhang MY, Albinson K, Kubek K, Kagan N, Beyer C, Lin Q,Dwye JM, Zaleska MM, Bowlby MR, Dunlop J, Monaghan M. Amiloride is neuroprotective in an MPTP model of Parkinson’s disease. Neurobiology of Disease. 2008;31:334-341. DOI: 10.1016/j.nbd.2008.05.008
  3. 3. Askwith CC, Wemmie JA, Price MP, Rokhlina T, Welsh MJ. Acid-sensing ion channel 2 (ASIC2) modulates ASIC1 H+-activated currents in hippocampal neurons. The Journal of Biological Chemistry. 2004;279(18):18296-18305. DOI: 10.1074/jbc.M312145200
  4. 4. Baron A, Diochot S, Salinas M, Deval E, Noël J, Lingueglia E. Venom toxins in the exploration of molecular, physiological and pathophysiological functions of acid-sensing ion channels. Toxicon. 2013;75:187-204. DOI: 10.1016/j.toxicon.2013.04.008
  5. 5. Baron A, Lingueglia E. Pharmacology of acid-sensing ion channels—Physiological and therapeutical perspectives. Neuropharmacology. 2015;94:19-35. DOI: 10.1016/j.neuropharma. 2015.01.005
  6. 6. Bohlen CJ, Chesler AT, Sharif-Naeini R, Medzihradszky KF, Zhou S, King D, Sanchez EE,Burlingame AL, Basbaum AI, Julius DA. Heteromeric Texas coral snake toxin targets acid-sensing ion channels to produce pain. Nature. 2011;479:410-414. DOI: 10.1038/nature10607
  7. 7. Bohlen CJ, Julius D. Receptor-targeting mechanisms of pain causing toxins: How ow? Toxicon. 2012;60:254-264. DOI: 10.1016/j.toxicon.2012.04.336
  8. 8. Chen CC, Zimmer A, Sun WH, Hall J, Brownstein MJ. A role for ASIC3 in the modulation of high-intensity pain stimuli. Proceedings of the National Academy of Sciences of the United States of America. 2002;99:8992-8997. DOI: 10.1073/pnas.122245999
  9. 9. Chu XP, Grasing KA, Wang JQ. Acid-sensing ion channels contribute to neurotoxicity. Translational Stroke Research. 2014;5(1):69-78. DOI: 10.1007/s12975-013-0305-y
  10. 10. Chu XP, Papasian CJ, Wang JQ, Xiong ZG. Modulation of acid-sensing ion channels: Molecular mechanisms and therapeutic potential. International Journal of Physiology, Pathophysiology and Pharmacology. 2011;3(4):288-309
  11. 11. Chu XP, Wemmie JA, Wang WZ, Zhu XM, Saugstad JA, Price MP, Simon RP, Xiong ZG. Subunit-dependent high-affinity zinc inhibition of acid sensing ion channels. The Journal of Neuroscience. 2004;24(40):8678-8689. DOI: 10.1523/JNEUROSCI.2844-04.2004
  12. 12. Chu XP, Xiong ZG. Physiological and pathological functions of acid-sensing ion channels in the central nervous system. Current Drug Targets. 2012;13(2):263-271
  13. 13. Chung WS, Farley JM, Swenson A, Barnard JM, Hamilton G, Chiposi R, Drummond HA. Extracellular acidosis activates ASIC-like channels in freshly isolated cerebral artery smooth muscle cells. American Journal of Physiology-Cell Physiology. 2010;298:C1198-C1208. DOI: 10.1152/ajpcell.00511.2009
  14. 14. Coryell MW, Wunsch AM, Haenfler JM, Allen JE, McBride JL, Davidson BL, Wemmie JA.Restoring acid-sensing ion channel-1a in the amygdala of knock-out mice rescues fear memory but not unconditioned fear responses. The Journal of Neuroscience. 2008;28:13738-13741. DOI: 10.1523/JNEUROSCI.3907-08.2008
  15. 15. Coryell MW, Wunsch AM, Haenfler JM, Allen JE, Schnizler M, Ziemann AE, Cook MN, Dunning JP, Price MP, Rainier JD, Liu Z, Light AR, Langbehn DR, Wemmie JA. Acid-sensing ion channel-1a in the amygdala, a novel therapeutic target in depression related behavior. The Journal of Neuroscience. 2009;29:5381-5388. DOI: 10.1523/JNEUROSCI.0360-09.2009
