1.A.4 The Transient Receptor Potential Ca2+/Cation Channel (TRP-CC) Family

TRP (transient receptor potential) channels represent a superfamily of cation channels conserved from worms to humans (Vennekens et al. 2012).  They comprise seven subfamilies (TRPC, TRPV, TRPM, TRPN, TRPA, TRPP, and TRPML) that can be consider to be in two classes, group I (TRPC/V/M/N/A) and TRPML (TRP MucoLipin) and TRPP (TRP Polycystin) making up group II (Fine et al. 2019). According to Latorre et al. (2009), TRP channels can be grouped into seven subfamilies based on their amino acid sequence homology: (1) the canonical or classic TRPs, (2) the vanilloid receptor TRPs, (3) the melastatin or long TRPs, (4) ankyrin (whose only member is the transmembrane protein 1 [TRPA1]), (5) TRPN after the nonmechanoreceptor potential C (nonpC), and the more distant cousins, the (6) polycystins and (7) mucolipins. Members of the VIC (1.A.1), RIR-CaC (2.A.3) and TRP-CC (1.A.4). Families have similar transmembrane domain structures, but very different cytosolic doman structures (Mio et al. 2008).  Because of their role as cellular sensors, polymodal activation and gating properties, many TRP channels are activated by a variety of different stimuli and function as signal integrators (Latorre et al., 2009; Montell, 2005; Ramsey et al., 2006). These mammalian proteins have been tabulated revealing their accepted designations, activators and inhibitors, putative interacting proteins and proposed functions (Clapham, 2007). The founding members of the TRP superfamily are the TRPC (TRP canonical) channels, which can be activated following the stimulation of phospholipase C and/or depletion of internal calcium stores (Montell, 2005). TRPC channels regulate nicotine-dependent behavior (Feng et al., 2006).  The intrinsic assembly domains that assure tetrameric TRP channel formation have been reviewed (Schindl and Romanin 2007). They also function by depolarising the membrane potential, which triggers the activation of voltage-gated Ca2+ channels.  Structural differences in the extracellular portions, transmembrane domains and the cytoplasmic domains of TRPC channels have been reviewed (Guo and Chen 2019). Cryo-EM has led to an explosion of TRP structures in the last few years (Cao 2020). These structures have confirmed that TRP channels assemble as tetramers and resemble voltage-gated ion channels in their overall architecture. However, each TRP subtype is endowed with a unique set of soluble domains that may confer diverse regulatory mechanisms. TRP channel structures have revealed sites and mechanisms of action of numerous synthetic and natural compounds, as well as those for endogenous ligands such as lipids, Ca2+, and calmodulin (Cao 2020). The role of TRP channels on taste and pain perception has been reviewed (Aroke et al. 2020). Genomic variability in the TRPV1 gene has been associated with alterations in various pain conditions, and polymorphisms of the TRPV1 gene have been associated with alterations in salty taste sensitivity and salt preference (Aroke et al. 2020). The involvement of  TRPM channels in human diseases has been reviewed (Jimenez et al. 2020). Interfacial binding sites for cholesterol are present on TRP ion channels (Lee 2019). It has been proposed that there is a close relationship between cholesterol binding and TRP channel function (Méndez-Reséndiz et al. 2020). The endogenous and pharmacological ligand-binding sites of TRP channels and their regulatory mechanisms have been reviewed (Zhao et al. 2021).

Alonso-Carbajo et al. 2017 reviewed the functions of these proteins in human vascular smooth muscle and cardiac striated muscle. The mammalian TRP superfamily includes at least 22 genes grouped into three major subfamilies based on sequence homology: TRPV (vanilloid), TRPC (canonical), and TRPM (melastatin). Three additional subfamilies (the 'distant TRPs'), TRPP (polycystin), TRPML (mucolipin), and TRPA bring the total number of TRP-related proteins to around 30. TRP proteins are six transmembrane-domain polypeptide subunits, and four subunits assemble in the plasma membrane to form functional channels. All TRP channels are cation permeable, and most are not selective for monovalent or divalent ions. However, TRPV5 and TRPV6, display specificity for Ca2+ ions, and TRPM4 and TRPM5 are highly selective for monovalent cations and impermeant to Ca2+.  TRP channels are activated by stimuli including changes in pressure, temperature, osmolarity, and intracellular Ca2+.  Fatty acids and receptor-dependent vasoconstrictor agonists also activate vascular TRP channels. Most channels assemble from four identical TRP subunits, but when multiple TRP subunits are coexpressed, heteromeric channels can form (Alonso-Carbajo et al. 2017). Over 75 structures of these proteins have been solved by cryo-EM (Madej and Ziegler 2018). They reveal a lack of an apparent general mechanism underlying channel opening and closing. Similarly, the structures reveal a surprising diversity in which chemical ligands bind TRP channels (Hilton et al. 2019). TRP channels in zhikong scallops (Chlamys farreri) are functional diversity (Peng et al. 2021).

The mammalian TRPM gene family can be subdivided into distinct categories of cation channels that are either highly permeable for Ca2+ (TRPM3/6/7), nonselective (TRPM2/8), or Ca2+ impermeable (TRPM4/5). TRPM6/7 are fused to alpha-kinase domains, whereas TRPM2 is linked to an ADP-ribose phosphohydrolase (Nudix domain). Phylogenetic evidence suggests that Nudix-linked channels represent an ancestral type of TRPM that is present in various phyla, ranging from protists to humans (Mederos y Schnitzler et al., 2008). The pore-forming segments of invertebrate TRPM2-like proteins display high sequence similarity to those of Ca2+-selective TRPMs. Restoration of only two 'ancient' pore residues in human TRPM2 (Q981E/P983Y) increased (4-fold) its permeability for Ca2+. Conversely, introduction of a 'modern' sequence motif into mouse TRPM7 (E1047Q/Y1049P) resulted in the loss of Ca2+ permeation and a linear TRPM2-like current-voltage relationship (Mederos y Schnitzler et al., 2008).  A cooperative knock-on mechanism underpins Ca2+-selective cation permeation in TRPV channels (Ives et al. 2023).

Volatile anaesthetics (VAs) are the most widely used compounds to induce reversible loss of consciousness and maintain general anaesthesia during surgical interventions. VAs depress central nervous system functions mainly through modulation of ion channels in the neuronal membrane, including 2-pore-domain K+ channels, GABA, NMDA receptors and nociceptive and thermosensitive TRP channels expressed in the peripheral nervous system, including TRPV1, TRPA1, TRPM3 and TRPM8 (Kelemen et al. 2020).  Comparison of structures determined of many Trp channels in the absence or presence of activating stimuli revealed similar constrictions in the central ion permeation pathway near the intracellular end of the S6 helices, pointing to a conserved cytoplasmic gate and suggesting that most available Trp channel structures represent non-conducting states (Huffer et al. 2020). Organisms consuming plants use TRP channel agonists as defense mechanisms (Gandhi et al. 2021). Since the expression pattern and ligand sensitivity of TRP channels varies between species, this presents an intriguing evolutionary adaptation to their specific habitat and life cycles. TRP channel levels influence symptoms of patients with pancreatic adenocarcinoma (Chelaru et al. 2022). The blockade of zinc translocation via the inhibition of the TRPC and TRPM channels was shown to be neuroprotective in brain disease (Hong et al. 2023).

The TRP-CC family includes a variety of channel/sensors that respond to temperature, touch, pain, osmolarity, pheromones, taste, and other stimuli (Clapham, 2003). It has also been called the store-operated calcium channel (SOC) family. These proteins are the prinicipal components in mechanosensitive channels in vertebrate hair cells (TRPA1; 1.A.4.6.1) and stretch-activated channels in various vertebrate cell types (TRPC1; 1.A.4.1.3) (Barritt and Rychkov, 2005). TRPA1 and TRPC1 may use different mechanisms of activation. (a) The functional TRPA1 channel is probably a tetramer that is composed of four identical TRPA1 polypeptide chains or a mixture of TRPA1 and another channel polypeptide. Each TRPA1 polypeptide has 17 ankyrin repeats at the cytoplasmic amino terminus. It is proposed that these are coupled to motor proteins or other regulatory proteins on the cytoplasmic face of the plasma membrane (Barritt and Rychkov, 2005). In response to the deflection of the mechanosensitive cilia bundle induced by sound, tension on the ankyrin repeat domains or changes in protein-protein interactions are altered and the channel opens to admit Ca2+ and other cations. (b) The functional TRPC1 channel is probably a tetramer that is composed of four identical TRPC1 polypeptides or a mixture of TRPC1 polypeptides and another polypeptide. Although each TRPC1 polypeptide contains 3 or 4 ankyrin domains at the N terminus, it is proposed that these are not directly involved in channel gating. In response to a stimulus, such as stretching of the membrane by an increase in the volume of the cell, the channel opens and admits Ca2+. It is possible that release of Ca2+ from the endoplasmic reticulum that is induced by thapsigargin also acts as a stimulus, which alters cell volume and therefore can activate TRPC1 through changes in tension of the phospholipid bilayer. The activation of TRP channels by polyunsaturated fatty acids has been examined, and residues involved have been identified (Riehle et al. 2018).

Prototypical members of the TRP-CC family include the Drosophila retinal proteins TRP and TRPL (Montell and Rubin, 1989; Hardie and Minke, 1993). The 81 aas integral membrane INAF-B protein forms a complex with TRP channels, and they stabilize each other (Cheng and Nash, 2007). SOC members of the family mediate the entry of extracellular Ca2+ into cells in response to depletion of intracellular Ca2+ stores (Clapham, 1996) and agonist stimulated production of inositol-1,4,5 trisphosphate (IP3). One member of the TRP-CC family, mammalian Htrp3, has been shown to form a tight complex with the IP3 receptor (TC #1.A.3.2.1). This interaction is apparently required for IP3 to stimulate Ca2+ release via Htrp3. The vanilloid receptor subtype 1 (VR1), which is the receptor for capsaicin (the 'hot' ingredient in chili peppers) and serves as a heat-activated ion channel in the pain pathway (Caterina et al., 1997), is also a member of this family, and is activated by cannabinoids (i.e., anandamide) and certain inflammatory metabolites of arachidonate such as prostaglandin E2 (Olah et al., 2001). The stretch-inhibitable non-selective cation channel (SIC) is identical to the vanilloid receptor throughout all of its first 700 residues, but it exhibits a different sequence in its last 100 residues. VR1 and SIC transport monovalent cations as well as Ca2+. VR1 is about 10x more permeable to Ca2+ than to monovalent ions. Ca2+ overload probably causes cell death after chronic exposure to capsaicin (McCleskey and Gold, 1999). Molecular characteristics and expression profiles of nine TRP channels in the brown planthopper, Nilaparvata lugens, have been described (Wang et al. 2021).

The proteins of the TRP-CC family exhibit the same topological organization with a probable KscA-type 3-dimensional structure (Dodier et al., 2004; Dohke et al., 2004). They consist of about 700-800 (VR1, SIC or ECaC) or 1300 (TRP proteins) amino acyl residues with six transmembrane spanners (TMSs) as well as a short hydrophobic 'loop' region between TMSs 5 and 6. This loop region may dip into the membrane and contribute to the ion permeation pathway (Hardie and Minke, 1993). An aspartate residue in the P-loop may form a ring of negative charges that modulate pore properties including ion selectivity and inhibitory characteristics (García-Martínez et al., 2000). VR1 forms homotetramers. In these respects, members of the TRP-CC family resemble those of the VIC family. When one member of the TRP-CC family, the IGF-regulated Ca2+ channel of Mus musculus (TC #1.A.4.2.4), was PSI-BLASTED, it retrieved a partial sequence of a Zea mays K+ channel protein (887 aas; gbY07632) that is clearly a member of the VIC family. The two homologous protein segments of 150 residues were 28% identical, 42% similar with a PSI-BLAST score (without iterations) of 2e6. This observation further suggests a common origin for certain domains in the TRP-CC and VIC families. Some TRP channels are involved in the development of pathological scarification (PS) and directly participate in PS fibrosis and re-epithelialization or indirectly activate immune cells to release cytokines and neuropeptides, which is subdivided into immune inflammation, fibrosis, pruritus and mechanical forces increased (Zheng et al. 2023). This review elaborates on the characteristics of TRP channels, the mechanism of PS, and how TRP channels mediate the development of PS, summarizes the important role of TRP channels in the different pathogenesis of PS and proposes that therapeutic strategies targeting TRP will be important for the prevention and treatment of PS (Zheng et al. 2023).

All members of the vanilloid family of TRP channels (TRPV) possess an N-terminal ankyrin repeat domain (ARD), which regulates calcium uptake and homeostasis. It is essential for channel assembly and regulation. The 1.7 Å crystal structure of the TRPV6-ARD revealed conserved structural elements unique to the ARDs of TRPV proteins. First, a large twist between the fourth and fifth repeats is induced by residues conserved in all TRPV ARDs. Second, the third finger loop is the most variable region in sequence, length and conformation. In TRPV6, a number of putative regulatory phosphorylation sites map to the base of this third finger. The TRPV6-ARD does not assemble as a tetramer and is monomeric in solution (Phelps et al., 2008). Voltage sensing in thermo-TRP channels has been reviewed by Brauchi and Orio, 2011. TRPV were named after the first identified member TRPV1, that is sensitive to the vanillylamide capsaicin. Six TRPV channel subtypes (TRPV1-6) are subdivided into the thermoTRPV (TRPV1-4) and the Ca2+-selective TRPV channels (TRPV5, TRPV6) (Seebohm and Schreiber 2021). They are not primarily voltage gated but have distinct properties and react to several endogenous ligands as well as different gating stimuli such as heat, pH, mechanical stress, and osmotic changes. Their physiological functions are diverse and tissue specific. They serve as sensors for different pain stimuli (heat, pressure, pH) and contribute to the homeostasis of electrolytes, the maintenance of barrier functions and the development of macrophages. Different aspects of their structures, different gating stimuli, expression patterns, physiological-pathological roles and modulating mechanisms of synthetic, natural and endogenous ligands have been reviewed (Seebohm and Schreiber 2021).

The transient receptor potential (TRP) family of ion channels participate in many signaling pathways. TRPV1 functions as a molecular integrator of noxious stimuli, including heat, low pH, and chemical ligands. The 19-A structure of TRPV1 determined by using single-particle electron cryomicroscopy exhibits fourfold symmetry and comprises two distinct regions: a large open basket-like domain, likely corresponding to the cytoplasmic N- and C-terminal portions, and a more compact domain, corresponding to the transmembrane portion (Moiseenkova et al., 2008). The assignment of transmembrane and cytoplasmic regions was supported by fitting crystal structures of the structurally homologous Kv1.2 channel and isolated TRPV1 ankyrin repeats into the TRPV1 structure. TRP channels assemble as tetramers and resemble voltage-gated ion channels in their overall architecture. But beyond the relatively conserved transmembrane core embedded within the lipid bilayer, each TRP subtype appears to be endowed with a unique set of soluble domains that may confer diverse regulatory mechanisms. TRP channel structures reveal sites and mechanisms of action of numerous synthetic and natural compounds, as well as those for endogenous ligands such as lipids, Ca2+, and calmodulin (Cao 2020). Ca2+-selective TRPV channel permeation occurs by a three-binding site knock-on mechanism, whereas a two-binding site knock-on mechanism is observed in non-selective TRPV channels (Ives et al. 2023). Each of the ion binding sites displays greater affinity for Ca2+ over Na+. As such, coupling to an extra binding site in the Ca2+-selective TRPV channels underpins their increased selectivity for Ca2+ over Na+ ions. Furthermore, analysis of all available TRPV channel structures shows that the selectivity filter entrance region is wider for the non-selective TRPV channels, slightly destabilizing ion binding at this site, which is likely to underlie mechanistic decoupling.

Most local anaesthetics used clinically are relatively hydrophobic molecules that gain access to their blocking site on the sodium channel by diffusing into or through the cell membrane. These anaesthetics block sodium channels and the excitability of neurons. Binshtok et al. (2007) tested the possibility that the excitability of primary sensory nociceptor (pain-sensing) neurons could be blocked by introducing the charged, membrane-impermeant lidocaine derivative QX-314 through the pore of the noxious-heat-sensitive TRPV1 channel (TC #1.A.4.2.1). They found that charged sodium-channel blockers can be targeted into nociceptors by the application of TRPV1 agonists to produce a pain-specific local anaesthesia. QX-314 applied externally had no effect on the activity of sodium channels in small sensory neurons when applied alone, but when applied in the presence of the TRPV1 agonist capsaicin, QX-314 blocked sodium channels and inhibited excitability (Binshtok et al., 2007).

The amino termini of TRP-CC proteins normally contain a proline-rich region and one or more ankyrin domains. VR1, for example, exhibits three such repeat domains in its amino terminal hydrophilic segment (432 amino acids). It also has a hydrophilic C-terminus that lacks recognizable motifs. The sequence similarity between VR1 and other TRP-CC family proteins is within and adjacent to the sixth TMS, including the hydrophobic 'loop' region. Unlike other TRP-CC family members, VR1 is not a SOC. Mammals appear to have multiple VR1 homologues. Fingerprint residues in TRP channels have been identified (Cabezas-Bratesco et al. 2022), revealing a conserved set of residues. This fingerprint is composed of twelve residues localized at equivalent three-dimensional positions in TRP channels from the different subtypes. These amino acids are arranged in three groups, connected by a set of aromatics located at the core of the transmembrane structure. Differences in the connectivities between these different groups of residues harbor the apparent differences in coupling strategies used by TRP subgroups (Cabezas-Bratesco et al. 2022).

One member of the TRP-CC family, TRP-PLIK (1862 aas; AF346629), has been implicated in the regulation of cell division. It has an N-terminal TRP-CC-like sequence and a C-terminal protein kinase-like sequence. It was shown to autophosphorylate and exhibits an ATP phosphorylation-dependent, non-selective, Ca2+-permeable, outward rectifying conductance (Runnels et al., 2001). Another long homologue, Melastatin, is associated with melanocytic tumor progression whereas another homologue, MTR1, is associated with Beckwith-Wiedemann syndrome and a predisposition for neoplasia. Each of these proteins may be present in the cell as several splice variants.

The rabbit kidney epithelial Ca2+ channel, ECaC, is a Ca2+-selective cation channel with monovalent cation transport activity sensitive to strong inhibition by low concentrations of Ca2+ or Mg2+. ECaC is >100 x more permeable to Ca2+ than Na+. Mutation of D542 to alanine (D542A) (not present in the TRP-CC homologue) abolishes Ca2+ permeation and divalent cation inhibition of monovalent cation permeation. The mutation does not inhibit the latter transport activity. The D542K mutation generates a nonfunctional channel. Thus, a single residue determines the characteristic cation selectivity of ECaC.

The ability to detect variations in humidity is critical for many animals. Birds, reptiles and insects all show preferences for specific humidities that influence their mating, reproduction and geographic distribution. Because of their large surface area to volume ratio, insects are particularly sensitive to humidity, and its detection can influence their survival. Two types of hygroreceptors exist in insects: one responds to an increase (moist receptor) and the other to a reduction (dry receptor) in humidity. Although previous data indicated that mechanosensation might contribute to hygrosensation, the cellular basis of hygrosensation and the genes involved in detecting humidity remain unknown. To understand better the molecular bases of humidity sensing,(Liu et al., 2007b) investigated several genes encoding channels associated with mechanosensation, thermosensing or water transport. They identified two Drosophila melanogaster transient receptor potential channels needed for sensing humidity: CG31284, named water witch (wtrw), which is required to detect moist air, and nanchung (nan), which is involved in detecting dry air. Neurons associated with specialized sensory hairs in the third segment of the antenna express these channels. Neurons expressing wtrw and nan project to central nervous system regions associated with mechanosensation (Liu et al., 2007b). The six TRP channels of dinoflagelates do not appear to be mechanoreceptors but rather are components of a mechanotransduction signaling pathway and may be activated via a PLC-dependent mechanism (Lindström et al. 2017).

TRP channels are calcium-permeable nonselective cation channels with six TMS domains and a putative pore loop between TMSs 5 and 6 (Hu et al., 2012). About 28 mammalian TRP channels have been identified, with different numbers of splicing variants for each channel gene. TRP channels have been classified into six different subgroups, including TRPV (1-6), TRPM (1-8), TRPC (1-7), TRPA1, TRPP (1-3), and TRPML (1-3), according to their sequence similarities. In general, TRP channels are involved in calcium handling (e.g., intracellular calcium mobilization and calcium reabsorption) and a broad range of sensory modalities, including pain, temperature, taste, etc. TRP channelopathies are part of important mechanisms in a variety of diseases such as neurodegenerative disorders, diabetes mellitus, inflammatory bowel diseases, epilepsy, cancer, etc. Several members of the TRP family, TRPV1-4, TRPM8, and TRPA1, also called 'ThermoTRPs,' are involved in the detection of temperature changes, thus acting as the molecular thermometers of our body. They are also polymodal nociceptors that integrate painful stimuli such as noxious temperatures and chemical insults. For example, the TRPV1 channel mediates thermal hyperalgesia and pain induced by capsaicin and acid. TRPA1 is a nociceptor that integrates many noxious environmental stimuli including oxidants and electrophilic agents. Gene deletion animals have been created to study the role of TRP channels in pain and nociception; involvement of TRPV1, TRPV3, TRPV4, and TRPA1 in nociception has been confirmed (Hu et al., 2012). 

A class of ion channels that belongs to the transient receptor potential (TRP) superfamily and is present in specialized neurons are temperature detectors. These channels are classified into subfamilies, namely canonical (TRPC), melastatin (TRPM), ankyrin (TRPA), and vanilloid (TRPV). Some of these channels are activated by heat (TRPM2/4/5, TRPV1-4), while others by cold (TRPA1, TRPC5, and TRPM8) (Baez et al. 2014). These channels resemble voltage-dependent K+ channels, with their subunits containing six transmembrane segments that form tetramers. Thermal TRP channels are polymodal receptors that can be activated by temperature, voltage, pH, lipids, and agonists. Their high temperature sensitivity is due to a large enthalpy change ( approximately 100 kcal/mol), which is about five times the enthalpy change in voltage-dependent gating. Pi-helices in TRP channels probably function in gating (Zubcevic and Lee 2019).

TRPV cation channels are polymodal sensors involved in a variety of physiological processes. TRPV2 is regulated by temperature, ligands such as probenecid and cannabinoids, and lipids. It may play a role in somatosensation, osmosensation and innate immunity. Zubcevic et al. 2016 presented the atomic model of rabbit TRPV2 in its putative desensitized state, as determined by cryo-EM at 4 Å resolution. TMS6 (S6), which is involved in gate opening, adopts a conformation different from the one observed in TRPV1. Structural comparisons of TRPV1 and TRPV2 indicate that a rotation of the ankyrin-repeat domain is coupled to pore opening via the TRP domain, and this pore opening can be modulated by rearrangements in the secondary structure of S6.  Non-canonical helical transitions and conformational switching are associated with characteristic regions of flexibility and disorder indices in both TRP and Kv channels (García-Morales and Balleza 2023). Some regions are flexible and disorder, and these regions with potential functional dynamism, and these regions can undergo conformational rearrangements that occur during ligand binding events, the compaction, and refolding of the outer pore loops in several TRP channels, as well as the well-known S4 motion in Kv channels (García-Morales and Balleza 2023).

Plasma membrane ion channels, and in particular TRPC channels, need a specific membrane environment and association with scaffolding, signaling, and cytoskeleton proteins in order to play their important functional roles. TRPC proteins are incorporated into macromolecular complexes including Ca2+ signaling proteins and proteins involved in vesicle trafficking, cytoskeletal interactions, and scaffolding. Association of TRPC with calmodulin (CaM), IP3R, PMCA, Gq/11, RhoA, and a variety of scaffolding proteins has been demonstrated. The interactions between TRPC channels and adaptor proteins determines their modes of regulation as well as their cellular localizations and functions. Adaptor proteins are involved in assembling Ca2+signaling complexes, in the correct sub-cellular localization of protein partners, and in the regulation of TRPC channelosome.The S4 - S5 linker is the gear box of TRP channel gating, and many pathogenic mutations occur in this region (Hofmann et al. 2017). High resolution structures are known for TRPV1, TRPV2, TRPV6, TRPA1, and TRPP2 (Hofmann et al. 2017).

Mechanosensory transduction for senses such as proprioception, touch, balance, acceleration, hearing and pain relies on mechanotransduction channels, which convert mechanical stimuli into electrical signals in specialized sensory cells. There are two major models. One is the membrane-tension model: force applied to the membrane generates a change in membrane tension that is sufficient to gate the channel, as in bacterial MscL channels  (TC# 1.A.22) and certain eukaryotic potassium channels (TC# 1.A.1). The other is the tether model: force is transmitted via a tether to gate the channel. The transient receptor potential (TRP) channel NOMPC is important for mechanosensation-related behaviours such as locomotion, touch and sound sensation across different species including Caenorhabditis elegans, Drosophila and zebrafish. NOMPC is the founding member of the TRPN subfamily, and is thought to be gated by tethering of its ankyrin repeat domain to microtubules of the cytoskeleton (Jin et al. 2017).

The generalized transport reaction catalyzed by TRP-CC family members is:

Ca2+ (out) ⇌ Ca2+ (in)

or

C+ and Ca2+ (out) ⇌ C+ and Ca2+ (in).



This family belongs to the VIC Superfamily.

