DOI QR코드

DOI QR Code

Acid-sensing ion channels (ASICs): therapeutic targets for neurological diseases and their regulation

  • Kweon, Hae-Jin (Department of Brain Science, Daegu Gyeongbuk Institute of Science and Technology (DGIST)) ;
  • Suh, Byung-Chang (Department of Brain Science, Daegu Gyeongbuk Institute of Science and Technology (DGIST))
  • Received : 2013.05.06
  • Published : 2013.06.30

Abstract

Extracellular acidification occurs not only in pathological conditions such as inflammation and brain ischemia, but also in normal physiological conditions such as synaptic transmission. Acid-sensing ion channels (ASICs) can detect a broad range of physiological pH changes during pathological and synaptic cellular activities. ASICs are voltage-independent, proton-gated cation channels widely expressed throughout the central and peripheral nervous system. Activation of ASICs is involved in pain perception, synaptic plasticity, learning and memory, fear, ischemic neuronal injury, seizure termination, neuronal degeneration, and mechanosensation. Therefore, ASICs emerge as potential therapeutic targets for manipulating pain and neurological diseases. The activity of these channels can be regulated by many factors such as lactate, $Zn^{2+}$, and Phe-Met-Arg-Phe amide (FMRFamide)-like neuropeptides by interacting with the channel's large extracellular loop. ASICs are also modulated by G protein-coupled receptors such as CB1 cannabinoid receptors and 5-$HT_2$. This review focuses on the physiological roles of ASICs and the molecular mechanisms by which these channels are regulated.

Keywords

INTRODUCTION

Tissue acidosis is a common feature in pain-generating pathological conditions such as inflammation, ischemic stroke, infections, and cancer. Tissue injury leads to the release of inflammatory mediators, including substance P, bradykinin, histamine, 5-hydroxytryptamin (5-HT or serotonin), glutamate, ATP, interleukin-1, nerve growth factor (NGF), and proton (1). Application of an acidic solution on human skin induces pain (2,3). It is well known that the extracellular pH levels drop to 5.4 in acute inflammation (4). Severe ischemia also induces the reduction of pH to 6.3 or even lower (5,6). In the central nervous system (CNS), the pH of the synaptic cleft can fall during neurotransmission, as the synaptic vesicles are acidic (7,8). Recently, local pH changes during normal brain activity were detected in mouse and human brains, although the extent of pH fluctuations still needs to be clarified (9).

Acid-sensing ion channels (ASICs) play a critical role in the perception of a wide range of pH changes in conditions related to tissue acidosis. ASICs are voltage-insensitive, proton-gated cation channels which are activated by extracellular pH fall. These channels are expressed throughout the central and peripheral nervous system. ASICs belong to the ENaC (Epithelial Na+ Channel)/DEG (Degenerin)/ASIC (Acid-sensing ion channel) superfamily of ion channels. The members of ENaC/DEG/ASIC (EDA) superfamily share the same topology, consisting of two hydrophobic transmembrane domains, a large cysteine-rich extracellular loop, and short intracellular N- and C- termini (Fig. 1A).

Four genes, including ACCN2, ACCN1, ACCN3, and ACCN4, encode at least six different ASIC subunits, ASIC1a, ASIC1b, ASIC2a, ASIC2b, ASIC3, and ASIC4 (Table 1). ASIC1a and ASIC1b are protein products of alternative splicing from the ACCN2 gene, and ASIC2a and ASIC2b are products of the alternative spliced ACCN1 gene. The CNS primarily expresses ASIC1a, ASIC2a, and ASIC2b, while all subunits are expressed in the peripheral nervous system (PNS) (6).

In 2007, Eric Gouaux et al. first showed the tridimensional structure of a chicken ASIC1 channel (10). They found that three subunits are required to form a functional channel (Fig. 1B). All ASIC subunits assemble to form homo- or hetero- trimeric channels, except for ASIC2b and ASIC4. ASIC2b and ASIC4 may only contribute to forming heteromeric channels with other ASIC subunits, and modulate the expression or properties of the channels (11,12). Although the canonical ligand for ASICs is the proton, the large extracellular loop of these channels allowed the possibility of the existence of other non-proton ligands (6,7). Some examples include 2-guanidine-4-methylquinazoline (GMQ) (13) and MitTx from the venom of the Texas coral snake (14) (Table 1). Moreover, amiloride, a non-specific blocker of ASICs, paradoxically triggered the sustained currents of ASIC3 at normal pH through binding to the non-proton ligand sensing domain of channels (15). Therefore, one possibility is that homomeric ASIC2b and homomeric ASIC4 have their own specific ligands (6).

Fig. 1.ASIC subunits and activation of ASICs by extracellular acidification. (A) Each subunit has two hydrophobic transmembrane domains, a large cysteine-rich extracellular loop, and short intracellular N- and C- termini. (B) Three subunits assemble to form a functional homo- or hetero- trimeric channel. (C) ASIC currents evoked by extracellular pH fall in tsA cells. ASIC1a, ASIC2a, and ASIC3 were activated by application of pH 6.0, 4.5, and 5.0 solution, respectively. The membrane potential was clamped to −70 mV. The dashed line indicates zero current level.

Table 1.Properties, inhibitors, and regulators of ASICs

 

BIOPHYSICAL PROPERTIES OF ASICs

All ASIC subunits have different pH sensitivities that enable them to detect a wide range of physiological pH. ASIC1a and ASIC3 are sensitive to slight extracellular acidosis, whereas ASIC2a requires more acidic pH value for activation. The pH values for half maximal activation (pH0.5 activation) of each subunit are described in Table 1. A slight discrepancy is apparent in the pH0.5 activation values depending on different studies. This discrepancy may be due to the different species and heterologous expression systems used by the various researchers. In addition, each subunit shows different biophysical properties in kinetics and ion selectivity. ASICs are mainly permeable to Na+ ion. However, ASIC1a homomeric channels, human ASIC1b homomeric channels, and ASIC1a/2b heteromeric channels are also permeable to Ca2+ ion (6,16,17). Activation of ASIC1a increased intracellular Ca2+ concentration in hippocampal, cortical, and dorsal root ganglion (DRG) neurons (18-20). Although Ca2+ permeability of ASICs is very low, the increase of intracellular Ca2+ through ASICs can induce neuronal injury during brain ischemia accompanied by prolonged acidosis (19,21,22).

A rapid drop in the extracellular pH from 7.4 to 6.0 for 10 seconds evoked transient ASIC1a inward currents that inactivated within seconds (Fig. 1C). ASIC2a homomers were activated by more acidic pH values, and displayed slow activating and slow inactivating currents. ASIC3 homomers have biphasic currents, a rapidly activating transient current followed by a sustained current that does not fully inactivate during prolonged acidosis. The increase in cation conductance that allows fluxes of K+, Cs+, and Na+ ions contributes to the sustained currents of ASIC3. This plays a role in the perception of non-adapting pain. ASIC1a homomeric channels have the unique property of tachyphylaxis. Tachyphylaxis means that the current amplitude is gradually reduced to successive acid stimuli, even though the time interval between repetitive acid stimulations is sufficient for recovery from desensitization (23). Tachyphylaxis occurs due to proton permeation through ASIC1a channels, and it is attenuated by the extracellular Ca2+ ion (23).

A number of pharmacological tools can be used for investigating the functions and properties of ASICs (Table 1). PcTx1, a polypeptide from the venom of spider toxin, specifically inhibits homomeric ASIC1a currents (16,24). ASIC3 channels and ASIC3-containing heteromeric channels are inhibited by APETx2, a sea anemone toxin (25). Non-steroidal anti-inflammatory (NSAIDs) drugs, such as ibuprofen and flurbiprofen, can inhibit the activities of ASIC1a and ASIC3 (26). Above these molecules, A-317567 (27) and mambalgin (28) can be used as inhibitors of ASICs. The activity of ASICs can be modulated by both divalent and trivalent metal ions, including Gd3+ (29), Pb2+ (30), Ni2+ (31), Cd2+ (31), Cu2+ (32), Ca2+ (33), and Zn2+ ions (34). Zinc (Zn2+) inhibits homomeric ASIC1a channels at nanomolar concentrations, whereas this metal ion can also potentiate homomeric ASIC2a, heteromeric ASIC1a/2a, and heteromeric ASIC2a/3 channels at micromolar concentrations (35).

The activity of ASICs can be regulated by a variety of molecules, such as lactate (36), arachidonic acid (37), spermine (38), dynorphins (39), nitric oxide (40), and Phe-Met-Arg-Phe amide (FMRFamide) (41,42) (Table 1). GMQ, a non-proton ligand for ASIC3 can also regulate the function of ASIC1a and ASIC1b, although it does not trigger ASIC1a and ASIC1b currents at normal pH. It has been reported that GMQ shifts the pH dependence of the ASIC1a and ASIC1b activation to lower pH values (43). The regulators of ASICs will be reviewed in a later section.