  16. 16. Coryell MW, Ziemann AE, Westmoreland PJ, Haenfler JM, Kurjakovic Z, Zha XM, Price M, Schnizler MK, Wemmie JA. Targeting ASIC1a reduces innate fear and alters neuronal activity in the fear circuit. Biological Psychiatry. 2007;62:1140-1148. DOI: 10.1016/j.biopsych.2007. 05.008
  17. 17. Cristofori-Armstrong B, Rash LD. Acid-sensing ion channel (ASIC) structure and function: Insights from spider, snake and sea anemone venoms. Neuropharmacology. 2017. pii: S0028-3908(17)30192-2. DOI: 10.1016/j.neuropharm.2017.04.042
  18. 18. Deval E, Gasull X, Noel J, Salinas M, Baron A, Diochot S, Lingueglia E. Acid-sensing ion channels (ASICs): Pharmacology and implication in pain. Pharmacology & Therapeutics. 2010;128:549-558. DOI: 10.1016/j.pharmthera.2010.08.006
  19. 19. Deval E, Noel J, Gasull X, Delaunay A, Alloui A, Friend V, Eschalier A, Lazdunski M, Lingueglia E. Acid-sensing ion channels in postoperative pain. The Journal of Neuroscience. 2011;31:6059-6066. DOI: 10.1523/JNEUROSCI.5266-10.2011
  20. 20. Deval E, Noel J, Lay N, Alloui A, Diochot S, Friend V, Jodar M, Lazdunski M, Lingueglia E.ASIC3, a sensor of acidic and primary inflammatory pain. The EMBO Journal. 2008;27:3047-3055. DOI: 10.1038/emboj.2008.213
  21. 21. Deval E, Salinas M, Baron A, Lingueglia E, Lazdunski M. ASIC2b-dependent regulation of ASIC3, an essential acid-sensing ion channel subunit in sensory neurons via the partner protein PICK-1. The Journal of Biological Chemistry. 2004;279:19531-19539. DOI: 10.1074/jbc.M313078200
  22. 22. Diochot S, Baron A, Rash LD, Deval E, Escoubas P, Scarzello S, Salinas M, Lazdunski MA.New sea anemone peptide, APETx2, inhibits ASIC3, a major acid-sensitive channel in sensory neurons. The EMBO Journal. 2004;23:1516-1525. DOI: 10.1038/sj.emboj.7600177
  23. 23. Diochot S, Baron A, Salinas M, Douguet D, Scarzello S, Dabert-Gay AS, Debayle D, Friend V, Alloui A, Lazdunski M, Lingueglia E. Black mamba venom peptides target acid-sensing ion channels to abolish pain. Nature. 2012;490:552-555. DOI: 10.1038/nature11494
  24. 24. Drew LJ, Rohrer DK, Price MP, Blaver KE, Cockayne DA, Cesare P, Wood JN. Acid-sensing ion channels ASIC2 and ASIC3 do not contribute to mechanically activated currents in mammalian sensory neurones. The Journal of Physiology. 2004;556:691-710. DOI: 10.1113/jphysiol.2003.058693
  25. 25. Drummond HA, Jernigan NL, Grifoni SC. Sensing tension: Epithelial sodium channel/acid-sensing ion channel proteins in cardiovascular homeostasis. Hypertension. 2008;51:1265-1271. DOI: 10.1161/HYPERTENSIONAHA.107.093401
  26. 26. Du J, Reznikov LR, Price MP, Zha XM, Lu Y, Moninger TO, Wemmie JA, Welsh MJ.Protons are a neurotransmitter that regulates synaptic plasticity in the lateral amygdala. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:8961-8966. DOI: 10.1073/pnas.1407018111
  27. 27. Dubé GR, Lehto SG, Breese NM, Baker SJ, Wang X, Matulenko MA, Honoré P, Stewart AO, Moreland RB, Brioni JD. Electrophysiological and in vivo characterization of A-317567, a novel blocker of acid sensing ion channels. Pain. 2005;117(1-2):88-96. DOI: 10.1016/j.pain.2005.05.021
  28. 28. Dutertre S, Lewis RJ. Use of venom peptides to probe ion channel structure and function. The Journal of Biological Chemistry. 2010;285:13315-13320. DOI: 10.1074/jbc.R109.076596
  29. 29. Dwyer JM, Rizzo SJ, Neal SJ, Lin Q, Jow F, Arias RL, Rosenzweig-Lipson S, Dunlop J,Beyer CE. Acid sensing ion channel (ASIC) inhibitors exhibit anxiolytic-like activity in preclinical pharmacological models. Psychopharmacology. 2009;203:41-52. DOI: 10.1007/s00213-008-1373-7