 

References:

Hu H, Bandell M, Grandl J, Petrus M. (2012) http://www.ncbi.nlm.nih.gov/pubmed?term=22593966



Acharya, T.K., R.P. Sahu, S. Kumar, S. Kumar, T.P. Rokade, R. Chakraborty, N.K. Dubey, D. Shikha, S. Chawla, and C. Goswami. (2022). Function and regulation of thermosensitive ion channel TRPV4 in the immune system. Curr Top Membr 89: 155-188.

Agosto, M.A., I.A. Anastassov, and T.G. Wensel. (2018). Differential epitope masking reveals synapse-specific complexes of TRPM1. Vis Neurosci 35: E001.

Agosto, M.A., Z. Zhang, F. He, I.A. Anastassov, S.J. Wright, J. McGehee, and T.G. Wensel. (2014). Oligomeric State of Purified Transient Receptor Potential Melastatin-1 (TRPM1), a Protein Essential for Dim Light Vision. J. Biol. Chem. 289: 27019-27033.

Ahmed, T., C.R. Nisler, E.C. Fluck, 3rd, S. Walujkar, M. Sotomayor, and V.Y. Moiseenkova-Bell. (2021). Structure of the ancient TRPY1 channel from Saccharomyces cerevisiae reveals mechanisms of modulation by lipids and calcium. Structure. [Epub: Ahead of Print]

Al-Bataineh, M.M., T.A. Sutton, and R.P. Hughey. (2017). Novel roles for mucin 1 in the kidney. Curr Opin Nephrol Hypertens 26: 384-391.

Alonso-Carbajo, L., M. Kecskes, G. Jacobs, A. Pironet, N. Syam, K. Talavera, and R. Vennekens. (2017). Muscling in on TRP channels in vascular smooth muscle cells and cardiomyocytes. Cell Calcium 66: 48-61.

Amantini, C., M. Mosca, M. Nabissi, R. Lucciarini, S. Caprodossi, A. Arcella, F. Giangaspero, and G. Santoni. (2007). Capsaicin-induced apoptosis of glioma cells is mediated by TRPV1 vanilloid receptor and requires p38 MAPK activation. J Neurochem 102: 977-990.

Amarouch, M.Y. and J. El Hilaly. (2020). Inherited Cardiac Arrhythmia Syndromes: Focus on Molecular Mechanisms Underlying TRPM4 Channelopathies. Cardiovasc Ther 2020: 6615038.

Amini, M., H. Wang, A. Belkacemi, M. Jung, A. Bertl, G. Schlenstedt, V. Flockerzi, and A. Beck. (2018). Identification of Inhibitory Ca Binding Sites in the Upper Vestibule of the Yeast Vacuolar TRP Channel. iScience 11: 1-12. [Epub: Ahead of Print]

Argudo, D., S. Capponi, N.P. Bethel, and M. Grabe. (2019). A multiscale model of mechanotransduction by the ankyrin chains of the NOMPC channel. J Gen Physiol. [Epub: Ahead of Print]

Arias-Darraz, L., D. Cabezas, C.K. Colenso, M. Alegría-Arcos, F. Bravo-Moraga, I. Varas-Concha, D.E. Almonacid, R. Madrid, and S. Brauchi. (2015). A transient receptor potential ion channel in Chlamydomonas shares key features with sensory transduction-associated TRP channels in mammals. Plant Cell 27: 177-188.

Aroke, E.N., K.L. Powell-Roach, R.B. Jaime-Lara, M. Tesfaye, A. Roy, P. Jackson, and P.V. Joseph. (2020). Taste the Pain: The Role of TRP Channels in Pain and Taste Perception. Int J Mol Sci 21:.

Arullampalam, P., M.C. Essers, M. Boukenna, S. Guichard, J.S. Rougier, and H. Abriel. (2023). Knockdown of the TRPM4 channel alters cardiac electrophysiology and hemodynamics in a sex- and age-dependent manner in mice. Physiol Rep 11: e15783.

Autzen, H.E., A.G. Myasnikov, M.G. Campbell, D. Asarnow, D. Julius, and Y. Cheng. (2018). Structure of the human TRPM4 ion channel in a lipid nanodisc. Science 359: 228-232.

Aydar, E. and C.P. Palmer. (2009). Polycystic kidney disease channel and synaptotagmin homologues play roles in schizosaccharomyces pombe cell wall synthesis/repair and membrane protein trafficking. J. Membr. Biol. 229: 141-152.

Baez, D., N. Raddatz, G. Ferreira, C. Gonzalez, and R. Latorre. (2014). Gating of thermally activated channels. Curr Top Membr 74: 51-87.

Bähner, M., S. Frechter, N. Da Silva, B. Minke, R. Paulsen, and A. Huber. (2002). Light-regulated subcellular translocation of Drosophila TRPL channels induces long-term adaptation and modifies the light-induced current. Neuron. 34: 83-93.

Bali, A., S.P. Schaefer, I. Trier, A.L. Zhang, L. Kabeche, and C.E. Paulsen. (2023). Molecular mechanism of hyperactivation conferred by a truncation of TRPA1. Nat Commun 14: 2867.

Barritt, G. and G. Rychkov. (2005). TRPs as mechanosensitive channels. Nat. Cell Biol. 7: 105-107.

Batool, L., K. Hariharan, Y. Xu, M. Kaßmann, D. Tsvetkov, B.O. Gohlke, S. Kaden, M. Gossen, B. Nürnberg, A. Kurtz, and M. Gollasch. (2023). An inactivating human TRPC6 channel mutation without focal segmental glomerulosclerosis. Cell Mol Life Sci 80: 265.

Bautista D.M., J. Siemens, J.M. Glazer, P.R. Tsuruda, A.I. Basbaum, C.L. Stucky, S.E. Jordt, D. Julius. (2007). The menthol receptor TRPM8 is the principal detector of environmental cold. Nature. 448: 204-208.

Benemei, S., R. Patacchini, M. Trevisani, and P. Geppetti. (2015). TRP channels. Curr Opin Pharmacol 22: 18-23.

Bertamino, A., N. Iraci, C. Ostacolo, P. Ambrosino, S. Musella, V. Di Sarno, T. Ciaglia, G. Pepe, M. Sala, M.V. Soldovieri, I. Mosca, S. Gonzalez-Rodriguez, A. Fernandez-Carvajal, A. Ferrer-Montiel, E. Novellino, M. Taglialatela, P. Campiglia, and I. Gomez-Monterrey. (2018). Identification of a Potent Tryptophan-Based TRPM8 Antagonist With in Vivo Analgesic Activity. J Med Chem. [Epub: Ahead of Print]

Bhardwaj, R., S. Lindinger, A. Neuberger, K.D. Nadezhdin, A.K. Singh, M.R. Cunha, I. Derler, G. Gyimesi, J.L. Reymond, M.A. Hediger, C. Romanin, and A.I. Sobolevsky. (2020). Inactivation-mimicking block of the epithelial calcium channel TRPV6. Sci Adv 6:.

Bidaux, G., A.S. Borowiec, C. Dubois, P. Delcourt, C. Schulz, F.V. Abeele, G. Lepage, E. Desruelles, A. Bokhobza, E. Dewailly, C. Slomianny, M. Roudbaraki, L. Héliot, J.L. Bonnal, B. Mauroy, P. Mariot, L. Lemonnier, and N. Prevarskaya. (2016). Targeting of short TRPM8 isoforms induces 4TM-TRPM8-dependent apoptosis in prostate cancer cells. Oncotarget. [Epub: Ahead of Print]

Bidaux, G., D. Gordienko, G. Shapovalov, V. Farfariello, A.S. Borowiec, O. Iamshanova, L. Lemonnier, M. Gueguinou, R. Guibon, G. Fromont, M. Paillard, Y. Gouriou, C. Chouabe, E. Dewailly, D. Gkika, P. López-Alvarado, J. Carlos Menéndez, L. Héliot, C. Slomianny, and N. Prevarskaya. (2018). 4TM-TRPM8 channels are new gatekeepers of the ER-mitochondria Ca transfer. Biochim. Biophys. Acta. 1865: 981-994.

Bidaux, G., M. Sgobba, L. Lemonnier, A.S. Borowiec, L. Noyer, S. Jovanovic, A.V. Zholos, and S. Haider. (2015). Functional and Modeling Studies of the Transmembrane Region of the TRPM8 Channel. Biophys. J. 109: 1840-1851.

BINET, L. (1960). [A rural center of medical biology]. Biol Med (Paris) 49: 165-177.

Binshtok, A.M., B.P. Bean, and C.J. Woolf. (2007). Inhibition of nociceptors by TRPV1-mediated entry of impermeant sodium channel blockers. Nature. 449(7162):607-610.

Bohlen, C.J., A. Priel, S. Zhou, D. King, J. Siemens, and D. Julius. (2010). A bivalent tarantula toxin activates the capsaicin receptor, TRPV1, by targeting the outer pore domain. Cell 141: 834-845.

Brauchi, S. and P. Orio. (2011). Voltage sensing in thermo-TRP channels. Adv Exp Med Biol 704: 517-530.

Brizzi, A., S. Maramai, F. Aiello, M.C. Baratto, F. Corelli, C. Mugnaini, M. Paolino, F. Scorzelli, C. Aldinucci, L. De Petrocellis, C. Signorini, and F. Pessina. (2022). Lipoic/Capsaicin-Related Amides: Synthesis and Biological Characterization of New TRPV1 Agonists Endowed with Protective Properties against Oxidative Stress. Int J Mol Sci 23:.

Burks, S.R., R.M. Lorsung, M.E. Nagle, T.W. Tu, and J.A. Frank. (2019). Focused ultrasound activates voltage-gated calcium channels through depolarizing TRPC1 sodium currents in kidney and skeletal muscle. Theranostics 9: 5517-5531.

Burns, D., V. Venditti, and D.A. Potoyan. (2023). Temperature Sensitive Contact Modes Allosterically Gate TRPV3. bioRxiv.

Cabezas-Bratesco D., Brauchi S., Gonzalez-Teuber V., Steinberg X., Valencia I. and Colenso C. (201). The Different Roles of The Channel-Kinases TRPM6 and TRPM7. Curr Med Chem. 22(25):2943-53.

Cabezas-Bratesco, D., F.A. Mcgee, C.K. Colenso, K. Zavala, D. Granata, V. Carnevale, J.C. Opazo, and S.E. Brauchi. (2022). Sequence and structural conservation reveal fingerprint residues in TRP channels. Elife 11:.

Caffrey M., Li D. and Dukkipati A. (2012). Membrane protein structure determination using crystallography and lipidic mesophases: recent advances and successes. Biochemistry. 51(32):6266-88.

Cai X., Srivastava S., Surindran S., Li Z. and Skolnik EY. (2014). Regulation of the epithelial Ca(2)(+) channel TRPV5 by reversible histidine phosphorylation mediated by NDPK-B and PHPT1. Mol Biol Cell. 25(8):1244-50.

Cai, R., X. Liu, R. Zhang, L. Hofmann, W. Zheng, M.R. Amin, L. Wang, Q. Hu, J.B. Peng, M. Michalak, V. Flockerzi, D.W. Ali, X.Z. Chen, and J. Tang. (2020). Autoinhibition of TRPV6 Channel and Regulation by PIP2. iScience 23: 101444. [Epub: Ahead of Print]

Callera, G.E., Y. He, A. Yogi, A.C. Montezano, T. Paravicini, G. Yao, and R.M. Touyz. (2009). Regulation of the novel Mg2+ transporter transient receptor potential melastatin 7 (TRPM7) cation channel by bradykinin in vascular smooth muscle cells. J Hypertens 27: 155-166.

Camacho Londoño, J.E., Q. Tian, K. Hammer, L. Schröder, J. Camacho Londoño, J.C. Reil, T. He, M. Oberhofer, S. Mannebach, I. Mathar, S.E. Philipp, W. Tabellion, F. Schweda, A. Dietrich, L. Kaestner, U. Laufs, L. Birnbaumer, V. Flockerzi, M. Freichel, and P. Lipp. (2015). A background Ca2+ entry pathway mediated by TRPC1/TRPC4 is critical for development of pathological cardiac remodelling. Eur Heart J 36: 2257-2266.

Cao, E. (2020). Structural mechanisms of transient receptor potential ion channels. J Gen Physiol 152:.

Cao, E., M. Liao, Y. Cheng, and D. Julius. (2013). TRPV1 structures in distinct conformations reveal activation mechanisms. Nature 504: 113-118.

Caterina, M.J., M.A. Schumacher, M. Tominaga, T.A. Rosen, J. D. Levine, and D. Julius. (1997). The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature 389: 816-824.

Chandel, A., K.K. Das, and A.K. Bachhawat. (2016). Glutathione depletion activates the yeast vacuolar TRP channel, Yvc1p by reversible glutathionylation of specific cysteines. Mol. Biol. Cell. [Epub: Ahead of Print]

Chang Y., Schlenstedt G., Flockerzi V. and Beck A. (2010). Properties of the intracellular transient receptor potential (TRP) channel in yeast, Yvc1. FEBS Lett. 584(10):2028-32.

Chelaru, N.R., A. Chiosa, A. Sorop, A. Spiridon, F. Cojocaru, D. Domocos, D. Cucu, I. Popescu, and S.O. Dima. (2022). The Association between TRP Channels Expression and Clinicopathological Characteristics of Patients with Pancreatic Adenocarcinoma. Int J Mol Sci 23:.

Chen, J., X.F. Zhang, M.E. Kort, J.R. Huth, C. Sun, L.J. Miesbauer, S.C. Cassar, T. Neelands, V.E. Scott, R.B. Moreland, R.M. Reilly, P.J. Hajduk, P.R. Kym, C.W. Hutchins, and C.R. Faltynek. (2008). Molecular determinants of species-specific activation or blockade of TRPA1 channels. J. Neurosci. 28: 5063-5071.

Cheng Y., Nash H.A. (2007). Drosophila TRP channels require a protein with a distinctive motif encoded by the inaF locus. Proc. Natl. Acad. Sci. U.S.A. 104: 17730-17734.

Cheng, K.T., X. Liu, H.L. Ong, and I.S. Ambudkar. (2008). Functional requirement for Orai1 in store-operated TRPC1-STIM1 channels. J. Biol. Chem. 283: 12935-12940.

Chernov-Rogan, T., E. Gianti, C. Liu, E. Villemure, A.P. Cridland, X. Hu, E. Ballini, W. Lange, H. Deisemann, T. Li, S.I. Ward, D.H. Hackos, S. Magnuson, B. Safina, M.L. Klein, M. Volgraf, V. Carnevale, and J. Chen. (2019). TRPA1 modulation by piperidine carboxamides suggests an evolutionarily conserved binding site and gating mechanism. Proc. Natl. Acad. Sci. USA 116: 26008-26019.

Chou, M.Z., T. Mtui, Y.D. Gao, M. Kohler, and R.E. Middleton. (2004). Resiniferatoxin binds to the capsaicin receptor (TRPV1) near the extracellular side of the S4 transmembrane domain. Biochemistry 43: 2501-2511.

Chu, X., Q. Tong, J. Wozney, W. Zhang, J.Y. Cheung, K. Conrad, V. Mazack, R. Stahl, D.L. Barber, and B.A. Miller. (2005). Identification of an N-terminal TRPC2 splice variant which inhibits calcium influx. Cell Calcium 37: 173-182.

Chu, Y., H. Zhang, M. Yang, and R. Yu. (2023). Molecular Dynamic Simulations Reveal the Activation Mechanisms of Oxidation-Induced TRPV1. Int J Mol Sci 24:.

Chubanov, V., K.P. Schlingmann, J. Waring, J. Heinzinger, S. Kaske, S. Waldegger, M.M. Schnitzler, and T. Gudermann. (2007). Hypomagnesemia with secondary hypocalcemia due to a missense mutation in the putative pore-forming region of TRPM6. J. Biol. Chem. 282: 7656-7667.

Chubanov, V., S. Waldegger, M.M. y Schnitzler, H. Vitzthum, M.C. Sassen, H.W. Seyberth, M. Konrad, and T. Gudermann. (2004). Disruption of TRPM6/TRPM7 complex formation by a mutation in the TRPM6 gene causes hypomagnesemia with secondary hypocalcemia. Proc. Natl. Acad. Sci. USA 101: 2894-2899.

Chyb, S., P. Raghu, and R.C. Hardie. (1999). Polyunsaturated fatty acids activate the Drosophila light-sensitive channels TRP and TRPL. Nature 397: 255-259.

Clapham D.E. (2007). SnapShot: mammalian TRP channels. Cell. 129: 220.

Clapham, D.E. (1996). TRP is cracked, but is CRAC TRP? Neuron 16: 1069-1072.

Clapham, D.E. (2003). TRP channels as cellular sensors. Nature 426: 517-524.

Clarke, A., K. Groschner, and T. Stockner. (2022). Exploring TRPC3 Interaction with Cholesterol through Coarse-Grained Molecular Dynamics Simulations. Biomolecules 12:.

Cruz-Torres, I., D.S. Backos, and P.S. Herson. (2020). Characterization and Optimization of the Novel Transient Receptor Potential Melastatin 2 Antagonist tatM2NX. Mol Pharmacol 97: 102-111.

Csanády, L. and B. Törocsik. (2009). Four Ca2+ ions activate TRPM2 channels by binding in deep crevices near the pore but intracellularly of the gate. J Gen Physiol 133: 189-203.

D'hoedt, D., G. Owsianik, J. Prenen, M.P. Cuajungco, C. Grimm, S. Heller, T. Voets, and B. Nilius. (2008). Stimulus-specific modulation of the cation channel TRPV4 by PACSIN 3. J. Biol. Chem. 283(10): 6272-6280.

Damak, S., M. Rong, K. Yasumatsu, Z. Kokrashvili, C.A. Pérez, N. Shigemura, R. Yoshida, B. Mosinger, Jr, J.I. Glendinning, Y. Ninomiya, and R.F. Margolskee. (2006). Trpm5 null mice respond to bitter, sweet, and umami compounds. Chem Senses 31: 253-264.

Damann, N., G. Bahrenberg, H. Stockhausen, C.J. Habermann, B. Lesch, R. Frank-Foltyn, J. Lee, J. Ann, and T. Christoph. (2020). In vitro characterization of the thermoneutral transient receptor potential vanilloid-1 (TRPV1) inhibitor GRTE16523. Eur J Pharmacol 871: 172934.

Demion, M., P. Bois, P. Launay, and R. Guinamard. (2007). TRPM4, a Ca2+-activated nonselective cation channel in mouse sino-atrial node cells. Cardiovasc Res 73: 531-538.

Diver, M.M., Y. Cheng, and D. Julius. (2019). Structural insights into TRPM8 inhibition and desensitization. Science. [Epub: Ahead of Print]

Dodier, Y., U. Banderali, H. Klein, O. Topalak, O. Dafi, M. Simoes, G. Bernatchez, R. Sauvé, and L. Parent. (2004). Outer pore topology of the ECaC-TRPV5 channel by cysteine scan mutagenesis. J. Biol. Chem. 279: 6853-6862.

Dohke, Y., Y.S. Oh, I.S. Ambudkar, and R.J. Turner. (2004). Biogenesis and topology of the transient receptor potential Ca2+ channel TRPC1. J. Biol. Chem. 279: 12242-12248.

Donate-Macian P., Bano-Polo M., Vazquez-Ibar JL., Mingarro I. and Peralvarez-Marin A. (2015). Molecular and topological membrane folding determinants of transient receptor potential vanilloid 2 channel. Biochem Biophys Res Commun. 462(3):221-6.

Du, E.J., T.J. Ahn, I. Kwon, J.H. Lee, J.H. Park, S.H. Park, T.M. Kang, H. Cho, T.J. Kim, H.W. Kim, Y. Jun, H.J. Lee, Y.S. Lee, J.Y. Kwon, and K. Kang. (2016). TrpA1 Regulates Defecation of Food-Borne Pathogens under the Control of the Duox Pathway. PLoS Genet 12: e1005773.

Duan, J., J. Li, B. Zeng, G.L. Chen, X. Peng, Y. Zhang, J. Wang, D.E. Clapham, Z. Li, and J. Zhang. (2018). Structure of the mouse TRPC4 ion channel. Nat Commun 9: 3102.

Eigenbrod, O., K.Y. Debus, J. Reznick, N.C. Bennett, O. Sánchez-Carranza, D. Omerbašić, D.W. Hart, A.J. Barker, W. Zhong, H. Lutermann, J.V. Katandukila, G. Mgode, T.J. Park, and G.R. Lewin. (2019). Rapid molecular evolution of pain insensitivity in multiple African rodents. Science 364: 852-859.

Fan, C., W. Choi, W. Sun, J. Du, and W. Lu. (2018). Structure of the human lipid-gated cation channel TRPC3. Elife 7:.

Feng, Z., W. Li, A. Ward, B.J. Piggott, E.R. Larkspur, P.W. Sternberg, and X.Z. Xu. (2006). A C. elegans model of nicotine-dependent behavior: regulation by TRP-family channels. Cell 127: 621-633.

Fine, M., X. Li, and S. Dang. (2019). Structural insights into group II TRP channels. Cell Calcium 86: 102107. [Epub: Ahead of Print]

Fu, H., Z. Jiao, Y. Li, J. Tian, L. Ren, F. Zhang, Q. Li, and S. Liu. (2021). Transient Receptor Potential (TRP) Channels in the Pacific Oyster (): Genome-Wide Identification and Expression Profiling after Heat Stress between and. Int J Mol Sci 22:.

Gandhi, A., R. Kariyat, A. Harikishore, M. Ayati, A. Bhunia, and N. Sahoo. (2021). Deciphering the Role of Ion Channels in Early Defense Signaling against Herbivorous Insects. Cells 10:.

Gao, X., C.W. Kuo, A. Main, E. Brown, F.J. Rios, L.L. Camargo, S. Mary, K. Wypijewski, C. Gök, R.M. Touyz, and W. Fuller. (2022). Palmitoylation regulates cellular distribution of and transmembrane Ca flux through TrpM7. Cell Calcium 106: 102639.

García-Martínez, C., C. Morenilla-Palao, R. Planells-Cases, J.M. Merino, and A. Ferrer-Montiel. (2000). Identification of an aspartic residue in the P-loop of the vanilloid receptor that modulates pore properties. J. Biol. Chem. 275: 32552-32558.

García-Morales, A. and D. Balleza. (2023). Non-canonical helical transitions and conformational switching are associated with characteristic flexibility and disorder indices in TRP and Kv channels. Channels (Austin) 17: 2212349.

García-Sanz, N., P. Valente, A. Gomis, A. Fernández-Carvajal, G. Fernández-Ballester, F. Viana, C. Belmonte, and A. Ferrer-Montiel. (2007). A role of the transient receptor potential domain of vanilloid receptor I in channel gating. J. Neurosci. 27: 11641-11650.

Gavva, N.R., L. Klionsky, Y. Qu, L. Shi, R. Tamir, S. Edenson, T.J. Zhang, V.N. Viswanadhan, A. Toth, L.V. Pearce, T.W. Vanderah, F. Porreca, P.M. Blumberg, J. Lile, Y. Sun, K. Wild, J.C. Louis, and J.J. Treanor. (2004). Molecular determinants of vanilloid sensitivity in TRPV1. J. Biol. Chem. 279: 20283-20295.

Gawalska, A., M. Kołaczkowski, and A. Bucki. (2022). Structural Modeling of TRPA1 Ion Channel-Determination of the Binding Site for Antagonists. Molecules 27:.

Gevaert, T., J. Vriens, A. Segal, W. Everaerts, T. Roskams, K. Talavera, G. Owsianik, W. Liedtke, D. Daelemans, I. Dewachter, F. van Leuven, T. Voets, D. de Ridder, and B. Nilius. (2007). Deletion of the transient receptor potential cation channel TRPV4 (Trp12) impairs murine bladder voiding. J. Clin. Invest. 117(11): 3453-3462.

Ghata, J. and B.D. Cowley, Jr. (2017). Polycystic Kidney Disease. Compr Physiol 7: 945-975.

Gochman, A., X.F. Tan, C. Bae, H. Chen, K.J. Swartz, and A. Jara-Oseguera. (2023). Cannabidiol sensitizes TRPV2 channels to activation by 2-APB. Elife 12:.

Goetzl, E.J., V.H. Srihari, M. Mustapic, D. Kapogiannis, and G.R. Heninger. (2022). Abnormal levels of mitochondrial Ca channel proteins in plasma neuron-derived extracellular vesicles of early schizophrenia. FASEB J. 36: e22466.

Gong, J., N.K. Nirala, J. Chen, F. Wang, P. Gu, Q. Wen, Y.T. Ip, and Y. Xiang. (2023). TrpA1 is a shear stress mechanosensing channel regulating intestinal stem cell proliferation in. Sci Adv 9: eadc9660.

Gopal, S., P. Søgaard, H.A. Multhaupt, C. Pataki, E. Okina, X. Xian, M.E. Pedersen, T. Stevens, O. Griesbeck, P.W. Park, R. Pocock, and J.R. Couchman. (2015). Transmembrane proteoglycans control stretch-activated channels to set cytosolic calcium levels. J. Cell Biol. 210: 1199-1211.

Groppi, S., F. Belotti, R.L. Brandão, E. Martegani, and R. Tisi. (2011). Glucose-induced calcium influx in budding yeast involves a novel calcium transport system and can activate calcineurin. Cell Calcium 49: 376-386.

Gualdani, R., P. Gailly, J.H. Yuan, X. Yerna, G. Di Stefano, A. Truini, G. Cruccu, S.D. Dib-Hajj, and S.G. Waxman. (2022). A TRPM7 mutation linked to familial trigeminal neuralgia: Omega current and hyperexcitability of trigeminal ganglion neurons. Proc. Natl. Acad. Sci. USA 119: e2119630119.

