Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Feb 1.
Published in final edited form as: J Cardiovasc Pharmacol. 2013 Feb;61(2):102–112. doi: 10.1097/FJC.0b013e318279ba20

Endothelial Small- and Intermediate-Conductance KCa Channels: An Update on Their Pharmacology and Usefulness as Cardiovascular Targets

Heike Wulff *, Ralf Köhler †,§
PMCID: PMC3565027  NIHMSID: NIHMS422298  PMID: 23107876

Abstract

Most cardiovascular researchers are familiar with intermediate-conductance KCa3.1 and small-conductance KCa2.3 channels because of their contribution to endothelium-derived hyperpolarization (EDH). However, to immunologists and neuroscientists these channels are primarily known for their role in lymphocyte activation and neuronal excitability. KCa3.1 is involved in the proliferation and migration of T cells, B cell, mast cells, macrophages, fibroblasts and dedifferentiated vascular smooth muscle cells and is, therefore, being pursued as a potential target for use in asthma, immunosuppression, and fibroproliferative disorders. In contrast, the three KCa2 channels (KCa2.1, KCa2.2 and KCa2.3) contribute to the neuronal medium afterhyperpolarization and, depending on the type of neuron, are involved in determining firing rates and frequencies or in regulating bursting. KCa2 activators are accordingly being studied as potential therapeutics for ataxia and epilepsy while KCa2 channel inhibitors like apamin have long been known to improve learning and memory in rodents. Given this background, we review the recent discoveries of novel KCa3.1 and KCa2.3 modulators and critically assess the potential of KCa activators for the treatment of diabetes and cardiovascular diseases by improving endothelium-derived hyperpolarizations.

Keywords: endothelium derived hyperpolarization, blood pressure, lymphocyte activation, action potential, afterhyperpolarization, small/intermediate-conductance calcium-activated K+ channel, KCa3.1, KCa2.3, gating modulation

INTRODUCTION

The human genome contains eight Ca2+-activated K+ channels, which can be divided into two well defined but only distantly related groups based on their genetic relationship, single channel conductances and molecular mechanisms of calcium-“sensing”1. The first group consists of the large-conductance Ca2+-activated KCa1.1 channel, also known as BK or Maxi-K, and the related KCa4.1 (Slack), KCa4.2 (Slick) and KCa5.1 (Slo3) channels, which are activated by increases in intracellular Na+ and/or Cl or by alkalization. The second group of channels, the KCa2/3 group, comprises the three small-conductance KCa2 channels, KCa2.1 (SK1), KCa2.2 (SK2) and KCa2.3 (SK3), as well as the intermediate-conductance KCa3.1 (IK1) channel, which all four share the same calmodulin-mediated gating mechanism. In contrast to KCa1.1, which directly binds Ca2+, KCa2 and KCa3.1 channels are constitutively associated with calmodulin2,3, which functions as their Ca2+-sensing β-subunit, and induces Ca2+-dependent channel opening with reported EC50 values ranging from 95 to 350 nM for KCa3.1 and from 320 to 750 for the KCa2 channels1. Another distinguishing biophysical characteristic of the KCa2/3 channels is that their S4 segment contains fewer charged residues than the S4 segment of KCa1.1 or Kv channels. This results in a lack of voltage-dependence, which enables these channels to remain open at negative membrane potentials and to thus hyperpolarize the membrane towards values near the K+ equilibrium potential of −89 mV. KCa3.1 and KCa2 channels are accordingly expressed in cells that need to be able to hyperpolarize in order to 1) sustain Ca2+ influx through inward rectifier Ca2+ channels like CRAC (Ca2+ release activated Ca2+ channel) during cellular activation and proliferation; 2) to regulate firing frequency by preventing an untimely or premature action potential initiation; or 3) to pass on hyperpolarization through gap junctions.

Modulation of Ca2+ influx

The first function, modulation of Ca2+ influx during cellular activation and proliferation together with a role in erythrocyte volume regulation and fluid and electrolyte secretion, which will not be discussed in this review and which the interested reader can find reviewed elsewhere4,5, is primarily fulfilled by KCa3.1. The channel is accordingly widely expressed throughout the body and found in cells of the hematopoietic system (i.e erythrocytes, platelets, T cells, B cells, mast cells, monocytes/macrophages, microglia); epithelial tissues in the lung and gastrointestinal tracts, as well as in vascular endothelial cells, fibroblasts and proliferating neointimal vascular smooth muscle cells510. In these cells KCa3.1 regulates Ca2+ entry and thereby modulates Ca2+ signaling as depicted in the cartoon in Figure 1. Following increases in intracellular Ca2+ after, for example T cell receptor engagement in T cells or FGF receptor activation in fibroblasts, KCa3.1 generates a negative membrane potential, which sustains calcium entry through CRAC or transient receptor potential (TRP) channels11,12. KCa3.1 thus drives cellular proliferation in T and B cells, fibroblasts and neointimal vascular smooth muscle cells, while KCa3.1 blockers accordingly inhibit the proliferation of these cell types710.

FIGURE 1.

FIGURE 1

Cartoon of the physiological role of KCa3.1. The channel is activated by increases in intracellular Ca2+ following Ca2+ release from the ER (endoplasmatic reticulum), and/or Ca2+ influx through inward-rectifier Ca2+ channels like CRAC (Ca2+ release activated Ca2+ channel) or TRP (transient receptor potential) channels. PLC; phospholipase C; IP3 inositol-triphosphate; CAM; calmodulin.

Regulation of firing frequency

In contrast to KCa3.1, which following its cloning in 19971315 was initially assumed to be completely absent from neuronal tissue but has recently been described to be expressed in cerebellar Purkinje cells where it apparently is involved in suppressing low frequency parallel fiber input16, KCa2 channels are widely expressed throughout the CNS and are probably best known for their contribution to the medium afterhyperpolarization (mAHP) in neurons17,18. While KCa2.1 and KCa2.2 are particularly enriched in the cortex and the hippocampus, KCa2.3 is most prominent in subcortical areas like the striatum, thalamus and monaminergic nuclei17,19,20. By contributing to the mAHP, KCa2 channels regulate neuronal firing frequency for example in hippocampal CA1 pyramidal neurons17,2123, in the subthalamic nucleus24, as well as in Purkinje neurons25. Further deserving mention is the overlapping expression of KCa2.3 and dopamine synthesizing enzymes26. The autonomous pacemaker and bursting activity of dopaminergic neurons in the substantia nigra has accordingly been shown to be largely regulated by KCa2.327,28. In addition to these roles in regulating somatic excitability, KCa2 channels have also been found to be localized in the postsynaptic membrane of glutamatergic synapses, where their activation and regulated trafficking modulates synaptic plasticity, thereby affecting learning and memory (for a recent excellent review please see Adelman 201218).

KCa2 channel activity intimately follows the increases and decreases in free Ca2+ concentration originating from nearby Ca2+ sources (Figure 2). Under “normal” conditions, when neuronal [Ca2+]i is not pathologically increased, the Ca2+ for the activation of KCa2 channels has been shown to come from several different sources: 1) voltage-gated Ca2+ channels; 2) Ca2+ permeable ligand-gated ion channels such as NMDA or nicotinic acetylcholine receptors (nAChR); 3) IP3 mediated Ca2+ release from the ER downstream of G-protein coupled receptors; or 4) Ca2+-induced Ca2+ release from ryanodine receptors18. For example, in CA1 hippocampal neurons KCa2 channels and L-type Cav channels have long been known to be located in microdomains in which they are only 50–100 nm apart29, while in cerebral Purkinje neurons KCa2 channels are coupled to P/Q type channels30. In contrast, in dendritic spines of hippocampal or amygdala neurons, KCa2 channels seem to form feedback loops with NMDA receptors and are located in close proximity31,32.

FIGURE 2.

FIGURE 2

Cartoon depicting the different Ca2+ sources involved in KCa2 channel activation. In neurons KCa2 channels are often localized in close proximity to NMDA receptors or to voltage-gated Ca2+ channels (Cav), which can induce Ca2+-induced Ca2+ release from ryanodine receptors (RyR). KCa2 channels can further be activated through G-protein induced Ca2+-release from inositol-triphosphate (IP3) receptors and/or Ca2+ influx through TRP (transient receptor potential) channels.

Outside of the nervous system, KCa2.3 channels are also expressed in the liver, where they are believed to be involved in metabolic stress responses33, and urinary bladder smooth muscle, where they play an important role in determining excitability and contractility34. Since KCa2 channel inhibition with apamin has been reported to prolong action potential duration in both human and mouse atrial myocytes35 and KCa2.3 variants have been associated with lone atrial fibrillation36, KCa2 channels are further increasingly being recognized as contributors to cardiac excitability and are currently being investigated as novel targets for atrial fibrillation37.

