Molecular Pharmacology of K2P Potassium Channels

 

Niels Dechera    Susanne Rinnéa    Mauricio Bedoyab,c    Wendy Gonzalezb,c

Aytug K. Kipera

 

aVegetative Physiology, Institute for Physiology and Pathophysiology, Philipps-University Marburg, Marburg, Germany, bCentro de Bioinformática y Simulación Molecular, Universidad de Talca, Talca, Chile, cMillennium Nucleus of Ion Channels-Associated Diseases (MiNICAD), Universidad de Talca, Talca, Chile

 

 

 

 

Key Words

Drug binding sites • K2P potassium channels • Ion channels • Molecular pharmacology

 

Abstract

Potassium channels of the tandem of two-pore-domain (K2P) family were among the last potassium channels cloned. However, recent progress in understanding their physiological relevance and molecular pharmacology revealed their therapeutic potential and thus these channels evolved as major drug targets against a large variety of diseases. However, after the initial cloning of the fifteen family members there was a lack of potent and/or selective modulators. By now a large variety of K2P channel modulators (activators and blockers) have been described, especially for TASK-1, TASK-3, TREK-1, TREK2, TRAAK and TRESK channels. Recently obtained crystal structures of K2P channels, alanine scanning approaches to map drug binding sites, in silico experiments with molecular dynamics simulations (MDs) combined with electrophysiological studies to reveal the mechanism of channel inhibition/activation, yielded a good understanding of the molecular pharmacology of these channels. Besides summarizing drugs that were identified to modulate K2P channels, the main focus of this article is on describing the differential binding sites and mechanisms of channel modulation that are utilized by the different K2P channel blockers and activators.

 

 

Introduction

 

Tandem of two-pore-domain potassium (K2P) channels belong to the latest family of potassium channels cloned, with TOK1 from Saccharomyces cerevisiae as the first channel discovered in 1995 [1]. The mammalian K2P potassium channel family contains 15 members with different subfamilies, characterized by mechanistic hallmarks like acid inhibition, stretch activation, alkaline activation and halothane inhibition (shown in Fig. 1a). K2P channels have four transmembrane domains and contain two pore loops and thus assemble as dimer in order that four pore loops form the potassium selectivity filter, similar as in other potassium channel families (shown in Fig. 1b, 1c). The channels have a large extracellular M1-P1 loop which has been identified in early studies as self-interacting domain (‘SID’) and is relevant for the dimerization of the channels (shown in Fig. 1b) [2]. In fact, the extracellular M1-P1 loop forms the so called ‘cap’ structure which is a structural hallmark of the K2P channels (shown in Fig. 1b, 1c) [3-9]. It has been postulated that the extracellular ‘cap’ prevents classical toxins from binding to the pore region [4]. However, whether there are other physiological functions that can be assigned to this unique structure has not been addressed yet. Although K2P channels were initially described as leak channels with an outward rectification that appeared to solely obey the Goldman-Hodgkin-Katz equation for a potassium selective hole, we know in the meantime that these channels are highly regulated by a plethora of different stimuli (shown in Fig. 1a). Moreover, these channels are in fact voltage sensitive with potassium acting as the actual voltage sensor, meaning similar to the CLC chloride channel family [10], the permeating ion actually also gates the channel [11]. Thus, K2P channels are potassium gated potassium channels with a potassium efflux increasing the open probability of the channel at depolarized potentials.

Initially, K2P channels were described to have a unique pharmacology compared to other potassium channel families, as they were less sensitive to the classical potassium channel blockers like TEA, 4-AP, Cs+ or Ba2+. Channels of the K2P family appeared to be drug resistant. However, classical open channel blockers of ion channels share a conserved binding site scheme with about two or more interacting residues of the pore forming helices and an additional binding to residues of the pore signature sequence (shown in Fig. 1d). This binding site pattern can be found in different potassium channel families, but also in voltage-gated sodium and calcium channels [12-20]. Thus, it appears unlikely that K2P channels in general should be resistant to classical pore blockers. In the meantime, we know that K2P channels are also highly sensitive to classical pore block by quaternary ammonium compounds (QA) (shown in Fig. 1e), however, presumably due to differences in the architecture of the central cavity providing more lateral space underneath the selectivity filter, the QAs need longer alkyl side chains, like TPenA or THexA (shown in Fig. 1e), to be stabilized underneath the selectivity filter by interact with the pore forming helices [21]. While the sensitivity of TASK-3 channels to extracellular polyamines and ruthenium red (RR) has been described very early after cloning of the TASK channels [22-29], there was for a longer time a lack of potent small compound inhibitors. A decade after cloning of the first K2P channels and TASK‑1 [23, 30, 31], we described the first potent K2P channel blocker A293 [32]. Strikingly, the development of a potent and selective TASK-1 channel blocker opened the door for functional studies of the channel in native tissue [33], leading for instance to the isolation of a whole cell TASK-1 current in ventricular myocytes [32]. However, for most of the K2P channels there are still no highly potent and selective blockers or activators available (shown in Table 1-8). Thus, we eagerly anticipate the development of small compound modulators for all K2P channels as this will, besides the generation and study of transgenic mouse models, definitely foster our research aiming to understand the physiological role of K2P channels in different organs.

However, while we still have a poor understanding of the pharmacology of TWIK, THIK and TALK channels, there are in the meantime many potent activators and blockers described for the channels of the TASK and TREK-subfamily, which we would like to summarize in separate sections, especially by focusing on the mechanistic models of channel modulation and the differential binding sites utilized by the drugs.

 

Fig. 1. The K2P channel family. (a) Dendrogram of K2P channels with their physiological or pharmacological key modulators. TWIK: Tandem of P-domains in a weak inward rectifying K+ channel, TREK: TWIK-related K+ channel, TRAAK: TWIK-related arachidonic acid activated K+ channel, TASK: TWIK-related acid-sensitive K+ channel, TALK: TWIK-related alkaline activated K+ channel, THIK: TWIK-related halothane inhibited K+ channel, TRESK: TWIK-related spinal cord K+ channel. (b) Membrane topology of K2P channels depicting transmembrane domains in blue, the pore helices in red and the extracellular 'cap' structure in purple. (c) Illustration of the crystal structure of the K2P channel TWIK-1 (PDB ID: 3UKM). Transmembrane domains are illustrated in blue, the pore helices in red and the extracellular 'cap' structure in purple. (d) Illustration of the conserved scheme of drug binding sites in voltage gated ion channels. The typical binding sites of classical pore blockers are indicated with purple stars. CC: central cavity; SF: selectivity filter (e) Chemical formula of quaternary ammonium compounds, which are classical pore blockers with different alkyl side chain lengths.

 

 

Pharmacology of K2P channels of the TWIK, THIK and TALK subfamilies

 

For K2P channels of the TWIK, THIK and TALK subfamilies a low potency block was observed for instance by local anesthetics or quinidine (shown in Table 1). Similar as for other K2P channels these subfamilies are also sensitive to volatile anesthetics (shown in Table 1, 2). In addition, many anti-arrhythmic drugs were reported to block TASK-4/(TALK-2) channels, albeit also with a very low potency (shown in Table 1). Thus, for the K2P channels of the TWIK, THIK and TALK subfamilies there are to our best knowledge no small compound activators or blockers reported that are active in the submicromolar range.

 

Table 1. List of inhibitors of the TWIK, THIK and TALK K2P channel subfamilies

Table 2. List of activators of the TWIK, THIK and TALK K2P channel subfamilies

 

 

Pharmacology of the TRESK channel

 

TRESK is a K2P channel that was reported to be primarily or almost exclusively expressed in the spinal cord and DRG neurons [34, 35] which posed this channel on the list of novel promising drugs targeting pain sensation. Noteworthy, TRESK was also found in other tissue like the heart and lung [36]. As a channel sharing 65% identity with human TRESK was isolated from mouse testis, the authors termed this as TRESK-2 [34], however later on it became clear, that this channel is the mouse orthologue of TRESK and that there is only one channel within this subfamily of K2P channels. In contrast to the channels of the TWIK, THIK and TALK subfamilies, TRESK appears to be more ‘druggable’ since many blockers and/or activators were already described (shown in Table 3, 4). TRESK is blocked by many small compound inhibitors and also with fairly low IC50 values (shown in Table 3), for instance by the antihistaminic drug loratadine, with an IC50 of about 1 µM [37]. Also many activators have been described (shown in Table 4), here some compounds exhibit even EC50 values in the submicromolar range. For instance, acetyl-B-methylcholine, oxotremorine and OXA-22 activate TRESK with an EC50 of about 700 nM, 300 nM and 100 nM, respectively [37] (shown in Table 4). Why it appears more feasible to identify modulators of TRESK channels than for members of the TWIK, THIK and TALK subfamilies is an open question.