 

PHYSIOLOGICAL ROLES OF ASICs

In the CNS

Synaptic plasticity: ASIC1a and ASIC2a are widely expressed throughout the brain areas, such as the hippocampus, cingulate cortex, striatum, amygdala, cerebellar cortex, and olfactory bulb (44,45). ASIC2a is predominantly localized in dendrites, dendritic spines, and synaptosomes (46). Localization of ASIC1a and ASIC2a in the excitatory synaptic area suggested that ASICs may contribute to synaptic plasticity (44,47). Mice lacking the ASIC1a gene displayed impaired long-term potentiation (LTP) and spatial memory (47). It was proposed that the activation of postsynaptic ASIC1a induces membrane depolarization which facilitates N-methyl-D-aspartate (NMDA) receptor activation and contributes to hippocampal LTP. However, the functional relevance of ASIC1a in synaptic transmission remains unclear (47,48). Wemmie et al. recently detected local pH changes during normal brain activity in mouse and human brains (9). This study directly supports the potential activation of ASICs during brain activity, although attempts to measure the ASIC-mediated currents during synaptic transmission in hippocampal neurons have not been successful (6,45). However, another recent report suggests that ASIC1a is not necessary for normal hippocampal LTP and spatial memory (48). Normal LTP was observed in ASIC1a-null mice, and disrupting the ASIC1a gene did not impair hippocampal spatial memory (48). Therefore, the role of ASIC1a in learning and memory remains controversial.

Brain ischemia: Severe ischemic stroke results in dramatic extracellular pH decrease to 6.0 (5,6). Deprivation of oxygen due to the reduction of blood flow increases anaerobic metabolism, which leads to the accumulation of lactic acid (36,49,50). Buildup of lactic acid, accompanying proton release from ATP hydrolysis, induces a dramatic decrease in pH value (49). Tissue acidosis from ischemic stroke has been known to induce neuronal injury in the brain, although the underlying mechanism is not clear. Ca2+ toxicity is also essential for ischemic neuronal injury (19,51). Excessive Ca2+ overload in the cell induces neuronal injury by activating cytotoxic cascades (19). Glutamate receptors were considered to be responsible for Ca2+ overload in the ischemic brain (51). However, while clinical trials to protect the brain from Ca2+ toxicity by using glutamate antagonists have failed (52), recent findings suggest that the activation of ASIC1a can induce neuronal injury during brain ischemia through the increase of [Ca2+]i (19,21,22). The infarct volume after transient middle cerebral artery occlusion (MCAO)-induced ischemia was surprisingly reduced by intracerebroventricular injection of ASIC1a blockers, amiloride and PcTx1 in rodents (19). Moreover, disrupting the ASIC1a gene protected the brain from ischemic neuronal injury (19). Therefore, the activation of Ca2+-permeable ASIC1a can induce neuronal injury during brain ischemia, and ASIC1a is a novel neuroprotective therapeutic target for brain ischemia (18,53).

Seizure: ASIC1a has a critical role in seizure termination. It is well known that seizure and intense neuronal excitation induce the decrease of pH levels in the brain (54), and acidosis terminates epileptic activity (55,56). Disrupting the ASIC1a gene enhanced the severity of the seizure, whereas overexpression of the ASIC1a gene displayed the opposite effect in mice (57). Acidosis-induced ASIC1a activation facilitates the function of inhibitory interneurons, which contributes to seizure termination by increasing the inhibitory tone (57). Therefore, ASIC1a activators could be potent therapeutic drugs for seizure.

Multiple sclerosis: ASIC1a also contributes to axonal degeneration during multiple sclerosis. Multiple sclerosis, which is an autoimmune inflammatory disease accompanying tissue acidosis in inflammatory lesions, leads to axonal degeneration in the CNS (58). Excessive accumulation of Ca2+ ion through ASIC1a activation during multiple sclerosis contributes to induce axonal degeneration. Disrupting the ASIC1a gene attenuates axonal degeneration during multiple sclerosis. Amiloride, a non-specific blocker of ASICs, has also displayed neuroprotective effects against axonal degeneration (58). These results suggest that ASIC1a could be an effective neuroprotective target for the treatment of neuronal degeneration.

In the PNS

Nociception: Noxious stimuli is mainly carried by the thin nerve fibers, such as thin myelinated Aδ-fibers and unmyelinated C-fibers, and the neurons participating in nociception are mostly small (59). Almost all ASIC subunits are abundantly expressed in the small and medium nociceptive sensory neurons, which are responsible for pain perception. Among the ASIC subunits, ASIC3, which is the most essential pH sensor for pain, is specifically localized in nociceptive fibers innervating the skeletal and cardiac muscles, joints, and bone (1,11,60). In these tissues, anaerobic metabolism during severe exercise or tissue injury induces the accumulation of lactate and protons, resulting in the activation of nociceptors. In addition, inflammation and tissue injury increase the expression levels of ASIC1a, ASIC2b, and ASIC3 mRNA in DRG neurons (26). Activation of ASICs, which is important for sensitizing cutaneous nociceptors, is upregulated by inflammatory mediators such as bradykinin, arachidonic acid (37,61), and nitric oxide (40). ASIC3-like currents were activated by tissue acidification in rat cutaneous sensory neurons, and acidosis-induced pain was suppressed by APETx2, a specific blocker of ASIC3, or knockdown of ASIC3 with siRNA (62).

Mechanosensation: ASICs have also been found in large mechanosensitive neurons (mechanoreceptors), which are responsible for the perception of mechanical stimuli or proprioception. It was initially proposed that ASICs have a role in mechanotransduction, since their phylogenetic homologues in the nematode Caenorhabditis elegans (C. elegans) (the mechanosensory abnormality 4- or 10- (MEC-4/MEC-10) proteins, which are expressed in touch receptor neurons in C. elegans), is important for touch sensation (63). ASIC subunits are localized in specialized nerve endings of skin and muscle spindles, and cutaneous mechanosensory structures such as Meissner corpuscles, Merkel cell neurite complexes, and Pacinian corpuscles (11,64). Therefore, ASICs have been considered to participate in neurosensory mechanotransduction, although the role of ASICs in mechanosensation is still controversial. The role of ASICs in mechanotransduction has been investigated in behavioral studies by using mice with the targeted gene deleted. Disrupting the ASIC2 gene significantly reduced the firing rates of Aβ-fibers in response to low-threshold mechanical stimuli. Thus, ASIC2 was proposed to be involved in the perception of light touch (65). However, disrupting the ASIC2 gene had no effect on the current amplitude or kinetics in response to mechanical stimuli in large DRG neurons (66). These conflicting results could be generated from the compensatory effects of multiple mechanosensitive ion channels (e.g. TRP channels) or receptors in ASIC knockout mice (64). The role of ASICs in echanosensation remains elusive.

 

REGULATION OF ASICs

Lactate: During brain ischemia, the concentration of lactate has been reported to increase up to 15 mM from the resting level of below 1 mM (36). It is well known that the buildup of lactic acid accompanying acidosis contributes to neuronal injury. Lactate significantly potentiated the amplitude of ASIC currents in rat sensory neurons innervating the heart (36). Potentiation of ASICs by lactate was also observed in excised outside-out membrane patches, indicating that the effect of lactate is not mediated by receptor activation or signaling cascade (36). One hypothesis suggested that lactate may potentiate ASIC currents by chelating extracellular divalent ions such as Ca2+ and Mg2+ ions, which in turn modulate the activities of membrane receptors and ion channels (49). The effect of lactate on ASICs was mimicked by reducing the concentration of Ca2+ and Mg2+ ions in the extracellular solution, even without treating lactate. Moreover, potentiation of ASIC currents by lactate was diminished by increasing the concentration of divalent ions (36). These results suggest that lactate potentiates the activity of ASICs by chelating extracellular divalent ions.

Arachidonic acid: Arachidonic acid (AA), a polyunsaturated fatty acid with a 20-carbon chain and four double bonds, is involved in cellular signaling activities as a lipid second messenger (37,61). AA is liberated from membrane phospholipids by the activation of phospholipase A2 (PLA2). The increase of intracellular Ca2+ concentration during brain ischemia leads to the activation of PLA2, resulting in increased production of AA. AA is also one of the proinflammatory factors playing a critical role in pathological conditions such as inflammation and neurological disorders including ischemic neuronal injury (67). Moreover, AA has been known to regulate the functions of various types of potassium channels, L-type and N-type Ca2+ channels (68), and transient receptor potential (TRP) channels (69). The activity of ASICs is also regulated by AA (37,61) (Fig. 2). The effects of AA can be mediated either by the direct action of AA or by the AA metabolites (70). However, inhibition of AA metabolism has no significant effects on AA-mediated potentiation of ASIC currents (37). Moreover, the regulation of ASICs by AA was not attributed to cell swelling or membrane stretch, both of which were induced by the bath application of hypotonic solution (37). These results suggest that AA increases the activity of ASICs by a direct action, and not by AA metabolism or membrane stretch.

