Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2013 Dec 17;170(8):1607–1651. doi: 10.1111/bph.12447

The Concise Guide to Pharmacology 2013/14: Ion Channels

Stephen PH Alexander 1,*, Helen E Benson 2, Elena Faccenda 2, Adam J Pawson 2, Joanna L Sharman 2, William A Catterall 3, Michael Spedding 4, John A Peters 5, Anthony J Harmar 2
PMCID: PMC3892289  PMID: 24528239

Abstract

The Concise Guide to PHARMACOLOGY 2013/14 provides concise overviews of the key properties of over 2000 human drug targets with their pharmacology, plus links to an open access knowledgebase of drug targets and their ligands (www.guidetopharmacology.org), which provides more detailed views of target and ligand properties. The full contents can be found at http://onlinelibrary.wiley.com/doi/10.1111/bph.12444/full.

Ion channels are one of the seven major pharmacological targets into which the Guide is divided, with the others being G protein-coupled receptors, ligand-gated ion channels, catalytic receptors, nuclear hormone receptors, transporters and enzymes. These are presented with nomenclature guidance and summary information on the best available pharmacological tools, alongside key references and suggestions for further reading. A new landscape format has easy to use tables comparing related targets.

It is a condensed version of material contemporary to late 2013, which is presented in greater detail and constantly updated on the website www.guidetopharmacology.org, superseding data presented in previous Guides to Receptors and Channels. It is produced in conjunction with NC-IUPHAR and provides the official IUPHAR classification and nomenclature for human drug targets, where appropriate. It consolidates information previously curated and displayed separately in IUPHAR-DB and the Guide to Receptors and Channels, providing a permanent, citable, point-in-time record that will survive database updates.

An Introduction to Ion Channels

Overview: Ion channels are pore-forming proteins that allow the flow of ions across membranes, either plasma membranes or the membranes of intracellular organelles (Hille, 2001). Many ion channels (such as most Na, K Ca and some Cl channels) are gated by voltage but others (such as certain K and Cl channels, TRP channels, ryanodine receptors and IP3 receptors) are relatively voltage-insensitive and are gated by second messengers and other intracellular and/or extracellular mediators. As such, there is some blurring of the boundaries between “ion channels” and “ligand-gated channels” which are compiled separately in the Concise Guide to PHARMACOLOGY 2013/14.

Resolution of ion channel structures, beginning with K channels (Doyle et al., 1998) then Cl channels (Dutzler et al., 2002) and most recently Na channels (Payandeh et al., 2011) has greatly improved understanding of the structural basis behind ion channel function. Many ion channels (e.g., K, Na, Ca, HCN and TRP channels) share several structural similarities. These channels are thought to have evolved from a common ancestor and have been classified together as the “voltage-gated-like (VGL) ion channel chanome” (see Yu et al., 2005). Other ion channels, however, such as Cl channels, aquaporins and connexins, have completely different structural properties to the VGL channels, having evolved quite separately.

Currently, ion channels (including ligand-gated ion channels) represent the second largest target for existing drugs after G protein-coupled receptors (Overington et al., 2006). However, the advent of novel, faster screening techniques for compounds acting on ion channels (Dunlop et al., 2008) suggests that these proteins represent promising targets for the development of additional, novel therapeutic agents in the near future.

Acknowledgments

We wish to acknowledge the tremendous help provided by the Consultants to the Guides past and present (see list in the Overview, p. 1452). We are also extremely grateful for the financial contributions from the British Pharmacological Society, the International Union of Basic and Clinical Pharmacology, the Wellcome Trust (099156/Z/12/Z]), which support the website and the University of Edinburgh, who host the guidetopharmacology.org website.

Conflict of interest

The authors state that there is no conflict of interest to disclose.

List of records presented

  1. 1609 Acid-sensing (proton-gated) ion channels (ASICs)

  2. 1611 Aquaporins

  3. 1612 CatSper and Two-Pore channels

  4. 1613 Chloride channels

  5. 1620 Connexins and Pannexins

  6. 1621 Cyclic nucleotide-regulated channels

  7. 1623 Epithelial sodium channels (ENaC)

  8. 1625 IP3 receptor

  9. 1626 Potassium channels

  10. 1630 Ryanodine receptor

  11. 1632 Sodium leak channel, non-selective

  12. 1633 Transient receptor potential channels

  13. 1643 Voltage-gated calcium channels

  14. 1645 Voltage-gated proton channel

  15. 1646 Voltage-gated sodium channels

Acid-sensing (proton-gated) ion channels (ASICs)

Overview

Acid-sensing ion channels (ASICs, provisional nomenclature; 27,47) are members of a Na+ channel superfamily that includes the epithelial Na+ channel (ENaC), the FMRF-amide activated channel (FaNaC) of invertebrates, the degenerins (DEG) of Caenorhabitis elegans, channels in Drosophila melanogaster and ‘orphan’ channels that include BLINaC 34 and INaC 35. ASIC subunits contain two TM domains and assemble as homo- or hetero-trimers 22,26 to form proton-gated, voltage-insensitive, Na+ permeable, channels (reviewed in 23). Splice variants of ASIC1 [provisionally termed ASIC1a (ASIC, ASICα, BNaC2α) 43, ASIC1b (ASICβ, BNaC2β) 8 and ASIC1b2 (ASICβ2) 39; note that ASIC1a is also permeable to Ca2+] and ASIC2 [provisionally termed ASIC2a (MDEG1, BNaC1α, BNC1α) 21,33,44 and ASIC2b (MDEG2, BNaC1β) 28 ] have been cloned. Unlike ASIC2a (listed in table), heterologous expression of ASIC2b alone does not support H+-gated currents. A third member, ASIC3 (DRASIC, TNaC1) 42, has been identified. A fourth mammalian member of the family (ASIC4/SPASIC) does not support a proton-gated channel in heterologous expression systems and is reported to down regulate the expression of ASIC1a and ASIC3 1,16,24. ASIC channels are primarily expressed in central and peripheral neurons including nociceptors where they participate in neuronal sensitivity to acidosis. They have also been detected in taste receptor cells (ASIC1-3), photoreceptors and retinal cells (ASIC1-3), cochlear hair cells (ASIC1b), testis (hASIC3), pituitary gland (ASIC4), lung epithelial cells (ASIC1a and -3), urothelial cells, adipose cells (ASIC3), vascular smooth muscle cells (ASIC1-3), immune cells (ASIC1,-3 and -4) and bone (ASIC1-3). The activation of ASIC1a within the central nervous system contributes to neuronal injury caused by focal ischemia 48 and to axonal degeneration in autoimmune inflammation in a mouse model of multiple sclerosis 20. However, activation of ASIC1a can terminate seizures 51. Peripheral ASIC3-containing channels play a role in post-operative pain 12. Further proposed roles for centrally and peripherally located ASICs are reviewed in 47 and 27. The relationship of the cloned ASICs to endogenously expressed proton-gated ion channels is becoming established 14,15,19,25,27,29,38,4547. Heterologously expressed heteromultimers form ion channels with altered kinetics, ion selectivity, pH- sensitivity and sensitivity to blockers that resemble some of the native proton activated currents recorded from neurones 3,6,19,28.

Subunits

Nomenclature ASIC1 ASIC2 ASIC3
HGNC, UniProt ASIC1, P78348 ASIC2, Q16515 ASIC3, Q9UHC3
Endogenous activators (EC50) Extracellular H+ (ASIC1a) (∼1.6x10-7 – 6.3x10-7 M), Extracellular H+ (ASIC1b) (∼6.3x10-7 – 8x10-6 M) Extracellular H+ (∼1x10-5 – 8x10-5 M) Extracellular H+ (transient component) (∼2x10-7 – 6.3x10-7 M), Extracellular H+ (sustained component) (∼5x10-5 – 3.5x10-4 M)
Activators (EC50) GMQ (largely non-desensitizing; at pH 7.4) (∼1x10-3 M), arcaine (at pH 7.4) (∼1.2x10-3 M), agmatine (at pH 7.4) (∼9.8x10-3 M)
Channel Blockers (IC50) psalmotoxin 1 (ASIC1a) (9x10-10 M), Zn2+ (ASIC1a) (∼7x10-9 M), Pb2+ (ASIC1b) (∼1.5x10-6 M), A317567 (ASIC1a) (∼2x10-6 M), Pb2+ (ASIC1a) (∼4x10-6 M), amiloride (ASIC1a) (1x10-5 M), benzamil (ASIC1a) (1x10-5 M), EIPA (ASIC1a) (1x10-5 M), nafamostat (ASIC1a) (∼1.3x10-5 M), amiloride (ASIC1b) (2.1x10-5 – 2.3x10-5 M), flurbiprofen (ASIC1a) (3.5x10-4 M), ibuprofen (ASIC1a) (∼3.5x10-4 M), Ni2+ (ASIC1a) (∼6x10-4 M) amiloride (2.8x10-5 M), A317567 (∼3x10-5 M), nafamostat (∼7x10-5 M), Cd2+ (∼1x10-3 M) APETx2 (transient component only) (6.3x10-8 M), nafamostat (transient component) (∼2.5x10-6 M), A317567 (∼1x10-5 M), amiloride (transient component only - sustained component enhanced by 200μM amiloride at pH 4) (1.6x10-5 – 6.3x10-5 M), Gd3+ (4x10-5 M), Zn2+ (6.1x10-5 M), aspirin (sustained component) (9.2x10-5 M), diclofenac (sustained component) (9.2x10-5 M), salicylic acid (sustained component) (2.6x10-4 M)
Radioligands (Kd) [125I]psalmotoxin 1 (ASIC1a) (2.13x10-10 M)
Functional characteristics ASIC1a: γ ∼14pS, PNa/PK = 5–13, PNa/PCa =2.5, rapid activation rate (5.8–13.7 ms), rapid inactivation rate (1.2–4 s) @ pH 6.0, slow recovery (5.3–13s) @ pH 7.4 ASIC1b: γ ∼ 19 pS, PNa/PK =14.0, PNa ≫ PCa, rapid activation rate (9.9 ms), rapid inactivation rate (0.9–1.7 s) @ pH 6.0, slow recovery (4.4–7.7 s) @ pH 7.4 γ∼10.4–13.4 pS, PNa/PK =10, PNa/PCa = 20, rapid activation rate, moderate inactivation rate (3.3–5.5 s) @ pH 5 γ∼ 13–15 pS; biphasic response consisting of rapidly inactivating transient and sustained components; very rapid activation (<5 ms) and inactivation (0.4s); fast recovery (0.4–0.6 s) @ pH 7.4, transient component partially inactivated at pH 7.2
Comment ASIC1a and ASIC1b are also blocked by diarylamidines (IC50 ∼3 μM for ASIC1a) ASIC2 is also blocked by diarylamidines ASIC3 is also blocked by diarylamidines

Comments

psalmotoxin 1 (PcTx1) inhibits ASIC1a by modifying activation and desensitization by H+, but promotes ASIC1b opening. PcTx1 has little effect upon ASIC2a, ASIC3, or ASIC1a expressed as a heteromultimer with either ASIC2a, or ASIC3 15,19 but does block ASIC1a expressed as a heteromultimer with ASIC2b 36. spermine, which apparently competes with PcTx1 for binding to ASIC1a, selectively enhances the function of the channel 17. Blockade of ASIC1a by PcTx1 activates the endogenous enkephalin pathway and has very potent analgesic effects in rodents 31. APETx2 most potently blocks homomeric ASIC3 channels, but also ASIC2b+ASIC3, ASIC1b+ASIC3, and ASIC1a+ASIC3 heteromeric channels with IC50 values of 117 nM, 900 nM and 2 μM, respectively. APETx2 has no effect on ASIC1a, ASIC1b, ASIC2a, or ASIC2a+ASIC3 14,15. IC50 values for A317567 are inferred from blockade of ASIC channels native to dorsal root ganglion neurones 18. The pEC50 values for proton activation of ASIC channels are influenced by numerous factors including extracellular di- and poly-valent ions, Zn2+, protein kinase C and serine proteases (reviewed in 29). Rapid acidification is required for activation of ASIC1 and ASIC3 due to fast inactivation/desensitization. pEC50 values for H+-activation of either transient, or sustained, currents mediated by ASIC3 vary in the literature and may reflect species and/or methodological differences 4,11,42. The transient and sustained current components mediated by rASIC3 are selective for Na+ 42; for hASIC3 the transient component is Na+ selective (PNa/PK > 10) whereas the sustained current appears non-selective (PNa/PK = 1.6) 4,11. The reducing agents dithiothreitol (DTT) and glutathione (GSH) increase ASIC1a currents expressed in CHO cells and ASIC-like currents in sensory ganglia and central neurons 2,9 whereas oxidation, through the formation of intersubunit disulphide bonds, reduces currents mediated by ASIC1a 50. ASIC1a is also irreversibly modulated by extracellular serine proteases, such as trypsin, through proteolytic cleavage 41. Non-steroidal anti-inflammatory drugs (NSAIDs) are direct blockers of ASIC currents at therapeutic concentrations (reviewed in 40). Extracellular Zn2+ potentiates proton activation of homomeric and heteromeric channels incorporating ASIC2a, but not homomeric ASIC1a or ASIC3 channels 5. However, removal of contaminating Zn2+ by chealation reveals a high affinity block of homomeric ASIC1a and heteromeric ASIC1a+ASIC2 channels by Zn2+ indicating complex biphasic actions of the divalent 10. NO potentiates submaximal currents activated by H+ mediated by ASIC1a, ASIC1b, ASIC2a and ASIC3 7. Ammonium activates ASIC channels (most likely ASIC1a) in midbrain dopaminergic neurones: that may be relevant to neuronal disorders associated with hyperammonemia 32. The positive modulation of homomeric, heteromeric and native ASIC channels by the peptide FMRFamide and related substances, such as neuropeptides FF and SF, is reviewed in detail in 29. Inflammatory conditions and particular pro-inflammatory mediators induce overexpression of ASIC-encoding genes, enhance ASIC currents 30, and in the case of arachidonic acid directly activate the channel 13,37. The sustained current component mediated by ASIC3 is potentiated by hypertonic solutions in a manner that is synergistic with the effect of arachidonic acid 13. Selective activation of ASIC3 by GMQ at a site separate from the proton binding site is potentiated by mild acidosis and reduced extracellular Ca2+ 49.

Aquaporins

Overview

Aquaporins and aquaglyceroporins (provisional nomenclature) are membrane channels that allow the permeation of water and certain other small solutes across the cell membrane. Since the isolation and cloning of the first aquaporin (AQP1) 54, 12 additional members of the family have been identified, although little is known about the functional properties of two of these (AQP11; Q8NBQ7 and AQP12A; Q8IXF9). The other 11 aquaporins can be divided into two families (aquaporins and aquaglyceroporins) depending on whether they are permeable to glycerol 53. One or more members of this family of proteins have been found to be expressed in almost all tissues of the body. Individual AQP subunits have six transmembrane domains with an inverted symmetry between the first three and last three domains 52. Functional AQPs exist as tetramers but, unusually, each subunit contains a separate pore, so each channel has four pores.

Subunits

Nomenclature AQP0 AQP1 AQP2 AQP3 AQP4 AQP5 AQP6 AQP7 AQP8 AQP9 AQP10
HGNC, UniProt MIP, P30301 AQP1, P29972 AQP2, P41181 AQP3, Q92482 AQP4, P55087 AQP5, P55064 AQP6, Q13520 AQP7, O14520 AQP8, O94778 AQP9, O43315 AQP10, Q96PS8
Permeability water (low) water (high) water (high) water (high), glycerol water (high) water (high) water (low), anions water (high), glycerol water (high) water (low), glycerol water (low), glycerol
Endogenous activators cGMP
Inhibitors Hg2+ Ag+, Hg2+, tetraethylammonium Hg2+ Hg2+ (also inhibited by acid pH) Hg2+ Hg2+ Hg2+ Hg2+ Hg2+, phloretin Hg2+
Comment AQP3 is also inhibited by acid pH AQP4 is inhibited by PKC activation AQP6 is an intracellular channel permeable to anions as well as water 55

CatSper and Two-Pore channels

Overview

CatSper channels (CatSper1–4; nomenclature as agreed by NC-IUPHAR, 59) are putative 6TM, voltage-gated, calcium permeant channels that are presumed to assemble as a tetramer of α-like subunits and mediate the current ICatSper. In mammals, CatSper subunits are structurally most closely related to individual domains of voltage-activated calcium channels (Cav) 71. CatSper1 71, CatSper2 70 and CatSpers 3 and 4 61,66,69, in common with a recently identified putative 2TM auxiliary CatSperβ protein 65 and two putative 1TM associated CatSperγ and CatSperδ proteins 58,73, are restricted to the testis and localised to the principle piece of sperm tail.

Subunits

Nomenclature CatSper1 CatSper2 CatSper3 CatSper4
HGNC, UniProt CATSPER1, Q8NEC5 CATSPER2, Q96P56 CATSPER3, Q86XQ3 CATSPER4, Q7RTX7
Activators CatSper1 is constitutively active, weakly facilitated by membrane depolarisation, strongly augmented by intracellular alkalinisation. In human, but not mouse, spermatozoa progesterone (EC50 ∼ 8 nM) also potentiates the CatSper current (ICatSper).
Channel Blockers (IC50) NNC55-0396 (2x10-6 – 1x10-5 M), ruthenium red (1x10-5 M), HC056456 (2x10-5 M), mibefradil (3x10-5 M), Cd2+ (2x10-4 M), Ni2+ (3x10-4 M)
Functional characteristics Calcium selective ion channel (Ba2+>Ca2+>>Mg2+>>Na+); quasilinear monovalent cation current in the absence of extracellular divalent cations; alkalinization shifts the voltage-dependence of activation towards negative potentials [V½ @ pH 6.0 = +87 mV (mouse); V½ @ pH 7.5 = +11mV (mouse) or pH 7.4 = +85 mV (human)] Required for ICatSper Required for ICatSper Required for ICatSper

Comments

CatSper channel subunits expressed singly, or in combination, fail to functionally express in heterologous expression systems 70,71. The properties of CatSper1 tabulated above are derived from whole cell voltage-clamp recordings comparing currents endogenous to spermatozoa isolated from the corpus epididymis of wild-type and Catsper1(-/-) mice 62 and also mature human sperm 63,72. ICatSper is also undetectable in the spermatozoa of Catsper2(-/-), Catsper3(-/-), or Catsper4(-/-) mice and CatSper 1 associates with CatSper 2, 3, or 4 in heterologous expression systems 69. Moreover, targeted disruption of Catsper1, 2, 3, or 4 genes results in an identical phenotype in which spermatozoa fail to exhibit the hyperactive movement (whip-like flagellar beats) necessary for penetration of the egg cumulus and zona pellucida and subsequent fertilization. Such disruptions are associated with a deficit in alkalinization and depolarization-evoked Ca2+ entry into spermatozoa 56,57,69. Thus, it is likely that the CatSper pore is formed by a heterotetramer of CatSpers1–4 69 in association with the auxiliary subunits (β, γ, δ) that are also essential for function 58. CatSper channels are required for the increase in intracellular Ca2+ concentration in sperm evoked by egg zona pellucida glycoproteins 74. The driving force for Ca2+ entry is principally determined by a mildly outwardly rectifying K+ channel (KSper) that, like CatSpers, is activated by intracellular alkalinization 68. Mouse KSper is encoded by mSlo3, a protein detected only in testis 67,68,75. In human sperm, such alkalinization may result from the activation of Hv1, a proton channel 64. Mutations in CatSpers are associated with syndromic and non-syndromic male infertility 60. In human ejaculated spermatozoa, progesterone (8,17]. In addition, certain prostaglandins (e.g. PGF1α, PGE1) also potentiate CatSper mediated currents 63,72.

Chloride channels

Overview

Chloride channels are a functionally and structurally diverse group of anion selective channels involved in processes including the regulation of the excitability of neurones, skeletal, cardiac and smooth muscle, cell volume regulation, transepithelial salt transport, the acidification of internal and extracellular compartments, the cell cycle and apoptosis (reviewed in 92). Excluding the transmitter-gated GABAA and glycine receptors (see separate tables), well characterised chloride channels can be classified as certain members of the voltage-sensitive ClC subfamily, calcium-activated channels, high (maxi) conductance channels, the cystic fibrosis transmembrane conductance regulator (CFTR) and volume regulated channels 155. No official recommendation exists regarding the classification of chloride channels. Functional chloride channels that have been cloned from, or characterised within, mammalian tissues are listed with the exception of several classes of intracellular channels (e.g. CLIC) that are reviewed by in 96.

ClC family

Overview

The mammalian ClC family (reviewed in 77,87,92,94,108) contains 9 members that fall, on the basis of sequence homology, into three groups; ClC-1, ClC-2, hClC-Ka (rClC-K1) and hClC-Kb (rClC-K2); ClC-3 to ClC-5, and ClC-6 and -7. ClC-1 and ClC-2 are plasma membrane chloride channels. ClC-Ka and ClC-Kb are also plasma membrane channels (largely expressed in the kidney and inner ear) when associated with barttin (BSND, Q8WZ55), a 320 amino acid 2TM protein 97. The localisation of the remaining members of the ClC family is likely to be predominantly intracellular in vivo, although they may traffic to the plasma membrane in overexpression systems. Numerous recent reports indicate that ClC-4, ClC-5, ClC-6 and ClC-7 (and by inference ClC-3) function as Cl-/H+ antiporters (secondary active transport), rather than classical Cl- channels 103,114,125,134,146; reviewed in 77,139). It has recently been reported that the activity of ClC-5 as a Cl-/H+ exchanger is important for renal endocytosis 127. Alternative splicing increases the structural diversity within the ClC family. The crystal structure of two bacterial ClC proteins has been described 95 and a eukaryotic ClC transporter (CmCLC) has recently been described at 3.5 Å resolution 99. Each ClC subunit, with a complex topology of 18 intramembrane segments, contributes a single pore to a dimeric ‘double-barrelled’ ClC channel that contains two independently gated pores, confirming the predictions of previous functional and structural investigations (reviewed in 87,94,108,139). As found for ClC-4, ClC-5, ClC-6 and ClC-7, the prokaryotic ClC homologue (ClC-ec1) and CmCLC function as H+/Cl antiporters, rather than as ion channels 151,99. The generation of monomers from dimeric ClC-ec1 has firmly established that each ClC subunit is a functional unit for transport and that cross-subunit interaction is not required for Cl-/H+ exchange in ClC transporters 141.