  30. 30. Escoubas P, Bernard C, Lambeau G, Lazdunski M, Darbon H. Recombinant production and solution structure of PcTx1, the specific peptide inhibitor of ASIC1a proton-gated cation channels. Protein Science. 2003;12:1332-1343. DOI: 10.1110/ps.0307003
  31. 31. Escoubas P, De Weille JR, Lecoq A, Diochot S, Waldmann R, Champigny G, Moinier D, Menez A, Lazdunski M. Isolation of a tarantula toxin specific for a class of proton-gated Na+ channels. The Journal of Biological Chemistry. 2000;275:25116-25121. DOI: 10.1074/jbc.M003643200
  32. 32. Feldman DH, Horiuchi M, Keachie K, McCauley E, Bannerman P, Itoh A, Itoh T, Pleasure D. Characterization of acid-sensing ion channel expression in oligodendrocyte-lineage cells. Glia. 2008;56:1238-1249. DOI: 10.1002/glia.20693
  33. 33. Friese MA, Craner MJ, Etzensperger R, Vergo S, Wemmie JA, Welsh MJ, Vincen A, Fugge L. Acid-sensing ion channel-1 contributes to axonal degeneration in autoimmune inflammation of the central nervous system. Nature Medicine. 2007;13:1483-1489. DOI: 10.1038/nm1668
  34. 34. Fromy B, Lingueglia E, Sigaudo-Roussel D, Saumet JL, Lazdunski M. Asic3 is a neuronal mechanosensor for pressure-induced vasodilation that protects against pressure ulcers. Nature Medicine. 2012;18:1205-1207. DOI: 10.1038/nm.2844
  35. 35. Gründer S, Chen X. Structure, function, and pharmacology of acid sensing ion channels (ASICs): Focus on ASIC1a. International Journal of Physiology, Pathophysiology and Pharmacology. 2010;2:73-94
  36. 36. Hayes SG, Kindig AE, Kaufman MP. Blockade of acid sensing ion channels attenuates the exercise pressor reflex in cats. The Journal of Physiology. 2007;581:1271-1282. DOI: 10.1113/jphysiol.2007.129197
  37. 37. Hesselager M, Timmermann DB, Ahring PK. pH dependency and desensitization kinetics of heterologously expressed combinations of acid-sensing ion channel subunits. The Journal of Biological Chemistry. 2004;279(12):11006-11015. DOI: 10.1074/jbc.M313507200
  38. 38. Housley DM, Housley GD, Liddell MJ, Jennings EA. Scorpion toxin peptide action at the ion channel subunit level. Neuropharmacology: pii: S0028-3908(16), 30454-3. 2016. DOI: 10.1016/j.neuropharm.2016.10.004
  39. 39. Hu R, Duan B, Wang D, Yu Y, Li W, Luo H, Lu P, Lin J, Zhu G, Wan Q, Feng H. Role of acid-sensing ion channel 1a in the secondary damage of traumatic spinal cord injury. Annals of Surgery. 2011;254:353-362. DOI: 10.1097/SLA.0b013e31822645b4
  40. 40. Huda R, Pollema-Mays SL, Chang Z, Alheid GF, McCrimmon D, Martina M. Acid-sensing ion channels contribute to chemosensitivity of breathing-related neurons of the nucleus of the solitary tract. The Journal of Physiology. 2012;590(19):4761-4775. DOI: 10.1113/jphysiol.2012.232470
  41. 41. Ikeuchi M, Kolker SJ, Burnes LA, Walder RY, Sluka KA. Role of ASIC3 in the primary and secondary hyperalgesia produced by joint inflammation in mice. Pain. 2008;137:662-669. DOI: 10.1016/j.pain.2008.01.020
  42. 42. Ikeuchi M, Kolker SJ, Sluka KA. Acid-sensing ion channel 3 expression in mouse knee joint afferents and effects of carrageenan induced arthritis. The Journal of Pain. 2009;10:336-342. DOI: 10.1016/j.jpain.2008.10.010
  43. 43. Izumi M, Ikeuchi M, Ji Q, Tani T. Local ASIC3 modulates pain and disease progression in a rat model of osteoarthritis. Journal of Biomedical Science. 2012;19:77. DOI: 10.1186/1423-0127-19-77
  44. 44. Jasti J, Furukawa H, Gonzales EB, Gouaux E. Structure of acid sensing ion channel 1 at 1.9A resolution and low pH. Nature. 2007;449:316-323. DOI: 10.1038/nature06163