FMRFamide-like neuropeptides: FMRFamide and structurally related peptides, which function as neurotransmitters and neuromodulators, are abundant in invertebrates, including C. elegans (71) and Drosophila melanogaster (72). Several FMRFamide-like neuropeptides such as Phe-Leu-Phe-Gln-Pro-Gln-Arg-Phe amide (neuropeptide FF) and A18Famide (neuropeptide AF) are present in mammals. Inflammation increases the expression level of FMRFamide-like neuropeptides that enhance pain (41). FMRFamide and related peptides were thought to elicit their effects through G protein-coupled receptors (GPCRs), especially, opioid receptors. However, some effects of these peptides were insensitive to naloxone, which is an opioid inverse agonist (73). Recently, FMRFamide and related neuropeptides have been found to directly modulate the activity of ASICs. While FMRFamide triggered no current at normal pH, it potentiated the current amplitude of heterologously expressed ASIC1a by slowing the inactivation rate and generating sustained currents during acidification (73). Neuropeptide FF showed only slight effects on ASIC1a, but significantly potentiated ASIC3 currents (73).

Fig. 2.ASICs are regulated by signal transduction pathways. 5-HT2 receptors activate PLCβ through the heterotrimeric Gq/11 proteins. PLCβ hydrolyzes membrane PI(4,5)P2 to two second messengers, IP3 and DAG. IP3 releases Ca2+ from the internal Ca2+ stores in ER. DAG activates PKC, which enhances the activity of ASICs by interaction with PICK1. CB1 receptors inhibit ASICs via suppression of AC/cAMP pathway. AC is inhibited by the heterotrimeric Gi/o proteins, and inhibition of AC leads to reduction of the cAMP levels, which in turn inhibits binding of PKA to AKAP150. TrkB activates PI3-K, and enhances the membrane expression of ASICs through Akt proteins. Arachidonic acid directly potentiates the amplitude of ASIC currents. Abbreviations: ASICs: Acid-sensing ion channels, 5-HT2R: 5-HT2 receptor, CB1R: Cannabinoid-1 receptor, PLCβ: phospholipase C β, PI(4,5)P2: phosphatidylinositol 4,5-bisphosphate, PI(3,4,5)P3: phosphatidylinositol 3,4,5-trisphosphate, IP3: inositol 1,4,5-trisphosphate, IP3R: inositol 1,4,5-trisphosphate receptor, DAG: diacylglycerol, PKC: protein kinase C, PICK1: protein interacting with C-kinase, ER: endoplasmic reticulum, AC: adenylyl cyclase, cAMP: cyclic AMP, PKA: protein kinase A, AKAP150: A-kinase anchoring protein 150, TrkB: tropomyosin-related kinase B, BDNF: brain-derived neurotrophic factor, PI3-K: phosphatidylinositol 3-kinase, Akt: protein kinase B, AA: arachidonic acid.

Dynorphin: The dynorphin opioid peptides are endogenous neuropeptides, abundantly expressed in the CNS. They have antinociceptive effects at physiological concentrations. However, the expression level of dynorphin A and its metabolite increases under pathological conditions, and dynorphin A at micromolar concentration shows excitotoxic effects leading to neuronal death through opioid and NMDA receptor activation (74). Therefore, dynorphin peptides have either a protective or nociceptive effect on neurons depending on their concentration. Dynorphin A and big dynorphin potentiate the activity of ASIC1a in cortical neurons (39). Dynorphins directly interact with channels to increase ASIC1a activity by reducing steady-state desensitization. Steady-state desensitization shows that channels enter the desensitized state from the closed state during prolonged exposure to moderate acidic pH (75). Big dynorphin also promoted acidosis-induced neuronal death through inhibiting steady-state desensitization of ASIC1a in cortical neurons (76).

Phosphoinositides: ENaC has been well known to be regulated by phosphoinositides such as phosphatidylinositol 4,5-bisphosphate [PI(4,5)P2] and phosphatidylinositol 3,4,5-trisphosphate [PI(3,4,5)P3]. PI(4,5)P2 is a cofactor for the function of ion channels including inwardly rectifying K+ (Kir) channels (77), KCNQ channels (78, 79), voltage-gated Ca2+ channels (VGCC) (80), and TRP channels (81). Several ion channels such as Kir1.1 channels, Kir6.2 channels (82), and transient receptor potential melastatin 4 (TRPM4) channels (83) have PI(3,4,5)P3 sensitivity, although they have different binding affinities to PI(3,4,5)P3. These are well described in a previous review (84). Phosphoinositides have been known to directly interact with ENaC to enhance the activity of channels (85). The negative-charged head groups of phosphoinositides putatively bind to conserved positive-charged residues in ENaC channels. The positive-charged residues in the intracellular C-terminus following the second transmembrane domain in β and γ ENaC are critical for PI(3,4,5)P3 sensitivity of ENaC, while the positive-charged residues in the extreme N-terminus of β and γ ENaC are important for the regulation of ENaC by PI(4,5)P2 (85). However, ASICs do not seem to have sensitivity to phosphoinositides. Supplementing PI(4,5)P2 analogue in the pipette solution had no effects on the activity of ASIC1a (86). Furthermore, activation of muscarinic receptors that hydrolyze endogenous membrane PI(4,5)P2 did not affect the amplitude of ASIC1a currents (86). This result conflicts with the previous study. The bath application of muscarinic receptor agonists strongly inhibited the amplitude of ASIC1a currents in Chinese hamster ovary (CHO) cells (87). Therefore, PI(4,5)P2 sensitivity of ASIC1a still needs to be clarified.

Phosphatidylinositol 3-kinase: Activation of tropomyosin-related kinase B (TrkB) by brain-derived neurotrophic factor (BDNF) induces pain hypersensitivity, although the underlying mechanism is not understood. A recent study found that TrkB activation increases the surface expression and the activity of ASIC1a via the phosphatidylinositol 3-kinase (PI3-K)/protein kinase B (PKB or Akt) pathway (88) (Fig. 2). BDNF-induced forward trafficking of ASIC1a is crucial for secondary mechanical hyperalgesia (88). Phosphorylation of Serine-25 at the N-terminus of ASIC1a is critical for BDNF-induced ASIC1a trafficking. Enhancement of ASIC1a surface expression by BDNF was abolished by the mutation of Serine-25 of ASIC1a to alanine (88). Therefore, ASIC1a could be a potent analgesic target for pain hypersensitivity.

 

MODULATION OF ASICs BY GPCRs

Gi/o protein-coupled receptors

Cannabinoid-1 receptor: Cannabinoid-1 (CB1) receptors, which modulate nociceptive pain, are highly expressed in nociceptive primary sensory neurons (89,90). Activation of CB1 receptors, which belong to the Gi/o protein-coupled receptor family, induces the inhibition of adenylyl cyclase (AC) leading to the reduction of the cyclic AMP (cAMP) level. WIN55,212-2, an agonist of CB1 receptors reversibly inhibited ASIC currents in rat primary sensory neurons (91). Furthermore, the mean number of action potentials induced by acid stimulus decreased following activation of CB1 receptors (91). Suppression of ASIC currents by CB1 receptors was abolished by the application of cAMP analogue or the AC activator forskolin (91). These results indicate that analgesic effects of cannabinoids are mediated by the inhibition of AC/cAMP-dependent pathway through CB1 receptors. A-kinase anchoring protein 150 (AKAP150) has been reported to be co-localized with ASIC1a and ASIC2a in rat cortical neurons, and regulates the activity of these channels. It has also been reported that the inhibition of protein kinase A (PKA) binding to AKAP150 reduces the amplitude of ASIC currents, suggesting that AKAP150 mediates the PKA-dependent phosphorylation of ASICs (92). Therefore, the reduction of PKA activity may be involved in the CB1 receptor-mediated analgesic effects of cannabinoids (Fig. 2).