Subunits

Nomenclature ClC-1 ClC-2 ClC-Ka ClC-Kb
HGNC, UniProt CLCN1, P35523 CLCN2, P51788 CLCNKA, P51800 CLCNKB, P51801
Endogenous activators arachidonic acid
Activators (EC50) lubiprostone, omeprazole niflumic acid (1x10-5 – 1x10-3 M) niflumic acid (1x10-5 – 1x10-3 M)
Channel Blockers (IC50) 9-A-C, Cd2+, fenofibric acid, S-(-)CPB, S-(-)CPP, niflumic acid, Zn2+ Cd2+, DPC, NPPB, Zn2+, GaTx2 (Kd 1.5x10-11 M) [voltage dependent-100.0 mV] 3-phenyl-CPP, DIDS, niflumic acid (>1x10-3 M) 3-phenyl-CPP, DIDS
Functional characteristics γ = 1–1.5 pS; voltage-activated (depolarization) (by fast gating of single protopores and a slower common gate allowing both pores to open simultaneously); inwardly rectifying; incomplete deactivation upon repolarization, ATP binding to cytoplasmic cystathionine β-synthetase related (CBS) domains inhibits ClC-1 (by closure of the common gate), depending on its redox status γ = 2–3 pS; voltage-activated by membrane hyperpolarization by fast protopore and slow cooperative gating; channels only open negative to ECl resulting in steady-state inward rectification; voltage dependence modulated by permeant anions; activated by cell swelling, PKA, and weak extracellular acidosis; potentiated by SGK1; inhibited by phosphorylation by p34(cdc2)/cyclin B; cell surface expression and activity increased by association with Hsp90 γ = 26 pS; linear current-voltage relationship except at very negative potentials; no time dependence; inhibited by extracellular protons (pK = 7.1); potentiated by extracellular Ca2+ Bidirectional rectification; no time dependence; inhibited by extracellular protons; potentiated by extracellular Ca2+
Comment CIC-1 is constitutively active CIC-2 is also activated by amidation CIC-Ka is constitutively active (when co-expressed with barttin), and can be blocked by benzofuran derivatives CIC-Kb is constitutively active (when co-expressed with barttin), and can be blocked by benzofuran derivatives
Nomenclature ClC-3 ClC-4 ClC-5 ClC-6 ClC-7
HGNC, UniProt CLCN3, P51790 CLCN4, P51793 CLCN5, P51795 CLCN6, P51797 CLCN7, P51798
Channel Blockers (IC50) phloretin (3x10-5 M) Zn2+ (5x10-5 M) 131, Cd2+ (6.8x10-5 M) 131 DIDS (1x10-3 M) DIDS (4x10-5 M) 149, NS5818 (5.2x10-5 M) 149, NPPB (1.56x10-4 M) 149
Functional characteristics Cl-/H+ antiporter 121; pronounced outward rectification; slow activation, fast deactivation; activity enhanced by CaM kinase II; inhibited by intracellular Ins(3,4,5,6)P4 and extracellular acidosis Cl-/H+ antiporter (2Cl-:1H+) 79,134,146; extreme outward rectification; voltage-dependent gating with midpoint of activation at +73 mV 130; rapid activation and deactivation; inhibited by extracellular acidosis; non-hydrolytic nucleotide binding required for full activity Cl-/H+ antiporter (2Cl-:1H+) 134,146,151,159; extreme outward rectification; voltage-dependent gating with midpoint of activation of 116.0 mV; rapid activation and deactivation; potentiated and inhibited by intracellular and extracellular acidosis, respectively; ATP binding to cytoplasmic cystathionine β-synthetase related (CBS) domains activates ClC-5 Cl-/H+ antiporter (2Cl-:1H+) 125; outward rectification, rapid activation and deactivation Cl-/H+ antiporter (2Cl-:1H+) 103,114,149; strong outward rectification; voltage-dependent gating with a threshold more positive than ∼ + 20 mV; very slow activation and deactivation
Comment insensitive to the channel blockers DIDS, NPPB and tamoxifen (10 μM) insensitive to the channel blockers DIDS (1 mM), DPC (1 mM), 9-A-C (2 mM), NPPB (0.5 mM) and niflumic acid (1 mM) active when co-expressed with Ostm1

Comments

ClC channels display the permeability sequence Cl- > Br- > I- (at physiological pH). ClC-1 has significant opening probability at resting membrane potential, accounting for 75% of the membrane conductance at rest in skeletal muscle, and is important for stabilization of the membrane potential. S-(-)CPP, 9-A-C and niflumic acid act intracellularly and exhibit a strongly voltage-dependent block with strong inhibition at negative voltages and relief of block at depolarized potentials (115 and reviewed in 138). Inhibition of ClC-2 by the peptide GaTx2, from Leiurus quinquestriatus herbareus venom, is likely to occur through inhibition of channel gating, rather than direct open channel blockade 153. Although ClC-2 can be activated by cell swelling, it does not correspond to the VRAC channel (see below). Alternative potential physiological functions for ClC-2 are reviewed in 137. Functional expression of human ClC-Ka and ClC-Kb requires the presence of barttin 97,147 reviewed in 98. The properties of ClC-Ka/barttin and ClC-Kb/barttin tabulated are those observed in mammalian expression systems: in oocytes the channels display time- and voltage-dependent gating. The rodent homologue (ClC-K1) of ClC-Ka demonstrates limited expression as a homomer, but its function is enhanced by barttin which increases both channel opening probablility in the physiological range of potentials 97,101,147 reviewed in 98). ClC-Ka is approximately 5 to 6-fold more sensitive to block by 3-phenyl-CPP and DIDS than ClC-Kb, while newly synthesized benzofuran derivatives showed the same blocking affinity (<10 μM) on both CLC-K isoforms 116. The biophysical and pharmacological properties of ClC-3, and the relationship of the protein to the endogenous volume-regulated anion channel(s) VRAC 78,105 are controversial and further complicated by the possibility that ClC-3 may function as both a Cl-/H+ exchanger and an ion channel 78,134,156. The functional properties tabulated are those most consistent with the close structural relationship between ClC-3, ClC-4 and ClC-5. Activation of heterologously expressed ClC-3 by cell swelling in response to hypotonic solutions is disputed, as are many other aspects of its regulation. Dependent upon the predominant extracellular anion (e.g. SCN- versus Cl-), CIC-4 can operate in two transport modes: a slippage mode in which behaves as an ion channel and an exchanger mode in which unitary transport rate is 10-fold lower 79. Similar findings have been made for ClC-5 158. ClC-7 associates with a β subunit, Ostm1, which increases the stability of the former 112 and is essential for its function 114.

CFTR

Overview

CFTR, a 12TM, ABC transporter-type protein, is a cAMP-regulated epithelial cell membrane Cl- channel involved in normal fluid transport across various epithelia. Of the 1700 mutations identified in CFTR, the most common is the deletion mutant ΔF508 (a class 2 mutation) which results in impaired trafficking of CFTR and reduces its incorporation into the plasma membrane causing cystic fibrosis (reviewed in 88). Channels carrying the ΔF508 mutation that do traffic to the plasma membrane demonstrate gating defects. Thus, pharmacological restoration the function of the ΔF508 mutant would require a compound that embodies ‘corrector’ (i.e. facilitates folding and trafficking to the cell surface) and ‘potentiator’ (i.e. promotes opening of channels at the cell surface) activities 88. In addition to acting as an anion channel per se, CFTR may act as a regulator of several other conductances including inhibition of the epithelial Na channel (ENaC), calcium activated chloride channels (CaCC) and volume regulated anion channel (VRAC), activation of the outwardly rectifying chloride channel (ORCC), and enhancement of the sulphonylurea sensitivity of the renal outer medullary potassium channel (ROMK2), (reviewed in 126). CFTR also regulates TRPV4, which provides the Ca2+ signal for regulatory volume decrease in airway epithelia 81. The activities of CFTR and the chloride-bicarbonate exchangers SLC26A3 (DRA) and SLC26A6 (PAT1) are mutually enhanced by a physical association between the regulatory (R) domain of CFTR and the STAS domain of the SCL26 transporters, an effect facilitated by PKA-mediated phosphorylation of the R domain of CFTR 109.

Channels

Nomenclature HGNC, UniProt Activators (EC50) Channel Blockers (IC50) Functional characteristics Comment
CFTR CFTR, P13569 apigenin (Potentiation), capsaicin (Potentiation), CBIQ (Potentiation), felodipine (Potentiation), genistein (Potentiation), nimodipine (Potentiation), NS004 (Potentiation), phenylglycine-01 (Potentiation), SF-01 (Potentiation), UCCF-029 (Potentiation), UCCF-339 (Potentiation), UCCF-853 (Potentiation), VX-770 (Potentiation) intracellular CFTRinh-172 (intracellular application prolongs mean closed time), GaTx1, glibenclamide, extracellular GlyH-101 γ = 6–10 pS; permeability sequence = Br- ≥ Cl- > I- > F-, (PI/PCl = 0.1–0.85); slight outward rectification; phosphorylation necessary for activation by ATP binding at binding nucleotide binding domains (NBD)1 and 2; positively regulated by PKC and PKGII (tissue specific); regulated by several interacting proteins including syntaxin 1A, Munc18 and PDZ domain proteins such as NHERF (EBP50) and CAP70 UCCF-339, UCCF-029, apigenin and genistein are examples of flavones. UCCF-853 and NS004 are examples of benzimidazolones. CBIQ is an example of a benzoquinoline. felodipine and nimodipine are examples of 1,4-dihydropyridines. phenylglycine-01 is an example of a phenylglycine. SF-01 is an example of a sulfonamide. Malonic acid hydrazide conjugates are also CFTR channel blockers (see Verkman and Galietta, 135 155)

Comments

In addition to the agents listed in the table, the novel small molecule, ataluren, induces translational read through of nonsense mutations in CFTR (reviewed in 150). Corrector compounds that aid the folding of DF508CFTR to increase the amount of protein expressed and potentially delivered to the cell surface include VX-532 (which is also a potentiator), VRT-325, KM11060, Corr-3a and Corr-4a see 155 for details and structures of Corr-3a and Corr-4a). Inhibition of CFTR by intracellular application of the peptide GaTx1, from Leiurus quinquestriatus herbareus venom, occurs preferentially for the closed state of the channel 102. CFTR contains two cytoplasmic nucleotide binding domains (NBDs) that bind ATP. A single open-closing cycle is hypothesised to involve, in sequence: binding of ATP at the N-terminal NBD1, ATP binding to the C-terminal NBD2 leading to the formation of an intramolecular NBD1-NBD2 dimer associated with the open state, and subsequent ATP hydrolysis at NBD2 facilitating dissociation of the dimer and channel closing, and the initiation of a new gating cycle 80,122. Phosphorylation by PKA at sites within a cytoplasmic regulatory (R) domain facilitates the interaction of the two NBD domains. PKC (and PKGII within intestinal epithelial cells via guanylinstimulated cGMP formation) positively regulate CFTR activity.

Calcium activated chloride channel

Overview

Chloride channels activated by intracellular calcium (CaCC) are widely expressed in excitable and non-excitable cells where they perform diverse functions 106. The molecular nature of CaCC has been uncertain with both CLCA, TWEETY and BEST genes having been considered as likely candidates 92,107,117. It is now accepted that CLCA expression products are unlikely to form channels per se and probably function as cell adhesion proteins, or are secreted 133. Similarly, TWEETY gene products do not recapictulate the properties of endogenous CaCC. The bestrophins encoded by genes BEST1-4 have a topology more consistent with ion channels 107 and form chloride channels that are activated by physiological concentrations of Ca2+, but whether such activation is direct is not known 107. However, currents generated by bestrophin over-expression do not resemble native CaCC currents. The evidence for and against bestrophin proteins forming CaCC is critically reviewed by Duran et al. in their 2010 paper 92. Recently, a new gene family, TMEM16 (anoctamin) consisting of 10 members (TMEM16A-K; anoctamin 1–10) has been identified and there is firm evidence that some of these members form chloride channels 91,110. TMEM16A (anoctamin 1; Ano 1) produces Ca2+-activated Cl- currents with kinetics similar to native CaCC currents recorded from different cell types 86,142,148,157. Knockdown of TMEM16A greatly reduces currents mediated by calcium-activated chloride channels in submandibular gland cells 157 and smooth muscle cells from pulmonary artery 118. In TMEM16A(-/-) mice secretion of Ca2+-dependent Cl- secretion by several epithelia is reduced 132,142. Alternative splicing regulates the voltage- and Ca2+- dependence of TMEM16A and such processing may be tissue-specific manner and thus contribute to functional diversity 100. There are also reports that TMEM16B (anoctamin 2; Ano 2) supports CaCC activity (e.g. 135) and in TMEM16B(-/-) mice Ca-activated Cl- currents in the main olfactory epithelium (MOE) and in the vomeronasal organ are virtually absent 85.

Subunits

Nomenclature HGNC, UniProt Endogenous activators (EC50) Endogenous channel blockers (IC50) Channel Blockers (IC50) Functional characteristics
CaCC ANO1, Q5XXA6 intracellular Ca2+ Ins(3,4,5,6)P4 9-A-C, DCDPC, DIDS, flufenamic acid, fluoxetine, mibefradil, niflumic acid, NPPB, SITS, tannic acid γ = 0.5–5 pS; permeability sequence, SCN- > NO3-> I- > Br- > Cl- > F-; relative permeability of SCN-:Cl- ∼8. I-:Cl- ∼3, aspartate:Cl- ∼0.15, outward rectification (decreased by increasing [Ca2+]i); sensitivity to activation by [Ca2+]i decreased at hyperpolarized potentials; slow activation at positive potentials (accelerated by increasing [Ca2+]i); rapid deactivation at negative potentials, deactivation kinetics modulated by anions binding to an external site; modulated by redox status

Comments

Blockade of ICl(Ca) by niflumic acid, DIDS and 9-A-C is voltage-dependent whereas block by NPPB is voltage-independent 106. Extracellular niflumic acid; DCDPC and 9-A-C (but not DIDS) exert a complex effect upon ICl(Ca) in vascular smooth muscle, enhancing and inhibiting inwardly and outwardly directed currents in a manner dependent upon [Ca2+]i (see 113 for summary). Considerable crossover in pharmacology with large conductance Ca2+-activated K+ channels also exists (see 104 for overview). Two novel compounds, CaCCinh-A01 and CaCCinh-B01 have recently been identified as blockers of calcium-activated chloride channels in T84 human intestinal epithelial cells 89 for structures). Significantly, other novel compounds totally block currents mediated by TMEM116A, but have only a modest effect upon total current mediated by CaCC native to T84 cells or human bronchial epithelial cells, suggesting that TMEM16A is not the predominant CaCC in such cells 124. CaMKII modulates CaCC in a tissue dependent manner (reviewed by 106,113). CaMKII inhibitors block activation of ICl(Ca) in T84 cells but have no effect in parotid acinar cells. In tracheal and arterial smooth muscle cells, but not portal vein myocytes, inhibition of CaMKII reduces inactivation of ICl(Ca). Intracellular Ins(3,4,5,6)P4may act as an endogenous negative regulator of CaCC channels activated by Ca2+, or CaMKII. Smooth muscle CaCC are also regulated positively by Ca2+-dependent phosphatase, calcineurin (see 113 for summary).

Maxi chloride channel

Overview

Maxi Cl- channels are high conductance, anion selective, channels initially characterised in skeletal muscle and subsequently found in many cell types including neurones, glia, cardiac muscle, lymphocytes, secreting and absorbing epithelia, macula densa cells of the kidney and human placenta syncytiotrophoblasts 144. The physiological significance of the maxi Cl- channel is uncertain, but roles in cell volume regulation and apoptosis have been claimed. Evidence suggests a role for maxi Cl- channels as a conductive pathway in the swelling-induced release of ATP from mouse mammary C127i cells that may be important for autocrine and paracrine signalling by purines 93,143. A similar channel mediates ATP release from macula densa cells within the thick ascending of the loop of Henle in response to changes in luminal NaCl concentration 83. A family of human high conductance Cl- channels (TTYH1-3) that resemble Maxi Cl- channels has been cloned 152, but alternatively, Maxi Cl- channels have also been suggested to correspond to the voltage-dependent anion channel, VDAC, expressed at the plasma membrane 82,128.

Channels

Nomenclature Activators (EC50) Endogenous channel blockers (IC50) Channel Blockers (IC50) Functional characteristics Comment
Maxi Cl- extracellular chlorpromazine, cytosolic GTPγS, extracellular tamoxifen, extracellular toremifene, extracellular triflupromazine intracellular arachidonic acid DPC, extracellular Gd3+, SITS, DIDS (4x10-5 M) 149, extracellular Zn2+ (5x10-5 M) 131, NPPB (1.56x10-4 M) 149 γ = 280–430 pS (main state); permeability sequence, I > Br > Cl > F > gluconate (PCIPCl = ∼1.5); ATP is a voltage dependent permeant blocker of single channel activity (PATP/PCl = 0.08–0.1); channel activity increased by patch-excision; channel opening probability (at steady-state) maximal within approximately ± 20 mV of 0 mV, opening probability decreased at more negative and (commonly) positive potentials yielding a bell-shaped curve; channel conductance and opening probability regulated by annexin 6 Maxi Cl- is also activated by G protein-coupled receptors and cell swelling. tamoxifen and toremifene are examples of triphenylethylene anti-oestrogens

Comments

Differing ionic conditions may contribute to variable estimates of γ reported in the literature. Inhibition by arachidonic acid (and cis-unsaturated fatty acids) is voltage-independent, occurs at an intracellular site, and involves both channel shut down (Kd = 4–5 μM) and a reduction of γ (Kd = 13–14 μM). Blockade of channel activity by SITS, DIDS, Gd3+ and arachidonic acid is paralleled by decreased swelling-induced release of ATP 93,143. Channel activation by anti-oestrogens in whole cell recordings requires the presence of intracellular nucleotides and is prevented by pre-treatment with 17β-estradiol, dibutyryl cAMP, or intracellular dialysis with GDPβS 90. Activation by tamoxifen is suppressed by low concentrations of okadaic acid, suggesting that a dephosphorylation event by protein phosphatase PP2A occurs in the activation pathway 90. In contrast, 17β-estradiol and tamoxifen appear to directly inhibit the maxi Cl- channel of human placenta reconstituted into giant liposomes and recorded in excised patches 140.

Volume regulated chloride channels

Overview

Volume activated chloride channels (also termed VSOAC, volume-sensitive organic osmolyte/anion channel; VRC, volume regulated channel and VSOR, volume expansion-sensing outwardly rectifying anion channel) participate in regulatory volume decrease (RVD) in response to cell swelling. VRAC may also be important for several other processes including the regulation of membrane excitability, transcellular Cl- transport, angiogenesis, cell proliferation, necrosis, apoptosis, glutamate release from astrocytes, insulin (INS, P01308) release from pancreatic β cells and resistance to the anti-cancer drug, cisplatin (reviewed by 84,123,126,129). VRAC may not be a single entity, but may instead represent a number of different channels that are expressed to a variable extent in different tissues and are differentially activated by cell swelling. In addition to ClC-3 expression products (see above) several former VRAC candidates including MDR1 P-glycoprotein, Icln, Band 3 anion exchanger and phospholemman are also no longer considered likely to fulfil this function (see reviews 126,145).

Channels

Nomenclature Activators (EC50) Endogenous channel blockers (IC50) Channel Blockers (IC50) Functional characteristics Comment
VRAC GTPγS arachidonic acid, extracellular Mg2+ 1,9-dideoxyforskolin, 9-A-C, carbenoxolone, clomiphene, DCPIB, diBA-(5)-C4, DIDS, gossypol, IAA-94, mefloquine, mibefradil, nafoxidine, NDGA, NPPB, NS3728, quinidine, quinine, tamoxifen γ = 10–20 pS (negative potentials), 50–90 pS (positive potentials); permeability sequence SCN > I > NO3- >Br- > Cl- > F- > gluconate; outward rectification due to voltage dependence of γ; inactivates at positive potentials in many, but not all, cell types; time dependent inactivation at positive potentials; intracellular ionic strength modulates sensitivity to cell swelling and rate of channel activation; rate of swelling-induced activation is modulated by intracellular ATP concentration; ATP dependence is independent of hydrolysis and modulated by rate of cell swelling; inhibited by increased intracellular free Mg2+ concentration; swelling induced activation of several intracellular signalling cascades may be permissive of, but not essential to, the activation of VRAC including: the Rho-Rho kinase-MLCK; Ras-Raf-MEK-ERK; PIK3-NOX-H2O2 and Src-PLCγ-Ca2+ pathways; regulation by PKCα required for optimal activity; cholesterol depletion enhances activity; activated by direct stretch of β1-integrin VRAC is also activated by cell swelling and low intracellular ionic strength. VRAC is also blocked by chromones, extracellular nucleotides and nucleoside analogues

Comments

In addition to conducting monovalent anions, in many cell types the activation of VRAC by a hypotonic stimulus can allow the efflux of organic osmolytes such as amino acids and polyols that may contribute to RVD.

Other chloride channels

In addition to some intracellular chloride channels that are not considered here, plasma membrane channels other than those listed have been functionally described. Many cells and tissues contain outwardly rectifying chloride channels (ORCC) that may correspond to VRAC active under isotonic conditions. A cAMP-activated Cl- channel that does not correspond to CFTR has been described in intestinal Paneth cells 154. A Cl channel activated by cGMP with a dependence on raised intracellular Ca2+ has been recorded in various vascular smooth muscle cells types, which has a pharmacology and biophysical characteristics very different from the ‘conventional’ CaCC 119,136. It has been proposed that bestrophin-3 (BEST3, Q8N1M1) is an essential component of the cGMP-activated channel 120. A proton-activated, outwardly rectifying anion channel has also been described 111.

Connexins and Pannexins

Overview

Gap junctions are essential for many physiological processes including cardiac and smooth muscle contraction, regulation of neuronal excitability and epithelial electrolyte transport 162164. Gap junction channels allow the passive diffusion of molecules of up to 1,000 Daltons which can include nutrients, metabolites and second messengers (such as IP3) as well as cations and anions. 21 connexin genes (Cx23, Cx25, Cx26, Cx30, Cx30.2, Cx30.3, Cx31, Cx31.1, Cx31.9, Cx32, Cx36, Cx37, Cx40, Cx40.1, Cx43, Cx45, Cx46, Cx47, Cx50, Cx59, Cx62) and 3 pannexin genes (Px1, Px2, Px3; which are structurally related to the invertebrate innexin genes) code for gap junction proteins (provisional nomenclature) in humans. Each connexin gap junction comprises 2 hemichannels or ‘connexons’ which are themselves formed from 6 connexin molecules. The various connexins have been observed to combine into both homomeric and heteromeric combinations, each of which may exhibit different functional properties. It is also suggested that individual hemichannels formed by a number of different connexins might be functional in at least some cells 165. Connexins have a common topology, with four α-helical transmembrane domains, two extracellular loops, a cytoplasmic loop, and N- and C-termini located on the cytoplasmic membrane face. In mice, the most abundant connexins in electrical synapses in the brain seem to be Cx36, Cx45 and Cx57 168. Mutations in connexin genes are associated with the occurrence of a number of pathologies, such as peripheral neuropathies, cardiovascular diseases and hereditary deafness. The pannexin genes Px1 and Px2 are widely expressed in the mammalian brain 169. Like the connexins, at least some of the pannexins can form hemichannels 162,166.