Endothelium-myocyte coupling

Although KCa3.1 and KCa2.3 typically have a very different tissue distribution (see above), one place, where they are expressed together, is vascular endothelium38. This co-expression was initially simply viewed as redundancy but it is becoming increasingly clear that the two endothelial KCa channels are located in spatially distinct microdomains and apparently participate in different signaling pathways38,39. While KCa3.1 channels are predominantly localized to endothelial cell projections traversing the internal elastic lamina and are activated by Ca2+ released from the ER in response to stimulation of muscarinic acetylcholine receptors and perhaps other GPCRs40,41, KCa2.3 is found at inter-endothelial junctions and has been shown co-localize in caveolae with TRP channels42 and, presumably, also inward rectifier K+ channels43. In this location KCa2.3 has been suggested to sense local Ca2+-increases produced by mechanical deformation during shear stress stimulation44 and to amplify endothelial hyperpolarization by relieving endogenous blockade of inward rectifiers45. For a more detailed review and some excellent cartoons depicting KCa3.1 and KCa2.3 localization and signaling the interested reader is referred to a number of recent reviews on EDH38,39,46,47. However, even though KCa3.1 and KCa2.3 seem to be part of different signal transduction pathways responding to different upstream stimuli, the activation of either channel initiates hyperpolarization and subsequent endothelium derived hyperpolarization (EDH) mediated vasodilatory responses in which the endothelial hyperpolarization spreads to the underlying vascular smooth muscle cell layer, closes voltage-gated calcium channels, and finally produces relaxation and vasodilation38,39,4648. The relevance of this KCa3.1/KCa2.3 EDH system for systemic cardiovascular regulation is demonstrated by the observation of a higher blood pressure in KCa3.1 and/or KCa2.3 deficient mice44. Even though the deficiency of either channel increases mean arterial blood pressure (MAP) by 7–9 mmHg, absence of KCa3.1 appears to increase MAP because of a higher systolic pressure (SP) and higher pulse pressure (PP) at a normal diastolic pressure (DP), while KCa2.3-deficiency increases MAP by increasing both SP and DP with no change in PP (Table 1, for extensive review see49). Combined deficiency increased systolic and diastolic pressure as well as PP, but, surprisingly, did not increase MAP further suggesting compensation by other mechanisms (Table 1). Interestingly, systemic challenge of the mice with the NO-synthase inhibitor L-NAME, increased MAP in all mice by 6–11 mmHg and lowered heart rate suggesting a conserved ability to produce NO and intact baroreceptor functions (Table 1). An unexpected finding in these animals was that the higher MAP in either KO-mice was more pronounced during activity at night44, suggesting a role of the endothelial KCa channels and the EDH system for adjusting blood pressure during exercise.

Table 1.

Aggravation of hypertension in KCa3.1/KCa2.3-deficient mice by Inhibition of NO-synthesis.

genotype n SP
before
SP
w L-NAME
DP
before
DP
w L-
NAME
MAP
before
MAP
w L-NAME
HR
before
HR
w L-NAME
KCa3.1+/+/KCa2.3+/+(wt) 6 111 ± 2 124 ± 2# 83 ± 2 90 ± 2# 97 ± 1 107 ± 2# 607 ± 16 556 ± 10#
KCa3.1+/+/ KCa2.3T/Tdox 6 125 ± 3* 138 ± 4# 94 ± 1* 102 ± 3#* 110 ± 2* 119 ± 3#* 583 ± 20 528 ± 19#
KCa3.1−/−/KCa2.3+/+ 4 124 ± 4* 134 ± 2# 91 ± 3 95 ± 3# 108 ± 3* 114 ± 2# 625 ± 20 543 ± 37#
KCa3.1−/−/ KCa2.3T/Tdox 6 122 ± 3* 135 ± 4# 91 ± 2* 99 ± 4# 107 ± 2* 118 ± 4# 565 ± 13 528 ± 18

SP = systolic blood pressure; DP = diastolic blood pressure; MAP = mean arterial pressure; HR = heart rate; PP = pulse pressure; Dox = doxycycline to turn off KCa2.3 expression; Telemetric data are 24-hrs means ± SEM.

*

P<0.05 vs. wild-type, One-way ANOVA and Bonferroni post-hoc test.

#

P<0.05,

24 hrs L-NAME (50 µg/ml drinking water) vs. before treatment, paired Student´s t-test.

PHARMACOLOGY

In contrast to many other ion channels, KCa3.1 and KCa2 channels have a relatively well-developed pharmacology and pharmacological tool compounds have accordingly greatly contributed to elucidating the role of these channels in EDH-type dilator responses and other physiological functions. Since KCa channel pharmacology has previously been reviewed by us in great detail and put into its historical perspective50, we will here only briefly mention the most commonly used modulators (Figure 3) and reserve more detailed discussions to compounds identified in the last three years.

FIGURE 3.

FIGURE 3

Structures and potencies of commonly used or recently developed KCa3.1 and KCa2 modulators.

Classical blockers

The “classical” blockers for KCa channels are, of course, charybdotoxin for KCa3.1 and apamin for the KCa2 channels50. The larger charybdotoxin is a typical α-KTX scorpion toxin that anchors itself by two salt-bridges and inserts a central lysine residue into the selectivity filter of KCa3.1 and thus occludes the channel’s outer vestibule51. However, charybdotoxin never was an ideal KCa3.1 blocker because it also inhibits KCa1.1 (BK) and the voltage-gated Kv1.3 channels, both cross-reactivities, which initially caused confusion concerning the role of KCa3.1 in T cells and in vascular endothelium. Since the description of the clotrimazole derivatives TRAM-3452 and ICA-1704353, which entered clinical trials for sickle cell anemia under the name senicapoc54, charybdotoxin has been largely replaced as a pharmacological tool compound by these more KCa3.1 selective small molecule blockers. A new, structurally different KCa3.1 blocker is the benzothiazinone NS6180, which was recently identified in a Ca2+-activated Tl+ influx based high-throughput screen at NeuroSearch A/S in Denmark55. NS6180 blocks KCa3.1 with an IC50 of 11 nM by binding to the same two residues, Thr250 and Val275, in the inner pore as TRAM-34 and clotrimazole56. Up to a concentration of 1 µM NS6180 and TRAM-34 are very selective KCa3.1 blockers and show no effect on the T-cell Ca2+ channel CRAC consisting of Orai-1 and STIM1 and a range of Kv, KCa, Nav, and TRP channels55. However, at 10 µM both compounds start to significantly affect other channels. While NS6180 inhibits KCa1.1 (BK), Kv1.3 and Kv11.1 (hERG) by more than 50%, TRAM-34 blocks Kv1.3, Kv1.4, Kv7.2+Kv7.3 and Nav1.4 at 10 µM55. Both TRAM-34 and the new NS6180 should therefore not be used at concentrations above 1 µM, especially since both compounds also have very limited solubility around 10 µM and tend to quickly precipitate in aqueous solutions.

As mentioned above, the most widely used KCa2 blocker is the 18-amino acid honey bee venom peptide apamin50, which is a remarkably selective blocker of KCa2 channels, with KCa2.2 being the most sensitive channel (IC50 ~200 pM) and KCa2.1 and KCa2.3 being slightly less sensitive (see Table in Figure 3). Although initially assumed to inhibit KCa2 channels by simply occluding the pore similar to the larger scorpion and snake toxins, apamin more recently has been shown to inhibit KCa2 currents via an allosteric mechanism involving an outer pore histidine57 and residues in the S3-S4 extracellular loop58. The same type of more allosteric blocking mechanism probably applies to the permanently charged bis-quinolinium cyclophane UCL 168459 and its derivatives, which were designed to mimic the charges of apamin, and which are also of remarkable potency and selectivity (for a review see50). A less potent but highly KCa2.2 selective blocker is Lei-Dab7, a derivative of the scorpion toxin leiurotoxin I, in which one residue in a crucial signature motif is replaced by the unnatural amino acid diaminobutonoic acid60.

Positive and negative gating modulators

In addition to these “straight” blockers, which are either binding in the outer pore like the peptide toxins or in the inner pore like the TRAMs and NS6180, KCa2/3 channels also possess positive and negative gating-modulators (Figure 3), which left-shift or right-shift the calcium-response curve of these calcium/calmodulin gated channels and apparently render the channels more or less calcium-sensitive. Interestingly, gating-modulation of KCa3.1 and KCa2 channels seems to be possible at several points along the chain of molecular events leading from Ca2+ binding to the calmodulin at the C-terminus to eventual opening of the physical gate of these channels, which seems to be deeply buried in the inner pore vestibule, close to or even overlapping with the K+ selectivity filter6164. While benzimidazolone-type KCa activators like EBIO, the structurally related oxime NS30965 or the benzothiazole SKA-3166, do not differentiate well between KCa2 and KCa3.1 channels, which is in keeping with their binding site being located in the C-terminal region, close to or at the calmodulin binding domain22, the negative KCa2 channel gating-modulator NS8593 and the KCa2.1 selective (−)CM-TMPF have been found to interact with positions deep within the inner pore vestibule67,68, where the gate of these channels is located and accordingly exhibit subtype selectivity. The negative gating modulator NS8593, which interacts with Ser507 in the pore-loop near the selectivity filter and A532 in an adjacent position in S6, exerts no effect on KCa3.1 channels but inhibits all three KCa2 channels at submicromolar concentrations67. The very recently described [1,2,4]triazolo[1,5-a]pyrimidine (−)CM-TMPF and the structurally related (−)B-TMPF interact with Ser293 in S5 and act as KCa2.1 selective negative and positive gating modulators with IC50 or EC50 values of 24 and 31 nM68. Both compounds are 40 to 100 times more potent than their respective (+)-enantiomers and display 10–20-fold selectivity over KCa2.2, KCa2.3 and KCa3.1, properties which make them valuable tool compounds for probing the currently not well defined physiological or pathophysiological role of KCa2.1. Another new and probably in the future very useful KCa2 channel modulator is the CyPPA derivative NS13001, which activates KCa2.2 and KCa2.3 with EC50 values of 2 µM and 140 nM, but has no effects on KCa2.1 and KCa3.169.

Of these compounds, the ones that have so far most commonly been used in vivo are apamin (which interestingly is able to cross the blood-brain barrier and induces seizures following intraperitoneal application70) and NS8593 as KCa2 channel inhibitors, TRAM-34 and senicapoc as KCa3.1 blockers, and SKA-31 as a mixed KCa2/3 channel activator, with ~10-fold selectivity for KCa3.166. Despite its high potency, NS309 is unfortunately not suitable for in vivo use due to its extremely short half-life and its 1 µM IC50 for Kv11.1 (hERG)65.