  45. 45. Jiang Q, Li MH, Papasian CJ, Branigan D, Xiong ZG, Wang JQ, Chu XP. Characterization of acid-sensing ion channels in medium spiny neurons of mouse striatum. Neuroscience. 2009;162(1):55-66. DOI: 10.1016/j.neuroscience.2009.04.029
  46. 46. Jones NG, Slater R, Cadiou H, McNaughton P, McMahon SB. Acid-induced pain and its modulation in humans. Journal of Neuroscience. 2004, 2004;24:10974-10979. DOI: 10.1523/JNEUROSCI.2619-04.2004
  47. 47. Karczewski J, Spencer RH, Garsky VM, Liang A, Leitl MD, Cato MJ, Cook SP, Kane S,Urban MO. Reversal of acid-induced and inflammatory pain by the selective ASIC3 inhibitor, APETx2. British Journal of Pharmacology. 2010;161:950-960. DOI: 10.1111/j.1476-5381.2010.00918.x
  48. 48. Krishtal OA, Pidoplichko VI. A receptor for protons in the nerve cell membrane. Neuroscience. 1980;5(12):2325-2327. DOI: 10.1016/0306-4522(81)90105-6
  49. 49. Krishtal OA, Pidoplichko VI. Receptor for protons in the membrane of sensory neurons. Brain Research. 1981;214(1):150-154. DOI: 10.1016/0006-8993(81)90446-7
  50. 50. Li WG, Xu TL. ASIC3 channels in multimodal sensory perception. ACS Chemical Neuroscience. 2011;2:26-37. DOI: 10.1021/cn100094b
  51. 51. Lingueglia E, de Weille JR, Bassilana F, Heurteaux C, Sakai H, Waldmann R, Lazdunski M.A modulatory subunit of acid-sensing ion channels in brain and dorsal root ganglion cells. The Journal of Biological Chemistry. 1997;272:29778-29783. DOI: 10.1074/jbc.272.47.29778
  52. 52. Mamet J, Baron A, Lazdunski M, Voilley N. Proinflammatory mediators, stimulators of sensory neuron excitability via the expression of acid-sensing ion channels. Journal of Neuroscience. 2002;22:10662-10670
  53. 53. Mamet J, Lazdunski M, Voilley N. How nerve growth factor drives physiological and inflammatory expressions of acid-sensing ion channel 3 in sensory neurons. The Journal of Biological Chemistry. 2003;278:48907-48913. DOI: 10.1074/jbc.M309468200
  54. 54. Mazzuca M, Heurteaux C, Alloui A, Diochot S, Baron A, Voilley N, Blondeau N, Escoubas P, Gelot A, Cupo A, Zimmer A, Zimmer AM, Eschalier A, Lazdunski M. A tarantula peptide against pain via ASIC1a channels and opioid mechanisms. Nature Neuroscience. 2007;10:943-945. DOI: 10.1038/nn1940
  55. 55. McCord JL, Tsuchimochi H, Kaufman MP. Acid-sensing ion channels contribute to the metaboreceptor component of the exercise pressor reflex. American Journal of Physiology. Heart and Circulatory Physiology. 2009;297:H443-H449. DOI: 10.1152/ajpheart.00328.2009
  56. 56. McMahon SB, Jones NG. Plasticity of pain signaling: Role of neurotrophic factors exemplified by acid-induced pain. Journal of Neurobiology. 2004;61:72-87. DOI: 10.1523/JNEUROSCI.2619-04.2004
  57. 57. Miljanich GP. Ziconotide: Neuronal calcium channel blocker for treating severe chronic pain. Current Medicinal Chemistry. 2004;11:3029-3040. DOI: 10.2174/0929867043363884
  58. 58. Molliver DC, Immke DC, Fierro L, Pare M, Rice FL, EW MC. ASIC3, an acid-sensing ion channel, is expressed in metaboreceptive sensory neurons. Molecular Pain. 2005;1:35. DOI: 10.1186/1744-8069-1-35
  59. 59. Noël J, Salinas M, Baron A, Diochot S, Deval E, Lingueglia E. Current perspectives on acid-sensing ion channels: New advances and therapeutic implications. Expert Review of Clinical Pharmacology. 2010;3:331-346. DOI: 10.1586/ecp.10.13
  60. 60. Pignataro G, Simon RP, Xiong ZG. Prolonged activation of ASIC1a and the time window for neuroprotection in cerebral ischaemia. Brain. 2007;130:151-158. DOI: 10.1093/brain/awl325