Subunits

Nomenclature Cx23, Cx25, Cx26, Cx30, Cx30.2, Cx30.3, Cx31, Cx31.1, Cx31.9, Cx32, Cx36, Cx37, Cx40, Cx40.1, Cx43, Cx45, Cx46, Cx47, Cx50, Cx59, Cx62 Px1, Px2, Px3
HGNC, UniProt GJE1, A6NN92; GJB7, Q6PEY0; GJB2, P29033; GJB6, O95452; GJC3, Q8NFK1; GJB4, Q9NTQ9; GJB3, O75712; GJB5, O95377; GJD3, Q8N144; GJB1, P08034; GJD2, Q9UKL4; GJA4, P35212; GJA5, P36382; GJD4, Q96KN9; GJA1, P17302; GJC1, P36383; GJA3, Q9Y6H8; GJC2, Q5T442; GJA8, P48165; GJA9, P57773; GJA10, Q969M2 PANX1, Q96RD7; PANX2, Q96RD6; PANX3, Q96QZ0
Endogenous inhibitors extracellular Ca2+ (blocked by raising external Ca2+)
Inhibitors carbenoxolone, flufenamic acid, octanol carbenoxolone, flufenamic acid (little block by flufenamic acid)
Comment The pannexins are unaffected by raising external Ca2+

Comments

Connexins are most commonly named according to their molecular weights, so, for example, Cx23 is the connexin protein of 23 kDa. This can cause confusion when comparing between species – for example, the mouse connexin Cx57 is orthologous to the human connexin Cx62. No natural toxin or specific inhibitor of junctional channels has been identified yet however two compounds often used experimentally to block connexins are carbenoxolone and flufenamic acid 167. At least some pannexin hemichannels are more sensitive to carbenoxolone than connexins but much less sensitive to flufenamic acid 161. It has been suggested that 2-aminoethoxydiphenyl borate (2-APB) may be a more effective blocker of some connexin channel subtypes (Cx26, Cx30, Cx36, Cx40, Cx45, Cx50) compared to others (Cx32, Cx43, Cx46, 160).

Cyclic nucleotide-regulated channels

Overview

Cyclic nucleotide-gated (CNG) channels are responsible for signalling in the primary sensory cells of the vertebrate visual and olfactory systems. A standardised nomenclature for CNG channels has been proposed by the NC-IUPHAR subcommittee on voltage-gated ion channels 176.

CNG channels are voltage-independent cation channels formed as tetramers. Each subunit has 6TM, with the pore-forming domain between TM5 and TM6. CNG channels were first found in rod photoreceptors 175,177, where light signals through rhodopsin and transducin to stimulate phosphodiesterase and reduce intracellular cGMP level. This results in a closure of CNG channels and a reduced ‘dark current’. Similar channels were found in the cilia of olfactory neurons 178 and the pineal gland 174. The cyclic nucleotides bind to a domain in the C terminus of the subunit protein: other channels directly binding cyclic nucleotides include HCN, eag and certain plant potassium channels.

Subunits

Nomenclature CNGA1 CNGA2 CNGA3 CNGA4 CNGB1 CNGB3
HGNC, UniProt CNGA1, P29973 CNGA2, Q16280 CNGA3, Q16281 CNGA4, Q8IV77 CNGB1, Q14028 CNGB3, Q9NQW8
Activators cGMP (EC50 ∼ 30 μM) >> cAMP cGMP ∼ cAMP (EC50 ∼ 1 μM) cGMP (EC50 ∼ 30 μM) >> cAMP
Inhibitors L-(cis)-diltiazem L-(cis)-diltiazem
Functional characteristics γ = 25–30 pS, PCa/PNa = 3.1 γ = 35 pS, PCa/PNa = 6.8 γ = 40 pS, PCa/PNa = 10.9

Comments

CNGA1, CNGA2 and CNGA3 express functional channels as homomers. Three additional subunits CNGA4 (Q8IV77), CNGB1 (Q14028) and CNGB3 (Q9NQW8) do not, and are referred to as auxiliary subunits. The subunit composition of the native channels is believed to be as follows. Rod: CNGA13/CNGB1a; Cone: CNGA32/CNGB32; Olfactory neurons: CNGA22/CNGA4/CNGB1b 180184.

Hyperpolarisation-activated, cyclic nucleotide-gated (HCN)

The hyperpolarisation-activated, cyclic nucleotide-gated (HCN) channels are cation channels that are activated by hyperpolarisation at voltages negative to ∼-50 mV. The cyclic nucleotides cAMP and cGMP directly activate the channels and shift the activation curves of HCN channels to more positive voltages, thereby enhancing channel activity. HCN channels underlie pacemaker currents found in many excitable cells including cardiac cells and neurons 173,179. In native cells, these currents have a variety of names, such as Ih, Iq and If. The four known HCN channels have six transmembrane domains and form tetramers. It is believed that the channels can form heteromers with each other, as has been shown for HCN1 and HCN4 170. A standardised nomenclature for HCN channels has been proposed by the NC-IUPHAR subcommittee on voltage-gated ion channels 176.

Nomenclature HCN1 HCN2 HCN3 HCN4
HGNC, UniProt HCN1, O60741 HCN2, Q9UL51 HCN3, Q9P1Z3 HCN4, Q9Y3Q4
Activators cAMP > cGMP (both weak) cAMP > cGMP cAMP > cGMP
Inhibitors Cs+, ivabradine, ZD7288 Cs+, ivabradine, ZD7288 Cs+, ivabradine, ZD7288 Cs+, ivabradine, ZD7288

Comments

HCN channels are permeable to both Na+ and K+ ions, with a Na+/K+ permeability ratio of about 0.2. Functionally, they differ from each other in terms of time constant of activation with HCN1 the fastest, HCN4 the slowest and HCN2 and HCN3 intermediate. The compounds ZD7288 171 and ivabradine 172 have proven useful in identifying and studying functional HCN channels in native cells. zatebradine and cilobradine are also useful blocking agents.

Epithelial sodium channels (ENaC)

Overview

The epithelial sodium channels (ENaC) mediates sodium reabsorption in the aldosterone-sensitive distal part of the nephron and the collecting duct of the kidney. ENaC is found on other tight epithelial tissues such as the airways, distal colon and exocrine glands. ENaC activity is tightly regulated in the kidney by aldosterone, angiotensin II (AGT, P01019), vasopressin, insulin (INS, P01308) and glucocorticoids; this fine regulation of ENaC is essential to maintain sodium balance between daily intake and urinary excretion of sodium, circulating volume and blood pressure. ENaC expression is also vital for clearance of foetal lung fluid, and to maintain air-surface-liquid 195,199. Sodium reabsorption is suppressed by the ‘potassium-sparing’ diuretics amiloride and triamterene. ENaC is a heteromultimeric channel made of homologous α β and γ subunits. The primary structure of αENaC subunit was identified by expression cloning 188; β and γ ENaC were identified by functional complementation of the α subunit 189. Each ENaC subunit contains 2 TM α helices connected by a large extracellular loop and short cytoplasmic amino- and carboxy-termini. The stoichiometry of the epithelial sodium channel in the kidney and related epithelia is, by homology with the structurally related channel ASIC1a, thought to be a heterotrimer of 1α:1β:1γ subunits 193.

Channels

Nomenclature Subunits Activators (EC50) Channel Blockers (IC50) Functional characteristics
ENaCαβγ ENaC α, ENaC β, ENaC γ S3969 (1.2x10-6 M) 200 P552-02 (7.6x10-9 M), benzamil (∼1x10-8 M), amiloride (1x10-7 – 2x10-7 M), triamterene (∼5x10-6 M) 189,196 γ ≈ 4–5 pS, PNa/PK > 20; tonically open at rest; expression and ion flux regulated by circulating aldosterone-mediated changes in gene transcription. The action of aldosterone, which occurs in ‘early’ (1.5–3 h) and ‘late’ (6–24 hr) phases is competitively antagonised by spironolactone, its active metabolites and eplerenone. Glucocorticoids are important functional regulators in lung/airways and this control is potentiated by thyroid hormone; but the mechanism underlying such potentiation is unclear 185,206,209. The density of channels in the apical membrane, and hence GNa, can be controlled via both serum and glucocorticoid-regulated kinases (SGK1, 2 and 3) 190,191 and via cAMP/PKA 203; and these protein kinases appear to act by inactivating Nedd-4/2, a ubiquitin ligase that normally targets the ENaC channel complex for internalization and degradation 186,190. ENaC is constitutively activated by soluble and membrane-bound serine proteases, such as furin, prostasin (CAP1), plasmin and elastase 197,198,204,207,208. The activation of ENaC by proteases is blocked by a protein, SPLUNC1, secreted by the airways and which binds specifically to ENaC to prevent its cleavage 192. Pharmacological inhibitors of proteases (e.g. camostat acting upon prostasin) reduce the activity of ENaC 202. Phosphatidylinositides such as PtIns(4,5)P2 and PtIns(3,4,5)P3) stabilise channel gating probably by binding to the β and γ ENaC subunits, respectively 201,205, whilst C terminal phosphorylation of β and γ-ENaC by ERK1/2 has been reported to inhibit the withdrawal of the channel complex from the apical membrane 212. This effect may contribute to the cAMP-mediated increase in sodium conductance.

Comments

Data in the table refer to the αβγ heteromer. There are several human diseases resulting from mutations in ENaC subunits. Liddle's syndrome (including features of salt-sensitive hypertension and hypokalemia), is associated with gain of function mutations in the β and γ subunits leading to defective ENaC ubiquitylation and increased stability of active ENaC at the cell surface 208,210,211. Enzymes that deubiquitylate ENaC increase its function in vivo. Pseudohypoaldosteronism type 1 (PHA-1) can occur through either mutations in the gene encoding the mineralocorticoid receptor, or loss of function mutations in genes encoding ENaC subunits 187. Regulation of ENaC by phosphoinositides may underlie insulin (INS, P01308)-evoked renal Na+ retention that can complicate the clinical management of type 2 diabetes using insulin-sensitizing thiazolidinedione drugs 194.

IP3 receptor

Overview

The inositol 1,4,5-trisphosphate (IP3) receptors (provisional nomenclature) are ligand-gated Ca2+-release channels on intracellular Ca2+ store sites (such as the endoplasmic reticulum). They are responsible for the mobilization of intracellular Ca2+ stores and play an important role in intracellular Ca2+ signalling in a wide variety of cell types. Three different gene products (types I–III) have been isolated, which assemble as large tetrameric structures. IP3 receptors are closely associated with certain proteins: calmodulin (CALM2, CALM3, CALM1, P62158) and FKBP (and calcineurin via FKBP). They are phosphorylated by PKA, PKC, PKG and CaMKII.

Subunits

Nomenclature IP3R1 IP3R2 IP3R3
HGNC, UniProt ITPR1, Q14643 ITPR2, Q14571 ITPR3, Q14573
Endogenous activators (EC50) cytosolic ATP (< mM range), IP3 (endogenous; nM - μM range), cytosolic Ca2+ (Concentration range = < 7.5x10-4 M) cytosolic Ca2+ (nM range), IP3 (endogenous; nM - μM range) cytosolic Ca2+ (nM range), IP3 (endogenous; nM - μM range)
Activators (EC50) adenophostin A (pharmacological; nM range), Ins(2,4,5)P3 (pharmacological; also activated by other InsP3 analogues) adenophostin A (pharmacological; nM range), Ins(2,4,5)P3 (pharmacological; also activated by other InsP3 analogues)
Endogenous antagonists (IC50) heparin (μg/ml) heparin (μg/ml) heparin (μg/ml)
Antagonists (IC50) caffeine (mM range), decavanadate (μM range), PIP2 (μM range), xestospongin C (μM range) decavanadate (μM range) decavanadate (μM range)
Functional characteristics Ca2+: (PBa/PK ∼6) single-channel conductance, ∼70 pS (50 mM Ca2+) Ca2+: single-channel conductance, ∼70 pS (50 mM Ca2+), ∼390 pS (220 mM Cs+) Ca2+: single-channel conductance, ∼88 pS (55 mM Ba2+)
Comment IP3 R1 is also antagonised by calmodulin at high cytosolic Ca2+ concentrations

Comments

The absence of a modulator of a particular isoform of receptor indicates that the action of that modulator has not been determined, not that it is without effect.

Potassium channels

Overview

Potassium channels are fundamental regulators of excitability. They control the frequency and the shape of action potential waveform, the secretion of hormones and neurotransmitters and cell membrane potential. Their activity may be regulated by voltage, calcium and neurotransmitters (and the signalling pathways they stimulate). They consist of a primary pore-forming a subunit often associated with auxiliary regulatory subunits. Since there are over 70 different genes encoding K channels α subunits in the human genome, it is beyond the scope of this guide to treat each subunit individually. Instead, channels have been grouped into families and subfamilies based on their structural and functional properties. The three main families are the 2TM (two transmembrane domain), 4TM and 6TM families. A standardised nomenclature for potassium channels has been proposed by the NC-IUPHAR subcommittees on potassium channels 213216.

Inwardly rectifying potassium channels

Overview

The 2TM domain family of K channels are also known as the inward-rectifier K channel family. This family includes the strong inward-rectifier K channels (KIR2.x), the G-protein-activated inward-rectifier K channels (KIR3.x) and the ATP-sensitive K channels (KIR6.x, which combine with sulphonylurea receptors (SUR)). The pore-forming a subunits form tetramers, and heteromeric channels may be formed within subfamilies (e.g. KIR3.2 with KIR3.3).

Subunits

Nomenclature Kir1.1 Kir2.1, Kir2.2, Kir2.3, Kir2.4 Kir3.1, Kir3.2, Kir3.3, Kir3.4 Kir4.1, Kir4.2 Kir5.1 Kir6.1, Kir6.2 Kir7.1
HGNC, UniProt KCNJ1, P48048 KCNJ2, P63252; KCNJ12, Q14500; KCNJ4, P48050; KCNJ14, Q9UNX9 KCNJ3, P48549; KCNJ6, P48051; KCNJ9, Q92806; KCNJ5, P48544 KCNJ10, P78508; KCNJ15, Q99712 KCNJ16, Q9NPI9 KCNJ8, Q15842; KCNJ11, Q14654 KCNJ13, O60928
Endogenous inhibitors intracellular Mg2+
Endogenous activators PIP2
Associated subunits SUR1, SUR2A, SUR2B
Activators cromakalim, diazoxide, minoxidil, nicorandil
Inhibitors glibenclamide, tolbutamide
Functional characteristics Inward-rectifier current IK1 in heart, ‘strong’ inward–rectifier current G-protein-activated inward-rectifier current Inward-rectifier current Inward-rectifier current ATP-sensitive, inward-rectifier current Inward-rectifier current
Comment KIR2.1 is also inhibited by intracellular polyamines, KIR2.2 is also inhibited by intracellular polyamines, KIR2.3 is also inhibited by intracellular polyamines, KIR2.4 is also inhibited by intracellular polyamines KIR3.1 is also activated by Gβγ, KIR3.2 is also activated by Gβγ, KIR3.3 is also activated by Gβγ, KIR3.4 is also activated by Gβγ

Two-P potassium channels

Overview

The 4TM family of K channels are thought to underlie many leak currents in native cells. They are open at all voltages and regulated by a wide array of neurotransmitters and biochemical mediators. The primary pore-forming αsubunit contains two pore domains (indeed, they are often referred to as two-pore domain K channels or K2P) and so it is envisaged that they form functional dimers rather than the usual K channel tetramers. There is some evidence that they can form heterodimers within subfamilies (e.g. K2P3.1 with K2P9.1). There is no current, clear, consensus on nomenclature of 4TM K channels, nor on the division into subfamilies 213. The suggested division into subfamilies, below, is based on similarities in both structural and functional properties within subfamilies.

Subunits

Nomenclature K2P1.1, K2P6.1, K2P7.1 K2P2.1, K2P10.1, K2P4.1 K2P3.1, K2P9.1, K2P15.1 K2P16.1, K2P5.1, K2P17.1 K2P13.1, K2P12.1 K2P18.1
HGNC, UniProt KCNK1, O00180; KCNK6, Q9Y257; KCNK7, Q9Y2U2 KCNK2, O95069; KCNK10, P57789; KCNK4, Q9NYG8 KCNK3, O14649; KCNK9, Q9NPC2; KCNK15, Q9H427 KCNK16, Q96T55; KCNK5, O95279; KCNK17, Q96T54 KCNK13, Q9HB14; KCNK12, Q9HB15 KCNK18, Q7Z418
Endogenous activators (EC50) arachidonic acid
Activators (EC50) halothane, riluzole halothane
Inhibitors anandamide, ruthenium red halothane
Endogenous inhibitors arachidonic acid
Functional characteristics Background current Background current Background current Background current Background current Background current
Comment K2P1.1 is inhibited by acid pHi, K2P6.1 is inhibited by acid pHi, K2P7.1 is inhibited by acid pHi K2P2.1 is also activated by stretch, heat and acid pHi, K2P10.1 is also activated by stretch, heat and acid pHi, K2P4.1 is also activated by stretch, heat and acid pHi K2P3.1 is also activated by alkakine pHo and inhibited by acid pHo, K2P9.1 is also inhibited by acid pHo, K2P15.1 is inhibited by acid pHo K2P16.1 is activated by alkaline pHo, K2P5.1 is activated by alkaline pHo, K2P17.1 is activated by alkaline pHo

Comments

The K2P7.1, K2P15.1 and K2P12.1 subtypes, when expressed in isolation, are nonfunctional. All 4TM channels are insensitive to the classical potassium channel blockers tetraethylammonium and 4-aminopyridine, but are blocked to varying degrees by Ba2+ ions.

Voltage-gated and calcium-activated potassium channels

Overview

The 6TM family of K channels comprises the voltage-gated KV subfamilies, the KCNQ subfamily the EAG subfamily (which includes herg channels), the Ca2+-activated Slo subfamily (actually with 7TM) and the Ca2+-activated SK subfamily. As for the 2TM family, the pore-forming a subunits form tetramers and heteromeric channels may be formed within subfamilies (e.g. KV1.1 with KV1.2; KCNQ2 with KCNQ3).

Subunits

Nomenclature Kv1.1, Kv1.2, Kv1.3, Kv1.4, Kv1.5, Kv1.6, Kv1.7, Kv1.8 Kv2.1, Kv2.2 Kv3.1, Kv3.2, Kv3.3, Kv3.4 Kv4.1, Kv4.2, Kv4.3 Kv7.1, Kv7.2, Kv7.3, Kv7.4, Kv7.5 Kv10.1, Kv10.2, Kv11.1, Kv11.2, Kv11.3, Kv12.1, Kv12.2, Kv12.3 KCa1.1, KCa4.1, KCa4.2, KCa5.1 KCa2.1, KCa2.2, KCa2.3, KCa3.1
HGNC, UniProt KCNA1, Q09470; KCNA2, P16389; KCNA3, P22001; KCNA4, P22459; KCNA5, P22460; KCNA6, P17658; KCNA7, Q96RP8; KCNA10, Q16322 KCNB1, Q14721; KCNB2, Q92953 KCNC1, P48547; KCNC2, Q96PR1; KCNC3, Q14003; KCNC4, Q03721 KCND1, Q9NSA2; KCND2, Q9NZV8; KCND3, Q9UK17 KCNQ1, P51787; KCNQ2, O43526; KCNQ3, O43525; KCNQ4, P56696; KCNQ5, Q9NR82 KCNH1, O95259; KCNH5, Q8NCM2; KCNH2, Q12809; KCNH6, Q9H252; KCNH7, Q9NS40; KCNH8, Q96L42; KCNH3, Q9ULD8; KCNH4, Q9UQ05 KCNMA1, Q12791; KCNT1, Q5JUK3; KCNT2, Q6UVM3; KCNU1, A8MYU2 KCNN1, Q92952; KCNN2, Q9H2S1; KCNN3, Q9UGI6; KCNN4, O15554
Associated subunits Kv β1 and Kv β2 Kv5.1, Kv6.1–6.4, Kv8.1–8.2 and Kv9.1–9.3 MiRP2 is an associated subunit for Kv3.4 KChIP and KChAP minK and MiRP2 minK and MiRP1
Activators (EC50) retigabine NS004, NS1619
Inhibitors α-dendrotoxin, margatoxin, noxiustoxin, tetraethylammonium (potent)tetraethylammonium (moderate)4-aminopyridine (potent) tetraethylammonium (moderate) 4-aminopyridine (potent), tetraethylammonium (potent)sea anemone toxin BDS-I linopirdine, tetraethylammonium, XE991 astemizole, E4031, terfenadine charybdotoxin, iberiotoxin, tetraethylammonium apamin, charybdotoxin
Functional characteristics KV, KA KV KV, KA KA cardiac IK5, M current, M current cardiac IKR Maxi KCa KNa (slack & slick) SKCa, IKCa

Ryanodine receptor

Overview

The ryanodine receptors (RyRs, provisional nomenclature) are found on intracellular Ca2+ storage/release organelles. The family of RyR genes encodes three highly related Ca2+ release channels: RyR1, RyR2 and RyR3, which assemble as large tetrameric structures. These RyR channels are ubiquitously expressed in many types of cells and participate in a variety of important Ca2+ signaling phenomena (neurotransmission, secretion, etc.). In addition to the three mammalian isoforms described below, various nonmammalian isoforms of the ryanodine receptor have been identified 218. The function of the ryanodine receptor channels may also be influenced by closely associated proteins such as the tacrolimus (FK506)-binding protein, calmodulin 219, triadin, calsequestrin, junctin and sorcin, and by protein kinases and phosphatases.

Subunits

Nomenclature RyR1 RyR2 RyR3
HGNC, UniProt RYR1, P21817 RYR2, Q92736 RYR3, Q15413
Endogenous activators (EC50) cytosolic ATP (endogenous; mM range), luminal Ca2+ (endogenous), cytosolic Ca2+ (endogenous; μM range) cytosolic ATP (endogenous; mM range), luminal Ca2+ (endogenous), cytosolic Ca2+ (endogenous; μM range) cytosolic ATP (endogenous; mM range), cytosolic Ca2+ (endogenous; μM range)
Activators (EC50) caffeine (pharmacological; mM range), ryanodine (pharmacological; nM - μM range), suramin (pharmacological; μM range) caffeine (pharmacological; mM range), ryanodine (pharmacological; nM - μM range), suramin (pharmacological; μM range) caffeine (pharmacological; mM range), ryanodine (pharmacological; nM - μM range)
Endogenous antagonists (IC50) cytosolic Mg2+ (mM range), cytosolic Ca2+ (Concentration range = > 1x10-4 M) cytosolic Mg2+ (mM range), cytosolic Ca2+ (Concentration range = > 1x10-3 M) cytosolic Mg2+ (mM range), cytosolic Ca2+ (Concentration range = > 1x10-3 M)
Antagonists (IC50) dantrolene dantrolene
Channel Blockers (IC50) procaine, ruthenium red, ryanodine (Concentration range = > 1x10-4 M) procaine, ruthenium red, ryanodine (Concentration range = > 1x10-4 M) ruthenium red
Functional characteristics Ca2+: (P Ca/P K∼6) single-channel conductance: ∼90 pS (50mM Ca2+), 770 pS (200 mM K+) Ca2+: (P Ca/P K∼6) single-channel conductance: ∼90 pS (50mM Ca2+), 720 pS (210 mM K+) Ca2+: (P Ca/PK∼6) single-channel conductance: ∼140 pS (50mM Ca2+), 777 pS (250 mM K+)
Comment RyR1 is also activated by depolarisation via DHP receptor, calmodulin at low cytosolic Ca2+ concentrations, CaM kinase and PKA; antagonised by calmodulin at high cytosolic Ca2+ concentrations RyR2 is also activated by CaM kinase and PKA; antagonised by calmodulin at high cytosolic Ca2+ concentrations RyR3 is also activated by calmodulin at low cytosolic Ca2+ concentrations; antagonised by calmodulin at high cytosolic Ca2+ concentrations

Comments

The modulators of channel function included in this table are those most commonly used to identify ryanodine-sensitive Ca2+ release pathways. Numerous other modulators of ryanodine receptor/channel function can be found in the reviews listed below. The absence of a modulator of a particular isoform of receptor indicates that the action of that modulator has not been determined, not that it is without effect. The potential role of cyclic ADP ribose as an endogenous regulator of ryanodine receptor channels is controversial. A region of RyR likely to be involved in ion translocation and selection has been identified 217,220.