CLINICAL AND PRECLINICAL EXPERIENCE WITH KCa3.1 BLOCKERS

Past experiences

Apart from a single nucleotide polymorphism associated with ileal Crohn’s disease in an Australian population71, KCa3.1 mutations have so far not been described to be involved in human diseases. Nevertheless, KCa3.1 blockade constitutes a relatively-well validated therapeutic approach for immunosuppression and for curbing vascular smooth muscle cell and fibroblast proliferation12. The oldest indication for KCa3.1 blockers is prevention of erythrocyte dehydration in sickle cell disease through inhibition of the so-called ‘Gàrdos channel’, the erythrocyte KCa3.1 channel. Early proof-of-concept studies from Carlo Brugnara’s group at the Children’s Hospital in Boston demonstrated that the unselective KCa3.1 blocker clotrimazole reduced erythrocyte dehydration in a transgenic mouse model72 and in 5 patients with sickle cell disease73. Following up on this, Icagen Inc. advanced the more selective senicapoc53 into clinical trials and reported that the compound significantly reduced hemolysis and increased hemoglobin levels in a 12-week, multicenter, randomized double-blind Phase-2 study54. However, in a subsequent Phase-3 study, which was designed to compare the rate of acute vaso-occlusive pain crisis occurring in sickle cell disease patients, senicapoc failed to reduce this desired clinical endpoint and was terminated early, despite patients in the senicapoc group again showing improvements in anaemia and haemolysis54. Since senicapoc was safe and well tolerated in these studies, Icagen Inc. next explored asthma as a potential therapeutic indication based on a large body of work demonstrating that KCa3.1 is expressed in various cell types involved in the pathogenesis of asthma (mast cell, macrophages, fibroblasts, airway epithelium and airway smooth muscle cells)74 and that KCa3.1 blockade inhibits mast cell degranulation and migration75,76 as well as airway smooth muscle proliferation9,77. Following an initial evaluation in allergen induced asthma in sheep78, senicapoc demonstrated encouraging results in a small Phase-2 allergen challenge study in patients with allergic asthma. However, in a second Phase-2 trial in exercise induced asthma the compound failed to demonstrate any significant improvement in lung function12.

KCa3.1 blockade and/or disruption of the KCa3.1 gene has further been found to ameliorate various autoimmune and cardiovascular disease models through a combination of inhibiting immune cell proliferation, infiltration and cytokine production as well as dampening smooth muscle cell, fibroblast and vascular endothelial cell proliferation. For example, TRAM-34 or senicapoc prevent MOG induced autoimmune encephalomyelitis79, anti-collagen antibody-induced arthritis80, and trinitrobenzene sulfonic acid-induced colitis in mice81, renal fibrosis following unilateral uretral obstruction in mice and rats82, angiogenesis in the mouse matrigel plug assay83, atherosclerosis development in ApoE−/− mice84, as well as angioplasty induced intimal smooth muscle hyperplasia (restenosis) in rats and pigs9,85. KCa3.1 blockade has further been found to reduce microglia activation86 and thus curb inflammatory responses and reduce neuronal damage in models of ischemic stroke87, traumatic brain injury88, optic nerve transection89, and traumatic spinal cord injury90.

Future potential

Despite the so far disappointing clinical trial results with KCa3.1 blockers, KCa3.1 remains an attractive pharmacological target for a variety of indications including postangioplasty restenosis, atherosclerosis, inflammatory bowel disease, and possibly neuroinflammation in the context of stroke, multiple sclerosis and Alzheimer’s disease91. Even asthma should probably not yet be dismissed as an indication since the performed studies were certainly not long enough to determine whether KCa3.1 blockers can prevent airway remodeling as would be expected from their inhibitory effects on airway smooth muscle and fibroblast proliferation74. However, given the important role that KCa3.1 channels play in EDH responses, proposing to develop KCa3.1 blockers for any of the above named indications of course raises the question of whether KCa3.1 blockers will increase blood pressure, particularly when considering the higher blood pressure in the KCa3.1 and/or KCa2.3-deficient mice as outlined above. Pharmacological inhibition of KCa3.1, in contrast, has not been observed to raise blood pressure in mice84 or in over 500 human volunteers and patients taking senicapoc for up to two years92, suggesting that combined blockade of endothelial KCa3.1 and KCa2.3 channels might be necessary to significantly raise blood pressure in humans. These findings, of course, do not discount that KCa3.1 blockade might not lead to increases in blood pressure and associated cardiovascular risk in situations like diabetes, hypertension and obesity, where EDH responses are known to be compromised39,93. Moreover, KCa3.1 blockers may worsen the vascular dementia and cerebrovascular ischemia in aged patients. It, however, should be kept in mind that the beneficial anti-inflammatory and anti-atherosclerotic effects of KCa3.1 blockers may outweigh the impairments in endothelial responses and thus decrease cardiovascular risk overall. Taken together, the potential therapeutic usage of KCa3.1 blockers should therefore be carefully considered on the basis of a patient’s individual cardiovascular risk profile, the intended indication and side-effect profiles of the therapeutic alternatives. For example, in situations where the alternatives are cyclosporine or glucocorticoids, both medications that are diabetogenic and induce hypertension and/or kidney toxicity, a KCa3.1 blocker might be viewed very favorably. However, it will require longer-term animal studies and ideally clinical trials to ultimately assess cardiovascular risks of KCa3.1 blockers.

IN VIVO EXPERIENCE WITH KCa2 MODULATORS

As described above, KCa2 channels play an important role in controlling neuronal excitability and firing frequency by determining the magnitude of the mAHP. Pharmacological modulation of KCa channels, therefore, offers the opportunity to significantly affect neuronal activity. While KCa2 channel blockers like apamin induce high frequency bursts in neurons that normally exhibit regular tonic action potential firing30 and elicit seizures in rodents94, KCa2 channel activators like EBIO reduce neuronal firing rates and suppress epileptiform activity in cultured hippocampal neurons22,23,95, suggesting that KCa2 channel activators might be useful for the treatment of CNS disorders that are characterized by hyperexcitability such as epilepsy and ataxia50. This therapeutic hypothesis is supported by the finding that transgenic mice overexpressing SK3-1B, a dominant negative alternative transcript of KCa2.3, which suppresses KCa2 channels, exhibit severe ataxia with in-coordination, tremor and altered gate due to increased excitability of their deep cerebellar neurons96. Conversely, acute KCa2 channel activation with SKA-31 partially corrects abnormal Purkinje cell firing and improves motor function in SCA3 mice97, a model of spinocerebellar ataxia type 3, while 3 weeks of oral administration of NS13001 has recently been shown to improve performance of aging SCA2 mice (a model of spinocerebellar ataxia type 2) in motor coordination assays and to reduce Purkinje cell degeneration69. Additional evidence supporting KCa2 channels as targets for ataxia comes from the observation that direct infusion of EBIO into the cerebellum significantly improves the motor coordination deficits in ducky and the dyskinesia and ataxia in tottering mice98, both harboring mutations in P/Q-type Cav channels leading to loss of precision in Purkinje cell firing. EBIO has further been reported to reduce electroshock and pentylenetetrazole induced seizures99 demonstrating that KCa2 activators are indeed anticonvulsants in vivo, a finding that has been recently confirmed by our own laboratory with SKA-31 in collaboration with the NIH Anticonvulsant Screening program (unpublished results). While potential blood pressure lowering effects of KCa2 activators, especially if KCa2.2 selective compounds are developed, are less of a concern for ataxia and epilepsy treatment, there is the possibility that KCa2 channel activators might affect learning and memory since KCa2 channels play an important role in synaptic plasticity and long-term potentiation (LTP)18,100. Apamin administration in rats or mice enhance performance in the Morris water maze, the object recognition test and the appetitive lever-pressing task101104, while mice overexpressing KCa2.2 exhibit impaired hippocampal-dependent learning and memory in both the Morris water maze and a contextual fear-conditioning paradigm105. Similarly, pharmacological KCa2 channel activation with EBIO or CyPPA has been reported to impair the encoding, but not the retrieval of object memory in a spontaneous object recognition task106. However, in this context it should also be mentioned that targeted overexpression of KCa2.2 in the basolateral amygdala has been found to reduce anxiety and stress-induced corticosterone secretion107 and that EBIO and CyPPA administration has recently been reported to reduce fear memory caused by repeated stress in rats, suggesting the use of KCa2 activators for alleviating psychiatric symptoms associated with stress induced amygdala hyperactivity108. There is also increasing evidence that neuronal KCa2 channel expression is highly plastic. For example, hippocampal KCa2.3 expression increases with age100, while decreased KCa2.3 expression in the nucleus accumbens is associated with alcohol dependence and infusion of EBIO into the nucleus accumbens reduces alcohol seeking behavior in alcohol dependent rats109. In contrast, pilocarpine-treated chronically epileptic rats exhibit decreased KCa2.1 and KCa2.2 expression in the hippocampus110 and reduced afterhyperpolarization in CA1 hippocampal neurons111.

In addition to reducing neuronal excitability, KCa2 activators might also be useful for treating erectile dysfunction, urinary incontinence and for reducing uterine contractions. Both overexpression of KCa2.3 and intravesicular instillation of NS309 improve bladder function34,112, while CyPPA has recently been shown to inhibit phasic uterine contractions and delays preterm birth in mice113, a finding which is in keeping with the observations made in transgenic mice that KCa2.3 is involved in respiration and parturition114.