 

Table 3. List of inhibitors of the TRESK K2P channel subfamily

Table 4. List of activators of the TRESK K2P channel subfamily

 

 

Pharmacology of channels from the TREK/TRAAK subfamily

 

TREK/TRAAK activators

A hallmark of the members of the TREK/TRAAK subfamily of K2P channels is the channel modulation by polyunsaturated fatty acids (PUFAs), such as arachidonic acid (AA) [38, 39] (shown in Fig. 2a and Table 5). Analyses of deletion constructs demonstrated that the C-terminus of TREK-1 is crucial for the response to AA [39]. In addition, replacing the C-terminus of TREK-2 with the C-terminus of TASK‑3 abolished the sensitivity to AA. However, replacing the C-terminus of TRAAK with that of TASK-1 or TASK-3 did not affect the response to AA. These results show that the mechanism of activation of TRAAK and TREK by fatty acids may be different [40].

Volatile anesthetics, including diethyl ether, halothane, isoflurane or chloroform, activate TREK-1 and TREK-2 (but not TRAAK) channels which requires the C-terminus of the channels [41] (shown in Table 5). Chloroform specifically and reversibly activates TREK channels [41], whereas halothane or isoflurane activates both, TASK and TREK channels [41] (shown in Fig. 2a and Table 5, 6). On the contrary, diethyl ether increased TREK-1 whereas it decreased TASK-1 currents [41]. In addition, anesthetic gases, as nitrous oxide, xenon and cyclopropane activate TREK-1 channels in clinically relevant concentrations, whereas TASK-3 is insensitive [42] (shown in Table 5). The Glu306 residue, also critical for TREK-1 modulation by AA, stretch or internal pH, was shown to be important for TREK-1 channel activation by these anesthetic gases [42].

In 2000 it was shown by Duprat et al. that the neuroprotective agent riluzole activates TREK‑1 and TRAAK channels (shown in Fig. 2a and Table 5). As TREK-1 is inhibited by increased cAMP levels via PKA phosphorylation and riluzole has the capacity of increasing cAMP levels, the activation of TREK-1 is only transient. In contrast, TRAAK channels, which lack a PKA inhibition, are permanently activated [43].

TREK-1 channels are discussed as novel drug targets against pain [44, 45]. Devilliers et al. demonstrated, that TREK-1 contributes to morphine-induced analgesia in mice. The channel was directly activated (independent of µ opioid receptor activation) leading to analgesia without adverse effects [46].

Another TREK-1 activator with therapeutic potential is BL-1249 [47] (shown in Fig. 2a and Table 5). Using a whole exome sequencing approach in a patient with right ventricular outflow tract tachycardia (RVOT-VT), Decher et al. identified the TREK-1I267T mutation, an amino acid exchange located directly before the selectivity filter of the second pore loop [47]. The mutation almost completely abolished outward currents through the channel and introduced a strong sodium permeability to the potassium channel [47]. Interestingly, application of BL-1249 rescued the potassium selectivity and loss-of-function of the TREK-1I267T channel. The fact that this fenamate-like compound was able to rescue the selectivity filter defect of TREK-1I267T indicated that this drug or maybe other similar activators directly stabilize the selectivity filter. Consequently, Schewe et al. identified negatively charged activators (NCAs) harboring a negatively charged tetrazole or carboxylate group (such as BL-1249, PD-118057 and NS11021) as activators of the mechano-gated K2P channels TREK-1 and TREK-2 [48] (shown in Fig. 2a and Table 5). Strikingly, these NCAs activate most of the K2P channels (shown in Table 2, 4, 5) and also activate other selectivity filter gated ion channels like hERG or BK channels with equal efficiency [48]. Thus, NCAs act with a common mechanism on selectivity filter gated channels, providing a universal ‘master key’ to unlock the selectivity filter gate by binding below the selectivity filter where their negative charge promotes K+ binding to the pore cavity (shown in Fig. 2b). This in turn alters the ion occupancy in the selectivity filter in a way that is known to promote activation of the filter gate [11, 48].

In contrast, 2-aminoethoxydiphenyl borate (2-APB) is a non NCA that activates channels of the TREK/TRAAK subfamily (shown in Fig. 2a and Table 5). TREK-2 is much more sensitive to modulation by 2-APB compared with TREK-1 or TRAAK [49] (shown in Table 5). 2-APB does not bind to either the binding site of NCAs or the ‘cryptic’ binding site described below. Zhuo et al. described that for TREK-2 channels the cytosolic C‑terminus plays a role in controlling the stimulatory effects of the compound. In particular the proximal C-terminus, including His368 as a key residue, was required for channel activation by 2-APB [50]. In addition, specific mutations in the M4 segment reduced the 2-APB efficiency and thus the authors proposed an allosteric coupling between the proximal C‑terminus and the selectivity filter induced by 2-APB [51]. This allosteric coupling is facilitated by the movement of M4 and thus mutations that reduce the flexibility of the M4 movements impair 2-APB activation [51].

Very recently a novel class of small molecule activators has been identified that utilizes a completely different, the ‘cryptic’, binding site, which is clearly distinct to that of NCAs [6] (shown in Fig. 2c). ML335 and ML402 bind to an L-shaped pocket behind the selectivity filter formed by the P1 pore helix and M4 transmembrane helix intrasubunit interface. The drugs activate the channels by acting as ‘molecular wedges‘, restricting the interdomain interface movement behind the selectivity filter [6]. Mechanistically, binding to the ‘cryptic’ binding site stabilizes the C-type gate in a more conductive ‘leak mode’ through a common set of hydrogen bonds, π-π, and cation-π interactions of ML335 and ML402 with the ‘P1 face’ and an ‘M4 face’ reducing P1/TM4 interface dynamics [6].

A high-throughput fluorescence-based thallium flux screen identified small molecules that selectively activated TREK-2 [52]. These novel compounds were subsequently proven to directly activate TREK-2 channels, using single channel measurements in excised membrane patches [52]. Strikingly, 11-deoxy PGF2α or T2A3 which were described in this study, activated TREK-2 while they blocked TREK-1 channels [52] (shown in Table 5, 7). For these compounds a region connecting the second pore loop to the M4 segment was proposed to determine the observed activation or inhibition [52].

 

Fig. 2. Drugs that modulate channels of the TREK subfamily of K2P channels and different binding sites identified. (a) Chemical formula of the most important activators for channels of the TREK/(TRAAK) subfamily. (b) Binding site of the 'NCA' BL-1249 mapped in the TREK-2 crystal structure template (PDB ID: 4XDJ) [48]. (c) ML402 bound in the 'cryptic' binding site of TREK-1 determined by a co-crystallization study (PDB ID: 6CQ9) [6]. (d) Binding site of norfluoxetine in TREK-2, see co-crystal structure PDB ID: 4XDL [5]. (e) Binding site of ruthenium red at the 'keystone' binding site, determined by co-crystallization with TREK-1I110D (PDB ID: 6V3C) [58]. (f) Chemical formula of the most important inhibitors for channels of the TREK/TRAAK subfamily. (b-e) Potassium ions are represented by black spheres. Red arrows indicate the positions of the respective drugs in top view.