Sodium leak channel, non-selective

Overview

The sodium leak channel, non selective (NC-IUPHAR tentatively recommends the nomenclature NaVi2.1) is structurally a member of the family of voltage-gated sodium channel family (Nav1.1 – Nav1.9) 221,228. In contrast to the latter, NaVi2.1, is voltage-insensitive (denoted in the subscript ‘vi’ in the tentative nomenclature) and possesses distinctive ion selectivity and pharmacological properties. NaVi2.1, which is insensitive to tetrodotoxin (10 μM), has been proposed to mediate the tetrodotoxin-resistant and voltage-insensitive Na+ leak current (IL-Na) observed in many types of neurone 222. However, whether NaVi2.1 is constitutively active has been challenged 226. NaVi2.1 is widely distributed within the central nervous system and is also expressed in the heart and pancreas specifically, in rodents, within the islets of Langerhans 221,222.

Subunits

Nomenclature HGNC, UniProt Activators Channel Blockers (IC50) Functional characteristics
Navi2.1 NALCN, Q8IZF0 Constitutively active (Lu et al., 2007), or activated downstream of Src family tyrosine kinases (SFKs) (Lu et al., 2009; Swayne et al., 2009); positively modulated by decreased extracellular Ca2+ concentration (Lu et al., 2010) 222224,226 Gd3+ (1.4x10-6 M), Cd2+ (1.5x10-4 M), Co2+ (2.6x10-4 M), verapamil (3.8x10-4 M) γ = 27 pS (by fluctuation analysis), PNa/PCs = 1.3, PK/PCs = 1.2, PCa/PCs = 0.5, linear current voltage-relationship, voltage-independent and non-inactivating

Comments

In native and recombinant expression systems NaVi2.1 can be activated by stimulation of NK1 (in hippocampal neurones), neurotensin (in ventral tegmental area neurones) and M3 muscarinic acetylcholine receptors (in MIN6 pancreatic β-cells) and in a manner that is independent of signalling through G-proteins 223,226. Pharmacological and molecular biological evidence indicates such modulation to occur though a pathway that involves the activation of Src family tyrosine kinases. It is suggested that NaVi2.1 exists as a macromolecular complex with M3 receptors 226 and peptide receptors 223, in the latter instance in association with the protein UNC-80, which recruits Src to the channel complex 223,227. By contrast, stimulation of Navi2.1 by decreased extracellular Ca2+ concentration is G-protein dependent and involves a Ca2+-sensing G protein-coupled receptor and UNC80 which links Navi2.1 to the protein UNC79 in the same complex 224. NaVi2.1 null mutant mice have severe disturbances in respiratory rhythm and die within 24 hours of birth 222. Navi2.1 heterozygous knockout mice display increased serum sodium concentrations in comparison to wildtype littermates and a role for the channel in osmoregulation has been postulated 225.

Transient receptor potential channels

Overview

The TRP superfamily of channels (nomenclature agreed by NC-IUPHAR; 244,351), whose founder member is the Drosophila Trp channel, exists in mammals as six families; TRPC, TRPM, TRPV, TRPA, TRPP and TRPML based on amino acid homologies. TRP subunits contain six putative transmembrane domains and assemble as homo- or hetero-tetramers to form cation selective channels with diverse modes of activation and varied permeation properties (reviewed by 314). Established, or potential, physiological functions of the individual members of the TRP families are discussed in detail in the recommended reviews and a compilation edited by Islam 271. The established, or potential, involvement of TRP channels in disease is reviewed in 279,307 and 308, together with a special edition of Biochemica et Biophysica Acta on the subject 307. The pharmacology of most TRP channels is poorly developed 351. Broad spectrum agents are listed in the tables along with more selective, or recently recognised, ligands that are flagged by the inclusion of a primary reference. Most TRP channels are regulated by phosphoinostides such as PtIns(4,5)P2 and IP3 although the effects reported are often complex, occasionally contradictory, and likely be dependent upon experimental conditions (reviewed by 309,327,343). Such regulation is generally not included in the tables.

TRPA (ankyrin) family 

TRPA1 is the sole mammalian member of this group (reviewed by 259). In some 234,274,330,334, but not other 272,303, studies TRPA1 is activated by noxious cold. One study suggests that activation of TRPA1 is secondary to a cold-induced elevation of [Ca2+]i 357, but this has been refuted 274. Additionally, TRPA1 has been proposed to be a component of a mechanosensitive transduction channel of vertebrate hair cells 246,303, but TRPA1(-/-) mice demonstrate no impairment in hearing, or vestibular function 238,284. There is consensus that TRPA1 acts as a nociceptor for environmental irritants 235.

TRPC (canonical) family 

Members of the TRPC subfamily (reviewed by 229,230,241,242,258,277,316,324) fall into the subgroups outlined below. TRPC2 (not tabulated) is a pseudogene in man. It is generally accepted that all TRPC channels are activated downstream of Gq/11-coupled receptors, or receptor tyrosine kinases (reviewed by 321,337,351). A comprehensive listing of G-protein coupled receptors that activate TRPC channels is given in 229. Hetero-oligomeric complexes of TRPC channels and their association with proteins to form signalling complexes are detailed in 230 and 278. TRPC channels have frequently been proposed to act as store-operated channels (SOCs) (or compenents of mulimeric complexes that form SOCs), activated by depletion of intracellular calcium stores (reviewed by 230,243,317,322,329,354), but this is controversial. All members of the TRPC family are blocked by 2-APB and SKF96356 265,266. Activation of TRPC channels by lipids is discussed by 241.

TRPC1/C4/C5 subgroup 

TRPC4/C5 may be distinguished from other TRP channels by their potentiation by micromolar concentrations of La3+.

TRPC3/C6/C7 subgroup 

All members are activated by diacylglycerol independent of protein kinase C stimulation 266.

TRPM (melastatin) family 

Members of the TRPM subfamily (reviewed by 257,265,317,356) fall into the five subgroups outlined below.

TRPM1/M3 subgroup 

TRPM1 exists as five splice variants and is involved in normal melanocyte pigmentation 312 and is also a visual transduction channel in retinal ON bipolar cells 283. TRPM3 (reviewed by 313) exists as multiple splice variants four of which (mTRPM3α1, mTRPM3α2, hTRPM3a and hTRPM31325) have been characterised and found to differ significantly in their biophysical properties. TRPM3 has recently been found to contribute to the detection of noxious heat 346.

TRPM2 

TRPM2 functions as a sensor of redox status in cells and is also activated by heat (reviewed by 353). Numerous splice variants of TRPM2 exist which differ in their activation mechanisms 254.

TRPM4/5 subgroup 

TRPM4 and TRPM5 are thermosensitive and have the distinction within all TRP channels of being impermeable to Ca2+ 351. A splice variant of TRPM4 (i.e.TRPM4b) and TRPM5 are molecular candidates for endogenous calcium-activated cation (CAN) channels 262. TRPM4 has been shown to be an important regulator of Ca2+ entry in to mast cells 339 and dendritic cell migration 236. TRPM5 in taste receptor cells of the tongue appears essential for the transduction of sweet, amino acid and bitter stimuli 289.

TRPM6/7 subgroup 

TRPM6 and 7 combine channel and enzymatic activities (‘chanzymes’) and are involved in Mg2+ homeostasis (reviewed by 237,318,328).

TRPM8 

Is a channel activated by cooling and pharmacological agents evoking a ‘cool’ sensation and participates in the thermosensation of cold temperatures 240,245,252 reviewed by 282,291,302,344.

TRPML (mucolipin) family 

The TRPML family 323,325,355 consists of three mammalian members (TRPML1-3). TRPML channels are probably restricted to intracellular vesicles and mutations in the gene (MCOLN1) encoding TRPML1 (mucolipin-1) are the cause of the neurodegenerative disorder mucolipidosis type IV (MLIV) in man. TRPML1 is a cation selective ion channel that is important for sorting/transport of endosomes in the late endocytotic pathway and specifically fusion between late endosome-lysosome hybrid vesicles. TRPML2 (MCLN2) remains to be functionally characterised in detail. TRPML3 is important for hair cell maturation, stereocilia maturation and intracellular vesicle transport. A naturally occurring gain of function mutation in TRPML3 (i.e. A419P) results in the varitint waddler (Va) mouse phenotype (reviewed by 310,325).

TRPP (polycystin) family 

The TRPP family (reviewed by 249,251,260,268,349) subsumes the polycystins that are divided into two structurally distinct groups, polycystic kidney disease 1-like (PKD1-like) and polycystic kidney disease 2-like (PKD2-like). Members of the PKD1-like group, in mammals, include PKD1 (reclassified as TRPP1), PDKREJ, PKD1L1, PKD1L2 and PKD1L3. The PKD2-like members comprise PKD2, PKD2L1 and PKD2L2, which have renamed TRPP2, TRPP3 and TRPP5, respectively 300. PKDREJ (Q9NTG1), PKD1L1 (Q8TDX9), mouse PKD1L2 (Q7TN88), PKD1L3 (Q7Z443) and TRPP5 (PKD2L2, Q9NZM6) are not listed in the table due to lack of functional data. Similarly, TRPP1 (P98161) is also omitted because although one study 233 has reported the induction of a cation conductance in CHO cells transfected with TRPP1, there is no unequivocal evidence that TRPP1 is a channel per se and in other studies (e.g. 250,264) TRPP1 is incapable of producing currents.

TRPV (vanilloid) family 

Members of the TRPV family (reviewed by 340) can broadly be divided into the theromosensitive, non-selective cation channels, TRPV1-4 and the calcium selective channels TRPV5 and TRPV6.

TRPV1-V4 subfamily 

TRPV1 is involved in the development of thermal hyperalgesia following inflammation and may contribute to the detection of noxius heat (reviewed by 320,333,335). Numerous splice variants of TRPV1 have been described, some of which modulate the activity of TRPV1, or act in a dominant negative manner when co-expressed with TRPV1 331. The pharmacology of TRPV1 channels is discussed in detail in 263 and 345. TRPV2 is probably not a thermosensor in man 315, but has recently been implicated in innate immunity 290. TRPV3 and TRPV4 are both thermosensitive, with the latter also having a mechanosensing function 255.

TRPV5/V6 subfamily 

Under physiological conditions, TRPV5 and TRPV6 are calcium selective channels involved in the absorption and reabsorption of calcium across intestinal and kidney tubule epithelia (reviewed by 248,348).

Subunits

Nomenclature TRPC1 TRPC4 TRPC5
HGNC, UniProt TRPC1, P48995 TRPC4, Q9UBN4 TRPC5, Q9UL62
Chemical activators NO-mediated cysteine S-nitrosylation
Physical activators membrane stretch (likely direct)
Other chemical activators NO-mediated cysteine S-nitrosylation, potentiation by extracellular protons NO-mediated cysteine S-nitrosylation (disputed), potentiation by extracellular protons
Physical activators membrane stretch (likely indirect)
Endogenous activators (EC50) lysophosphatidylcholine, intracellular Ca2+ (at negative potentials) (6.35x10-7 M)
Activators (EC50) La3+ (μM range) 7,4'-dihydroxyisoflavone, genistein (independent of tyrosine kinase inhibition) 350, La3+ (μM range), Gd3+ (Concentration range = 1x10-4 M), Pb2+ (Concentration range = 5x10-6 M)
Channel Blockers (IC50) 2-APB, Gd3+, GsMTx-4, La3+, SKF96365 2-APB, La3+ (mM range), niflumic acid, SKF96365, ML204 (2.9x10-6 M) 299 2-APB, BTP2, chlorpromazine, flufenamic acid, GsMTx-4, KB-R7943, La3+ (mM range), SKF96365, ML204 (∼1x10-5 M) 299
Functional characteristics γ = 16 pS (fluctuation analysis), conducts mono- and di-valent cations non-selectively; monovalent cation current suppressed by extracellular Ca2+; non-rectifying, or mildly inwardly rectifying; non-inactivating γ = 30 –41 pS, conducts mono and di-valent cations non-selectively (PCa/PNa = 1.1 – 7.7); dual (inward and outward) rectification γ = 41-63 pS; conducts mono-and di-valent cations non-selectively (PCa/PNa = 1.8 – 9.5); dual rectification (inward and outward) as a homomer, outwardly rectifying when expressed with TRPC1 or TRPC4
Nomenclature TRPC3 TRPC6 TRPC7
HGNC, UniProt TRPC3, Q13507 TRPC6, Q9Y210 TRPC7, Q9HCX4
Chemical activators diacylglycerols diacylglycerols
Other chemical activators diacylglycerols
Physical activators membrane stretch (likely indirect)
Endogenous activators (EC50) 20-HETE, arachidonic acid, lysophosphatidylcholine
Activators (EC50) 2,4 diahexanxoylphloroglucinol 287, flufenamate, hyperforin 288
Channel Blockers (IC50) 2-APB, ACAA, BTP2, Gd3+, KB-R7943, La3+, Ni2+, Pyr3 280, SKF96365 2-APB, ACAA, amiloride, Cd2+, Gd3+, GsMTx-4, Extracellular H+, KB-R7943, ML9, SKF96365, La3+ (∼6x10-6 M) 2-APB, amiloride, La3+, SKF96365
Functional characteristics γ = 66 pS; conducts mono and di-valent cations non-selectively (PCa/PNa = 1.6); monovalent cation current suppressed by extracellular Ca2+; dual (inward and outward) rectification γ = 28-37 pS; conducts mono and divalent cations with a preference for divalents (PCa/PNa = 4.5–5.0); monovalent cation current suppressed by extracellular Ca2+ and Mg2+, dual rectification (inward and outward), or inward rectification γ = 25–75 pS; conducts mono and divalent cations with a preference for divalents (PCa/ PCs = 5.9); modest outward rectification (monovalent cation current recorded in the absence of extracellular divalents); monovalent cation current suppressed by extracellular Ca2+ and Mg2+
Nomenclature TRPM1 TRPM3
HGNC, UniProt TRPM1, Q7Z4N2 TRPM3, Q9HCF6
Physical activators heat (Q10 = 7.2 between 15 - 25°C; Vriens et al., 2011), hypotonic cell swelling 346
Endogenous activators (EC50) pregnenolone sulphate 285 epipregnanolone sulphate 294, pregnenolone sulphate 347
Activators (EC50) dihydro-D-erythrosphingosine, nifedipine, sphingosine
Endogenous channel blockers (IC50) Zn2+ (1x10-6 M) intracellular Mg2+, extracellular Na+ (TRPM3α2 only)
Channel Blockers (IC50) 2-APB, Gd3+, La3+, mefenamic acid 281, pioglitazone (independent of PPAR-γ) 295, rosiglitazone, troglitazone
Functional characteristics Conducts mono- and di-valent cations non-selectively, dual rectification (inward and outward) TRPM31235: γ = 83 pS (Na+ current), 65 pS (Ca2+ current); conducts mono and di-valent cations non-selectively (PCa/PNa = 1.6) TRPM3α1: selective for monovalent cations (PCa/PCs∼0.1); TRPM3α2: conducts mono- and di-valent cations non-selectively (PCa/PCs = 1–10); Outwardly rectifying (magnitude varies between spice variants)
Nomenclature HGNC, UniProt Other chemical activators Physical activators Endogenous activators (EC50) Activators (EC50) Endogenous channel blockers (IC50) Channel Blockers (IC50) Functional characteristics
TRPM2 TRPM2, O94759 agents producing reactive oxygen (e.g. H2O2) and nitrogen (e.g. GEA 3162) species heat ∼ 35°C intracellular ADP ribose, arachidonic acid (Potentiation), intracellular cADPR, intracellular Ca2+ (via calmodulin), H2O2 GEA 3162 extracellular H+, Zn2+ (1x10-6 M) 2-APB, ACAA, clotrimazole, econazole, flufenamic acid, miconazole γ = 52-60 pS at negative potentials, 76 pS at positive potentials; conducts mono- and di-valent cations non-selectively (PCa/PNa = 0.6–0.7); non-rectifying; inactivation at negative potentials; activated by oxidative stress probably via PARP-1, PARP inhibitors reduce activation by oxidative stress, activation inhibited by suppression of APDR formation by glycohydrolase inhibitors
Nomenclature TRPM4 TRPM5
HGNC, UniProt TRPM4, Q8TD43 TRPM5, Q9NZQ8
Other channel blockers intracellular nucleotides including ATP, ADP, AMP and AMP-PNP with an IC50 range of 1.3–1.9 μM
Physical activators membrane depolarization (V½ = -20 mV to + 60 mV dependent upon conditions) in the presence of elevated [Ca2+]i, heat (Q10 = 8.5 @ +25 mV between 15 and 25°C) membrane depolarization (V½ = 0 to + 120 mV dependent upon conditions), heat (Q10 = 10.3 @ -75 mV between 15 and 25°C)
Endogenous activators (EC50) intracellular Ca2+ (transient activation of whole cell current) (3x10-7 – 2x10-5 M) intracellular Ca2+ (transient activation) (6.35x10-7 – 8.4x10-7 M)
Activators (EC50) BTP2 (Potentiation), decavanadate rosiglitazone 295
Channel Blockers (IC50) 9-phenanthrol, clotrimazole, flufenamic acid (2.8x10-6 M), intracellular spermine (3.5x10-5 – 6.1x10-5 M), adenosine (6.3x10-4 M) flufenamic acid (2.4x10-5 M), intracellular spermine (3.7x10-5 M), Extracellular H+ (6.3x10-4 M)
Functional characteristics γ = 23 pS (within the range 60 to +60 mV); permeable to monovalent cations; impermeable to Ca2+; strong outward rectification; slow activation at positive potentials, rapid deactivation at negative potentials, deactivation blocked by decavanadate γ = 15-25 pS; conducts monovalent cations selectively (PCa/PNa = 0.05); strong outward rectification; slow activation at positive potentials, rapid inactivation at negative potentials; activated and subsequently desensitized by [Ca2+]I
Comment TRPM5 is not blocked by ATP
Nomenclature TRPM6 TRPM7
HGNC, UniProt TRPM6, Q9BX84 TRPM7, Q96QT4
Other chemical activators constitutively active, activated by reduction of intracellular Mg2+ activation of PKA
Endogenous activators (EC50) extracellular H+ (Potentiation, μM range), intracellular Mg2+ intracellular ATP (Potentiation), cAMP (elevated cAMP levels), Extracellular H+ (Potentiation)
Activators (EC50) 2-APB (Potentiation) 2-APB (mM range)
Endogenous channel blockers (IC50) Mg2+ (inward current mediated by monovalent cations is blocked) (1.1x10-6 – 3.4x10-6 M), Ca2+ (inward current mediated by monovalent cations is blocked) (4.8x10-6 – 5.4x10-6 M) Mg2+
Channel Blockers (IC50) ruthenium red (1x10-7 M) [voltage dependent -120.0 mV] 2-APB (μM range), carvacrol, La3+, spermine (permeant blocker)
Functional characteristics γ = 40–87 pS; permeable to mono- and di-valent cations with a preference for divalents (Mg2+ > Ca2+; PCa/PNa = 6.9), conductance sequence Zn2+ > Ba2+ > Mg2+ = Ca2+ = Mn2+ > Sr2+ > Cd2+> Ni2+; strong outward rectification abolished by removal of extracellular divalents, inhibited by intracellular Mg2+ (IC50 = 0.5 mM) and ATP γ = 40-105 pS at negative and positive potentials respectively; conducts mono-and di-valent cations with a preference for monovalents (PCa/PNa = 0.34); conductance sequence Ni2+ > Zn2+ > Ba2+ = Mg2+ > Ca2+ = Mn2+ > Sr2+ > Cd2+; outward rectification, decreased by removal of extracellular divalent cations; inhibited by intracellular Mg2+, Ba2+, Sr2+, Zn2+, Mn2+ and Mg.ATP (disputed); activated by and intracellular alkalinization; sensitive to osmotic gradients
Nomenclature HGNC, UniProt Other chemical activators Physical activators Activators (EC50) Channel Blockers (IC50) Functional characteristics Comment
TRPM8 TRPM8, Q7Z2W7 agonist activities are temperature dependent and potentiated by cooling depolarization (V½ ∼ +50 mV at 15°C), cooling (< 22–26°C) icilin (requires intracellular Ca2+ as a co-factor for full agonist activity), (-)-menthol (inhibited by intracellular Ca2+), WS-12 2-APB, 5-benzyloxytryptamine, ACAA, AMTB 286, anandamide, BCTC, cannabidiol, capsazepine, clotrimazole, Δ9-tetrahydrocannabinol, La3+, linoleic acid, NADA γ = 40-83 pS at positive potentials; conducts mono- and di-valent cations non-selectively (PCa/PNa = 1.0–3.3); pronounced outward rectification; demonstrates densensitization to chemical agonists and adaptation to a cold stimulus in the presence of Ca2+; modulated by lysophospholipids and PUFAs cannabidiol and Δ9-tetrahydrocannabinol are examples of cannabinoids. TRPM8 is insensitive to ruthenium red

Comments

Ca2+ activates all splice variants of TRPM2, but other activators listed are effective only at the full length isoform 254. Inhibition of TRPM2 by clotrimazole, miconazole, econazole, flufenamic acid is largely irreversible. TRPM4 exists as multiple spice variants: data listed are for TRPM4b. The sensitivity of TRPM4b and TRPM5 to activation by [Ca2+]i demonstrates a pronounced and time-dependent reduction following excision of inside-out membrane patches 338. The V½ for activation of TRPM4 and TRPM5 demonstrates a pronounced negative shift with increasing temperature. Activation of TRPM8 by depolarization is strongly temperature-dependent via a channel-closing rate that decreases with decreasing temperature. The V½ is shifted in the hyperpolarizing direction both by decreasing temperature and by exogenous agonists, such as (-)-menthol 342 whereas antagonists produce depolarizing shifts in V½ 301. The V½ for the native channel is far more positive than that of heterologously expressed TRPM8 301. It should be noted that (-)-menthol and structurally related compounds can elicit release of Ca2+ from the endoplasmic reticulum independent of activation of TRPM8 293. Intracellular pH modulates activation of TRPM8 by cold and icilin, but not (-)-menthol 231.