Furthermore, KCa2 channels are involved in cardiac excitability and are expressed at higher levels in the atria than the ventricles of mice, rats, rabbits and humans35,115. KCa2 inhibitors have accordingly been proposed as potential new drugs to treat atrial fibrillation (AF)37. As in other tissues, KCa2 channels in the heart seem to be tightly coupled to Ca2+ sources and KCa2.2 have been found to associate with Cav1.2 and Cav1.3 channels in atrial myocytes through the cytoskeletal protein α-actinin2116. Changes in KCa2 channel expression like the reported increase of KCa2.3 expression in failing rabbit ventricles117, which leads to action potential shortening, or the KCa2.3 variations found in lone AF36, have been associated with ventricular or atrial fibrillation. In keeping with the idea that KCa2 blockers are antiarrhythmic, apamin increases action potential duration in human, mouse and rabbit atrial myocytes35,117 and both the NS8593 and UCL 1684 have been reported to terminate acetylcholine-induced AF in rat, guinea pig and rabbit hearts118 and to reduce AF duration in normotensive and hypertensive rats119.

DO ACTVATORS OF ENDOTHELIAL KCa CHANNELS HAVE THERAPEUTIC POTENTIAL AS NOVEL ANTIHYPERTENSIVES?

There is no doubt that KCa3.1/KCa2 activators exert cardiovascular effects in vivo. In fact, i.p. injections of SKA-31 at concentrations of 10–30 mg/kg lower blood pressure in mice, a depressor response that is considerable and last for 60–90 minutes (Delta approx. −30 mmHg)120. Blood pressure lowering effects have been further documented in murine models of hypertension like angiotensin-II induced hypertension66 or in transgenic mice expressing human angiotensinogen and renin (hAGN/hR)121. SKA-31 is also active in large mammals with a heart rate and blood pressure more similar to humans. In instrumented, conscious dogs i.v. injection of 2 mg/kg SKA-31 or of its derivative SKA-2066 produced an immediate and strong (−30 mmHg), but short-lived (10–20 sec) depressor response122. These studies in dogs and mice showed that KCa3.1 and KCa2.3 channels are principally available as antihypertensive drug targets. Moreover, the experiments demonstrated for the first time that the meanwhile “classical” EDH-system can be targeted to lower blood pressure in vivo and could thus constitute an alternative to other antihypertensives like Ca2+ channel antagonist or NO-donors acting directly on the smooth muscle.

Unresolved issues

However, there are several unresolved issues that need to be addressed before it is possible to determine whether KCa3.1 and KCa2.3 activators will be clinically viable and useful antihypertensives. 1) We currently do not know whether continuous stimulation of endothelial KCa2/3 channels will result in desensitization either through “feed-back” modulation of Ca2+ sensitivity and channel activity or reduction in channel expression. That the latter phenomenon is a possibility has been suggested by a study on keratinocytes where EBIO treatment unexpectedly down-regulated KCa3.1 expression and inhibited proliferation123. 2) Sedation and possible impairment of learning and memory through activation of neuronal KCa2 channels could constitute a serious enough side-effect to prohibit the use of KCa2 activators for treating hypertension. Indeed, we recently showed that i.p. injections of SKA-31 at 30 mg/kg produce substantial and lasting sedation (over 1 hour) in mice and 10 and 30 mg/kg SKA-31 reduce voluntary home cage activity over 24 hours124. This problem could of course be avoided by either developing KCa3.1 selective activators or mixed KCa2.3/KCa3.1 activators that are not brain penetrant. 3) KCa2 activators could potentially induce cardiac side effects by causing blockade of AV-transmission or slowing down heart frequency through prolongation of atrial and/or ventricular APs. Indeed, at 30 mg/kg SKA-31 produces pronounced bradycardia in both wild-type and KCa3.1-deficient mice120 suggesting that the effect is most likely caused by activation of cardiac KCa2.X channels. On the other hand, acute i.v. administration of SKA-31 in dogs produces reflex tachycardia to the depressor response122. The latter indicates that KCa3.1/KCa2.3 activators elicit very species-specific cardiovascular effects, which in this particular case are most probably related to the different heart rates of mice and dogs and their different cardiovascular regulatory mechanisms. 4) Another principal concern is that endothelial vasodilator function is often compromised in patients with essential hypertension, obesity or diabetes39,93 and presumably also neurodegenerative disorders like Alzheimer’s disease, and that the endothelial KCa channels might not be sufficiently available as targets or their activation not effective enough. 5) Lastly, there is the possibility that KCa3.1 activators could stimulate immune cell functions like migration or cytokine and reactive oxygen species production and thus exacerbate inflammation. However, we think that this is an unlikely scenario since lymphocytes will probably counter-regulate. We think it equally improbable that KCa3.1 activators would stimulate hyperplasia and fibrosis. Nevertheless, these possibilities should be experimentally investigated before KCa3.1 activators could enter clinical development.

CONCLUSIONS

Considering all these potential caveats, we think it unlikely that a major pharmaceutical company would, at the current stage, invest the resources to develop KCa3.1/KCa2.3 activators for long-term treatment of essential hypertension given the availability of many potent and highly effective antihypertensives such as angiotensin converting enzyme inhibitors or angiotensin receptor blockers. However, KCa activators, especially KCa3.1 selective compounds or mixed KCa3.1/KCa2.3 activators that do not enter the brain, certainly have therapeutic potential for short term applications like intra-surgical hypertension, acute vasospasm or local application after stent implantations. One especially interesting application might be preservation of endothelial function in large vascular organs like hearts or kidneys or in vessel grafts during storage and transplantation. For this usage KCa3.1/KCa2.3 activators could simply be included in the cardioplegia solution. KCa3.1 activators might further be useful for combination therapy in situations where existing antihypertensives are not sufficiently effective or when it is considered desirable to increase blood flow in the microcirculation.

Acknowledgments

Supported by NIH grants R21NS072585 and R01GM076063 to H.W. and a grant of Novo Nordisk Fonden to R.K.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

