Two-Pore Domain TASK Potassium Channels: Comparison
Please note this is a comparison between Version 2 by Vivi Li and Version 1 by Jin Liu.

TWIK-related acid-sensitive K+ (TASK) channels, including TASK-1, TASK-3, and TASK-5, are important members of the two-pore domain potassium (K2P) channel family. TASK-5 is not functionally expressed in the recombinant system. TASK channels are very sensitive to changes in extracellular pH and are active during all membrane potential periods. They are similar to other K2P channels in that they can create and use background-leaked potassium currents to stabilize resting membrane conductance and repolarize the action potential of excitable cells. TASK channels are expressed in both the nervous system and peripheral tissues, including excitable and non-excitable cells, and are widely engaged in pathophysiological phenomena, such as respiratory stimulation, pulmonary hypertension, arrhythmia, aldosterone secretion, cancers, anesthesia, neurological disorders, glucose homeostasis, and visual sensitivity. Therefore, they are important targets for innovative drug development. 

  • TASK channels
  • biophysical properties
  • gating profiles
  • biological roles
  • targeted compounds

1. Introduction

Ion channels play essential roles in a variety of cellular functions and are involved in the pathophysiology of a wide range of diseases. Currently, mammalian potassium channels are the most numerous, broadly distributed, and complicated class of ion channels that have been identified. They are classified into three main categories based on the topology of the transmembrane regions: voltage-gated potassium (Kv) channels, inwardly rectifying potassium (Kir) channels, and two-pore domain potassium (K2P) channels [1,2,3][1][2][3]. Among them, the family containing Kv channels is the largest. It includes approximately 40 genes, which are divided into 12 subfamilies (Kv1–Kv12), and the channels mainly consist of six transmembrane fragments and a pore structural domain [4]. Furthermore, the calcium-activated potassium channels (KCa) have either seven or six transmembrane segments and can link K+ dynamics to intracellular calcium signaling [5]. Kir channels are defined as channels that permit the inward flow of K+ under hyperpolarization but permit little or no outward K+ flow under depolarization. They can be grouped into seven subfamilies (Kir1–7), and these channels have two transmembrane fragments and a pore structural domain [6]. K2P channels have a dimeric structure instead of a tetrameric structure, and each subunit consists of four transmembrane fragments and two pore domains. These subunits are organized into six subgroups based on their various architectures and functions [7]. K2P channels can be regulated by diverse chemical, physical, and metabolic factors, such as extra- or intra-cellular pH, temperature, mechanical stimuli, G proteins, and signaling lipids [8]. They play an important role in stabilizing resting membrane conductance and repolarizing the action potential of excitable cells [9].
The tandem of pore domains in a weakly inward rectifying K+ channel 1 (TWIK-1) was the first K2P channel to be discovered and identified [10], and TWIK-related acid-sensitive K+ channel 1 (TASK-1) was the first cloned mammalian potassium channel that was found to create the characteristic background potassium currents [11]. TASK channels comprise TASK-1 (K2P3.1 or KCNK3), TASK-3 (K2P9.1 or KCNK9), and TASK-5 (K2P15.1 or KCNK15), which are composed of 394, 374, and 330 amino acids, respectively, in humans [11,12,13][11][12][13]. Among them, TASK-1 shares 58.9% amino acid sequence homology with TASK-3 and 51.4% homology with TASK-5, whereas TASK-3 shares 55.1% homology with TASK-5 [14[14][15][16],15,16], and TASK-5 was not found to be functionally expressed in the recombinant system [13]. TASK channels are important members of the K2P channel family and are widely involved in pathophysiological conditions. Therefore, an improved understanding of the biological mechanisms and functions of TASK channels will greatly assist in the future development of novel, potent, and selective target compounds.

2. The Structure, Localization, and Electrophysiological Properties of TASK Channels

2.1. Structure

As the closest homologue, the crystal structure of the human TASK-1 (hTASK-1) channel has currently been resolved at a resolution of 3.0 Å using X-ray crystallography [101][17]. Similar to other members of the K2P family, TASK channels have two-fold symmetry, with each subunit containing four transmembrane α-helices (M1–M4), an inner pore region located between M1 and M2, an inner pore region located between M3 and M4, two pore helices (PH1 and PH2) suspended like cradles in their respective pores, and two selectivity filters (SF1 and SF2) shaped by four pore loops that interact with each other to determine the ionic selectivity of the channels [102][18] (Figure 1). The structures of TASK-1 channel are from protein data bank (PDB) and are generated by PyMOL 2.5.
Figure 1. The structure and topology of the hTASK-1 channel. (A) Side view from the cell membrane. The hTASK-1 channel (PDB code 6RV2) consists of two homologous subunits, which are shown in blue and olive. The four transmembrane fragments (M1–M4) and two extracellular cap regions are labeled. (B) Top view of the hTASK-1 channel from the extracellular face, with the two pore helices (PH1 and PH2) highlighted in red and magenta in each subunit, respectively. The purple spheres are potassium ions. (C) Schematic topology of TASK channels: EC1 and EC2 are extracellular caps 1 and 2; P1 and P2 are pore regions 1 and 2; SF1 and SF2 are the selectivity filter of P1 and P2; OUT and IN represent the extracellular and intracellular surface, respectively.

2.2. Expression and Localization

The TASK-1 channel is abundantly expressed in the human central nervous system (CNS) and in peripheral tissues. In the human CNS, the highest expression levels of TASK-1 are in the cerebellum, thalamus, and pituitary gland, whereas the lowest expression level is in the corpus callosum. In human peripheral tissues, TASK-1 expression level is the highest in the pancreas, placenta, lungs and pulmonary arteries [17,18][19][20]. It is lower in the prostate, stomach, small intestine, and heart and lowest in the liver, spleen, skeletal muscles, and testes [17][19]. In rats and mice, the highest levels of TASK-1 mRNA in the CNS were observed in the cerebellum and somatic motoneurons, while the lowest levels were observed in the septum and striatum [16,110][16][21]. In humans, the TASK-3 channel is predominantly found in the CNS and is expressed in almost all regions of the brain, with the highest expression in the cerebellum and higher expression in the cerebral cortex, thalamus, nucleus accumbens, hippocampus, and hypothalamus. Its expression is lowest in the spinal cord, caudate nucleus, and corpus callosum. In addition, small amounts are expressed in peripheral tissues, such as the stomach, testes, skeletal muscles, uterus, kidney, spleen, pancreas, prostate, and small intestine. Very low levels are expressed in the heart, liver, and lungs [17][19]. In rat tissues, TASK-3 mRNA is widely expressed in the brain (with the highest expression levels found in the cerebellum), kidney, liver, lungs, colon, stomach, spleen, testes, and skeletal muscles. Its expression level is lowest in the heart and small intestine [71,110,111][21][22][23]. In humans, the TASK-5 channel is highly expressed in the pancreas and is also relatively abundantly expressed in the liver, kidney, lungs, ovary, testes, and heart [13].