Nomenclature HGNC, UniProt Other chemical activators Physical activators Activators (EC50) Channel Blockers (IC50) Functional characteristics
TRPA1 TRPA1, O75762 isothiocyanates (covalent) and 1,4-dihydropyridines (non-covalent) cooling (<17°C) (disputed) chlorobenzylidene malononitrile (Activation, covalent), cinnamaldehyde (Activation, covalent), formalin (Activation, covalent), icilin (Activation, non-covalent), (-)-menthol (Activation, non-covalent) (Concentration range = 1x10-6 - 1x10-4 M), thymol (Activation, non-covalent) (Concentration range = 1x10-6 - 1x10-4 M), acrolein (Agonist, covalent) (5.011x10-6 M) [Physiological voltage] 238, allicin (Agonist, covalent) (7.943x10-6 M) [Physiological voltage] 239, Δ9-tetrahydrocannabinol (Agonist, non-covalent) (1.259x10-5 M) [-60.0 mV] 272, nicotine (Activation, non-covalent) (∼2x10-5 M), URB597 (Agonist, non-covalent) (2.4x10-5 M) 306 ruthenium red (Inhibition) (<1x10-6 – 3x10-6 M), AP18 (Inhibition) (3.1x10-6 M) 319, HC030031 (Inhibition) (6.2x10-6 M) 296 γ = 87–100 pS; conducts mono- and di-valent cations non-selectively (PCa/PNa = 0.84); outward rectification; activated by elevated intracellular Ca2+

Comments

Agents activating TRPA1 in a covalent manner are thiol reactive electrophiles that bind to cysteine and lysine residues within the cytoplasmic domain of the channel 267,292. TRPA1 is activated by a wide range of endogenous and exogenous compounds and only a few representative examples are mentioned in the table: an exhaustive listing can be found in 235. In addition, TRPA1 is potently activated by intracellular zinc (EC50 = 8 nM) 232,269.

Nomenclature TRPV1 TRPV2 TRPV3 TRPV4
HGNC, UniProt TRPV1, Q8NER1 TRPV2, Q9Y5S1 TRPV3, Q8NET8 TRPV4, Q9HBA0
Other chemical activators NO-mediated cysteine S-nitrosylation NO-mediated cysteine S-nitrosylation epoxyeicosatrieonic acids and NO-mediated cysteine S-nitrosylation
Physical activators depolarization (V½ ∼ 0 mV at 35°C), noxious heat (> 43°C at pH 7.4) noxious heat (> 35°C; rodent, not human) 305 depolarization (V½ ∼ +80 mV, reduced to more negative values following heat stimuli), heat (23°C - 39°C, temperature threshold reduces with repeated heat challenge) constitutively active, heat (> 24°C - 32°C), mechanical stimuli
Endogenous activators (EC50) 12S-HPETE, 15S-HPETE, 5S-HETE, LTB4, Extracellular H+ (at 37°C) (3.98x10-6 M)
Activators (EC50) camphor, capsaicin, diphenylboronic anhydride, DkTx (Irreversible agonist), olvanil, phenylacetylrinvanil, resiniferatoxin Δ9-tetrahydrocannabinol, diphenylboronic anhydride, probenecid, 2-APB (1x10-5 M - Rat) 305,326, cannabidiol (Activation) (3.17x10-5 M) 326 2-APB, 6-tert-butyl-m-cresol, camphor, carvacrol, diphenylboronic anhydride, eugenol, incensole acetate, (-)-menthol, thymol 4α-PDD, 4α-PDH, bisandrographolide, phorbol 12-myristate 13-acetate, GSK1016790A (2.1x10-9 M) 336, RN1747 (7.7x10-7 M) 341
Selective activators (EC50) 2-APB (Agonist) (Mouse) [Physiological voltage] 270,273,297,298
Channel Blockers (IC50) 2-APB, allicin, anandamide, NADA, SB452533, AMG517 (9x10-10 M), 5'-iodoresiniferatoxin (3.9x10-9 M), AMG628 (3.7x10-9 M), SB705498 (3x10-9 – 6x10-9 M), A425619 (5x10-9 M), A778317 (5x10-9 M), 6-iodo-nordihydrocapsaicin (1x10-8 M), JYL1421 (9.2x10-9 M), BCTC (6x10-9 – 3.5x10-8 M), SB366791 (1.8x10-8 M), JNJ17203212 (6.5x10-8 M), capsazepine (4x10-8 – 2.8x10-7 M), ruthenium red (9x10-8 – 2.2x10-7 M) amiloride, La3+, SKF96365, TRIM, ruthenium red (6x10-7 M) diphenyltetrahydrofuran (Concentration range = 6x10-6 - 1x10-5 M), ruthenium red (Concentration range = < 1x10-6 M) Gd3+, La3+, ruthenium red [voltage dependent], HC067047 (1.7x10-8 M) 256, RN1734 (2.3x10-6 M) 341
Radioligands (Kd) [125I]resiniferatoxin, [3H]resiniferatoxin, [3H]A778317 (3.4x10-9 M)
Functional characteristics γ = 35 pS at – 60 mV; 77 pS at + 60 mV, conducts mono and di-valent cations with a selectivity for divalents (PCa/PNa = 9.6); voltage- and time- dependent outward rectification; potentiated by ethanol; activated/potentiated/upregulated by PKC stimulation; extracellular acidification facilitates activation by PKC; desensitisation inhibited by PKA; inhibited by Ca2+/ calmodulin; cooling reduces vanilloid-evoked currents; may be tonically active at body temperature Conducts mono- and di-valent cations (PCa/PNa = 0.9–2.9); dual (inward and outward) rectification; current increases upon repetitive activation by heat; translocates to cell surface in response to IGF-1 to induce a constitutively active conductance, translocates to the cell surface in response to membrane stretch γ = 197 pS at = +40 to +80 mV, 48 pS at negative potentials; conducts mono- and di-valent cations; outward rectification; potentiated by arachidonic acid γ = ∼60 pS at –60 mV, ∼90-100 pS at +60 mV; conducts mono- and di-valent cations with a preference for divalents (PCa/PNa = 6–10); dual (inward and outward) rectification; potentiated by intracellular Ca2+ via Ca2+/ calmodulin; inhibited by elevated intracellular Ca2+ via an unknown mechanism (IC50 = 0.4 μM)
Nomenclature TRPV5 TRPV6
HGNC, UniProt TRPV5, Q9NQA5 TRPV6, Q9H1D0
Activators constitutively active (with strong buffering of intracellular Ca2+) constitutively active (with strong buffering of intracellular Ca2+)
Activators (EC50) 2-APB (Potentiation)
Other channel blockers Pb2+ = Cu2+ = Gd3+ > Cd2+ > Zn2+ > La3+ > Co2+ > Fe2+
Channel Blockers (IC50) econazole, Mg2+, miconazole, ruthenium red (1.21x10-7 M) Cd2+, La3+, Mg2+, ruthenium red (9x10-6 M)
Functional characteristics γ = 59–78 pS for monovalent ions at negative potentials, conducts mono- and di-valents with high selectivity for divalents (PCa/PNa > 107); voltage- and time- dependent inward rectification; inhibited by intracellular Ca2+ promoting fast inactivation and slow downregulation; feedback inhibition by Ca2+ reduced by calcium binding protein 80-K-H; inhibited by extracellular and intracellular acidosis; upregulated by 1,25-dihydrovitamin D3 γ = 58–79 pS for monovalent ions at negative potentials, conducts mono- and di-valents with high selectivity for divalents (PCa/PNa > 130); voltage- and time-dependent inward rectification; inhibited by intracellular Ca2+ promoting fast and slow inactivation; gated by voltage-dependent channel blockade by intracellular Mg2+; slow inactivation due to Ca2+-dependent calmodulin binding; phosphorylation by PKC inhibits Ca2+-calmodulin binding and slow inactivation; upregulated by 1,25-dihydroxyvitamin D3

Comments

Activation of TRPV1 by depolarisation is strongly temperature-dependent via a channel opening rate that increases with increasing temperature. The V½ is shifted in the hyperpolarizing direction both by increasing temperature and by exogenous agonists 342. The sensitivity of TRPV4 to heat, but not 4α-PDD is lost upon patch excision. TRPV4 is activated by anandamide and arachidonic acid following P450 epoxygenase-dependent metabolism to 5,6-epoxyeicosatrienoic acid (reviewed by 311). Activation of TRPV4 by cell swelling, but not heat, or phorbol esters, is mediated via the formation of epoxyeicosatrieonic acids. Phorbol esters bind directly to TRPV4. TRPV5 preferentially conducts Ca2+ under physiological conditions, but in the absence of extracellular Ca2+, conducts monovalent cations. Single channel conductances listed for TRPV5 and TRPV6 were determined in divalent cation-free extracellular solution. Ca2+-induced inactivation occurs at hyperpolarized potentials when Ca2+ is present extracellularly. Single channel events cannot be resolved (probably due to greatly reduced conductance) in the presence of extracellular divalent cations. Measurements of PCa/PNa for TRPV5 and TRPV6 are dependent upon ionic conditions due to anomalous mole fraction behaviour. Blockade of TRPV5 and TRPV6 by extracellular Mg2+ is voltage-dependent. Intracellular Mg2+ also exerts a voltage dependent block that is alleviated by hyperpolarization and contributes to the time-dependent activation and deactivation of TRPV6 mediated monovalent cation currents. TRPV5 and TRPV6 differ in their kinetics of Ca2+-dependent inactivation and recovery from inactivation. TRPV5 and TRPV6 function as homo- and hetero-tetramers.

Nomenclature TRPML1 TRPML2 TRPML3
HGNC, UniProt MCOLN1, Q9GZU1 MCOLN2, Q8IZK6 MCOLN3, Q8TDD5
Activators TRPML1Va: Constitutively active, current potentiated by extracellular acidification (equivalent to intralysosomal acidification) TRPML2Va: Constitutively active, current potentiated by extracellular acidification (equivalent to intralysosomal acidification) TRPML3Va: Constitutively active, current inhibited by extracellular acidification (equivalent to intralysosomal acidicification), Wild type TRPML3: Activated by Na+-free extracellular (extracytosolic) solution and membrane depolarization, current inhibited by extracellular acidification (equivalent to intralysosomal acidicification)
Channel Blockers (IC50) Gd3+
Functional characteristics TRPML1Va: γ = 40 pS and 76-86 pS at very negative holding potentials with Fe2+ and monovalent cations as charge carriers, respectively; conducts Na+≅ K+>Cs+ and divalent cations (Ba2+>Mn2+>Fe2+>Ca2+> Mg2+> Ni2+>Co2+> Cd2+>Zn2+>>Cu2+) protons; monovalent cation flux suppressed by divalent cations (e.g. Ca2+, Fe2+); inwardly rectifying TRPML1Va: Conducts Na+; monovalent cation flux suppressed by divalent cations; inwardly rectifying TRPML3Va: γ = 49 pS at very negative holding potentials with monovalent cations as charge carrier; conducts Na+ > K+ > Cs+ with maintained current in the presence of Na+, conducts Ca2+ and Mg2+, but not Fe2+, impermeable to protons; inwardly rectifying Wild type TRPML3: γ = 59 pS at negative holding potentials with monovalent cations as charge carrier; conducts Na+ > K+ > Cs+ and Ca2+ (PCa/PK ≅ 350), slowly inactivates in the continued presence of Na+ within the extracellular (extracytosolic) solution; outwardly rectifying

Comments

Data in the table are for TRPML proteins mutated (i.e TRPML1Va, TRPML2Va and TRPML3Va) at loci equivalent to TRPML3 A419P to allow plasma membrane expression when expressed in HEK-293 cells and subsequent characterisation by patch-clamp recording 253,261,275,304,352. Data for wild type TRPML3 are also tabulated 275,276,304,352. It should be noted that alternative methodologies, particularly in the case of TRPML1, have resulted in channels with differing biophysical characteristics (reviewed by 323).

Nomenclature TRPP2 TRPP3
HGNC, UniProt PKD2L1, Q9P0L9 PKD2L2, Q9NZM6
Activators Low constitutive activity, enhanced by membrane depolarization; changes in cell volume affect voltage-dependent gating (increased channel opening probability with cell swelling)
Channel Blockers (IC50) flufenamate, Gd3+, La3+, phenamil (1.4x10-7 M), benzamil (1.1x10-6 M), EIPA (1.05x10-5 M), amiloride (1.43x10-4 M)
Functional characteristics γ = 105–137 pS (outward conductance) 184–399 pS (inward conductance), conducts mono- and di-valent cations with a preference for divalents (PCa/PNa = 4.0–4.3); steady state currents rectify outwardly, whereas instantaneous currents show strong inward rectification; activated and subsequently inactivated by intracellular Ca2+ (human, but not mouse); inhibited by extracellular acidification and potentiated by extracellular alkalization

Comments

Data in the table are extracted from 247,251 and 332. Broadly similar single channel conductance, mono- and di-valent cation selectivity and sensitivity to blockers are observed for TRPP2 co-expressed with TRPP1 250. Ca2+, Ba2+ and Sr2+ permeate TRPP3, but reduce inward currents carried by Na+. Mg2+ is largely impermeant and exerts a voltage dependent inhibition that increases with hyperpolarization.

Voltage-gated calcium channels

Overview

Calcium (Ca2+) channels are voltage-gated ion channels present in the membrane of most excitable cells. The nomenclature for Ca2+ channels was proposed by 359 and approved by the NC-IUPHAR subcommittee on Ca2+ channels 358. Ca2+ channels form hetero-oligomeric complexes. The α1 subunit is pore-forming and provides the extracellular binding site(s) for practically all agonists and antagonists. The 10 cloned α-subunits can be grouped into three families: (1) the high-voltage activated dihydropyridine-sensitive (L-type, CaV1.x) channels; (2) the high-voltage activated dihydropyridine-insensitive (CaV2.x) channels and (3) the low-voltage-activated (T-type, CaV3.x) channels. Each α1 subunit has four homologous repeats (I–IV), each repeat having six transmembrane domains and a pore-forming region between transmembrane domains S5 and S6. Gating is thought to be associated with the membrane-spanning S4 segment, which contains highly conserved positive charges. Many of the α1-subunit genes give rise to alternatively spliced products. At least for high-voltage activated channels, it is likely that native channels comprise co-assemblies of α1, β and α2–δ subunits. The γ subunits have not been proven to associate with channels other than α1s. The α2–δ1 and α2–δ2 subunits bind gabapentin and pregabalin.

Subunits

Nomenclature Cav1.1 Cav1.2 Cav1.3 Cav1.4 Cav2.1
HGNC, UniProt CACNA1S, Q13698 CACNA1C, Q13936 CACNA1D, Q01668 CACNA1F, O60840 CACNA1A, O00555
Activators (EC50) FPL64176, (-)-(S)-BayK8644, SZ(+)-(S)-202-791 FPL64176, (-)-(S)-BayK8644, SZ(+)-(S)-202-791 (-)-(S)-BayK8644 (-)-(S)-BayK8644
Channel Blockers (IC50) calciseptine, diltiazem, nifedipine, verapamil calciseptine, diltiazem, nifedipine, verapamil verapamil (less sensitive to dihydropyridine antagonists) ω-agatoxin IVB, ω-conotoxin MVIIC, ω-agatoxin IVA (P current component) (∼1x10-9 M), ω-agatoxin IVA (Q current component) (∼9x10-8 M)
Functional characteristics High voltage-activated, slow inactivation High voltage-activated, slow inactivation (Ca2+ dependent) Low-moderate voltage-activated, slow inactivation (Ca2+ dependent) Moderate voltage-activated, slow inactivation (Ca2+ independent) Moderate voltage-activated, moderate inactivation
Comment nifedipine, diltiazem, verapamil and calciseptine are examples of dihydropyridine antagonists nifedipine, diltiazem, verapamil and calciseptine are examples of dihydropyridine antagonists verapamil is an example of a dihydropyridine antagonist Cav1.4 is less sensitive to dihydropyridine antagonists
Nomenclature Cav2.2 Cav3.1 Cav3.2 Cav3.3
HGNC, UniProt CACNA1B, Q00975 CACNA1G, O43497 CACNA1H, O95180 CACNA1I, Q9P0X4
Channel Blockers (IC50) ω-conotoxin GVIA, ω-conotoxin MVIIC kurtoxin, mibefradil, Ni2+ (low sensitivity to Ni2+), SB209712 kurtoxin, mibefradil, Ni2+ (high sensitivity to Ni2+), SB209712 kurtoxin, mibefradil, Ni2+ (low sensitivity to Ni2+), SB209712
Functional characteristics High voltage-activated, moderate inactivation Low voltage-activated, fast inactivation Low voltage-activated, fast inactivation Low voltage-activated, moderate inactivation

Comments

In many cell types, P and Q current components cannot be adequately separated and many researchers in the field have adopted the terminology ‘P/Q-type’ current when referring to either component. Ziconotide (a synthetic peptide equivalent to ω-conotoxin MVIIA) has been approved for the treatment of chronic pain 360.

Voltage-gated proton channel

Overview

The voltage-gated proton channel (provisionally denoted Hv1) is a putative 4TM proton-selective channel gated by membrane depolarization and which is sensitive to the transmembrane pH gradient 361363,372,374. The structure of Hv1 is homologous to the voltage sensing domain (VSD) of the superfamily of voltage-gated ion channels (i.e. segments S1 to S4) and contains no discernable pore region 372,374. Proton flux through Hv1 is instead most likely mediated by a water wire completed in a crevice of the protein when the voltage-sensing S4 helix moves in response to a change in transmembrane potential 371,377. Hv1 expresses largely as a dimer mediated by intracellular C-terminal coiled-coil interactions 367 but individual promoters nonetheless support gated H+ flux via separate conduction pathways 365,366,370,375. Within dimeric structures, the two protomers do not function independently, but display co-operative interactions during gating resulting in increased voltage sensitivity, but slower activation, of the dimeric, versus monomeric, complexes 364,376.

Subunits

Nomenclature HGNC, UniProt Channel Blockers (IC50) Functional characteristics
Hv1 HVCN1, Q96D96 Zn2+ (∼5x10-7 – 2x10-6 M), Cd2+ (∼1x10-5 M) Activated by membrane depolarization mediating macroscopic currents with time-, voltage- and pH-dependence; outwardly rectifying; voltage dependent kinetics with relatively slow current activation sensitive to extracellular pH and temperature, relatively fast deactivation; voltage threshold for current activation determined by pH gradient (ΔpH = pHo -pHi) across the membrane

Comments

The voltage threshold (Vthr) for activation of Hv1 is not fixed but is set by the pH gradient across the membrane such that Vthr is positive to the Nernst potential for H+, which ensures that only outwardly directed flux of H+ occurs under physiological conditions 361363. Phosphorylation of Hv1 within the N-terminal domain by PKC enhances the gating of the channel 368. Tabulated IC50 values for Zn2+ and Cd2+ are for heterologously expressed human and mouse Hv1 372,374. Zn2+ is not a conventional pore blocker, but is coordinated by two, or more, external protonation sites involving histamine residues 372. Zn2+ binding may occur at the dimer interface between pairs of histamine residues from both monomers where it may interfere with channel opening 369. Mouse knockout studies demonstrate that Hv1 participates in charge compensation in granulocytes during the respiratory burst of NADPH oxidase-dependent reactive oxygen species production that assists in the clearance of bacterial pathogens 373. Additional physiological functions of Hv1 are reviewed by 361.

Voltage-gated sodium channels

Overview

Sodium channels are voltage-gated sodium-selective ion channels present in the membrane of most excitable cells. Sodium channels comprise of one pore-forming α subunit, which may be associated with either one or two β subunits 380. α-Subunits consist of four homologous domains (I–IV), each containing six transmembrane segments (S1–S6) and a pore-forming loop. The positively charged fourth transmembrane segment (S4) acts as a voltage sensor and is involved in channel gating. The crystal structure of the bacterial NavAb channel has revealed a number of novel structural features compared to earlier potassium channel structures including a short selectivity filter with ion selectivity determined by interactions with glutamate side chains 381. Interestingly, the pore region is penetrated by fatty acyl chains that extend into the central cavity which may allow the entry of small, hydrophobic pore-blocking drugs 381. Auxiliary β1, β2, β3 and β4 subunits consist of a large extracellular N-terminal domain, a single transmembrane segment and a shorter cytoplasmic domain.

The nomenclature for sodium channels was proposed by Goldin et al., (2000) 379 and approved by the NC-IUPHAR subcommittee on sodium channels (Catterall et al., 12, 378).

Subunits

Nomenclature Nav1.1 Nav1.2 Nav1.3 Nav1.4 Nav1.5 Nav1.6 Nav1.7 Nav1.8 Nav1.9
HGNC, UniProt SCN1A, P35498 SCN2A, Q99250 SCN3A, Q9NY46 SCN4A, P35499 SCN5A, Q14524 SCN8A, Q9UQD0 SCN9A, Q15858 SCN10A, Q9Y5Y9 SCN11A, Q9UI33
Activators (EC50) batrachotoxin, veratridine batrachotoxin, veratridine batrachotoxin, veratridine batrachotoxin, veratridine batrachotoxin, veratridine batrachotoxin, veratridine batrachotoxin, veratridine
Channel Blockers (IC50) saxitoxin, tetrodotoxin (Concentration range = 1x10-8 M) saxitoxin,tetrodotoxin (Concentration range = 1x10-8 M) saxitoxin, tetrodotoxin (Concentration range = 2x10-9 - 1.5x10-8 M) μ-conotoxin GIIIA, saxitoxin, tetrodotoxin (Concentration range = 5x10-9 M) tetrodotoxin (Concentration range = 2x10-6 M) saxitoxin, tetrodotoxin (Concentration range = 6x10-9 M) saxitoxin, tetrodotoxin (Concentration range = 4x10-9 M) tetrodotoxin (Concentration range = 6x10-5 M) tetrodotoxin (Concentration range = 4x10-5 M)
Functional characteristics Fast inactivation (0.7 ms) Fast inactivation (0.8 ms) Fast inactivation (0.8 ms) Fast inactivation (0.6 ms) Fast inactivation (1 ms) Fast inactivation (1 ms) Fast inactivation (0.5 ms) Slow inactivation (6 ms) Slow inactivation (16 ms)

Comments

Sodium channels are also blocked by local anaesthetic agents, antiarrythmic drugs and antiepileptic drugs. There are two clear functional fingerprints for distinguishing different subtypes. These are sensitivity to tetrodotoxin (NaV1.5, NaV1.8 and NaV1.9 are much less sensitive to block) and rate of inactivation (NaV1.8 and particularly NaV1.9 inactivate more slowly).