Table 5. List of activators of the TREK/(TRAAK) K2P channel subfamily

Table 6. List of activators of the TASK K2P channel subfamily

Table 7. List of inhibitors of the TREK/(TRAAK) K2P channel subfamily

 

TREK/TRAAK inhibitors

In terms of blockers for channels of the TREK/TRAAK subfamily, spadin is presumably the best known blocker of TREK channels described (shown in Fig. 2f). Spadin is a 17 amino acid sortilin-derived peptide targeting TREK-1 channels with an IC50 of 70 nM [53] (shown in Fig. 2f and Table 7). It is discussed as a putative antidepressant [53, 54], although the binding site and mechanism of inhibition are not known so far. However, it has been postulated that spadin should bind to the ‘down state’ of the channel to specifically antagonize activation of TREK-1 by AA, utilizing an allosteric mechanism of inhibition [55].

Dong et al. described the crystal structure of TREK-2 in complex with norfluoxetine [5] (shown in Fig. 2d, 2f). Here several residues in the side fenestrations, including Ile194 and Pro198 of the M2 segment, Cys249 and Val253 in M3, Phe316 and Leu320 in M4, as well as Val276, Leu279 and Thr280 of the second pore helix, close to the selectivity filter, were proposed to interact with the compound [5]. Norfluoxetine binds to the channel in the ‘down state’, presumably impairing the transition to the ‘up state’ from which the channel opening preferentially occurs [56]. Surprisingly, although norfluoxetine appears to preferentially bind to the ‘down state’ [5] this state dependence is not reflected by a voltage-dependent inhibition of TREK channels [57].

RR inhibits a number of ion channels including members of the K2P channel family [28, 29] (shown in Table 7, 8) with E70 in TASK-3 [29] and D135 in TREK-2 [28] as key residue for RR action. In contrast, TREK-1 is not sensitive to RR [28]. Using X-ray crystal structures of a RR sensitive TREK-1 mutant (TREK-1I110D) alone or complexed with RR revealed that a negatively charged residue at this specific site provides the key inhibitor site in the extracellular ion pathway (‘EIP’) above the selectivity filter which is formed by the ‘cap’ structure (shown in Fig. 2e). Binding of RR to this site occludes the ‘EIP’ and thus the current flux [58].

The methanesulfonamide TKDC was described as another small molecule inhibitor of the TREK/TRAAK subfamily (shown in Fig. 2f and Table 7). However, TKDC blocks with a novel and unusual allosteric mechanism. Luo et al. identified an allosteric ligand-binding site located in the extracellular ‘cap’ of the channels. From this site the ligands are supposed to induce an allosteric conformational transition which ultimately leads to an obstruction of the ‘EIP’ [59].

 

Table 8. List of inhibitors of the TASK K2P channel subfamily

 

 

Pharmacology of TASK channel subfamily members

 

TASK-1 and TASK-3 blockers

TASK-1 transcripts and currents are upregulated under atrial fibrillation (AF) [60, 61], variants of KCNK3, encoding TASK-1, are associated with AF [62] and genetic ablation of KCNK3 by a dominant negative viral approach suppresses AF in a pacemaker-induced AF model of the pig [63]. Furthermore, both, TASK-1 and TASK-3, are expressed in the carotid body and brain stem regions associated with respiratory control, and mice lacking these channels have impaired carotid body function [64]. Thus, TASK channel inhibition is for instance a promising therapeutic approach for the treatment of AF (DOCTOS Trial) or breathing disorders like obstructive sleep apnea (OSA) (SANDMAN Trial) [65-67].

In 2007 Putzke et al. described A293 (shown in Fig. 3a), the first potent K2P channel blocker [32] with an IC50 of 222 nM on TASK-1 expressed in Xenopus oocytes (shown in Table 8), enabling the isolation of the first native whole cell current of a K2P channel, the ITASK‑1 in rat, mouse and human cardiomyocytes [32, 68, 69]. A few years later we described the first highly potent and selective TASK-1 channel blocker being active in the one digit nanomolar range. The IC50 of A1899 on TASK-1 expressed in CHO cell was 7 nM and in oocytes 35.1 nM [70] (shown in Fig. 3a and Table 8). A1899 was in the submicromolar range not active on a plethora of ion channels tested [70] and thus A1899 is an even more promising tool than A293 in terms of specificity. Subsequently, using this compound we described the first drug binding site of a K2P channel which helped understanding the pore structure of these channels, as these were not crystallized at this time [70]. Using an alanine scanning mutagenesis approach we found that the M2 and M4 segments form the inner pore of K2P channels and which residues actually face into the central cavity [70]. The A1899 binding site is formed by Thr93 of the first pore loop, Ile118 and Leu122 of the M2 segment, Thr199 of the second pore loop, Ile235, Gly236, Leu239 and Asn240 of the M4 segment and Val243 and Met247 of the halothane response element (‘HRE’) [70]. Noteworthy, the IC50 of A1899 for TASK-3 is 10-fold higher than that of TASK-1 [70] (shown in Table 8). The binding site of A1899 in TASK-1 is fully conserved in the TASK-1/3/5 subfamily except for one residue. This amino acid variation is located in the ‘HRE’ of TASK-3, which is 243VLRFMT248 for TASK-1 and 243VLRFLT248 for TASK-3. This sequence variation might contribute to the different drug affinities of TASK-1 and TASK-3, since the TASK-1M247L mutation causes a 3.3-fold increase in IC50 for A1899 [70].

A few years later we found that blockers of the Kv1.5 channel which were developed as antiarrhythmic compounds to treat or prevent AF, are much more potent inhibitors of TASK-1 than Kv1.5 [65]. Note that A1899 was initially developed by Sanofi as a blocker of Kv1.5 (S0200951) and A293 was initially described as the Kv1.5 blocker AVE1231 which was under clinical investigation against AF [65]. However, both compounds were about 70-fold more potent on TASK-1, making them TASK selective when low doses of the compounds are applied [65]. These data suggest that the real channel, effectively targeted against AF by Kv1.5 blockers was TASK-1 and not Kv1.5, further supporting the notion that TASK-1 might be a promising drug target against AF and OSA [65]. However, it also raised the question how different compounds can efficiently block both, TASK-1 and Kv1.5 channels, albeit they belong to only very remotely related families of potassium channel. Surprisingly, in silico analyses comparing the binding sites in TASK-1 and Kv1.5 revealed important similarities. For both channels, the drug binding sites are formed by a ring of threonine residues at the signature sequence of the selectivity filter plus three layers of lipophilic residues facing the central cavity underneath the selectivity filter [65]. We proposed that the accessibility of the drug to the pore and the more lipophilic environment of TASK-1 might be the reason why most of the Kv1.5 blockers are even more potent on TASK-1 [65].

The binding mode of A1899 to the TASK-1 channel pore was initially modelled on a KvAP-based homology model that has a fourfold symmetric pore [70]. However, since K2P channels do not have such a fourfold symmetry in the central cavity, Ramirez et al. re-evaluated the A1899 binding site in TASK-1 using a pore homology model of TASK-1 based on TWIK-1 (shown in Fig 3b). TWIK-1 was, together with TRAAK, one of the first K2P channels crystallized and TWIK-1 is more closely related to TASK-1 than TRAAK. The TWIK-1 based homology model, combined with docking experiments and MD simulations revealed that A1899 binds to residues located in the side fenestrations providing a physical ‘anchor’, reflecting an energetically favorable binding mode that after pore occlusion stabilizes the closed state of the channel [71] (shown in Fig. 3b).

Subsequently we reported the binding site of the antiarrhythmic compounds A293 in TASK‑1 which partially overlaps with the A1899 binding site [72] (shown in Fig. 3c). Although, the A293 binding site has not been studied as detailed as that of A1899, it appears that A293 binds at a lower position within the central cavity, nearby the opening of the lateral fenestrations, however without parts of the drugs extending downwards to the halothane response element (‘T’ shaped binding mode of A1899) (compare Fig. 3b versus Fig. 3c).