2.3. Electrophysiological Properties

K2P channels are active across the entire physiological membrane potential range and generate background leaked potassium currents [112][24]. TASK channels are important members of the K2P channel family and have a current–voltage (I–V) relationship curve similar to the Goldman–Hodgkin–Katz equation at extra- and intra-cellularly different K+ concentrations, indicating that the generation of TASK channel currents correlates with K+ concentrations on both sides of the cell membrane. This also suggests that these channels show a high level of voltage-dependent gating despite the absence of a traditional voltage-activating domain because they have an ionic check valve located at the SF [113][25]. TASK channel currents are outwardly rectified and insensitive to potassium channel blockers, such as Ba2+, Cs+, TEA, and 4-AP [114][26]. TASK-1 and TASK-3 have different sensitivities to ruthenium red (RR or RuR), zinc, and anandamide. For instance, RR and zinc selectively block the TASK-3 channel but not the TASK-1 channel [115[27][28],116], whereas anandamide has a more potent blocking effect on TASK-1 than the TASK-3 channel [117][29]. In addition, TASK-1 and TASK-3 channel proteins can form functional heterodimers and are insensitive to mechanical forces [71,118][22][30]. The second messenger diacylglycerol can also inhibit TASK channels by activating G protein-coupled receptors [119,120][31][32]. TASK channels are sensitive to changes in extracellular pH but not intracellular pH, and can be blocked by external protons [121][33]. The histidine residue at position 98, which is near SF1, and the aspartic acid residue at position 204, which is near SF2, are the key sites for sensing extracellular pH variations [104,122][34][35]. Extracellular membrane acidification significantly inhibits TASK channels, whereas alkalization weakly activates them (Figure 2). Variations in extracellular pH also affect the ion selectivity of TASK channels, with TASK-1, TASK-3, and TWIK-1 channels becoming less permeable to potassium ions and more permeable to sodium ions when the extracellular solution becomes acidified [123,124,125][36][37][38].
Figure 2. Representative illustration of the current−voltage (I−V) relationship for TASK channels under different extracellular pH conditions. TASK channels are sensitive to changes in extracellular pH; they are significantly inhibited by acids and are weakly activated by alkalis.

3. The Biological Roles of TASK Channels

3.1. Breathing Rhythm

TASK channels are sensitive to extracellular pH and are expressed in multiple chemosensory nuclei or carotid bodies, which can sense PCO2 shifts in the early stages of acidosis [19,20,21][39][40][41]. The respiratory stimulant doxapram inhibits the activities of human TASK-1 and TASK-3 channels with effective half doses of 4 and 2.5 µM, respectively, suggesting that TASK channels are involved in regulating breathing rhythm [22][42]. Pharmacological and single-channel experiments revealed that TASK-1, TASK-3, and TASK-1/TASK-3 heterodimers play critical roles in the chemosensation of the carotid body and are major sensors of hypoxia and metabolic acidosis [23][43]. In neural recordings from electrodes implanted at the carotid body/sinus from TASK-1−/− mice, a significant reduction (49% and 68%, respectively) in chemoafferent cell firing induced by hypoxia (10% O2) and hypercapnia (3–6% CO2) was found along with a blunted ventilatory response to hypoxia. No changes in TASK-3−/− mice were observed under the same conditions, suggesting that TASK-1 in particular plays a key role in peripheral chemosensing. However, the TASK-3 channel can mediate hypoxic depolarization of normal glomus cells [24][44]. Studies using TASK-1 channel knockout (KO) mice also showed that pH sensitivity in serotonergic raphe neurons was abolished but was maintained in the retrotrapezoid nucleus (RTN). Because the RTN is central to the normal ventilatory response to CO2, this indicates that TASK-1 channels are not involved in the regulation of central respiratory chemosensitivity [25][45]. Alkaline-sensitive TASK-2 channels are widely distributed in Phox2b-expressing neurons in the brainstem RTN, which are the principal central respiratory chemoreceptors. TASK-2 KO mice displayed a diminished ventilatory reaction to CO2/H+ in vivo, indicating that TASK-2 channels play a major regulatory role in central respiratory chemoreception [1,20,137,146][1][40][46][47].

3.2. Pulmonary Artery Hypertension

Hypoxic pulmonary vasoconstriction (HPV) is an autoregulatory mechanism of the pulmonary vessels in response to hypoxia, which can reduce the ratio of ventilation and blood flow in the hypoxic alveolar region to ensure its oxygen supply. Thus, it plays an overriding role in maintaining the local ventilation/blood flow ratio and a constant arterial partial pressure of oxygen. Potassium channels play an important role in pulmonary arterial smooth muscle cells (PASMCs), and reduced potassium channel activity increases their resistance to apoptosis, cell proliferation, and vascular constriction, leading to vascular remodeling [147,148,149][48][49][50]. Hypoxia inhibits certain potassium channel activities, leading to cell membrane depolarization, which activates voltage-dependent calcium channels. This triggers inward extracellular Ca2+ flow and the contraction of pulmonary vascular smooth muscle, which ultimately causes increased pulmonary vascular resistance and the initiation of HPV. There are five main potassium channels in PASMCs: voltage-gated potassium (Kv) channels, Ca2+-activated potassium (KCa) channels, ATP-sensitive potassium (KATP) channels, inward rectifier potassium (Kir) channels, and K2P channels [150,151][51][52]. Human PASMCs express only TASK-1 potassium channels [26][53].

3.3. Cardiac Arrhythmia

TASK-1 potassium channels are atrial-specific, and protein blotting experiments have shown that the TASK-1 channel in the human heart is expressed at 14–16-fold higher levels in the atria than that in the ventricles [152,153,154][54][55][56]. The TASK-1 channel is known to be involved in the pathophysiology of cardiovascular diseases [37,38,39][57][58][59]. It has been found that TASK-1 KO mice have a significantly prolonged action potential duration [40][60] and that the use of the TASK-1 blocker A293 significantly prolonged the action potential duration in rat ventricular muscle. This suggests that TASK-1 channels are not only involved in maintaining the resting membrane potential of cardiac myocytes but also help in regulating the outward current during the plateau phase of the action potential [41][61]. Class III antiarrhythmic drugs prolong action potentials and inhibit atrial and ventricular arrhythmias. Among these drugs, amiodarone inhibits open and closed TASK-1 channels in a concentration-dependent manner, with an IC50 value of 0.4 μM in oocytes. This is a possible mechanism for the efficacy of these agents [42,43][62][63].