REFERENCES

  • 1.Wei AD, Gutman GA, Aldrich R, Chandy KG, Grissmer S, Wulff H. International Union of Pharmacology. LII. Nomenclature and molecular relationships of calcium-activated potassium channels. Pharmacol Rev. 2005;57:463–472. doi: 10.1124/pr.57.4.9. [DOI] [PubMed] [Google Scholar]
  • 2.Xia XM, Fakler B, Rivard A, Wayman G, Johnson-Pais T, Keen JE, Ishii T, Hirschberg B, Bond CT, Lutsenko S, Maylie J, Adelman JP. Mechanism of calcium gating in small-conductance calcium-activated potassium channels. Nature. 1998;395:503–507. doi: 10.1038/26758. [DOI] [PubMed] [Google Scholar]
  • 3.Fanger CM, Ghanshani S, Logsdon NJ, Rauer H, Kalman K, Zhou J, Beckingham K, Chandy KG, Cahalan MD, Aiyar J. Calmodulin mediates calcium-dependent activation of the intermediate conductance KCa channel, IKCa1. J Biol Chem. 1999;274:5746–5754. doi: 10.1074/jbc.274.9.5746. [DOI] [PubMed] [Google Scholar]
  • 4.Brugnara C. Sickle cell disease: From membrane pathophysiology to novel therapies for prevention of erythrocyte dehydration. J Pediatr Hematol Oncol. 2003;25:927–933. doi: 10.1097/00043426-200312000-00004. [DOI] [PubMed] [Google Scholar]
  • 5.Heitzmann D, Warth R. Physiology and pathophysiology of potassium channels in gastrointestinal epithelia. Physiol Rev. 2008;88:1119–1182. doi: 10.1152/physrev.00020.2007. [DOI] [PubMed] [Google Scholar]
  • 6.Vandorpe DH, Shmukler BE, Jiang L, Lim B, Maylie J, Adelman JP, de Franceschi L, Cappellini MD, Brugnara C, Alper SL. Cdna cloning and functional characterization of the mouse Ca2+-gated K+ channel, mIK1. Roles in regulatory volume decrease and erythroid differentiation. J Biol Chem. 1998;273:21542–21553. doi: 10.1074/jbc.273.34.21542. [DOI] [PubMed] [Google Scholar]
  • 7.Pena TL, Chen SH, Konieczny SF, Rane SG. Ras/mek/erk up-regulation of the fibroblast KCa channel FIK is a common mechanism for basic fibroblast growth factor and transforming growth factor-beta suppression of myogenesis. J Biol Chem. 2000;275:13677–13682. doi: 10.1074/jbc.275.18.13677. [DOI] [PubMed] [Google Scholar]
  • 8.Ghanshani S, Wulff H, Miller MJ, Rohm H, Neben A, Gutman GA, Cahalan MD, Chandy KG. Up-regulation of the IKCa1 potassium channel during T-cell activation: Molecular mechanism and functional consequences. J Biol Chem. 2000;275:37137–37149. doi: 10.1074/jbc.M003941200. [DOI] [PubMed] [Google Scholar]
  • 9.Kohler R, Wulff H, Eichler I, Kneifel M, Neumann D, Knorr A, Grgic I, Kampfe D, Si H, Wibawa J, Real R, Borner K, Brakemeier S, Orzechowski HD, Reusch HP, Paul M, Chandy KG, Hoyer J. Blockade of the intermediate-conductance calcium-activated potassium channel as a new therapeutic strategy for restenosis. Circulation. 2003;108:1119–1125. doi: 10.1161/01.CIR.0000086464.04719.DD. [DOI] [PubMed] [Google Scholar]
  • 10.Wulff H, Knaus HG, Pennington M, Chandy KG. K+ channel expression during B-cell differentiation: Implications for immunomodulation and autoimmunity. J Immunol. 2004;173:776–786. doi: 10.4049/jimmunol.173.2.776. [DOI] [PubMed] [Google Scholar]
  • 11.Cahalan MD, Chandy KG. The functional network of ion channels in T lymphocytes. Immunol Rev. 2009;231:59–87. doi: 10.1111/j.1600-065X.2009.00816.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Wulff H, Castle NA. Therapeutic potential of KCa3.1 blockers: Recent advances and promissing trends. Expert Rev Clin Pharmacol. 2010;3:385–396. doi: 10.1586/ecp.10.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Joiner WJ, Wang LY, Tang MD, Kaczmarek LK. hSK4, a member of a novel subfamily of calcium-activated potassium channels. Proc Natl Acad Sci USA. 1997;94:11013–11018. doi: 10.1073/pnas.94.20.11013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ishii TM, Silvia C, Hirschberg B, Bond CT, Adelman JP, Maylie J. A human intermediate conductance calcium-activated potassium channel. Proc Natl Acad Sci USA. 1997;94:11651–11656. doi: 10.1073/pnas.94.21.11651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Logsdon NJ, Kang J, Togo JA, Christian EP, Aiyar J. A novel gene, hKCa4, encodes the calcium-activated potassium channel in human T lymphocytes. J Biol Chem. 1997;272:32723–32726. doi: 10.1074/jbc.272.52.32723. [DOI] [PubMed] [Google Scholar]
  • 16.Engbers JD, Anderson D, Asmara H, Rehak R, Mehaffey WH, Hameed S, McKay BE, Kruskic M, Zamponi GW, Turner RW. Intermediate conductance calcium-activated potassium channels modulate summation of parallel fiber input in cerebellar purkinje cells. Proc Natl Acad Sci USA. 2012;109:2601–2606. doi: 10.1073/pnas.1115024109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Stocker M. Ca2+-activated K+ channels: Molecular determinants and function of the SK family. Nat Rev Neurosci. 2004;5:758–770. doi: 10.1038/nrn1516. [DOI] [PubMed] [Google Scholar]
  • 18.Adelman JP, Maylie J, Sah P. Small-conductance Ca2+-activated K+ channels: Form and function. Annu Rev Physiol. 2012;74:245–269. doi: 10.1146/annurev-physiol-020911-153336. [DOI] [PubMed] [Google Scholar]
  • 19.Stocker M, Pedarzani P. Differential distribution of three Ca2+-activated K+ channel subunits, SK1, SK2, and SK3, in the adult rat central nervous system. Mol Cell Neurosci. 2000;15:476–493. doi: 10.1006/mcne.2000.0842. [DOI] [PubMed] [Google Scholar]
  • 20.Pedarzani P, Stocker M. Molecular and cellular basis of small- and intermediate-conductance, calcium-activated potassium channel function in the brain. Cell Mol Life Sci. 2008;65:3196–3217. doi: 10.1007/s00018-008-8216-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Stocker M, Krause M, Pedarzani P. An apamin-sensitive Ca2+-activated K+ current in hippocampal pyramidal neurons. Proc Natl Acad Sci USA. 1999;96:4662–4667. doi: 10.1073/pnas.96.8.4662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Pedarzani P, Mosbacher J, Rivard A, Cingolani LA, Oliver D, Stocker M, Adelman JP, Fakler B. Control of electrical activity in central neurons by modulating the gating of small conductance Ca2+-activated K+ channels. J Biol Chem. 2001;276:9762–9769. doi: 10.1074/jbc.M010001200. [DOI] [PubMed] [Google Scholar]
  • 23.Pedarzani P, McCutcheon JE, Rogge G, Jensen BS, Christophersen P, Hougaard C, Strobaek D, Stocker M. Specific enhancement of SK channel activity selectively potentiates the afterhyperpolarizing current I(AHP) and modulates the firing properties of hippocampal pyramidal neurons. J Biol Chem. 2005;280:41404–41411. doi: 10.1074/jbc.M509610200. [DOI] [PubMed] [Google Scholar]
  • 24.Hallworth NE, Wilson CJ, Bevan MD. Apamin-sensitive small conductance calcium-activated potassium channels, through their selective coupling to voltage-gated calcium channels, are critical determinants of the precision, pace, and pattern of action potential generation in rat subthalamic nucleus neurons in vitro. J Neurosci. 2003;23:7525–7542. doi: 10.1523/JNEUROSCI.23-20-07525.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Cingolani LA, Gymnopoulos M, Boccaccio A, Stocker M, Pedarzani P. Developmental regulation of small-conductance Ca2+-activated K+ channel expression and function in rat Purkinje neurons. J Neurosci. 2002;22:4456–4467. doi: 10.1523/JNEUROSCI.22-11-04456.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Sarpal D, Koenig JI, Adelman JP, Brady D, Prendeville LC, Shepard PD. Regional distribution of SK3 mRNA-containing neurons in the adult and adolescent rat ventral midbrain and their relationship to dopamine-containing cells. Synapse. 2004;53:104–113. doi: 10.1002/syn.20042. [DOI] [PubMed] [Google Scholar]
  • 27.Wolfart J, Neuhoff H, Franz O, Roeper J. Differential expression of the small-conductance, calcium-activated potassium channel SK3 is critical for pacemaker control in dopaminergic midbrain neurons. J Neurosci. 2001;21:3443–3456. doi: 10.1523/JNEUROSCI.21-10-03443.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Waroux O, Massotte L, Alleva L, Graulich A, Thomas E, Liegeois JF, Scuvee-Moreau J, Seutin V. SK channels control the firing pattern of midbrain dopaminergic neurons in vivo. Eur J Neurosci. 2005;22:3111–3121. doi: 10.1111/j.1460-9568.2005.04484.x. [DOI] [PubMed] [Google Scholar]
  • 29.Marrion NV, Tavalin SJ. Selective activation of Ca2+-activated K+ channels by co-localized Ca2+ channels in hippocampal neurons. Nature. 1998;395:900–905. doi: 10.1038/27674. [DOI] [PubMed] [Google Scholar]
  • 30.Edgerton JR, Reinhart PH. Distinct contributions of small and large conductance Ca2+-activated K+ channels to rat purkinje neuron function. J Physiol. 2003;548:53–69. doi: 10.1113/jphysiol.2002.027854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Faber ES, Delaney AJ, Sah P. SK channels regulate excitatory synaptic transmission and plasticity in the lateral amygdala. Nat Neurosci. 2005;8:635–641. doi: 10.1038/nn1450. [DOI] [PubMed] [Google Scholar]
  • 32.Ngo-Anh TJ, Bloodgood BL, Lin M, Sabatini BL, Maylie J, Adelman JP. SK channels and nmda receptors form a Ca2+-mediated feedback loop in dendritic spines. Nat Neurosci. 2005;8:642–649. doi: 10.1038/nn1449. [DOI] [PubMed] [Google Scholar]