Gq/11 protein-coupled receptors

5-HT2 receptor: 5-HT (or serotonin), an important proinflammatory mediator, is released from platelets, mast cells, and endothelial cells during tissue injury accompanied by inflammation. 5-HT establishes pain sensation and hyperalgesia through sensitizing nociceptive afferents (93). Proinflammatory mediators such as 5-HT, bradykinin, NGF, and interleukin-1 are known to enhance the activity and the expression levels of ASICs. A mixture of these mediators enhanced the amplitude of ASIC-like currents and the number of neurons expressing ASICs in rat DRG neurons (94). Fourteen mammalian 5-HT receptor subtypes are divided into 7 subgroups of 5-HT receptors (5-HT1-7). 5-HT-induced hyperalgesia is mediated by 5-HT2 receptor subtype, which belongs to the Gq/11 protein-coupled receptor family. 5-HT2 agonists, excluding the 5-HT3 receptor agonists, significantly induced hyperalgesia (95). Another study supported the involvement of 5-HT2 receptors in 5-HT-induced hyperalgesia, which was inhibited by 5-HT2 antagonists in DRG neurons (96). Moreover, activation of 5-HT2 receptors potentiated the activity of ASICs in rat DRG neurons via protein kinase C (PKC)-dependent signaling pathways (93). ASICs have a PDZ-binding domain at their C-termini, and interact with PDZ-containing proteins. The combined proteins regulate the surface expression and the activity of ASICs (97). Protein interacting with C-kinase (PICK1), which was shown to co-localize with ASICs, directly interacts with ASICs through the PDZ-binding domain (98). Therefore, the PKC signaling pathway may be involved in the 5-HT2 receptor-mediated potentiation of ASICs (Fig. 2).

Neurokinin-1 receptor: Activation of neurokinin-1 (NK1) receptors by substance P, which is released from nociceptive nerve fibers, has an antinociceptive effect in acid-induced chronic muscle pain (99). Mice, lacking substance P and neurokinin A production by disrupting the tachykinin 1 (Tac1) gene or injection of a selective NK1 receptor antagonist, displayed persistent long-lasting hyperalgesia (99). Substance P significantly reduced ASIC3-mediated currents in muscle afferent DRG neurons. The inhibitory effect of substance P was restrictively observed in neurons expressing ASIC3. While NK1 receptor is a Gq/11 protein-coupled receptor, the inhibition of ASIC3-mediated currents by substance P is mediated by the unconventional G protein-independent, tyrosine kinase-dependent pathway (99). Replacing GTP with a non-hydrolysable analog in the internal pipette solution, had no effects on the inhibition of ASIC3-mediated currents by substance P (99). However, ASIC3-mediated currents were significantly diminished by a bath application of genistein, a phosphotyrosine kinase (PTK) inhibitor (99). Furthermore, M channel-like activity is also involved in the antinociceptive effect of substance P. During tissue acidosis in the muscle, ASIC3 channels are activated by protons, and depolarize the muscle afferent neurons, resulting in the firing of action potentials and the release of substance P. Substance P activates NK1 receptors, and the activation of NK1 receptors leads to the unconventional pathway, which activates tyrosine kinase and the M channel in the muscle afferent neurons. ASIC3-mediated action potentials are inhibited by M channel activation (99).

 

CONCLUSION

Acid-sensing ion channels are pH sensors for detecting a wide range of pH fluctuations during normal physiological processes and pathological conditions. The functions and the properties of ASICs have been investigated using various pharmacological tools and genetic techniques. ASICs are implicated in a number of physiological activities, including pain perception, synaptic plasticity, learning and memory, fear, ischemic neuronal injury, and mechanosensation. However, questions remain about the involvement of ASICs in mechanosensation and learning and memory. Many signaling molecules that regulate the activity of ASICs exist, and ASICs can be modulated by GPCRs. However, the regulation of ASICs remains poorly understood. Further investigation into the regulatory mechanisms of ASICs is necessary for developing effective therapeutic strategies for the treatment of pain and neurological diseases.