A1899 and PK-THPP are effective breathing stimulants in rats and thus both compounds may have therapeutic potential for treating breathing disorders [67]. PK-THPP which more potently blocks TASK-3 than TASK-1 channels (shown in Fig. 3a and Table 8), also shared several residues of the A1899 binding site in the central cavity [73-76] (shown in Fig. 3d). While L122, L239, G236 and L247 (M247 in TASK-1) were identified to be part of both binding sites, there were also some novel residues identified as relevant for PK-THPP inhibition (Q126, G231, A237, V242, L244 and T248) [73, 74]. On the other hand, other residues were found to be important exclusively for A1899 binding (I118, I235, N240 and V243) [70]. The binding of PK-THPP to TASK-3 depends on the state of the fenestration, as PK-THPP exclusively binds to the open state [74]. Whether differences in the state dependence and affinity towards TASK-1 and TASK-3 (shown in Table 8) depends on the different binding mode of PK-THPP and A1899, especially the different set of residues identified in the late M4 segment and halothane response element, remains an open question.

Following a high throughput fluorescent screen and structure activity relationship analysis of active compounds, ML365, a bisamide, was identified as another promising TASK channel blocker [76] (shown in Table 8). ML365 has an IC50 of 4 nM in a thallium flux fluorescent assay and an IC50 of 16 nM in an automated electrophysiology assay [76]. The small molecule inhibitor displayed little or no effect on more distantly related potassium channels like Kir2.1, KCNQ2 or hERG after application of 30 μM of the compound [76]. ML365 showed a 62-fold more potent IC50 for TASK-1, than for the closely related TASK-3 channel, thus displaying the strongest ‘split’ in pharmacology between TASK-1 and TASK-3 channels that was described so far [76] (shown in Table 8). However, the molecular explanation for this phenomenon was not addressed so far.

Interestingly, channel inhibition by low-affinity antiarrhythmic compounds, such as carvedilol, propafenone and amiodarone was also affected by mutations of the A1899 and A293 binding site [70, 72]. Also, the respiratory stimulant doxapram is a blocker of TASK channels [77, 78] (shown in Fig. 3a and Table 8) that acts at this common intracellular binding site, located in the inner vestibule of TASK-1 [76] and TASK-3 [73]. Hence, there is an overlap for residues in the TM2 and TM4 regions for different compounds arguing for a conserved common site of action, however with substance specific variations in the binding mode that appear to modulate affinity and/or specificity.

Rinné et al. described a novel binding site for the local anesthetic bupivacaine, differing from those described above which results in an allosteric and voltage-dependent inhibition of TASK-1 and TASK-3 channels [79] (shown in Fig. 3a, 3e). A large alanine scanning mutagenesis approach identified residues that include I118 in M2 and I235, G236, L239 and N240 in M4, which were also part of the A1899 binding site, however there were several novel 'hits' in the M2 segment (C110, M111, A114, Q126, S127) as well as in the M4 segment (V234A and F238A). Bupivacaine was located laterally underneath the pore helices, in the side fenestrations, previously described for other K2P channels (shown in Fig. 3e, top view). Thus, bupivacaine was proposed to act by an allosteric mechanism disrupting the voltage-dependent K+-flux gating [11] at the selectivity filter [79].

Recently, the TASK-1 channel crystal structure was reported revealing an unexpected second gate located at the entrance to the inner vestibule which was not observed in the structures of other K2P channels crystallized so far [7] (shown in Fig. 3f). This observation was very unexpected as K2P channels were thought to be exclusively gated at the selectivity filter. This lower gate was created by interaction of two M4 helices for which the C-terminal ends crossed underneath the central cavity, prompting us to term it ‘X-gate’ [7]. Strikingly, the ‘X-gate’ is actually formed by amino acid residues of the ‘HRE’ motif, V243 to T248, which was previously described to be essential for the regulation of the channel by Gq pathways, volatile anesthetics and drugs [70, 80]. Two bends can be observed in the M4 segment, one before the ‘X-gate’ which is supposedly a gating ‘hinge’ relevant for the positioning of the extended alpha helix that actually forms the ‘X-gate’ and a second bend observed following the ‘X-gate’ at residue Asn250, allowing the distal end of M4 to adopt an α-helical structure which forms extensive interactions with the early M1 and late M2 segment [7]. This region which stabilizes the ‘X-gate’ and thus the closed state of the channel was named ‘latch’. In the progress of this study our co-workers at Bayer identified by ultra-high throughput screening (uHTS) novel highly potent TASK blockers, namely BAY1000493 (shown in Fig. 3a, 3f and Table 8) and BAY2341237 (shown in Table 8), which were subsequently co-crystallized with TASK-1 [7]. Both drugs were bound in a planar orientation directly below the selectivity filter, interacting with several key residues previously described by Streit et al. for the binding of A1899 [7, 70]. The planar orientation of the compound is stabilized by an interaction of the drugs with L122 of both subunits. Rödström et al. proposed, that the blockers get trapped by the ‘X-gate’ within the inner vestibule, explaining the slow wash-out rates and almost irreversible inhibition of TASK-1 by these blockers [7]. Strikingly, mutations that are thought to destabilize the closed ‘X-gate’ did not only cause an increased open probability of the channel, but also impaired the ability of the ‘X-gate’ to trap BAY1000493 [7]. The higher potency of BAY1000493 in comparison to A1899 (shown in Table 8) might be reflected by the T-shaped binding mode of the A1899 (shown in Fig. 3b) for which parts of the compound extend all the way down to span the narrowest restriction point of the ‘X-gate’ (at residue L244) [7] and to interact with M247 of the ‘HRE’ motif [70, 71]. This binding mode might prevent an efficient trapping of A1899. Compound trapping in the central cavity combined with a complementarity between the shapes of the inhibitor and the upper vestibule appear to be important for high-affinity drug binding in TASK-1 channels.

 

Fig. 3. Drugs that modulate channels of the TASK subfamily of K2P channels and different binding sites identified. (a) Chemical formula of the most important inhibitors of channels of the TASK subfamily. (b) Binding site of A1899 in a TASK-1 homology model [70, 71]. (c) Binding site of A293 in a TASK-1 homology model [72]. (d) Binding site of PK-THPP in a TASK-3 homology model [74]. (e) Illustration of the allosteric bupivacaine binding site in the side fenestration of a TASK-1 homology model [79]. (f) Binding site of BAY1000493 in TASK-1 determined by co-crystallization (PDB ID: 6RV3) [7]. (g) Chemical formula of the most important activators of channels of the TASK subfamily. (b-f) Potassium ions are represented by black spheres. Red arrows indicate the positions of the respective drugs in top view. *: homology model.

 

TASK-1 and TASK-3 activators

In contrast to blockers, TASK channels activators would be of therapeutic interest against pulmonary arterial hypertension (PAH), Birk-Barel mental retardation syndrome and some manifestations of pain [81-83]. However, only a few TASK-1 or TASK-3 channel activators are described so far (shown in Fig. 3g and Table 6). For TASK channels molecular modeling studies suggested an anesthetic binding pocket [84], including the HRE [41, 80] in the late M4 and M159 [85] in the late M3 segment [84]. Halogenated ether, alcohol, and alkane anesthetics were reported to be positioned near the side fenestrations of TASK-3 channels, in proximity to L239. However, mutating the pore facing L122 residue completely eliminated the activation by isoflurane [86]. The authors suggest that these effects are caused by altered channel gating by the L122 mutant that is in close proximity to L239 and the side fenestrations. This hypothesis is supported by mutations at an equivalent residue in TWIK-1 (Leu146) that activate TWIK-1 and by molecular dynamic studies which suggest that this region is important for pore hydration and/or the lipid access into the pore [87]. Alternatively, it could be also possible that volatile anesthetics are also bound to residues of the central cavity and act from the pore on the selectivity filter, similar as described for the ‘master key’ mechanism [48].

However, there are also examples of synthetic small molecule activators already: The guanylate cyclase stimulator riociguat, licensed for the treatment of PAH, enhances TASK-1 currents [82]. Albeit, this activation occurs after incubation of transiently transfected tsA201 cells with the compound and a direct channel modulation has not been demonstrated yet [82]. In addition, TASK-3 for instance is activated by flufenamic acid [83], terbinafine [88], CHET3 [81] or NPBA [89] (shown in Fig. 3g and Table 6). Interestingly, the TASK-3G236R mutation which causes Birk-Barel mental retardation conducts only very little currents [90]. This electrophysiological phenotype could be partially rescued by the TASK activators flufenamic acid [83] or terbinafine [88]. Moreover, Garcia et al. demonstrated that intrathecal pre-treatment with terbinafine, reduced the formalin-induced flinching and allodynia/hyperalgesia in rat, further supporting the putative future clinical relevance of TASK activators [91].