  • 33.Barfod ET, Moore AL, Lidofsky SD. Cloning and functional expression of a liver isoform of the small conductance Ca2+-activated K+ channel SK3. Am J Physiol Cell Physiol. 2001;280:C836–C842. doi: 10.1152/ajpcell.2001.280.4.C836. [DOI] [PubMed] [Google Scholar]
  • 34.Herrera GM, Pozo MJ, Zvara P, Petkov GV, Bond CT, Adelman JP, Nelson MT. Urinary bladder instability induced by selective suppression of the murine small conductance calcium-activated potassium (SK3) channel. J Physiol. 2003;551:893–903. doi: 10.1113/jphysiol.2003.045914. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Xu Y, Tuteja D, Zhang Z, Xu D, Zhang Y, Rodriguez J, Nie L, Tuxson HR, Young JN, Glatter KA, Vazquez AE, Yamoah EN, Chiamvimonvat N. Molecular identification and functional roles of a Ca2+-activated K+ channel in human and mouse hearts. J Biol Chem. 2003;278:49085–49094. doi: 10.1074/jbc.M307508200. [DOI] [PubMed] [Google Scholar]
  • 36.Ellinor PT, Lunetta KL, Glazer NL, Pfeufer A, Alonso A, Chung MK, Sinner MF, de Bakker PI, Mueller M, Lubitz SA, Fox E, Darbar D, Smith NL, Smith JD, Schnabel RB, Soliman EZ, Rice KM, Van Wagoner DR, Beckmann BM, van Noord C, Wang K, Ehret GB, Rotter JI, Hazen SL, Steinbeck G, Smith AV, Launer LJ, Harris TB, Makino S, Nelis M, Milan DJ, Perz S, Esko T, Kottgen A, Moebus S, Newton-Cheh C, Li M, Mohlenkamp S, Wang TJ, Kao WH, Vasan RS, Nothen MM, MacRae CA, Stricker BH, Hofman A, Uitterlinden AG, Levy D, Boerwinkle E, Metspalu A, Topol EJ, Chakravarti A, Gudnason V, Psaty BM, Roden DM, Meitinger T, Wichmann HE, Witteman JC, Barnard J, Arking DE, Benjamin EJ, Heckbert SR, Kaab S. Common variants in KCNN3 are associated with lone atrial fibrillation. Nature Genetics. 2010;42:240–244. doi: 10.1038/ng.537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Grunnet M, Bentzen BH, Sorensen US, Diness JG. Cardiac ion channels and mechanisms for protection against atrial fibrillation. Rev Physiol, Biochem & Pharmacol. 2012;162:1–58. doi: 10.1007/112_2011_3. [DOI] [PubMed] [Google Scholar]
  • 38.Edwards G, Feletou M, Weston AH. Endothelium-derived hyperpolarising factors and associated pathways: A synopsis. Pflugers Arch. 2010;459:863–879. doi: 10.1007/s00424-010-0817-1. [DOI] [PubMed] [Google Scholar]
  • 39.Grgic I, Kaistha BP, Hoyer J, Kohler R. Endothelial Ca2+-activated K+ channels in normal and impaired EDHF-dilator responses--relevance to cardiovascular pathologies and drug discovery. Br J Pharmacol. 2009;157:509–526. doi: 10.1111/j.1476-5381.2009.00132.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Dora KA, Gallagher NT, McNeish A, Garland CJ. Modulation of endothelial cell KCa3.1 channels during endothelium-derived hyperpolarizing factor signaling in mesenteric resistance arteries. Circ Res. 2008;102:1247–1255. doi: 10.1161/CIRCRESAHA.108.172379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Ledoux J, Taylor MS, Bonev AD, Hannah RM, Solodushko V, Shui B, Tallini Y, Kotlikoff MI, Nelson MT. Functional architecture of inositol 1,4,5-trisphosphate signaling in restricted spaces of myoendothelial projections. Proc Natl Acad Sci USA. 2008;105:9627–9632. doi: 10.1073/pnas.0801963105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Saliez J, Bouzin C, Rath G, Ghisdal P, Desjardins F, Rezzani R, Rodella LF, Vriens J, Nilius B, Feron O, Balligand JL, Dessy C. Role of caveolar compartmentation in endothelium-derived hyperpolarizing factor-mediated relaxation: Ca2+ signals and gap junction function are regulated by caveolin in endothelial cells. Circulation. 2008;117:1065–1074. doi: 10.1161/CIRCULATIONAHA.107.731679. [DOI] [PubMed] [Google Scholar]
  • 43.Absi M, Burnham MP, Weston AH, Harno E, Rogers M, Edwards G. Effects of methyl beta-cyclodextrin on EDHF responses in pig and rat arteries association between SK(Ca) channels and caveolin-rich domains. Br J Pharmacol. 2007;151:332–340. doi: 10.1038/sj.bjp.0707222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Brahler S, Kaistha A, Schmidt VJ, Wolfle SE, Busch C, Kaistha BP, Kacik M, Hasenau AL, Grgic I, Si H, Bond CT, Adelman JP, Wulff H, de Wit C, Hoyer J, Kohler R. Genetic deficit of SK3 and IK1 channels disrupts the endothelium-derived hyperpolarizing factor vasodilator pathway and causes hypertension. Circulation. 2009;119:2323–2332. doi: 10.1161/CIRCULATIONAHA.108.846634. [DOI] [PubMed] [Google Scholar]
  • 45.Weston AH, Porter EL, Harno E, Edwards G. Impairment of endothelial SK(Ca) channels and of downstream hyperpolarizing pathways in mesenteric arteries from spontaneously hypertensive rats. Br J Pharmacol. 2010;160:836–843. doi: 10.1111/j.1476-5381.2010.00657.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Kohler R, Kaistha BP, Wulff H. Vascular KCa-channels as therapeutic targets in hypertension and restenosis disease. Expert Opin Ther Targets. 2010;14:143–155. doi: 10.1517/14728220903540257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Dalsgaard T, Kroigaard C, Simonsen U. Calcium-activated potassium channels - a therapeutic target for modulating nitric oxide in cardiovascular disease? Expert Opin Ther Targets. 2010;14:825–837. doi: 10.1517/14728222.2010.500616. [DOI] [PubMed] [Google Scholar]
  • 48.Eichler I, Wibawa J, Grgic I, Knorr A, Brakemeier S, Pries AR, Hoyer J, Kohler R. Selective blockade of endothelial Ca2+-activated small- and intermediate-conductance K+-channels suppresses EDHF-mediated vasodilation. Br J Pharmacol. 2003;138:594–601. doi: 10.1038/sj.bjp.0705075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kohler R, Ruth P. Endothelial dysfunction and blood pressure alterations in K+-channel transgenic mice. Pflugers Arch. 2010;459:969–976. doi: 10.1007/s00424-010-0819-z. [DOI] [PubMed] [Google Scholar]
  • 50.Wulff H, Kolski-Andreaco A, Sankaranarayanan A, Sabatier JM, Shakkottai V. Modulators of small- and intermediate-conductance calcium-activated potassium channels and their therapeutic indications. Curr Med Chem. 2007;14:1437–1457. doi: 10.2174/092986707780831186. [DOI] [PubMed] [Google Scholar]
  • 51.Rauer H, Lanigan MD, Pennington MW, Aiyar J, Ghanshani S, Cahalan MD, Norton RS, Chandy KG. Structure-guided transformation of charybdotoxin yields an analog that selectively targets Ca2+-activated over voltage-gated K+ channels. J Biol Chem. 2000;275:1201–1208. doi: 10.1074/jbc.275.2.1201. [DOI] [PubMed] [Google Scholar]
  • 52.Wulff H, Miller MJ, Haensel W, Grissmer S, Cahalan MD, Chandy KG. Design of a potent and selective inhibitor of the intermediate-conductance Ca2+-activated K+ channel, ikca1: A potential immunosuppressant. Proc Natl Acad Sci USA. 2000;97:8151–8156. doi: 10.1073/pnas.97.14.8151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Stocker JW, De Franceschi L, McNaughton-Smith GA, Corrocher R, Beuzard Y, Brugnara C. ICA-17043, a novel Gardos channel blocker, prevents sickled red blood cell dehydration in vitro and in vivo in SAD mice. Blood. 2003;101:2412–2418. doi: 10.1182/blood-2002-05-1433. [DOI] [PubMed] [Google Scholar]
  • 54.Ataga KI, Reid M, Ballas SK, Yasin Z, Bigelow C, James LS, Smith WR, Galacteros F, Kutlar A, Hull JH, Stocker JW. Improvements in haemolysis and indicators of erythrocyte survival do not correlate with acute vaso-occlusive crises in patients with sickle cell disease: A phase III randomized, placebo-controlled, double-blind study of the Gardos channel blocker senicapoc (ICA-17043) Br J Haematol. 2011;153:92–104. doi: 10.1111/j.1365-2141.2010.08520.x. [DOI] [PubMed] [Google Scholar]
  • 55.Strobaek D, Brown DT, Jenkins DP, Chen Y, Coleman N, Ando Y, Chiu P, Jorgensen S, Demnitz J, Wulff H, Christophersen P. NS6180, a new KCa3.1 channel inhibitor prevents T-cell activation and inflammation in a rat model of inflammatory bowel disease. Br J Pharmacol. 2012 doi: 10.1111/j.1476-5381.2012.02143.x. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Wulff H, Gutman GA, Cahalan MD, Chandy KG. Delination of the clotrimazole/TRAM-34 binding site on the intermediate conductance calcium-activated potassium channel IKCa1. J Biol Chem. 2001;276:32040–32045. doi: 10.1074/jbc.M105231200. [DOI] [PubMed] [Google Scholar]
  • 57.Lamy C, Goodchild SJ, Weatherall KL, Jane DE, Liegeois JF, Seutin V, Marrion NV. Allosteric block of KCa2 channels by apamin. J Biol Chem. 2010;285:27067–27077. doi: 10.1074/jbc.M110.110072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Weatherall KL, Seutin V, Liegeois JF, Marrion NV. Crucial role of a shared extracellular loop in apamin sensitivity and maintenance of pore shape of small-conductance calcium-activated potassium (SK) channels. Proc Natl Acad Sci USA. 2011;108:18494–18499. doi: 10.1073/pnas.1110724108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Campos Rosa J, Galanakis D, Piergentili A, Bhandari K, Ganellin CR, Dunn PM, Jenkinson DH. Synthesis, molecular modeling, and pharmacological testing of bis-quinolinium cyclophanes: Potent, non-peptidic blockers of the apamin-sensitive Ca2+-activated K+ channel. J Med Chem. 2000;43:420–431. doi: 10.1021/jm9902537. [DOI] [PubMed] [Google Scholar]