References

  1. Basbaum, A. I., Bautista, D. M., Scherrer, G. and Julius, D. (2009) Cellular and molecular mechanisms of pain. Cell 139, 267-284. https://doi.org/10.1016/j.cell.2009.09.028
  2. Steen, K. H., Issberner, W. and Reeh, P. W. (1995) Pain due to experimental acidosis in human skin: evidence for non-adapting nociceptor excitation. Neurosci. Lett. 199, 29-32. https://doi.org/10.1016/0304-3940(95)12002-L
  3. Jones, N. G., Slater, R., Cadiou, H., McNaughton, P. and McMahon, S. B. (2004) Acid-induced pain and its modulation in humans. J. Neurosci. 24, 10974-10979. https://doi.org/10.1523/JNEUROSCI.2619-04.2004
  4. Steen, K. H., Reeh, P. W., Anton, F. and Handwerker, H. O. (1992) Protons selectively induce lasting excitation and sensitization to mechanical stimulation of nociceptors in rat skin, in vitro. J. Neurosci. 12, 86-95.
  5. Smith, M. L., Von Hanwehr, R. and Siesjo, B. K. (1986) Changes in extra- and intracellular pH in the brain during and following ischemia in hyperglycemic and in moderately hypoglycemic rats. J. Cereb. Blood Flow Metab. 6, 574-583. https://doi.org/10.1038/jcbfm.1986.104
  6. Zha, X. M. (2013) Acid-sensing ion channels: trafficking and synaptic function. Mol. Brain 6, 1. https://doi.org/10.1186/1756-6606-6-1
  7. Wemmie, J. A., Price, M. P. and Welsh, M. J. (2006) Acid-sensing ion channels: advances, questions and therapeutic opportunities. Trends. Neurosci. 29, 578-586. https://doi.org/10.1016/j.tins.2006.06.014
  8. Miesenbock, G., De Angelis, D. A. and Rothman, J. E. (1998) Visualizing secretion and synaptic transmission with pH-sensitive green fluorescent proteins. Nature 394, 192-195. https://doi.org/10.1038/28190
  9. Magnotta, V. A., Heo, H. Y., Dlouhy, B. J., Dahdaleh, N. S., Follmer, R. L., Thedens, D. R., Welsh, M. J. and Wemmie, J. A. (2012) Detecting activity-evoked pH changes in human brain. Proc. Natl. Acad. Sci. U. S. A. 109, 8270-8273. https://doi.org/10.1073/pnas.1205902109
  10. Jasti, J., Furukawa, H., Gonzales, E. B. and Gouaux, E. (2007) Structure of acid-sensing ion channel 1 at 1.9 A resolution and low pH. Nature 449, 316-323. https://doi.org/10.1038/nature06163
  11. Noel, J., Salinas, M., Baron, A., Diochot, S., Deval, E. and Lingueglia, E. (2010) Current perspectives on acid-sensing ion channels: new advances and therapeutic implications. Exp. Rev. Clin. Pharmacol. 3, 331-346. https://doi.org/10.1586/ecp.10.13
  12. Lingueglia, E., de Weille, J. R., Bassilana, F., Heurteaux, C., Sakai, H., Waldmann, R. and Lazdunski, M. (1997) A modulatory subunit of acid sensing ion channels in brain and dorsal root ganglion cells. J. Biol. Chem. 272, 29778-29783. https://doi.org/10.1074/jbc.272.47.29778
  13. Yu, Y., Chen, Z., Li, W. G., Cao, H., Feng, E. G., Yu, F., Liu, H., Jiang, H. and Xu, T. L. (2010) A nonproton ligand sensor in the acid-sensing ion channel. Neuron 68, 61-72. https://doi.org/10.1016/j.neuron.2010.09.001
  14. Bohlen, C. J., Chesler, A. T., Sharif-Naeini, R., Medzihradszky, K. F., Zhou, S., King, D., Sanchez, E. E., Burlingame, A. L., Basbaum, A. I. and Julius, D. (2011) A heteromeric Texas coral snake toxin targets acid-sensing ion channels to produce pain. Nature 479, 410-414. https://doi.org/10.1038/nature10607
  15. Li, W. G., Yu, Y., Huang, C., Cao, H. and Xu, T. L. (2011) Nonproton ligand sensing domain is required for paradoxical stimulation of acid-sensing ion channel 3 (ASIC3) channels by amiloride. J. Biol. Chem. 286, 42635-42646. https://doi.org/10.1074/jbc.M111.289058
  16. Sherwood, T. W., Lee, K. G., Gormley, M. G. and Askwith, C. C. (2011) Heteromeric acid-sensing ion channels (ASICs) composed of ASIC2b and ASIC1a display novel channel properties and contribute to acidosis-induced neuronal death. J. Neurosci. 31, 9723-9734. https://doi.org/10.1523/JNEUROSCI.1665-11.2011
  17. Hoagland, E. N., Sherwood, T. W., Lee, K. G., Walker, C. J. and Askwith, C. C. (2010) Identification of a calcium permeable human acid-sensing ion channel 1 transcript variant. J. Biol. Chem. 285, 41852-41862. https://doi.org/10.1074/jbc.M110.171330
  18. Yermolaieva, O., Leonard, A. S., Schnizler, M. K., Abboud, F. M. and Welsh, M. J. (2004) Extracellular acidosis increases neuronal cell calcium by activating acid-sensing ion channel 1a. Proc. Natl. Acad. Sci. U.S.A. 101, 6752-6757. https://doi.org/10.1073/pnas.0308636100
  19. Xiong, Z. G., Zhu, X. M., Chu, X. P., Minami, M., Hey, J., Wei, W. L., MacDonald, J. F., Wemmie, J. A., Price, M. P., Welsh, M. J. and Simon, R. P. (2004) Neuroprotection in ischemia: blocking calcium-permeable acid-sensing ion channels. Cell 118, 687-698. https://doi.org/10.1016/j.cell.2004.08.026
  20. Samways, D. S., Harkins, A. B. and Egan, T. M. (2009) Native and recombinant ASIC1a receptors conduct negligible Ca2+ entry. Cell Calcium 45, 319-325. https://doi.org/10.1016/j.ceca.2008.12.002
  21. Pignataro, G., Simon, R. P. and Xiong, Z. G. (2007) Prolonged activation of ASIC1a and the time window for neuroprotection in cerebral ischaemia. Brain 130, 151-158.
  22. Gao, J., Duan, B., Wang, D. G., Deng, X. H., Zhang, G. Y., Xu, L. and Xu, T. L. (2005) Coupling between NMDA receptor and acid-sensing ion channel contributes to ischemic neuronal death. Neuron 48, 635-646. https://doi.org/10.1016/j.neuron.2005.10.011
  23. Chen, X. and Grunder, S. (2007) Permeating protons contribute to tachyphylaxis of the acid-sensing ion channel (ASIC) 1a. J. Physiol. 579, 657-670. https://doi.org/10.1113/jphysiol.2006.120733
  24. Escoubas, P., De Weille, J. R., Lecoq, A., Diochot, S., Waldmann, R., Champigny, G., Moinier, D., Menez, A. and Lazdunski, M. (2000) Isolation of a tarantula toxin specific for a class of proton-gated Na+ channels. J. Biol. Chem. 275, 25116-25121. https://doi.org/10.1074/jbc.M003643200
  25. Diochot, S., Baron, A., Rash, L. D., Deval, E., Escoubas, P., Scarzello, S., Salinas, M. and Lazdunski, M. (2004) A new sea anemone peptide, APETx2, inhibits ASIC3, a major acid-sensitive channel in sensory neurons. EMBO J. 23, 1516-1525. https://doi.org/10.1038/sj.emboj.7600177
  26. Voilley, N., de Weille, J., Mamet, J. and Lazdunski, M. (2001) Nonsteroid Anti-Inflammatory Drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J. Neurosci. 21, 8026-8033.
  27. Dube, G. R., Lehto, S. G., Breese, N. M., Baker, S. J., Wang, X., Matulenko, M. A., Honore, P., Stewart, A. O., Moreland, R. B. and Brioni, J. D. (2005) Electrophysiological and in vivo characterization of A-317567, a novel blocker of acid sensing ion channels. Pain 117, 88-96. https://doi.org/10.1016/j.pain.2005.05.021
  28. Diochot, S., Baron, A., Salinas, M., Douguet, D., Scarzello, S., Dabert-Gay, A. S., Debayle, D., Friend, V., Alloui, A., Lazdunski, M. and Lingueglia, E. (2012) Black mamba venom peptides target acid-sensing ion channels to abolish pain. Nature 490, 552-555. https://doi.org/10.1038/nature11494
  29. Babinski, K., Catarsi, S., Biagini, G. and Seguela, P. (2000) Mammalian ASIC2a and ASIC3 subunits co-assemble into heteromeric proton-gated channels sensitive to $Gd^{3+}$. J. Biol. Chem. 275, 28519-28525. https://doi.org/10.1074/jbc.M004114200
  30. Wang, W., Duan, B., Xu, H., Xu, L. and Xu, T. L. (2006) Calcium-permeable acid-sensing ion channel is a molecular target of the neurotoxic metal ion lead. J. Biol Chem. 281, 2497-2505. https://doi.org/10.1074/jbc.M507123200
  31. Staruschenko, A., Dorofeeva, N. A., Bolshakov, K. V. and Stockand, J. D. (2007) Subunit-dependent cadmium and nickel inhibition of acid-sensing ion channels. J. Neurobiol. 67, 97-107. https://doi.org/10.1002/neu.20338
  32. Wang, W., Yu, Y. and Xu, T. L. (2007) Modulation of acid-sensing ion channels by $Cu^{2+}$ in cultured hypothalamic neurons of the rat. Neuroscience 145, 631-641. https://doi.org/10.1016/j.neuroscience.2006.12.009
  33. De Weille, J. and Bassilana, F. (2001) Dependence of the acid-sensitive ion channel, ASIC1a, on extracellular $Ca^{2+}$ ions. Brain Res. 900, 277-281. https://doi.org/10.1016/S0006-8993(01)02345-9
  34. Chu, X. P., Wemmie, J. A., Wang, W. Z., Zhu, X. M., Saugstad, J. A., Price, M. P., Simon, R. P. and Xiong, Z. G. (2004) Subunit-dependent high-affinity zinc inhibition of acid-sensing ion channels. J. Neurosci. 24, 8678-8689. https://doi.org/10.1523/JNEUROSCI.2844-04.2004
  35. Baron, A., Schaefer, L., Lingueglia, E., Champigny, G. and Lazdunski, M. (2001) $Zn^{2+}$ and $H^{+}$ are coactivators of acid-sensing ion channels. J. Biol. Chem. 276, 35361-35367. https://doi.org/10.1074/jbc.M105208200
  36. Immke, D. C. and McCleskey, E. W. (2001) Lactate enhances the acid-sensing $Na^{+}$ channel on ischemia-sensing neurons. Nat. Neurosci. 4, 869-870. https://doi.org/10.1038/nn0901-869