Guo, C., X. Yang, H. Shi, C. Chen, Z. Hu, X. Zheng, X. Yang, and C. Xie. (2022). Identification of VdASP F2-interacting protein as a regulator of microsclerotial formation in Verticillium dahliae. Microb Biotechnol. [Epub: Ahead of Print]

Guo, J., J. She, W. Zeng, Q. Chen, X.C. Bai, and Y. Jiang. (2017). Structures of the calcium-activated, non-selective cation channel TRPM4. Nature 552: 205-209.

Guo, W. and L. Chen. (2019). Recent progress in structural studies on canonical TRP ion channels. Cell Calcium 83: 102075. [Epub: Ahead of Print]

Guo, W., Q. Tang, M. Wei, Y. Kang, J.X. Wu, and L. Chen. (2022). Structural mechanism of human TRPC3 and TRPC6 channel regulation by their intracellular calcium-binding sites. Neuron. [Epub: Ahead of Print]

Guo, Y., Y. Song, W. Liu, T. Wang, X. Ma, and Z. Yu. (2023). Novel Insights into the Role of Keratinocytes-Expressed TRPV3 in the Skin. Biomolecules 13:.

Hagmann, H., N.H. Khayyat, C. Oezel, A. Papadakis, A. Kuczkowski, T. Benzing, E. Gulbins, S. Dryer, and P.T. Brinkkoetter. (2022). Paraoxonase 2 (PON2) Deficiency Reproduces Lipid Alterations of Diabetic and Inflammatory Glomerular Disease and Affects TRPC6 Signaling. Cells 11:.

Hagmann, H., N.H. Khayyat, M. Matin, C. Oezel, H. Chen, A. Schauss, C. Schell, T. Benzing, S. Dryer, and P.T. Brinkkoetter. (2023). Capsazepine (CPZ) Inhibits TRPC6 Conductance and Is Protective in Adriamycin-Induced Nephropathy and Diabetic Glomerulopathy. Cells 12:.

Haladyna, J.N., T. Pastuer, S.S. Riedel, A.L. Perraud, and K.M. Bernt. (2016). Transient potential receptor melastatin-2 (Trpm2) does not influence murine MLL-AF9-driven AML leukemogenesis or in vitro response to chemotherapy. Exp Hematol. [Epub: Ahead of Print]

Hardie, R.C. and B. Minke. (1993). Novel Ca2+ channels underlying transduction in Drosophila photoreceptors: implications for phosphoinositide-mediated Ca2+ mobilization. Trends Neurosci 16: 371-376.

He, Y., G. Yao, C. Savoia, and R.M. Touyz. (2005). Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: role of angiotensin II. Circ Res 96: 207-215.

He, Z., C. Yang, D. Jiang, X. Wang, Z. Xing, S. Yu, Q. Yang, and L. Wang. (2022). The expression profile of a multi-stress inducible transient receptor potential vanilloid 4 (TRPV4) in Pacific oyster. Fish Shellfish Immunol Rep 3: 100064.

Held, K., F. Gruss, V.D. Aloi, A. Janssens, C. Ulens, T. Voets, and J. Vriens. (2018). Mutations in the voltage-sensing domain affect the alternative ion permeation pathway in the TRPM3 channel. J. Physiol. [Epub: Ahead of Print]

Hellwig, N., N. Albrecht, C. Harteneck, G. Schultz, and M. Schaefer. (2005). Homo- and heteromeric assembly of TRPV channel subunits. J Cell Sci 118: 917-928.

Hilton, J.K., M. Kim, and W.D. Van Horn. (2019). Structural and Evolutionary Insights Point to Allosteric Regulation of TRP Ion Channels. Acc Chem Res. [Epub: Ahead of Print]

Hilton, J.K., T. Salehpour, N.J. Sisco, P. Rath, and W.D. Van Horn. (2018). Phosphoinositide-interacting regulator of TRP (PIRT) has opposing effects on human and mouse TRPM8 ion channels. J. Biol. Chem. [Epub: Ahead of Print]

Hoenderop, J.G., A.W. van der Kemp, A. Hartog, S.F. van de Graaf, C.H. van Os, P.H. Willems, and R.J. Bindels. (1999). Molecular identification of the apical Ca2+ channel in 1, 25-dihydroxyvitamin D3-responsive epithelia. J. Biol. Chem. 274: 8375-8378.

Hoenderop, J.G.J., T. Voets, S. Hoefs, F. Weidema, J. Prenen, B. Nilius, and R.J.M. Bindels. (2003). Homo- and heterotetrameric architecture of the epithelial Ca2+ channels TRPV5 and TRPV6. EMBO J. 22: 776-785.

Hofmann, L., H. Wang, A. Beck, U. Wissenbach, and V. Flockerzi. (2016). A conserved gating element in TRPV6 channels. Cell Calcium. [Epub: Ahead of Print]

Hofmann, L., H. Wang, W. Zheng, S.E. Philipp, P. Hidalgo, A. Cavalié, X.Z. Chen, A. Beck, and V. Flockerzi. (2017). The S4---S5 linker - gearbox of TRP channel gating. Cell Calcium. [Epub: Ahead of Print]

Hong, D.K., A.R. Kho, S.H. Lee, B.S. Kang, M.K. Park, B.Y. Choi, and S.W. Suh. (2023). Pathophysiological Roles of Transient Receptor Potential (Trp) Channels and Zinc Toxicity in Brain Disease. Int J Mol Sci 24:.

Hough, A., C. Criswell, A. Faruk, J.E. Cavanaugh, B.J. Kolber, and K.J. Tidgewell. (2023). Barbamide Displays Affinity for Membrane-Bound Receptors and Impacts Store-Operated Calcium Entry in Mouse Sensory Neuron.s. Mar Drugs 21:.

Huang, W., Y. Wu, N. Luo, X. Shuai, J. Guo, C. Wang, F. Yang, L. Liu, S. Liu, and Z. Cheng. (2023). Identification of TRPM2 as a prognostic factor correlated with immune infiltration in ovarian cancer. J Ovarian Res 16: 169.

Huang, Y., B. Roth, W. Lü, and J. Du. (2019). Ligand recognition and gating mechanism through three ligand-binding sites of human TRPM2 channel. Elife 8:.

Huffer, K.E., A.A. Aleksandrova, A. Jara-Oseguera, L.R. Forrest, and K.J. Swartz. (2020). Global alignment and assessment of TRP channel transmembrane domain structures to explore functional mechanisms. Elife 9:.

Hughes, T.E., J.S. Del Rosario, A. Kapoor, A.T. Yazici, Y. Yudin, E.C. Fluck, 3rd, M. Filizola, T. Rohacs, and V.Y. Moiseenkova-Bell. (2019). Structure-based characterization of novel TRPV5 inhibitors. Elife 8:.

Iannone, L.F., R. Nassini, R. Patacchini, P. Geppetti, and F. De Logu. (2023). Neuron.al and non-neuronal TRPA1 as therapeutic targets for pain and headache relief. Temperature (Austin) 10: 50-66.

Inoue, K., D. Branigan, and Z.G. Xiong. (2010). Zinc-induced neurotoxicity mediated by transient receptor potential melastatin 7 channels. J. Biol. Chem. 285: 7430-7439.

Ives, C.M., N.J. Thomson, and U. Zachariae. (2023). A cooperative knock-on mechanism underpins Ca2+-selective cation permeation in TRPV channels. J Gen Physiol 155:.

Jang, W., M. Oh, E.H. Cho, M. Baek, and C. Kim. (2023). Drosophila pain sensitization and modulation unveiled by a novel pain model and analgesic drugs. PLoS One 18: e0281874.

Jimenez, I., Y. Prado, F. Marchant, C. Otero, F. Eltit, C. Cabello-Verrugio, O. Cerda, and F. Simon. (2020). TRPM Channels in Human Diseases. Cells 9:.

Jin, P., D. Bulkley, Y. Guo, W. Zhang, Z. Guo, W. Huynh, S. Wu, S. Meltzer, T. Cheng, L.Y. Jan, Y.N. Jan, and Y. Cheng. (2017). Electron cryo-microscopy structure of the mechanotransduction channel NOMPC. Nature 547: 118-122.

Jirku, M., Z. Lansky, L. Bednarova, M. Sulc, L. Monincova, P. Majer, L. Vyklicky, J. Vondrasek, J. Teisinger, and K. Bousova. (2016). The characterization of a novel S100A1 binding site in the N-terminus of TRPM1. Int J Biochem. Cell Biol. [Epub: Ahead of Print]

Jo, A.O., M. Lakk, A.M. Frye, T.T. Phuong, S.N. Redmon, R. Roberts, B.A. Berkowitz, O. Yarishkin, and D. Križaj. (2016). Differential volume regulation and calcium signaling in two ciliary body cell types is subserved by TRPV4 channels. Proc. Natl. Acad. Sci. USA 113: 3885-3890.

John Haynes, W., X.L. Zhou, Z.W. Su, S.H. Loukin, Y. Saimi, and C. Kung. (2008). Indole and other aromatic compounds activate the yeast TRPY1 channel. FEBS Lett. 582: 1514-1518.

Jordt, S.-E. and D. Julius. (2002). Molecular basis for species-specific sensitivity to "hot" chili peppers. Cell 108: 421-430.

Jordt, S.E., D.M. Bautista, H.H. Chuang, D.D. McKemy, P.M. Zygmunt, E.D. Hogestatt, I.D. Meng, and D. Julius. (2004). Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature 427: 260-265.

Jorgensen, C. and C. Domene. (2018). Location and Character of Volatile General Anesthetics Binding Sites in the Transmembrane Domain of TRPV1. Mol Pharm 15: 3920-3930.

Kádár, K., V. Juhász, A. Földes, R. Rácz, Y. Zhang, H. Löchli, E. Kató, L. Köles, M.C. Steward, P. DenBesten, G. Varga, and &.#.1.9.3.;. Zsembery. (2021). TRPM7-Mediated Calcium Transport in HAT-7 Ameloblasts. Int J Mol Sci 22:.

Kang, L., J. Gao, W.R. Schafer, Z. Xie, and X.Z. Xu. (2010). C. elegans TRP family protein TRP-4 is a pore-forming subunit of a native mechanotransduction channel. Neuron. 67: 381-391.

Kashio, M., S. Masubuchi, and M. Tominaga. (2022). Protein kinase C-mediated phosphorylation of transient receptor potential melastatin type 2 Thr738 counteracts the effect of cytosolic Ca and elevates the temperature threshold. J. Physiol. [Epub: Ahead of Print]

Katz, B., T. Oberacker, D. Richter, H. Tzadok, M. Peters, B. Minke, and A. Huber. (2013). Drosophila TRP and TRPL are assembled as homomultimeric channels in vivo. J Cell Sci 126: 3121-3133.

Kedei, N., T. Szabo, J.D. Lile, J.J. Treanor, Z. Olah, M.J. Iadarola, and P.M. Blumberg. (2001). Analysis of the native quaternary structure of vanilloid receptor 1. J. Biol. Chem. 276: 28613-28619.

Kelemen, B., E. Lisztes, A. Vladár, M. Hanyicska, J. Almássy, A. Oláh, A.G. Szöllősi, Z. Pénzes, J. Posta, T. Voets, T. Bíró, and B.I. Tóth. (2020). Volatile anaesthetics inhibit the thermosensitive nociceptor ion channel transient receptor potential melastatin 3 (TRPM3). Biochem Pharmacol 174: 113826.

Kemp, B.J., D.L. Church, J. Hatzold, B. Conradt, and E.J. Lambie. (2009). Gem-1 encodes an SLC16 monocarboxylate transporter-related protein that functions in parallel to the gon-2 TRPM channel during gonad development in Caenorhabditis elegans. Genetics 181: 581-591.

Kim, J., Y.D. Chung, D. Park, S. Choi, D.W. Shin, H. Soh, H.W. Lee, W. Son, J. Yim, C.-S. Park, M.J. Kernan, and C. Kim. (2003). A TRPV family ion channel required for hearing in Drosophila. Nature 424: 81-82.

Kim, S.J., G.H. Park, D. Kim, J. Lee, H. Min, E. Wall, C.J. Lee, M.I. Simon, S.J. Lee, and S.K. Han. (2011). Analysis of cellular and behavioral responses to imiquimod reveals a unique itch pathway in transient receptor potential vanilloid 1 (TRPV1)-expressing neurons. Proc. Natl. Acad. Sci. USA 108: 3371-3376.

Kim, S.J., Y.S. Kim, J.P. Yuan, R.S. Petralia, P.F. Worley, and D.J. Linden. (2003). Activation of the TRPC1 cation channel by metabotropic glutamate receptor mGluR1. Nature 426: 285-291.

Kiselyov, K., X. Xu, G. Mozhayeva, T. Kuo, I. Pessah, G. Mignery, X. Zhu, L. Birnbaumer, and S. Muallem. (1998). Functional interaction between InsP3 receptors and store-operated Htrp3 channels. Nature 396: 478-482.

Knowles, H., J.W. Heizer, Y. Li, K. Chapman, C.A. Ogden, K. Andreasen, E. Shapland, G. Kucera, J. Mogan, J. Humann, L.L. Lenz, A.D. Morrison, and A.L. Perraud. (2011). Transient Receptor Potential Melastatin 2 (TRPM2) ion channel is required for innate immunity against Listeria monocytogenes. Proc. Natl. Acad. Sci. USA 108: 11578-11583.

Ko, K.D., G. Bhardwaj, Y. Hong, G.S. Chang, K. Kiselyov, D.B. van Rossum, and R.L. Patterson. (2009). Phylogenetic profiles reveal structural/functional determinants of TRPC3 signal-sensing antennae. Commun Integr Biol 2: 133-137.

Kon, S., A. Takaku, F. Toyama, E. Takayama-Watanabe, and A. Watanabe. (2019). Acrosome reaction-inducing substance triggers two different pathways of sperm intracellular signaling in newt fertilization. Int J Dev Biol 63: 589-595.

Krapivinsky, G., L. Krapivinsky, Y. Manasian, and D.E. Clapham. (2014). The TRPM7 Chanzyme Is Cleaved to Release a Chromatin-Modifying Kinase. Cell 157: 1061-1072.

Kremeyer, B., F. Lopera, J.J. Cox, A. Momin, F. Rugiero, S. Marsh, C.G. Woods, N.G. Jones, K.J. Paterson, F.R. Fricker, A. Villegas, N. Acosta, N.G. Pineda-Trujillo, J.D. Ramírez, J. Zea, M.W. Burley, G. Bedoya, D.L. Bennett, J.N. Wood, and A. Ruiz-Linares. (2010). A gain-of-function mutation in TRPA1 causes familial episodic pain syndrome. Neuron. 66: 671-680.

Kühn, F.J., G. Knop, and A. Lückhoff. (2007). The transmembrane segment S6 determines cation versus anion selectivity of TRPM2 and TRPM8. J. Biol. Chem. 282: 27598-27609.

Kumar, A., A.K. Mishra, V. Singh, S. Yadav, A. Saxena, S.K. Garg, and D.K. Swain. (2019). Molecular and functional insights into Transient Receptor Potential Vanilloid 1 (TRPV1) in bull spermatozoa. Theriogenology 128: 207-217.

Kurganov, E., S. Saito, C.T. Saito, and M. Tominaga. (2017). Requirement of extracellular Ca2+ binding to specific amino acids for heat-evoked activation of TRPA1. J. Physiol. [Epub: Ahead of Print]

Lai, Y.H., W. Bäumer, C. Meneses, D.M. Roback, J.B. Robertson, S.K. Mishra, B.D.X. Lascelles, and M.W. Nolan. (2021). Irradiation of the Normal Murine Tongue Causes Upregulation and Activation of Transient Receptor Potential (TRP) Ion Channels. Radiat Res. [Epub: Ahead of Print]

Lambers, T.T., A.F. Weidema, B. Nilius, J.G. Hoenderop, and R.J. Bindels. (2004). Regulation of the mouse epithelial Ca2(+) channel TRPV6 by the Ca2+-sensor calmodulin. J. Biol. Chem. 279: 28855-28861.

Lan, L., H. Brereton, and G.J. Barritt. (1998). The role of calmodulin-binding sites in the regulation of the Drosophila TRPL cation channel expressed in Xenopus laevis oocytes by ca2+, inositol 1,4,5-trisphosphate and GTP-binding proteins. Biochem. J. 330(Pt3): 1149-1158.

Lansky, S., J.M. Betancourt, J. Zhang, Y. Jiang, E.D. Kim, N. Paknejad, C.M. Nimigean, P. Yuan, and S. Scheuring. (2023). A pentameric TRPV3 channel with a dilated pore. Nature 621: 206-214.

Latorre, R., C. Zaelzer, and S. Brauchi. (2009). Structure-functional intimacies of transient receptor potential channels. Q. Rev. Biophys. 42: 201-246.

Launay, P., A. Fleig, A.-L. Perraud, A.M. Scharenberg, R. Penner, and J.-P. Kinet. (2002). TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell 109: 397-407.

Laursen, W.J., E.O. Anderson, L.J. Hoffstaetter, S.N. Bagriantsev, and E.O. Gracheva. (2015). Species-specific temperature sensitivity of TRPA1. Temperature (Austin) 2: 214-226.

Laursen, W.J., S.N. Bagriantsev, and E.O. Gracheva. (2014). TRPA1 channels: chemical and temperature sensitivity. Curr Top Membr 74: 89-112.

Lee, A.G. (2019). Interfacial Binding Sites for Cholesterol on TRP Ion Channels. Biophys. J. 117: 2020-2033.

Lee, G., J. Choi, Y.J. Nam, M.J. Song, J.K. Kim, W.J. Kim, P. Kim, J.S. Lee, S. Kim, K.T. No, J.H. Lee, J.K. Lee, and Y. Choi. (2019). Identification and characterization of saikosaponins as antagonists of transient receptor potential A1 channel. Phytother Res. [Epub: Ahead of Print]

Lee, Y., Y. Lee, J. Lee, S. Bang, S. Hyun, J. Kang, S.T. Hong, E. Bae, B.K. Kaang, and J. Kim. (2005). Pyrexia is a new thermal transient receptor potential channel endowing tolerance to high temperatures in Drosophila melanogaster. Nat. Genet. 37: 305-310.

Leffler, A., A. Lattrell, S. Kronewald, F. Niedermirtl, and C. Nau. (2011). Activation of TRPA1 by membrane permeable local anesthetics. Mol Pain 7: 62.

Leffler, A., M.J. Fischer, D. Rehner, S. Kienel, K. Kistner, S.K. Sauer, N.R. Gavva, P.W. Reeh, and C. Nau (2008). The vanilloid receptor TRPV1 is activated and sensitized by local anesthetics in rodent sensory neurons. J Cl- in Invest 118: 763-776.

Lei, M., P. Wang, H. Li, X. Liu, J. Shu, Q. Zhang, C. Cai, D. Li, and Y. Zhang. (2022). Case Report: Recurrent Hemiplegic Migraine Attacks Accompanied by Intractable Hypomagnesemia Due to a Gene Variant. Front Pediatr 10: 880242.

Li, L., C. Chen, C. Chiang, T. Xiao, Y. Chen, Y. Zhao, and D. Zheng. (2021). The Impact of TRPV1 on Cancer Pathogenesis and Therapy: A Systematic Review. Int J Biol Sci 17: 2034-2049.

Li, L., L. Ma, Z. Luo, X. Wei, Y. Zhao, C. Zhou, A. Mou, Z. Lu, M. You, C. He, H. Ma, Q. Zhou, L. Wang, T. Cao, Y. Gu, P. Gao, and Z. Zhu. (2022). Lack of TRPV1 aggravates obesity-associated hypertension through the disturbance of mitochondrial Ca2+ homeostasis in brown adipose tissue. Hypertens Res. [Epub: Ahead of Print]

Li, M., E. Liu, Q. Zhou, S. Li, X. Wang, Y. Liu, L. Wang, D. Sun, J. Ye, Y. Gao, X. Yang, J. Liu, Y. Yang, and J.Z. Wang. (2018). TRPC1 Null Exacerbates Memory Deficit and Apoptosis Induced by Amyloid-β. J Alzheimers Dis 63: 761-772.

Li, M., J. Du, J. Jiang, W. Ratzan, L.T. Su, L.W. Runnels, and L. Yue. (2007). Molecular Determinants of Mg2+ and Ca2+ Permeability and pH Sensitivity in TRPM6 and TRPM7. J. Biol. Chem. 282(35):25817-25830.

Li, W., Y. Ding, C. Smedley, Y. Wang, S. Chaudhari, L. Birnbaumer, and R. Ma. (2017). Increased glomerular filtration rate and impaired contractile function of mesangial cells in TRPC6 knockout mice. Sci Rep 7: 4145.

Liao, B.K., A.N. Deng, S.C. Chen, M.Y. Chou, and P.P. Hwang. (2007). Expression and water calcium dependence of calcium transporter isoforms in zebrafish gill mitochondrion-rich cells. BMC Genomics. 8: 354.

Liao, M., E. Cao, D. Julius, and Y. Cheng. (2013). Structure of the TRPV1 ion channel determined by electron cryo-microscopy. Nature 504: 107-112.

Lichtenegger, M., T. Stockner, M. Poteser, H. Schleifer, D. Platzer, C. Romanin, and K. Groschner. (2013). A novel homology model of TRPC3 reveals allosteric coupling between gate and selectivity filter. Cell Calcium 54: 175-185.

Liedtke, W., Y. Choe, M.A. Martí-Renom, A.M. Bell, C.S. Denis, A. Sali, A.J. Hudspeth, J.M. Friedman and S. Heller (2000). Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 103: 525-535.

Lindström, J.B., N.T. Pierce, and M.I. Latz. (2017). Role of TRP Channels in Dinoflagellate Mechanotransduction. Biol Bull 233: 151-167.

Liu X., K.T. Cheng, B.C. Bandyopadhyay, B. Pani, A. Dietrich, B.C. Paria, W.D. Swaim, D. Beech, E. Yildrim, B.B. Singh, L. Birnbaumer, I.S. Ambudkar. (2007a). Attenuation of store-operated Ca2+ current impairs salivary gland fluid secretion in TRPC1(-/-) mice. Proc Natl Acad Sci U S A. 104: 17542-17547.

Liu, B. and F. Qin. (2021). Identification of a helix-turn-helix motif for high temperature dependence of vanilloid receptor TRPV2. J. Physiol. [Epub: Ahead of Print]

Liu, C., R. Miao, F. Raza, H. Qian, and X. Tian. (2023). Research progress and challenges of TRPV1 channel modulators as a prospective therapy for diabetic neuropathic pain. Eur J Med Chem 245: 114893.

Liu, L., Y. Li, R. Wang, C. Yin, Q. Dong, H. Hing, C. Kim, and M.J. Welsh. (2007). Drosophila hygrosensation requires the TRP channels water witch and nanchung. Nature 450: 294-298.

Liu, Q., S. Li, Y. Qiu, J. Zhang, F.J. Rios, Z. Zou, and R.M. Touyz. (2023). Cardiovascular toxicity of tyrosine kinase inhibitors during cancer treatment: Potential involvement of TRPM7. Front Cardiovasc Med 10: 1002438.

Liu, S., C. Guo, Z. Dang, and X. Liang. (2016). Comparative proteomics reveal the mechanism of Tween80 enhanced phenanthrene biodegradation by Sphingomonas sp. GY2B. Ecotoxicol Environ Saf 137: 256-264. [Epub: Ahead of Print]

Liu, X., B.B. Singh, and I.S. Ambudkar. (2003). TRPC1 is required for functional store-operated Ca2+ channels. Role of acidic amino acid residues in the S5-S6 region. J. Biol. Chem. 278: 11337-11343.

Liu, X., B.C. Bandyopadhyay, B.B. Singh, K. Groschner, and I.S. Ambudkar. (2005). Molecular analysis of a store-operated and 2-acetyl-sn-glycerol-sensitive non-selective cation channel. Heteromeric assembly of TRPC1-TRPC3. J. Biol. Chem. 280: 21600-21606.

Loukin, S., Z. Su, X. Zhou, and C. Kung. (2010). Forward genetic analysis reveals multiple gating mechanisms of TRPV4. J. Biol. Chem. 285: 19884-19890.

Luo, J. and H. Hu. (2014). Thermally activated TRPV3 channels. Curr Top Membr 74: 325-364.

Luo, Y., S. Chen, F. Wu, C. Jiang, and M. Fang. (2022). The identification of the key residues E829 and R845 involved in transient receptor potential melastatin 2 channel gating. Front Aging Neurosci 14: 1033434.

Ma, H.T., Z. Peng, T. Hiragun, S. Iwaki, A.M. Gilfillan, and M.A. Beaven. (2008). Canonical transient receptor potential 5 channel in conjunction with Orai1 and STIM1 allows Sr2+ entry, optimal influx of Ca2+, and degranulation in a rat mast cell line. J. Immunol. 180: 2233-2239.

Ma, Y., R. Sugiura, A. Koike, H. Ebina, S.O. Sio, and T. Kuno. (2011). Transient receptor potential (TRP) and Cch1-Yam8 channels play key roles in the regulation of cytoplasmic Ca2+ in fission yeast. PLoS One 6: e22421.