3.4. Aldosterone Secretion

It has been found that TASK−/− channel KO mice exhibit non-tumorigenic primary hyperaldosteronism with overproduction of autonomous aldosterone that was neither suppressed by high dietary sodium intake nor corrected using the angiotensin II receptor 1 antagonist candesartan. These results suggested that TASK channels are possible therapeutic targets for primary hyperaldosteronism [52][64]. Furthermore, statistical analyses indicated that KCNK3 (TASK-1) variants are associated with hyperaldosteronism and hypertension [53][65]. Another study reported that deleting TASK-1 and TASK-3 channels in the zona glomerulosa of the adrenal tissue causes aldosterone-driven angiotensin II-independent hypertension [54][66] Several other studies also indicated that TASK channels participate in the modulation of aldosterone secretion [55,56,57][67][68][69].

3.5. Pain

Tissue acidification can cause acute and chronic pain, and TASK channels are very sensitive to extracellular acidification. For example, exposing the channel to an extracellular pH of 6.0 can reduce its current by 96% [71,158][22][70]. Acid inhibition of TASK channels may be an important mechanism for the sustained depolarization of cells due to tissue acidification. This suggests that they could also be potential targets for tissue acidification-induced nociceptive transmission [159,160,161,162][71][72][73][74]. Several studies have shown that TASK channels are closely associated with pain [58,59,72][75][76][77].
Marsh et al. [60][78] found that the mRNA levels of TASK channels in the spinal dorsal root ganglion (DRG) were correlated with spontaneous foot lifting after intradermal complete Freund’s adjuvant injections, indicating that TASK channels are associated with inflammation-induced pain. The activity of TASK channels is also altered in pathological models of nerve injury. The mRNA expression of TASK-3 and TWIK-1 channels was found to be downregulated in L4-L5 DRG ipsilateral to the neuropathic lesion one week after spared nerve injury surgery, whereas TASK-1 channel expression remained unchanged [73][79]. In addition, it has been reported that the TASK-3 channel co-locates with the transient receptor potential cation channel subfamily M member 8 (TRPM8) in sensory neurons. TRPM8 is a cold- and menthol-activated ion channel that participates in thermosensation [163][80]. Liao et al. [74][81] designed a peripherally acting selective agonist, CHET3, for the TASK-3 channel and used it to show that the TASK-3 channel is an intrinsic regulator of pain sensation. They proved that CHET3 could attenuate thermal hyperalgesia and mechanical allodynia in different rodent pain models. They also predicted that CHET3 would bind under the SF and close to the M2 and M4 regions, which would alter the gating activity of the channel by affecting the SF conformation. García et al. [61][82] found that spinal TASK-1 and TASK-3 channels are involved in the regulation of inflammatory and neuropathic pain and that intrathecal pretreatment with the activator terbinafine could reduce formalin-induced flinching and nociceptive hyperalgesia in rats with neuropathy.

3.6. Anesthetics

3.6.1. Volatile Anesthetics

TASK channels are expressed in CNS sites relevant to anesthetic actions and are the targets of many general anesthetics. Volatile anesthetics were the first group of TASK channel modulators to be extensively studied. Volatile anesthetics, such as halothane and isoflurane, activate TASK channels at clinically relevant concentrations to hyperpolarize cell membranes and reduce excitability [62][83]. This is especially true for brainstem motor and thalamocortical neurons and helps explain the loss of consciousness and motor ability that general anesthetics induce [164,165][84][85] It has been reported that the hyperpolarizing effects of halothane and isoflurane were reduced in TASK-1/TASK-3 KO mice, and the corresponding anesthetic effects of sedation, hypnosis, and immobilization were also diminished [64][86]. However, it has also been reported that high concentrations of isoflurane (0.8 mM) activate the TASK-3 channel and the heterodimeric TASK-1/TASK-3 channel, but inhibit the TASK-1 channel. This suggested that different concentrations of volatile anesthetics can regulate the activities of TASK channels in both directions [65][87] Studies have shown the presence of a binding region for volatile anesthetics in the structures of TASK channels, which includes the residue M159 at the cytoplasmic terminus of the M3 segment [66][88] and, in particular, residues 243 to 248 at the beginning of the cytoplasmic C-terminus, a region also known as the halothane-responsive region [63,101][17][89]. In addition, it has been shown that halothane and isoflurane can inhibit calcium current in smooth muscle cells of the coronary arteries to directly dilate coronary arteries and increase coronary flow during anesthesia [166][90].

3.7. Cancers

The TASK-3 channel is highly expressed in breast cancer, ovarian carcinoma, and melanoma cells, and it has been suggested that it promotes tumor growth and proliferation [75,76,77,78,79,80,81,82][91][92][93][94][95][96][97][98]. Thus far, only the TASK-3 channel in the K2P family has been identified to be expressed in the inner mitochondrial membrane, and silencing it induces apoptosis of human melanoma cells [83,84][99][100]. In addition, a study using a short hairpin RNA (shRNA)-mediated knockdown of the TASK-3 channel confirmed that the TASK-3 channel is involved in migration and cell survival in gastric cancer. Thus, it represents a potential therapeutic target for gastric cancer treatment [85][101]. Moreover, the addition of Y4, a monoclonal antibody against the TASK-3 channel extracellular domain, to KCNK9-expressing carcinoma cells can inhibit tumor growth and metastasis [86][102]. These confirmed that the TASK-3 channel is a promising target for the treatment of malignancies that express KCNK9 [87][103].

3.8. Neurological Activities and/or Disorders

3.8.1. Sleep

Synchronized burst firing and tonic action potential generation in the thalamocortical system of the brain mainly occur during sleep and waking states, respectively [168][104]. The switch between these two firing modes is critically modulated by the bidirectional regulation of TASK and TWIK-related K+ channels (TREK) in thalamic relay neurons.
It has been found that the genetic KO of TASK-1 did not change sleep/wake times in mice [169][105]. One study analyzed the locomotor activity and circadian rhythms of TASK-3 KO mice and found that, when compared to wild-type litter controls, both had normal circadian rhythms. However, TASK-3 KO mice had significantly increased nocturnal activity (38%), suggesting that the TASK-3 channel plays an important role in the regulation of sleep [88][106]. Another study based on continuous electroencephalogram and electromyogram recordings of TASK-3 KO mice revealed that they had a slower transition from wakefulness to sleep and more fragmented sleep episodes and rapid eye movement (REM) theta wave (4–9 Hz) oscillations during sleep [89,90][107][108].