Further reading

  1. Doyle D. Morais Cabral J. Pfuetzner RA. Kuo A. Gulbis JM. Cohen SL. Chait B. MacKinnon R. The structure of the potassium channel : molecular basis of potassium conduction and selectivity. Science. 1998;280:69–77. doi: 10.1126/science.280.5360.69. [DOI] [PubMed] [Google Scholar]
  2. Dunlop J. Bowlby M. Peri R. Vasilyev D. Arias R. High-throughput electrophysiology: an emerging paradigm for ion channel screening and physiology. Nat Rev Drug Discov. 2008;7:358–368. doi: 10.1038/nrd2552. [DOI] [PubMed] [Google Scholar]
  3. Dutzler R. Campbell EB. Cadene M. Chait B. MacKinnon R. X-ray structure of a ClC chloride channel at 3.0A reveals the molecular basis of ion selectivity. Nature. 2002;415:287–294. doi: 10.1038/415287a. [DOI] [PubMed] [Google Scholar]
  4. Hille B. 3rd Edition. Sunderland MA: Sinauer Associates; 2006. Ion channel of excitable membranes. [Google Scholar]
  5. Overington JP. Al-Lazikani B. Hopkins AL. Nat Rev Drug Discov. Vol. 5. 2001. How many drug targets are there? pp. 993–996. [DOI] [PubMed] [Google Scholar]
  6. Payandeh J. Scheuer T. Zheng N. Catterall WA. Nature. Vol. 475. 2011. The crystal structure of a voltage-gated sodium channel; pp. 353–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Yu FH. Yarov-Yarovoy V. Gutman GA. Catterall WA. Pharmacol Rev. Vol. 57. 2005. Overview of molecular relationships in the voltage-gated ion channel superfamily; pp. 387–395. [DOI] [PubMed] [Google Scholar]

Further reading

  1. Chen X. Orser BA. MacDonald JF. Eur J Pharmacol. Vol. 648. 2010. Design and screening of ASIC inhibitors based on aromatic diamidines for combating neurological disorders; pp. 15–23. [PMID:20854810] [DOI] [PubMed] [Google Scholar]
  2. Deval E. Gasull X. Noël J. Salinas M. Baron A. Diochot S. Lingueglia E. Pharmacol Ther. Vol. 128. 2010. Acid-sensing ion channels (ASICs): pharmacology and implication in pain; pp. 549–558. [PMID:20807551] [DOI] [PubMed] [Google Scholar]
  3. Waldmann R. Lazdunski M. Curr Opin Neurobiol. Vol. 8. 1998. H(+)-gated cation channels: neuronal acid sensors in the NaC/DEG family of ion channels; pp. 418–424. [PMID:9687356] [DOI] [PubMed] [Google Scholar]
  4. Wemmie JA. Price MP. Welsh MJ. Trends Neurosci. Vol. 29. 2006. Acid-sensing ion channels: advances, questions and therapeutic opportunities; pp. 578–586. [PMID:16891000] [DOI] [PubMed] [Google Scholar]
  5. Xiong ZG. Pignataro G. Li M. Chang SY. Simon RP. Curr Opin Pharmacol. Vol. 8. 2008. Acid-sensing ion channels (ASICs) as pharmacological targets for neurodegenerative diseases; pp. 25–32. [PMID:17945532] [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Xu TL. Duan B. Prog Neurobiol. Vol. 87. 2009. Calcium-permeable acid-sensing ion channel in nociceptive plasticity: a new target for pain control; pp. 171–180. [PMID:19388206] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Amiry-Moghaddam M. Ottersen OP. Nat Rev Neurosci. Vol. 4. 2003. The molecular basis of water transport in the brain; pp. 991–1001. [PMID:14682361] [DOI] [PubMed] [Google Scholar]
  2. Carbrey JM. Agre P. Handb Exp Pharmacol. Vol. 190. 2009. Discovery of the aquaporins and development of the field; pp. 3–28. [PMID:19096770] [DOI] [PubMed] [Google Scholar]
  3. Castle NA. Aquaporins as targets for drug discovery. Drug Discov Today. 2005;10:485–493. doi: 10.1016/S1359-6446(05)03390-8. [PMID:15809194] [DOI] [PubMed] [Google Scholar]
  4. Kimelberg HK. Water homeostasis in the brain: basic concepts. Neuroscience. 2004;129:851–860. doi: 10.1016/j.neuroscience.2004.07.033. [PMID:15561403] [DOI] [PubMed] [Google Scholar]
  5. King LS. Kozono D. Agre P. Nat Rev Mol Cell Biol. Vol. 5. 2004. From structure to disease: the evolving tale of aquaporin biology; pp. 687–698. [PMID:15340377] [DOI] [PubMed] [Google Scholar]
  6. Rojek A. Praetorius J. Frøkiaer J. Nielsen S. Fenton RA. Annu Rev Physiol. Vol. 70. 2008. A current view of the mammalian aquaglyceroporins; pp. 301–327. [PMID:17961083] [DOI] [PubMed] [Google Scholar]
  7. Verkman AS. Aquaporins: translating bench research to human disease. J Exp Biol. 2009;212:1707–1715. doi: 10.1242/jeb.024125. (Pt 11): [PMID:19448080] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Clapham DE. Garbers DL. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. L. Nomenclature and structure-function relationships of CatSper and two-pore channels; pp. 451–454. [PMID:16382101] [DOI] [PubMed] [Google Scholar]
  2. Hildebrand MS. Avenarius MR. Fellous M. Zhang Y. Meyer NC. Auer J. Serres C. Kahrizi K. Najmabadi H. Beckmann JS, et al. Eur J Hum Genet. Vol. 18. 2010. Genetic male infertility and mutation of CATSPER ion channels; pp. 1178–1184. [PMID:20648059] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Kirichok Y. Lishko PV. Mol Hum Reprod. Vol. 17. 2011. Rediscovering sperm ion channels with the patch-clamp technique; pp. 478–499. [PMID:21642646] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Lishko PV. Kirichok Y. J Physiol (Lond) Vol. 588. 2010. The role of Hv1 and CatSper channels in sperm activation; pp. 4667–4672. (Pt 23): [PMID:20679352] [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Ren D. Xia J. Physiology (Bethesda) Vol. 25. 2010. Calcium signaling through CatSper channels in mammalian fertilization; pp. 165–175. [PMID:20551230] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Accardi A. Picollo A. Biochim Biophys Acta. Vol. 1798. 2010. CLC channels and transporters: proteins with borderline personalities; pp. 1457–1464. [PMID:20188062] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alekov AK. Fahlke C. Curr Biol. Vol. 18. 2008. Anion channels: regulation of ClC-3 by an orphan second messenger; pp. R1061–R1064. [PMID:19036336] [DOI] [PubMed] [Google Scholar]
  3. Aleksandrov AA. Aleksandrov LA. Riordan JR. Pflugers Arch. Vol. 453. 2007. CFTR (ABCC7) is a hydrolyzable-ligand-gated channel; pp. 693–702. [PMID:17021796] [DOI] [PubMed] [Google Scholar]
  4. Amaral MD. Kunzelmann K. Trends Pharmacol Sci. Vol. 28. 2007. Molecular targeting of CFTR as a therapeutic approach to cystic fibrosis; pp. 334–341. [PMID:17573123] [DOI] [PubMed] [Google Scholar]
  5. Aromataris EC. Rychkov GY. Clin Exp Pharmacol Physiol. Vol. 33. 2006. ClC-1 chloride channel: Matching its properties to a role in skeletal muscle; pp. 1118–1123. [PMID:17042925] [DOI] [PubMed] [Google Scholar]
  6. Ashlock MA. Olson ER. Annu Rev Med. Vol. 62. 2011. Therapeutics development for cystic fibrosis: a successful model for a multisystem genetic disease; pp. 107–125. [PMID:21226613] [DOI] [PubMed] [Google Scholar]
  7. Best L. Brown PD. Sener A. Malaisse WJ. Islets. Vol. 2. 2010. Electrical activity in pancreatic islet cells: The VRAC hypothesis; pp. 59–64. [PMID:21099297] [DOI] [PubMed] [Google Scholar]
  8. Chen TY. Structure and function of clc channels. Annu Rev Physiol. 2005;67:809–839. doi: 10.1146/annurev.physiol.67.032003.153012. [PMID:15709979] [DOI] [PubMed] [Google Scholar]
  9. Chen TY. Hwang TC. Physiol Rev. Vol. 88. 2008. CLC-0 and CFTR: chloride channels evolved from transporters; pp. 351–387. [PMID:18391167] [DOI] [PubMed] [Google Scholar]
  10. Cuthbert AW. New horizons in the treatment of cystic fibrosis. Br J Pharmacol. 2011;163:173–183. doi: 10.1111/j.1476-5381.2010.01137.x. [PMID:21108631] [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Duan D. Phenomics of cardiac chloride channels: the systematic study of chloride channel function in the heart. J Physiol (Lond) 2009;587:2163–2177. doi: 10.1113/jphysiol.2008.165860. (Pt 10): [PMID:19171656] [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Duan DD. The ClC-3 chloride channels in cardiovascular disease. Acta Pharmacol Sin. 2011;32:675–684. doi: 10.1038/aps.2011.30. [PMID:21602838] [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Duran C. Hartzell HC. Acta Pharmacol Sin. Vol. 32. 2011. Physiological roles and diseases of Tmem16/Anoctamin proteins: are they all chloride channels? pp. 685–692. [PMID:21642943] [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Duran C. Thompson CH. Xiao Q. Hartzell HC. Annu Rev Physiol. Vol. 72. 2010. Chloride channels: often enigmatic, rarely predictable; pp. 95–121. [PMID:19827947] [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Dutzler R. A structural perspective on ClC channel and transporter function. FEBS Lett. 2007;581:2839–2844. doi: 10.1016/j.febslet.2007.04.016. [PMID:17452037] [DOI] [PubMed] [Google Scholar]
  16. Edwards JC. Kahl CR. FEBS Lett. Vol. 584. 2010. Chloride channels of intracellular membranes; pp. 2102–2111. [PMID:20100480] [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Fahlke C. Fischer M. Front Physiol. Vol. 1. 2010. Physiology and pathophysiology of ClC-K/barttin channels; p. 155. [PMID:21423394] [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Ferrera L. Caputo A. Galietta LJ. Physiology (Bethesda) Vol. 25. 2010. TMEM16A protein: a new identity for Ca(2+)-dependent Cl− channels; pp. 357–363. [PMID:21186280] [DOI] [PubMed] [Google Scholar]
  19. Gadsby DC. Vergani P. Csanády L. Nature. Vol. 440. 2006. The ABC protein turned chloride channel whose failure causes cystic fibrosis; pp. 477–483. [PMID:16554808] [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Galietta LJ. The TMEM16 protein family: a new class of chloride channels? Biophys J. 2009;97:3047–3053. doi: 10.1016/j.bpj.2009.09.024. [PMID:20006941] [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Greenwood IA. Leblanc N. Trends Pharmacol Sci. Vol. 28. 2007. Overlapping pharmacology of Ca2+-activated Cl- and K+ channels; pp. 1–5. [PMID:17150263] [DOI] [PubMed] [Google Scholar]
  22. Guan YY. Wang GL. Zhou JG. Trends Pharmacol Sci. Vol. 27. 2006. The ClC-3 Cl- channel in cell volume regulation, proliferation and apoptosis in vascular smooth muscle cells; pp. 290–296. [PMID:16697056] [DOI] [PubMed] [Google Scholar]
  23. Hartzell C. Putzier I. Arreola J. Annu Rev Physiol. Vol. 67. 2005. Calcium-activated chloride channels; pp. 719–758. [PMID:15709976] [DOI] [PubMed] [Google Scholar]
  24. Hartzell HC. Qu Z. Yu K. Xiao Q. Chien LT. Physiol Rev. Vol. 88. 2008. Molecular physiology of bestrophins: multifunctional membrane proteins linked to best disease and other retinopathies; pp. 639–672. [PMID:18391176] [DOI] [PubMed] [Google Scholar]
  25. Jentsch TJ. CLC chloride channels and transporters: from genes to protein structure, pathology and physiology. Crit Rev Biochem Mol Biol. 2008;43:3–36. doi: 10.1080/10409230701829110. [PMID:18307107] [DOI] [PubMed] [Google Scholar]
  26. Kirk KL. Wang W. J Biol Chem. Vol. 286. 2011. A unified view of cystic fibrosis transmembrane conductance regulator (CFTR) gating: combining the allosterism of a ligand-gated channel with the enzymatic activity of an ATP-binding cassette (ABC) transporter; pp. 12813–12819. [PMID:21296873] [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Krämer BK. Bergler T. Stoelcker B. Waldegger S. Nat Clin Pract Nephrol. Vol. 4. 2008. Mechanisms of Disease: the kidney-specific chloride channels ClCKA and ClCKB, the Barttin subunit, and their clinical relevance; pp. 38–46. [PMID:18094726] [DOI] [PubMed] [Google Scholar]
  28. Kunzelmann K. Tian Y. Martins JR. Faria D. Kongsuphol P. Ousingsawat J. Thevenod F. Roussa E. Rock J. Schreiber R. Pflugers Arch. Vol. 462. 2011. Anoctamins; pp. 195–208. [PMID:21607626] [DOI] [PubMed] [Google Scholar]
  29. Leblanc N. Ledoux J. Saleh S. Sanguinetti A. Angermann J. O'Driscoll K. Britton F. Perrino BA. Greenwood IA. Can J Physiol Pharmacol. Vol. 83. 2005. Regulation of calcium-activated chloride channels in smooth muscle cells: a complex picture is emerging; pp. 541–556. [PMID:16091780] [DOI] [PubMed] [Google Scholar]
  30. Muallem D. Vergani P. Philos Trans R Soc Lond, B, Biol Sci. Vol. 364. 2009. Review. ATP hydrolysis-driven gating in cystic fibrosis transmembrane conductance regulator; pp. 247–255. [PMID:18957373] [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Mulligan SJ. MacVicar BA. Sci STKE. Vol. 2006. 2006. VRACs CARVe a path for novel mechanisms of communication in the CNS; p. pe42. [PMID:17047222] [DOI] [PubMed] [Google Scholar]
  32. Noy E. Senderowitz H. ChemMedChem. Vol. 6. 2011. Combating cystic fibrosis: in search for CF transmembrane conductance regulator (CFTR) modulators; pp. 243–251. [PMID:21275046] [DOI] [PubMed] [Google Scholar]
  33. Okada Y. Cell volume-sensitive chloride channels: phenotypic properties and molecular identity. Contrib Nephrol. 2006;152:9–24. doi: 10.1159/000096285. [PMID:17065805] [DOI] [PubMed] [Google Scholar]
  34. Okada Y. Sato K. Numata T. J Physiol (Lond) Vol. 587. 2009. Pathophysiology and puzzles of the volume-sensitive outwardly rectifying anion channel; pp. 2141–2149. (Pt 10): [PMID:19171657] [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Patel AC. Brett TJ. Holtzman MJ. Annu Rev Physiol. Vol. 71. 2009. The role of CLCA proteins in inflammatory airway disease; pp. 425–449. [PMID:18954282] [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Planells-Cases R. Jentsch TJ. Biochim Biophys Acta. Vol. 1792. 2009. Chloride channelopathies; pp. 173–189. [PMID:19708126] [DOI] [PubMed] [Google Scholar]
  37. Plans V. Rickheit G. Jentsch TJ. Pflugers Arch. Vol. 458. 2009. Physiological roles of CLC Cl(-)/H (+) exchangers in renal proximal tubules; pp. 23–37. [PMID:18853181] [DOI] [PubMed] [Google Scholar]
  38. Puljak L. Kilic G. Biochim Biophys Acta. Vol. 1762. 2006. Emerging roles of chloride channels in human diseases; pp. 404–413. [PMID:16457993] [DOI] [PubMed] [Google Scholar]
  39. Pusch M. Zifarelli G. Murgia AR. Picollo A. Babini E. Exp Physiol. Vol. 91. 2006. Channel or transporter? The CLC saga continues; pp. 149–152. [PMID:16179405] [DOI] [PubMed] [Google Scholar]
  40. Riordan JR. Assembly of functional CFTR chloride channels. Annu Rev Physiol. 2005;67:701–718. doi: 10.1146/annurev.physiol.67.032003.154107. [PMID:15709975] [DOI] [PubMed] [Google Scholar]
  41. Riquelme G. Placental chloride channels: a review. Placenta. 2009;30:659–669. doi: 10.1016/j.placenta.2009.06.002. [PMID:19604577] [DOI] [PubMed] [Google Scholar]
  42. Sabirov RZ. Okada Y. J Physiol Sci. Vol. 59. 2009. The maxi-anion channel: a classical channel playing novel roles through an unidentified molecular entity; pp. 3–21. [PMID:19340557] [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Sloane PA. Rowe SM. Curr Opin Pulm Med. Vol. 16. 2010. Cystic fibrosis transmembrane conductance regulator protein repair as a therapeutic strategy in cystic fibrosis; pp. 591–597. [PMID:20829696] [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Verkman AS. Galietta LJ. Nat Rev Drug Discov. Vol. 8. 2009. Chloride channels as drug targets; pp. 153–171. [PMID:19153558] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Cruciani V. Mikalsen SO. Cell Mol Life Sci. Vol. 63. 2006. The vertebrate connexin family; pp. 1125–1140. [PMID:16568237] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Evans WH. De Vuyst E. Leybaert L. Biochem J. Vol. 397. 2006. The gap junction cellular internet: connexin hemichannels enter the signalling limelight; pp. 1–14. [PMID:16761954] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Kandouz M. Batist G. Expert Opin Ther Targets. Vol. 14. 2010. Gap junctions and connexins as therapeutic targets in cancer; pp. 681–692. [PMID:20446866] [DOI] [PubMed] [Google Scholar]
  4. MacVicar BA. Thompson RJ. Trends Neurosci. Vol. 33. 2010. Non-junction functions of pannexin-1 channels; pp. 93–102. [PMID:20022389] [DOI] [PubMed] [Google Scholar]
  5. Meşe G. Richard G. White TW. J Invest Dermatol. Vol. 127. 2007. Gap junctions: basic structure and function; pp. 2516–2524. [PMID:17934503] [DOI] [PubMed] [Google Scholar]
  6. Shestopalov VI. Panchin Y. Cell Mol Life Sci. Vol. 65. 2008. Pannexins and gap junction protein diversity; pp. 376–394. [PMID:17982731] [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Söhl G. Maxeiner S. Willecke K. Nat Rev Neurosci. Vol. 6. 2005. Expression and functions of neuronal gap junctions; pp. 191–200. [PMID:15738956] [DOI] [PubMed] [Google Scholar]