Mack, K. and M.J.M. Fischer. (2017). Disrupting sensitization of TRPV4. Neuroscience 352: 1-8. [Epub: Ahead of Print]

Macpherson, L.J., A.E. Dubin, M.J. Evans, F. Marr, P.G. Schultz, B.F. Cravatt, and A. Patapoutian. (2007). Noxious compounds activate TRPA1 ion channels through covalent modification of cysteines. Nature 445: 541-545.

Madej, M.G. and C.M. Ziegler. (2018). Dawning of a new era in TRP channel structural biology by cryo-electron microscopy. Pflugers Arch 470: 213-225.

Mahmuda, N.A., S. Yokoyama, T. Munesue, K. Hayashi, K. Yagi, C. Tsuji, and H. Higashida. (2020). One Single Nucleotide Polymorphism of the Channel Gene Identified as a Risk Factor in Bipolar Disorder Associates with Autism Spectrum Disorder in a Japanese Population. Diseases 8:.

Mammadova-Bach, E., M. Nagy, J.W.M. Heemskerk, B. Nieswandt, and A. Braun. (2019). Store-operated calcium entry in thrombosis and thrombo-inflammation. Cell Calcium 77: 39-48.

Mao, F., L. Guo, M. Jin, X.M. Qiao, G.Y. Ye, and J. Huang. (2018). Molecular cloning and characterization of TRPVs in two rice pests: Nilaparvata lugens (Stål) and Nephotettix cincticeps (Uhler). Pest Manag Sci. [Epub: Ahead of Print]

Maruyama, Y., T. Ogura, K. Mio, S. Kiyonaka, K. Kato, Y. Mori, and C. Sato. (2007). Three-dimensional Reconstruction Using Transmission Electron Microscopy Reveals a Swollen, Bell-shaped Structure of Transient Receptor Potential Melastatin Type 2 Cation Channel. J. Biol. Chem. 282: 36961-36970.

Matta, J.A. and G.P. Ahern. (2007). Voltage is a partial activator of rat thermosensitive TRP channels. J. Physiol. 585(Pt 2):469-482.

Matta, J.A., P.M. Cornett, R.L. Miyares, K. Abe, N. Sahibzada, and G.P. Ahern. (2008). General anesthetics activate a nociceptive ion channel to enhance pain and inflammation. Proc. Natl. Acad. Sci. USA 105: 8784-8789.

McCleskey E.W. and M.S. Gold. (1999). Ion channels of nociception. Annu. Rev. Physiol. 61: 835-856.

McGoldrick, L.L., A.K. Singh, K. Saotome, M.V. Yelshanskaya, E.C. Twomey, R.A. Grassucci, and A.I. Sobolevsky. (2017). Opening of the human epithelial calcium channel TRPV6. Nature. [Epub: Ahead of Print]

Mederos y Schnitzler, M., J. Wäring, T. Gudermann, and V. Chubanov. (2008). Evolutionary determinants of divergent calcium selectivity of TRPM channels. FASEB J. 22(5): 1540-1551.

Memon, T., O. Yarishkin, C.A. Reilly, D. Krizaj, B.M. Olivera, and R.W. Teichert. (2019). trans-Anethole of Fennel oil is a selective and non-electrophilic agonist of the TRPA1 ion channel. Mol Pharmacol. [Epub: Ahead of Print]

Méndez-Reséndiz, K.A., &.#.2.1.1.;. Enciso-Pablo, R. González-Ramírez, R. Juárez-Contreras, T. Rosenbaum, and S.L. Morales-Lázaro. (2020). Steroids and TRP Channels: A Close Relationship. Int J Mol Sci 21:.

Mercado, J., A. Gordon-Shaag, W.N. Zagotta, and S.E. Gordon. (2010). Ca2+-dependent desensitization of TRPV2 channels is mediated by hydrolysis of phosphatidylinositol 4,5-bisphosphate. J. Neurosci. 30: 13338-13347.

Minke, B. and B. Cook. (2002). TRP channel proteins and signal transduction. Physiol. Rev. 82: 429-472.

Mio, K., T. Ogura, and C. Sato. (2008). Structure of six-transmembrane cation channels revealed by single-particle analysis from electron microscopic images. J Synchrotron Radiat 15: 211-214.

Mio, K., T. Ogura, S. Kiyonaka, Y. Hiroaki, Y. Tanimura, Y. Fujiyoshi, Y. Mori, and C. Sato. (2007). The TRPC3 channel has a large internal chamber surrounded by signal sensing antennas. J. Mol. Biol. 367: 373-383.

Moiseenkova-Bell, V.Y., L.A. Stanciu, I.I. Serysheva, B.J. Tobe, and T.G. Wensel. (2008). Structure of TRPV1 channel revealed by electron cryomicroscopy. Proc. Natl. Acad. Sci. USA 105: 7451-7455.

Montell, C. (2005). The TRP superfamily of cation channels. Science STKE 272: 1-24.

Montell, C. and G.M. Rubin. (1989). Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction. Neuron 2: 1313-1323.

Montell, C., L. Birnbaumer, and V. Flockerzi. (2002). The TRP channels, a remarkably functional family. Cell 108: 595-598.

Moparthi, L., S. Kjellström, P. Kjellbom, M.R. Filipovic, P.M. Zygmunt, and U. Johanson. (2020). Electrophile-Induced Conformational Switch of the Human TRPA1 Ion Channel Detected by Mass Spectrometry. Int J Mol Sci 21:.

Moparthi, L., V. Sinica, V.K. Moparthi, M. Kreir, T. Vignane, M.R. Filipovic, V. Vlachova, and P.M. Zygmunt. (2022). The human TRPA1 intrinsic cold and heat sensitivity involves separate channel structures beyond the N-ARD domain. Nat Commun 13: 6113.

Morris, Z., D. Sinha, A. Poddar, B. Morris, and Q. Chen. (2019). Fission yeast TRP channel Pkd2p localizes to the cleavage furrow and regulates cell separation during cytokinesis. Mol. Biol. Cell 30: 1791-1804.

Motter, A.L. and G.P. Ahern. (2012). TRPA1 Is a Polyunsaturated Fatty Acid Sensor in Mammals. PLoS One 7: e38439.

Moussaieff, A., N. Rimmerman, T. Bregman, A. Straiker, C.C. Felder, S. Shoham, Y. Kashman, S.M. Huang, H. Lee, E. Shohami, K. Mackie, M.J. Caterina, J.M. Walker, E. Fride, and R. Mechoulam. (2008). Incensole acetate, an incense component, elicits psychoactivity by activating TRPV3 channels in the brain. FASEB J. 22: 3024-3034.

Mukerji, N., T.V. Damodaran, and M.P. Winn. (2007). TRPC6 and FSGS: the latest TRP channelopathy. Biochim. Biophys. Acta. 1772: 859-868.

Mulukala, S.K.N., S.S. Irukuvajjula, K. Kumar, K. Garai, P. Venkatesu, R. Vadrevu, and A.K. Pasupulati. (2020). Structural features and oligomeric nature of human podocin domain. Biochem Biophys Rep 23: 100774.

Murillo-Rodriguez, E., J.C. Pastrana-Trejo, M. Salas-Crisóstomo, and M. de-la-Cruz. (2017). The endocannabinoid system modulating levels of consciousness, emotions and likely dream contents. CNS Neurol Disord Drug Targets. [Epub: Ahead of Print]

Nadezhdin, K.D., I.A. Talyzina, A. Parthasarathy, A. Neuberger, D.X. Zhang, and A.I. Sobolevsky. (2023). Structure of human TRPV4 in complex with GTPase RhoA. Nat Commun 14: 3733.

Nadezhdin, K.D., L. Correia, C. Narangoda, D.S. Patel, A. Neuberger, T. Gudermann, M.G. Kurnikova, V. Chubanov, and A.I. Sobolevsky. (2023). Structural mechanisms of TRPM7 activation and inhibition. Nat Commun 14: 2639.

Nadler, M.J.S., M.C. Hermosura, K. Inabe, A.-L. Perraud, Q. Zhu, A.J. Stokes, T. Kurosaki, J.-P. Kinet, R. Penner, A.M. Scharenberg, and A. Fleig. (2001). LTRPC7 is a Mg·ATP-regulated divalent cation channel required for cell viability. Nature 411: 590-594.

Neuberger, A., K.D. Nadezhdin, E. Zakharian, and A.I. Sobolevsky. (2021). Structural mechanism of TRPV3 channel inhibition by the plant-derived coumarin osthole. EMBO Rep e53233. [Epub: Ahead of Print]

Neuberger, A., Y.A. Trofimov, M.V. Yelshanskaya, J. Khau, K.D. Nadezhdin, L.S. Khosrof, N.A. Krylov, R.G. Efremov, and A.I. Sobolevsky. (2023). Molecular pathway and structural mechanism of human oncochannel TRPV6 inhibition by the phytocannabinoid tetrahydrocannabivarin. Nat Commun 14: 4630.

Neuberger, A., Y.A. Trofimov, M.V. Yelshanskaya, K.D. Nadezhdin, N.A. Krylov, R.G. Efremov, and A.I. Sobolevsky. (2023). Structural mechanism of human oncochannel TRPV6 inhibition by the natural phytoestrogen genistein. Nat Commun 14: 2659.

Nilius, B., R. Vennekens, J. Prenen, J.G. Hoenderop, G. Droogmans, and R.J. Bindels. (2001). The single pore residue Asp542 determines Ca2+ permeation and Mg2+ block of the epithelial Ca2+ channel. J. Biol. Chem. 276: 1020-1025.

Nonaka, K., X. Han, H. Kato, H. Sato, H. Yamaza, Y. Hirofuji, and K. Masuda. (2019). Novel gain-of-function mutation of associated with accelerated chondrogenic differentiation of dental pulp stem cells derived from a patient with metatropic dysplasia. Biochem Biophys Rep 19: 100648.

Noroozbabaee, L., P.J. Blanco, S. Safaei, and D.P. Nickerson. (2022). A modular and reusable model of epithelial transport in the proximal convoluted tubule. PLoS One 17: e0275837.

Numata, T. and Y. Okada. (2008). Proton Conductivity through the Human TRPM7 Channel and Its Molecular Determinants. J. Biol. Chem. 283: 15097-15103.

Ogawa, Y., L. Zhou, S. Kaneko, and Y. Kusakabe. (2022). Agonistic/antagonistic properties of lactones in food flavors on the sensory ion channels TRPV1 and TRPA1. Chem Senses 47:.

Ohara, K., T. Fukuda, H. Okada, S. Kitao, Y. Ishida, K. Kato, C. Takahashi, M. Katayama, K. Uchida, and M. Tominaga. (2015). Identification of Significant Amino Acids in Multiple Transmembrane Domains of Human Transient Receptor Potential Ankyrin 1 (TRPA1) for Activation by Eudesmol, an Oxygenized Sesquiterpene in Hop Essential Oil. J. Biol. Chem. 290: 3161-3171.

Ojha, N.K., F. Zufall, and T. Leinders-Zufall. (2023). Optical Activation of Photoswitchable TRPC Ligands in the Mammalian Olfactory System Using Laser Scanning Confocal Microscopy. Methods Mol Biol 2710: 31-47.

Okumura, R., K. Shima, T. Muramatsu, K. Nakagawa, M. Shimono, T. Suzuki, H. Magloire, and Y. Shibukawa. (2005). The odontoblast as a sensory receptor cell? The expression of TRPV1 (VR-1) channels. Arch Histol Cytol 68: 251-257.

Olah, Z., L. Karai, and M.J. Iadarola. (2001). Anandamide activates vanilloid receptor 1 (VR1) at acidic pH in dorsal root ganglia neurons and cells ectopically expressing VR1. J. Biol. Chem. 276: 31163-31170.

Pabon, J., M.K. Law, and A. August. (2017). Drebrin Regulation of Calcium Signaling in Immune Cells. Adv Exp Med Biol 1006: 281-290.

Park, J.Y., E.M. Hwang, O. Yarishkin, J.H. Seo, E. Kim, J. Yoo, G.S. Yi, D.G. Kim, N. Park, C.M. Ha, J.H. La, D. Kang, J. Han, U. Oh, and S.G. Hong. (2008). TRPM4b channel suppresses store-operated Ca2+ entry by a novel protein-protein interaction with the TRPC3 channel. Biochem. Biophys. Res. Commun. 368: 677-683.

Parrasia, S., A. Mattarei, A. Furlan, M. Zoratti, and L. Biasutto. (2019). Small-Molecule Modulators of Mitochondrial Channels as Chemotherapeutic Agents. Cell Physiol Biochem 53: 11-43.

Peier, A.M., A. Moqrich, A.C. Hergarden, A.J. Reeve, D.A. Andersson, G.M. Story, T.J. Earley, I Dragoni, P. McIntyre, S. Bevan, and A. Patapoutian. (2002). A TRP channel that senses cold stimuli and menthol. Cell 108: 705-715.

Peng, C., Z. Yang, Z. Liu, S. Wang, H. Yu, C. Cui, Y. Hu, Q. Xing, J. Hu, X. Huang, and Z. Bao. (2021). A Systematical Survey on the TRP Channels Provides New Insight into Its Functional Diversity in Zhikong Scallop (). Int J Mol Sci 22:.

Peng, J.B., X.Z. Chen, U.V. Berger, P.M. Vassilev, H. Tsukaguchi, E.M. Brown, and M.A. Hediger. (1999). Molecular cloning and characterization of a channel-like transporter mediating intestinal calcium absorption. J. Biol. Chem. 274: 22739-22746.

Pereira, G.C., E. Piton, J. Bornholdt, B.M. Dos Santos, A.S. de Almeida, D.P. Dalenogare, M.F.P. Fialho, G. Becker, E. da Silva Brum, T.B. Sampaio, S.M. Oliveira, M.S. Oliveira, G. Trevisan, and G.V. Bochi. (2023). TRPA1 participation in behavioral impairment induced by chronic corticosterone administration. Psychopharmacology (Berl) 240: 157-169.

Perraud, A.-L., A. Fleig, C.A. Dunn, L.A. Bagley, P. Launay, C. Schmitz, A.J. Stokes, Q. Zhu, M.J. Bessman, R. Penner, J.-P. Kinet, and A.M. Scharenberg. (2001). ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature 411: 594-599.

Pertusa, M., B. Rivera, A. González, G. Ugarte, and R. Madrid. (2018). Critical role of the pore domain in the cold response of TRPM8 channels identified by ortholog functional comparison. J. Biol. Chem. 293: 12454-12471.

Pertusa, M., J. Solorza, and R. Madrid. (2023). Molecular determinants of TRPM8 function: key clues for a cool modulation. Front Pharmacol 14: 1213337.

Peters, F., J. Kopp, J. Fischer, and I. Tantcheva-Poór. (2020). Mutation in TRPV3 causes painful focal plantar keratoderma. J Eur Acad Dermatol Venereol. [Epub: Ahead of Print]

Petrova, R.S., N. Nair, N. Bavana, Y. Chen, K.L. Schey, and P.J. Donaldson. (2023). Modulation of Membrane Trafficking of AQP5 in the Lens in Response to Changes in Zonular Tension Is Mediated by the Mechanosensitive Channel TRPV1. Int J Mol Sci 24:.

Phelps, C.B., R.J. Huang, P.V. Lishko, R.R. Wang, and R. Gaudet (2008). Structural analyses of the ankyrin repeat domain of TRPV6 and related TRPV ion channels. Biochemistry 47: 2476-2484.

Prawitt, D., M.K. Monteilh-Zoller, L. Brixel, C. Spangenberg, B. Zabel, A. Fleig, and R. Penner. (2003). TRPM5 is a transient Ca2+-activated cation channel responding to rapid changes in [Ca2+]i. Proc. Natl. Acad. Sci. USA 100: 15166-15171.

Prawitt, D., T. Enklaar, G. Klemm, B. Gärtner, C. Spangenberg, A. Winterpacht, M. Higgins, J. Pelletier, and B. Zabel. (2000). Identification and characterization of MTR1, a novel gene with homology to melastatin (MLSN1) and the trp gene family located in the BWS-WT2 critical region on chromosome 11p15.5 and showing allele-specific expression. Hum Mol Genet 9: 203-216.

Premkumar, L.S. (2001). Interaction between vanilloid receptors and purinergic metabotropic receptors: pain perception and beyond. Proc. Natl. Acad. Sci. USA 98: 6537-6539.

Putney, J.W., Jr. and R.R. McKay. (1999). Capacitative calcium entry channels. BioEssays 21: 38-46.

Qiu, A. and C. Hogstrand. (2004). Functional characterisation and genomic analysis of an epithelial calcium channel (ECaC) from pufferfish, Fugu rubripes. Gene 342: 113-123.

Ramsey, I.S., M. Delling, and D.E. Clapham. ((2006)). An introduction to TRP channels. Annu. Rev. Physiol. 68: 619–647.

Riehle, M., D. Tsvetkov, B.O. Gohlke, R. Preissner, C. Harteneck, M. Gollasch, and B. Nürnberg. (2018). Molecular basis for the sensitivity of TRP channels to polyunsaturated fatty acids. Naunyn Schmiedebergs Arch Pharmacol. [Epub: Ahead of Print]

Riera, C.E., M.O. Huising, P. Follett, M. Leblanc, J. Halloran, R. Van Andel, C.D. de Magalhaes Filho, C. Merkwirth, and A. Dillin. (2014). TRPV1 Pain Receptors Regulate Longevity and Metabolism by Neuropeptide Signaling. Cell 157: 1023-1036.

Rish, A.D., Z. Shen, and T.M. Fu. (2022). It takes two to Tango: Two gates orchestrate the opening of human TRPM2. Cell Calcium 101: 102523.

Rixecker, T., I. Mathar, R. Medert, S. Mannebach, A. Pfeifer, P. Lipp, V. Tsvilovskyy, and M. Freichel. (2016). TRPM4-mediated control of FcεRI-evoked Ca2+ elevation comprises enhanced plasmalemmal trafficking of TRPM4 channels in connective tissue type mast cells. Sci Rep 6: 32981.

Rock, M.J., J. Prenen, V.A. Funari, T.L. Funari, B. Merriman, S.F. Nelson, R.S. Lachman, W.R. Wilcox, S. Reyno, R. Quadrelli, A. Vaglio, G. Owsianik, A. Janssens, T. Voets, S. Ikegawa, T. Nagai, D.L. Rimoin, B. Nilius, and D.H. Cohn. (2008). Gain-of-function mutations in TRPV4 cause autosomal dominant brachyolmia. Nat. Genet. 40: 999-1003.

Roessingh, S., W. Wolfgang, and R. Stanewsky. (2015). Loss of Drosophila melanogaster TRPA1 Function Affects "Siesta" Behavior but Not Synchronization to Temperature Cycles. J Biol Rhythms 30: 492-505.

Ruan, Z., E. Haley, I.J. Orozco, M. Sabat, R. Myers, R. Roth, J. Du, and W. Lü. (2021). Structures of the TRPM5 channel elucidate mechanisms of activation and inhibition. Nat Struct Mol Biol 28: 604-613.

Runnels, L.W., L. Yue, and D.E. Clapham. (2001). TRP-PLIK, a bifunctional protein with kinase and ion channel activities. Science 291: 1043-1046.

Saldías, M.P., P. Cruz, I. Silva, O. Orellana-Serradell, B. Lavanderos, D. Maureira, R. Pinto, and O. Cerda. (2023). The Cytoplasmic Region of SARAF Reduces Triple-Negative Breast Cancer Metastasis through the Regulation of Store-Operated Calcium Entry. Int J Mol Sci 24:.

Saotome, K., A.K. Singh, M.V. Yelshanskaya, and A.I. Sobolevsky. (2016). Crystal structure of the epithelial calcium channel TRPV6. Nature. [Epub: Ahead of Print]

Saqib, U., S. Munjuluri, S. Sarkar, S. Biswas, O. Mukherjee, H. Satsangi, M.S. Baig, A.G. Obukhov, and K. Hajela. (2023). Transient Receptor Potential Canonical 6 (TRPC6) Channel in the Pathogenesis of Diseases: A Jack of Many Trades. Inflammation 1-17. [Epub: Ahead of Print]

Schäffers, O.J.M., J.G.J. Hoenderop, R.J.M. Bindels, and J.H.F. de Baaij. (2018). The rise and fall of novel renal magnesium transporters. Am. J. Physiol. Renal Physiol 314: F1027-F1033.

Schindl, R. and C. Romanin. (2007). Assembly domains in TRP channels. Biochem Soc Trans 35: 84-85.

Schmitz, C., F. Deason, and A.L. Perraud. (2007). Molecular components of vertebrate Mg2+-homeostasis regulation. Magnes. Res. 20: 6-18.

Schoeber, J.P., C.N. Topala, X. Wang, R.J. Diepens, T.T. Lambers, J.G. Hoenderop, and R.J. Bindels. (2006). RGS2 inhibits the epithelial Ca2+ channel TRPV6. J. Biol. Chem. 281: 29669-29674.

Seebohm, G. and J.A. Schreiber. (2021). Beyond Hot and Spicy: TRPV Channels and their Pharmacological Modulation. Cell Physiol Biochem 55: 108-130.

Shannon, A.H., C.T. Elder, G. Lu, G. Su, A. Mast, M.D. Salmon, W.G. Montgomery, M.D. Spinosa, G.R. Upchurch, Jr, and A.K. Sharma. (2020). Pharmacologic inhibition of transient receptor channel vanilloid 4 attenuates abdominal aortic aneurysm formation. FASEB J. [Epub: Ahead of Print]

Sidi, S., R.W. Friedrich, and T. Nicolson. (2003). NompC TRP channel required for vertebrate sensory hair cell mechanotransduction. Science 301: 96-99.

Sierra-Valdez, F., C.M. Azumaya, L.O. Romero, T. Nakagawa, and J.F. Cordero-Morales. (2018). Structure-function analyses of the ion channel TRPC3 reveal that its cytoplasmic domain allosterically modulates channel gating. J. Biol. Chem. [Epub: Ahead of Print]

Simard C., Hof T., Keddache Z., Launay P. and Guinamard R. (2013). The TRPM4 non-selective cation channel contributes to the mammalian atrial action potential. J Mol Cell Cardiol. 59:11-9.

Singaravelu, G., I. Chatterjee, S. Rahimi, M.K. Druzhinina, L. Kang, X.Z. Xu, and A. Singson. (2012). The sperm surface localization of the TRP-3/SPE-41 Ca2+ -permeable channel depends on SPE-38 function in Caenorhabditis elegans. Dev Biol 365: 376-383.

Singh, A.K., K. Saotome, and A.I. Sobolevsky. (2017). Swapping of transmembrane domains in the epithelial calcium channel TRPV6. Sci Rep 7: 10669.

Singh, A.K., K. Saotome, L.L. McGoldrick, and A.I. Sobolevsky. (2018). Structural bases of TRP channel TRPV6 allosteric modulation by 2-APB. Nat Commun 9: 2465.

Sonkusare, S.K., A.D. Bonev, J. Ledoux, W. Liedtke, M.I. Kotlikoff, T.J. Heppner, D.C. Hill-Eubanks, and M.T. Nelson. (2012). Elementary Ca2+ signals through endothelial TRPV4 channels regulate vascular function. Science 336: 597-601.

Souza Bomfim, G.H., V. Costiniti, Y. Li, Y. Idaghdour, and R.S. Lacruz. (2020). TRPM7 activation potentiates SOCE in enamel cells but requires ORAI. Cell Calcium 87: 102187. [Epub: Ahead of Print]

Starkus, J.G., A. Fleig, and R. Penner. (2010). The calcium-permeable non-selective cation channel TRPM2 is modulated by cellular acidification. J. Physiol. 588: 1227-1240.

Startek, J.B., B. Boonen, A. López-Requena, A. Talavera, Y.A. Alpizar, D. Ghosh, N. Van Ranst, B. Nilius, T. Voets, and K. Talavera. (2019). Mouse TRPA1 function and membrane localization are modulated by direct interactions with cholesterol. Elife 8:.

Stokes, A.J., C. Wakano, K.A. Del Carmen, M. Koblan-Huberson, and H. Turner. (2005). Formation of a physiological complex between TRPV2 and RGA protein promotes cell surface expression of TRPV2. J. Cell. Biochem. 94: 669-683.

Story, G.M., A.M. Peier, A.J. Reeve, S.R. Eid, J. Mosbacher, T.R. Hricik, T.J. Earley, A.C. Hergarden, D.A. Andersson, S.W. Hwang, P. McIntyre, T. Jegla, S. Bevan, and A. Patapoutian. (2003). ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 112: 819-829.

Studer, M. and P.A. McNaughton. (2010). Modulation of single-channel properties of TRPV1 by phosphorylation. J. Physiol. 588: 3743-3756.

Stumpf, T., Q. Zhang, D. Hirnet, U. Lewandrowski, A. Sickmann, U. Wissenbach, J. Dörr, C. Lohr, J.W. Deitmer, and C. Fecher-Trost. (2008). The human TRPV6 channel protein is associated with cyclophilin B in human placenta. J. Biol. Chem. 283: 18086-18098.

Suresh K., Servinsky L., Reyes J., Baksh S., Undem C., Caterina M., Pearse DB. and Shimoda LA. (2015). Hydrogen peroxide-induced calcium influx in lung microvascular endothelial cells involves TRPV4. Am J Physiol Lung Cell Mol Physiol. 309(12):L1467-77.

Suzuki, M., J. Sato, K. Kutsuwada, G. Ooki, and M. Imai. (1999). Cloning of a stretch-inhibitable nonselective cation channel. J. Biol. Chem. 274: 6330-6335.