  37. Smith, E. S., Cadiou, H. and McNaughton, P. A. (2007) Arachidonic acid potentiates acid-sensing ion channels in rat sensory neurons by a direct action. Neuroscience 145, 686-698. https://doi.org/10.1016/j.neuroscience.2006.12.024
  38. Duan, B., Wang, Y. Z., Yang, T., Chu, X. P., Yu, Y., Huang, Y., Cao, H., Hansen, J., Simon, R. P., Zhu, M. X., Xiong, Z. G. and Xu, T. L. (2011) Extracellular spermine exacerbates ischemic neuronal injury through sensitization of ASIC1a channels to extracellular acidosis. J. Neurosci. 31, 2101-2112. https://doi.org/10.1523/JNEUROSCI.4351-10.2011
  39. Sherwood, T., Franke, R., Conneely, S., Joyner, J., Arumugan, P. and Askwith, C. (2009) Identification of protein domains that control proton and calcium sensitivity of ASIC1a. J. Biol. Chem. 284, 27899-27907. https://doi.org/10.1074/jbc.M109.029009
  40. Cadiou, H., Studer, M., Jones, N. G., Smith, E. S., Ballard, A., McMahon, S. B. and McNaughton, P. A. (2007) Modulation of acid-sensing ion channel activity by nitric oxide. J. Neurosci. 27, 13251-13260. https://doi.org/10.1523/JNEUROSCI.2135-07.2007
  41. Askwith, C. C., Cheng, C., Ikuma, M., Benson, C., Price, M. P. and Welsh, M. J. (2000) Neuropeptide FF and FMRFamide potentiate acid-evoked currents from sensory neurons and proton-gated DEG/ENaC channels. Neuron 26, 133-141. https://doi.org/10.1016/S0896-6273(00)81144-7
  42. Sherwood, T. W. and Askwith, C. C. (2008) Endogenous arginine-phenylalanine-amide-related peptides alter steadystate desensitization of ASIC1a. J. Biol. Chem. 283, 1818-1830. https://doi.org/10.1074/jbc.M705118200
  43. Alijevic, O. and Kellenberger, S. (2012) Subtype-specific modulation of acid-sensing ion channel (ASIC) function by 2-guanidine-4-methylquinazoline. J. Biol. Chem. 287, 36059-36070. https://doi.org/10.1074/jbc.M112.360487
  44. Wemmie, J. A., Askwith, C. C., Lamani, E., Cassell, M. D., Freeman, J. H. Jr., and Welsh, M. J. (2003) Acid-sensing ion channel 1 is localized in brain Rregions with high synaptic density and contributes to fear conditioning. J. Neurosci. 23, 5496-5502.
  45. de la Rosa, D. A., Krueger, S. R., Kolar, A., Shao, D., Fitzsimonds, R. M. and Canessa, C. M. (2003) Distribution, subcellular localization and ontogeny of ASIC1 in the mammalian central nervous system. J. Physiol. 546, 77-87. https://doi.org/10.1113/jphysiol.2002.030692
  46. Zha, X. M., Costa, V., Harding, A. M., Reznikov, L., Benson, C. J. and Welsh, M. J. (2009) ASIC2 subunits target acid-sensing ion channels to the synapse via an association with PSD-95. J. Neurosci. 29, 8438-8446. https://doi.org/10.1523/JNEUROSCI.1284-09.2009
  47. Wemmie, J. A., Chen, J., Askwith, C. C., Hruska-Hageman, A. M., Price, M. P., Nolan, B. C., Yoder, P. G., Lamani, E., Hoshi, T., Freeman, J. H. Jr., and Welsh, M. J. (2002) The acid-activated ion channel ASIC contributes to synaptic plasticity, learning, and memory. Neuron 34, 463-477. https://doi.org/10.1016/S0896-6273(02)00661-X
  48. Wu, P. Y., Huang, Y. Y., Chen, C. C., Hsu, T. T., Lin, Y. C., Weng, J. Y., Chien, T. C., Cheng, I. H. and Lien, C. C. (2013) Acid-sensing ion channel-1a is not required for normal hippocampal LTP and spatial memory. J. Neurosci. 33, 1828-1832. https://doi.org/10.1523/JNEUROSCI.4132-12.2013
  49. Chu, X. P. and Xiong, Z. G. (2013) Acid-sensing ion channels in pathological conditions. Adv. Exp. Med. Biol. 961, 419-431. https://doi.org/10.1007/978-1-4614-4756-6_36
  50. Ahn, Y. J., Kim, H., Lim, H., Lee, M., Kang, Y., Moon, S., Kim, H. S. and Kim, H. H. (2012) AMP-activated protein kinase: implications on ischemic diseases. BMB Rep. 45, 489-495. https://doi.org/10.5483/BMBRep.2012.45.9.169
  51. Choi, D. W. (1988) Glutamate neurotoxicity and diseases of the nervous system. Neuron 1, 623-634. https://doi.org/10.1016/0896-6273(88)90162-6
  52. Wahlgren, N. G. and Ahmed, N. (2004) Neuroprotection in cerebral ischaemia: facts and fancies--the need for new approaches. Cerebrovasc Dis. 17(Suppl 1), 153-166.
  53. Xiong, Z. G., Pignataro, G., Li, M., Chang, S. Y. and Simon, R. P. (2008) Acid-sensing ion channels (ASICs) as pharmacological targets for neurodegenerative diseases. Curr. Opin. Pharmacol. 8, 25-32. https://doi.org/10.1016/j.coph.2007.09.001
  54. Somjen, G. G. (1984) Acidification of interstitial fluid in hippocampal formation caused by seizures and by spreading depression. Brain Res. 311, 186-188. https://doi.org/10.1016/0006-8993(84)91416-1
  55. Mitchell, W. G. and Grubbs, R. C. (1956) Inhibition of audiogenic seizures by carbon dioxide. Science 123, 223-224. https://doi.org/10.1126/science.123.3189.223
  56. Velisek, L., Dreier, J. P., Stanton, P. K., Heinemann, U. and Moshe, S. L. (1994) Lowering of extracellular pH suppresses low-$Mg^{2+}$-induces seizures in combined entorhinal cortex-hippocampal slices. Exp. Brain Res. 101, 44-52. https://doi.org/10.1007/BF00243215
  57. Ziemann, A. E., Schnizler, M. K., Albert, G. W., Severson, M. A., Howard, M. A. 3rd, Welsh, M. J. and Wemmie, J. A. (2008) Seizure termination by acidosis depends on ASIC1a. Nat. Neurosci. 11, 816-822. https://doi.org/10.1038/nn.2132
  58. Friese, M. A., Craner, M. J., Etzensperger, R., Vergo, S., Wemmie, J. A., Welsh, M. J., Vincent, A. and Fugger, L. (2007) Acid-sensing ion channel-1 contributes to axonal degeneration in autoimmune inflammation of the central nervous system. Nat. Med. 13, 1483-1489. https://doi.org/10.1038/nm1668
  59. Krishtal, O. A. and Pidoplichko, V. I. (1981) A receptor for protons in the membrane of sensory neurons may participate in nociception. Neuroscience 6, 2599-2601. https://doi.org/10.1016/0306-4522(81)90105-6
  60. Ikeuchi, M., Kolker, S. J., Burnes, L. A., Walder, R. Y. and Sluka, K. A. (2008) Role of ASIC3 in the primary and secondary hyperalgesia produced by joint inflammation in mice. Pain 137, 662-669. https://doi.org/10.1016/j.pain.2008.01.020
  61. Allen, N. J. and Attwell, D. (2002) Modulation of ASIC channels in rat cerebellar purkinje neurons by ischaemia-related signals. J. Physiol. 543, 521-529. https://doi.org/10.1113/jphysiol.2002.020297
  62. Deval, E., Noel, J., Lay, N., Alloui, A., Diochot, S., Friend, V., Jodar, M., Lazdunski, M. and Lingueglia, E. (2008) ASIC3, a sensor of acidic and primary inflammatory pain. EMBO J. 27, 3047-3055. https://doi.org/10.1038/emboj.2008.213
  63. Delmas, P., Hao, J. and Rodat-Despoix, L. (2011) Molecular mechanisms of mechanotransduction in mammalian sensory neurons. Nat. Rev. Neurosci. 12, 139-153. https://doi.org/10.1038/nrn2993
  64. Chen, C. C. and Wong, C. W. (2013) Neurosensory mechanotransduction through acid-sensing ion channels. J. Cell. Mol. Med. 17, 337-349. https://doi.org/10.1111/jcmm.12025
  65. Price, M. P., Lewin, G. R., Mcllwrath, S. L., Cheng, C., Xie, J., Heppenstall, P. A., Stucky, C. L., Mannsfeldt, A. G., Brennan, T. J., Drummond, H. A., Qiao, J., Benson, C. J., Tarr, D. E., Hrstka, R. F., Yang, B., Williamson, R. A. and Welsh, M. J. (2000) The mammalian sodium channel BNC1 is required for normal touch sensation. Nature 407, 1007-1011. https://doi.org/10.1038/35039512
  66. Drew, L. J., Rohrer, D. K., Price, M. P., Blaver, K. E., Cockayne, D. A., Cesare, P. and Wood, J. N. (2004) Acid-sensing ion channels ASIC2 and ASIC3 do not contribute to mechanically activated currents in mammalian sensory neurones. J. Physiol. 556, 691-710. https://doi.org/10.1113/jphysiol.2003.058693
  67. Farooqui, A. A., Ong, W. Y. and Horrocks, L. A. (2006) Inhibitors of brain phospholipase A2 activity: their neuropharmacological effects and therapeutic importance for the treatment of neurologic disorders. Pharmacol. Rev. 58, 591-620. https://doi.org/10.1124/pr.58.3.7
  68. Liu, L. and Rittenhouse, A. R. (2000) Effects of arachidonic acid on unitary calcium currents in rat sympathetic neurons. J. Physiol. 525, 391-404. https://doi.org/10.1111/j.1469-7793.2000.00391.x
  69. Chyb, S., Raghu, P. and Hardie, R. C. (1999) Polyunsaturated fatty acids activate the Drosophila light-sensitive channels TRP and TRPL. Nature 397, 255-259. https://doi.org/10.1038/16703
  70. Meves, H. (2008) Arachidonic acid and ion channels: an update. Br. J. pharmacol. 155, 4-16. https://doi.org/10.1038/bjp.2008.216
  71. Nelson, L. S., Kim, K., Memmott, J. E. and Li, C. (1998) FMRFamide-related gene family in the nematode, Caenorhabditis elegans. Brain. Res. Mol. Brain. Res. 58, 103-111. https://doi.org/10.1016/S0169-328X(98)00106-5
  72. Schneider, L. E. and Taghert, P. H. (1988) Isolation and characterization of a Drosophila gene that encodes multiple neuropeptides related to Phe-Met-Arg-Phe-NH2 (FMRFamide). Proc. Natl. Acad. Sci. U. S. A. 85, 1993-1997. https://doi.org/10.1073/pnas.85.6.1993
  73. Gayton, R. J. (1982) Mammalian neuronal actions of FMRFamide and the structurally related opioid Met-enkephalin-Arg6-Phe7. Nature 298, 275-276. https://doi.org/10.1038/298275a0