In terms of the binding sites, amino acid residues in the early M2 and late M3 segment, which are not conserved in TASK-1, were important for TASK-3 channel activation by NPBA [89]. In contrast for CHET3 a binding site in TASK-3 was found underneath the selectivity filter close to the M2 and M4 segments altering channel gating by affecting the selectivity filter conformation [81].

 

 

Outlook - K2P channel modulators in human diseases

 

As briefly discussed above, K2P channel modulators carry a huge therapeutic potential, in many diseases, as became evident from their involvement in inherited ‘channelopathies’ [47, 90, 92-94] and by the fact that ion channels are known to be very good ‘druggable’. However, despite the multitude of K2P channel modulators known by now, we clearly require more K2P modulating compounds, not only to further increase potency and selectivity to avoid toxicity and/or specific side effects, but also to have compounds with the right pharmacokinetics of Liberation, Absorption, Distribution, Metabolism and Excretion (LADME). Thus, given the development of further compounds from distinct chemical structural classes, presumably primarily possible by pharmaceutical industry, one can think of many future applications of K2P channel modulators in human diseases. TASK-1 activators might rescue PAH and be even beneficial in those cases of PAH in which the disease was not caused by a KCNK3 loss-of-function mutation as described by Ma et al. [94, 95]. TASK‑3 activators might be able to rescue some aspects of the Birk-Barel mental retardation [90], when applied early in juvenile development. TASK‑1 blockers are already under clinical investigations against OSA (DOCTOS Trial) and AF (SANDMAN Trial). Also, as a lesson learned from a KCNK17 mutation causing progressive cardiac conduction disorder (PCCD) [93], TASK-4 blockers might be advantageous for the treatment of specific forms of PCCD. Furthermore, TRESK modulators might be beneficial in migraine [92, 96] and novel potent TREK-1 modulators could be effective against pain without the classical side effects of opioids [46]. These putative future therapeutic applications are, as we think, a strong motivation to further study the molecular pharmacology of K2P channels.

 

 

Conclusion

 

K2P channels notoriously suffered from a poor pharmacologic profile which was a drawback for studies aiming to address the physiological role of these channels. However, recent advances in the understanding of the molecular pharmacology of K2P channels, provided an understanding about a large variety of complex mechanisms that can cause modulation of these potassium channels. While the gating and molecular pharmacology of TREK, TRESK and TASK channel subfamily members was the subject of many excellent studies, we eagerly anticipate new drugs and more mechanistic insights for the molecular modulation of channels of the TWIK, THIK and TALK subfamilies.

 

 

Acknowledgements

 

All authors contributed to the draft of the manuscript and to the design of the figures. The authors are grateful to all members of their laboratory and for useful discussion with all our collaboration partners.

 

Funding

A.K.K. is supported by the von-Behring-Röntgen Stiftung (67-0015). N.D. is supported by the Deutsche Forschungsgemeinschaft DFG (DE 1482/9-1). W.G. is supported by Fondecyt (1191133).

 

 

Disclosure Statement

 

The authors have no conflicts of interest to declare.

 

 

References

 

1 Ketchum KA, Joiner WJ, Sellers AJ, Kaczmarek LK, Goldstein SA: A new family of outwardly rectifying potassium channel proteins with two pore domains in tandem. Nature 1995;376:690-695.
https://doi.org/10.1038/376690a0

 

2 Lesage F, Reyes R, Fink M, Duprat F, Guillemare E, Lazdunski M: Dimerization of TWIK-1 K+ channel subunits via a disulfide bridge. EMBO J 1996;15:6400-6407.
https://doi.org/10.1002/j.1460-2075.1996.tb01031.x

 

3 Brohawn SG, Campbell EB, MacKinnon R: Physical mechanism for gating and mechanosensitivity of the human TRAAK K+ channel. Nature 2014;516:126-130.
https://doi.org/10.1038/nature14013

 

4 Brohawn SG, del Marmol J, MacKinnon R: Crystal structure of the human K2P TRAAK, a lipid- and mechano-sensitive K+ ion channel. Science 2012;335:436-441.
https://doi.org/10.1126/science.1213808

 

5 Dong YY, Pike AC, Mackenzie A, McClenaghan C, Aryal P, Dong L, et al.: K2P channel gating mechanisms revealed by structures of TREK-2 and a complex with Prozac. Science 2015;347:1256-1259.
https://doi.org/10.1126/science.1261512

 

6 Lolicato M, Arrigoni C, Mori T, Sekioka Y, Bryant C, Clark KA, et al.: K2P2.1 (TREK-1)-activator complexes reveal a cryptic selectivity filter binding site. Nature 2017;547:364-368.
https://doi.org/10.1038/nature22988

 

7 Rödström KEJ, Kiper AK, Zhang W, Rinné S, Pike ACW, Goldstein M, et al.: A lower X-gate in TASK channels traps inhibitors within the vestibule. Nature 2020;582:443-447.
https://doi.org/10.1038/s41586-020-2250-8

 

8 Miller AN, Long SB: Crystal structure of the human two-pore domain potassium channel K2P1. Science 2012;335:432-436.
https://doi.org/10.1126/science.1213274

 

9 Li B, Rietmeijer RA, Brohawn SG: Structural basis for pH gating of the two-pore domain K+ channel TASK2. Nature 2020;586:457-462.
https://doi.org/10.1038/s41586-020-2770-2

 

10 Chen TY, Miller C: Nonequilibrium gating and voltage dependence of the ClC-0 Cl- channel. J Gen Physiol 1996;108:237-250.
https://doi.org/10.1085/jgp.108.4.237

 

11 Schewe M, Nematian-Ardestani E, Sun H, Musinszki M, Cordeiro S, Bucci G, et al.: A Non-canonical Voltage-Sensing Mechanism Controls Gating in K2P K+ Channels. Cell 2016;164:937-949.
https://doi.org/10.1016/j.cell.2016.02.002

 

12 Baukrowitz T, Yellen G: Two functionally distinct subsites for the binding of internal blockers to the pore of voltage-activated K+ channels. Proc Natl Acad Sci USA 1996;93:13357-13361.
https://doi.org/10.1073/pnas.93.23.13357

 

13 Choi KL, Mossman C, Aube J, Yellen G: The internal quaternary ammonium receptor site of Shaker potassium channels. Neuron 1993;10:533-541.
https://doi.org/10.1016/0896-6273(93)90340-W

 

14 Hockerman GH, Johnson BD, Scheuer T, Catterall WA: Molecular determinants of high affinity phenylalkylamine block of L-type calcium channels. J Biol Chem 1995;270:22119-22122.
https://doi.org/10.1074/jbc.270.38.22119

 

15 Hockerman GH, Peterson BZ, Sharp E, Tanada TN, Scheuer T, Catterall WA: Construction of a high-affinity receptor site for dihydropyridine agonists and antagonists by single amino acid substitutions in a non-L-type Ca2+ channel. Proc Natl Acad Sci USA 1997;94:14906-14911.
https://doi.org/10.1073/pnas.94.26.14906

 

16 Mitcheson JS, Chen J, Lin M, Culberson C, Sanguinetti MC: A structural basis for drug-induced long QT syndrome. Proc Natl Acad Sci U S A 2000;97:12329-12333.
https://doi.org/10.1073/pnas.210244497

 

17 Peterson BZ, Johnson BD, Hockerman GH, Acheson M, Scheuer T, Catterall WA: Analysis of the dihydropyridine receptor site of L-type calcium channels by alanine-scanning mutagenesis. J Biol Chem 1997;272:18752-18758.
https://doi.org/10.1074/jbc.272.30.18752

 

18 Ragsdale DS, McPhee JC, Scheuer T, Catterall WA: Common molecular determinants of local anesthetic, antiarrhythmic, and anticonvulsant block of voltage-gated Na+ channels. Proc Natl Acad Sci USA 1996;93:9270-9275.
https://doi.org/10.1073/pnas.93.17.9270