Suzuki, Y., D. Chitayat, H. Sawada, M.A. Deardorff, H.M. McLaughlin, A. Begtrup, K. Millar, J. Harrington, K. Chong, M. Roifman, K. Grand, M. Tominaga, F. Takada, S. Shuster, M. Obara, H. Mutoh, R. Kushima, and G. Nishimura. (2018). TRPV6 Variants Interfere with Maternal-Fetal Calcium Transport through the Placenta and Cause Transient Neonatal Hyperparathyroidism. Am J Hum Genet 102: 1104-1114.

Suzuki, Y., H. Sawada, T. Tokumasu, S. Suzuki, S. Ninomiya, M. Shirai, T. Mukai, C.T. Saito, G. Nishimura, and M. Tominaga. (2020). Novel TRPV6 mutations in the spectrum of transient neonatal hyperparathyroidism. J. Physiol. Sci 70: 33.

Szabó, T., L. Ambrus, N. Zákány, G. Balla, and T. Bíró. (2015). Regulation of TRPC6 ion channels in podocytes - Implications for focal segmental glomerulosclerosis and acquired forms of proteinuric diseases. Acta Physiol Hung 102: 241-251.

Szallasi, A. (2023). The Vanilloid (Capsaicin) Receptor TRPV1 in Blood Pressure Regulation: A Novel Therapeutic Target in Hypertension? Int J Mol Sci 24:.

Taga, A., M.A. Peyton, B. Goretzki, T.Q. Gallagher, A. Ritter, A. Harper, T.O. Crawford, U.A. Hellmich, C.J. Sumner, and B.A. McCray. (2022). TRPV4 mutations causing mixed neuropathy and skeletal phenotypes result in severe gain of function. Ann Clin Transl Neurol. [Epub: Ahead of Print]

Takahashi, K., K. Araki, H. Miyamoto, R. Shirakawa, T. Yoshida, and M. Wakamori. (2021). Capsaicin and Proton Differently Modulate Activation Kinetics of Mouse Transient Receptor Potential Vanilloid-1 Channel Induced by Depolarization. Front Pharmacol 12: 672157.

Tang, Q., W. Guo, L. Zheng, J.X. Wu, M. Liu, X. Zhou, X. Zhang, and L. Chen. (2018). Structure of the receptor-activated human TRPC6 and TRPC3 ion channels. Cell Res. [Epub: Ahead of Print]

Tedeschi, V., M.J. Sisalli, A. Pannaccione, I. Piccialli, P. Molinaro, L. Annunziato, and A. Secondo. (2022). Na/Ca exchanger isoform 1 (NCX1) and canonical transient receptor potential channel 6 (TRPC6) are recruited by STIM1 to mediate Store-Operated Calcium Entry in primary cortical neurons. Cell Calcium 101: 102525.

Thébault, S., G. Cao, H. Venselaar, Q. Xi, R.J. Bindels, and J.G. Hoenderop. (2008). Role of the α-kinase domain in transient receptor potential melastatin 6 channel and regulation by intracellular ATP. J. Biol. Chem. 283: 19999-20007.

Thompson, V., M. Moshirfar, T. Clinch, S. Scoper, S.H. Linn, A. McIntosh, Y. Li, M. Eaton, M. Ferriere, and K. Stasi. (2023). Topical Ocular TRPV1 Antagonist SAF312 (Libvatrep) for Postoperative Pain After Photorefractive Keratectomy. Transl Vis Sci Technol 12: 7.

Toft-Bertelsen, T.L., D. Krízaj, and N. MacAulay. (2017). When size matters: transient receptor potential vanilloid 4 channel as a volume-sensor rather than an osmo-sensor. J. Physiol. [Epub: Ahead of Print]

Ton, H.T., T.X. Phan, A.M. Abramyan, L. Shi, and G.P. Ahern. (2017). Identification of a putative binding site critical for general anesthetic activation of TRPA1. Proc. Natl. Acad. Sci. USA 114: 3762-3767.

Topala, C.N., W.T. Groenestege, S. Thébault, D. van den Berg, B. Nilius, J.G. Hoenderop, and R.J. Bindels. (2007). Molecular determinants of permeation through the cation channel TRPM6. Cell Calcium 41: 513-523.

Tóth, B. and L. Csanády. (2012). Pore collapse underlies irreversible inactivation of TRPM2 cation channel currents. Proc. Natl. Acad. Sci. USA 109: 13440-13445.

Tousova, K., K. Susankova, J. Teisinger, L. Vyklicky, and V. Vlachova. (2004). Oxidizing reagent copper-o-phenanthroline is an open channel blocker of the vanilloid receptor TRPV1. Neuropharmacology 47: 273-285.

Trofimov, Y.A., N.A. Krylov, and R.G. Efremov. (2019). Confined Dynamics of Water in Transmembrane Pore of TRPV1 Ion Channel. Int J Mol Sci 20:.

Tseng, H.H., C.T. Vong, Y.W. Kwan, S.M. Lee, and M.P. Hoi. (2016). TRPM2 regulates TXNIP-mediated NLRP3 inflammasome activation via interaction with p47 phox under high glucose in human monocytic cells. Sci Rep 6: 35016.

Tseng, W.C., D.C. Pryde, K.E. Yoger, K.M. Padilla, B.M. Antonio, S. Han, V. Shanmugasundaram, and A.C. Gerlach. (2018). TRPA1 ankyrin repeat six interacts with a small molecule inhibitor chemotype. Proc. Natl. Acad. Sci. USA 115: 12301-12306.

van de Graaf, S.F.J., J.G.J. Hoenderop, D. Gkika, D. Lamers, J. Prenen, U. Rescher, V. Gerke, O. Staub, B. Nilius, and R.J.M. Bindels. (2003). Functional expression of the epithelial Ca2+ channels (TRPV5 and TRPV6) requires association of the S100A10-annexin 2 complex. EMBO J. 22: 1478-1487.

van Krugten, J., N. Danné, and E.J.G. Peterman. (2022). A local interplay between diffusion and intraflagellar transport distributes TRPV-channel OCR-2 along C. elegans chemosensory cilia. Commun Biol 5: 720.

Vanden Abeele, F., A. Zholos, G. Bidaux, Y. Shuba, S. Thebault, B. Beck, M. Flourakis, Y. Panchin, R. Skryma, and N. Prevarskaya. (2006). Ca2+-independent phospholipase A2-dependent gating of TRPM8 by lysophospholipids. J. Biol. Chem. 281: 40174-40182.

Vennekens, R., A. Menigoz, and B. Nilius. (2012). TRPs in the Brain. Rev Physiol Biochem Pharmacol 163: 27-64.

Viswanath, V., G.M. Story, A.M. Peier, M.J. Petrus, V.M. Lee, S.W. Hwang, A. Patapoutian, and T. Jegla. (2003). Ion channels: opposite thermosensor in fruitfly and mouse. Nature 423: 822-823.

Voets, T., B. Nilius, S. Hoefs, A.W.C.M. van der Kemp, G. Droogmans, R.J.M. Bindels, and J.G.J. Hoenderop. (2004). TRPM6 forms the Mg2+ influx channel involved in intestinal and renal Mg2+ absorption. J. Biol. Chem. 279: 19-25.

Vriens, J., H. Watanabe, A. Janssens, G. Droogmans, T. Voets, and B. Nilius. (2004). Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc. Natl. Acad. Sci. USA 101: 396-401.

Walker, V. and G.W. Vuister. (2023). Biochemistry and pathophysiology of the Transient Potential Receptor Vanilloid 6 (TRPV6) calcium channel. Adv Clin Chem 113: 43-100.

Wang, H., P. Yang, Y. Lu, J. Wang, J. Jeon, Q. Wang, J.B. Tian, B. Zang, Y. Yu, and M.X. Zhu. (2021). Mechanisms of proton inhibition and sensitization of the cation channel TRPV3. J Gen Physiol 153:.

Wang, H., Q. Chen, S. Zhang, and L. Lu. (2021). A Transient Receptor Potential-like Calcium Ion Channel in the Filamentous Fungus. J Fungi (Basel) 7:.

Wang, H., Z. Xu, B.H. Lee, S. Vu, L. Hu, M. Lee, D. Bu, X. Cao, S. Hwang, Y. Yang, J. Zheng, and Z. Lin. (2018). Gain-of-function mutations in TRPM4 activation gate cause progressive symmetric erythrokeratoderma. J Invest Dermatol. [Epub: Ahead of Print]

Wang, L., R.P. Holmes, and J.B. Peng. (2017). The L530R variation associated with recurrent kidney stones impairs the structure and function of TRPV5. Biochem. Biophys. Res. Commun. 492: 362-367.

Wang, L., T.M. Fu, Y. Zhou, S. Xia, A. Greka, and H. Wu. (2018). Structures and gating mechanism of human TRPM2. Science 362:.

Wang, L.X., C.D. Niu, S.F. Wu, and C.F. Gao. (2021). Molecular characterizations and expression profiles of transient receptor potential channels in the brown planthopper, Nilaparvata lugens. Pestic Biochem Physiol 173: 104780.

Wang, Y., J. Yang, R. Miao, Y. Kang, and Z. Qi. (2020). A novel zinc transporter essential for Arabidopsis zinc and iron-dependent growth. J Plant Physiol. 256: 153296. [Epub: Ahead of Print]

Wang, Y., L. Tan, K. Jiao, C. Xue, Q. Tang, S. Jiang, Y. Ren, H. Chen, T.M.A. El-Aziz, K.N.M. Abdelazeem, Y. Yu, F. Zhao, M.X. Zhu, and Z. Cao. (2022). Scutellarein attenuates atopic dermatitis by selectively inhibiting transient receptor potential vanilloid 3 channels. Br J Pharmacol. [Epub: Ahead of Print]

Wang, Y.Y., R.B. Chang, and E.R. Liman. (2010). TRPA1 is a component of the nociceptive response to CO2. J. Neurosci. 30: 12958-12963.

Wei, S., J. Behn, C.P. Poore, S.W. Low, B. Nilius, H. Fan, and P. Liao. (2022). Binding epitope for recognition of human TRPM4 channel by monoclonal antibody M4M. Sci Rep 12: 19562.

Weissgerber, P., U. Kriebs, V. Tsvilovskyy, J. Olausson, O. Kretz, C. Stoerger, S. Mannebach, U. Wissenbach, R. Vennekens, R. Middendorff, V. Flockerzi, and M. Freichel. (2012). Excision of Trpv6 gene leads to severe defects in epididymal Ca2+ absorption and male fertility much like single D541A pore mutation. J. Biol. Chem. 287: 17930-17941.

Wheeler, G.L. and C. Brownlee. (2008). Ca2+ signalling in plants and green algae--changing channels. Trends Plant Sci. 13: 506-514.

Wilkinson, J.A., J.L. Scragg, J.P. Boyle, B. Nilius, and C. Peers. (2008). H2O 2-stimulated Ca2+ influx via TRPM2 is not the sole determinant of subsequent cell death. Pflugers Arch 455: 1141-1151.

Winn, M.P., P.J. Conlon, K.L. Lynn, M.K. Farrington, T. Creazzo, A.F. Hawkins, N. Daskalakis, S.Y. Kwan, S. Ebersviller, J.L. Burchette, M.A. Pericak-Vance, D.N. Howell, J.M. Vance, and P.B. Rosenberg. (2005). A mutation in the TRPC6 cation channel causes familial focal segmental glomerulosclerosis. Science 308: 1801-1804.

Woll, K.A., K.A. Skinner, E. Gianti, N.V. Bhanu, B.A. Garcia, V. Carnevale, R.G. Eckenhoff, and R. Gaudet. (2017). Sites Contributing to TRPA1 Activation by the Anesthetic Propofol Identified by Photoaffinity Labeling. Biophys. J. [Epub: Ahead of Print]

Won, J., J. Kim, H. Jeong, J. Kim, S. Feng, B. Jeong, M. Kwak, J. Ko, W. Im, I. So, and H.H. Lee. (2023). Molecular architecture of the Gα-bound TRPC5 ion channel. Nat Commun 14: 2550.

Wong, F., E.L. Schaefer, B.C. Roop, J.N. LaMendola, D. Johnson-Seaton, and D. Shao. (1989). Proper function of the Drosophila trp gene product during pupal development is important for normal visual transduction in the adult. Neuron 3: 81-94.

Woo SK., Kwon MS., Ivanov A., Geng Z., Gerzanich V. and Simard JM. (2013). Complex N-glycosylation stabilizes surface expression of transient receptor potential melastatin 4b protein. J Biol Chem. 288(51):36409-17.

Xia, R., Z.Z. Mei, H.J. Mao, W. Yang, L. Dong, H. Bradley, D.J. Beech, and L.H. Jiang. (2008). Identification of pore residues engaged in determining divalent cationic permeation in transient receptor potential melastatin subtype channel 2. J. Biol. Chem. 283: 27426-27432.

Xiao, B., A.E. Dubin, B. Bursulaya, V. Viswanath, T.J. Jegla, and A. Patapoutian. (2008). Identification of transmembrane domain 5 as a critical molecular determinant of menthol sensitivity in mammalian TRPA1 channels. J. Neurosci. 28: 9640-9651.

Xiao, R. and X.Z. Xu. (2009). Function and regulation of TRP family channels in C. elegans. Pflugers Arch 458: 851-860.

Xiao, R., B. Zhang, Y. Dong, J. Gong, T. Xu, J. Liu, and X.Z. Xu. (2013). A genetic program promotes C. elegans longevity at cold temperatures via a thermosensitive TRP channel. Cell 152: 806-817.

Xie, B. and X.Y. Li. (2018). Inflammatory mediators causing cutaneous chronic itch in some diseases via transient receptor potential channel subfamily V member 1 and subfamily A member 1. J Dermatol. [Epub: Ahead of Print]

Xu, H., I.S. Ramsey, S.A. Kotecha, M.M. Moran, J.A. Chong, D. Lawson, P. Ge, J. Lilly, I. Silos-Santiago, Y. Xie, P.S. DiStefano, R. Curtis, and D.E. Clapham. (2002). TRPV3 is a calcium-permeable temperature-sensitive cation channel. Nature 418: 181-186.

Xu, L., Y. Han, X. Chen, A. Aierken, H. Wen, W. Zheng, H. Wang, X. Lu, Z. Zhao, C. Ma, P. Liang, W. Yang, S. Yang, and F. Yang. (2020). Molecular mechanisms underlying menthol binding and activation of TRPM8 ion channel. Nat Commun 11: 3790.

Xu, X.Z., and P.W. Sternberg. (2003). A C. elegans sperm TRP protein required for sperm-egg interactions during fertilization. Cell 114: 285-297.

Xu, X.Z., F. Chien, A. Butler, L. Salkoff, and C. Montell. (2000). TRPgamma, a drosophila TRP-related subunit, forms a regulated cation channel with TRPL. Neuron. 26: 647-657.

Yan, J., C.P. Bengtson, B. Buchthal, A.M. Hagenston, and H. Bading. (2020). Coupling of NMDA receptors and TRPM4 guides discovery of unconventional neuroprotectants. Science 370:.

Yang, F. and J. Zheng. (2017). Understand spiciness: mechanism of TRPV1 channel activation by capsaicin. Protein Cell. [Epub: Ahead of Print]

Yang, F., Y. Cui, K. Wang, and J. Zheng. (2010). Thermosensitive TRP channel pore turret is part of the temperature activation pathway. Proc. Natl. Acad. Sci. USA 107: 7083-7088.

Yang, P.L., X.H. Li, J. Wang, X.F. Ma, B.Y. Zhou, Y.F. Jiao, W.H. Wang, P. Cao, M.X. Zhu, P.W. Li, Z.H. Xiao, C.Z. Li, C.R. Guo, Y.T. Lei, and Y. Yu. (2021). GSK1702934A and M085 directly activate TRPC6 via a mechanism of stimulating the extracellular cavity formed by the pore helix and transmembrane helix S6. J. Biol. Chem. 101125. [Epub: Ahead of Print]

Yang, T., S. Zhao, Y. Yuan, X. Zhao, F. Bu, Z. Zhang, Q. Li, Y. Li, Z. Wei, X. Sun, Y. Zhang, and J. Xie. (2023). Platycodonis Radix Alleviates LPS-Induced Lung Inflammation through Modulation of TRPA1 Channels. Molecules 28:.

Yao, J., B. Liu, and F. Qin. (2011). Modular thermal sensors in temperature-gated transient receptor potential (TRP) channels. Proc. Natl. Acad. Sci. USA 108: 11109-11114.

Ye, L., S. Kleiner, J. Wu, R. Sah, R.K. Gupta, A.S. Banks, P. Cohen, M.J. Khandekar, P. Boström, R.J. Mepani, D. Laznik, T.M. Kamenecka, X. Song, W. Liedtke, V.K. Mootha, P. Puigserver, P.R. Griffin, D.E. Clapham, and B.M. Spiegelman. (2012). TRPV4 is a regulator of adipose oxidative metabolism, inflammation, and energy homeostasis. Cell 151: 96-110.

Yelshanskaya, M.V., K.D. Nadezhdin, M.G. Kurnikova, and A.I. Sobolevsky. (2020). Structure and function of the calcium-selective TRP channel TRPV6. J. Physiol. [Epub: Ahead of Print]

Yin, Y., F. Zhang, S. Feng, K.J. Butay, M.J. Borgnia, W. Im, and S.Y. Lee. (2022). Activation mechanism of the mouse cold-sensing TRPM8 channel by cooling agonist and PIP. Science 378: eadd1268.

Yin, Y., M. Wu, L. Zubcevic, W.F. Borschel, G.C. Lander, and S.Y. Lee. (2018). Structure of the cold- and menthol-sensing ion channel TRPM8. Science 359: 237-241.

Yin, Y.L., H.H. Wang, Z.C. Gui, S. Mi, S. Guo, Y. Wang, Q.Q. Wang, R.Z. Yue, L.B. Lin, J.X. Fan, X. Zhang, B.Y. Mao, T.H. Liu, G.R. Wan, H.Q. Zhan, M.L. Zhu, L.H. Jiang, and P. Li. (2022). Citronellal Attenuates Oxidative Stress-Induced Mitochondrial Damage through TRPM2/NHE1 Pathway and Effectively Inhibits Endothelial Dysfunction in Type 2 Diabetes Mellitus. Antioxidants (Basel) 11:.

Yoshida, Y., K. Saitoh, Y. Aihara, S. Okada, T. Misaka, and K. Abe. (2007). Transient receptor potential channel M5 and phospholipaseC-beta2 colocalizing in zebrafish taste receptor cells. Neuroreport 18: 1517-1520.

Zakharian, E., C. Cao, and T. Rohacs. (2010). Gating of transient receptor potential melastatin 8 (TRPM8) channels activated by cold and chemical agonists in planar lipid bilayers. J. Neurosci. 30: 12526-12534.

Zayats V., Samad A., Minofar B., Roelofs KE., Stockner T. and Ettrich R. (2013). Regulation of the transient receptor potential channel TRPA1 by its N-terminal ankyrin repeat domain. J Mol Model. 19(11):4689-700.

Zernov, N., A.V. Veselovsky, V.V. Poroikov, D. Melentieva, A. Bolshakova, and E. Popugaeva. (2022). New Positive TRPC6 Modulator Penetrates Blood-Brain Barrier, Eliminates Synaptic Deficiency and Restores Memory Deficit in 5xFAD Mice. Int J Mol Sci 23:.

Zhai, K., A. Liskova, P. Kubatka, and D. Büsselberg. (2020). Calcium Entry through TRPV1: A Potential Target for the Regulation of Proliferation and Apoptosis in Cancerous and Healthy Cells. Int J Mol Sci 21:.

Zhang, F., A. Jara-Oseguera, T.H. Chang, C. Bae, S.M. Hanson, and K.J. Swartz. (2017). Heat activation is intrinsic to the pore domain of TRPV1. Proc. Natl. Acad. Sci. USA. [Epub: Ahead of Print]

Zhang, L., C. Simonsen, L. Zimova, K. Wang, L. Moparthi, R. Gaudet, M. Ekoff, G. Nilsson, U.A. Hellmich, V. Vlachova, P. Gourdon, and P.M. Zygmunt. (2022). Cannabinoid non-cannabidiol site modulation of TRPV2 structure and function. Nat Commun 13: 7483.

Zhang, Y. and Y. Wang. (2017). [TRPV1: an important molecule involved in the peripheral sensitization during chronic pain and central pain modulation]. Sheng Li Xue Bao 69: 677-684.

Zhang, Z., H. Okawa, Y. Wang, and E.R. Liman. (2005). Phosphatidylinositol 4,5-bisphosphate rescues TRPM4 channels from desensitization. J. Biol. Chem. 280: 39185-39192.

Zhao, Y., B.M. McVeigh, and V.Y. Moiseenkova-Bell. (2021). Structural Pharmacology of TRP Channels. J. Mol. Biol. 166914. [Epub: Ahead of Print]

Zheng, Y., Q. Huang, Y. Zhang, L. Geng, W. Wang, H. Zhang, X. He, and Q. Li. (2023). Multimodal roles of transient receptor potential channel activation in inducing pathological tissue scarification. Front Immunol 14: 1237992.

Zhou, X., Z. Su, A. Anishkin, W.J. Haynes, E.M. Friske, S.H. Loukin, C. Kung, and Y. Saimi. (2007). Yeast screens show aromatic residues at the end of the sixth helix anchor transient receptor potential channel gate. Proc. Natl. Acad. Sci. USA. 104: 15555-15559.

Zhou, X.L., S.H. Loukin, R. Coria, C. Kung, and Y. Saimi. (2005). Heterologously expressed fungal transient receptor potential channels retain mechanosensitivity in vitro and osmotic response in vivo. Eur Biophys. J. 34: 413-422.

Zhou, Y., P. Castonguay, E.H. Sidhom, A.R. Clark, M. Dvela-Levitt, S. Kim, J. Sieber, N. Wieder, J.Y. Jung, S. Andreeva, J. Reichardt, F. Dubois, S.C. Hoffmann, J.M. Basgen, M.S. Montesinos, A. Weins, A.C. Johnson, E.S. Lander, M.R. Garrett, C.R. Hopkins, and A. Greka. (2017). A small-molecule inhibitor of TRPC5 ion channels suppresses progressive kidney disease in animal models. Science 358: 1332-1336.

Zhu, C., K. Huang, Y. Wang, K. Alanis, W. Shi, and L.A. Baker. (2021). Imaging with Ion Channels. Anal Chem 93: 5355-5359.

Zimmermann, K., J.K. Lennerz, A. Hein, A.S. Link, J.S. Kaczmarek, M. Delling, S. Uysal, J.D. Pfeifer, A. Riccio, and D.E. Clapham. (2011). Transient receptor potential cation channel, subfamily C, member 5 (TRPC5) is a cold-transducer in the peripheral nervous system. Proc. Natl. Acad. Sci. USA 108: 18114-18119.

Zouharova, M., P. Herman, K. Hofbauerová, J. Vondrasek, and K. Bousova. (2019). TRPM6 N-Terminal CaM- and S100A1-Binding Domains. Int J Mol Sci 20:.

Zsidó, B.Z., R. Börzsei, E. Pintér, and C. Hetényi. (2021). Prerequisite Binding Modes Determine the Dynamics of Action of Covalent Agonists of Ion Channel TRPA1. Pharmaceuticals (Basel) 14:.

Zubcevic, L. and S.Y. Lee. (2019). The role of π-helices in TRP channel gating. Curr. Opin. Struct. Biol. [Epub: Ahead of Print]

Zubcevic, L., M.A. Herzik, Jr, B.C. Chung, Z. Liu, G.C. Lander, and S.Y. Lee. (2016). Cryo-electron microscopy structure of the TRPV2 ion channel. Nat Struct Mol Biol 23: 180-186.

Zubcevic, L., W.F. Borschel, A.L. Hsu, M.J. Borgnia, and S.Y. Lee. (2019). Regulatory switch at the cytoplasmic interface controls TRPV channel gating. Elife 8:.

Examples:

TC#NameOrganismal TypeExample
1.A.4.1.1

Transient receptor potential (TRP) protein.  Assembles in vivo as a homomultimeric channel, not as a heteromeric channel with TrpL as the subunit (Katz et al. 2013).

Animals

TRP protein of Drosophila melanogaster (P19334)

 
1.A.4.1.10

Trp-2 channel; controls nicotine-dependent behavior (Xiao and Xu 2009).  The TRPC orthologues TRP-1 and -2 genetically complement the loss of syndecan by suppressing neuronal guidance and locomotory defects related to increases in neuronal calcium levels. The widespread and conserved syndecan-TRPC axis therefore fine tunes cytoskeletal organization and cell behavior (Gopal et al. 2015).

Animals

Trp-2 of Caenorhabditis elegans

 
1.A.4.1.11

TRP channel homologue, Trp1, of 766 aas and 6 - 9 TMSs.  Contains Ankyrin - PKD1 - TrpC channel domains.  Exhibits properties of mammalian signal transduction Trp channels (Arias-Darraz et al. 2015).  Photoswitchable reagents are used for investigating various types of TRPC channels, including TRPC2, TRPC3, TRPC5, and TRPC6, to gain new insights into the gating mechanisms and functions of these channels (Ojha et al. 2023).