3.9. Other Roles

Several studies have shown that the TASK-1 channels found in the pancreas participate in the regulation of glucose homeostasis [69,70][109][110]. Using the highly specific “RNAscope” method of in situ mRNA hybridization in combination with pharmacological antagonists and genetic deletion tools, Wen et al. [100][111] reported the first evidence that the TASK-3 channel is abundantly expressed in retinal ganglion cells and plays a critical role in visual processing in the retina.

References

  1. Sepúlveda, F.V.; Pablo Cid, L.; Teulon, J.; Niemeyer, M.I. Molecular aspects of structure, gating, and physiology of pH-sensitive background K2P and Kir K+-transport channels. Physiol. Rev. 2015, 95, 179–217.
  2. Kuang, Q.; Purhonen, P.; Hebert, H. Structure of potassium channels. Cell. Mol. Life Sci. CMLS 2015, 72, 3677–3693.
  3. González, C.; Baez-Nieto, D.; Valencia, I.; Oyarzún, I.; Rojas, P.; Naranjo, D.; Latorre, R. K+ channels: Function-structural overview. Compr. Physiol. 2012, 2, 2087–2149.
  4. Kim, D.M.; Nimigean, C.M. Voltage-Gated Potassium Channels: A Structural Examination of Selectivity and Gating. Cold Spring Harb. Perspect. Biol. 2016, 8, a029231.
  5. Trombetta-Lima, M.; Krabbendam, I.E.; Dolga, A.M. Calcium-activated potassium channels: Implications for aging and age-related neurodegeneration. Int. J. Biochem. Cell Biol. 2020, 123, 105748.
  6. Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their structure, function, and physiological roles. Physiol. Rev. 2010, 90, 291–366.
  7. Lesage, F.; Lazdunski, M. Molecular and functional properties of two-pore-domain potassium channels. Am. J. Physiol. Renal Physiol. 2000, 279, F793–F801.
  8. Goldstein, S.A.; Bockenhauer, D.; O’Kelly, I.; Zilberberg, N. Potassium leak channels and the KCNK family of two-P-domain subunits. Nat. Rev. Neurosci. 2001, 2, 175–184.
  9. Huang, L.; Xu, G.; Jiang, R.; Luo, Y.; Zuo, Y.; Liu, J. Development of Non-opioid Analgesics Targeting Two-pore Domain Potassium Channels. Curr. Neuropharmacol. 2022, 20, 16–26.
  10. Lesage, F.; Guillemare, E.; Fink, M.; Duprat, F.; Lazdunski, M.; Romey, G.; Barhanin, J. TWIK-1, a ubiquitous human weakly inward rectifying K+ channel with a novel structure. EMBO J. 1996, 15, 1004–1011.
  11. 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.
  12. Rajan, S.; Wischmeyer, E.; Xin Liu, G.; Preisig-Müller, R.; Daut, J.; Karschin, A.; Derst, C. TASK-3, a novel tandem pore domain acid-sensitive K+ channel. An extracellular histiding as pH sensor. J. Biol. Chem. 2000, 275, 16650–16657.
  13. Ashmole, I.; Goodwin, P.A.; Stanfield, P.R. TASK-5, a novel member of the tandem pore K+ channel family. Pflugers Arch. 2001, 442, 828–833.
  14. Brohawn, S.G.; del Mármol, J.; MacKinnon, R. Crystal structure of the human K2P TRAAK, a lipid- and mechano-sensitive K+ ion channel. Science 2012, 335, 436–441.
  15. Cotten, J.F. TASK-1 (KCNK3) and TASK-3 (KCNK9) tandem pore potassium channel antagonists stimulate breathing in isoflurane-anesthetized rats. Anesth. Analg. 2013, 116, 810–816.
  16. Talley, E.M.; Lei, Q.; Sirois, J.E.; Bayliss, D.A. TASK-1, a two-pore domain K+ channel, is modulated by multiple neurotransmitters in motoneurons. Neuron 2000, 25, 399–410.
  17. Rödström, K.E.J.; Kiper, A.K.; Zhang, W.; Rinné, S.; Pike, A.C.W.; Goldstein, M.; Conrad, L.J.; Delbeck, M.; Hahn, M.G.; Meier, H.; et al. A lower X-gate in TASK channels traps inhibitors within the vestibule. Nature 2020, 582, 443–447.
  18. Enyedi, P.; Czirják, G. Molecular background of leak K+ currents: Two-pore domain potassium channels. Physiol. Rev. 2010, 90, 559–605.
  19. Medhurst, A.D.; Rennie, G.; Chapman, C.G.; Meadows, H.; Duckworth, M.D.; Kelsell, R.E.; Gloger, I.I.; Pangalos, M.N. Distribution analysis of human two pore domain potassium channels in tissues of the central nervous system and periphery. Brain Res. Mol. Brain Res. 2001, 86, 101–114.
  20. Olschewski, A.; Veale, E.L.; Nagy, B.M.; Nagaraj, C.; Kwapiszewska, G.; Antigny, F.; Lambert, M.; Humbert, M.; Czirják, G.; Enyedi, P.; et al. TASK-1 (KCNK3) channels in the lung: From cell biology to clinical implications. Eur. Respir. J. 2017, 50, 1700754.
  21. Talley, E.M.; Solorzano, G.; Lei, Q.; Kim, D.; Bayliss, D.A. Cns distribution of members of the two-pore-domain (KCNK) potassium channel family. J. Neurosci. 2001, 21, 7491–7505.
  22. 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.
  23. Rusznák, Z.; Pocsai, K.; Kovács, I.; Pór, A.; Pál, B.; Bíró, T.; Szücs, G. Differential distribution of TASK-1, TASK-2 and TASK-3 immunoreactivities in the rat and human cerebellum. Cell. Mol. Life Sci. 2004, 61, 1532–1542.
  24. Zilberberg, N.; Ilan, N.; Goldstein, S.A. KCNKØ: Opening and closing the 2-P-domain potassium leak channel entails “C-type” gating of the outer pore. Neuron 2001, 32, 635–648.
  25. Schewe, M.; Nematian-Ardestani, E.; Sun, H.; Musinszki, M.; Cordeiro, S.; Bucci, G.; de Groot, B.L.; Tucker, S.J.; Rapedius, M.; Baukrowitz, T. A Non-canonical Voltage-Sensing Mechanism Controls Gating in K2P K+ Channels. Cell 2016, 164, 937–949.
  26. O’Connell, A.D.; Morton, M.J.; Sivaprasadarao, A.; Hunter, M. Selectivity and interactions of Ba2+ and Cs+ with wild-type and mutant TASK1 K+ channels expressed in Xenopus oocytes. J. Physiol. 2005, 562 Pt 3, 687–696.
  27. Wang, X.; Guan, R.; Zhao, X.; Zhu, D.; Song, N.; Shen, L. TASK1 and TASK3 Are Coexpressed With ASIC1 in the Ventrolateral Medulla and Contribute to Central Chemoreception in Rats. Front. Cell. Neurosci. 2018, 12, 285.
  28. Clarke, C.E.; Veale, E.L.; Green, P.J.; Meadows, H.J.; Mathie, A. Selective block of the human 2-P domain potassium channel, TASK-3, and the native leak potassium current, IKSO, by zinc. J. Physiol. 2004, 560 Pt 1, 51–62.
  29. Maingret, F.; Patel, A.J.; Lazdunski, M.; Honoré, E. The endocannabinoid anandamide is a direct and selective blocker of the background K+ channel TASK-1. EMBO J. 2001, 20, 47–54.
  30. Czirják, G.; Enyedi, P. Formation of functional heterodimers between the TASK-1 and TASK-3 two-pore domain potassium channel subunits. J. Biol. Chem. 2002, 277, 5426–5432.
  31. Wilke, B.U.; Lindner, M.; Greifenberg, L.; Albus, A.; Kronimus, Y.; Bünemann, M.; Leitner, M.G.; Oliver, D. Diacylglycerol mediates regulation of TASK potassium channels by Gq-coupled receptors. Nat. Commun. 2014, 5, 5540.
  32. Bista, P.; Pawlowski, M.; Cerina, M.; Ehling, P.; Leist, M.; Meuth, P.; Aissaoui, A.; Borsotto, M.; Heurteaux, C.; Decher, N.; et al. Differential phospholipase C-dependent modulation of TASK and TREK two-pore domain K+ channels in rat thalamocortical relay neurons. J. Physiol. 2015, 593, 127–144.
  33. Lesage, F.; Barhanin, J. Molecular physiology of pH-sensitive background K2P channels. Physiology (Bethesda) 2011, 26, 424–437.
  34. González, W.; Zúñiga, L.; Cid, L.P.; Arévalo, B.; Niemeyer, M.I.; Sepúlveda, F.V. An extracellular ion pathway plays a central role in the cooperative gating of a K2P K+ channel by extracellular pH. J. Biol. Chem. 2013, 288, 5984–5991.
  35. Yuill, K.; Ashmole, I.; Stanfield, P.R. The selectivity filter of the tandem pore potassium channel TASK-1 and its pH-sensitivity and ionic selectivity. Pflugers Arch. 2004, 448, 63–69.
  36. Chen, H.; Chatelain, F.C.; Lesage, F. Altered and dynamic ion selectivity of K+ channels in cell development and excitability. Trends Pharmacol. Sci. 2014, 35, 461–469.
  37. Ma, L.; Zhang, X.; Zhou, M.; Chen, H. Acid-sensitive TWIK and TASK two-pore domain potassium channels change ion selectivity and become permeable to sodium in extracellular acidification. J. Biol. Chem. 2012, 287, 37145–37153.
  38. Chatelain, F.C.; Bichet, D.; Douguet, D.; Feliciangeli, S.; Bendahhou, S.; Reichold, M.; Warth, R.; Barhanin, J.; Lesage, F. TWIK1, a unique background channel with variable ion selectivity. Proc. Natl. Acad. Sci. USA 2012, 109, 5499–5504.
  39. Koizumi, H.; Smerin, S.E.; Yamanishi, T.; Moorjani, B.R.; Zhang, R.; Smith, J.C. TASK channels contribute to the K+-dominated leak current regulating respiratory rhythm generation in vitro. J. Neurosci. 2010, 30, 4273–4284.
  40. Bayliss, D.A.; Barhanin, J.; Gestreau, C.; Guyenet, P.G. The role of pH-sensitive TASK channels in central respiratory chemoreception. Pflugers Arch. 2015, 467, 917–929.
  41. Buckler, K.J. TASK channels in arterial chemoreceptors and their role in oxygen and acid sensing. Pflügers Arch.-Eur. J. Physiol. 2015, 467, 1013–1025.
  42. Cunningham, K.P.; MacIntyre, D.E.; Mathie, A.; Veale, E.L. Effects of the ventilatory stimulant, doxapram on human TASK-3 (KCNK9, K2P9.1) channels and TASK-1 (KCNK3, K2P3.1) channels. Acta Physiol. 2020, 228, e13361.
  43. Kim, D.; Cavanaugh, E.J.; Kim, I.; Carroll, J.L. Heteromeric TASK-1/TASK-3 is the major oxygen-sensitive background K+ channel in rat carotid body glomus cells. J. Physiol. 2009, 587 Pt 12, 2963–2975.
  44. Trapp, S.; Aller, M.I.; Wisden, W.; Gourine, A.V. A role for TASK-1 (KCNK3) channels in the chemosensory control of breathing. J. Neurosci. 2008, 28, 8844–8850.
  45. Mulkey, D.K.; Talley, E.M.; Stornetta, R.L.; Siegel, A.R.; West, G.H.; Chen, X.; Sen, N.; Mistry, A.M.; Guyenet, P.G.; Bayliss, D.A. TASK channels determine pH sensitivity in select respiratory neurons but do not contribute to central respiratory chemosensitivity. J. Neurosci. 2007, 27, 14049–14058.
  46. Li, B.; Rietmeijer, R.A.; Brohawn, S.G. Structural basis for pH gating of the two-pore domain K+ channel TASK2. Nature 2020, 586, 457–462.
  47. Guyenet, P.G.; Bayliss, D.A.; Stornetta, R.L.; Ludwig, M.G.; Kumar, N.N.; Shi, Y.; Burke, P.G.; Kanbar, R.; Basting, T.M.; Holloway, B.B.; et al. Proton detection and breathing regulation by the retrotrapezoid nucleus. J. Physiol. 2016, 594, 1529–1551.
  48. Hayabuchi, Y. The Action of Smooth Muscle Cell Potassium Channels in the Pathology of Pulmonary Arterial Hypertension. Pediatr. Cardiol. 2017, 38, 1–14.
  49. Le Ribeuz, H.; Capuano, V.; Girerd, B.; Humbert, M.; Montani, D.; Antigny, F. Implication of Potassium Channels in the Pathophysiology of Pulmonary Arterial Hypertension. Biomolecules 2020, 10, 1261.
  50. Southgate, L.; Machado, R.D.; Gräf, S.; Morrell, N.W. Molecular genetic framework underlying pulmonary arterial hypertension. Nat. Rev. Cardiol. 2019, 17, 85–95.
  51. Dogan, M.F.; Yildiz, O.; Arslan, S.O.; Ulusoy, K.G. Potassium channels in vascular smooth muscle: A pathophysiological and pharmacological perspective. Fundam. Clin. Pharmacol. 2019, 33, 504–523.
  52. Tennant, B.P.; Cui, Y.; Tinker, A.; Clapp, L.H. Functional expression of inward rectifier potassium channels in cultured human pulmonary smooth muscle cells: Evidence for a major role of Kir2.4 subunits. J. Membr. Biol. 2006, 213, 19–29.
  53. Olschewski, A.; Li, Y.; Tang, B.; Hanze, J.; Eul, B.; Bohle, R.M.; Wilhelm, J.; Morty, R.E.; Brau, M.E.; Weir, E.K.; et al. Impact of TASK-1 in human pulmonary artery smooth muscle cells. Circ. Res. 2006, 98, 1072–1080.
  54. Wiedmann, F.; Frey, N.; Schmidt, C. Two-Pore-Domain Potassium (K2P-) Channels: Cardiac Expression Patterns and Disease-Specific Remodelling Processes. Cells 2021, 10, 2914.
  55. Rinné, S.; Kiper, A.K.; Schlichthörl, G.; Dittmann, S.; Netter, M.F.; Limberg, S.H.; Silbernagel, N.; Zuzarte, M.; Moosdorf, R.; Wulf, H.; et al. TASK-1 and TASK-3 may form heterodimers in human atrial cardiomyocytes. J. Mol. Cell. Cardiol. 2015, 81, 71–80.
  56. Staudacher, K.; Staudacher, I.; Ficker, E.; Seyler, C.; Gierten, J.; Kisselbach, J.; Rahm, A.K.; Trappe, K.; Schweizer, P.A.; Becker, R.; et al. Carvedilol targets human K2P 3.1 (TASK1) K+ leak channels. Br. J. Pharmacol. 2011, 163, 1099–1110.
  57. Besana, A.; Barbuti, A.; Tateyama, M.A.; Symes, A.J.; Robinson, R.B.; Feinmark, S.J. Activation of protein kinase C epsilon inhibits the two-pore domain K+ channel, TASK-1, inducing repolarization abnormalities in cardiac ventricular myocytes. J. Biol. Chem. 2004, 279, 33154–33160.
  58. Gurney, A.; Manoury, B. Two-pore potassium channels in the cardiovascular system. Eur. Biophys. J. 2009, 38, 305–318.
  59. Liang, B.; Soka, M.; Christensen, A.H.; Olesen, M.S.; Larsen, A.P.; Knop, F.K.; Wang, F.; Nielsen, J.B.; Andersen, M.N.; Humphreys, D.; 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.
  60. Donner, B.C.; Schullenberg, M.; Geduldig, N.; Hüning, A.; Mersmann, J.; Zacharowski, K.; Kovacevic, A.; Decking, U.; Aller, M.I.; Schmidt, K.G. Functional role of TASK-1 in the heart: Studies in TASK-1-deficient mice show prolonged cardiac repolarization and reduced heart rate variability. Basic Res. Cardiol. 2011, 106, 75–87.
  61. Putzke, C.; Wemhöner, K.; Sachse, F.B.; Rinné, S.; Schlichthörl, G.; Li, X.T.; Jaé, L.; Eckhardt, I.; Wischmeyer, E.; Wulf, H.; et al. The acid-sensitive potassium channel TASK-1 in rat cardiac muscle. Cardiovasc. Res. 2007, 75, 59–68.
  62. Gierten, J.; Ficker, E.; Bloehs, R.; Schweizer, P.A.; Zitron, E.; Scholz, E.; Karle, C.; Katus, H.A.; Thomas, D. The human cardiac K2P3.1 (TASK-1) potassium leak channel is a molecular target for the class III antiarrhythmic drug amiodarone. Naunyn Schmiedebergs Arch. Pharmacol. 2010, 381, 261–270.
  63. Kraft, M.; Büscher, A.; Wiedmann, F.; L’Hoste, Y.; Haefeli, W.E.; Frey, N.; Katus, H.A.; Schmidt, C. Current Drug Treatment Strategies for Atrial Fibrillation and TASK-1 Inhibition as an Emerging Novel Therapy Option. Front. Pharmacol. 2021, 12, 638445.
  64. Davies, L.A.; Hu, C.; Guagliardo, N.A.; Sen, N.; Chen, X.; Talley, E.M.; Carey, R.M.; Bayliss, D.A.; Barrett, P.Q. TASK channel deletion in mice causes primary hyperaldosteronism. Proc. Natl. Acad. Sci. USA 2008, 105, 2203–2208.
  65. Manichaikul, A.; Rich, S.S.; Allison, M.A.; Guagliardo, N.A.; Bayliss, D.A.; Carey, R.M.; Barrett, P.Q. KCNK3 Variants Are Associated With Hyperaldosteronism and Hypertension. Hypertension 2016, 68, 356–364.
  66. Guagliardo, N.A.; Yao, J.; Stipes, E.J.; Cechova, S.; Le, T.H.; Bayliss, D.A.; Breault, D.T.; Barrett, P.Q. Adrenal Tissue-Specific Deletion of TASK Channels Causes Aldosterone-Driven Angiotensin II-Independent Hypertension. Hypertension 2019, 73, 407–414.
  67. Penton, D.; Bandulik, S.; Schweda, F.; Haubs, S.; Tauber, P.; Reichold, M.; Cong, L.D.; El Wakil, A.; Budde, T.; Lesage, F.; et al. Task3 potassium channel gene invalidation causes low renin and salt-sensitive arterial hypertension. Endocrinology 2012, 153, 4740–4748.
  68. Chen, A.X.; Nishimoto, K.; Nanba, K.; Rainey, W.E. Potassium channels related to primary aldosteronism: Expression similarities and differences between human and rat adrenals. Mol. Cell. Endocrinol. 2015, 417, 141–148.
  69. Barrett, P.Q.; Guagliardo, N.A.; Bayliss, D.A. Ion Channel Function and Electrical Excitability in the Zona Glomerulosa: A Network Perspective on Aldosterone Regulation. Annu. Rev. Physiol. 2021, 83, 451–475.
  70. Meadows, H.J.; Randall, A.D. Functional characterisation of human TASK-3, an acid-sensitive two-pore domain potassium channel. Neuropharmacology 2001, 40, 551–559.
  71. Reeh, P.W.; Kress, M. Molecular physiology of proton transduction in nociceptors. Curr. Opin. Pharmacol. 2001, 1, 45–51.
  72. Waldmann, R.; Champigny, G.; Bassilana, F.; Heurteaux, C.; Lazdunski, M. A proton-gated cation channel involved in acid-sensing. Nature 1997, 386, 173–177.
  73. Plant, L.D. A Role for K2P Channels in the Operation of Somatosensory Nociceptors. Front. Mol. Neurosci. 2012, 5, 21.
  74. Sun, W.H.; Chen, C.C. Roles of Proton-Sensing Receptors in the Transition from Acute to Chronic Pain. J. Dent. Res. 2016, 95, 135–142.
  75. Li, X.Y.; Toyoda, H. Role of leak potassium channels in pain signaling. Brain Res. Bull. 2015, 119 Pt A, 73–79.
  76. Gada, K.; Plant, L.D. Two-pore domain potassium channels: Emerging targets for novel analgesic drugs: IUPHAR Review 26. Br. J. Pharmacol. 2019, 176, 256–266.
  77. Ren, W.J.; Ulrich, H.; Semyanov, A.; Illes, P.; Tang, Y. TASK-3: New Target for Pain-Relief. Neurosci. Bull. 2020, 36, 951–954.
  78. Marsh, B.; Acosta, C.; Djouhri, L.; Lawson, S.N. Leak K⁺ channel mRNAs in dorsal root ganglia: Relation to inflammation and spontaneous pain behaviour. Mol. Cell. Neurosci. 2012, 49, 375–386.
  79. Pollema-Mays, S.L.; Centeno, M.V.; Ashford, C.J.; Apkarian, A.V.; Martina, M. Expression of background potassium channels in rat DRG is cell-specific and down-regulated in a neuropathic pain model. Mol. Cell. Neurosci. 2013, 57, 1–9.
  80. Morenilla-Palao, C.; Luis, E.; Fernández-Peña, C.; Quintero, E.; Weaver, J.L.; Bayliss, D.A.; Viana, F. Ion channel profile of TRPM8 cold receptors reveals a role of TASK-3 potassium channels in thermosensation. Cell Rep. 2014, 8, 1571–1582.
  81. Liao, P.; Qiu, Y.; Mo, Y.; Fu, J.; Song, Z.; Huang, L.; Bai, S.; Wang, Y.; Zhu, J.J.; Tian, F.; 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.
  82. García, G.; Noriega-Navarro, R.; Martínez-Rojas, V.A.; Gutiérrez-Lara, E.J.; Oviedo, N.; Murbartián, J. Spinal TASK-1 and TASK-3 modulate inflammatory and neuropathic pain. Eur. J. Pharmacol. 2019, 862, 172631.
  83. Patel, A.J.; Honoré, E.; Lesage, F.; Fink, M.; Romey, G.; Lazdunski, M. Inhalational anesthetics activate two-pore-domain background K+ channels. Nat. Neurosci. 1999, 2, 422–426.
  84. Sirois, J.E.; Lynch, C., 3rd; Bayliss, D.A. Convergent and reciprocal modulation of a leak K+ current and I(h) by an inhalational anaesthetic and neurotransmitters in rat brainstem motoneurones. J. Physiol. 2002, 541 Pt 3, 717–729.
  85. Meuth, S.G.; Budde, T.; Kanyshkova, T.; Broicher, T.; Munsch, T.; Pape, H.C. Contribution of TWIK-related acid-sensitive K+ channel 1 (TASK1) and TASK3 channels to the control of activity modes in thalamocortical neurons. J. Neurosci. 2003, 23, 6460–6469.
  86. Lazarenko, R.M.; Willcox, S.C.; Shu, S.; Berg, A.P.; Jevtovic-Todorovic, V.; Talley, E.M.; Chen, X.; Bayliss, D.A. Motoneuronal TASK channels contribute to immobilizing effects of inhalational general anesthetics. J. Neurosci. 2010, 30, 7691–7704.
  87. Berg, A.P.; Talley, E.M.; Manger, J.P.; Bayliss, D.A. Motoneurons express heteromeric TWIK-related acid-sensitive K+ (TASK) channels containing TASK-1 (KCNK3) and TASK-3 (KCNK9) subunits. J. Neurosci. 2004, 24, 6693–6702.
  88. Conway, K.E.; Cotten, J.F. Covalent modification of a volatile anesthetic regulatory site activates TASK-3 (KCNK9) tandem-pore potassium channels. Mol. Pharmacol. 2012, 81, 393–400.
  89. Talley, E.M.; Bayliss, D.A. 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.
  90. Buljubasic, N.; Rusch, N.J.; Marijic, J.; Kampine, J.P.; Bosnjak, Z.J. Effects of halothane and isoflurane on calcium and potassium channel currents in canine coronary arterial cells. Anesthesiology 1992, 76, 990–998.
  91. Mu, D.; Chen, L.; Zhang, X.; See, L.H.; Koch, C.M.; Yen, C.; Tong, J.J.; Spiegel, L.; Nguyen, K.C.; Servoss, A.; et al. Genomic amplification and oncogenic properties of the KCNK9 potassium channel gene. Cancer Cell 2003, 3, 297–302.
  92. Pei, L.; Wiser, O.; Slavin, A.; Mu, D.; Powers, S.; Jan, L.Y.; Hoey, T. Oncogenic potential of TASK3 (Kcnk9) depends on K+ channel function. Proc. Natl. Acad. Sci. USA 2003, 100, 7803–7807.
  93. Zúñiga, R.; Valenzuela, C.; Concha, G.; Brown, N.; Zúñiga, L. TASK-3 Downregulation Triggers Cellular Senescence and Growth Inhibition in Breast Cancer Cell Lines. Int. J. Mol. Sci 2018, 19, 1033.
  94. Zúñiga, R.; Concha, G.; Cayo, A.; Cikutović-Molina, R.; Arevalo, B.; González, W.; Catalán, M.A.; Zúñiga, L. Withaferin A suppresses breast cancer cell proliferation by inhibition of the two-pore domain potassium (K2P9) channel TASK-3. Biomed. Pharmacother. 2020, 129, 110383.
  95. Innamaa, A.; Jackson, L.; Asher, V.; Van Shalkwyk, G.; Warren, A.; Hay, D.; Bali, A.; Sowter, H.; Khan, R. Expression and prognostic significance of the oncogenic K2P potassium channel KCNK9 (TASK-3) in ovarian carcinoma. Anticancer Res. 2013, 33, 1401–1408.
  96. Rusznák, Z.; Bakondi, G.; Kosztka, L.; Pocsai, K.; Dienes, B.; Fodor, J.; Telek, A.; Gönczi, M.; Szucs, G.; Csernoch, L. Mitochondrial expression of the two-pore domain TASK-3 channels in malignantly transformed and non-malignant human cells. Virchows Arch. 2008, 452, 415–426.
  97. Nagy, D.; Gönczi, M.; Dienes, B.; Szöőr, Á.; Fodor, J.; Nagy, Z.; Tóth, A.; Fodor, T.; Bai, P.; Szücs, G.; et al. Silencing the KCNK9 potassium channel (TASK-3) gene disturbs mitochondrial function, causes mitochondrial depolarization, and induces apoptosis of human melanoma cells. Arch. Dermatol. Res. 2014, 306, 885–902.
  98. Bachmann, M.; Rossa, A.; Antoniazzi, G.; Biasutto, L.; Carrer, A.; Campagnaro, M.; Leanza, L.; Gonczi, M.; Csernoch, L.; Paradisi, C.; et al. Synthesis and cellular effects of a mitochondria-targeted inhibitor of the two-pore potassium channel TASK-3. Pharmacol. Res. 2021, 164, 105326.
  99. Wrzosek, A.; Gałecka, S.; Żochowska, M.; Olszewska, A.; Kulawiak, B. Alternative Targets for Modulators of Mitochondrial Potassium Channels. Molecules 2022, 27, 299.
  100. Yao, J.; McHedlishvili, D.; McIntire, W.E.; Guagliardo, N.A.; Erisir, A.; Coburn, C.A.; Santarelli, V.P.; Bayliss, D.A.; Barrett, P.Q. Functional TASK-3-Like Channels in Mitochondria of Aldosterone-Producing Zona Glomerulosa Cells. Hypertension 2017, 70, 347–356.
  101. Cikutović-Molina, R.; Herrada, A.A.; González, W.; Brown, N.; Zúñiga, L. TASK-3 Gene Knockdown Dampens Invasion and Migration and Promotes Apoptosis in KATO III and MKN-45 Human Gastric Adenocarcinoma Cell Lines. Int. J. Mol. Sci. 2019, 20, 6077.
  102. Sun, H.; Luo, L.; Lal, B.; Ma, X.; Chen, L.; Hann, C.L.; Fulton, A.M.; Leahy, D.J.; Laterra, J.; Li, M. A monoclonal antibody against KCNK9 K+ channel extracellular domain inhibits tumour growth and metastasis. Nat. Commun. 2016, 7, 10339.
  103. Bedoya, M.; Rinné, S.; Kiper, A.K.; Decher, N.; González, W.; Ramírez, D. TASK Channels Pharmacology: New Challenges in Drug Design. J. Med. Chem. 2019, 62, 10044–10058.
  104. Bista, P.; Cerina, M.; Ehling, P.; Leist, M.; Pape, H.C.; Meuth, S.G.; Budde, T. The role of two-pore-domain background K+ (K2P) channels in the thalamus. Pflugers Arch. 2015, 467, 895–905.
  105. Meuth, S.G.; Aller, M.I.; Munsch, T.; Schuhmacher, T.; Seidenbecher, T.; Meuth, P.; Kleinschnitz, C.; Pape, H.C.; Wiendl, H.; Wisden, W.; et al. The contribution of TWIK-related acid-sensitive K+-containing channels to the function of dorsal lateral geniculate thalamocortical relay neurons. Mol. Pharmacol. 2006, 69, 1468–1476.
  106. Linden, A.M.; Sandu, C.; Aller, M.I.; Vekovischeva, O.Y.; Rosenberg, P.H.; Wisden, W.; Korpi, E.R. TASK-3 knockout mice exhibit exaggerated nocturnal activity, impairments in cognitive functions, and reduced sensitivity to inhalation anesthetics. J. Pharmacol. Exp. Ther. 2007, 323, 924–934.
  107. Pang, D.S.; Robledo, C.J.; Carr, D.R.; Gent, T.C.; Vyssotski, A.L.; Caley, A.; Zecharia, A.Y.; Wisden, W.; Brickley, S.G.; Franks, N.P. An unexpected role for TASK-3 potassium channels in network oscillations with implications for sleep mechanisms and anesthetic action. Proc. Natl. Acad. Sci. USA 2009, 106, 17546–17551.
  108. Gotter, A.L.; Santarelli, V.P.; Doran, S.M.; Tannenbaum, P.L.; Kraus, R.L.; Rosahl, T.W.; Meziane, H.; Montial, M.; Reiss, D.R.; Wessner, K.; et al. TASK-3 as a potential antidepressant target. Brain Res. 2011, 1416, 69–79.
  109. Dadi, P.K.; Vierra, N.C.; Jacobson, D.A. Pancreatic β-cell-specific ablation of TASK-1 channels augments glucose-stimulated calcium entry and insulin secretion, improving glucose tolerance. Endocrinology 2014, 155, 3757–3768.
  110. Dadi, P.K.; Luo, B.; Vierra, N.C.; Jacobson, D.A. TASK-1 Potassium Channels Limit Pancreatic α-Cell Calcium Influx and Glucagon Secretion. Mol. Endocrinol. 2015, 29, 777–787.
  111. Wen, X.; Liao, P.; Luo, Y.; Yang, L.; Yang, H.; Liu, L.; Jiang, R. Tandem pore domain acid-sensitive K channel 3 (TASK-3) regulates visual sensitivity in healthy and aging retina. Sci. Adv. 2022, 8, eabn8785.
More
Video Production Service