 

19 Decher N, Pirard B, Bundis F, Peukert S, Baringhaus KH, Busch AE, et al.: Molecular basis for Kv1.5 channel block: conservation of drug binding sites among voltage-gated K+ channels. J Biol Chem 2004;279:394-400.
https://doi.org/10.1074/jbc.M307411200

 

20 Strutz-Seebohm N, Gutcher I, Decher N, Steinmeyer K, Lang F, Seebohm G: Comparison of potent Kv1.5 potassium channel inhibitors reveals the molecular basis for blocking kinetics and binding mode. Cell Physiol Biochem 2007;20:791-800.
https://doi.org/10.1159/000110439

 

21 Piechotta PL, Rapedius M, Stansfeld PJ, Bollepalli MK, Ehrlich G, Andres-Enguix I, et al.: The pore structure and gating mechanism of K2P channels. EMBO J 2011;30:3607-3619.
https://doi.org/10.1038/emboj.2011.268

 

22 Ashmole I, Goodwin PA, Stanfield PR: TASK-5, a novel member of the tandem pore K+ channel family. Pflugers Arch 2001;442:828-833.
https://doi.org/10.1007/s004240100620

 

23 Duprat F, Lesage F, Fink M, Reyes R, Heurteaux C, Lazdunski M: TASK, a human background K+ channel to sense external pH variations near physiological pH. EMBO J 1997;16:5464-5471.
https://doi.org/10.1093/emboj/16.17.5464

 

24 Kim D, Gnatenco C: TASK-5, a new member of the tandem-pore K+ channel family. Biochem Biophys Res Commun 2001;284:923-930.
https://doi.org/10.1006/bbrc.2001.5064

 

25 Kim Y, Bang H, Kim D: TASK-3, a new member of the tandem pore K+ channel family. J Biol Chem 2000;275:9340-9347.
https://doi.org/10.1074/jbc.275.13.9340

 

26 Musset B, Meuth SG, Liu GX, Derst C, Wegner S, Pape HC, et al.: Effects of divalent cations and spermine on the K+ channel TASK-3 and on the outward current in thalamic neurons. J Physiol 2006;572:639-657.
https://doi.org/10.1113/jphysiol.2006.106898

 

27 Rajan S, Wischmeyer E, Xin Liu G, Preisig-Müller R, Daut J, Karschin A, et al.: TASK-3, a novel tandem pore domain acid-sensitive K+ channel. An extracellular histiding as pH sensor. J Biol Chem 2000;275:16650-16657.
https://doi.org/10.1074/jbc.M000030200

 

28 Braun G, Lengyel M, Enyedi P, Czirjak G: Differential sensitivity of TREK-1, TREK-2 and TRAAK background potassium channels to the polycationic dye ruthenium red. Br J Pharmacol 2015;172:1728-1738.
https://doi.org/10.1111/bph.13019

 

29 Czirjak G, Enyedi P: Ruthenium red inhibits TASK-3 potassium channel by interconnecting glutamate 70 of the two subunits. Mol Pharmacol 2003;63:646-652.
https://doi.org/10.1124/mol.63.3.646

 

30 Lesage F, Guillemare E, Fink M, Duprat F, Lazdunski M, Romey G, et al.: TWIK-1, a ubiquitous human weakly inward rectifying K+ channel with a novel structure. EMBO J 1996;15:1004-1011.
https://doi.org/10.1002/j.1460-2075.1996.tb00437.x

 

31 Fink M, Duprat F, Lesage F, Reyes R, Romey G, Heurteaux C, et al.: Cloning, functional expression and brain localization of a novel unconventional outward rectifier K+ channel. EMBO J 1996;15:6854-6862.
https://doi.org/10.1002/j.1460-2075.1996.tb01077.x

 

32 Putzke C, Wemhöner K, Sachse FB, Rinné S, Schlichthörl G, Li XT, et al.: The acid-sensitive potassium channel TASK-1 in rat cardiac muscle. Cardiovasc Res 2007;75:59-68.
https://doi.org/10.1016/j.cardiores.2007.02.025

 

33 Charpentier F: Understanding the cardiac role of K2P channels: a new TASK for electrophysiologists. Cardiovasc Res 2007;75:5-6.
https://doi.org/10.1016/j.cardiores.2007.05.011

 

34 Kang D, Mariash E, Kim D: Functional expression of TRESK-2, a new member of the tandem-pore K+ channel family. J Biol Chem 2004;279:28063-28070.
https://doi.org/10.1074/jbc.M402940200

 

35 Sano Y, Inamura K, Miyake A, Mochizuki S, Kitada C, Yokoi H, et al.: A novel two-pore domain K+ channel, TRESK, is localized in the spinal cord. J Biol Chem 2003;278:27406-27412.
https://doi.org/10.1074/jbc.M206810200

 

36 Dobler T, Springauf A, Tovornik S, Weber M, Schmitt A, Sedlmeier R, et al.: TRESK two-pore-domain K+ channels constitute a significant component of background potassium currents in murine dorsal root ganglion neurones. J Physiol 2007;585:867-879.
https://doi.org/10.1113/jphysiol.2007.145649

 

37 Bruner JK, Zou B, Zhang H, Zhang Y, Schmidt K, Li M: Identification of novel small molecule modulators of K2P18.1 two-pore potassium channel. Eur J Pharmacol 2014;740:603-610.
https://doi.org/10.1016/j.ejphar.2014.06.021

 

38 Fink M, Lesage F, Duprat F, Heurteaux C, Reyes R, Fosset M, et al.: A neuronal two P domain K+ channel stimulated by arachidonic acid and polyunsaturated fatty acids. EMBO J 1998;17:3297-3308.
https://doi.org/10.1093/emboj/17.12.3297

 

39 Patel AJ, Honore E, Maingret F, Lesage F, Fink M, Duprat F, et al.: A mammalian two pore domain mechano-gated S-like K+ channel. EMBO J 1998;17:4283-4290.
https://doi.org/10.1093/emboj/17.15.4283

 

40 Kim Y, Bang H, Gnatenco C, Kim D: Synergistic interaction and the role of C-terminus in the activation of TRAAK K+ channels by pressure, free fatty acids and alkali. Pflugers Arch 2001;442:64-72.
https://doi.org/10.1007/s004240000496

 

41 Patel AJ, Honore E, Lesage F, Fink M, Romey G, Lazdunski M: Inhalational anesthetics activate two-pore-domain background K+ channels. Nat Neurosci 1999;2:422-426.
https://doi.org/10.1038/8084

 

42 Gruss M, Bushell TJ, Bright DP, Lieb WR, Mathie A, Franks NP: Two-pore-domain K+ channels are a novel target for the anesthetic gases xenon, nitrous oxide, and cyclopropane. Mol Pharmacol 2004;65:443-452.
https://doi.org/10.1124/mol.65.2.443

 

43 Duprat F, Lesage F, Patel AJ, Fink M, Romey G, Lazdunski M: The neuroprotective agent riluzole activates the two P domain K+ channels TREK-1 and TRAAK. Mol Pharmacol 2000;57:906-912.