  • 60.Shakkottai VG, Regaya I, Wulff H, Fajloun Z, Tomita H, Fathallah M, Cahalan MD, Gargus JJ, Sabatier JM, Chandy KG. Design and characterization of a highly selective peptide inhibitor of the small conductance calcium-activated K+ channel, SKCa2. J Biol Chem. 2001;276:43145–43151. doi: 10.1074/jbc.M106981200. [DOI] [PubMed] [Google Scholar]
  • 61.Bruening-Wright A, Schumacher MA, Adelman JP, Maylie J. Localization of the activation gate for small conductance Ca2+-activated K+ channels. J Neurosci. 2002;22:6499–6506. doi: 10.1523/JNEUROSCI.22-15-06499.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Bruening-Wright A, Lee WS, Adelman JP, Maylie J. Evidence for a deep pore activation gate in small conductance Ca2+-activated K+ channels. J Gen Physiol. 2007;130:601–610. doi: 10.1085/jgp.200709828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Garneau L, Klein H, Banderali U, Longpre-Lauzon A, Parent L, Sauve R. Hydrophobic interactions as key determinants to the KCa3.1 channel closed configuration. An analysis of KCa3.1 mutants constitutively active in zero Ca2+ J Biol Chem. 2009;284:389–403. doi: 10.1074/jbc.M805700200. [DOI] [PubMed] [Google Scholar]
  • 64.Klein H, Garneau L, Banderali U, Simoes M, Parent L, Sauve R. Structural determinants of the closed KCa3.1 channel pore in relation to channel gating: Results from a substituted cysteine accessibility analysis. J Gen Physiol. 2007;129:299–315. doi: 10.1085/jgp.200609726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Strobaek D, Teuber L, Jorgensen TD, Ahring PK, Kjaer K, Hansen RS, Olesen SP, Christophersen P, Skaaning-Jensen B. Activation of human IK and SK Ca2+ -activated K+ channels by NS309 (6,7-dichloro-1H-indole-2,3-dione 3-oxime) Biochim Biophys Acta. 2004;1665:1–5. doi: 10.1016/j.bbamem.2004.07.006. [DOI] [PubMed] [Google Scholar]
  • 66.Sankaranarayanan A, Raman G, Busch C, Schultz T, Zimin PI, Hoyer J, Kohler R, Wulff H. Naphtho[1,2-d]thiazol-2-ylamine (SKA-31), a new activator of KCa2 and KCa3.1 potassium channels, potentiates the endothelium-derived hyperpolarizing factor response and lowers blood pressure. Mol Pharmacol. 2009;75:281–295. doi: 10.1124/mol.108.051425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Jenkins DP, Strobaek D, Hougaard C, Jensen ML, Hummel R, Sorensen US, Christophersen P, Wulff H. Negative gating modulation by (R)-N-(benzimidazol-2-yl)-1,2,3,4-tetrahydro-1-naphthylamine (NS8593) depends on residues in the inner pore vestibule: Pharmacological evidence of deep-pore gating of KCa2 channels. Mol Pharmacol. 2011;79:899–909. doi: 10.1124/mol.110.069807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Hougaard C, Hammami S, Eriksen BL, Sorensen US, Jensen ML, Strobaek D, Christophersen P. Evidence for a common pharmacological interaction site on KCa2 channels providing both selective activation and selective inhibition of the human KCa2.1 subtype. Mol Pharmacol. 2012;81:210–219. doi: 10.1124/mol.111.074252. [DOI] [PubMed] [Google Scholar]
  • 69.Kasumu AW, Hougaard C, Rode F, Jacobson TA, Sabatier JM, Eriksen BL, Strobaek D, Liang X, Egorova P, Vorontsova D, Christophersen P, Ronn LCB, Bezprozvanny I. Novel selective positive modulator of calcium-activated potassium channels exerts beneficial effects in a mouse model of spinocerebellar ataxia type 2. Chemistry & Biology. 2012 doi: 10.1016/j.chembiol.2012.07.013. Epub ahead of Print. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Cheng-Raude D, Treloar M, Habermann E. Preparation and pharmacokinetics of labeled derivatives of apamin. Toxicon. 1976;14:467–476. doi: 10.1016/0041-0101(76)90064-7. [DOI] [PubMed] [Google Scholar]
  • 71.Simms LA, Doecke JD, Roberts RL, Fowler EV, Zhao ZZ, McGuckin MA, Huang N, Hayward NK, Webb PM, Whiteman DC, Cavanaugh JA, McCallum R, Florin TH, Barclay ML, Gearry RB, Merriman TR, Montgomery GW, Radford-Smith GL. KCNN4 gene variant is associated with ileal Crohn's disease in the Australian and New Zealand population. Am J Gastroenterol. 2010;105:2209–2217. doi: 10.1038/ajg.2010.161. [DOI] [PubMed] [Google Scholar]
  • 72.De Franceschi L, Saadane N, Trudel M, Alper SL, Brugnara C, Beuzard Y. Treatment with oral clotrimazole blocks Ca2+-activated K+ transport and reverses erythrocyte dehydration in transgenic sad mice. A model for therapy of sickle cell disease. J Clin Invest. 1994;93:1670–1676. doi: 10.1172/JCI117149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Brugnara C, Gee B, Armsby CC, Kurth S, Sakamoto M, Rifai N, Alper SL, Platt OS. Therapy with oral clotrimazole induces inhibition of the Gardos channel and reduction of erythrocyte dehydration in patients with sickle cell disease. J Clin Invest. 1996;97:1227–1234. doi: 10.1172/JCI118537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Bradding P, Wulff H. The K+ channels KCa3.1 and Kv1.3 as novel targets for asthma therapy. Br J Pharmacol. 2009;157:1330–1339. doi: 10.1111/j.1476-5381.2009.00362.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Duffy SM, Berger P, Cruse G, Yang W, Bolton SJ, Bradding P. The K+ channel IKCa1 potentiates Ca2+ influx and degranulation in human lung mast cells. J Allergy Clin Immunol. 2004;114:66–72. doi: 10.1016/j.jaci.2004.04.005. [DOI] [PubMed] [Google Scholar]
  • 76.Cruse G, Duffy SM, Brightling CE, Bradding P. Functional KCa3.1 K+ channels are required for human lung mast cell migration. Thorax. 2006;61:880–885. doi: 10.1136/thx.2006.060319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Shepherd MC, Duffy SM, Harris T, Cruse G, Schuliga M, Brightling CE, Neylon CB, Bradding P, Stewart AG. KCa3.1 Ca2+ activated K+ channels regulate human airway smooth muscle proliferation. Am J Respir Cell Mol Biol. 2007;37:525–531. doi: 10.1165/rcmb.2006-0358OC. [DOI] [PubMed] [Google Scholar]
  • 78.Robinette L, Abraham WM, Bradding P, Krajewski J, Antonio B, Shelton T, Bannon AW, Rigdon G, Krafte D, Wagoner K, Castle NA. Senicapoc (ICA-17043), a potent and selective KCa3.1 K+ channel blocker, attenuates allergen-induced asthma in sheep. 15th International Conference of the Inflammation Research Association. 2008 [Google Scholar]
  • 79.Reich EP, Cui L, Yang L, Pugliese-Sivo C, Golovko A, Petro M, Vassileva G, Chu I, Nomeir AA, Zhang LK, Liang X, Kozlowski JA, Narula SK, Zavodny PJ, Chou CC. Blocking ion channel KCNN4 alleviates the symptoms of experimental autoimmune encephalomyelitis in mice. Eur J Immunol. 2005;35:1027–1036. doi: 10.1002/eji.200425954. [DOI] [PubMed] [Google Scholar]
  • 80.Chou CC, Lunn CA, Murgolo NJ. KCa3.1: Target and marker for cancer, autoimmune disorder and vascular inflammation? Expert Rev Mol Diagn. 2008;8:179–187. doi: 10.1586/14737159.8.2.179. [DOI] [PubMed] [Google Scholar]
  • 81.Di L, Srivastava S, Zhdanova O, Ding Y, Li Z, Wulff H, Lafaille M, Skolnik EY. Inhibition of the K+ channel KCa3.1 ameliorates T cell-mediated colitis. Proc Natl Acad Sci USA. 2010;107:1541–1546. doi: 10.1073/pnas.0910133107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Grgic I, Kiss E, Kaistha BP, Busch C, Kloss M, Sautter J, Muller A, Kaistha A, Schmidt C, Raman G, Wulff H, Strutz F, Grone HJ, Kohler R, Hoyer J. Renal fibrosis is attenuated by targeted disruption of KCa3.1 potassium channels. Proc Natl Acad Sci USA. 2009;106:14518–14523. doi: 10.1073/pnas.0903458106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Grgic I, Eichler I, Heinau P, Si H, Brakemeier S, Hoyer J, Kohler R. Selective blockade of the intermediate-conductance Ca2+-activated K+ channel suppresses proliferation of microvascular and macrovascular endothelial cells and angiogenesis in vivo. Arterioscler Thromb Vasc Biol. 2005;25:704–709. doi: 10.1161/01.ATV.0000156399.12787.5c. [DOI] [PubMed] [Google Scholar]
  • 84.Toyama K, Wulff H, Chandy KG, Azam P, Raman G, Saito T, Fujiwara Y, Mattson DL, Das S, Melvin JE, Pratt PF, Hatoum OA, Gutterman DD, Harder DR, Miura H. The intermediate-conductance calcium-activated potassium channel KCa3.1 contributes to atherogenesis in mice and humans. J Clin Invest. 2008;118:3025–3037. doi: 10.1172/JCI30836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Tharp DL, Wamhoff BR, Wulff H, Raman G, Cheong A, Bowles DK. Local delivery of the KCa3.1 blocker, TRAM-34, prevents acute angioplasty-induced coronary smooth muscle phenotypic modulation and limits stenosis. Arterioscler Thromb Vasc Biol. 2008;28:1084–1089. doi: 10.1161/ATVBAHA.107.155796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Khanna R, Roy L, Zhu X, Schlichter LC. K+ channels and the microglial respiratory burst. Am J Physiol Cell Physiol. 2001;280:C796–C806. doi: 10.1152/ajpcell.2001.280.4.C796. [DOI] [PubMed] [Google Scholar]
  • 87.Chen YJ, Raman G, Bodendiek S, O'Donnell ME, Wulff H. The KCa3.1 blocker TRAM-34 reduces infarction and neurological deficit in a rat model of ischemia/reperfusion stroke. J Cereb Blood Flow Metab. 2011;31:2363–2374. doi: 10.1038/jcbfm.2011.101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Mauler F, Hinz V, Horvath E, Schuhmacher J, Hofmann HA, Wirtz S, Hahn MG, Urbahns K. Selective intermediate-/small-conductance calcium-activated potassium channel (KCNN4) blockers are potent and effective therapeutics in experimental brain oedema and traumatic brain injury caused by acute subdural haematoma. Eur J Neurosci. 2004;20:1761–1768. doi: 10.1111/j.1460-9568.2004.03615.x. [DOI] [PubMed] [Google Scholar]