  74. Hauser, K. F., Foldes, J. K. and Turbek, C. S. (1999) Dynorphin A (1-13) neurotoxicity in vitro: opioid and non-opioid mechanisms in mouse spinal cord neurons. Exp. Neurolo. 160, 361-375. https://doi.org/10.1006/exnr.1999.7235
  75. Babini, E., Paukert, M., Geisler, H. S. and Grunder, S. (2002) Alternative splicing and interaction with di- and polyvalent cations control the dynamic range of acid-sensing ion channel 1 (ASIC1). J. Biol. Chem. 277, 41597-41603. https://doi.org/10.1074/jbc.M205877200
  76. Sherwood, T. W. and Askwith, C. C. (2009) Dynorphin opioid peptides enhance acid-sensing ion channel 1a activity and acidosis-induced neuronal death. J. Neurosci. 29, 14371-14380. https://doi.org/10.1523/JNEUROSCI.2186-09.2009
  77. Hilgemann, D. W. and Ball, R. (1996) Regulation of cardiac $Na^{+}$, $Ca^{2+}$ Exchange and KATP Potassium Channels by $PIP_{2}$. Science 273, 956-959. https://doi.org/10.1126/science.273.5277.956
  78. Suh, B. C. and Hille, B. (2002) Recovery from muscarinic modulation of M current channels requires Phosphatidylinositol 4,5-Bisphosphate synthesis. Neuron 35, 507-520. https://doi.org/10.1016/S0896-6273(02)00790-0
  79. Suh, B. C., Inoue, T., Meyer, T. and Hille, B. (2006) Rapid chemically induced changes of PtdIns(4,5)$P_{2}$ gate KCNQ ion channels. Science 314, 1454-1457. https://doi.org/10.1126/science.1131163
  80. Suh, B. C., Leal, K., and Hille, B. (2010) Modulation of high-voltage activated $Ca^{2+}$ channels by membrane phosphatidylinositol 4,5-bisphosphate. Neuron 67, 224-238. https://doi.org/10.1016/j.neuron.2010.07.001
  81. Rohacs, T. and Nilius, B. (2007) Regulation of transient receptor potential (TRP) channels by phosphoinositides. Pflugers. Archiv. 455, 157-168. https://doi.org/10.1007/s00424-007-0275-6
  82. Rohacs, T., Lopes, C. M., Jin, T., Ramdya, P. P., Molnar, Z. and Logothetis, D. E. (2003) Specificity of activation by phosphoinositides determines lipid regulation of Kir channels. Proc. Natl. Acad. Sci. U. S. A. 100, 745-750. https://doi.org/10.1073/pnas.0236364100
  83. Zhang, Z., Okawa, H., Wang, Y. and Liman, E. R. (2005) Phosphatidylinositol 4,5-bisphosphate rescues TRPM4 channels from desensitization. J. Biol. Chem. 280, 39185-39192. https://doi.org/10.1074/jbc.M506965200
  84. Suh, B. C. and Hille, B. (2008) $PIP_{2}$ is a necessary cofactor for ion channel function: how and why? Annu. Rev. Biophys. 37, 175-195. https://doi.org/10.1146/annurev.biophys.37.032807.125859
  85. Pochynyuk, O., Tong, Q., Medina, J., Vandewalle, A., Staruschenko, A., Bugaj, V. and Stockand, J. D. (2007) Molecular determinants of PI(4,5)$P_{2}$ and PI(3,4,5))$P_{3}$ regulation of the epithelial $Na^{+}$ channel. J. Gen. Physiol. 130, 399-413. https://doi.org/10.1085/jgp.200709800
  86. Li, T., Yang, Y. and Canessa, C. M. (2012) Impact of recovery from desensitization on acid-sensing ion channel-1a (ASIC1a) current and response to high frequency stimulation. J. Biol. Chem. 287, 40680-40689. https://doi.org/10.1074/jbc.M112.418400
  87. Dorofeeva, N. A., Karpushev, A. V., Nikolaev M. V., Bolshakov, K. V. and Stockand, J. D. (2009) Muscarinic M1 modulation of acid-sensing ion channels. Neuroreport 20, 1386-1391. https://doi.org/10.1097/WNR.0b013e3283318912
  88. Duan, B., Liu, D. S., Huang, Y., Zeng, W. Z., Wang, X., Yu, H., Zhu, M. X., Chen, Z. Y. and Xu, T. L. (2012) PI3-kinase/Akt pathway-regulated membrane insertion of acid-sensing ion channel 1a underlies BDNF-induced pain hypersensitivity. J. Neurosci. 32, 6351-6363. https://doi.org/10.1523/JNEUROSCI.4479-11.2012
  89. Ahluwalia, J., Urban, L., Capogna, M., Bevan, S. and Nagy, I. (2000) Cannabinoid 1 receptors are expressed in nociceptive primary sensory neurons. Neuroscience 100, 685-688. https://doi.org/10.1016/S0306-4522(00)00389-4
  90. Hohmann, A. G. and Herkenham, M. (1999) Localization of central cannabinoid CB1 receptor messenger RNA in neuronal subpopulations of rat dorsal root ganglia: a double- abel in situ hybridization study. Neuroscience 90, 923-931. https://doi.org/10.1016/S0306-4522(98)00524-7
  91. Liu, Y. Q., Qiu, F., Qiu, C. Y., Cai, Q., Zou, P., Wu, H. and Hu, W. P. (2012) Cannabinoids inhibit acid-sensing ion channel currents in rat dorsal root ganglion neurons. PLoS One 7, e45531. https://doi.org/10.1371/journal.pone.0045531
  92. Chai, S., Li, M., Lan, J., Xiong, Z. G., Saugstad, J. A., and Simon, R. P. (2007) A kinase-anchoring protein 150 and calcineurin are involved in regulation of acid-sensing ion channels ASIC1a and ASIC2a. J. Biol. Chem. 282, 22668-22677. https://doi.org/10.1074/jbc.M703624200
  93. Qiu, F., Qiu, C. Y., Liu, Y. Q., Wu, D., Li, J. D. and Hu, W. P. (2012) Potentiation of acid-sensing ion channel activity by the activation of 5-$HT_{2}$ receptors in rat dorsal root ganglion neurons. Neuropharmacology 63, 494-500. https://doi.org/10.1016/j.neuropharm.2012.04.034
  94. Mamet, J., Baron, A., Lazdunski, M. and Voilley, N. (2002) ProInflammatory mediators, stimulators of sensory neuron excitability via the expression of acid-sensing ion channels. J. Neurosci. 22, 10662-10670.
  95. Tokunaga, A., Saika, M. and Senba, E. (1998) $5-HT_{2A}$ receptor subtype is involved in the thermal hyperalgesic mechanism of serotonin in the periphery. Pain 76, 349-355. https://doi.org/10.1016/S0304-3959(98)00066-9
  96. Lin, S. Y., Chang, W. J., Lin, C. S., Huang, C. Y., Wang, H. F. and Sun, W. H. (2011) Serotonin receptor 5-HT2B mediates serotonin-induced mechanical hyperalgesia. J. Neurosci. 31, 1410-1418. https://doi.org/10.1523/JNEUROSCI.4682-10.2011
  97. Lingueglia, E. (2007) Acid-sensing ion channels in sensory perception. J. Biol. Chem. 282, 17325-17329. https://doi.org/10.1074/jbc.R700011200
  98. Duggan, A., Garcia-Anoveros, J., and Corey, D. P. (2002) The PDZ domain protein PICK1 and the sodium channels BNaC1 interact and localize at mechanosensory terminals of dorsal root ganglion neurons and dendrites of central neurons. J. Biol. Chem. 277, 5203-5208. https://doi.org/10.1074/jbc.M104748200
  99. Lin, C. C., Chen, W. N., Chen, C. J., Lin, Y. W., Zimmer, A. and Chen, C. C. (2012) An antinociceptive role for substance P in acid-induced chronic muscle pain. PNAS 109, E76-E83. https://doi.org/10.1073/pnas.1108903108
  100. Benson, C. J., Xie, J., Wemmie, J. A., Price, M. P., Henss, J. M., Welsh, M. J., and Snyder, P. M. (2002) Heteromultimers of DEG/ENaC subunits form $H^{+}$-gated channels in mouse sensory neurons. Proc. Natl. Acad. Sci. U. S. A. 99, 2338-2343. https://doi.org/10.1073/pnas.032678399
  101. Sutherland, S. P., Benson, C. J., Adelman, J. P. and McCleskey, E. W. (2001) Acid-sensing ion channel 3 matches the acid-gated current in cardiac ischemia-sensing neurons. Proc. Natl. Acad. Sci. U. S. A. 98, 711-716. https://doi.org/10.1073/pnas.98.2.711
  102. Waldmann, R., Champigny, G., Bassilana, F., Heurteaux, C. and Lazdunski, M. (1997) A proton-gated cation channel involved in acid-sensing. Nature 386, 173-177. https://doi.org/10.1038/386173a0
  103. Waldmann, R. and Lazdunski, M. (1998) $H^{+}$-gated cation channels: neuronal acid sensors in the NaC/DEG family of ion channels. Curr. Opin. Neurobiol. 8, 418-424. https://doi.org/10.1016/S0959-4388(98)80070-6
  104. Baron, A., Voilley, N., Lazdunski, M. and Lingueglia, E. (2008) Acid sensing ion channels in dorsal spinal cord neurons. J. Neurosci. 28, 1498-1508. https://doi.org/10.1523/JNEUROSCI.4975-07.2008
  105. Ettaiche, M., Deval, E., Cougnon, M., Lazdunski, M. and Voilley, N. (2006) Silencing acid-sensing ion channel 1a alters cone-mediated retinal function. J. Neurosci. 26, 5800-5809. https://doi.org/10.1523/JNEUROSCI.0344-06.2006
  106. Walder, R. Y., Rasmussen, L. A., Rainier, J. D., Light, A. R., Wemmie, J. A. and Sluka, K. A. (2010) ASIC1 and ASIC3 play different roles in the development of hyperalgesia after inflammatory muscle injury. J. Pain 11, 210-218. https://doi.org/10.1016/j.jpain.2009.07.004
  107. Ettaiche, M., Guy, N., Hofman, P., Lazdunski, M. and Waldmann, R. (2004) Acid-sensing ion channel 2 is important for retinal function and protects against light-induced retinal degeneration. J. Neurosci. 24, 1005-1012. https://doi.org/10.1523/JNEUROSCI.4698-03.2004
  108. Ugawa, S., Minami, Y., Guo, W., Saishin, Y., Takatsuji, K., Yamamoto, T., Tohyama, M. and Shimada, S. (1998) Receptor that leaves a sour taste in the mouth. Nature 395, 555-556. https://doi.org/10.1038/26882
  109. Lu, Y., Ma, X., Sabharwal, R., Snitsarev, V., Morgan, D., Rahmouni, K., Drummond, H. A., Whiteis, C. A., Costa, V., Price, M., Benson, C., Welsh, M. J., Chapleau, M. W. and Abboud, F. M. (2009) The ion channel ASIC2 is required for baroreceptor and autonomic control of the circulation. Neuron 64, 885-897. https://doi.org/10.1016/j.neuron.2009.11.007