Green algae

TRP channel homologue of Chlamydomonas reinhardtii (Chlamydomonas smithii)

 
1.A.4.1.12

TrpC4 of 977aas.  In epidermal keratinocytes, a syndecan-TRPC4 complex controls adhesion, adherens junction composition, and early differentiation in vivo and in vitro (Gopal et al. 2015).  Constitutively active TRPC1/C4-dependent background Ca2+ entry fine-tunes Ca2+ cycling in beating adult cardiomyocytes. Double TRPC1/C4-gene inactivation protects against development of maladaptive cardiac remodelling without altering cardiac or extracardiac functions contributing to this pathogenesis (Camacho Londoño et al. 2015). A cryo-EM structure of TRPC4 in its unliganded (apo) state has beeen solved to an overall resolution of 3.3 A. It reveals a unique architecture with a long pore loop stabilized by a disulfide bond. Beyond the shared tetrameric six-transmembrane fold, the TRPC4 structure deviates from other TRP channels with a unique N-terminal cytosolic domain which forms extensive aromatic contacts with the TRP and the C-terminal domains (Duan et al. 2018).

TrpC4 of Homo sapiens

 
1.A.4.1.13

Transient receptor potential ion channel protein, TRP6, of 2341 aas and 6 - 9 TMSs.

TRP6 OF Chlamydomonas reinhardtii (Chlamydomonas smithii)

 
1.A.4.1.14

Flagellar associated calcium channel protein of 1,729 aas, FAP148 (Wheeler and Brownlee 2008).

FAP148 of Chlamydomonas reinhardtii

 
1.A.4.1.15

Transient potential protein-gamma, Trpγ, of 1128 aas and 10 TMSs.  A light-sensitive cation/calcium channel that is required for inositide-mediated Ca2+ entry in the retina during phospholipase C (PLC)-mediated phototransduction. It forms a regulated cation channel when heteromultimerized with TrpL (Xu et al. 2000).

TrpL of Drosophila melanogaster (Fruit fly)

 
1.A.4.1.16

Short transient receptor potential channel 7 of 722 aas and 7 or 8 TMSs plus the P-loop. TRPC3.6, TRPC3.7, and TRPV4.7 are important for thermal regulation in oysters (Fu et al. 2021).

TRP channel 7 of Crassostrea gigas (Pacific oyster) (Crassostrea angulata)

 
1.A.4.1.2

TRP7 receptor-activated capacitative Ca2+ entry channel

Animals

TRP7 of Mus musculus (Q9WVC5)

 
1.A.4.1.3

TRPC1 store-operated Ca2+ channel (Liu et al., 2003) (activated by the metabotropic [G- protein-dependent] glutamate receptor, mGluR1) (Kim et al., 2003) (controls salivary gland fluid secretion in mice (Liu et al., 2007a).  Constitutively active TRPC1/C4-dependent background Ca2+ entry fine-tunes Ca2+ cycling in beating adult cardiomyocytes. Double TRPC1/C4-gene inactivation protects against development of maladaptive cardiac remodelling without altering cardiac or extracardiac functions contributing to this pathogenesis (Camacho Londoño et al. 2015). Regulated by drebrin (DBN1; 649 aas; Q16643) (Pabon et al. 2017). TRPC1 null mutations exacerbate memory loss and apoptosis induced by amyloid-beta (Li et al. 2018). Pulsed focused ultrasound (pFUS) acoustic radiation forces mechanically activate a Na+-containing TRPC1 (TC# 1.A.4..1.3) channel generating current upstream of voltage-gated Ca2+ channels (VGCC) rather than directly opening VGCC (Burks et al. 2019).

Animals

TRPC1 of Homo sapiens (P48995)

 
1.A.4.1.4

TRPC3 store-operated non-selective cation channel (activated by thapsigargin and 2 acyl glycerol; forms a heteromeric channel with TrpC1, TC #1.A.4.1.3) (Liu et al., 2005).  A  structural model of the TRPC3 permeation pathway based on a sodium channel (TC# 1.A.1.14.5) with a localized selectivity filter and an occluding gate with evidence for allosteric coupling between the gate and the selectivity filter has been proposed (Ko et al. 2009; Lichtenegger et al. 2013). The channel may have a large internal chamber surrounded by signal sensing antennas (Mio et al. 2007). TRPC channels are involved in store-operated calcium entry and calcium homeostasis, and they are implicated in human diseases such as neurodegenerative disease, cardiac hypertrophy, and spinocerebellar ataxia (Fan et al. 2018). The structure in a lipid-occupied, closed state has been solved at 3.3 Å resolution. TRPC3 has four elbow-like membrane reentrant helices prior to the first transmembrane helix. The TRP helix is perpendicular to, and thus disengaged from, the pore-lining S6, suggesting a different gating mechanism from other TRP subfamily channels. The third transmembrane helix S3 is remarkably long, shaping a unique transmembrane domain, and constituting an extracellular domain that may serve as a sensor of external stimuli. Fan et al. 2018 identified two lipid binding sites, one being sandwiched between the pre-S1 elbow and the S4-S5 linker, and the other being close to the ion-conducting pore, where the conserved LWF motif of the TRPC family is located. The cytoplasmic domain allosterically modulates channel gating (Sierra-Valdez et al. 2018). This channel may be present in mitochondria (Parrasia et al. 2019). TRPC3 and TRPC6 channels are calcium-permeable non-selective cation channels. The gain-of-function (GOF) mutations of TRPC6 lead to familial focal segmental glomerulosclerosis (FSGS) in humans. Guo et al. 2022 reported the cryo-EM structures of human TRPC3 in both high-calcium and low-calcium conditions. They identified both inhibitory and activating calcium-binding sites in TRPC3 that couple intracellular calcium concentrations to the basal channel activity. These calcium sensors are structurally and functionally conserved in TRPC6. The GOF mutations of TRPC6 activate the channel by allosterically abolishing the inhibitory effects of intracellular calcium. Structures of human TRPC6 in complex with two chemically distinct inhibitors bound at different ligand-binding pockets revealed different conformations of the transmembrane domain (Guo et al. 2022). TRPC3 is primarily gated by lipids, and its surface expression is dependent on cholesterol (Clarke et al. 2022). Regulating the activity of the SOCE response via SARAF activity may allow therapeutic strategies for triple-negative breast cancer (Saldías et al. 2023).

Animals

TRPC3 of Homo sapiens (Q13507)

 
1.A.4.1.5

Transient receptor potential canonical-6, TRPC6,  a non-selective cation channel that is directly activated by diacylglycerol (DAG (Szabó et al. 2015). Mutation causes a particularly aggressive form of familial focal segmental glomerulosclerosis (Winn et al., 2005; Mukerji et al., 2007). Tang et al. 2018 presented the structure of the human TRPC6 homotetramer in complex with a high-affinity inhibitor, BTDM, solved by single-particle cryo-EM to 3.8 Å resolution. The structure shows a two-layer architecture in which the bell-shaped cytosolic layer holds the transmembrane layer. Extensive inter-subunit interactions of cytosolic domains, including the N-terminal ankyrin repeats and the C-terminal coiled-coil, contribute to the tetramer assembly. The high-affinity inhibitor BTDM wedges between the S5-S6 pore domain and voltage sensor-like domain to inhibit channel opening (Tang et al. 2018).  TRPC6 may regulate the glomerular filtration rate by modulating mesangial cell contractile function through multiple Ca2+ signaling pathways (Li et al. 2017). Several proteins including podocin (8.A.21.1.2), nephrin (8.A.23.1.33), CD2AP (8.A.34.1.5) and TRPC6 form a macromolecular assembly that constitutes the slit-diaphragm in podocytes that resembles tight junctions (Mulukala et al. 2020). Two small molecules, GSK1702934A and M085, directly activate TRPC6 via a mechanism involving stimulation of the extracellular cavity formed by the pore helix and transmembrane helix S6 (Yang et al. 2021). Na+/Ca2+ exchanger, NCX1, and canonical transient receptor potential channel 6 (TRPC6) are recruited by STIM1 to mediate Store-Operated Calcium Entry in primary cortical neurons (Tedeschi et al. 2022). Guo et al. 2022 reported the cryo-EM structures of human TRPC3 in both high-calcium and low-calcium conditions. They identified both inhibitory and activating calcium-binding sites in TRPC3 that couple intracellular calcium concentrations to the basal channel activity. These calcium sensors are structurally and functionally conserved in TRPC6. The GOF mutations of TRPC6 activate the channel by allosterically abolishing the inhibitory effects of intracellular calcium. Structures of human TRPC6 in complex with two chemically distinct inhibitors bound at different ligand-binding pockets revealed different conformations of the transmembrane domain (Guo et al. 2022). The selective TRPC6 agonist, 3-(3-,4-Dihydro-6,7-dimethoxy-3,3-dimethyl-1-isoquinolinyl)-2H-1-benzopyran-2-one (C20) binds to the extracellular agonist binding site of  TRPC6, protects hippocampal mushroom spines from amyloid toxicity in vitro, efficiently recovers synaptic plasticity in 5xFAD brain slices, penetrates the blood-brain barrier and recovers cognitive deficits in 5xFAD mice. Thus, C20 is the novel TRPC6-selective drug suitable to treat synaptic deficiency in Alzheimer's disease-affected hippocampal neurons (Zernov et al. 2022). Paraoxonase 2 (PON2) deficiency reproduces lipid alterations of diabetic and inflammatory glomerular disease while affecting TRPC6 signaling (Hagmann et al. 2022). Capsazepine (CPZ) inhibits TRPC6 conductance and is protective in adriamycin-induced nephropathy and diabetic glomerulopathy (Hagmann et al. 2023). The mammalian TRPC subfamily comprises seven transmembrane proteins (TRPC1-7) forming cation channels in the plasma membrane of mammalian cells. TRPC channels mediate Ca2+ and Na+ influx into cells. Amongst TRPCs, TRPC6 deficiency or increased activity due to gain-of-function mutations has been associated with multiple diseases, such as kidney, pulmonary, and neurological diseases. Indeed, the TRPC6 protein is expressed in various organs and is involved in diverse signalling pathways. The last decade saw a surge in studies concerning the physiological roles of TRPC6 and describing the development of new pharmacological tools modulating TRPC6 activity (Saqib et al. 2023). One defective TRPC6 gene copy is not sufficient to cause focal segmental glomerulosclerosis (FSGS), which is inherited as an autosomal dominant disease. Increased rather than reduced calcium influx through TRPC6 is required for podocyte cell death (Batool et al. 2023).

Animals

TRPC6 of Homo sapiens (Q9Y210)

 
1.A.4.1.6

Sperm TRP-3 (SPE-41) Ca2+-permeable channel. Translocated from vesicles to the plasma membrane upon sperm activation in a process dependent on the 4TMS SPE-38 protein (8.A.36.1.1) (Singaravelu et al., 2012) during sperm-egg interactions leading to fertilization (Xu et al., 2003).

Animals

TRP-3 of Caenorhabditis elegans (AAQ22724)

 
1.A.4.1.7

Short transient receptor channel 5 (TrpC5 or Htrp5) (transports Ca2+ and Sr2+ in the presence of Orai1 and STIM1 (TC# 1.A.52.1.1) (Ma et al., 2008). It is a cold-transducer in the peripheral nervous system (Zimmermann et al., 2011). A small-molecule inhibitor suppresses progressive kidney disease in rats (Zhou et al. 2017).  ORAI and TRP, and the transmembrane Ca2+ sensors, stromal interaction molecules (STIMs), are involved in thrombosis and thrombo-inflammation in platelets and immune cells. Disregulated store-operated Ca2+ (SOCE) fluxes in platelets and immune cells are responsible, and the potential of SOCE inhibition as a therapeutic option to prevent or treat arterial thrombosis as well as thrombo-inflammatory disease states such as ischemic stroke have been considered (Mammadova-Bach et al. 2019). The molecular architecture of the Galpha(i)-bound TRPC5 ion channel has been solved (Won et al. 2023). G-protein coupled receptors (GPCRs) and ion channels serve as key molecular switches through which extracellular stimuli are transformed into intracellular effects, and it has long been postulated that ion channels are direct effector molecules of the alpha subunit of G-proteins (Galpha; see TC family 8.A.43). Won et al. 2023 presented cryo-EM structures of the human TRPC5-Galpha(i3) complexes with a 4:4 stoichiometry in lipid nanodiscs. Galpha(i3) binds to the ankyrin repeat edge of TRPC5 ~ 50 Å away from the cell membrane. Electrophysiological analyses showed that Galpha(i3) increases the sensitivity of TRPC5 to phosphatidylinositol 4,5-bisphosphate (PIP(2)), thereby rendering TRPC5 more easily opened in the cell membrane, where the concentration of PIP(2) is physiologically regulated. These observations show that ion channels are one of the direct effector molecules of Galpha proteins triggered by GPCR activation-providing a structural framework for unraveling the crosstalk between two major classes of transmembrane proteins: GPCRs and ion channels (Won et al. 2023).

Animals

TrpC5 of Homo sapiens (Q9UL62)

 
1.A.4.1.8

TrpL (Trp-like), isoform A (1124 aas). A light-sensitive calcium channel that is required for inositide-mediated Ca2+ entry in the retina during phospholipase C (PLC)-mediated phototransduction (Lan et al. 1998; Chyb et al. 1999). It is required for vision in the dark and in dim light. and binds calmodulin. Trp and TrpL act together in the light response (Bähner et al. 2002). TrpL assembles in vivo as a homo-multimeric channe, not as a hetero-meric channels as reported previously (Katz et al. 2013).

Animals

TrpL of Drosophila melanogaster (P48994)

 
1.A.4.1.9

Trp-1 isoform channel; controls nicotne-dependent behavior (Xiao and Xu 2009). TRPC orthologues TRP-1 and -2 genetically complement the loss of syndecan by suppressing neuronal guidance and locomotory defects related to increases in neuronal calcium levels. The widespread and conserved syndecan-TRPC axis therefore fine tunes cytoskeletal organization and cell behavior (Gopal et al. 2015).

Animals

Trp-1 of Caenorhabditis elegans

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.10.1

TRP cation-slective channel homologue of 1177 aas

Green algae

TRP channel homologue of Chlamydomonas reinhardtii (Chlamydomonas smithii)

 
1.A.4.10.2

TRP channel homologue of 962 aas

Alveolata

TRP channel homologue of Oxytricha trifallax

 
1.A.4.10.3

TRP channel homologue of 1486 aas

Green algae

TRP channel homologue of Volvox carteri

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.2.1

Vanilloid receptor subtype 1 (VR1 or TRPV1) (noxious, heat-sensitive [opens with increasing temperatures; e.g., >42°C]; also sensitive to acidic pH and voltage and inflamation; serves as the receptor for the alkaloid irritant, capsaicin, for resiniferatoxin and for endo-cannabinoids (Murillo-Rodriguez et al. 2017). Resiniferatoxin binds to the capsaicin receptor (TRPV1) near the extracellular side of the S4 transmembrane domain (Chou et al. 2004). It is regulated by bradykinin and prostaglandin E2) (contains a C-terminal region, adjacent to the channel gate, that determines the coupling of stimulus sensing and channel opening) (Garcia-Sanz et al., 2007; Matta and Ahern, 2007). It is activated and sensitized by local anesthetics in sensory neurons (Leffler et al., 2008). A bivalent tarantula toxin activates the capsaicin receptor (TRPV1) by targeting the outer pore domain (Bohlen et al., 2010). Single-channel properties of TRPV1 are modulated by phosphorylation (Studer and McNaughton, 2010). TRPV1 mediates an itch associated response (Kim et al., 2011). The thermosensitive TRP channel pore turret is part of the temperature activation apparatus (Yang et al., 2010). Modular thermal sensors in temperature-gated transient receptor potential (TRP) channels have been identified (Yao et al., 2011). TRPV1 opening is associated with major structural rearrangements in the outer pore, including the pore helix and selectivity filter, as well as pronounced dilation of a hydrophobic constriction at the lower gate, suggesting a dual gating mechanism (Cao et al. 2013). Allosteric coupling between upper and lower gates may account for modulation exhibited by TRPV1 and other TRP channels (Liao et al. 2013).  TRPV1 regulates longevity and metabolism by neuropeptides in mice (Riera et al. 2014). The pore of TRPV1 contains the structural elements sufficient for activation by noxious heat (Zhang et al. 2017). In bull sperm, TRPV1 functions in the regulation of motility and the  acrosome reaction (Kumar et al. 2019). The dynamics of water in the transmembrane pore of TRPV1 have been studied (Trofimov et al. 2019). TRPV1 - 6 channel subunits do not combine arbitrarily. With the exception of TRPV5 and TRPV6, TRPV channel subunits preferentially assemble into homomeric complexes (Hellwig et al. 2005). TrpV1-gated ion channels have been used as sensors for imaging applications (Zhu et al. 2021). Capsaicin and protons differently modulate the activation kinetics of the mouse TrpV1 channel induced by depolarization (Takahashi et al. 2021). The impact of TRPV1 on cancer pathogenesis and therapy has been reviewed (Li et al. 2021). TRPV1 may be an analgesic target for patients experiencing pain after oral irradiation (Lai et al. 2021). The vanilloid (capsaicin) receptor TRPV1 functions in blood pressure regulation and may be a therapeutic target in hypertension (Szallasi 2023). Chu et al. 2023 elucidated the redox state of C387-C391 mediated long-range allostery of TRPV1, which provided new insights into the activation mechanism of TRPV1.

Animals

TrpV1 or VR1 of Rattus norvegicus

 
1.A.4.2.10

TRPV5 epithelial Ca2+ channel (ECaC1) (forms homo- and heterotetrameric channels with TRPV6; requires the S100A10-annexin 2 complex for routing to the plasma membrane) (Hoenderop et al., 2003; van de Graaf et al., 2003).  The kidney maintains whole body calcium homoeostasis due to the reabsorption of Ca2+ filtered by the kidney glomerulus. TRPV5 regulates urinary Ca2+ excretion by mediating active Ca2+ reabsorption in the distal convoluted tubule of the kidney. The histidine kinase, nucleoside diphosphate kinase B (NDPK-B), activates TRPV5 channel activity and Ca2+ flux, and this activation requires histidine 711 in the carboxy terminal tail of TRPV5. In addition, the histidine phosphatase, protein histidine phosphatase 1 (PHPT1), inhibits NDPK-B activated TRPV5 (Cai et al. 2014).  TRPV5 also transports cadmium (Cd2+). The L530R mutation is associated with recurrent kidney stones (Wang et al. 2017).  May be stabilized by Mucin-1 (Muc1; P15941) (Al-Bataineh et al. 2017). TRPV5 inhibitors have been identified (Hughes et al. 2019). A modular and reusable model of epithelial transport in the proximal convoluted tubule of the kidney has appeared (Noroozbabaee et al. 2022). Only TrpV5 and TrpV6 are calcium selective, while others are general for inorganic cations, and an explanatory mechanism has been proposed (Ives et al. 2023).

Animals

TRPV5 of Homo sapiens (NP_062815)

 
1.A.4.2.11

TRPV6 epithelial Ca2+ channel (ECaC2) (forms homo- and heterotetrameric channels with TRPV5; requires the S100A10-annexin 2 complex for routing to the plasma membrane) (Hoenderop et al., 2003; van de Graaf et al., 2003). Epithelial TrpV6, but not TrpV5, is inhibited by the regulator of G-protein signaling 2 (RGS2; Q9JHX0; 211 aas) by direct binding (Schoeber et al., 2006). Calmodulin (CaM) positively affects TRPV6 activity upon Ca2+ binding to EF-hands 3 and 4, located in the high Ca2+ affinity CaM C-terminus (Lambers et al. 2004). Cyclophilin B is an accessory activating protein (Stumpf et al., 2008).  The crystal structure of rat TRPV6 at 3.25 A resolution revealed shared and unique features compared with other TRP channels (Saotome et al. 2016). Intracellular domains engage in extensive interactions to form an intracellular 'skirt' involved in allosteric modulation. In the K+ channel-like transmembrane domain, Ca2+ selectivity is determined by direct coordination of Ca2+ by a ring of aspartate side chains in the selectivity filter (Saotome et al. 2016).  Replacing Gly-516 within the cytosolic S4-S5 linker (conserved in all TRP channel proteins) by ser forces the channels into an open conformation, thereby enhancing constitutive Ca2+ entry and preventing inactivation (Hofmann et al. 2016). Tetrameric ion channels have either swapped or non-swapped arrangements of the S1-S4 and pore domains. Singh et al. 2017 showed that mutations in the transmembrane domain can result in conversion from a domain-swapped to the non-swapped fold. These results raise the possibility that a single ion channel subtype can fold into either arrangement in vivo, affecting its function in normal or disease states. Cryo-EM structures of human TRPV6 in the open and closed states shows that the channel selectivity filter adopts similar conformations in both states, consistent with its explicit role in ion permeation. The iris-like channel opening is accompanied by an alpha-to-pi-helical transition in the pore-lining transmembrane helix S6 at an alanine hinge just below the selectivity filter. As a result of this transition, the S6 helices bend and rotate, exposing different residues to the ion channel pore in the open and closed states (McGoldrick et al. 2017). TRPV6 is an epithelial Ca2+-selective channel associated with transient neonatal hyperparathyroidism (TNHP), an autosomal-recessive disease caused by TRPV6 mutations that affect maternal-fetal calcium transport (Suzuki et al. 2018). TRPV6 mediates calcium uptake in epithelia, and its expression increases in numerous types of cancer while inhibitors suppress tumor growth. Singh et al. 2018 presented crystal and cryo-EM structures of human and rat TRPV6 bound to 2-aminoethoxydiphenyl borate (2-APB), a TRPV6 inhibitor and modulator of numerous TRP channels. 2-APB binds to TRPV6 in a pocket formed by the cytoplasmic half of the S1-S4 transmembrane helix bundle. 2-APB induces TRPV6 channel closure by modulating protein-lipid interactions. The 2-APB binding site may be present in other members of vanilloid subfamily TRP channels. The crystal structure has been determined (see 30299652 and Yelshanskaya et al. 2020). Novel mutations in TRPV6 give rise to the spectrum of transient neonatal hyperparathyroidism (Suzuki et al. 2020). TRPV6) plays roles in calcium absorption in epithelia and bone and is involved in human diseases including vitamin-D deficiency, osteoporosis, and cancer. Cai et al. 2020 showed that the TRPV6 intramolecular S4-S5 linker to the C-terminal TRP helix (L/C) and N-terminal pre-S1 helix to TRP helix (N/C) interactions, mediated by Arg470:Trp593 and Trp321:Ile597 bonding, respectively, are autoinhibitory and are required for maintaining TRPV6 at basal states. Disruption of either interaction by mutations or blocking peptides activates TRPV6. The N/C interaction depends on the L/C interaction but not inversely. Three cationic residues in S5 or the C terminus are involved in binding PIP2 to suppress both interactions, thereby activating TRPV6 (Cai et al. 2020). The biochemistry and pathophysiology of TRPV6 calcium channels have been reviewed (Walker and Vuister 2023). The structure of human TRPV6 in complex with the plant-derived phytoestrogen genistein, extracted from Styphnolobium japonicum, inhibits cell invasion and metastasis of cancer cells. Cryo-EM combined with other techniques revealed that genistein binds in the intracellular half of the TRPV6 pore and acts as an ion channel blocker and gating modifier. Genistein binding to the open channel causes pore closure and a two-fold symmetrical conformational rearrangement in the S4-S5 and S6-TRP helix regions (Neuberger et al. 2023). TRPV6 is also inhibited by the phytocannabinoid tetrahydrocannabivarin (Neuberger et al. 2023).

Animals

TRPV6 of Homo sapiens (NP_071858)

 
1.A.4.2.12Epithelial calcium channel, ECaC (Liao et al., 2007). Animals ECaC of Danio rerio (Q6JQN0)
 
1.A.4.2.13

TrpV1 of 839 aas and ~ 6 TMSs. Molecular determinants of vanilloid sensitivity have been examined (Gavva et al. 2004).  Ligand-activated non-selective calcium permeant cation channel involved in detection of noxious chemical and thermal stimuli. TRPV1 channels are present in odontoblasts, suggesting that odontoblasts may directly respond to noxious stimuli such as a thermal-heat stimulus (Okumura et al. 2005). It may mediate proton influx and be involved in intracellular acidosis in nociceptive neurons. It is also involved in mediating inflammatory pain and hyperalgesia (Benemei et al. 2015).  The 3.4 Å resolution structure shows that the overall fold is the same as for voltage-gated ion channels (TC# 1.A.1) (Liao et al. 2013). Capsaicin-induced apoptosis in glioma cells is mediated by TRPV1 (Amantini et al. 2007). Capsaicin binds to a pocket formed by the channel's TMSs, where it takes a ""tail-up, head-down"" configuration. Binding is mediated by both hydrogen bonds and van der Waals interactions. Upon binding, capsaicin stabilizes the open state of TRPV1 by ""pull-and-contact"" with the S4-S5 linker (Yang and Zheng 2017). Several protein kinases, including PKD1 (protein kinase D1), Cdk5 (cyclin-dependent kinase 5) and LIMK (LIM- motif containing kinase) regulate TRPV1 and inflammatory thermal hyperalgesia (Zhang and Wang 2017). TrpV1 and TrpA1 are inflammatory mediators causing cutaneous chronic itch in several diseases (Xie and Li 2018). The locations and characteristics of volatile general anesthetic binding sites in the transmembrane domain of TRPV1 have been examined (Jorgensen and Domene 2018). The TRPV1 ion channel is a neuronal sensor that plays an important role in nociception and neuropathic as well as inflammatory pain. In clinical trials, hyperthermia and thermo-hypoaesthesia are major side effects of TRPV1 antagonists (Damann et al. 2020). The TRPV1 ion channel is a polymodal sensor integrating stimuli from molecular modulators with temperature, pH and transmembrane potential. Temperature-dependent gating may constitute the molecular basis for its role in heat sensation and body temperature regulation. Damann et al. 2020 characterized the prototypic small molecule TRPV1 inhibitors GRT12360V and GRTE16523. The oxidizing reagent copper-o-phenanthroline is an open channel blocker of TRPV1 (Tousova et al. 2004). Lack of TRPV1 aggravates obesity-associated hypertension through the disturbance of mitochondrial Ca2+ homeostasis in brown adipose tissue (Li et al. 2022). Lipoic/Capsaicin-related amides are TRPV1 agonists endowed with protective properties against oxidative stress (Brizzi et al. 2022). Agonistic/antagonistic properties of lactones in food flavors on the sensory ion channels, TRPV1 and TRPA1 have been reviewed (Ogawa et al. 2022). TRPV1 channel modulators provide a prospective therapy for diabetic neuropathic pain (Liu et al. 2023). Drosophila appear to possess intricate pain sensitization and modulation mechanisms similar to those in mammals (Jang et al. 2023). Barbamide enhances the effect of the TRPV1 agonist capsaicin and enhanced store-operated calcium entry (SOCE) responses in mice after depletion of intracellular calcium (Hough et al. 2023). The safety and efficacy of topical ocular SAF312 (Libvatrep) in post-photorefractive keratectomy (PRK) pain, an inhibitor of TRPV1, has been evaluated (Thompson et al. 2023). Modulation of membrane trafficking of AQP5 in the lens in response to changes in zonular tension is mediated by TRPV1 (Petrova et al. 2023).