  8. Zoidl G. Dermietzel R. Pflugers Arch. Vol. 460. 2010. Gap junctions in inherited human disease; pp. 451–466. [PMID:20140684] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Baruscotti M. Bottelli G. Milanesi R. DiFrancesco JC. DiFrancesco D. Pflugers Arch. Vol. 460. 2010. HCN-related channelopathies; pp. 405–415. [PMID:20213494] [DOI] [PubMed] [Google Scholar]
  2. Baruscotti M. Bucchi A. Difrancesco D. Pharmacol Ther. Vol. 107. 2005. Physiology and pharmacology of the cardiac pacemaker ("funny") current; pp. 59–79. [PMID:15963351] [DOI] [PubMed] [Google Scholar]
  3. Biel M. Michalakis S. Handb Exp Pharmacol. Vol. 191. 2009. Cyclic nucleotide-gated channels; pp. 111–136. [PMID:19089328] [DOI] [PubMed] [Google Scholar]
  4. Biel M. Wahl-Schott C. Michalakis S. Zong X. Physiol Rev. Vol. 89. 2009. Hyperpolarization-activated cation channels: from genes to function; pp. 847–885. [PMID:19584315] [DOI] [PubMed] [Google Scholar]
  5. Bois P. Guinamard R. Chemaly AE. Faivre JF. Bescond J. Curr Pharm Des. Vol. 13. 2007. Molecular regulation and pharmacology of pacemaker channels; pp. 2338–2349. [PMID:17692005] [DOI] [PubMed] [Google Scholar]
  6. Bradley J. Reisert J. Frings S. Curr Opin Neurobiol. Vol. 15. 2005. Regulation of cyclic nucleotide-gated channels; pp. 343–349. [PMID:15922582] [DOI] [PubMed] [Google Scholar]
  7. Brown RL. Strassmaier T. Brady JD. Karpen JW. Curr Pharm Des. Vol. 12. 2006. The pharmacology of cyclic nucleotide-gated channels: emerging from the darkness; pp. 3597–3613. [PMID:17073662] [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Craven KB. Zagotta WN. Annu Rev Physiol. Vol. 68. 2006. CNG and HCN channels: two peas, one pod; pp. 375–401. [PMID:16460277] [DOI] [PubMed] [Google Scholar]
  9. Cukkemane A. Seifert R. Kaupp UB. Trends Biochem Sci. Vol. 36. 2011. Cooperative and uncooperative cyclic-nucleotide-gated ion channels; pp. 55–64. [PMID:20729090] [DOI] [PubMed] [Google Scholar]
  10. DiFrancesco D. The role of the funny current in pacemaker activity. Circ Res. 2010;106:434–446. doi: 10.1161/CIRCRESAHA.109.208041. [PMID:20167941] [DOI] [PubMed] [Google Scholar]
  11. Dunlop J. Vasilyev D. Lu P. Cummons T. Bowlby MR. Curr Pharm Des. Vol. 15. 2009. Hyperpolarization-activated cyclic nucleotide-gated (HCN) channels and pain; pp. 1767–1772. [PMID:19442189] [DOI] [PubMed] [Google Scholar]
  12. Hofmann F. Biel M. Kaupp UB. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. LI. Nomenclature and structure-function relationships of cyclic nucleotide-regulated channels; pp. 455–462. [PMID:16382102] [DOI] [PubMed] [Google Scholar]
  13. Maher MP. Wu NT. Guo HQ. Dubin AE. Chaplan SR. Wickenden AD. Comb Chem High Throughput Screen. Vol. 12. 2009. HCN channels as targets for drug discovery; pp. 64–72. [PMID:19149492] [DOI] [PubMed] [Google Scholar]
  14. Mazzolini M. Marchesi A. Giorgetti A. Torre V. Pflugers Arch. Vol. 459. 2010. Gating in CNGA1 channels; pp. 547–555. [PMID:19898862] [DOI] [PubMed] [Google Scholar]
  15. Meldrum BS. Rogawski MA. Neurotherapeutics. Vol. 4. 2007. Molecular targets for antiepileptic drug development; pp. 18–61. [PMID:17199015] [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Tardif JC. Ivabradine: I(f) inhibition in the management of stable angina pectoris and other cardiovascular diseases. Drugs Today. 2008;44:171–181. doi: 10.1358/dot.2008.44.1193864. [PMID:18536779] [DOI] [PubMed] [Google Scholar]
  17. Wahl-Schott C. Biel M. Cell Mol Life Sci. Vol. 66. 2009. HCN channels: structure, cellular regulation and physiological function; pp. 470–494. [PMID:18953682] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Bhalla V. Hallows KR. J Am Soc Nephrol. Vol. 19. 2008. Mechanisms of ENaC regulation and clinical implications; pp. 1845–1854. [PMID:18753254] [DOI] [PubMed] [Google Scholar]
  2. Bubien JK. Epithelial Na+ channel (ENaC), hormones, and hypertension. J Biol Chem. 2010;285:23527–23531. doi: 10.1074/jbc.R109.025049. [PMID:20460373] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Butterworth MB. Regulation of the epithelial sodium channel (ENaC) by membrane trafficking. Biochim Biophys Acta. 2010;1802:1166–1177. doi: 10.1016/j.bbadis.2010.03.010. [PMID:20347969] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Hamm LL. Feng Z. Hering-Smith KS. Curr Opin Nephrol Hypertens. Vol. 19. 2010. Regulation of sodium transport by ENaC in the kidney; pp. 98–105. [PMID:19996890] [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Kitamura K. Tomita K. Clin Exp Nephrol. Vol. 14. 2010. Regulation of renal sodium handling through the interaction between serine proteases and serine protease inhibitors; pp. 405–410. [PMID:20535627] [DOI] [PubMed] [Google Scholar]
  6. Kleyman TR. Carattino MD. Hughey RP. J Biol Chem. Vol. 284. 2009. ENaC at the cutting edge: regulation of epithelial sodium channels by proteases; pp. 20447–20451. [PMID:19401469] [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Loffing J. Korbmacher C. Pflugers Arch. Vol. 458. 2009. Regulated sodium transport in the renal connecting tubule (CNT) via the epithelial sodium channel (ENaC) pp. 111–135. [PMID:19277701] [DOI] [PubMed] [Google Scholar]
  8. Ma HP. Chou CF. Wei SP. Eaton DC. Pflugers Arch. Vol. 455. 2007. Regulation of the epithelial sodium channel by phosphatidylinositides: experiments, implications, and speculations; pp. 169–180. [PMID:17605040] [DOI] [PubMed] [Google Scholar]
  9. Planès C. Caughey GH. Curr Top Dev Biol. Vol. 78. 2007. Regulation of the epithelial Na+ channel by peptidases; pp. 23–46. [PMID:17338914] [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Pochynyuk O. Bugaj V. Stockand JD. Curr Opin Nephrol Hypertens. Vol. 17. 2008. Physiologic regulation of the epithelial sodium channel by phosphatidylinositides; pp. 533–540. [PMID:18695396] [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Rossier BC. Stutts MJ. Annu Rev Physiol. Vol. 71. 2009. Activation of the epithelial sodium channel (ENaC) by serine proteases; pp. 361–379. [PMID:18928407] [DOI] [PubMed] [Google Scholar]
  12. Rotin D. Schild L. Curr Drug Targets. Vol. 9. 2008. ENaC and its regulatory proteins as drug targets for blood pressure control; pp. 709–716. [PMID:18691017] [DOI] [PubMed] [Google Scholar]
  13. Schild L. The epithelial sodium channel and the control of sodium balance. Biochim Biophys Acta. 2010;1802:1159–1165. doi: 10.1016/j.bbadis.2010.06.014. [PMID:20600867] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Barker CJ. Berggren PO. Pharmacol Rev. Vol. 65. 2013. New horizons in cellular regulation by inositol polyphosphates: insights from the pancreatic β-cell; pp. 641–669. [PMID:23429059] [DOI] [PubMed] [Google Scholar]
  2. Decrock E. De Bock M. Wang N. Gadicherla AK. Bol M. Delvaeye T. Vandenabeele P. Vinken M. Bultynck G. Krysko DV, et al. Biochim Biophys Acta. Vol. 1833. 2013. IP3, a small molecule with a powerful message; pp. 1772–1786. [PMID:23291251] [DOI] [PubMed] [Google Scholar]
  3. Foskett JK. Inositol trisphosphate receptor Ca2+ release channels in neurological diseases. Pflugers Arch. 2010;460:481–494. doi: 10.1007/s00424-010-0826-0. [PMID:20383523] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Kiviluoto S. Vervliet T. Ivanova H. Decuypere JP. De Smedt H. Missiaen L. Bultynck G. Parys JB. Biochim Biophys Acta. Vol. 1833. 2013. Regulation of inositol 1,4,5-trisphosphate receptors during endoplasmic reticulum stress; pp. 1612–1624. [PMID:23380704] [DOI] [PubMed] [Google Scholar]
  5. Rossi AM. Tovey SC. Rahman T. Prole DL. Taylor CW. Biochim Biophys Acta. Vol. 1820. 2012. Analysis of IP3 receptors in and out of cells; pp. 1214–1227. [PMID:22033379] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Ahern CA. Kobertz WR. Biochemistry. Vol. 48. 2009. Chemical tools for K(+) channel biology; pp. 517–526. [PMID:19113860] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Bayliss DA. Barrett PQ. Trends Pharmacol Sci. Vol. 29. 2008. Emerging roles for two-pore-domain potassium channels and their potential therapeutic impact; pp. 566–575. [PMID:18823665] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bean BP. The action potential in mammalian central neurons. Nat Rev Neurosci. 2007;8:451–465. doi: 10.1038/nrn2148. [PMID:17514198] [DOI] [PubMed] [Google Scholar]
  4. Dalby-Brown W. Hansen HH. Korsgaard MP. Mirza N. Olesen SP. Curr Top Med Chem. Vol. 6. 2006. K(v)7 channels: function, pharmacology and channel modulators; pp. 999–1023. [PMID:16787276] [DOI] [PubMed] [Google Scholar]
  5. Enyedi P. Czirják G. Physiol Rev. Vol. 90. 2010. Molecular background of leak K+ currents: two-pore domain potassium channels; pp. 559–605. [PMID:20393194] [DOI] [PubMed] [Google Scholar]
  6. Goldstein SA. Bayliss DA. Kim D. Lesage F. Plant LD. Rajan S. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. LV. Nomenclature and molecular relationships of two-P potassium channels; pp. 527–540. [PMID:16382106] [DOI] [PubMed] [Google Scholar]
  7. Gutman GA. Chandy KG. Grissmer S. Lazdunski M. McKinnon D. Pardo LA. Robertson GA. Rudy B. Sanguinetti MC. Stühmer W. Wang X. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. LIII. Nomenclature and molecular relationships of voltage-gated potassium channels; pp. 473–508. [PMID:16382104] [DOI] [PubMed] [Google Scholar]
  8. Hancox JC. McPate MJ. El Harchi A. Zhang YH. Pharmacol Ther. Vol. 119. 2008. The hERG potassium channel and hERG screening for drug-induced torsades de pointes; pp. 118–132. [PMID:18616963] [DOI] [PubMed] [Google Scholar]
  9. Hansen JB. Towards selective Kir6.2/SUR1 potassium channel openers, medicinal chemistry and therapeutic perspectives. Curr Med Chem. 2006;13:361–376. doi: 10.2174/092986706775527947. [PMID:16475928] [DOI] [PubMed] [Google Scholar]
  10. Honoré E. The neuronal background K2P channels: focus on TREK1. Nat Rev Neurosci. 2007;8:251–261. doi: 10.1038/nrn2117. [PMID:17375039] [DOI] [PubMed] [Google Scholar]
  11. Jenkinson DH. Potassium channels–multiplicity and challenges. Br J Pharmacol. 2006;147(Suppl 1):S63–S71. doi: 10.1038/sj.bjp.0706447. [PMID:16402122] [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Judge SI. Bever CT. Pharmacol Ther. Vol. 111. 2006. Potassium channel blockers in multiple sclerosis: neuronal Kv channels and effects of symptomatic treatment; pp. 224–259. [PMID:16472864] [DOI] [PubMed] [Google Scholar]
  13. Kannankeril P. Roden DM. Darbar D. Pharmacol Rev. Vol. 62. 2010. Drug-induced long QT syndrome; pp. 760–781. [PMID:21079043] [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Kobayashi T. Ikeda K. Curr Pharm Des. Vol. 12. 2006. G protein-activated inwardly rectifying potassium channels as potential therapeutic targets; pp. 4513–4523. [PMID:17168757] [DOI] [PubMed] [Google Scholar]
  15. Kubo Y. Adelman JP. Clapham DE. Jan LY. Karschin A. Kurachi Y. Lazdunski M. Nichols CG. Seino S. Vandenberg CA. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. LIV. Nomenclature and molecular relationships of inwardly rectifying potassium channels; pp. 509–526. [PMID:16382105] [DOI] [PubMed] [Google Scholar]
  16. Lawson K. McKay NG. Curr Pharm Des. Vol. 12. 2006. Modulation of potassium channels as a therapeutic approach; pp. 459–470. [PMID:16472139] [DOI] [PubMed] [Google Scholar]
  17. Lüscher C. Slesinger PA. Nat Rev Neurosci. Vol. 11. 2010. Emerging roles for G protein-gated inwardly rectifying potassium (GIRK) channels in health and disease; pp. 301–315. [PMID:20389305] [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Mannhold R. Structure-activity relationships of K(ATP) channel openers. Curr Top Med Chem. 2006;6:1031–1047. doi: 10.2174/156802606777323647. [PMID:16787278] [DOI] [PubMed] [Google Scholar]
  19. Mathie A. Veale EL. Curr Opin Investig Drugs. Vol. 8. 2007. Therapeutic potential of neuronal two-pore domain potassium-channel modulators; pp. 555–562. [PMID:17659475] [PubMed] [Google Scholar]
  20. Nardi A. Olesen SP. Curr Med Chem. Vol. 15. 2008. BK channel modulators: a comprehensive overview; pp. 1126–1146. [PMID:18473808] [DOI] [PubMed] [Google Scholar]
  21. Pongs O. Schwarz JR. Physiol Rev. Vol. 90. 2010. Ancillary subunits associated with voltage-dependent K+ channels; pp. 755–796. [PMID:20393197] [DOI] [PubMed] [Google Scholar]
  22. Salkoff L. Butler A. Ferreira G. Santi C. Wei A. Nat Rev Neurosci. Vol. 7. 2006. High-conductance potassium channels of the SLO family; pp. 921–931. [PMID:17115074] [DOI] [PubMed] [Google Scholar]
  23. Stocker M. Ca(2+)-activated K+ channels: molecular determinants and function of the SK family. Nat Rev Neurosci. 2004;5:758–770. doi: 10.1038/nrn1516. [PMID:15378036] [DOI] [PubMed] [Google Scholar]
  24. Takeda M. Tsuboi Y. Kitagawa J. Nakagawa K. Iwata K. Matsumoto S. Mol Pain. Vol. 7. 2011. Potassium channels as a potential therapeutic target for trigeminal neuropathic and inflammatory pain; p. 5. [PMID:21219657] [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Trimmer JS. Rhodes KJ. Annu Rev Physiol. Vol. 66. 2004. Localization of voltage-gated ion channels in mammalian brain; pp. 477–519. [PMID:14977411] [DOI] [PubMed] [Google Scholar]
  26. Wang H. Tang Y. Wang L. Long CL. Zhang YL. Curr Med Chem. Vol. 14. 2007. ATP-sensitive potassium channel openers and 2,3-dimethyl-2-butylamine derivatives; pp. 133–155. [PMID:17266574] [DOI] [PubMed] [Google Scholar]
  27. Weatherall KL. Goodchild SJ. Jane DE. Marrion NV. Prog Neurobiol. Vol. 91. 2010. Small conductance calcium-activated potassium channels: from structure to function; pp. 242–255. [PMID:20359520] [DOI] [PubMed] [Google Scholar]
  28. Wei AD. Gutman GA. Aldrich R. Chandy KG. Grissmer S. Wulff H. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. LII. Nomenclature and molecular relationships of calcium-activated potassium channels; pp. 463–472. [PMID:16382103] [DOI] [PubMed] [Google Scholar]
  29. Wickenden AD. McNaughton-Smith G. Curr Pharm Des. Vol. 15. 2009. Kv7 channels as targets for the treatment of pain; pp. 1773–1798. [PMID:19442190] [DOI] [PubMed] [Google Scholar]
  30. Witchel HJ. The hERG potassium channel as a therapeutic target. Expert Opin Ther Targets. 2007;11:321–336. doi: 10.1517/14728222.11.3.321. [PMID:17298291] [DOI] [PubMed] [Google Scholar]
  31. Wulff H. Castle NA. Pardo LA. Nat Rev Drug Discov. Vol. 8. 2009. Voltage-gated potassium channels as therapeutic targets; pp. 982–1001. [PMID:19949402] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Betzenhauser MJ. Marks AR. Pflugers Arch. Vol. 460. 2010. Ryanodine receptor channelopathies; pp. 467–480. [PMID:20179962] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Gaburjakova M. Bal NC. Gaburjakova J. Periasamy M. Cell Mol Life Sci. Vol. 70. 2013. Functional interaction between calsequestrin and ryanodine receptor in the heart; pp. 2935–2945. [PMID:23109100] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Kushnir A. Betzenhauser MJ. Marks AR. FEBS Lett. Vol. 584. 2010. Ryanodine receptor studies using genetically engineered mice; pp. 1956–1965. [PMID:20214899] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. MacMillan D. FK506 binding proteins: cellular regulators of intracellular Ca2+ signalling. Eur J Pharmacol. 2013;700:181–193. doi: 10.1016/j.ejphar.2012.12.029. [PMID:23305836] [DOI] [PubMed] [Google Scholar]
  5. McCauley MD. Wehrens XH. Trends Cardiovasc Med. Vol. 21. 2011. Ryanodine receptor phosphorylation, calcium/calmodulin-dependent protein kinase II, and life-threatening ventricular arrhythmias; pp. 48–51. [PMID:22578240] [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Niggli E. Ullrich ND. Gutierrez D. Kyrychenko S. Poláková E. Shirokova N. Biochim Biophys Acta. Vol. 1833. 2013. Posttranslational modifications of cardiac ryanodine receptors: Ca(2+) signaling and EC-coupling; pp. 866–875. [PMID:22960642] [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Van Petegem F. Ryanodine receptors: structure and function. J Biol Chem. 2012;287:31624–31632. doi: 10.1074/jbc.R112.349068. [PMID:22822064] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Gilon P. Rorsman P. EMBO Rep. Vol. 10. 2009. NALCN: a regulated leak channel; pp. 963–964. [PMID:19662077] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Swayne LA. Mezghrani A. Lory P. Nargeot J. Monteil A. Islets. Vol. 2. 2010. The NALCN ion channel is a new actor in pancreatic β-cell physiology; pp. 54–56. [PMID:21099296] [DOI] [PubMed] [Google Scholar]