 

44 Vivier D, Soussia IB, Rodrigues N, Lolignier S, Devilliers M, Chatelain FC, et al.: Development of the First Two-Pore Domain Potassium Channel TWIK-Related K+ Channel 1-Selective Agonist Possessing in Vivo Antinociceptive Activity. J Med Chem 2017;60:1076-1088.
https://doi.org/10.1021/acs.jmedchem.6b01285

 

45 Gada K, Plant LD: Two-pore domain potassium channels: emerging targets for novel analgesic drugs: IUPHAR Review 26. Br J Pharmacol 2019;176:256-266.
https://doi.org/10.1111/bph.14518

 

46 Devilliers M, Busserolles J, Lolignier S, Deval E, Pereira V, Alloui A, et al.: Activation of TREK-1 by morphine results in analgesia without adverse side effects. Nat Commun 2013;4:2941.
https://doi.org/10.1038/ncomms3941

 

47 Decher N, Ortiz-Bonnin B, Friedrich C, Schewe M, Kiper AK, Rinné S, et al.: Sodium permeable and "hypersensitive" TREK-1 channels cause ventricular tachycardia. EMBO Mol Med 2017;9:403-414.
https://doi.org/10.15252/emmm.201606690

 

48 Schewe M, Sun H, Mert U, Mackenzie A, Pike ACW, Schulz F, et al.: A pharmacological master key mechanism that unlocks the selectivity filter gate in K+ channels. Science 2019;363:875-880.
https://doi.org/10.1126/science.aav0569

 

49 Beltran L, Beltran M, Aguado A, Gisselmann G, Hatt H: 2-Aminoethoxydiphenyl borate activates the mechanically gated human KCNK channels KCNK2 (TREK-1), KCNK4 (TRAAK), and KCNK10 (TREK-2). Front Pharmacol 2013;4:63.
https://doi.org/10.3389/fphar.2013.00063

 

50 Zhuo RG, Liu XY, Zhang SZ, Wei XL, Zheng JQ, Xu JP, et al.: Insights into the stimulatory mechanism of 2-aminoethoxydiphenyl borate on TREK-2 potassium channel. Neuroscience 2015;300:85-93.
https://doi.org/10.1016/j.neuroscience.2015.05.012

 

51 Zhuo RG, Peng P, Liu XY, Yan HT, Xu JP, Zheng JQ, et al.: Allosteric coupling between proximal C-terminus and selectivity filter is facilitated by the movement of transmembrane segment 4 in TREK-2 channel. Sci Rep 2016;6:21248.
https://doi.org/10.1038/srep21248

 

52 Dadi PK, Vierra NC, Days E, Dickerson MT, Vinson PN, Weaver CD, et al.: Selective Small Molecule Activators of TREK-2 Channels Stimulate Dorsal Root Ganglion c-Fiber Nociceptor Two-Pore-Domain Potassium Channel Currents and Limit Calcium Influx. ACS Chem Neurosci 2017;8:558-568.
https://doi.org/10.1021/acschemneuro.6b00301

 

53 Mazella J, Petrault O, Lucas G, Deval E, Beraud-Dufour S, Gandin C, et al.: Spadin, a sortilin-derived peptide, targeting rodent TREK-1 channels: a new concept in the antidepressant drug design. PLoS Biol 2010;8:e1000355.
https://doi.org/10.1371/journal.pbio.1000355

 

54 Djillani A, Pietri M, Mazella J, Heurteaux C, Borsotto M: Fighting against depression with TREK-1 blockers: Past and future. A focus on spadin. Pharmacol Ther 2019;194:185-198.
https://doi.org/10.1016/j.pharmthera.2018.10.003

 

55 Ma R, Lewis A: Spadin Selectively Antagonizes Arachidonic Acid Activation of TREK-1 Channels. Front Pharmacol 2020;11:434.
https://doi.org/10.3389/fphar.2020.00434

 

56 McClenaghan C, Schewe M, Aryal P, Carpenter EP, Baukrowitz T, Tucker SJ: Polymodal activation of the TREK-2 K2P channel produces structurally distinct open states. J Gen Physiol 2016;147:497-505.
https://doi.org/10.1085/jgp.201611601

 

57 Kennard LE, Chumbley JR, Ranatunga KM, Armstrong SJ, Veale EL, Mathie A: Inhibition of the human two-pore domain potassium channel, TREK-1, by fluoxetine and its metabolite norfluoxetine. Br J Pharmacol 2005;144:821-829.
https://doi.org/10.1038/sj.bjp.0706068

 

58 Pope L, Lolicato M, Minor DL, Jr.: Polynuclear Ruthenium Amines Inhibit K2P Channels via a "Finger in the Dam" Mechanism. Cell chemical biology 2020;27:511-524 e514.
https://doi.org/10.1016/j.chembiol.2020.01.011

 

59 Luo Q, Chen L, Cheng X, Ma Y, Li X, Zhang B, et al.: An allosteric ligand-binding site in the extracellular cap of K2P channels. Nat Commun 2017;8:378.
https://doi.org/10.1038/s41467-017-00499-3

 

60 Schmidt C, Wiedmann F, Voigt N, Zhou XB, Heijman J, Lang S, et al.: Upregulation of K2P3.1 K+ Current Causes Action Potential Shortening in Patients With Chronic Atrial Fibrillation. Circulation 2015;132:82-92.
https://doi.org/10.1161/CIRCULATIONAHA.114.012657

 

61 Barth AS, Merk S, Arnoldi E, Zwermann L, Kloos P, Gebauer M, et al.: Functional profiling of human atrial and ventricular gene expression. Pflugers Arch 2005;450:201-208.
https://doi.org/10.1007/s00424-005-1404-8

 

62 Liang B, Soka M, Christensen AH, Olesen MS, Larsen AP, Knop FK, et al.: Genetic variation in the two-pore domain potassium channel, TASK-1, may contribute to an atrial substrate for arrhythmogenesis. J Mol Cell Cardiol 2014;67:69-76.
https://doi.org/10.1016/j.yjmcc.2013.12.014

 

63 Schmidt C, Wiedmann F, Beyersdorf C, Zhao Z, El-Battrawy I, Lan H, et al.: Genetic Ablation of TASK-1 (Tandem of P Domains in a Weak Inward Rectifying K+ Channel-Related Acid-Sensitive K+ Channel-1) (K2P3.1) K+ Channels Suppresses Atrial Fibrillation and Prevents Electrical Remodeling. Circulation Arrhythmia and electrophysiology 2019;12:e007465.
https://doi.org/10.1161/CIRCEP.119.007465

 

64 Ortega-Saenz P, Levitsky KL, Marcos-Almaraz MT, Bonilla-Henao V, Pascual A, Lopez-Barneo J: Carotid body chemosensory responses in mice deficient of TASK channels. J Gen Physiol 2010;135:379-392.
https://doi.org/10.1085/jgp.200910302

 

65 Kiper AK, Rinné S, Rolfes C, Ramirez D, Seebohm G, Netter MF, et al.: Kv1.5 blockers preferentially inhibit TASK-1 channels: TASK-1 as a target against atrial fibrillation and obstructive sleep apnea? Pflugers Arch 2015;467:1081-1090.
https://doi.org/10.1007/s00424-014-1665-1

 

66 Gurges P, Liu H, Horner RL: Modulation of TASK-1/3 Channels at the Hypoglossal Motoneuron Pool and Effects on Tongue Motor Output and Responses to Excitatory Inputs In-Vivo: Implications for Strategies for Obstructive Sleep Apnea Pharmacotherapy. Sleep 2020; DOI:10.1093/sleep/zsaa144.
https://doi.org/10.1093/sleep/zsaa144

 

67 Cotten JF: TASK-1 (KCNK3) and TASK-3 (KCNK9) tandem pore potassium channel antagonists stimulate breathing in isoflurane-anesthetized rats. Anesth Analg 2013;116:810-816.
https://doi.org/10.1213/ANE.0b013e318284469d

 

68 Decher N, Wemhöner K, Rinné S, Netter MF, Zuzarte M, Aller MI, et al.: Knock-out of the potassium channel TASK-1 leads to a prolonged QT interval and a disturbed QRS complex. Cell Physiol Biochem 2011;28:77-86.
https://doi.org/10.1159/000331715

 

69 Limberg SH, Netter MF, Rolfes C, Rinné S, Schlichthörl G, Zuzarte M, et al.: TASK-1 channels may modulate action potential duration of human atrial cardiomyocytes. Cell Physiol Biochem 2011;28:613-624.
https://doi.org/10.1159/000335757

 

70 Streit AK, Netter MF, Kempf F, Walecki M, Rinné S, Bollepalli MK, et al.: A specific two-pore domain potassium channel blocker defines the structure of the TASK-1 open pore. J Biol Chem 2011;286:13977-13984.
https://doi.org/10.1074/jbc.M111.227884

 

71 Ramirez D, Arevalo B, Martinez G, Rinné S, Sepulveda FV, Decher N, et al.: Side Fenestrations Provide an "Anchor" for a Stable Binding of A1899 to the Pore of TASK-1 Potassium Channels. Mol Pharm 2017;14:2197-2208.
https://doi.org/10.1021/acs.molpharmaceut.7b00005