Animals

TrpV1 of Homo sapiens

 
1.A.4.2.14

Epithelial calcium channel 2, ECaC2; TrpV6 of 719 aas and 6 TMSs. It displays all structural features typical for mammalian ECaCs including three ankyrin repeats, six transmembrane domains, and a putative pore region between TM V and TM VI (Qiu and Hogstrand 2004). 

ECaC2 of Takifugu rubripes (Japanese pufferfish) (Fugu rubripes)

 
1.A.4.2.2

Stretch-inhibitable non-selective cation channel, SIC

Animals

SIC of Rattus norvegicus

 
1.A.4.2.3

Vitamin D-responsive, apical, epithelial Ca2+ channel, ECaC

Animals

ECaC of Oryctolagus cuniculus

 
1.A.4.2.4

Insulin-like growth factor I-regulated Ca2+ channel

Animals

IGF-regulated Ca2+ channel of Mus musculus

 
1.A.4.2.5

Vanilloid receptor-related, osmotically activated channel, VR-OAC (also called TRPV4, VRL2, VROAC and Trp12); required for bladder voiding in mice (Gevaert et al., 2007). Regulated by Pacsin3 via its SH3 domain which affects its subcellular localization and inhibits its activity in a stimulus-specific fashion (D'hoedt et al., 2008). Responsible for autosomal dominant brachyolmia (Rock et al., 2008). Multiple gating mechanisms have been demonstrated for TRPV4 (Loukin et al., 2010). TRPV4 Ca2+ signalling regulates endothelial vascular function (Sonkusare et al., 2012) and adipose oxidative metabolism, inflammation and energy homeostasis (Ye et al. 2012).  H2O2 induces Ca2+ influx into microvascular endothelial cells via TrpV4 (Suresh et al. 2015). TrpV4 orthologs are volume-sensors, rather than osmo-sensors (Toft-Bertelsen et al. 2017) that mediate fluid secretion by the ciliary body. They are important for vertebrate vision by providing nutritive support to the cornea and lens, and by maintaining intraocular pressure (Jo et al. 2016). Interacts with the A-kinase anchor protein 5 (AKAP5 or AKAP79 of 427 aas; TC# 8.A.28.1.6; P24588) (Mack and Fischer 2017). Mutations in TRPV4 are associated with accelerated chondrogenic differentiation of dental pulp stem cells (Nonaka et al. 2019). The homolog in Cynops pyrrhogaster (85% identical) is inhibited by RN1734 and may play a role in the sperm acrosome reaction (Kon et al. 2019). TRPV4 antagonism attenuates aortic inflammation and remodeling via decreased smooth muscle cell activation and neutrophil transendothelial migration (Shannon et al. 2020). It forms a tight complex with CD98hc (TC# 8.A.9.2.2) and beta1 integrin (TC# 9.B.87.1.8) in focal adhesions where mechanochemical conversion takes place. CD98hc knock down inhibits TRPV4-mediated calcium influx induced by mechanical forces, but not by chemical activators, thus confirming the mechanospecificity of this signaling response. Molecular analysis revealed that forces applied to beta1 integrin must be transmitted from its cytoplasmic C-terminus via the CD98hc cytoplasmic tail to the ankyrin repeat domain of TRPV4 in order to produce ultra-rapid, force-induced, channel activation within the focal adhesion (Potla et al. 2020). TRPV4 mutations, resulting in severe gain of function, cause mixed neuropathy and skeletal phenotypes in humans (Taga et al. 2022). Cell swelling, heat, and chemical agonists use distinct pathways for the activation of TRPV4 (Vriens et al. 2004). Human TRPV4 is involved in immune activation, and because of its diverse engagement in the neuronal and immune systems, it is a potential therapeutic target for several immune-related disorders (Acharya et al. 2022). It is one of the major non-selective cation channel proteins that plays a crucial role in sensing biotic and abiotic stresses, such as pathogen infection, temperature, mechanical pressure and osmotic pressure changes by regulating Ca2+ homeostasis (He et al. 2022). The structure of human TRPV4 in complex with GTPase RhoAhas been determined, providing a template for the design of future therapeutics for treatment of TRPV4-related diseases (Nadezhdin et al. 2023).

 

Animals

VR-OAC (TrpV4) of Rattus norvegicus

 
1.A.4.2.6Osmosensitive transient receptor potential channel 3, O-TRP3 Animals O-TRP3 of Mus musculus
 
1.A.4.2.7

Intestinal endocyte Ca2+ (Sr2+; Ba2+) entry channel, CaT1. Excision of the Trpv6 gene leads to severe defects in epididymal Ca2+ absorption and male fertility as does the single D541A pore mutation (Weissgerber et al., 2012).

Animals

CaT1 of Rattus norvegicus

 
1.A.4.2.8

The noxious heat (>52°C)-sensitive vanilloid-like receptor cation selective channel, TRPV2. Ca2+-dependent desensitization of TRPV2 channels is mediated by hydrolysis of phosphatidylinositol 4,5-bisphosphate (Mercado et al., 2010).  Deleting the first N-terminal 74 residues preceding the ankyrin repeat domain (ARD) shows a key role for this region in targeting the protein to the membrane. Co-translational insertion of the membrane-embedded region occurs with the TM1-TM4 and TM5-TM6 regions assembling as independent folding domains. ARD is not required for TM domain insertion into the membrane (Doñate-Macian et al. 2015).  The TRPV2 structure has been solved at 4 Å resolution by cryoEM (Zubcevic et al. 2016). Formation of a physical complex between mouse TRPV2 (GRC) and the mouse RGA protein promotes cell surface expression of TRPV2 (Stokes et al. 2005). The role of Ca2+ infllux via TRPV1 in cell death and survival related to cancer has been evaluated (Zhai et al. 2020). A helix-turn-helix motif for high temperature dependence of TRPV2 has been identified (Liu and Qin 2021). As noted above, TRPV2 is a ligand-operated temperature sensor. Zhang et al. 2022 combined calcium imaging and patch-clamp electrophysiology with cryo-EM to explore how TRPV2 activity is modulated by the phytocannabinoid Δ9-tetrahydrocannabiorcol (C16) and by probenecid. C16 and probenecid act in concert to stimulate TRPV2 responses including histamine release from mast cells. Each ligand causes distinct conformational changes in TRPV2. Although the binding for probenecid remains elusive, C16 associates within the vanilloid pocket. As such, the C16 binding location is distinct from that of cannabidiol, partially overlapping with the binding site of the TRPV2 inhibitor piperlongumine (Zhang et al. 2022). The cation-permeable TRPV2 channel is important for cardiac and immune cell function (Gochman et al. 2023). Cannabidiol (CBD), a non-psychoactive cannabinoid of clinical relevance, is one of the few molecules known to activate TRPV2. Using the patch-clamp technique, Gochman et al. 2023 discovered that CBD can sensitize current responses of the rat TRPV2 channel to the synthetic agonist 2-aminoethoxydiphenyl borate (2-APB) by over two orders of magnitude, without sensitizing channels to activation by moderate (40°C) heat. Using cryo-EM, Gochman et al. 2023 uncovered a new small-molecule binding site in the pore domain of rTRPV2 in addition to a nearby CBD site.

Animals

TRPV2 of Homo sapiens

 
1.A.4.2.9

The temperature (heat; >39°C)-sensitive, capsaicin-insensitive receptor cation-selective channel, TRPV3 or TRL3 (may form heterooligomers with VR1 (TRPV1; TC #1.A.4.2.1)). Incensole acetate, an incense component, elicits psychoactivity by activating TRPV3 channels in the brain (Moussaieff et al., 2008).  TRPV3 is activated by synthetic small-molecule chemicals and natural compounds from plants as well as warm temperatures. Its function is regulated by a variety of physiological factors including extracellular divalent cations and acidic pH, intracellular ATP, membrane voltage, and arachidonic acid. It shows a broad expression pattern in both neuronal and non-neuronal tissues including epidermal keratinocytes, epithelial cells in the gut, endothelial cells in blood vessels, and neurons in dorsal root ganglia and the CNS. TRPV3 null mice exhibit abnormal hair morphogenesis and compromised skin barrier function, and it may play critical roles in inflammatory skin disorders, itch, and pain sensation (Luo and Hu 2014). TRPV3 gating involves large rearrangements at the cytoplasmic inter-protomer interface, and this motion triggers coupling between cytoplasmic and transmembrane domains, priming the channel for opening (Zubcevic et al. 2019). Mutations in TRPV3 cause painful focal plantar keratoderma (Peters et al. 2020). TRPV3 is a temperature-sensitive, nonselective cation channel expressed prominently in skin keratinocytes that plays important roles in hair morphogenesis and maintenance of epidermal barrier function. Mechanisms of proton inhibition and sensitization have been discussed (Wang et al. 2021). Mechanisms of proton inhibition and sensitization of TRPV3 have been considered (Wang et al. 2021). TRPV3 is predominantly expressed in skin keratinocytes and has been implicated in cutaneous sensation and associated with numerous skin pathologies and cancers. TRPV3 is inhibited by the natural coumarin derivative osthole, an active ingredient of Cnidium monnieri, which has been used in traditional Chinese medicine for the treatment of various human diseases. Neuberger et al. 2021 presented cryo-EM structures of TRPV3 in complex with osthole, revealing two types of osthole binding sites in the transmembrane region of TRPV3 that coincide with the binding sites of agonist 2-APB. Osthole binding converts the channel pore into a previously unidentified conformation with a widely open selectivity filter and closed intracellular gate. The structures provide insight into competitive inhibition of TRPV3 by osthole (Neuberger et al. 2021). Scutellarein attenuates atopic dermatitis by selectively inhibiting TRP Vanilloid 3 (Wang et al. 2022). TRPV3 involvement in itching, heat pain, hair development, and TRPV3-related skin diseases has been reviewed (Guo et al. 2023).  Temperature-sensitive contact modes allosterically gate TRPV3 (Burns et al. 2023). More than 210 structures from more than 20 different TRP channels have been determined, and all are tetramers. TrpV3 exhibits the pore-dilation phenomenon, whereby prolonged activation leads to increased conductance, permeability to large ions and loss of rectification (Lansky et al. 2023). TRPV3 can exist in a pentameric state which is in dynamic equilibrium with the canonical tetramer through membrane diffusive protomer exchange. The pentamer population increased upon diphenylboronic anhydride (DPBA) addition, an agonist that has been shown to induce TRPV3 pore dilation with a larger pore size (Lansky et al. 2023).

Animals

TRPV3 of Homo sapiens

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.3.1

Olfactory, mechanosensitive channel. Forms a complex with Stim1 and Orai1 (TC# 1.A.52.1.1) which is required for SOC currents (Cheng et al., 2008) (most similar to 1.A.4.8.1, but both are most closely related to 1.A.4.2).  Serves as a chemo-, osmo- and touch sensation receptor (Xiao and Xu 2009).

Animals

Olfactory channel of Caenorhabditis elegans

 
1.A.4.3.2

The Nanchung (Nan) hearing ion channel; mediates hypo-osmotically activated Ca2+ influx in chordotonal neurons of insects (Kim et al., 2003). Nanchung is the "dry" humidity receptor, one of two hygrosensation receptors. These two transient receptor potential channels are needed for sensing humidity.  The other is Water witch (Wtrw), involved in detecting moist air. Neurons associated with specialized sensory hairs in the third segment of the antenna express these channels, and neurons expressing Wtrw and Nan project to central nervous system regions associated with mechanosensation. Construction of the hygrosensing system with opposing receptors may allow an organism to very sensitively detect changes in environmental humidity (Liu et al. 2007). Two commercial insecticides, pymetrozine and pyrifluquinazon, target the heteromeric TRPV ion channel complex which is specifically expressed in the chordotonal organ neurons in Drosophila species and may play roles in male-specific behavior (Mao et al. 2018).

Animals

Nan of Drosophila melanogaster (833 aas; Q9VUD5)

 
1.A.4.3.3

TrpV-type Osm-2 (OSM2) chemo-, osmo- and touch sensation receptor channel (Xiao and Xu 2009). It is also called OCR-2. To survive, C. elegans depends on sensing soluble chemicals with transmembrane proteins (TPs) in the cilia of its chemosensory neurons. Cilia rely on intraflagellar transport (IFT) to facilitate the distribution of cargo, such as TPs, along the ciliary axoneme (van Krugten et al. 2022). IFT and diffusion in ciliary dynamics contribute to ciliary signal transduction and chemosensing.

Animals

Osm-2 of Caenorhabditis elegans

 
1.A.4.3.4

TRP channel homologue of 1240 aas

Brown algae

TRP channel homologue of Ectocarpus siliculosus

 
1.A.4.3.5

TRP channel homologue of 1724 aas

Stramenopiles

TRP channel homologue of Ectocarpus siliculosus (Brown alga)

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.4.1

Vacuolar, voltage-dependent cation-selective, Ca2+-activated channel, YVC1. (Yeast vacuolar conductance protein 1; also called TrpY1; Yor088w) (Chang et al., 2009). Activated by stretch to release vacuolar Ca2+ into the cytoplasm upon osmotic upshock (Zhou et al. 2005). (Also activated by glucose, indole and other aromatic compounds (Haynes et al., 2008; Groppi et al. 2011)).  Glutathione activates by reversible glutathionylation of specific cysteyl residues in YVC1 (Chandel et al. 2016). Channel activity is activated by cytoplasmic Ca2+ and inhibited by vacuolar lumen Ca2+, and two residues, D401 and D405, are involved in Ca2+ sensing in the lumen (Amini et al. 2018). The cryoEM structure of TRPY1 at 3.1 Å resolution in a closed state has been determined (Ahmed et al. 2021). The structure, despite containing an evolutionarily conserved and archetypical transmembrane domain, reveals distinctive structural folds for the cytosolic N and C termini compared with other eukaryotic TRP channels. An inhibitory phosphatidylinositol 3-phosphate (PI(3)P) lipid-binding site, along with two Ca2+-binding sites were identified: a cytosolic site, implicated in channel activation, and a vacuolar lumen site, implicated in inhibition. TRPY1 channel modulation by lipids and Ca2+ have been revealed, and the molecular evolution of TRP channels has been suggested (Ahmed et al. 2021).

Yeast

YVC1 or TrpY1 (Yor088w) of Saccharomyces cerevisiae (Q12324)

 
1.A.4.4.2

Yvc1 or TrpY2 of 678 aas and 9 apparent TMSs. It has the same mechanosenstivity as does the S. cereviseae ortholog (Zhou et al. 2005). 45% identical to the latter protein.

Yvc1 of Kluyveromyces lactis

 
1.A.4.4.3

Yvc-1, Yvc1 or TrpY3 of 676 aas and 9 apparent TMSs.  It has the same mechanosensitive properties of the S. cerevisiae ortholog with TC# 1.A.4.4.1 (Zhou et al. 2005). 57% identical to the latter protein.

TrpY3 of Candida albicans

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.5.1

Mg2+-selective channel/kinase-1; Mg2+-ATP-regulated divalent cation channel, LTRPC7, TRPM7, or TRP-PLIK, of 1862 aas. Bradykinin regulates TRPM7 and its downstream target annexin-1 through a phospholipase C-dependent, protein kinase C-dependent and c-Src-dependent pathway that is cAMP-independent; effects are mediated through the bradykinin type 2 receptor (Callera et al. 2009).  TRPM7 is a Mg2+ sensor and transducer of signaling pathways during stressful environmental conditions. Its kinase can act on its own in chromatin remodeling processes, but TRPM6's kinase activity regulates intracellular trafficking of TRPM7 and TRPM7-dependent cell growth (Cabezas-Bratesco et al. 2015).  Syndecans (proteoglycans) regulate TRPC channels to control cytosolic calcium equilibria and consequent cell behavior. In fibroblasts, ligand interactions with heparan sulfate of syndecan-4 recruit cytoplasmic protein kinase C to target serine714 of TRPC7 with subsequent control of the cytoskeleton and the myofibroblast phenotype (Gopal et al. 2015).  May be associated with melanocytic tumors.  Phenanthrenes, naltriben derivatives, are stimulatory agonist of the TRPM7 channel (Liu et al. 2016). TRP7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells, and its activity is positively regulated by aldosterone and angiotensin II (He et al. 2005). TRPM7 plays an important role in cellular Ca2+, Zn2+ and Mg2+ homeostasis. The protein is  abundantly expressed in ameloblasts and, in the absence of TRPM7, dental enamel is hypomineralized. A role of TRPM7 channels in Ca2+ transport during amelogenesis is likely as it serves both as a modulator of Orai-dependent Ca2+ uptake and as an independent Ca2+ entry pathway, sensitive to pH (Kádár et al. 2021). Recurrent hemiplegic migraine attacks are accompanied by intractable hypomagnesemia due to a de novo TRPM7 gene variant (Lei et al. 2022). TrpM6 is palmitoylated on the C terminal side of its Trp domain, and palmitoylation controls ion channel activity of TrpM7; TrpM7 trafficking is also dependant on its palmitoylation (Gao et al. 2022). The TRPM7-A931T mutation, located in the S3 segment at the interface with the transmembrane region S4, generates an omega current that carries Na+ influx under physiological conditions. A931T produces hyperexcitability and a sustained Na+ influx in trigeminal ganglion neurons that may underlie pain in this kindred with trigeminal neuralgia (Gualdani et al. 2022).

 

Animals

Channel-kinase-1 (LTRPC7) of Homo sapiens

 
1.A.4.5.10
TrpCC family member, Gon2.  Required for initiation and continuation of postembryonic mitotic cell division of gonadal cells Z1 and Z4. Zygotic expression is necessary for hermaphrodite fertility. Probably a cation channel that functions together with Gem1 (TC#2.A.1.13.22) (Kemp et al. 2009).

Animals

Gon-2 of Caenorhabditis elegans

 
1.A.4.5.12

TrpM4 of 1213 aas and 6 TMSs. Calcium-activated non selective cation channel that mediates membrane depolarization. While it is activated by increases in intracellular Ca2+, it is impermeable to it. It does mediate transport of monovalent cations (Na+ > K+ > Cs+ > Li+), leading to depolarize the membrane. It thereby plays a central role in  the function of cardiomyocytes, neurons from entorhinal cortex, dorsal root and vomeronasal neurons, endocrine pancreas cells, kidney epithelial cells, cochlea hair cells etc. It also participates in T-cell activation by modulating Ca2+ oscillations after T lymphocyte activation (Demion et al. 2007). The structure has been determined by cryo EM both with and without ATP (Guo et al. 2017). It consists of multiple transmembrane and cytosolic domains, which assemble into a three-tiered architecture. The N-terminal nucleotide-binding domain and the C-terminal coiled-coil participate in the tetrameric assembly of the channel; ATP binds at the nucleotide-binding domain to inhibit channel activity. TRPM4 has an exceptionally wide filter although it is only permeable to monovalent cations; filter residue Gln973 is essential in defining monovalent selectivity. The S1-S4 domain and the post-S6 TRP domain form the central gating apparatus that probably houses the Ca2+- and PtdIns(4,5)P2-binding sites (Guo et al. 2017).  TRPM4 currents are activated by micromolar concentrations of cytoplasmic Ca2+and progressively desensitized. Zhang et al. 2005 showed that desensitization can be explained by a loss of phosphatidylinositol 4,5-bisphosphate (PI(4,5)P(2)) from the channels. TrpM4 interacts directly with glutamate N-methyl-D-aspartate receptor channels (NMDARs) to promote excitotoxicity. Small-molecule interface inhibitors prevent NMDAR-TRPM4 physical coupling and eliminate excitotoxicity. They are therefore neuroprotectants (Yan et al. 2020). Knockdown of the TRPM4 channel alters cardiac electrophysiology and hemodynamics in a sex- and age-dependent manner in mice (Arullampalam et al. 2023).

TRPM4 of Mus musculus

 
1.A.4.5.13

TRPM8 of the collared flycatcher of 1103 aas.  It is 83% identical to the human ortholog. Its structure has been determined to ~4.1 Å resolution by cryo EM (Yin et al. 2018). The structure reveals a three-layered architecture. The amino-terminal domain with a fold distinct among known TRP structures, together with the carboxyl-terminal region, forms a large two-layered cytosolic ring that extensively interacts with the transmembrane channel layer. The structure suggests that the menthol-binding site is located within the voltage-sensor-like domain and thus provides a structural glimpse of the design principle of the molecular transducer for cold and menthol sensation (Yin et al. 2018). TrpM8 is the primary cold and menthol receptor in humans.  The structure has been solved for the collared flycatcher at 4.1 Å resolution (6BPQ_A - D (Yin et al. 2017). Transient receptor potential cation channel subfamily M member 8, TrpM8, the primary cold and menthol receptor in humans.  The structure has been solved for the collared flycatcher TrpM8 at 4.1 Å resolution (6BPQ_A - D (Yin et al. 2017). Transient receptor potential cation channel subfamily M member 8, TrpM8 is the primary cold and menthol receptor in humans.  The structure has been solved for the collared flycatcher at 4.1 Å resolution (6BPQ_A - D (Yin et al. 2017). Cold thermoreceptor neurons detect temperature drops with highly sensitive molecular machinery concentrated in their peripheral free nerve endings. The main molecular entity responsible for cold transduction in these neurons is the thermo-TRP channel TRPM8. Cold, cooling compounds such as menthol, voltage, and osmolality rises activate this polymodal ion channel. Dysregulation of TRPM8 activity underlies several physiopathological conditions, including painful cold hypersensitivity in response to axonal damage, migraine, dry-eye disease, an overactive bladder, and several forms of cancer. TRPM8 could be an attractive target for treating these highly prevalent diseases. Different mutagenesis approaches have allowed the identification of specific amino acids in the cavity comprised of the S1-S4 and TRP domains that determine modulation by chemical ligands (Pertusa et al. 2023). Different studies revealing specific regions within the N- and C-termini and the transmembrane domain contribute to cold-dependent TRPM8 gating. Pertusa et al. 2023 highlight the milestones in the field: cryo-EM structures of TRPM8 that have provided a better comprehension of the 21 years of research on this ion channel, shedding light on the molecular bases underlying its modulation, and promoting the future rational design of novel drugs to selectively regulate abnormal TRPM8 activity under pathophysiological conditions (Pertusa et al. 2023).

TRP8 of Ficedula albicollis (Collared flycatcher) (Muscicapa albicollis)

 
1.A.4.5.14

TrpM5 of 1165 aas and 8 - 10 TMSs. cryo-EM structures have been solved in an apo closed state, a Ca2+-bound open state, and an antagonist-bound inhibited state. Ruan et al. 2021 defined two novel ligand binding sites: a Ca2+ site (CaICD) in the intracellular domain and an antagonist site in the transmembrane domain (TMD). The CaICD site is unique to TRPM5 and has two roles: modulating the voltage dependence and promoting Ca2+ binding to the CaTMD site, which is conserved throughout TRPM channels. Conformational changes initialized from both Ca2+ sites cooperatively open the ion-conducting pore. The antagonist NDNA wedges into the space between the S1-S4 domain and the pore domain, stabilizing the transmembrane domain in an apo-like closed state (Ruan et al. 2021). It and phospholipase C-β2 colocalize in taste receptor cells (Yoshida et al. 2007).

TrpM45 of Danio rerio (Zebrafish) (Brachydanio rerio)

 
1.A.4.5.2

Melastatin 1 or transient receptor potential melastatin-1 (TRPM1; LTRPC1, MLSN, MLSN1) (a non-selective, Ca2+-permeable cation channel, implicated in cell death (Wilkinson et al., 2008).  Required for dim light vision.  Purified TRPM1 is mostly dimeric. The three-dimensional structure of TRPM1 dimers is characterized by a small putative transmembrane domain and a larger domain with a hollow cavity (Agosto et al. 2014). Since dimers are not likely to be functional ion channels, the authors suggested that additional partner subunits participate in forming the transduction channel required for dim light vision and the ON pathway.  The N-terminal region of TRPM1 (residues L242 to E344) regulates activity by direct interaction by the S100A1 calcium-binding protein (TC# 8.A.81) (Jirku et al. 2016). TRPM1 is required for synaptic transmission between photoreceptors and the ON subtype of bipolar cells (Agosto et al. 2018). Abnormal levels occur in plasma neuron-derived extracellular vesicles of early schizophrenia and other neurodevelopmental diseases (Goetzl et al. 2022).