Cited by

  1. Proton-induced currents in substantia gelatinosa neurons of the rat trigeminal subnucleus caudalis vol.762, 2015, https://doi.org/10.1016/j.ejphar.2015.04.052
  2. Involvement of acid-sensing ion channel 1α in hepatic carcinoma cell migration and invasion vol.36, pp.6, 2015, https://doi.org/10.1007/s13277-015-3070-6
  3. The RS685012 Polymorphism of ACCN2, the Human Ortholog of Murine Acid-Sensing Ion Channel (ASIC1) Gene, is Highly Represented in Patients with Panic Disorder vol.18, pp.1, 2016, https://doi.org/10.1007/s12017-015-8380-8
  4. MOLECULAR BASIS OF DENTAL SENSITIVITY: THE ODONTOBLASTS ARE MULTISENSORY CELLS AND EXPRESS MULTIFUNCTIONAL ION CHANNELS 2017, https://doi.org/10.1016/j.aanat.2017.09.006
  5. Novel Potent Orthosteric Antagonist of ASIC1a Prevents NMDAR-Dependent LTP Induction vol.58, pp.11, 2015, https://doi.org/10.1021/jm5017329
  6. ASIC2a-dependent increase of ASIC3 surface expression enhances the sustained component of the currents vol.49, pp.10, 2016, https://doi.org/10.5483/BMBRep.2016.49.10.057
  7. Lancemaside A fromCodonopsis lanceolataModulates the Inflammatory Responses Mediated by Monocytes and Macrophages vol.2014, 2014, https://doi.org/10.1155/2014/405158
  8. Morphine inhibits acid-sensing ion channel currents in rat dorsal root ganglion neurons vol.1554, 2014, https://doi.org/10.1016/j.brainres.2014.01.042
  9. The Regulatory Role of Activating Transcription Factor 2 in Inflammation vol.2014, 2014, https://doi.org/10.1155/2014/950472
  10. Regulating Factors in Acid-Sensing Ion Channel 1a Function vol.41, pp.4, 2016, https://doi.org/10.1007/s11064-015-1768-x
  11. Expression and functions of ASIC1 in the zebrafish retina vol.455, pp.3-4, 2014, https://doi.org/10.1016/j.bbrc.2014.11.021
  12. Inhibition of acid-sensing ion channels by chlorogenic acid in rat dorsal root ganglion neurons vol.567, 2014, https://doi.org/10.1016/j.neulet.2014.03.027
  13. Effects of pH on the response of peripheral nerve to anoxia vol.125, pp.3, 2015, https://doi.org/10.3109/00207454.2014.922971
  14. Differential Regulation of Proton-Sensitive Ion Channels by Phospholipids: A Comparative Study between ASICs and TRPV1 vol.10, pp.3, 2015, https://doi.org/10.1371/journal.pone.0122014
  15. Acid-Sensing Ion Channel 2a (ASIC2a) Promotes Surface Trafficking of ASIC2b via Heteromeric Assembly vol.6, pp.1, 2016, https://doi.org/10.1038/srep30684
  16. Potentiation of acid-sensing ion channel activity by peripheral group I metabotropic glutamate receptor signaling vol.107, 2016, https://doi.org/10.1016/j.phrs.2016.02.018
  17. Modulation of Acid-sensing Ion Channel 1a by Intracellular pH and Its Role in Ischemic Stroke vol.291, pp.35, 2016, https://doi.org/10.1074/jbc.M115.713636
  18. Enhancement of acid-sensing ion channel activity by metabotropic P2Y UTP receptors in primary sensory neurons vol.12, pp.1, 2016, https://doi.org/10.1007/s11302-015-9479-y
  19. Glucose Deficiency Elevates Acid-Sensing Ion Channel 2a Expression and Increases Seizure Susceptibility in Temporal Lobe Epilepsy vol.7, pp.1, 2017, https://doi.org/10.1038/s41598-017-05038-0
  20. Characterization of proton-induced currents in rat trigeminal mesencephalic nucleus neurons vol.1583, 2014, https://doi.org/10.1016/j.brainres.2014.08.009
  21. Up-regulation of acid-sensing ion channels in the capsule of the joint in frozen shoulder vol.97-B, pp.6, 2015, https://doi.org/10.1302/0301-620X.97B6.35254
  22. 4-Isopropyl-2,6-bis(1-phenylethyl)aniline 1, an Analogue of KTH-13 Isolated fromCordyceps bassiana, Inhibits the NF-κB-Mediated Inflammatory Response vol.2015, 2015, https://doi.org/10.1155/2015/143025
  23. Selective Activation of hTRPV1 by N-Geranyl Cyclopropylcarboxamide, an Amiloride-Insensitive Salt Taste Enhancer vol.9, pp.2, 2014, https://doi.org/10.1371/journal.pone.0089062
  24. 21-O-Angeloyltheasapogenol E3, a Novel Triterpenoid Saponin from the Seeds of Tea Plants, Inhibits Macrophage-Mediated Inflammatory Responses in a NF-κB-Dependent Manner vol.2014, 2014, https://doi.org/10.1155/2014/658351
  25. Prolactin potentiates the activity of acid-sensing ion channels in female rat primary sensory neurons vol.103, 2016, https://doi.org/10.1016/j.neuropharm.2015.07.016
  26. NRA-2, a Nicalin Homolog, Regulates Neuronal Death by Controlling Surface Localization of ToxicCaenorhabditis elegansDEG/ENaC Channels vol.289, pp.17, 2014, https://doi.org/10.1074/jbc.M113.533695
  27. Sensitization of ASIC3 by proteinase-activated receptor 2 signaling contributes to acidosis-induced nociception vol.14, pp.1, 2017, https://doi.org/10.1186/s12974-017-0916-4
  28. Scorpion Toxin, BmP01, Induces Pain by Targeting TRPV1 Channel vol.7, pp.9, 2015, https://doi.org/10.3390/toxins7093671
  29. Neuronal calcium signaling in chronic pain vol.357, pp.2, 2014, https://doi.org/10.1007/s00441-014-1942-5
  30. Oxytocin inhibits the activity of acid-sensing ion channels through the vasopressin, V1Areceptor in primary sensory neurons vol.171, pp.12, 2014, https://doi.org/10.1111/bph.12635
  31. Smooth muscle acid-sensing ion channel 1: pathophysiological implication in hypoxic pulmonary hypertension vol.100, pp.2, 2015, https://doi.org/10.1113/expphysiol.2014.081612
  32. ASICs Mediate Pain and Inflammation in Musculoskeletal Diseases vol.30, pp.6, 2015, https://doi.org/10.1152/physiol.00030.2015
  33. ASIC1a Promotes Acid-Induced Autophagy in Rat Articular Chondrocytes through the AMPK/FoxO3a Pathway vol.18, pp.10, 2017, https://doi.org/10.3390/ijms18102125
  34. ASIC1a involves acidic microenvironment-induced activation and autophagy of pancreatic stellate cells vol.8, pp.54, 2018, https://doi.org/10.1039/C8RA03679A
  35. Anticancer, antimicrobial, and analgesic activities of spider venoms vol.7, pp.3, 2018, https://doi.org/10.1039/C8TX00022K