 

72 Wiedmann F, Kiper AK, Bedoya M, Ratte A, Rinné S, Kraft M, et al.: Identification of the A293 (AVE1231) Binding Site in the Cardiac Two-Pore-Domain Potassium Channel TASK-1: a Common Low Affinity Antiarrhythmic Drug Binding Site. Cell Physiol Biochem 2019;52:1223-1235.
https://doi.org/10.33594/000000083

 

73 Chokshi RH, Larsen AT, Bhayana B, Cotten JF: Breathing Stimulant Compounds Inhibit TASK-3 Potassium Channel Function Likely by Binding at a Common Site in the Channel Pore. Mol Pharmacol 2015;88:926-934.
https://doi.org/10.1124/mol.115.100107

 

74 Ramirez D, Bedoya M, Kiper AK, Rinné S, Morales-Navarro S, Hernandez-Rodriguez EW, et al.: Structure/Activity Analysis of TASK-3 Channel Antagonists Based on a 5,6,7,8 tetrahydropyrido[4,3-d]pyrimidine. Int J Mol Sci 2019;20:2252.
https://doi.org/10.3390/ijms20092252

 

75 Coburn CA, Luo Y, Cui M, Wang J, Soll R, Dong J, et al.: Discovery of a pharmacologically active antagonist of the two-pore-domain potassium channel K2P9.1 (TASK-3). ChemMedChem 2012;7:123-133.
https://doi.org/10.1002/cmdc.201100351

 

76 Flaherty DP, Simpson DS, Miller M, Maki BE, Zou B, Shi J, et al.: Potent and selective inhibitors of the TASK-1 potassium channel through chemical optimization of a bis-amide scaffold. Bioorg Med Chem Lett 2014;24:3968-3973.
https://doi.org/10.1016/j.bmcl.2014.06.032

 

77 Cotten JF, Keshavaprasad B, Laster MJ, Eger EI, 2nd, Yost CS: The ventilatory stimulant doxapram inhibits TASK tandem pore (K2P) potassium channel function but does not affect minimum alveolar anesthetic concentration. Anesth Analg 2006;102:779-785.
https://doi.org/10.1213/01.ane.0000194289.34345.63

 

78 O'Donohoe PB, Huskens N, Turner PJ, Pandit JJ, Buckler KJ: A1899, PK-THPP, ML365, and Doxapram inhibit endogenous TASK channels and excite calcium signaling in carotid body type-1 cells. Physiol Rep 2018;6:e13876.
https://doi.org/10.14814/phy2.13876

 

79 Rinné S, Kiper AK, Vowinkel KS, Ramirez D, Schewe M, Bedoya M, et al.: The molecular basis for an allosteric inhibition of K+-flux gating in K2P channels. Elife 2019;8:e39476.
https://doi.org/10.7554/eLife.39476

 

80 Talley EM, Bayliss DA: Modulation of TASK-1 (Kcnk3) and TASK-3 (Kcnk9) potassium channels: volatile anesthetics and neurotransmitters share a molecular site of action. J Biol Chem 2002;277:17733-17742.
https://doi.org/10.1074/jbc.M200502200

 

81 Liao P, Qiu Y, Mo Y, Fu J, Song Z, Huang L, et al.: Selective activation of TWIK-related acid-sensitive K+ 3 subunit-containing channels is analgesic in rodent models. Sci Transl Med 2019;11:eaaw8434.
https://doi.org/10.1126/scitranslmed.aaw8434

 

82 Cunningham KP, Holden RG, Escribano-Subias PM, Cogolludo A, Veale EL, Mathie A: Characterization and regulation of wild-type and mutant TASK-1 two pore domain potassium channels indicated in pulmonary arterial hypertension. J Physiol 2019;597:1087-1101.
https://doi.org/10.1113/JP277275

 

83 Veale EL, Hassan M, Walsh Y, Al-Moubarak E, Mathie A: Recovery of current through mutated TASK3 potassium channels underlying Birk Barel syndrome. Mol Pharmacol 2014;85:397-407.
https://doi.org/10.1124/mol.113.090530

 

84 Bertaccini EJ, Dickinson R, Trudell JR, Franks NP: Molecular modeling of a tandem two pore domain potassium channel reveals a putative binding site for general anesthetics. ACS Chem Neurosci 2014;5:1246-1252.
https://doi.org/10.1021/cn500172e

 

85 Conway KE, Cotten JF: Covalent modification of a volatile anesthetic regulatory site activates TASK-3 (KCNK9) tandem-pore potassium channels. Mol Pharmacol 2012;81:393-400.
https://doi.org/10.1124/mol.111.076281

 

86 Luethy A, Boghosian JD, Srikantha R, Cotten JF: Halogenated Ether, Alcohol, and Alkane Anesthetics Activate TASK-3 Tandem Pore Potassium Channels Likely through a Common Mechanism. Mol Pharmacol 2017;91:620-629.
https://doi.org/10.1124/mol.117.108290

 

87 Aryal P, Abd-Wahab F, Bucci G, Sansom MS, Tucker SJ: A hydrophobic barrier deep within the inner pore of the TWIK-1 K2P potassium channel. Nat Commun 2014;5:4377.
https://doi.org/10.1038/ncomms5377

 

88 Wright PD, Veale EL, McCoull D, Tickle DC, Large JM, Ococks E, et al.: Terbinafine is a novel and selective activator of the two-pore domain potassium channel TASK3. Biochem Biophys Res Commun 2017;493:444-450.
https://doi.org/10.1016/j.bbrc.2017.09.002

 

89 Tian F, Qiu Y, Lan X, Li M, Yang H, Gao Z: A Small-Molecule Compound Selectively Activates K2P Channel TASK-3 by Acting at Two Distant Clusters of Residues. Mol Pharmacol 2019;96:26-35.
https://doi.org/10.1124/mol.118.115303

 

90 Barel O, Shalev SA, Ofir R, Cohen A, Zlotogora J, Shorer Z, et al.: Maternally inherited Birk Barel mental retardation dysmorphism syndrome caused by a mutation in the genomically imprinted potassium channel KCNK9. Am J Hum Genet 2008;83:193-199.
https://doi.org/10.1016/j.ajhg.2008.07.010

 

91 Garcia G, Noriega-Navarro R, Martinez-Rojas VA, Gutierrez-Lara EJ, Oviedo N, Murbartian J: Spinal TASK-1 and TASK-3 modulate inflammatory and neuropathic pain. Eur J Pharmacol 2019;862:172631.
https://doi.org/10.1016/j.ejphar.2019.172631

 

92 Lafreniere RG, Cader MZ, Poulin JF, Andres-Enguix I, Simoneau M, Gupta N, et al.: A dominant-negative mutation in the TRESK potassium channel is linked to familial migraine with aura. Nat Med 2010;16:1157-1160.
https://doi.org/10.1038/nm.2216

 

93 Friedrich C, Rinné S, Zumhagen S, Kiper AK, Silbernagel N, Netter MF, et al.: Gain-of-function mutation in TASK-4 channels and severe cardiac conduction disorder. EMBO Mol Med 2014;6:937-951.
https://doi.org/10.15252/emmm.201303783

 

94 Ma L, Roman-Campos D, Austin ED, Eyries M, Sampson KS, Soubrier F, et al.: A novel channelopathy in pulmonary arterial hypertension. N Engl J Med 2013;369:351-361.
https://doi.org/10.1056/NEJMoa1211097

 

95 Olschewski A, Veale EL, Nagy BM, Nagaraj C, Kwapiszewska G, Antigny F, et al.: TASK-1 (KCNK3) channels in the lung: from cell biology to clinical implications. Eur Respir J 2017;50:1700754.
https://doi.org/10.1183/13993003.00754-2017

 

96 Andres-Bilbe A, Castellanos A, Pujol-Coma A, Callejo G, Comes N, Gasull X: The Background K+ Channel TRESK in Sensory Physiology and Pain. Int J Mol Sci 2020;21:5206.
https://doi.org/10.3390/ijms21155206