  61. 61. Prashanth JR, Hasaballah N, Vetter I. Pharmacological screening technologies for venom peptide discovery. Neuropharmacology S0028-3908 (17), 30130-2. 2017. DOI: 10.1016/j.neuropharm.2017.03.038
  62. 62. Price MP, Gong H, Parsons MG, Kundert JR, Reznikov LR, Bernardinelli L, Chaloner K, Buchanan GF, Wemmie JA, Richerson GB, Cassell MD, Welsh MJ. Localization and behaviors in null mice suggest that ASIC1 and ASIC2 modulate responses to aversive stimuli. Genes, Brain, and Behavior. 2014;13(2):179-194. DOI: 10.1111/gbb.12108
  63. 63. Price MP, McIlwrath SL, Xie J, Cheng C, Qiao J, Tarr DE, Sluka KA, Brennan TJ, Lewin GR, Welsh MJ. The DRASIC cation channel contributes to the detection of cutaneous touch and acid stimuli in mice. Neuron. 2001;32:1071-1083. DOI: 10.1016/S0896-6273(01)00547-5
  64. 64. Rodríguez AA, Salceda E, Garateix AG, Zaharenko AJ, Peigneur S, López O, Pons T, Richardson M, Díaz M, Hernández Y, Ständker L, Tytgat J, Soto E. A novel sea anemone peptide that inhibits acid-sensing ion channels. Peptides. 2014;53:3-12. DOI: 10.1016/j.peptides.2013.06.003
  65. 65. Sarfo-Poku C, Eshun O, Lee KH. Medical application of scorpion venom to breast cancer: A mini-review. Toxicon. 2016;122:109-112. DOI: 10.1016/j.toxicon.2016.09.005
  66. 66. Schmidtko A, Lotsch J, Freynhagen R, Geisslinger G. Ziconotide for treatment of severe chronic pain. Lancet. 2010;375:1569-1577. DOI: 10.1016/S0140-6736(10)60354-6
  67. 67. Sherwood TW, Lee KG, Gormley MG, Askwith CC. Heteromeric acid-sensing ion channels (ASICs) composed of ASIC2b and ASIC1a display novel channel properties and contribute to acidosis-induced neuronal death. The Journal of Neuroscience. 2011;31(26):9723-9734. DOI: 10.1523/JNEUROSCI.1665-11.2011
  68. 68. Sluka KA, Price MP, Breese NM, Stucky CL, Wemmie JA, Welsh MJ. Chronic hyperalgesia induced by repeated acid injections in muscle is abolished by the loss of ASIC3, but not ASIC1. Pain. 2003;106:229-239
  69. 69. Sluka KA, Radhakrishnan R, Benson CJ, Eshcol JO, Price MP, Babinski K, Audette KM,Yeomans DC, Wilson SP. ASIC3 in muscle mediates mechanical, but not heat, hyperalgesia associated with muscle inflammation. Pain. 2007;129:102-112. DOI: 10.1186/1744-8069-1-35
  70. 70. Sluka KA, Winter OC, Wemmie JA. Acid-sensing ion channels: A new target for pain and CNS diseases. Current Opinion in Drug Discovery & Development. 2009;12:693-704
  71. 71. Song N, Guan R, Jiang Q, Hassanzadeh CJ, Chu Y, Zhao X, Wang X, Yang D, Du Q, Chu XP, Shen L. Acid-sensing ion channels are expressed in the ventrolateral medulla and contribute to central chemoreception. Scientific Reports. 2016;6:38777. DOI: 10.1038/srep38777
  72. 72. Song N, Zhang G, Geng W, Liu Z, Jin W, Li L, Cao Y, Zhu D, Yu J, Shen L. Acid sensing ion channel 1 in lateral hypothalamus contributes to breathing control. PloS One. 2012;7:e39982. DOI: 10.1371/journal.pone.0039982
  73. 73. Steen KH, Issberner U, Reeh PW. Pain due to experimental acidosis in human skin: Evidence for non-adapting nociceptor excitation. Neuroscience Letters. 1995;199:29-32. DOI: 10.1016/0304-3940(95)12002-L
  74. 74. Tan J, Ye X, Xu Y, Wang H, Sheng M, Wang F. Acid-sensing ion channel 1a is involved in retinal ganglion cell death induced by hypoxia. Molecular Vision. 2011;17:3300-3308