Further reading

  1. Abramowitz J. Birnbaumer L. FASEB J. Vol. 23. 2009. Physiology and pathophysiology of canonical transient receptor potential channels; pp. 297–328. [PMID:18940894] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Ambudkar IS. Ong HL. Pflugers Arch. Vol. 455. 2007. Organization and function of TRPC channelosomes; pp. 187–200. [PMID:17486362] [DOI] [PubMed] [Google Scholar]
  3. Baraldi PG. Preti D. Materazzi S. Geppetti P. J Med Chem. Vol. 53. 2010. Transient receptor potential ankyrin 1 (TRPA1) channel as emerging target for novel analgesics and anti-inflammatory agents; pp. 5085–5107. [PMID:20356305] [DOI] [PubMed] [Google Scholar]
  4. Bates-Withers C. Sah R. Clapham DE. Adv Exp Med Biol. Vol. 704. 2011. TRPM7, the Mg(2+) inhibited channel and kinase; pp. 173–183. [PMID:21290295] [DOI] [PubMed] [Google Scholar]
  5. Beech DJ. Acta Physiol (Oxf) 2009. Integration of transient receptor potential canonical channels with lipids. [Epub ahead of print]. [PMID:21624095] [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Birnbaumer L. The TRPC class of ion channels: a critical review of their roles in slow, sustained increases in intracellular Ca(2+) concentrations. Annu Rev Pharmacol Toxicol. 2011;49:395–426. doi: 10.1146/annurev.pharmtox.48.113006.094928. [PMID:19281310] [DOI] [PubMed] [Google Scholar]
  7. Cheng KT. Ong HL. Liu X. Ambudkar IS. Adv Exp Med Biol. Vol. 704. 2011. Contribution of TRPC1 and Orai1 to Ca(2+) entry activated by store depletion; pp. 435–449. [PMID:21290310] [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Clapham DE. Montell C. Schultz G. Julius D. Pharmacol Rev. Vol. 55. 2003. International Union of Pharmacology International Union of Pharmacology. XLIII. Compendium of voltage-gated ion channels: transient receptor potential channels; pp. 591–596. [PMID:14657417] [DOI] [PubMed] [Google Scholar]
  9. Groot T. Bindels RJ. Hoenderop JG. Kidney Int. Vol. 74. 2008. TRPV5: an ingeniously controlled calcium channel; pp. 1241–1246. [PMID:18596722] [DOI] [PubMed] [Google Scholar]
  10. Delmas P. Polycystins: polymodal receptor/ion-channel cellular sensors. Pflugers Arch. 2005;451:264–276. doi: 10.1007/s00424-005-1431-5. [PMID:15889307] [DOI] [PubMed] [Google Scholar]
  11. Delmas P. Padilla F. Osorio N. Coste B. Raoux M. Crest M. Biochem Biophys Res Commun. Vol. 322. 2004. Polycystins, calcium signaling, and human diseases; pp. 1374–1383. [PMID:15336986] [DOI] [PubMed] [Google Scholar]
  12. Dhaka A. Viswanath V. Patapoutian A. Annu Rev Neurosci. Vol. 29. 2006. Trp ion channels and temperature sensation; pp. 135–161. [PMID:16776582] [DOI] [PubMed] [Google Scholar]
  13. Everaerts W. Nilius B. Owsianik G. Prog Biophys Mol Biol. Vol. 103. 2010. The vanilloid transient receptor potential channel TRPV4: from structure to disease; pp. 2–17. [PMID:19835908] [DOI] [PubMed] [Google Scholar]
  14. Fleig A. Penner R. Trends Pharmacol Sci. Vol. 25. 2004. The TRPM ion channel subfamily: molecular, biophysical and functional features; pp. 633–639. [PMID:15530641] [DOI] [PubMed] [Google Scholar]
  15. Freichel M. Vennekens R. Olausson J. Stolz S. Philipp SE. Weissgerber P. Flockerzi V. J Physiol (Lond) Vol. 567. 2005. Functional role of TRPC proteins in native systems: implications from knockout and knock-down studies; pp. 59–66. (Pt 1): [PMID:15975974] [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. García-Añoveros J. Nagata K. Handb Exp Pharmacol. Vol. 179. 2007. TRPA1; pp. 347–362. [PMID:17217068] [DOI] [PubMed] [Google Scholar]
  17. Giamarchi A. Padilla F. Coste B. Raoux M. Crest M. Honoré E. Delmas P. EMBO Rep. Vol. 7. 2006. The versatile nature of the calcium-permeable cation channel TRPP2; pp. 787–793. [PMID:16880824] [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Guinamard R. Sallé L. Simard C. Adv Exp Med Biol. Vol. 704. 2011. The non-selective monovalent cationic channels TRPM4 and TRPM5; pp. 147–171. [PMID:21290294] [DOI] [PubMed] [Google Scholar]
  19. Gunthorpe MJ. Chizh BA. Drug Discov Today. Vol. 14. 2009. Clinical development of TRPV1 antagonists: targeting a pivotal point in the pain pathway; pp. 56–67. [PMID:19063991] [DOI] [PubMed] [Google Scholar]
  20. Harteneck C. Function and pharmacology of TRPM cation channels. Naunyn Schmiedebergs Arch Pharmacol. 2005;371:307–314. doi: 10.1007/s00210-005-1034-x. [PMID:15843919] [DOI] [PubMed] [Google Scholar]
  21. Harteneck C. Gollasch M. Curr Pharm Biotechnol. Vol. 12. 2011. Pharmacological modulation of diacylglycerol-sensitive TRPC3/6/7 channels; pp. 35–41. [PMID:20932261] [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Harteneck C. Klose C. Krautwurst D. Adv Exp Med Biol. Vol. 704. 2011. Synthetic modulators of TRP channel activity; pp. 87–106. [PMID:21290290] [DOI] [PubMed] [Google Scholar]
  23. Hofherr A. Köttgen M. Adv Exp Med Biol. Vol. 704. 2011. TRPP channels and polycystins; pp. 287–313. [PMID:21290302] [DOI] [PubMed] [Google Scholar]
  24. Islam MS. TRP channels of islets. Adv Exp Med Biol. 2011;704:811–830. doi: 10.1007/978-94-007-0265-3_42. [PMID:21290328] [DOI] [PubMed] [Google Scholar]
  25. Kiselyov K. Patterson RL. Front Biosci. Vol. 14. 2009. The integrative function of TRPC channels; pp. 45–58. [PMID:19273053] [DOI] [PubMed] [Google Scholar]
  26. Kiselyov K. Shin DM. Kim JY. Yuan JP. Muallem S. Handb Exp Pharmacol. Vol. 179. 2007. TRPC channels: interacting proteins; pp. 559–574. [PMID:17217079] [DOI] [PubMed] [Google Scholar]
  27. Kiselyov K. Soyombo A. Muallem S. J Physiol (Lond) Vol. 578. 2007. TRPpathies; pp. 641–653. (Pt 3): [PMID:17138610] [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Knowlton WM. McKemy DD. Curr Pharm Biotechnol. Vol. 12. 2011. TRPM8: from cold to cancer, peppermint to pain; pp. 68–77. [PMID:20932257] [DOI] [PubMed] [Google Scholar]
  29. Koike C. Numata T. Ueda H. Mori Y. Furukawa T. Cell Calcium. Vol. 48. 2010. TRPM1: a vertebrate TRP channel responsible for retinal ON bipolar function; pp. 95–101. [PMID:20846719] [DOI] [PubMed] [Google Scholar]
  30. Liman ER. TRPM5 and taste transduction. Handb Exp Pharmacol. 2007;179:287–298. doi: 10.1007/978-3-540-34891-7_17. [PMID:17217064] [DOI] [PubMed] [Google Scholar]
  31. Liu Y. Qin N. Adv Exp Med Biol. Vol. 704. 2011. TRPM8 in health and disease: cold sensing and beyond; pp. 185–208. [PMID:21290296] [DOI] [PubMed] [Google Scholar]
  32. Moran MM. Xu H. Clapham DE. Curr Opin Neurobiol. Vol. 14. 2004. TRP ion channels in the nervous system; pp. 362–369. [PMID:15194117] [DOI] [PubMed] [Google Scholar]
  33. Mälkiä A. Morenilla-Palao C. Viana F. Curr Pharm Biotechnol. Vol. 12. 2011. The emerging pharmacology of TRPM8 channels: hidden therapeutic potential underneath a cold surface; pp. 54–67. [PMID:20932258] [DOI] [PubMed] [Google Scholar]
  34. Nilius B. TRP channels in disease. Biochim Biophys Acta. 2007;1772:805–812. doi: 10.1016/j.bbadis.2007.02.002. [PMID:17368864] [DOI] [PubMed] [Google Scholar]
  35. Nilius B. Owsianik G. Pflugers Arch. Vol. 460. 2010. Transient receptor potential channelopathies; pp. 437–450. [PMID:20127491] [DOI] [PubMed] [Google Scholar]
  36. Nilius B. Owsianik G. Voets T. EMBO J. Vol. 27. 2008. Transient receptor potential channels meet phosphoinositides; pp. 2809–2816. [PMID:18923420] [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Nilius B. Owsianik G. Voets T. Peters JA. Physiol Rev. Vol. 87. 2007. Transient receptor potential cation channels in disease; pp. 165–217. [PMID:17237345] [DOI] [PubMed] [Google Scholar]
  38. Nilius B. Talavera K. Owsianik G. Prenen J. Droogmans G. Voets T. J Physiol (Lond) Vol. 567. 2005. Gating of TRP channels: a voltage connection? pp. 35–44. (Pt 1): [PMID:15878939] [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Nilius B. Voets T. Pflugers Arch. Vol. 451. 2005. TRP channels: a TR(I)P through a world of multifunctional cation channels; pp. 1–10. [PMID:16012814] [DOI] [PubMed] [Google Scholar]
  40. Nilius B. Vriens J. Prenen J. Droogmans G. Voets T. Am J Physiol, Cell Physiol. Vol. 286. 2004. TRPV4 calcium entry channel: a paradigm for gating diversity; pp. C195–205. [PMID:14707014] [DOI] [PubMed] [Google Scholar]
  41. Oberwinkler J. Phillipp SE. Handb Exp Pharmacol. Vol. 179. 2007. TRPM3; pp. 253–267. [PMID:17217062] [DOI] [PubMed] [Google Scholar]
  42. Owsianik G. D'hoedt D. Voets T. Nilius B. Rev Physiol Biochem Pharmacol. Vol. 156. 2006. Structure-function relationship of the TRP channel superfamily; pp. 61–90. [PMID:16634147] [PubMed] [Google Scholar]
  43. Owsianik G. Talavera K. Voets T. Nilius B. Annu Rev Physiol. Vol. 68. 2006. Permeation and selectivity of TRP channels; pp. 685–717. [PMID:16460288] [DOI] [PubMed] [Google Scholar]
  44. Patel A. Sharif-Naeini R. Folgering JR. Bichet D. Duprat F. Honoré E. Pflugers Arch. Vol. 460. 2010. Canonical TRP channels and mechanotransduction: from physiology to disease states; pp. 571–581. [PMID:20490539] [DOI] [PubMed] [Google Scholar]
  45. Pedersen SF. Owsianik G. Nilius B. Cell Calcium. Vol. 38. 2005. TRP channels: an overview; pp. 233–252. [PMID:16098585] [DOI] [PubMed] [Google Scholar]
  46. Penner R. Fleig A. Handb Exp Pharmacol. Vol. 179. 2007. The Mg2+ and Mg(2+)-nucleotide-regulated channel-kinase TRPM7; pp. 313–328. [PMID:17217066] [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Pingle SC. Matta JA. Ahern GP. Handb Exp Pharmacol. Vol. 179. 2007. Capsaicin receptor: TRPV1 a promiscuous TRP channel; pp. 155–171. [PMID:17217056] [DOI] [PubMed] [Google Scholar]
  48. Plant TD. Schaefer M. Cell Calcium. Vol. 33. 2003. TRPC4 and TRPC5: receptor-operated Ca2+-permeable nonselective cation channels; pp. 441–450. [PMID:12765689] [DOI] [PubMed] [Google Scholar]
  49. Potier M. Trebak M. Pflugers Arch. Vol. 457. 2008. New developments in the signaling mechanisms of the store-operated calcium entry pathway; pp. 405–415. [PMID:18536932] [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Puertollano R. Kiselyov K. Am J Physiol Renal Physiol. Vol. 296. 2009. TRPMLs: in sickness and in health; pp. F1245–F1254. [PMID:19158345] [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Putney JW. Physiological mechanisms of TRPC activation. Pflugers Arch. 2005;451:29–34. doi: 10.1007/s00424-005-1416-4. [PMID:16133266] [DOI] [PubMed] [Google Scholar]
  52. Qian F. Noben-Trauth K. Pflugers Arch. Vol. 451. 2005. Cellular and molecular function of mucolipins (TRPML) and polycystin 2 (TRPP2) pp. 277–285. [PMID:15971078] [DOI] [PubMed] [Google Scholar]
  53. Ramsey IS. Delling M. Clapham DE. Annu Rev Physiol. Vol. 68. 2006. An introduction to TRP channels; pp. 619–647. [PMID:16460286] [DOI] [PubMed] [Google Scholar]
  54. Rohacs T. Phosphoinositide regulation of non-canonical transient receptor potential channels. Cell Calcium. 2009;45:554–565. doi: 10.1016/j.ceca.2009.03.011. [PMID:19376575] [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Runnels LW. TRPM6 and TRPM7: A Mul-TRP-PLIK-cation of channel functions. Curr Pharm Biotechnol. 2011;12:42–53. doi: 10.2174/138920111793937880. [PMID:20932259] [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Salido GM. Sage SO. Rosado JA. Biochim Biophys Acta. Vol. 1793. 2009. TRPC channels and store-operated Ca(2+) entry; pp. 223–230. [PMID:19061922] [DOI] [PubMed] [Google Scholar]
  57. Schumacher MA. Eilers H. Front Biosci. Vol. 15. 2010. TRPV1 splice variants: structure and function; pp. 872–882. [PMID:20515731] [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Starowicz K. Nigam S. Di Marzo V. Pharmacol Ther. Vol. 114. 2007. Biochemistry and pharmacology of endovanilloids; pp. 13–33. [PMID:17349697] [DOI] [PubMed] [Google Scholar]
  59. Szallasi A. Cortright DN. Blum CA. Eid SR. Nat Rev Drug Discov. Vol. 6. 2007. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept; pp. 357–372. [PMID:17464295] [DOI] [PubMed] [Google Scholar]
  60. Trebak M. Lemonnier L. Smyth JT. Vazquez G. Putney JW. Handb Exp Pharmacol. Vol. 179. 2007. Phospholipase C-coupled receptors and activation of TRPC channels; pp. 593–614. [PMID:17217081] [DOI] [PubMed] [Google Scholar]
  61. Vay L. Gu C. McNaughton PA. Br J Pharmacol. 2007. The thermo-TRP ion channel family: properties and therapeutic implications. [Epub ahead of print]. [PMID:21797839] [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Venkatachalam K. Montell C. Annu Rev Biochem. Vol. 76. 2011. TRP channels; pp. 387–417. [PMID:17579562] [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Vennekens R. Nilius B. Handb Exp Pharmacol. Vol. 179. 2007. Insights into TRPM4 function, regulation and physiological role; pp. 269–285. [PMID:17217063] [DOI] [PubMed] [Google Scholar]
  64. Vennekens R. Owsianik G. Nilius B. Curr Pharm Des. Vol. 14. 2008. Vanilloid transient receptor potential cation channels: an overview; pp. 18–31. [PMID:18220815] [DOI] [PubMed] [Google Scholar]
  65. Vincent F. Duncton MA. Curr Top Med Chem. Vol. 11. 2011. TRPV4 agonists and antagonists; pp. 2216–2226. [PMID:21671873] [DOI] [PubMed] [Google Scholar]
  66. Voets T. Nilius B. J Physiol (Lond) Vol. 582. 2007. Modulation of TRPs by PIPs; pp. 939–944. (Pt 3): [PMID:17395625] [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Voets T. Owsianik G. Nilius B. Handb Exp Pharmacol. Vol. 179. 2007. TRPM8; pp. 329–344. [PMID:17217067] [DOI] [PubMed] [Google Scholar]
  68. Voets T. Talavera K. Owsianik G. Nilius B. Nat Chem Biol. Vol. 1. 2005. Sensing with TRP channels; pp. 85–92. [PMID:16408004] [DOI] [PubMed] [Google Scholar]
  69. Vriens J. Appendino G. Nilius B. Mol Pharmacol. Vol. 75. 2009. Pharmacology of vanilloid transient receptor potential cation channels; pp. 1262–1279. [PMID:19297520] [DOI] [PubMed] [Google Scholar]
  70. Wissenbach U. Niemeyer BA. Handb Exp Pharmacol. Vol. 179. 2007. TRPV6; pp. 221–234. [PMID:17217060] [DOI] [PubMed] [Google Scholar]
  71. Witzgall R. TRPP2 channel regulation. Handb Exp Pharmacol. 2007;179:363–375. doi: 10.1007/978-3-540-34891-7_22. [PMID:17217069] [DOI] [PubMed] [Google Scholar]
  72. Wu LJ. Sweet TB. Clapham DE. Pharmacol Rev. Vol. 62. 2010. International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family; pp. 381–404. [PMID:20716668] [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Yamamoto S. Takahashi N. Mori Y. Prog Biophys Mol Biol. Vol. 103. 2010. Chemical physiology of oxidative stress-activated TRPM2 and TRPC5 channels; pp. 18–27. [PMID:20553742] [DOI] [PubMed] [Google Scholar]
  74. Yuan JP. Kim MS. Zeng W. Shin DM. Huang G. Worley PF. Muallem S. Channels (Austin) Vol. 3. 2009. TRPC channels as STIM1-regulated SOCs; pp. 221–225. [PMID:19574740] [DOI] [PubMed] [Google Scholar]
  75. Zeevi DA. Frumkin A. Bach G. Biochim Biophys Acta. Vol. 1772. 2007. TRPML and lysosomal function; pp. 851–858. [PMID:17306511] [DOI] [PubMed] [Google Scholar]
  76. Zholos A. Pharmacology of transient receptor potential melastatin channels in the vasculature. Br J Pharmacol. 2010;159:1559–1571. doi: 10.1111/j.1476-5381.2010.00649.x. [PMID:20233227] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Bauer CS. Tran-Van-Minh A. Kadurin I. Dolphin AC. Curr Opin Neurobiol. Vol. 20. 2010. A new look at calcium channel α2δ subunits; pp. 563–571. [PMID:20579869] [DOI] [PubMed] [Google Scholar]
  2. Belardetti F. Zamponi GW. Curr Opin Investig Drugs. Vol. 9. 2008. Linking calcium-channel isoforms to potential therapies; pp. 707–715. [PMID:18600576] [PubMed] [Google Scholar]
  3. Buraei Z. Yang J. Physiol Rev. Vol. 90. 2010. The ß subunit of voltage-gated Ca2+ channels; pp. 1461–1506. [PMID:20959621] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16:521–555. doi: 10.1146/annurev.cellbio.16.1.521. [PMID:11031246] [DOI] [PubMed] [Google Scholar]
  5. Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol. 2011;3:a003947. doi: 10.1101/cshperspect.a003947. [PMID:21746798] [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Catterall WA. Dib-Hajj S. Meisler MH. Pietrobon D. J Neurosci. Vol. 28. 2008. Inherited neuronal ion channelopathies: new windows on complex neurological diseases; pp. 11768–11777. [PMID:19005038] [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Catterall WA. Perez-Reyes E. Snutch TP. Striessnig J. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels; pp. 411–425. [PMID:16382099] [DOI] [PubMed] [Google Scholar]
  8. Davies A. Hendrich J. Van Minh AT. Wratten J. Douglas L. Dolphin AC. Trends Pharmacol Sci. Vol. 28. 2007. Functional biology of the alpha(2)delta subunits of voltage-gated calcium channels; pp. 220–228. [PMID:17403543] [DOI] [PubMed] [Google Scholar]
  9. Dolphin AC. G protein modulation of voltage-gated calcium channels. Pharmacol Rev. 2003;55:607–627. doi: 10.1124/pr.55.4.3. [PMID:14657419] [DOI] [PubMed] [Google Scholar]
  10. Dolphin AC. Calcium channel diversity: multiple roles of calcium channel subunits. Curr Opin Neurobiol. 2009;19:237–244. doi: 10.1016/j.conb.2009.06.006. [PMID:19559597] [DOI] [PubMed] [Google Scholar]
  11. Elmslie KS. Calcium channel blockers in the treatment of disease. J Neurosci Res. 2004;75:733–741. doi: 10.1002/jnr.10872. [PMID:14994334] [DOI] [PubMed] [Google Scholar]
  12. Ertel EA. Campbell KP. Harpold MM. Hofmann F. Mori Y. Perez-Reyes E. Schwartz A. Snutch TP. Tanabe T. Birnbaumer L. Tsien RW. Catterall WA. Neuron. Vol. 25. 2000. Nomenclature of voltage-gated calcium channels; pp. 533–535. [PMID:10774722] [DOI] [PubMed] [Google Scholar]
  13. Gao L. An update on peptide drugs for voltage-gated calcium channels. Recent Pat CNS Drug Discov. 2010;5:14–22. doi: 10.2174/157488910789753558. [PMID:19751208] [DOI] [PubMed] [Google Scholar]
  14. Han TS. Teichert RW. Olivera BM. Bulaj G. Curr Pharm Des. Vol. 14. 2008. Conus venoms - a rich source of peptide-based therapeutics; pp. 2462–2479. [PMID:18781995] [DOI] [PubMed] [Google Scholar]
  15. Hofmann F. Lacinová L. Klugbauer N. Rev Physiol Biochem Pharmacol. Vol. 139. 1999. Voltage-dependent calcium channels: from structure to function; pp. 33–87. [PMID:10453692] [DOI] [PubMed] [Google Scholar]
  16. Kochegarov AA. Pharmacological modulators of voltage-gated calcium channels and their therapeutical application. Cell Calcium. 2003;33:145–162. doi: 10.1016/s0143-4160(02)00239-7. [PMID:12600802] [DOI] [PubMed] [Google Scholar]
  17. Lewis RJ. Garcia ML. Nat Rev Drug Discov. Vol. 2. 2003. Therapeutic potential of venom peptides; pp. 790–802. [PMID:14526382] [DOI] [PubMed] [Google Scholar]
  18. Lory P. Chemin J. Expert Opin Ther Targets. Vol. 11. 2007. Towards the discovery of novel T-type calcium channel blockers; pp. 717–722. [PMID:17465728] [DOI] [PubMed] [Google Scholar]
  19. Nelson MT. Todorovic SM. Perez-Reyes E. Curr Pharm Des. Vol. 12. 2006. The role of T-type calcium channels in epilepsy and pain; pp. 2189–2197. [PMID:16787249] [DOI] [PubMed] [Google Scholar]
  20. Perez-Reyes E. Molecular physiology of low-voltage-activated t-type calcium channels. Physiol Rev. 2003;83:117–161. doi: 10.1152/physrev.00018.2002. [PMID:12506128] [DOI] [PubMed] [Google Scholar]
  21. Pexton T. Moeller-Bertram T. Schilling JM. Wallace MS. Expert Opin Investig Drugs. Vol. 20. 2011. Targeting voltage-gated calcium channels for the treatment of neuropathic pain: a review of drug development; pp. 1277–1284. [PMID:21740292] [DOI] [PubMed] [Google Scholar]
  22. Taylor CP. Angelotti T. Fauman E. Epilepsy Res. Vol. 73. 2007. Pharmacology and mechanism of action of pregabalin: the calcium channel alpha2-delta (alpha2-delta) subunit as a target for antiepileptic drug discovery; pp. 137–150. [PMID:17126531] [DOI] [PubMed] [Google Scholar]
  23. Terlau H. Olivera BM. Physiol Rev. Vol. 84. 2004. Conus venoms: a rich source of novel ion channel-targeted peptides; pp. 41–68. [PMID:14715910] [DOI] [PubMed] [Google Scholar]
  24. Triggle DJ. L-type calcium channels. Curr Pharm Des. 2006;12:443–457. doi: 10.2174/138161206775474503. [PMID:16472138] [DOI] [PubMed] [Google Scholar]
  25. Triggle DJ. Calcium channel antagonists: clinical uses–past, present and future. Biochem Pharmacol. 2007;74:1–9. doi: 10.1016/j.bcp.2007.01.016. [PMID:17276408] [DOI] [PubMed] [Google Scholar]
  26. Trimmer JS. Rhodes KJ. Annu Rev Physiol. Vol. 66. 2004. Localization of voltage-gated ion channels in mammalian brain; pp. 477–519. [PMID:14977411] [DOI] [PubMed] [Google Scholar]
  27. Williams JA. Day M. Heavner JE. Expert Opin Pharmacother. Vol. 9. 2008. Ziconotide: an update and review; pp. 1575–1583. [PMID:18518786] [DOI] [PubMed] [Google Scholar]
  28. Yamamoto T. Takahara A. Curr Top Med Chem. Vol. 9. 2009. Recent updates of N-type calcium channel blockers with therapeutic potential for neuropathic pain and stroke; pp. 377–395. [PMID:19442208] [DOI] [PubMed] [Google Scholar]
  29. Yu FH. Catterall WA. Sci STKE. Vol. 2004. 2004. The VGL-chanome: a protein superfamily specialized for electrical signaling and ionic homeostasis; p. re15. [PMID:15467096] [DOI] [PubMed] [Google Scholar]
  30. Zamponi GW. Lewis RJ. Todorovic SM. Arneric SP. Snutch TP. Brain Res Rev. Vol. 60. 2009. Role of voltage-gated calcium channels in ascending pain pathways; pp. 84–89. [PMID:19162069] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Capasso M. DeCoursey TE. Dyer MJ. Trends Cell Biol. Vol. 21. 2011. pH regulation and beyond: unanticipated functions for the voltage-gated proton channel, HVCN1; pp. 20–28. [PMID:20961760] [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. DeCoursey TE. Voltage-gated proton channels. Cell Mol Life Sci. 2008;65:2554–2573. doi: 10.1007/s00018-008-8056-8. [PMID:18463791] [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. DeCoursey TE. Voltage-gated proton channels: what's next? J Physiol (Lond) 2008;586:5305–5324. doi: 10.1113/jphysiol.2008.161703. (Pt 22): [PMID:18801839] [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. DeCoursey TE. Cherny VV. Curr Pharm Des. Vol. 13. 2007. Pharmacology of voltage-gated proton channels; pp. 2400–2420. [PMID:17692009] [DOI] [PubMed] [Google Scholar]
  5. Tombola F. Ulbrich MH. Isacoff EY. J Physiol (Lond) Vol. 587. 2009. Architecture and gating of Hv1 proton channels; pp. 5325–5329. (Pt 22): [PMID:19915215] [DOI] [PMC free article] [PubMed] [Google Scholar]

Further reading

  1. Andavan GS. Lemmens-Gruber R. Curr Med Chem. Vol. 18. 2011. Voltage-gated sodium channels: mutations, channelopathies and targets; pp. 377–397. [PMID:21143119] [DOI] [PubMed] [Google Scholar]
  2. Baker MD. Wood JN. Trends Pharmacol Sci. Vol. 22. 2001. Involvement of Na+ channels in pain pathways; pp. 27–31. [PMID:11165669] [DOI] [PubMed] [Google Scholar]
  3. Bean BP. The action potential in mammalian central neurons. Nat Rev Neurosci. 2007;8:451–465. doi: 10.1038/nrn2148. [PMID:17514198] [DOI] [PubMed] [Google Scholar]
  4. Cantrell AR. Catterall WA. Nat Rev Neurosci. Vol. 2. 2001. Neuromodulation of Na+ channels: an unexpected form of cellular plasticity; pp. 397–407. [PMID:11389473] [DOI] [PubMed] [Google Scholar]
  5. Catterall WA. Structure and function of voltage-gated ion channels. Annu Rev Biochem. 1995;64:493–531. doi: 10.1146/annurev.bi.64.070195.002425. [PMID:7574491] [DOI] [PubMed] [Google Scholar]
  6. Catterall WA. From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron. 2000;26:13–25. doi: 10.1016/s0896-6273(00)81133-2. [PMID:10798388] [DOI] [PubMed] [Google Scholar]
  7. Catterall WA. Dib-Hajj S. Meisler MH. Pietrobon D. J Neurosci. Vol. 28. 2008. Inherited neuronal ion channelopathies: new windows on complex neurological diseases; pp. 11768–11777. [PMID:19005038] [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Catterall WA. Goldin AL. Waxman SG. Pharmacol Rev. Vol. 57. 2005. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels; pp. 397–409. [PMID:16382098] [DOI] [PubMed] [Google Scholar]
  9. Dib-Hajj SD. Cummins TR. Black JA. Waxman SG. Annu Rev Neurosci. Vol. 33. 2010. Sodium channels in normal and pathological pain; pp. 325–347. [PMID:20367448] [DOI] [PubMed] [Google Scholar]
  10. England S. Groot MJ. Br J Pharmacol. Vol. 158. 2009. Subtype-selective targeting of voltage-gated sodium channels; pp. 1413–1425. [PMID:19845672] [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Fozzard HA. Hanck DA. Physiol Rev. Vol. 76. 1996. Structure and function of voltage-dependent sodium channels: comparison of brain II and cardiac isoforms; pp. 887–926. [PMID:8757791] [DOI] [PubMed] [Google Scholar]
  12. Fozzard HA. Lee PJ. Lipkind GM. Curr Pharm Des. Vol. 11. 2005. Mechanism of local anesthetic drug action on voltage-gated sodium channels; pp. 2671–2686. [PMID:16101448] [DOI] [PubMed] [Google Scholar]
  13. George AL. Inherited disorders of voltage-gated sodium channels. J Clin Invest. 2005;115:1990–1999. doi: 10.1172/JCI25505. [PMID:16075039] [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Goldin AL. Resurgence of sodium channel research. Annu Rev Physiol. 2001;63:871–894. doi: 10.1146/annurev.physiol.63.1.871. [PMID:11181979] [DOI] [PubMed] [Google Scholar]
  15. Han TS. Teichert RW. Olivera BM. Bulaj G. Curr Pharm Des. Vol. 14. 2008. Conus venoms - a rich source of peptide-based therapeutics; pp. 2462–2479. [PMID:18781995] [DOI] [PubMed] [Google Scholar]
  16. Isom LL. Sodium channel beta subunits: anything but auxiliary. Neuroscientist. 2001;7:42–54. doi: 10.1177/107385840100700108. [PMID:11486343] [DOI] [PubMed] [Google Scholar]
  17. Kyle DJ. Ilyin VI. J Med Chem. Vol. 50. 2007. Sodium channel blockers; pp. 2583–2588. [PMID:17489575] [DOI] [PubMed] [Google Scholar]
  18. Lai J. Porreca F. Hunter JC. Gold MS. Annu Rev Pharmacol Toxicol. Vol. 44. 2004. Voltage-gated sodium channels and hyperalgesia; pp. 371–397. [PMID:14744251] [DOI] [PubMed] [Google Scholar]
  19. Lewis RJ. Garcia ML. Nat Rev Drug Discov. Vol. 2. 2003. Therapeutic potential of venom peptides; pp. 790–802. [PMID:14526382] [DOI] [PubMed] [Google Scholar]
  20. Mantegazza M. Curia G. Biagini G. Ragsdale DS. Avoli M. Lancet Neurol. Vol. 9. 2010. Voltage-gated sodium channels as therapeutic targets in epilepsy and other neurological disorders; pp. 413–424. [PMID:20298965] [DOI] [PubMed] [Google Scholar]
  21. Matulenko MA. Scanio MJ. Kort ME. Curr Top Med Chem. Vol. 9. 2009. Voltage-gated sodium channel blockers for the treatment of chronic pain; pp. 362–376. [PMID:19442207] [DOI] [PubMed] [Google Scholar]
  22. Priest BT. Kaczorowski GJ. Expert Opin Ther Targets. Vol. 11. 2007. Blocking sodium channels to treat neuropathic pain; pp. 291–306. [PMID:17298289] [DOI] [PubMed] [Google Scholar]
  23. Priestley T. Voltage-gated sodium channels and pain. Curr Drug Targets CNS Neurol Disord. 2004;3:441–456. doi: 10.2174/1568007043336888. [PMID:15578963] [DOI] [PubMed] [Google Scholar]
  24. Terlau H. Olivera BM. Physiol Rev. Vol. 84. 2004. Conus venoms: a rich source of novel ion channel-targeted peptides; pp. 41–68. [PMID:14715910] [DOI] [PubMed] [Google Scholar]
  25. Trimmer JS. Rhodes KJ. Annu Rev Physiol. Vol. 66. 2004. Localization of voltage-gated ion channels in mammalian brain; pp. 477–519. [PMID:14977411] [DOI] [PubMed] [Google Scholar]
  26. Wood JN. Boorman J. Curr Top Med Chem. Vol. 5. 2005. Voltage-gated sodium channel blockers; target validation and therapeutic potential; pp. 529–537. [PMID:16022675] [DOI] [PubMed] [Google Scholar]
  27. Yu FH. Catterall WA. Sci STKE. Vol. 2004. 2004. The VGL-chanome: a protein superfamily specialized for electrical signaling and ionic homeostasis; p. re15. [PMID:15467096] [DOI] [PubMed] [Google Scholar]

References


Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES