Skip to main content
Brain Pathology logoLink to Brain Pathology
. 2010 May 5;20(6):999–1009. doi: 10.1111/j.1750-3639.2010.00407.x

From the Background to the Spotlight: TASK Channels in Pathological Conditions

Stefan Bittner 1, Thomas Budde 2, Heinz Wiendl 3, Sven G Meuth 1,3,
PMCID: PMC8094868  PMID: 20529081

Abstract

TWIK‐related acid‐sensitive potassium channels (TASK1–3) belong to the family of two‐pore domain (K2P) potassium channels. Emerging knowledge about an involvement of TASK channels in cancer development, inflammation, ischemia and epilepsy puts the spotlight on a leading role of TASK channels under these conditions. TASK3 has been especially linked to cancer development. The pro‐oncogenic potential of TASK3 could be shown in cell lines and in various tumor entities. Pathophysiological hallmarks in solid tumors (e.g. low pH and oxygen deprivation) regulate TASK3 channels. These conditions can also be found in (autoimmune) inflammation. Inhibition of TASK1,2,3 leads to a reduction of T cell effector function. It could be demonstrated that TASK1−/− mice are protected from experimental autoimmune inflammation while the same animals display increased infarct volumes after cerebral ischemia. Furthermore, TASK channels have both an anti‐epileptic as well as a pro‐epileptic potential. The relative contribution of these opposing influences depends on their cell type‐specific expression and the conditions of the cellular environment. This indicates that TASK channels are per se neither protective nor detrimental but their functional impact depends on the “pathophysiological” scenario. Based on these findings TASK channels have evolved from “mere background” channels to key modulators in pathophysiological conditions.

Keywords: cancer, epilepsy, inflammation, ischemia, K2P channels, oncology, TASK channels

INTRODUCTION

It was only about 10 years ago (1996–2003) that a whole new group of potassium channels named two‐pore domain potassium channels (K2P channels) was identified. They contribute to the setting and modulation of the resting membrane potential and were initially regarded as background channels. As these channels are mostly time‐ and voltage‐independent, they are especially suited to carry a background K+ conductance at negative membrane potentials, whereas other potassium channels like voltage‐gated or inward‐rectifying K+ channels generally show less persistent activity [for recent general reviews about K2P channels, see for example 30, 48, 65, 102, 125].

Compared to other K+ channels, the family of K2P channels is unique in their membrane topology as their channel subunits display two‐pore and four‐transmembrane domains (2P/4TM). Functional K2P channels consist of a homo‐ or heterodimer and the four‐pore domains build a K+‐selective pore region. The human K2P family has 15 members—of which 12 were functionally expressed in expression systems—and can be further divided into six subgroups with regard to their phylogenetic relationships (Fig. 1; the widely accepted popular nomenclature will generally be used within this review). K2P channels are under a complex control of a wide spectrum of chemical and physical stimuli as well as unusual pharmacological modulations and clinically relevant drugs [e.g. pH values 24, 40, O2 partial pressure (15), membrane stretch (104), temperature (77), G proteins 89, 125, fatty acids 66, 79 and inhalation anesthetics 16, 85, 103, 120]. Along with their selective and differential expression pattern, K2P channels are likely to be key players in a variety of physiological and pathological conditions.

Figure 1.

Figure 1

Organization of K2P channels. K2P channel members are displayed in a phylogenetic tree and sorted into six subfamilies. Nonfunctional members are displayed in grey and channels of special interest for this review are highlighted. The three common classification systems [popular name based on acronyms describing distinct properties of the channels, the KCNK classification of the Human Genome Organization (HUGO) and the K2PX.1 nomenclature of the International Union of Pharmacology (IUPHAR)] are displayed on the right side. Abbreviations: TALK = TWIK‐related alkaline‐pH‐activated K+ channel; TASK = TWIK‐related acid‐sensitive K+ channel; THIK = tandem pore domain halothane‐inhibited K+ channel; TRAAK = TWIK‐related arachidonic‐acid‐stimulated K+ channel; TREK = TWIK‐related K+‐channel gene; TRESK =  TWIK‐related spinal‐cord K+ channel; TWIK = tandem of pore domains in a weak inwardly rectifying potassium channel.

This review will especially focus on pH‐sensitive K2P channel family members [acid‐sensitive: tandem of pore domains in a weak inwardly rectifying potassium channel [TWIK]‐related acid‐sensitive K+ channel‐1 (TASK1, K2P3.1) and TASK3 (K2P9.1); alkaline‐activated: TASK2 (K2P5.1)] in light of increased research effort during the last years, which provided us with a variety of interesting findings. TASK channels are among others inhibited by extracellular acidification, hypoxia, local anesthetics and the endocannabinoid anandamide and are activated by inhalational anesthetics, but experimental studies are often challenging due to the absence of selective blockers (3). This gap could at least partially be filled by usage of knockout animals. An involvement of TASK channels could, for example, be implicated in cell proliferation, ischemic and inflammatory tissue damage and epileptic activity. Here, we summarize findings concerning the role of TASK1, TASK2 and TASK3 as signal integrators under pathophysiological conditions like tumorigenesis, ischemia, (autoimmune) inflammation and epilepsia (see Table 1 for a selection of key findings discussed in this review). Other K2P channels are only mentioned briefly when interesting findings increase the level of our knowledge and understanding of putative properties of TASK channels.

Table 1.

Literature overview. Abbreviations: TASK = tandem of pore domains in a weak inwardly rectifying potassium channel (TWIK)‐related acid‐sensitive K+ channel; TRESK = TWIK‐related spinal‐cord K+ channel.

Topic Channel Experimental finding Reference
Oncology TASK3 Overexpression in cancer samples (91)
TASK3 Anti‐apoptotic effect of TASK3 expression 75, 106
TASK3 Implication in mitochondrial dysfunction 108, 116
Ischemia TREK1 Neuroprotective effect of channel activation 37, 38
TASK1 Increased ischemic damage in TASK1−/− mice (88)
TASK1 Neuroprotective effect of channel inhibition (9)
Inflammation TASK1,3 Critical role for T cell activation 9, 83
TRESK Putative role on T lymphocytes (109)
TASK2 Influences on T cell effector functions (8)
Epilepsia TASK1 Expression on interneurons dampens neuronal acitivity (126)
TASK1 Downregulation in astrocytes has an anti‐epileptic effect (50)
TASK2 Protective effect upon channel upregulation (51)

TASK CHANNELS IN ONCOLOGY

Background

Potassium channels are critically involved in the signaling pathways that regulate cell proliferation and cell death 57, 58, 123, 131. Inhibition of these channels has been shown to decrease cell proliferation in human lymphocytes 70, 83, 111, embryonic stem cells (90), human melanoma cells (97), hepatocarcinoma cells (142), glioma cells (41), small cell lung (98), breast (133) and prostate (121) cancer cells. The mechanisms by which these channels might alter the proliferative state of cells has been an open question for several years and initial studies linked the functional role of K+ channels to their impact on the intracellular Ca2+ concentration 118, 132, to the regulation of the cell volume (114) or to direct activation of intracellular signaling cascades 131, 139. However, it has also become clear that K+‐currents play a necessary and pivotal role in apoptosis 11, 42 and oncogenesis (100).

Experimental data

Recently, one member of the K2P family, namely TASK3, has been shown to be amplified 3–10‐fold in 10% of human breast cancer samples (91) and thus was identified as a novel proto‐oncogene in breast cancer. Furthermore, TASK3 was found to be overexpressed in 10%–50% of breast cancer, lung cancer and colorectal cancer cells 47, 91. Overexpression of TASK3 in NmuMG (cell line derived from mammary epithelial tissue) or NIH‐3T3 (fibroblasts) cells did not result in transformation of these cell lines. However, mice injected with NmuMG cells overexpressing TASK3 developed tumors within 3 months (91).

Other studies have investigated the oncogenic potential of TASK3 in more detail. Transfection of TASK3 in the partially transformed mouse embryonic fibroblast cell line C8 strengthens the oncogenic potential of C8 cells and especially increases cell survival under pro‐apoptotic conditions [low serum conditions, hypoxia; 75, 91]. Liu et al. additionally showed that overexpression of TASK3 as well as TASK1 and TASK2 enhanced cell viability of transfected C8 cells by inhibiting intracellular apoptosis pathways. In support of these findings, C8 cells transfected with a mutated TASK3 construct with abolished TASK3 channel activity showed reduced oncogenic properties, which directly link TASK3 channel function and tumorigenesis (106). These results are nicely extended by other studies focusing on K+ channel‐dependent cell death using high extracellular K+ concentrations 61, 86, which can be found in vivo under pathophysiological conditions with impaired K+ homeostasis 12, 63. In this model, rat granule neurons and several human glioma cell lines undergo cell death when cultured in a medium with physiological K+ concentration (5 mM), but cell survival is promoted by high extracellular K+ concentrations (25 mM). Using pharmacological modulators and changes in the extracellular pH value, this effect could be linked to TASK3 channel activity 61, 86. The biophysical properties of TASK3 (pH dependency, extracellular K+ concentration‐dependent conductance changes) may be especially relevant in the hypoxic and acidic micro milieu often occurring in the center of solid tumors (105): TASK1 and TASK3 channels are both inhibited by acidosis [pK values: TASK1: 7.3; TASK3: 6.7; (105)] and in addition TASK1 channels are directly inhibited by hypoxia. The conflicting data of O2 sensitivity of TASK1 channels 44, 68 may be explained by a recent report (62) proposing NOX4 as O2 sensor of TASK1 channels. Differential NOX4 expression could explain the inhomogeneous O2 sensitivity of TASK1. However, other channels will also be affected by hypoxia like tandem pore domain halothane‐inhibited K+ channel (THIK1) (17), voltage‐gated potassium channels (101) and also a variety of other targets. Thus more research effort is needed to dissect the pleiotropic effects of hypoxia in tumors in vivo.

TASK3 channels have also been reported on melanoma cells 108, 116. Thorough investigations showed that TASK3 was expressed especially in mitochondria. Several potassium channels (124) have been shown to play a role in mitochondrial dysfunction. While the exact role of TASK3 channels in this context is yet unclear, their involvement in mitochondrial damage might be another piece of the puzzle toward understanding the role of TASK3 in apoptosis and tumorigenesis.

Interestingly, genomic investigations have revealed that TASK3 is an imprinted gene 76, 115, and proposing that the regulation and diversity of TASK3 channel function might even be more complex than thought before. Several diseases like Prader–Willi syndrome or Angelman syndrome have their cause in epigenetic dysfunction and a recent report of the maternally inherited so‐called Birk Barel mental retardation dysmorphism syndrome (2) revealed a similar hereditary disease for TASK3. As several tumors like Wilms tumor have also been linked to imprinting errors, it will be interesting to see this new evolvement of TASK channel research in the future.

Besides TASK3, differential expression of TASK1, another member of the K2P family with comparable biophysical properties, could be correlated with prolonged survival of patients with advanced colorectal cancer (Dukes stages C and D) 5 years after surgery (18). This finding supports the hypothesis that both members of the K2P family (TASK1 and TASK3) who are also able to form heterodimers 19, 49 can play a critical role in a variety of cancer entities (see Fig. 2).

Figure 2.

Figure 2

Involvement of TASK channels in the underlying pathophysiological mechanisms of A. oncology, B. ischemia, C. (autoimmune) inflammation and D. epilepsy. A. A schematic of a neuron is shown. 1. TASK3 channels influence cell survival and proliferation on different levels including cell volume, [Ca2+]I and apoptosis. A low external pH value might block TASK3 channels. 2. TASK3 channels are also implicated in mitochondrial dysfunction. 3. In vitro experiments show that high local [K+]e is associated with a reduced cell survival by blocking TASK3 channels. 4. Besides TASK3, other K2P channels (TASK1, TREK1, TASK5 and THIK2) have been linked to tumor development. B. A neuron under ischemic conditions is shown. 1. TREK1 channels are activated by lower internal pH values, PUFAs, riluzole and inhalational anesthetics. 2. TASK1 channels are inhibited by anandamide, lower external pH values, hypoxia and activated by inhalational anesthetics. 3. Activation of potassium channels limits a harmful overload of the internal calcium concentration. 4. Depletion of potassium can, on the other hand, induce apoptosis. C. A number of ion channels are expressed on T lymphocytes. 1. TCR activation leads to an elevation of the internal calcium concentration necessary for further gene‐transcription and proliferation. 2. Potassium channels provide the counterbalancing driving force mandatory for this calcium influx. 3. TASK channels are inhibited by low external pH value and hypoxia, which might occur under inflammatory conditions. D. 1. TASK channels on interneurons may suppress intrinsic pacemaking and spontaneous firing in pyramidal cells. 2. An elevated TASK2 expression may dampen epileptic activity in response to the early seizure‐related alkalinization. 3. Neurotransmitters hyperpolarize pyramidal cells by activating TWIK1 and TREK2 channels. 4. TASK1 downregulation in astrocytes dampens seizure activity. Abbrevations: Ac = astrocyte; [Ca2+]i = internal calcium concentration; CaM = Calmodulin; CRAC =  Calcium release‐activated calcium channel; GABA = gamma‐aminobutyric acid; 5‐HT = serotonin; IKCa1; intermediate calcium‐dependent potassium channel; IKIR = inward rectifier K+ channel; In = interneuron; [K+]e = external potassium concentration; NA =  noradrenalin; pH↓ = lower pH value; pH↑ = increased pH value; PUFA = polyunsaturated fatty acids; Py = pyramidal cell; O2↓ = hypoxia; TASK = TWIK‐related acid‐sensitive K+ channel; TCR = T‐cell receptor; THIK = tandem pore domain halothane‐inhibited K+ channel; TRESK =  TWIK‐related spinal‐cord K+ channel. Dashed lines indicate receptor‐mediated signaling pathways.

It should briefly be mentioned here that other members of the K2P family as well have been linked to different cancer entities, like TWIK‐related K+‐channel gene‐1 (TREK1) to prostate carcinoma (130) and ALL(127), THIK2 to tubular breast carcinoma, synovial sarcoma and malignant peripheral nerve sheet tumor 94, 95, 113 and TASK5 to colon cancer and acute lymphoblastic leukemia (ALL) (119).

Taken together, the influence of TASK channels on tumors depends on the exact channel subunit, expressing cell type and surrounding pathophysiological stimuli (105). Increasing evidence points to an involvement of TASK channels in cell proliferation, apoptosis and tumorigenesis.

TASK CHANNELS AND ISCHEMIA

Background

Neuronal damage caused by ischemic or inflammatory events is an important cause for persistent disability and death in clinical practice. So far, the molecular mechanisms leading to neuronal death are poorly understood. Mild hypoxia or inflammation can, for example, induce neuroprotective signaling cascades preventing neuronal death 23, 34, 107. The role of K2P channels during cellular stress remains unclear and is discussed controversially.

Activation of potassium channels results in membrane hyperpolarization, which has important implications for the function and survival of neurons: decreased neuronal activity and thereby lower metabolic demands could enhance neuronal survival under critical conditions. Moreover, the opening of voltage‐dependent Ca2+ channels in presynaptic neurons would be limited. Alongside an enhancement of the Mg2+ block of NMDA receptors (N‐methyl D‐aspartate) in postsynaptic neurons (80), this would protect against glutamate excitotoxicity 27, 39. Additionally, an immediate effect of neuronal depolarization might be a direct excitotoxic neuronal damage caused by prolonged influx of Ca2+ ions via different routes, including a reversal of the Na+/Ca2+ antiporter. The protective effect of K+ channels against an internal overload of Ca2+ ions, which would lead to apoptosis by activation of proteases, lipases and endonucleases, has been previously linked to other K+ channels [e.g. large conductance Ca2+‐activated K+ channels and adenosine triphosphate (ATP)‐sensitive K+ channels 36, 60]. This protective effect might therefore be true for different potassium channels remaining active during potentially harmful pathophysiological conditions.

Experimental data

So far, research has focused on a putative neuroprotective role of TREK and TASK channels. The open probability of TREK1 and TREK2 channels is strongly increased upon reduction of the intracellular pH value 1, 78, which can be observed in ischemia or inflammation (73). Polyunsaturated fatty acids (PUFAs)—known to provide neuroprotection from cerebral ischemia (59)—as well as riluzole—the first approved neuroprotective agent for the treatment of amyotrophic lateral sclerosis—exert their neuroprotective potential at least partially through activation of TREK1 channels 25, 38. Alpha linelonic acid and lysophosphatidylcholine (TREK1 activators) showed no additional effect in TREK1−/− mice proofing the specificity of the protective effect. TREK1−/− mice displayed significantly reduced neuronal survival rates in a model of cerebral ischemia (37).

These findings were supported by the observation of a worse outcome in a transient middle cerebral artery occlusion model of cerebral ischemia in TASK1−/− mice compared to controls (88). TASK1 and 3 exert an important contribution to the standing outward current (ISO) of neurons, which regulates, among others, the activity of neurons and shapes the action potential generation. In brain slices, pH reduction and O2 deprivation led to a ∼40% reduction of the whole cell outward current resulting in membrane depolarization and concomitantly to an increase in neuronal activity. In vivo, TASK1−/− mice showed significantly increased infarct volumes and this situation could be mimicked by the TASK1/3 inhibitor anandamide (88). Both TASK1 and TASK3 are inhibited by low extracellular pH values. TASK1 is also sensitive to hypoxic conditions (101), whereas reports are still inconsistent regarding TASK3 [(14) versus (35)]. Detailed single‐channel electrophysiological studies in rat glomus cells revealed recently that all TASK1, TASK3 and TASK1/3 heterodimers are inhibited by hypoxia (49). It was suggested that species differences might be responsible for different reports concerning TASK channels and oxygen sensitivity (136). Of notice, the phenotype of TASK2−/− mice upon ischemic brain damage has not been reported so far, while the infarct volume in TASK3−/− mice was unchanged (92).

This line of evidence is supported by experimental data describing a neuroprotective role of inhalational anesthetics, which is thought to be mediated by an inhibition of excitatory activity. Besides direct antagonistic mechanism on NMDA and amino‐3‐hydroxy‐5‐methylisoxazole‐4‐propionic acid (AMPA) receptors and a potentiation of gamma‐aminobutyric acid (GABA) receptors, the neuroprotective effects of inhalation anesthetics have also been connected to an activation of K2P channels 27, 54. Many but not all K2P channels are opened by volatile anesthetics [TREK1, TREK2, TASK1, TASK3, TASK2, TWIK‐related alkaline‐pH‐activated K+ channel‐2 (TALK2) and TWIK‐related spinal‐cord K+ channel (TRESK) 29, 31, 67, 74, 81, 103]. Supportively, knockout of the predominantly neuronal expressed channels TREK1, TASK1 and TASK3 37, 71, 72 results in moderate resistance to inhalational anesthesia, while this effect is not observed in TASK2−/− or KCNK7−/− mice 28, 140.

On the other hand, an intracellular depletion of K+ is also a typical sign of early apoptosis 61, 75, resulting in neuronal cell death. This situation might especially occur under longer lasting pathophysiological conditions as in chronic (autoimmune) inflammation. Recently, TASK1−/− mice were shown to be less susceptible to neuronal degeneration in myelin oligodendrocyte glycoprotein (MOG) peptide‐induced experimental autoimmune encephalomyelitis (EAE), an animal model of multiple sclerosis (MS) (9). Axonal loss was significantly reduced in TASK1−/− mice. The chosen experimental model, however, does not allow to clearly dissect between central nervous system‐related neuroprotective effects of TASK1 deficiency and an indirect neuroprotection via an attenuated inflammation‐mediated neurodegeneration. Therefore, experiments were performed with acute brain slices from naïve wild‐type and TASK1−/− mice which were cocultured with CD4+ T lymphocytes isolated from immunized wild‐type EAE mice at disease maximum. This approach ensured an identical inflammatory component under both conditions. A significant reduction of neuronal apoptosis in brain slices from TASK1−/− mice argued in favor of a direct neuroprotective impact of TASK channel inhibition on neurons under these experimental conditions. As a note of caution, it should be kept in mind that the axonal degeneration in EAE is rather acute and mostly driven by T cell infiltration. Targeting the role of TASK1 in mouse models of Parkinson's, Alzheimer's or motor neuron disease should provide further insight into the role of TASK channels in the future.

The described phenotype is in contrast to the increased susceptibility of TASK1−/− mice to acute ischemic brain damage as discussed above. It might be speculated that the diverse contribution of TASK channels on neuronal survival depends on the strength and duration of the pathophysiological stimulus. While TASK channel might contribute to neuronal hyperpolarization in the early phase of a pathological stimulus (acute ischemia), the situation might change as soon as a certain threshold of damage severity is reached (chronic inflammatory ischemia). Now the role of TASK channels switches to being a player in neuronal apoptosis. Therefore, inhibition of TASK channels might be neuroprotective in the long‐term course by preventing neuronal apoptosis. It should also be mentioned here that lowering of the extracellular pH value will not only affect TASK channels but also a variety of other targets as well, including several ion channels like transient receptor potential vanilloid 1 (TRPV1), acid sensing ion channels (ASICs) and NMDA receptors. This makes it difficult to predict the overall effect of acidosis in vivo.

TASK CHANNELS IN (AUTOIMMUNE) INFLAMMATION

Background

Human T lymphocytes are key players in the pathogenesis of autoimmune diseases like MS, rheumatoid arthritis (RA) or type I diabetes mellitus (7). The underlying pathophysiological processes of these diseases are still not completely understood but a major contribution of ion channels to T lymphocyte functions has been suggested. Therefore ion channels were appraised as potential target structures for future pharmacological approaches 53, 135. Upon T cell receptor activation, a longer lasting elevation of the intracellular calcium concentration is mandatory for transcription‐dependent steps of T cell activation 69, 96, 110 and potassium channels provide the counterbalancing hyperpolarizing current, which preserves the electrochemical driving force for an influx of Ca2+ ions (99).

So far, three different potassium currents could be identified to play an important role on T lymphocytes: a voltage‐gated K+ channel [KV1.3 (20)], a Ca2+‐activated K+ channel [IKCa1 or KCa3.1 (64)] and the recently identified K2P channels TASK1, TASK2, TASK3 and TRESK 8, 83, 109. Selective blockade of KV1.3 influences T cell effector functions leading to reduced cytokine production and suppressed proliferation rates 5, 7. KV1.3 knockout mice did not show the expected phenotype because of compensating chloride currents (55) but in vivo blockade of KV1.3 in rats and mini swine significantly ameliorated the disease course of EAE and delayed‐type hypersensitivity response 6, 56. EAE animals showed a later disease onset and reduced disease maximum after immunization. Electrophysiological characterization of human T cell subtypes was used to investigate differential regulation of KV1.3 and IKCa1 after antigen stimulation: Unstimulated T cells express about 200–300 copies of KV1.3 and chronic T cell stimulation in vitro results in a selective and stable upregulation of KV1.3 channels [1.500/cell (4)] on a special CD4+ T cell subtype named CCR7CD45RA effector memory T cells (TEM), which are thought to have an substantial impact in autoimmune diseases (134).

Experimental data

Most recently, three members of the K2P family, TASK1, TASK2 and TASK3 could be identified and functionally characterized in the context of T cell activation 8, 9, 83. Human, rat and mouse T lymphocytes constitutively express TASK1, TASK2 and TASK3 and pharmacological inhibition leads to reduced cytokine production and proliferation after T cell receptor stimulation. Patch‐clamp experiments of stimulated T lymphocytes revealed an important anandamide‐sensitive component of the whole cell outward current and other current characteristics indicative of TASK channels. Based on these measurements and a numerical model analysis, it could be shown that around 40% of the total outward current is mediated by TASK channels. The in vivo relevance of TASK1/3 for T cell activation and effector functions could be confirmed in an active transfer EAE model in rats. Selective blockade of TASK channels on MBP specific T cells before i.v. transfer resulted in a reduced disease activity (83). Supportingly, TASK1−/− mice showed a significantly ameliorated disease course upon MOG35‐55 immunization, which was accompanied by a reduced activation of the immune system while naïve T lymphocytes per se were not severely impaired in TASK1−/− mice. This EAE phenotype could be reproduced by the TASK channel inhibitor anandamide in wildtype mice when administered daily resulting in a delayed disease onset and reduced disease severity over 50 days. Remarkably, anandamide was equally effective when treatment only started after occurrence of first disease symptoms (9) or using a preventive treatment strategy (starting from the day of disease induction). Differential regulation of TASK1,3 during MOG‐induced EAE in rats was confirmed by immunohistochemistry and in situ hybridization revealing increased neuronal expression of TASK1,3 in the optic nerve and decreased expression in the spinal cord and thalamus. Lesion analysis of human brain autopsies displayed a similar downregulation of TASK channels emphasizing the regulatory role for TASK channels in neurons under pathophysiological conditions (87). Furthermore, human as well as rodent T lymphocytes functionally express TASK2 (see Fig. 2), which was found to be upregulated in MS patients with active disease.

A putative role of other members of the K2P family in the immune system still needs to be addressed. Recently, an expression of the calcineurin‐regulated K2P channel TRESK on Jurkat cells was postulated 109, 129 but a functional role for TRESK still lacks confirmation on primary T cells. This point is critical because of known differences in channel expression between leukemic and normal T cells [e.g. small conductance Ca2+‐activated K+ channels (SK2) 32, 43] and further studies are warranted to clarify the relevance of TRESK channel expression on immune cells in vivo. Additionally, an expression of TREK2 in mouse B lymphocytes has been described recently (141).

Taken together, TASK channels have been demonstrated to play a role in the regulation of the membrane potential of T lymphocytes and for immunmodulation in vitro and in vivo and might therefore be an interesting new molecular target for pharmacological approaches in autoimmune diseases. The rising plurality of different endogenous potassium channels on T lymphocytes, combined with multiple mechanisms of regulation, underlines the importance of a tightly controlled membrane potential of these cells. However, the picture is also becoming more complex demanding further studies on the role of ion channels in the immune system.

TASK CHANNELS IN EPILEPSY/SLEEP

Background

Epilepsy is a common chronic neurological condition that is characterized by the spontaneous occurrence of local or generalized episodes of neuronal discharges (112). These recurrent seizures are correlated with abnormal excessive and synchronous neuronal activity in the brain. Some types of frontal, generalized and infantile seizure syndromes have been associated with mutations in genes coding for protein subunits of ion channels, abnormal synaptic transmission, modulation of transmembrane K+ gradients and activity‐dependent shifts in extracellular pH values. The extracellular pH in the brain is tightly regulated and displays only minimal changes (45). However there are physiological as well as pathological conditions in which the extracellular pH of the brain changes. Seizure activity for example results in a biphasic pH shift, consisting of an initial extracellular alkalinization followed by a delayed acidification (138). Since K2P channels are highly regulated by pH, several subtypes have been investigated in neurons as well as glia cells with respect to their possible involvement in the generation of and adaptation to epilepsy (see Fig. 2).

Experimental data

Most information is available for the role of TASK channels in hippocampal and entorhinal cortical regions representing key structures in temporal lobe epilepsy (93). In the rat hippocampus, TASK channels are differentially expressed in neurons and glia cells of different subregions 50, 52. In the CA1 region, TASK1 expression has been found to some extent in pyramidal cells and horizontal interneurons, while TASK3 was strongly expressed in pyramidal cells (126). The resulting hyperpolarizing effect of TASK channels may suppress intrinsic pacemaking and spontaneous firing in pyramidal cells, an electrophysiological behavior that is otherwise typical for interneurons (see Fig. 2). A second consequence may be the cell type‐specific sensitivity to extracellular acidification in conditions such as epilepsy. In addition, TASK2 expression was found in CA1‐3 pyramidal cells, dentate gyrus granula cells and perivascular astrocytes (51).

Astroglial expression of TASK1 was also described in the hippocampus of gerbils (51), a well‐established animal model of inherited epilepsy (13). In seizure‐sensitive strains, offspring of epileptic parents have a high probability of developing epilepsy in later life and individual seizures can be induced by mechanical stimulation of the back. Characteristic changes in TASK channel expression occur before and after the onset/induction of epileptic seizures. In gerbils, hippocampal expression of TASK1 and TASK2 was not different between young seizure‐sensitive and seizure‐resistant individuals (50). In adult seizure‐sensitive animals, the expression of TASK1 in astrocytes was higher compared to epilepsy‐free animals. After an individual seizure, TASK1 expression in astrocytes was downregulated. The general upregulation of TASK1 in astrocytes of seizure‐sensitive animals may diminish astroglial K+ buffering through an increased K+ outward rectification leading to slightly increased extracellular K+ levels and thus increased excitability. The rapid downregulation of TASK1 expression in astrocytes during an individual seizure is part of a fast adaptation process that helps to dampen seizure activity by reducing astroglial outward rectification. In this respect, it is interesting to note that several anti‐epileptic drugs reduce the expression of TASK1 channels in astrocytes (50). In rat hippocampus, experimentally induced temporal lobe epilepsy is associated with an increase of TASK2 expression in the CA3 pyramidal layer, dentate granule layer, and the endfeet of perivascular astrocytes (50). The elevated TASK2 expression in neurons probably represents a rapid adaptation that results in hyperpolarization and dampening of epileptic activity in response to the early seizure‐related alkalinization (see Fig. 2). The upregulated TASK2 expression in perivascular regions may contribute to hippocampal damage by contributing to abnormal blood flow and/or blood–brain barrier abnormalities.

Findings in interneurons of the entorhinal cortex add to our knowledge concerning the role of TASK3 channels in temporal lobe epilepsy. The entorhinal cortex receives strong serotonergic inputs from the raphe nuclei (10) and serotonin inhibits neuronal excitability (33). In GABA‐ergic interneurons of the entorhinal cortex, application of serotonin generates membrane depolarization mediated by inhibition of TASK3 channels and an increase in action potential firing (21). In consequence, the release of GABA by interneurons is increased, resulting in decreased pyramidal cell activity and inhibition of low‐Mg2+‐induced seizure activity. Two other types of K2P channels present on entorhinal pyramidal cells, namely TWIK1 and TREK2, may be the molecular basis for the anti‐epileptic effects of monoamines (see Fig. 2). Serotonin and noradrenalin hyperpolarize entorhinal pyramidal cells by activating TWIK1 and TREK2 channels, respectively 22, 137. As a result, neuronal excitability in the entorhinal cortex and its hippocampal projection areas is inhibited.

The thalamocortical neuronal network is characterized by two fundamentally different states of activity (122): whereas low‐frequency (≤15 Hz) oscillatory activity occurs during natural sleep, deep anesthesia and absence epilepsy, high‐frequency oscillatory activity in the gamma range and tonic activity prevail during wakefulness. The switch between the two states of thalamic activity is accompanied by a depolarizing shift of the membrane potential of thalamocortical relay neurons and is governed by the neurotransmitter‐induced inhibition of TASK1 and TASK3 channels 82, 84. Therefore TASK3 has been regarded as a promising candidate gene for absence epilepsy in humans. After localizing the human TASK3 in the chromosomal region 8q24, a mutation analysis of the TASK3 gene in absence epilepsy patients revealed one exon‐2 polymorphism, which, however, was not associated with the disease (46). In another attempt to map a genetic cause of idiopathic generalized epilepsies in humans, a balanced translocation breakpoint involving chromosome 6p21 has been found (117). A sequence analysis revealed the genes of TALK1 and TALK2 at a distance of 3.5 kb downstream of the breakpoint. No malfunction of these K2P channels was found.

In addition to the K2P subtypes discussed above, TREK1 channels are involved in neuroprotection against epilepsy (37). TREK1‐deficient mice are hypersensitive to drug‐induced seizures and show no PUFA‐dependent neuroprotection against epilepsy.

It is concluded that TASK channels might have both an anti‐epileptic as well as a pro‐epileptic potential; the relative contribution of these opposing influences depend on their cell type‐specific expression and the exact conditions of the cellular environment. In addition genetic studies in humans were not able to demonstrate a coupling between mutated K2P genes and epilepsy. In comparison to their role in ischemia, which involves a long‐term anti‐apoptotic effect, the influence of TASK channels on cellular excitability acts rather in the short term. Based on the findings discussed here the therapeutic potential of K2P channels was emphasized in many cases 37, 50, 128.

CONCLUSION

Pathophysiological conditions of different etiology share comparable stimuli on the cellular level albeit their relative contribution to a specific disorder differs in strength and duration. It is particularly worth noting that oxygen deprivation, extra‐ and/or intracellular acidification, endogenous cannabinoids, disturbances of K+ homeostasis or actions by neurotransmitters (e.g. glutamate) belong to a multifarious repertoire of alterations that are particularly observed during potentially harmful processes. Based on their biophysical properties, expression pattern and complex physicochemical regulation, TASK channels represent target structures that are influenced by the stimuli mentioned above. Therefore, they have the potential to integrate these signals on the molecular level, resulting in modulation of basal cellular parameters (e.g. K+ homeostasis, resting membrane potential, action potential generation).

A dual influence of TASK channels on neuronal survival evolves during different pathophysiological situations: TASK channels contribute to neuronal hyperpolarization in the early phase of a moderate enduring pathophysiological stimulus (e.g. under ischemic condition), thus decreasing neuronal activity and preventing excessive Ca2+ overload. The situation changes once a certain threshold of damage severity is reached (e.g. long‐lasting inflammation). Now the role of TASK channels switches to being a player in neuronal apoptosis as K+ efflux can be a characteristic initial feature of early apoptosis. Therefore, inhibition of TASK channels might be neuroprotective in the long‐term course by preventing neuronal cell death.

TASK channels have also been demonstrated to be relevant for the regulation of the membrane potential of T lymphocytes and for immunmodulation in vitro and in vivo. Therefore they might be an interesting new molecular target for pharmacological approaches in autoimmune diseases. However many yet unsolved questions still await clarification: Possible off‐target effects of TASK channel modulation? Are TASK channels differently expressed on various immune cell subtypes (e.g. T effector cells, regulatory T cells)? How does antigen‐specific stimulation influence TASK channel expression? Are TASK channel functions regulated intracellularly in immune cells? What is the contribution of TASK channels in different autoimmune disorders (e.g. MS, RA)? Future research efforts will help to shed light on these and further unsolved questions (26).

Furthermore, TASK channels have an important influence on cellular excitability, thereby influencing epileptic conditions. In this context, they seem to have both an anti‐epileptic, as well as a pro‐epileptic potential. The relative contribution of these opposing influences most likely depends on their cell type‐specific expression and the exact micromilieu. However, genetic analysis has so far revealed no clear association between epilepsy and TASK channel expression and function.

Finally, we do not want to argue that TASK channels represent the Rosetta stone for diagnosis and treatment of neurological disorders. However, based on their expression pattern, biophysical regulation and their impact on basal cellular parameters like the resting membrane potential, they seem to be critically involved in pathophysiological scenarios shared by different neurological disorders. Highly specific inhibitors and activators of these channels (e.g. small peptide ligands and blocking antibodies), specific antibodies and further knockout animals (e.g. TRESK) are warranted to complete the scientific armamentarium for the investigation of the pathophysiological role of these interesting—potentially therapeutic and useful—molecular structures.

ACKNOWLEDGEMENTS

This work was supported by the German Research Society (SFB 581, TP A10) to SGM and DFG K2P‐Forschergruppe 1086 (BU1019/9‐1 to TB and ME3283/1‐1 to SGM). The authors have no conflicting financial interests.

REFERENCES

  • 1. Bang H, Kim Y, Kim D (2000) TREK‐2, a new member of the mechanosensitive tandem‐pore K+ channel family. J Biol Chem 275:17412–17419. [DOI] [PubMed] [Google Scholar]
  • 2. Barel O, Shalev SA, Ofir R, Cohen A, Zlotogora J, Shorer Z et al. (2008) Maternally inherited Birk Barel mental retardation dysmorphism syndrome caused by a mutation in the genomically imprinted potassium channel KCNK9. Am J Hum Genet 83:193–199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Bayliss DA, Barrett PQ (2008) Emerging roles for two‐pore‐domain potassium channels and their potential therapeutic impact. Trends Pharmacol Sci 29:566–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Beeton C, Chandy KG (2005) Potassium channels, memory T cells, and multiple sclerosis. Neuroscientist 11:550–562. [DOI] [PubMed] [Google Scholar]
  • 5. Beeton C, Pennington MW, Wulff H, Singh S, Nugent D, Crossley G et al. (2005) Targeting effector memory T cells with a selective peptide inhibitor of Kv1.3 channels for therapy of autoimmune diseases. Mol Pharmacol 67:1369–1381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Beeton C, Wulff H, Barbaria J, Clot‐Faybesse O, Pennington M, Bernard D et al. (2001) Selective blockade of T lymphocyte K(+) channels ameliorates experimental autoimmune encephalomyelitis, a model for multiple sclerosis. Proc Natl Acad Sci USA 98:13942–13947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Beeton C, Wulff H, Standifer NE, Azam P, Mullen KM, Pennington MW et al. (2006) Kv1.3 channels are a therapeutic target for T cell‐mediated autoimmune diseases. Proc Natl Acad Sci U S A 103:17414–17419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Bittner S, Bobak N, Herrmann AM, Göbel K, Meuth P, Höhn KG et al. (2010) Upregulation of K2P5.1 potassium channels in multiple sclerosis. Ann Neurol 9999:DOI: 10.1002/ana.22010. [DOI] [PubMed] [Google Scholar]
  • 9. Bittner S, Meuth SG, Gobel K, Melzer N, Herrmann AM, Simon OJ et al. (2009) TASK1 modulates inflammation and neurodegeneration in autoimmune inflammation of the central nervous system. Brain 132(Pt 9):2501–2516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Bobillier P, Pettijean F, Salvert D, Ligier M, Seguin S (1975) Differential projections of the nucleus raphe dorsalis and nucleus raphe centralis as revealed by autoradiography. Brain Res 85:205–210. [DOI] [PubMed] [Google Scholar]
  • 11. Bortner CD, Cidlowski JA (2004) The role of apoptotic volume decrease and ionic homeostasis in the activation and repression of apoptosis. Pflugers Arch 448:313–318. [DOI] [PubMed] [Google Scholar]
  • 12. Branston NM, Strong AJ, Symon L (1977) Extracellular potassium activity, evoked potential and tissue blood flow. Relationships during progressive ischaemia in baboon cerebral cortex. J Neurol Sci 32:305–321. [DOI] [PubMed] [Google Scholar]
  • 13. Buchhalter JR (1993) Animal models of inherited epilepsy. Epilepsia 34(Suppl. 3):S31–S41. [DOI] [PubMed] [Google Scholar]
  • 14. Buckler KJ, Honore E (2005) The lipid‐activated two‐pore domain K+ channel TREK‐1 is resistant to hypoxia: implication for ischaemic neuroprotection. J Physiol 562(Pt 1):213–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Buckler KJ, Williams BA, Honore E (2000) An oxygen‐, acid‐ and anaesthetic‐sensitive TASK‐like background potassium channel in rat arterial chemoreceptor cells. J Physiol 525(Pt 1):135–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Budde T, Coulon P, Pawlowski M, Meuth P, Kanyshkova T, Japes A et al. (2008) Reciprocal modulation of I (h) and I (TASK) in thalamocortical relay neurons by halothane. Pflugers Arch 456:1061–1073. [DOI] [PubMed] [Google Scholar]
  • 17. Campanucci VA, Brown ST, Hudasek K, O’Kelly IM, Nurse CA, Fearon IM (2005) O2 sensing by recombinant TWIK‐related halothane‐inhibitable K+ channel‐1 background K+ channels heterologously expressed in human embryonic kidney cells. Neuroscience 135:1087–1094. [DOI] [PubMed] [Google Scholar]
  • 18. Cavalieri D, Dolara P, Mini E, Luceri C, Castagnini C, Toti S et al. (2007) Analysis of gene expression profiles reveals novel correlations with the clinical course of colorectal cancer. Oncol Res 16:535–548. [DOI] [PubMed] [Google Scholar]
  • 19. Czirjak G, Enyedi P (2002) Formation of functional heterodimers between the TASK‐1 and TASK‐3 two‐pore domain potassium channel subunits. J Biol Chem 277:5426–5432. [DOI] [PubMed] [Google Scholar]
  • 20. DeCoursey TE, Chandy KG, Gupta S, Cahalan MD (1984) Voltage‐gated K+ channels in human T lymphocytes: a role in mitogenesis? Nature 307:465–468. [DOI] [PubMed] [Google Scholar]
  • 21. Deng PY, Lei S (2008) Serotonin increases GABA release in rat entorhinal cortex by inhibiting interneuron TASK‐3 K+ channels. Mol Cell Neurosci 39:273–284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Deng PY, Poudel SK, Rojanathammanee L, Porter JE, Lei S (2007) Serotonin inhibits neuronal excitability by activating two‐pore domain k+ channels in the entorhinal cortex. Mol Pharmacol 72:208–218. [DOI] [PubMed] [Google Scholar]
  • 23. Dirnagl U, Simon RP, Hallenbeck JM (2003) Ischemic tolerance and endogenous neuroprotection. Trends Neurosci 26:248–254. [DOI] [PubMed] [Google Scholar]
  • 24. Duprat F, Lesage F, Fink M, Reyes R, Heurteaux C, Lazdunski M (1997) TASK, a human background K+ channel to sense external pH variations near physiological pH. EMBO J 16:5464–5471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Duprat F, Lesage F, Patel AJ, Fink M, Romey G, Lazdunski M (2000) The neuroprotective agent riluzole activates the two P domain K(+) channels TREK‐1 and TRAAK. Mol Pharmacol 57:906–912. [PubMed] [Google Scholar]
  • 26. Enyedi P, Czirják G (2010) Molecular background of leak K+ currents: two‐pore domain potassium channels. Physiol Rev 90:559–605. [DOI] [PubMed] [Google Scholar]
  • 27. Franks NP, Honore E (2004) The TREK K2P channels and their role in general anaesthesia and neuroprotection. Trends Pharmacol Sci 25:601–608. [DOI] [PubMed] [Google Scholar]
  • 28. Gerstin KM, Gong DH, Abdallah M, Winegar BD, Eger EI 2nd, Gray AT (2003) Mutation of KCNK5 or Kir3.2 potassium channels in mice does not change minimum alveolar anesthetic concentration. Anesth Analg 96:1345–1349. [DOI] [PubMed] [Google Scholar]
  • 29. Girard C, Duprat F, Terrenoire C, Tinel N, Fosset M, Romey G et al. (2001) Genomic and functional characteristics of novel human pancreatic 2P domain K(+) channels. Biochem Biophys Res Commun 282:249–256. [DOI] [PubMed] [Google Scholar]
  • 30. Goldstein SA, Bockenhauer D, O’Kelly I, Zilberberg N (2001) Potassium leak channels and the KCNK family of two‐P‐domain subunits. Nat Rev Neurosci 2:175–184. [DOI] [PubMed] [Google Scholar]
  • 31. Gray AT, Zhao BB, Kindler CH, Winegar BD, Mazurek MJ, Xu J et al. (2000) Volatile anesthetics activate the human tandem pore domain baseline K+ channel KCNK5. Anesthesiology 92:1722–1730. [DOI] [PubMed] [Google Scholar]
  • 32. Grissmer S, Lewis RS, Cahalan MD (1992) Ca(2+)‐activated K+ channels in human leukemic T cells. J Gen Physiol 99:63–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Grünschlag CR, Haas HL, Stevens DR (1997) 5‐HT inhibits lateral entorhinal cortical neurons of the rat in vitro by activation of potassium channel‐coupled 5‐HT1A receptors. Brain Res 770:10–17. [DOI] [PubMed] [Google Scholar]
  • 34. Hallenbeck JM (2002) The many faces of tumor necrosis factor in stroke. Nat Med 8:1363–1368. [DOI] [PubMed] [Google Scholar]
  • 35. Hartness ME, Lewis A, Searle GJ, O’Kelly I, Peers C, Kemp PJ (2001) Combined antisense and pharmacological approaches implicate hTASK as an airway O(2) sensing K(+) channel. J Biol Chem 276:26499–26508. [DOI] [PubMed] [Google Scholar]
  • 36. Heurteaux C, Bertaina V, Widmann C, Lazdunski M (1993) K+ channel openers prevent global ischemia‐induced expression of c‐fos, c‐jun, heat shock protein, and amyloid beta‐protein precursor genes and neuronal death in rat hippocampus. Proc Natl Acad Sci USA 90:9431–9435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Heurteaux C, Guy N, Laigle C, Blondeau N, Duprat F, Mazzuca M et al. (2004) TREK‐1, a K(+) channel involved in neuroprotection and general anesthesia. EMBO J 23:2684–2695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Heurteaux C, Laigle C, Blondeau N, Jarretou G, Lazdunski M (2006) Alpha‐linolenic acid and riluzole treatment confer cerebral protection and improve survival after focal brain ischemia. Neuroscience 137:241–251. [DOI] [PubMed] [Google Scholar]
  • 39. Honore E (2007) The neuronal background K2P channels: focus on TREK1. Nat Rev Neurosci 8:251–261. [DOI] [PubMed] [Google Scholar]
  • 40. Honore E, Maingret F, Lazdunski M, Patel AJ (2002) An intracellular proton sensor commands lipid‐ and mechano‐gating of the K(+) channel TREK‐1. EMBO J 21:2968–2976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Huang L, Li B, Li W, Guo H, Zou F (2009) ATP‐sensitive potassium channels control glioma cells proliferation by regulating ERK activity. Carcinogenesis 30:737–744. [DOI] [PubMed] [Google Scholar]
  • 42. Hughes FM Jr, Cidlowski JA (1999) Potassium is a critical regulator of apoptotic enzymes in vitro and in vivo . Adv Enzyme Regul 39:157–171. [DOI] [PubMed] [Google Scholar]
  • 43. Jager H, Adelman JP, Grissmer S (2000) SK2 encodes the apamin‐sensitive Ca(2+)‐activated K(+) channels in the human leukemic T cell line, Jurkat. FEBS Lett 469:196–202. [DOI] [PubMed] [Google Scholar]
  • 44. Johnson RP, O’Kelly IM, Fearon IM (2004) System‐specific O2 sensitivity of the tandem pore domain K+ channel TASK‐1. Am J Physiol Cell Physiol 286:C391–C397. [DOI] [PubMed] [Google Scholar]
  • 45. Kaila K, Ransom BR (1998) pH and Brain Function. Wiley‐Liss: New York. [Google Scholar]
  • 46. Kananura C, Sander T, Rajan S, Preisig‐Muller R, Grzeschik KH, Daut J et al. (2002) Tandem pore domain K(+)‐channel TASK‐3 (KCNK9) and idiopathic absence epilepsies. Am J Med Genet 114:227–229. [DOI] [PubMed] [Google Scholar]
  • 47. Kim CJ, Cho YG, Jeong SW, Kim YS, Kim SY, Nam SW et al. (2004) Altered expression of KCNK9 in colorectal cancers. APMIS 112:588–594. [DOI] [PubMed] [Google Scholar]
  • 48. Kim D (2005) Physiology and pharmacology of two‐pore domain potassium channels. Curr Pharm Des 11:2717–2736. [DOI] [PubMed] [Google Scholar]
  • 49. Kim D, Cavanaugh EJ, Kim I, Carroll JL (2009) Heteromeric TASK‐1/TASK‐3 is the major oxygen‐sensitive background K+ channel in rat carotid body glomus cells. J Physiol 587(Pt 12):2963–2975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Kim DS, Kim JE, Kwak SE, Choi HC, Song HK, Kimg YI et al. (2007) Up‐regulated astroglial TWIK‐related acid‐sensitive K+ channel‐1 (TASK‐1) in the hippocampus of seizure‐sensitive gerbils: a target of anti‐epileptic drugs. Brain Res 1185:346–358. [DOI] [PubMed] [Google Scholar]
  • 51. Kim JE, Kwak SE, Kang TC (2009) Upregulated TWIK‐related acid‐sensitive K+ channel‐2 in neurons and perivascular astrocytes in the hippocampus of experimental temporal lobe epilepsy. Epilepsia 50:654–663. [DOI] [PubMed] [Google Scholar]
  • 52. Kindler CH, Pietruck C, Yost CS, Sampson ER, Gray AT (2000) Localization of the tandem pore domain K+ channel TASK‐1 in the rat central nervous system. Brain Res Mol Brain Res 80:99–108. [DOI] [PubMed] [Google Scholar]
  • 53. Kleinschnitz C, Meuth SG, Kieseier BC, Wiendl H (2007) Immunotherapeutic approaches in MS: update on pathophysiology and emerging agents or strategies 2006. Endocr Metab Immune Disord Drug Targets 7:35–63. [DOI] [PubMed] [Google Scholar]
  • 54. Koerner IP, Brambrink AM (2006) Brain protection by anesthetic agents. Curr Opin Anaesthesiol 19:481–486. [DOI] [PubMed] [Google Scholar]
  • 55. Koni PA, Khanna R, Chang MC, Tang MD, Kaczmarek LK, Schlichter LC, Flavella RA (2003) Compensatory anion currents in Kv1.3 channel‐deficient thymocytes. J Biol Chem 278:39443–39451. [DOI] [PubMed] [Google Scholar]
  • 56. Koo GC, Blake JT, Talento A, Nguyen M, Lin S, Sirotina A et al. (1997) Blockade of the voltage‐gated potassium channel Kv1.3 inhibits immune responses in vivo . J Immunol 158:5120–5128. [PubMed] [Google Scholar]
  • 57. Kunzelmann K (2005) Ion channels and cancer. J Membr Biol 205:159–173. [DOI] [PubMed] [Google Scholar]
  • 58. Lang F, Foller M, Lang KS, Lang PA, Ritter M, Gulbins E et al. (2005) Ion channels in cell proliferation and apoptotic cell death. J Membr Biol 205:147–157. [DOI] [PubMed] [Google Scholar]
  • 59. Lauritzen I, Blondeau N, Heurteaux C, Widmann C, Romey G, Lazdunski M (2000) Polyunsaturated fatty acids are potent neuroprotectors. EMBO J 19:1784–1793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Lauritzen I, De Weille JR, Lazdunski M (1997) The potassium channel opener (‐)‐cromakalim prevents glutamate‐induced cell death in hippocampal neurons. J Neurochem 69:1570–1579. [DOI] [PubMed] [Google Scholar]
  • 61. Lauritzen I, Zanzouri M, Honore E, Duprat F, Ehrengruber MU, Lazdunski M, Patel AJ (2003) K+‐dependent cerebellar granule neuron apoptosis. Role of task leak K+ channels. J Biol Chem 278:32068–32076. [DOI] [PubMed] [Google Scholar]
  • 62. Lee YM, Kim BJ, Chun YS, So I, Choi H, Kim MS, Park JW (2006) NOX4 as an oxygen sensor to regulate TASK‐1 activity. Cell Signal 18:499–507. [DOI] [PubMed] [Google Scholar]
  • 63. Leis JA, Bekar LK, Walz W (2005) Potassium homeostasis in the ischemic brain. Glia 50:407–416. [DOI] [PubMed] [Google Scholar]
  • 64. Leonard RJ, Garcia ML, Slaughter RS, Reuben JP (1992) Selective blockers of voltage‐gated K+ channels depolarize human T lymphocytes: mechanism of the antiproliferative effect of charybdotoxin. Proc Natl Acad Sci USA 89:10094–10098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Lesage F, Lazdunski M (2000) Molecular and functional properties of two‐pore‐domain potassium channels. Am J Physiol Renal Physiol 279:F793–F801. [DOI] [PubMed] [Google Scholar]
  • 66. Lesage F, Maingret F, Lazdunski M (2000) Cloning and expression of human TRAAK, a polyunsaturated fatty acids‐activated and mechano‐sensitive K(+) channel. FEBS Lett 471:137–140. [DOI] [PubMed] [Google Scholar]
  • 67. Lesage F, Terrenoire C, Romey G, Lazdunski M (2000) Human TREK2, a 2P domain mechano‐sensitive K+ channel with multiple regulations by polyunsaturated fatty acids, lysophospholipids, and Gs, Gi, and Gq protein‐coupled receptors. J Biol Chem 275:28398–28405. [DOI] [PubMed] [Google Scholar]
  • 68. Lewis A, Hartness ME, Chapman CG, Fearon IM, Meadows HJ, Peers C, Kemp PJ (2001) Recombinant hTASK1 is an O(2)‐sensitive K(+) channel. Biochem Biophys Res Commun 285:1290–1294. [DOI] [PubMed] [Google Scholar]
  • 69. Lewis RS (2001) Calcium signaling mechanisms in T lymphocytes. Annu Rev Immunol 19:497–521. [DOI] [PubMed] [Google Scholar]
  • 70. Lin CS, Boltz RC, Blake JT, Nguyen M, Talento A, Fischer PA et al. (1993) Voltage‐gated potassium channels regulate calcium‐dependent pathways involved in human T lymphocyte activation. J Exp Med 177:637–645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Linden AM, Aller MI, Leppa E, Vekovischeva O, Aitta‐Aho T, Veale EL et al. (2006) The in vivo contributions of TASK‐1‐containing channels to the actions of inhalation anesthetics, the alpha(2) adrenergic sedative dexmedetomidine, and cannabinoid agonists. J Pharmacol Exp Ther 317:615–626. [DOI] [PubMed] [Google Scholar]
  • 72. Linden AM, Sandu C, Aller MI, Vekovischeva OY, Rosenberg PH, Wisden W, Korpi ER (2007) TASK‐3 knockout mice exhibit exaggerated nocturnal activity, impairments in cognitive functions, and reduced sensitivity to inhalation anesthetics. J Pharmacol Exp Ther 323:924–934. [DOI] [PubMed] [Google Scholar]
  • 73. Lipton P (1999) Ischemic cell death in brain neurons. Physiol Rev 79:1431–1568. [DOI] [PubMed] [Google Scholar]
  • 74. Liu C, Au JD, Zou HL, Cotten JF, Yost CS (2004) Potent activation of the human tandem pore domain K channel TRESK with clinical concentrations of volatile anesthetics. Anesth Analg 99:1715–1722, table of contents. [DOI] [PubMed] [Google Scholar]
  • 75. Liu C, Cotten JF, Schuyler JA, Fahlman CS, Au JD, Bickler PE, Yost CS (2005) Protective effects of TASK‐3 (KCNK9) and related 2P K channels during cellular stress. Brain Res 1031:164–173. [DOI] [PubMed] [Google Scholar]
  • 76. Luedi PP, Dietrich FS, Weidman JR, Bosko JM, Jirtle RL, Hartemink AJ (2007) Computational and experimental identification of novel human imprinted genes. Genome Res 17:1723–1730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Maingret F, Lauritzen I, Patel AJ, Heurteaux C, Reyes R, Lesage F et al. (2000) TREK‐1 is a heat‐activated background K(+) channel. EMBO J 19:2483–2491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Maingret F, Patel AJ, Lesage F, Lazdunski M, Honore E (1999) Mechano‐ or acid stimulation, two interactive modes of activation of the TREK‐1 potassium channel. J Biol Chem 274:26691–26696. [DOI] [PubMed] [Google Scholar]
  • 79. Maingret F, Patel AJ, Lesage F, Lazdunski M, Honore E (2000) Lysophospholipids open the two‐pore domain mechano‐gated K(+) channels TREK‐1 and TRAAK. J Biol Chem 275:10128–10133. [DOI] [PubMed] [Google Scholar]
  • 80. Mayer ML, Westbrook GL, Guthrie PB (1984) Voltage‐dependent block by Mg2+ of NMDA responses in spinal cord neurones. Nature 309:261–263. [DOI] [PubMed] [Google Scholar]
  • 81. Meadows HJ, Randall AD (2001) Functional characterisation of human TASK‐3, an acid‐sensitive two‐pore domain potassium channel. Neuropharmacology 40:551–559. [DOI] [PubMed] [Google Scholar]
  • 82. Meuth SG, Aller MI, Munsch T, Schuhmacher T, Seidenbecher T, Kleinschnitz C et al. (2006) The contribution of TASK‐1‐containing channels to the function of dorsal lateral geniculate thalamocortical relay neurons. Mol Pharmacol 69:1468–1476. [DOI] [PubMed] [Google Scholar]
  • 83. Meuth SG, Bittner S, Meuth P, Simon OJ, Budde T, Wiendl H (2008) TWIK‐related acid‐sensitive K+ channel 1 (TASK1) and TASK3 critically influence T lymphocyte effector functions. J Biol Chem 283:14559–14570. [DOI] [PubMed] [Google Scholar]
  • 84. Meuth SG, Budde T, Kanyshkova T, Broicher T, Munsch T, Pape H‐C (2003) Contribution of TWIK‐related acid‐sensitive K+ channel 1 (TASK1) and TASK3 channels to the control of activity modes in thalamocortical neurons. J Neurosci 23:6460–6469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Meuth SG, Budde T, Kanyshkova T, Broicher T, Munsch T, Pape HC (2003) Contribution of TWIK‐related acid‐sensitive K+ channel 1 (TASK1) and TASK3 channels to the control of activity modes in thalamocortical neurons. J Neurosci 23:6460–6469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Meuth SG, Herrmann AM, Ip CW, Kanyshkova T, Bittner S, Weishaupt A et al. (2008) The two‐pore domain potassium channel TASK3 functionally impacts glioma cell death. J Neurooncol 87:263–270. [DOI] [PubMed] [Google Scholar]
  • 87. Meuth SG, Kanyshkov T, Melzer N, Bittner S, Kieseier BC, Budde T, Wiendl H (2008) Altered neuronal expression of TASK1 and TASK3 potassium channels in rodent and human autoimmune CNS inflammation. Neurosci Lett 446:133–138. [DOI] [PubMed] [Google Scholar]
  • 88. Meuth SG, Kleinschnitz C, Broicher T, Austinat M, Braeuninger S, Bittner S et al. (2009) The neuroprotective impact of the leak potassium channel TASK1 on stroke development in mice. Neurobiol Dis 33:1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Millar JA, Barratt L, Southan AP, Page KM, Fyffe RE, Robertson B, Mathie A (2000) A functional role for the two‐pore domain potassium channel TASK‐1 in cerebellar granule neurons. Proc Natl Acad Sci USA 97:3614–3618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Morokuma J, Blackiston D, Adams DS, Seebohm G, Trimmer B, Levin M (2008) Modulation of potassium channel function confers a hyperproliferative invasive phenotype on embryonic stem cells. Proc Natl Acad Sci USA 105:16608–16613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Mu D, Chen L, Zhang X, See LH, Koch CM, Yen C et al. (2003) Genomic amplification and oncogenic properties of the KCNK9 potassium channel gene. Cancer Cell 3:297–302. [DOI] [PubMed] [Google Scholar]
  • 92. Muhammad S, Aller MI, Maser‐Gluth C, Schwaninger M, Wisden W (2010) Expression of the KCNK3 potassium channel gene lessens the injury from cerebral ischemia, most likely by a general influence on blood pressure. Neuroscience 167:758–764. [DOI] [PubMed] [Google Scholar]
  • 93. Nadler JV (2003) The recurrent mossy fiber pathway of the epileptic brain. Neurochem Res 28:1649–1658. [DOI] [PubMed] [Google Scholar]
  • 94. Nakagawa Y, Numoto K, Yoshida A, Kunisada T, Ohata H, Takeda K et al. (2006) Chromosomal and genetic imbalances in synovial sarcoma detected by conventional and microarray comparative genomic hybridization. J Cancer Res Clin Oncol 132:444–450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Nakagawa Y, Yoshida A, Numoto K, Kunisada T, Wai D, Ohata N et al. (2006) Chromosomal imbalances in malignant peripheral nerve sheath tumor detected by metaphase and microarray comparative genomic hybridization. Oncol Rep 15:297–303. [PubMed] [Google Scholar]
  • 96. Negulescu PA, Shastri N, Cahalan MD (1994) Intracellular calcium dependence of gene expression in single T lymphocytes. Proc Natl Acad Sci USA 91:2873–2877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Nilius B, Wohlrab W (1992) Potassium channels and regulation of proliferation of human melanoma cells. J Physiol 445:537–548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Pancrazio JJ, Tabbara IA, Kim YI (1993) Voltage‐activated K+ conductance and cell proliferation in small‐cell lung cancer. Anticancer Res 13:1231–1234. [PubMed] [Google Scholar]
  • 99. Panyi G, Varga Z, Gaspar R (2004) Ion channels and lymphocyte activation. Immunol Lett 92:55–66. [DOI] [PubMed] [Google Scholar]
  • 100. Pardo LA, Del Camino D, Sanchez A, Alves F, Bruggemann A, Beckh S, Stuhmer W (1999) Oncogenic potential of EAG K(+) channels. EMBO J 18:5540–5547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Patel AJ, Honore E (2001) Molecular physiology of oxygen‐sensitive potassium channels. Eur Respir J 18:221–227. [DOI] [PubMed] [Google Scholar]
  • 102. Patel AJ, Honore E (2001) Properties and modulation of mammalian 2P domain K+ channels. Trends Neurosci 24:339–346. [DOI] [PubMed] [Google Scholar]
  • 103. Patel AJ, Honore E, Lesage F, Fink M, Romey G, Lazdunski M (1999) Inhalational anesthetics activate two‐pore‐domain background K+ channels. Nat Neurosci 2:422–426. [DOI] [PubMed] [Google Scholar]
  • 104. Patel AJ, Honore E, Maingret F, Lesage F, Fink M, Duprat F, Lazdunski M (1998) A mammalian two pore domain mechano‐gated S‐like K+ channel. EMBO J 17:4283–4290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Patel AJ, Lazdunski M (2004) The 2P‐domain K+ channels: role in apoptosis and tumorigenesis. Pflugers Arch 448:261–273. [DOI] [PubMed] [Google Scholar]
  • 106. Pei L, Wiser O, Slavin A, Mu D, Powers S, Jan LY, Hoey T (2003) Oncogenic potential of TASK3 (Kcnk9) depends on K+ channel function. Proc Natl Acad Sci USA 100:7803–7807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. Plant LD, Kemp PJ, Peers C, Henderson Z, Pearson HA (2002) Hypoxic depolarization of cerebellar granule neurons by specific inhibition of TASK‐1. Stroke 33:2324–2328. [DOI] [PubMed] [Google Scholar]
  • 108. Pocsai K, Kosztka L, Bakondi G, Gonczi M, Fodor J, Dienes B et al. (2006) Melanoma cells exhibit strong intracellular TASK‐3‐specific immunopositivity in both tissue sections and cell culture. Cell Mol Life Sci 63:2364–2376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109. Pottosin II, Bonales‐Alatorre E, Valencia‐Cruz G, Mendoza‐Magana ML, Dobrovinskaya OR (2008) TRESK‐like potassium channels in leukemic T cells. Pflugers Arch 456:1037–1048. [DOI] [PubMed] [Google Scholar]
  • 110. Prakriya M, Lewis RS (2003) CRAC channels: activation, permeation, and the search for a molecular identity. Cell Calcium 33:311–321. [DOI] [PubMed] [Google Scholar]
  • 111. Rader RK, Kahn LE, Anderson GD, Martin CL, Chinn KS, Gregory SA (1996) T cell activation is regulated by voltage‐dependent and calcium‐activated potassium channels. J Immunol 156:1425–1430. [PubMed] [Google Scholar]
  • 112. Rakhade SN, Jensen FE (2009) Epileptogenesis in the immature brain: emerging mechanisms. Nat Rev 5:380–391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Riener MO, Nikolopoulos E, Herr A, Wild PJ, Hausmann M, Wiech T et al. (2008) Microarray comparative genomic hybridization analysis of tubular breast carcinoma shows recurrent loss of the CDH13 locus on 16q. Hum Pathol 39:1621–1629. [DOI] [PubMed] [Google Scholar]
  • 114. Rouzaire‐Dubois B, Dubois JM (1998) K+ channel block‐induced mammalian neuroblastoma cell swelling: a possible mechanism to influence proliferation. J Physiol 510(Pt 1):93–102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115. Ruf N, Bahring S, Galetzka D, Pliushch G, Luft FC, Nurnberg P et al. (2007) Sequence‐based bioinformatic prediction and QUASEP identify genomic imprinting of the KCNK9 potassium channel gene in mouse and human. Hum Mol Genet 16:2591–2599. [DOI] [PubMed] [Google Scholar]
  • 116. Rusznak Z, Bakondi G, Kosztka L, Pocsai K, Dienes B, Fodor J et al. (2008) Mitochondrial expression of the two‐pore domain TASK‐3 channels in malignantly transformed and non‐malignant human cells. Virchows Arch 452:415–426. [DOI] [PubMed] [Google Scholar]
  • 117. Saez‐Hernandez L, Peral B, Sanz R, Gomez‐Garre P, Ramos C, Ayuso C, Serratosa JM (2003) Characterization of a 6p21 translocation breakpoint in a family with idiopathic generalized epilepsy. Epilepsy Res 56:155–163. [DOI] [PubMed] [Google Scholar]
  • 118. Santella L, De Riso L, Gragnaniello G, Kyozuka K (1998) Separate activation of the cytoplasmic and nuclear calcium pools in maturing starfish oocytes. Biochem Biophys Res Commun 252:1–4. [DOI] [PubMed] [Google Scholar]
  • 119. Shu J, Jelinek J, Chang H, Shen L, Qin T, Chung W et al. (2006) Silencing of bidirectional promoters by DNA methylation in tumorigenesis. Cancer Res 66:5077–5084. [DOI] [PubMed] [Google Scholar]
  • 120. Sirois JE, Lei Q, Talley EM, Lynch C 3rd, Bayliss DA (2000) The TASK‐1 two‐pore domain K+ channel is a molecular substrate for neuronal effects of inhalation anesthetics. J Neurosci 20:6347–6354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Skryma RN, Prevarskaya NB, Dufy‐Barbe L, Odessa MF, Audin J, Dufy B (1997) Potassium conductance in the androgen‐sensitive prostate cancer cell line, LNCaP: involvement in cell proliferation. Prostate 33:112–122. [DOI] [PubMed] [Google Scholar]
  • 122. Steriade M, Jones EG, McCormick DA (1997) Thalamus, 1st edn. Elsevier: Amsterdam. [Google Scholar]
  • 123. Stutzin A, Hoffmann EK (2006) Swelling‐activated ion channels: functional regulation in cell‐swelling, proliferation and apoptosis. Acta Physiol (Oxf) 187:27–42. [DOI] [PubMed] [Google Scholar]
  • 124. Szewczyk A, Jarmuszkiewicz W, Kunz WS (2009) Mitochondrial potassium channels. IUBMB Life 61:134–143. [DOI] [PubMed] [Google Scholar]
  • 125. Talley EM, Sirois JE, Lei Q, Bayliss DA (2003) Two‐pore‐Domain (KCNK) potassium channels: dynamic roles in neuronal function. Neuroscientist 9:46–56. [DOI] [PubMed] [Google Scholar]
  • 126. Taverna S, Tkatch T, Metz AE, Martina M (2005) Differential expression of TASK channels between horizontal interneurons and pyramidal cells of rat hippocampus. J Neurosci 25:9162–9170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127. Taylor KH, Pena‐Hernandez KE, Davis JW, Arthur GL, Duff DJ, Shi H et al. (2007) Large‐scale CpG methylation analysis identifies novel candidate genes and reveals methylation hotspots in acute lymphoblastic leukemia. Cancer Res 67:2617–2625. [DOI] [PubMed] [Google Scholar]
  • 128. Tsai SJ (2008) Sipatrigine could have therapeutic potential for major depression and bipolar depression through antagonism of the two‐pore‐domain K+ channel TREK‐1. Medical Hypotheses 70:548–550. [DOI] [PubMed] [Google Scholar]
  • 129. Valencia‐Cruz G, Shabala L, Delgado‐Enciso I, Shabala S, Bonales‐Alatorre E, Pottosin II, Dobrovinskaya OR (2009) Kbg and Kv1.3 channels mediate potassium efflux in the early phase of apoptosis in Jurkat T lymphocytes. Am J Physiol Cell Physiol. [DOI] [PubMed] [Google Scholar]
  • 130. Voloshyna I, Besana A, Castillo M, Matos T, Weinstein IB, Mansukhani M et al. (2008) TREK‐1 is a novel molecular target in prostate cancer. Cancer Res 68:1197–1203. [DOI] [PubMed] [Google Scholar]
  • 131. Wang Z (2004) Roles of K+ channels in regulating tumour cell proliferation and apoptosis. Pflugers Arch 448:274–286. [DOI] [PubMed] [Google Scholar]
  • 132. Whitfield JF, Bird RP, Chakravarthy BR, Isaacs RJ, Morley P (1995) Calcium‐cell cycle regulator, differentiator, killer, chemopreventor, and maybe, tumor promoter. J Cell Biochem Suppl 22:74–91. [PubMed] [Google Scholar]
  • 133. Woodfork KA, Wonderlin WF, Peterson VA, Strobl JS (1995) Inhibition of ATP‐sensitive potassium channels causes reversible cell‐cycle arrest of human breast cancer cells in tissue culture. J Cell Physiol 162:163–171. [DOI] [PubMed] [Google Scholar]
  • 134. Wulff H, Calabresi PA, Allie R, Yun S, Pennington M, Beeton C, Chandy KG (2003) The voltage‐gated Kv1.3 K(+) channel in effector memory T cells as new target for MS. J Clin Invest 111:1703–1713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135. Wulff H, Zhorov BS (2008) K+ channel modulators for the treatment of neurological disorders and autoimmune diseases. Chem Rev 108:1744–1773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Wyatt CN, Peers C (2009) Hetero or homo, hypoxia has them all. J Physiol 587(Pt 12):2717–2718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Xiao Z, Deng PY, Rojanathammanee L, Yang C, Grisanti L, Permpoonputtana K et al. (2009) Noradrenergic depression of neuronal excitability in the entorhinal cortex via activation of TREK‐2 K+ channels. J Biol Chem 284:10980–10991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Xiong ZQ, Stringer JL (2000) Extracellular pH responses in CA1 and the dentate gyrus during electrical stimulation, seizure discharges, and spreading depression. J Neurophysiol 83:3519–3524. [DOI] [PubMed] [Google Scholar]
  • 139. Xu D, Wang L, Dai W, Lu L (1999) A requirement for K+‐channel activity in growth factor‐mediated extracellular signal‐regulated kinase activation in human myeloblastic leukemia ML‐1 cells. Blood 94:139–145. [PubMed] [Google Scholar]
  • 140. Yost CS, Oh I, Eger EI 2nd, Sonner JM (2008) Knockout of the gene encoding the K(2P) channel KCNK7 does not alter volatile anesthetic sensitivity. Behav Brain Res 193:192–196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141. Zheng H, Nam JH, Pang B, Shin DH, Kim JS, Chun YS et al. (2009) Identification of the large‐conductance background K+ channel in mouse B cells as TREK‐2. Am J Physiol Cell Physiol 297:C188–C197. [DOI] [PubMed] [Google Scholar]
  • 142. Zhou Q, Kwan HY, Chan HC, Jiang JL, Tam SC, Yao X (2003) Blockage of voltage‐gated K+ channels inhibits adhesion and proliferation of hepatocarcinoma cells. Int J Mol Med 11:261–266. [PubMed] [Google Scholar]

Articles from Brain Pathology are provided here courtesy of Wiley

RESOURCES