  75. 75. Tsuchimochi H, Yamauchi K, McCord JL, Kaufman MP. Blockade of acid sensing ion channels attenuates the augmented exercise pressor reflex in rats with chronic femoral artery occlusion. The Journal of Physiology. 2011;589:6173-6189. DOI: 10.1113/jphysiol.2011.217851
  76. 76. Ugawa S, Ueda T, Ishida Y, Nishigaki M, Shibata Y, Shimada S. Amiloride-blockable acid-sensing ion channels are leading acid sensors expressed in human nociceptors. The Journal of Clinical Investigation. 2002;110:1185-1190. DOI: 10.1172/JCI15709
  77. 77. Voilley N, de Weille J, Mamet J, Lazdunski M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation induced expression of acid-sensing ion channels in nociceptors. The Journal of Neuroscience. 2001;21:8026-8033
  78. 78. Walder RY, Gautam M, Wilson SP, Benson CJ, Sluka KA. Selective targeting of ASIC3 using artificial miRNAs inhibits primary and secondary hyperalgesia after muscle inflammation. Pain. 2011;152:2348-2356. DOI: 10.1016/j.pain.2011.06.027
  79. 79. Walder RY, Rasmussen LA, Rainier JD, Light AR, Wemmie JA, Sluka KA. ASIC1 and ASIC3 play different roles in the development of hyperalgesia after inflammatory muscle injury. The Journal of Pain. 2010;11:210-218. DOI: 10.1016/j.jpain.2009.07.004
  80. 80. Waldmann R, Champigny G, Bassilana F, Heurteaux C, Lazdunski M. A proton-gated cation channel involved in acid-sensing. Nature. 1997;386(6621):173-177. DOI: 10.1038/386173a0
  81. 81. Wang YZ, Xu TL. Acidosis acid-sensing ion channels, and neuronal cell death. Molecular Neurobiology. 2011;44:350-358. DOI: 10.1007/s12035-011-8204-2
  82. 82. Wemmie JA, Askwith CC, Lamani E, Cassell MD, Freeman JH Jr, Welsh MJ. Acid-sensing ion channel 1 is localized in brain regions with high synaptic density and contributes to fear conditioning. Journal of Neuroscience. 2003;23:5496-5502
  83. 83. Wemmie JA, Chen J, Askwith CC, Hruska-Hageman AM, Price MP, Nolan BC, Yoder PG, Lamani E, Hoshi T, Freeman JH Jr, Welsh MJ. The acid-activated ion channel ASIC contributes to synaptic plasticity, learning, and memory. Neuron. 2002;34:463-477. DOI: 10.1016/S0896-6273(02)00661-X
  84. 84. Wemmie JA, Price MP, Welsh MJ. Acid-sensing ion channels: Advances, questions and therapeutic opportunities. Trends in Neurosciences. 2006;29(10):578-586. DOI: 10.1016/j.tins.2006.06.014
  85. 85. Wemmie JA, Coryell MW, Askwith CC, Lamani E, Leonard AS, Sigmund CD, Welsh MJ. Overexpression of acid-sensing ion channel 1a in transgenic mice increases acquired fear-related behavior. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:3621-3626. DOI: 10.1073/pnas.0308753101
  86. 86. WL W, Cheng CF, Sun WH, Wong CW, Chen CC. Targeting ASIC3 for pain, anxiety, and insulin resistance. Pharmacology & Therapeutics. 2012;134:127-138. DOI: 10.1016/j.pharmathera. 2011.12.009
  87. 87. Xiong ZG, Chu XP, Simon RP. Acid sensing ion channels–Novel therapeutic targets for ischemic brain injury. Frontiers in Bioscience. 2007;12:1376-1386
  88. 88. Xiong ZG, Pignataro G, Li M, Chang SY, Simon RP. Acid sensing ion channels (ASICs) as pharmacological targets for neurodegenerative diseases. Current Opinion in Pharmacology. 2008;8:25-32. DOI: 10.1016/j.coph.2007.09.001
  89. 89. Xiong ZG, Zhu XM, Chu XP, Minami M, Hey J, Wei WL, MacDonald JF, Wemmie JA, Price MP, Welsh MJ, Simon RP. Neuroprotection in ischemia: Blocking calcium-permeable acid-sensing ion channels. Cell. 2004;118:687-698. DOI: 10.1016/j.cell.2004.08.026

Written By

Chengchong Li, Yuhua Wang and Xiang-Ping Chu

Submitted: 27 October 2016 Reviewed: 31 August 2017 Published: 20 December 2017