Animals

Melastatin 1 of Homo sapiens

 
1.A.4.5.3

MLSN1- and TRP-related MTR1 (TrpM5; LTRPC5) of 1165 aas and 6 TMSs.  Associated with the Beckman-Wiedemann Syndrum and causes a predisposition for neoplasia (Prawitt et al. 2000). Involved in taste to bitter, sweet and umami, but not absolutely required for these. Thus, TrpM5-dependent and TrpM5-independent pathways underlie bitter, sweet, and umami tastes (Damak et al. 2006). It plays a role in insulin secretion. It and phospholipase C-β2 colocalize in taste receptor cells of zebrafish (Yoshida et al. 2007). It is a voltage-modulated, Ca2+-activated, monovalent cation (Na+, K+, Cs+) channel (VCAM) that mediates transient membrane depolarization. It is blocked by extracellular acidification but activated by arachidonic acid (Prawitt et al. 2003). The cryoEM structure of TrpM5 in Zebrafish is known (See TC# 1.A.4.5.14).

Animals

MTR1 of Homo sapiens

 
1.A.4.5.4

Intracellular Ca2+-activated nonselective monovalent cation (Na+ and K+) channel (non-permeable to Ca2+), TRPM4b, involved in inherited cardiac arrhythmia syndromes (Amarouch and El Hilaly 2020). It interacts with the TRPC3 channel and suppresses store-operated Ca+ entry (Park et al., 2008).  Contributes to the mammalian atrial action potential (Simard et al. 2013). TRPM4 is widely expressed and is associated with a variety of cardiovascular disorders. Autzen et al. 2018 presented two structures of full-length human TRPM4 embedded in lipid nanodiscs at ~3-angstrom resolution, as determined by single-particle cryo-electron microscopy. These structures, with and without calcium bound, reveal the general architecture for this major subfamily of TRP channels and a well-defined calcium-binding site within the intracellular side of the S1-S4 domain. The structures correspond to two distinct closed states. Calcium binding induces conformational changes that likely prime the channel for voltage-dependent opening (Autzen et al. 2018). TRPM4 functions as a limiting factor for antigen evoked calcium rise in connective tissue type mast cells, and concurrent translocation of TRPM4 into the plasma membrane is part of this mechanism (Rixecker et al. 2016).  Gain-of-function mutations in the TRPM4 activation gate caused progressive symmetric erythrokeratoderma (Wang et al. 2018). Substitution of the 4 residue motif, EPGF, with other amino acids reduced cation binding affinity. Analysis of the human TRPM4 structure indicated that EPGF is located externally to the channel pore (Wei et al. 2022). 

Animals

TRPM4b of Homo sapiens

 
1.A.4.5.5

ADP-ribose/NAD/pyrimidine nucleotide-gated Ca2+ permeable, cation nonselective, long transient receptor potential channel-2, LTRPC2; Melastatin 2; TRPM2 (ATP inhibitable). The 3-D structure resembles a swollen bell shaped structure (Maruyama et al., 2007). It can be converted to an anion-selective channel by introducing a lysyl residue in TMS 6 (Kuhn et al., 2007). It transports Ca2+ and Mg2+ with equal facility (Xia et al., 2008).  Four Ca2+ ions activate TRPM2 channels by binding in deep crevices near the pore but intracellularly of the gate (Csanády and Törocsik, 2009). Protons also regulate activity (Starkus et al., 2010). It is present in the plasma membrane and lysosomes, and plays a role in ROS-induced inflammatory processes and cell death. Melastatin is required for innate immunity against Listeria monocytogenes (Knowles et al., 2011). It functions in pathogen-evoked phagocyte activation, postischemic neuronal apoptosis, and glucose-evoked insulin secretion, by linking these cellular responses to oxidative stress (Tóth and Csanády, 2012).  Pore collapse upon prolonged stimulation underlies irreversible inactivation (Tóth and Csanády 2012).  TRPM2 is preferentially expressed in cells of the myeloid lineage and modulates signaling pathways converging into NF-kB but does not seem to play a major role in myeloid leukemogenesis. Its loss does not augment the cytotoxicity of standard AML chemotherapeutic agents (Haladyna et al. 2016).  TrpM2, expressed in hypothalamic neurons in the brain is a thermosensitive, redox-sensitive channel, required for thermoregulation.  It regulates body temperature, limiting fever and driving hypothermia (Song et al. 2016). Tseng et al. 2016 suggested a mechanistic link between TRPM2-mediated Ca2+ influx and p47 phox signaling to induce excess ROS production and TXNIP-mediated NLRP3 inflammasome activation under high gllucose in Type 2 diabetes Mellitus. The cryoEM strcuture reveals a C-terminal NUDT9 homology (NUDT9H) domain responsible for binding ADP-ribose(ADPR) (Wang et al. 2018). Both ADPR and Ca2+ are required for TRPM2 activation, and structures with ADPR and Ca2+ show both intra- and inter-subunit interactions with the N-terminal TRPM homology region (MHR1/2/3) in the apo state, but undergoing conformational changes upon ADPR binding, resulting in rotation of MHR1/2 and disruption of the inter-subunit interaction. Ca2+ binding further engages transmembrane helices and the conserved TRP helix to cause conformational changes at the MHR arm and the lower gating pore to potentiate channel opening (Wang et al. 2018). Consecutive structural rearrangements and channel activation are induced by binding of ADPR in two indispensable locations, and the binding of Ca2+ in the transmembrane domain (Huang et al. 2019). An N-terminal TRPC2 splice variant of 213 aas inhibits calcium influx (Chu et al. 2005). An antogonists of channel function has been identified (Cruz-Torres et al. 2020). A point mutant of TrpM2 (rs93315) has been identified as a risk factor for bipolar disorder (Mahmuda et al. 2020). Two gates orchestrate the opening of human TRPM2 (Rish et al. 2022). Protein kinase C (PKC)-mediated phosphorylation of TRPM2 Thr738 counteracts the effect of cytosolic Ca2+ and elevates the temperature threshold (Kashio et al. 2022). Citronellal suppresses the expression of NHE1 and TPRM2, alleviates oxidative stress-induced mitochondrial damage, and imposes a protective effect on endothelial dysfunction in type 2 diabetes mellitus rats (Yin et al. 2022). Key residues, E829 and R845, are involved in TRPM2 channel gating (Luo et al. 2022).  TRPM2 is a prognostic factor correlated with immune infiltration in ovarian cancer (Huang et al. 2023).

Animals

LTRPC2 of Homo sapiens

 
1.A.4.5.6

Transient receptor potential cation channel subfamily, member 3, TRPM3. It is subject to muscarinic receptor activation. An alternative ion permeation pathway in TRPM3 allows large inward currents upon hyperpolarization, independently of the central pore.  Four residues in S4 (W982, R985, D988 and G991) are determinants of the properties of the alternative ion permeation pathway (Held et al. 2018).  TRPM3 is a thermosensitive TRP channel, playing a central role in noxious heat sensation. Volitile anesthetics (VAs) inhibit TRPM3-mediated transmembrane currents. Chloroform, halothane, isoflurane and sevoflurane inhibited both the agonist-induced (pregnenolone sulfate, CIM0216) and heat-activated Ca2+ signals and transmembrane currents in a concentration dependent way in cells overexpressing recombinant TRPM3 (Kelemen et al. 2020). Among the tested VAs, halothane was the most potent blocker (IC50=0.52+/-0.05 mM). VAs exerted their effects on native TRPM3 channels expressed in sensory neurons of the dorsal root ganglia. While volatile anesthetics activate certain sensory neurons independently of TRPM3, they strongly and reversibly inhibit the agonist-induced TRPM3 activity (Kelemen et al. 2020).

Animals

TrpM3 of Homo sapiens (Q9HCF6)

 
1.A.4.5.7

Cold-sensitive (opens with decreasing temperatures; e.g., <22°C) and menthol-sensitive cation-selective channel, transient receptor potential melastatin 8 (TRPM8). TRPM8 is activated by low temperatures and cooling agents such as menthol. It underlies the cold-induced excitation of sensory neurons. Its gating is regulated by voltage and lysophospholipids which induce prolonged channel opening (Vanden Abeele et al., 2006; Bautista et al., 2007; Matta and Ahern, 2007). It can be converted to an anion-selective channel by introducing a lysyl residue in TMS 6 (Kuhn et al., 2007). Gating of TRPM8 channels is activated by cold and chemical agonists in planar lipid bilayers (Zakharian et al., 2010). Residues involved in intra- and intersubunit interactions have been identified, and their link with channel activity, sensitivity to icilin, menthol and cold, and their impact on channel oligomerization have been measured (Bidaux et al. 2015).  Targeting the small isoform of TRPM8 may be useful to fight prostate cancer (Bidaux et al. 2016). The human isoform is 83% identical to the TRPM8 of the collared flycatcher (TC# 1.A.4.5.13), the structure of which has been characterized to 4.1 Å resolution (Yin et al. 2018). Activation of TRPM8 by cooling compounds relies on allosteric actions of agonist and the membrane lipid, phosphatidylinositol 4,5-bisphosphate (PIP2). The cryoEM structures of TRPM8 in complex with the synthetic cooling compound icilin, PIP2, and Ca2+, as well as in complex with the menthol analog WS-12 and PIP2 revealed the binding sites for cooling agonists and PIP2 in TRPM8. PIP2 binds to TRPM8 in two different modes, which illustrate the mechanism of allosteric coupling between PIP2 and agonists.
     4TMS-TRPM8 isoforms form functional channels in the ER and participate in regulation of the steady-state Ca2+ concentration in mitochondria and the ER. 4TMS-TRPM8 isoforms are ER Ca2+ release channels (Bidaux et al. 2018). Human PIRT (TC# 8.A.64) attenuates human TPRM8 conductance, unlike mouse PIRT, which enhances mouse TRPM8 conductance (Hilton et al. 2018). PIRT and the TRPM8 S1–S4 domain interact with a 1:1 binding stoichiometry, suggesting that a functional tetrameric TRPM8 channel has four PIRT-binding sites (Hilton et al. 2018). TRPM8 has been implicated in nociception and pain and is regarded as an attractive target for the pharmacological treatment of neuropathic pain syndromes. A series of analogues of N,N'-dibenzyl tryptamine 1, a potent TRPM8 antagonist, were made and studied. Molecular modeling studies identified the putative binding mode of these antagonists, suggesting that they could influence an interaction network between the S1-4 transmembrane segments and the TRP domains of the channel subunits (Bertamino et al. 2018). Cold sensitivity is due to nonconserved residues located within the pore loop (residues 526 - 556) (Pertusa et al. 2018). Diver et al. 2019 have presented cryo-EM structures of TRPM8 in ligand-free, antagonist- or calcium-bound forms, revealing how robust conformational changes give rise to two non-conducting states, closed and desensitized. A malleable ligand-binding pocket accommodates drugs of diverse chemical structures, and delineates the ion permeation pathway, including the contribution of lipids to pore architecture. Direct calcium binding mediates stimulus-evoked desensitization. Large rearrangements within the S4-S5 linker reposition the S1-S4 and pore domains relative to the TRP helix, suggesting a model for modulation of TRPM8 and possibly other TRP channels (Diver et al. 2019). Menthol binding induces wide-spread conformational rearrangements within the transmembrane domains. A temporal sequence of conformational changes in the S6 bundle crossing and the selectivity filter leading to channel activation was estimated (Xu et al. 2020). Yin et al. 2022 presented cryo-electron microscopy structures of mouse TRPM8 in closed, intermediate, and open states along the ligand- and PIP2-dependent gating pathways. The results uncover two discrete agonist sites, state-dependent rearrangements in the gate positions, and a disordered-to-ordered transition of the gate-forming S6 - elucidating the molecular basis of chemically induced cool sensation in mammals.

Animals

TRPM8 of Homo sapiens

 
1.A.4.5.8

The intestinal/renal Mg2+ absorption Mg2+ influx channel, Melastatin6 or TRPM6 (5x higher affinity for Mg2+ than Ca2+; regulated by internal Mg2+) (Voets et al., 2004). TRPM6 and its closest homologue TRPM7 (also a Mg2+-permeable cation channel) assemble to form a functional heterooligomeric channel (Chubanov et al., 2004).  Mutations in TRPM6 promotes hypomagnesemia with secondary hypocalcemia (Chubanov et al., 2007). TRPM6 and the closely related TRPM7 are large channel-kinase proteins (Li et al., 2007; Schmitz et al., 2007). TRPM7 also transports protons competitively with Mg2+ and Ca2+ (Numata and Okada, 2008). Intracellular ATP regulates TRPM6 channel activity via its α-kinase domain independently of α-kinase activity (Thébault et al., 2008). Also plays a role in Zn2+ homeostasis and Zn2+- mediated neuronal injury (Inoue et al., 2010).  The protein is cleaved to release a chromatin-modifying kinase (Krapivinsky et al. 2014).  TRPM7 is a Mg2+ sensor and transducer of signaling pathways under stressful environmental conditions. Its kinase can act on its own in chromatin remodeling processes, but TRPM6's kinase activity regulates intracellular trafficking of TRPM7 and TRPM7-dependent cell growth (Cabezas-Bratesco et al. 2015).  Residues involved in cation selectivity have been identified (Topala et al. 2007); reviewed by Schäffers et al. 2018.  Calmodulin (CaM) and S100A1 share the same binding domain at the TRPM6 N-terminus although the ligand-binding mechanisms may be different (Zouharova et al. 2019). TRPM7 activation potentiates store-operated Ca2+ entry (SOCE) in enamel cells but requires ORAI (Souza Bomfim et al. 2020). TRPM7 is a cation channel that regulates transmembrane Mg2+ and Ca2+ and is involved in a variety of (patho)physiological processes in the cardiovascular system, contributing to hypertension, cardiac fibrosis, inflammation, and atrial arrhythmias (Liu et al. 2023). TRPM7 is a master regulator of the organismal balance of divalent cations that plays an essential role in embryonic development, immune responses, cell mobility, proliferation, and differentiation. It is implicated in neuronal and cardiovascular disorders, tumor progression and is a drug target. Cryo-EM, functional analysis, and molecular dynamics simulations uncovered two distinct structural mechanisms of TRPM7 activation (Nadezhdin et al. 2023).

 

 

Animals

TRPM6 of Homo sapiens (NP_060132)
TRPM7 of Homo sapiens (TC #1.A.4.5.1)

 
1.A.4.5.9

Transient receptor potential cation channel TrpM

Animals

T9.a.14.4.12

rpM of Drosophila melanogaster

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.6.1

Cold-activated cation channel in nociceptive sensory neurons, ANKTM1 (TRPA1; the Wasabi receptor), with lower activation temperature (in the noxious cold range) than TRPM8 (TC #1.A.4.5.7) (Story et al., 2003). Also called TRPA1 (Acc #AAS78661) which translates sound into electric signals in the ear. It sits at the tips of cilia in the inner ear and allows passage of K+ and Ca2+ into the cell. Vibrations in the hair cause the channel to open and close. The frequency of the sound waves generate an electrical signal of the same frequency (Jordt et al., 2004). (Shows 25% identity with α-latrotoxin precursor (TC #1.C.6.3.1.1) in its N-terminal half.) TRPA1 is a polyunsaturated fatty acid sensor in mammals, but not in flies and fish (Motter and Ahern, 2012). TRPA1 is regulated by its N-terminal ankyrin repeat domain (Zayats et al., 2012). Agonistic/antagonistic properties of lactones in food flavors on the sensory ion channels, TRPV1 and TRPA1 have been reviewed (Ogawa et al. 2022).

Animals

ANKTM1 of Mus musculus (Q8BLA8)

 
1.A.4.6.2

Warm-activated thermosensory cation channel of insects, ThermoTRPV, ANKTM1 or TrpA1 (Viswanath et al., 2003). It is required to control activity during the warm part of the day (Roessingh et al. 2015). The TrpA1(A) transcript spliced with exon10b (TrpA1(A)10b) that is present in a subset of midgut enteroendocrine cells (EECs) is critical for uracil-dependent defecation of microorganisms (Du et al. 2016). TrpA1 is a shear stress mechanosensing channel regulating intestinal stem cell proliferation in Drosophila (Gong et al. 2023).

Animals

ANKTM1 of Drosophila melanogaster (1197 aas; Q7Z020)

 
1.A.4.6.3

The nociceptive neuron TRPA1 (Trp-ankyrin 1) (also called the Wasabi Receptor) senses peripheral damage by transmitting pain signals (activated by cold temperatures, pungent compounds and environmental irritants). Noxious compounds also activate through covalent modification of cysteyl residues (Macpherson et al., 2007). TRPA1 is an excitatory, nonselective cation channel implicated in somatosensory function, pain, and neurogenic inflammation. Through covalent modification of cysteine and lysine residues, TRPA1 can be activated by electrophilic compounds, including active ingredients of pungent natural products (e.g., allyl isothiocyanate), environmental irritants (e.g., acrolein), and endogenous ligands (4-hydroxynonenal) (Chen et al., 2008). General anesthetics activate TRPA1 nociceptive ion channels to enhance pain and inflammation (Matta et al., 2008; Leffler et al., 2011). TMS5 is a critical molecular determinant of menthol sensitivity (Xiao et al., 2008) and a variety of inhibitors which are analgesics.  Another class of inhibitors are in the thiadiazole structural class of compounds, and they bind to the TRPA1 ankyrin repeat 6 (Tseng et al. 2018). Inhibitors are potential analgesics. The majority of TRPA1 inhibitors interact with the S5 transmembrane helices, forming part of the pore region of the channel. TRPA1 is a component of the nociceptive response to CO2 (Wang et al., 2010). TRPA1 is a polyunsaturated fatty acid sensor in mammals but not in flies and fish (Motter and Ahern, 2012). It  is regulated by its N-terminal ankyrin repeat domain (Zayats et al., 2012).  Mutations in TrpA1 cause alterred pain perception (Kremeyer et al. 2010). The hop compound, eudesmol, an oxygenated sesquiterpene, activates the channel (Ohara et al. 2015).  These channels regulate heat and cold perception, mechanosensitivity, hearing, inflammation, pain, circadian rhythms, chemoreception, and other processes (Laursen et al. 2014).  TRPA1 is a polymodal ion channel sensitive to temperature and chemical stimuli, but its resposes are species specific (Laursen et al. 2015). A probable binding site for general anesthetics has been identified (Ton et al. 2017), and specific residues involved in binding of the anesthetic, propofol, are known (Woll et al. 2017). TrpV1 and TrpA1 are inflammatory mediators causing cutaneous chronic itch in several diseases (Xie and Li 2018). TRPA1 is specifically activated by natural products including allyl isothiocyanate (mustard oil), cinnamaldehyde (cinnamon), allicin (garlic) and trans-anethole in Fennel Oil (FO) (Memon et al. 2019). Mutations in TRPA1 result in insensitivity to pain promoting algogens such as capsaicin, acid, and allyl isothiocyanate (AITC), have been documented (Eigenbrod et al. 2019). TRPA1 transduces noxious chemical stimuli into nociceptor electrical excitation and neuropeptide release, leading to pain and neurogenic inflammation. It is regulated by the membrane environment.  Startek et al. 2019 found that mouse TRPA1 localizes to cholesterol-rich domains, and that cholesterol depletion decreases channel sensitivity to chemical agonists. Two structural motifs in TMSs 2 and 4 are involved in cholesterol interactions that are necessary for normal agonist sensitivity and plasma membrane localization. TRPA1 is an irritant sensor and a therapeutic target for treating pain, itch, and respiratory diseases. It can be activated by electrophilic compounds such as allyl isothiocyanate (AITC). A class of piperidine carboxamides (PIPCs) are potent noncovalent agonists (Chernov-Rogan et al. 2019). Saikosaponins are channel antogonists (Lee et al. 2019). hTRPA1 is activated by electrophiles such as N-methyl maleimide (NMM). A conformational switch of the protein, possibly associated with activation or desensitization of the ion channel, involves covalent derivatization of several cysteyl and lysyl residues in the transmembrane domain and the proximal N-terminal region as targets for electrophilic activation (Moparthi et al. 2020). Altering expression of the genes encoding Kv1.1, Piezo2, and TRPA1 regulate the response of mechanosensitive muscle nociceptors (Nagaraja et al. 2021). As a polymodal nocisensor, TRPA1 can be activated by thermal and mechanical stimuli as well as a wide range of chemically damaging molecules including small volatile environmental toxicants and endogenous algogenic lipids (Zsidó et al. 2021). After activation by such compounds, the ion channel opens up, allowing calcium influx into the cytosol, inducing signal transduction pathways. Then, calcium influx desensitizes irritant evoked responses and results in an inactive state of the ion channel. It was shown how reversible interactions with binding sites contribute to structural changes of TRPA1, leading to covalent bonding of agonists (Zsidó et al. 2021). The binding site(s) for antagonists have been determined for the TRPA1 ion channel (Gawalska et al. 2022). The hTRPA1 C-terminial domain (CTD) harbors cold and heat sensitive domains allosterically coupled to the S5-S6 pore region and the VSLD, respectively (Moparthi et al. 2022). TRPA1 is a sensor for inflammation and oxidative stress which contribute to the pathophysiology of major depressive disorder (MDD), and TRPA1 channels appear crucial to mediate behavioral impairment induced by chronic corticosterone administration (CCA) (Pereira et al. 2023). Neuronal and non-neuronal TRPA1 are therapeutic targets for pain and headache relief (Iannone et al. 2023). A TRPA1 mutant (R919*), identified in CRAMPT syndrome patients, confers hyperactivity when co-expressed with wild type TRPA1. The R919* mutant co-assembles with WT TRPA1 subunits into heteromeric channels at the plasma membrane. The R919* mutant hyperactivates channels by enhancing agonist sensitivity and calcium permeability, which could account for the observed neuronal hypersensitivity-hyperexcitability symptoms. Possibly, R919* TRPA1 subunits contribute to heteromeric channel sensitization by altering pore architecture and lowering energetic barriers to channel activation (Bali et al. 2023). Platycodonis Radix, a widely consumed herbal food produces a bioactive constituents, platycodins, alleviates LPS-induced lung inflammation through modulation of TRPA1 channels (Yang et al. 2023).

Animals

TRPA1 of Homo sapiens (O75762)

 
1.A.4.6.4The Pyrexia (Pyx) thermal TRP channel allowing increased tolerance to high temperature (Lee et al., 2005)AnimalsPyx of Drosophila melanogaster (Q9W0T5)
 
1.A.4.6.5

Thermosensitive TPR channel TRPA1 (TrpA-1) of 1211 aas.  Detects a temperature drop promoting increased longevity.  This requires TPRA1-mediated Ca2+ influx and activation of protein kinase C.  Human TRPA1 (TC# 1.A.4.6.3) can functionally substitute for worm TRPA-1 in promoting longevity (Xiao et al. 2013).  Also mediates touch sensation.

Animals

TRPA1 of Caenorhabditis elegans

 
1.A.4.6.6

Water witch (Wtrw) of 986 aas, the "moist" humidity receptor, one of two hygrosensation receptors. These two transient receptor potential channels are needed for sensing humidity.  The other is Nanchung (Nan), involved in detecting dry air. Neurons associated with specialized sensory hairs in the third segment of the antenna express these channels, and neurons expressing Wtrw and Nan project to central nervous system regions associated with mechanosensation. Construction of the hygrosensing system with opposing receptors may allow an organism to very sensitively detect changes in environmental humidity (Liu et al. 2007).

WtrW of Drosophila melanogaster

 
1.A.4.6.7

TRP ankyrin 1 (TRPA1 of 1188 aas).  It is a homotetrameric, non-selective, cation channel with multiple ankyrin repeats at the N-terminus.  The systems from insects to birds are heat activatable, and this activation is dependent on an extracellular Ca2+ binding site near the vestibule surface. Neutralization of acidic amino acids by extracellular Ca2+ seems to be important for heat-evoked activation (Kurganov et al. 2017).

TRPA1 of Anolis carolinensis (Green anole) (American chameleon)

 
Examples:

TC#NameOrganismal TypeExample
1.A.4.7.1The mechanically gated hearing and balance ion channel in sensory hair cells of the vertebrate inner ear, NompC (Sidi et al., 2003)AnimalsNompC of Danio rerio (zebrafish) (1614 aas; Q7T1G6)
 
1.A.4.7.2

The sensory ion channel in tactile bristles of insects, NompC.  The atomic structure of Drosophila NOMPC has been determined by single-particle electron cryo-microscopy. Structural analyses suggested that the ankyrin repeat domain (29 repeats) of NOMPC resembles a helical spring, suggesting its role of linking mechanical displacement of the cytoskeleton to the opening of the channel (Jin et al. 2017). Compression of the ankyrin chains imparts a rotational torque on the TRP domain, which may result in channel opening (Argudo et al. 2019).

Animals

NompC of Drosophila melanogaster (1619 aas; AAF59842)

 
1.A.4.7.3

The pore forming subunit, Trp-4, a mechanosensitive cation/Ca2+ channel. Present in ciliated mechanosensitive neurons; Activation and latency occur in the microsecond range. trp-4 mutations alter ion selectivity (Kang et al., 2010; Xiao and Xu 2009). 

Animals

Trp-4 of Caenorhabditis elegans (Q9GRV5)

 
Examples:

TC#NameOrganismal TypeExample
Examples:

TC#NameOrganismal TypeExample