  • 89.Kaushal V, Koeberle PD, Wang Y, Schlichter LC. The Ca2+-activated K+ channel KCNN4/KCa3.1 contributes to microglia activation and nitric oxide-dependent neurodegeneration. J Neurosci. 2007;27:234–244. doi: 10.1523/JNEUROSCI.3593-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Bouhy D, Ghasemlou N, Lively S, Redensek A, Rathore KI, Schlichter LC, David S. Inhibition of the Ca2+-dependent K+ channel, KCNN4/KCa3.1, improves tissue protection and locomotor recovery after spinal cord injury. J Neurosci. 2011;31:16298–16308. doi: 10.1523/JNEUROSCI.0047-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Maezawa I, Jenkins DP, Jin BE, Wulff H. Microglial KCa3.1 channels as a potential therapeutic target for Alzheimer's disease. International Journal of Alzheimer's Disease. 2012;2012 doi: 10.1155/2012/868972. Epub ahead of print. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Ataga KI, Smith WR, De Castro LM, Swerdlow P, Saunthararajah Y, Castro O, Vichinsky E, Kutlar A, Orringer EP, Rigdon GC, Stocker JW. Efficacy and safety of the Gardos channel blocker, senicapoc (ICA-17043), in patients with sickle cell anemia. Blood. 2008;111:3991–3997. doi: 10.1182/blood-2007-08-110098. [DOI] [PubMed] [Google Scholar]
  • 93.Gao X, Martinez-Lemus LA, Zhang C. Endothelium-derived hyperpolarizing factor and diabetes. World J Cardiol. 2011;3:25–31. doi: 10.4330/wjc.v3.i1.25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.McCown TJ, Breese GR. Effects of apamin and nicotinic acetylcholine receptor antagonists on inferior collicular seizures. Eur J Pharmacol. 1990;187:49–58. doi: 10.1016/0014-2999(90)90339-8. [DOI] [PubMed] [Google Scholar]
  • 95.Kobayashi K, Nishizawa Y, Sawada K, Ogura H, Miyabe M. K+-channel openers suppress epileptiform activities induced by 4-aminopyridine in cultured rat hippocampal neurons. J Pharmacol Sci. 2008;108:517–528. doi: 10.1254/jphs.08214fp. [DOI] [PubMed] [Google Scholar]
  • 96.Shakkottai VG, Chou CH, Oddo S, Sailer CA, Knaus HG, Gutman GA, Barish ME, LaFerla FM, Chandy KG. Enhanced neuronal excitability in the absence of neurodegeneration induces cerebellar ataxia. J Clin Invest. 2004;113:582–590. doi: 10.1172/JCI20216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Shakkottai VG, do Carmo Costa M, Dell'Orco JM, Sankaranarayanan A, Wulff H, Paulson HL. Early changes in cerebellar physiology accompany motor dysfunction in the polyglutamine disease spinocerebellar ataxia type 3. J Neurosci. 2011;31:13002–13014. doi: 10.1523/JNEUROSCI.2789-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Walter JT, Alvina K, Womack MD, Chevez C, Khodakhah K. Decreases in the precision of Purkinje cell pacemaking cause cerebellar dysfunction and ataxia. Nat Neurosci. 2006;9:389–397. doi: 10.1038/nn1648. [DOI] [PubMed] [Google Scholar]
  • 99.Anderson NJ, Slough S, Watson WP. In vivo characterisation of the small-conductance KCa (SK) channel activator 1-ethyl-2-benzimidazolinone (1-EBIO) as a potential anticonvulsant. Eur J Pharmacol. 2006;546:48–53. doi: 10.1016/j.ejphar.2006.07.007. [DOI] [PubMed] [Google Scholar]
  • 100.Blank T, Nijholt I, Kye MJ, Radulovic J, Spiess J. Small-conductance, Ca2+-activated K+ channel SK3 generates age-related memory and LTP deficits. Nat Neurosci. 2003;6:911–912. doi: 10.1038/nn1101. [DOI] [PubMed] [Google Scholar]
  • 101.Deschaux O, Bizot JC, Goyffon M. Apamin improves learning in an object recognition task in rats. Neurosci Lett. 1997;222:159–162. doi: 10.1016/s0304-3940(97)13367-5. [DOI] [PubMed] [Google Scholar]
  • 102.Fournier C, Kourrich S, Soumireu-Mourat B, Mourre C. Apamin improves reference memory but not procedural memory in rats by blocking small conductance Ca2+-activated K+ channels in an olfactory discrimination task. Behav Brain Res. 2001;121:81–93. doi: 10.1016/s0166-4328(00)00387-9. [DOI] [PubMed] [Google Scholar]
  • 103.Messier C, Mourre C, Bontempi B, Sif J, Lazdunski M, Destrade C. Effect of apamin, a toxin that inhibits Ca2+-dependent K+ channels, on learning and memory processes. Brain Res. 1991;551:322–326. doi: 10.1016/0006-8993(91)90950-z. [DOI] [PubMed] [Google Scholar]
  • 104.Bond CT, Maylie J, Adelman JP. SK channels in excitability, pacemaking and synaptic integration. Curr Opin Neurobiol. 2005;15:305–311. doi: 10.1016/j.conb.2005.05.001. [DOI] [PubMed] [Google Scholar]
  • 105.Hammond RS, Bond CT, Strassmaier T, Ngo-Anh TJ, Adelman JP, Maylie J, Stackman RW. Small-conductance Ca2+-activated K+ channel type 2 (SK2) modulates hippocampal learning, memory, and synaptic plasticity. J Neurosci. 2006;26:1844–1853. doi: 10.1523/JNEUROSCI.4106-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Vick KAt, Guidi M, Stackman RW., Jr In vivo pharmacological manipulation of small conductance Ca2+-activated K+ channels influences motor behavior, object memory and fear conditioning. Neuropharmacol. 2010;58:650–659. doi: 10.1016/j.neuropharm.2009.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Mitra R, Ferguson D, Sapolsky RM. SK2 potassium channel overexpression in basolateral amygdala reduces anxiety, stress-induced corticosterone secretion and dendritic arborization. Mol Psychiatry. 2009;14:847–8455. doi: 10.1038/mp.2009.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Atchley D, Hankosky ER, Gasparotto K, Rosenkranz JA. Pharmacological enhancement of calcium-activated potassium channel function reduces the effects of repeated stress on fear memory. Behav Brain Res. 2012;232:37–43. doi: 10.1016/j.bbr.2012.03.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Hopf FW, Bowers MS, Chang SJ, Chen BT, Martin M, Seif T, Cho SL, Tye K, Bonci A. Reduced nucleus accumbens SK channel activity enhances alcohol seeking during abstinence. Neuron. 65:682–694. doi: 10.1016/j.neuron.2010.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Oliveira MS, Skinner F, Arshadmansab MF, Garcia I, Mello CF, Knaus HG, Ermolinsky BS, Otalora LF, Garrido-Sanabria ER. Altered expression and function of small-conductance (SK) Ca2+-activated K+ channels in pilocarpine-treated epileptic rats. Brain Res. 2010;1348:187–199. doi: 10.1016/j.brainres.2010.05.095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Schulz R, Kirschstein T, Brehme H, Porath K, Mikkat U, Kohling R. Network excitability in a model of chronic temporal lobe epilepsy critically depends on SK channel-mediated AHP currents. Neurobiol Disease. 2012;45:337–347. doi: 10.1016/j.nbd.2011.08.019. [DOI] [PubMed] [Google Scholar]
  • 112.Pandita RK, Ronn LC, Jensen BS, Andersson KE. Urodynamic effects of intravesical administration of the new small/intermediate conductance calcium activated potassium channel activator NS309 in freely moving, conscious rats. J Urol. 2006;176:1220–1224. doi: 10.1016/j.juro.2006.04.081. [DOI] [PubMed] [Google Scholar]
  • 113.Skarra DV, Cornwell T, Solodushko V, Brown A, Taylor MS. CyPPA a positive modulator of small-conductance Ca2+-activated K+ channels, inhibits phasic uterine contractions and delays preterm birth in mice. Am J Physiol Cell Physiol. 2011;301:C1027–C1035. doi: 10.1152/ajpcell.00082.2011. [DOI] [PubMed] [Google Scholar]
  • 114.Bond CT, Sprengel R, Bissonnette JM, Kaufman WA, Pribnow D, Neelands T, Storck T, Baetscher M, Jerecic J, Maylie J, Knaus HG, Seeburg P, Adelman JP. Respiration and parturition affected by conditional overexpression of the Ca2+ activated K+ channel subunit, SK3. Science. 2000;289:1942–1946. doi: 10.1126/science.289.5486.1942. [DOI] [PubMed] [Google Scholar]
  • 115.Tuteja D, Xu D, Timofeyev V, Lu L, Sharma D, Zhang Z, Xu Y, Nie L, Vazquez AE, Young JN, Glatter KA, Chiamvimonvat N. Differential expression of small-conductance Ca2+-activated K+ channels SK1, SK2, and SK3 in mouse atrial and ventricular myocytes. Am J Physiol Heart Circ Physiol. 2005;289:H2714–H2723. doi: 10.1152/ajpheart.00534.2005. [DOI] [PubMed] [Google Scholar]
  • 116.Lu L, Zhang Q, Timofeyev V, Zhang Z, Young JN, Shin HS, Knowlton AA, Chiamvimonvat N. Molecular coupling of a Ca2+-activated K+ channel to L-type Ca2+ channels via alpha-actinin2. Circ Res. 2007;100:112–120. doi: 10.1161/01.RES.0000253095.44186.72. [DOI] [PubMed] [Google Scholar]
  • 117.Chua SK, Chang PC, Maruyama M, Turker I, Shinohara T, Shen MJ, Chen Z, Shen C, Rubart-von der Lohe M, Lopshire JC, Ogawa M, Weiss JN, Lin SF, Ai T, Chen PS. Small-conductance calcium-activated potassium channel and recurrent ventricular fibrillation in failing rabbit ventricles. Circ Res. 2011;108:971–979. doi: 10.1161/CIRCRESAHA.110.238386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Diness JG, Sorensen US, Nissen JD, Al-Shahib B, Jespersen T, Grunnet M, Hansen RS. Inhibition of small-conductance Ca2+-activated K+ channels terminates and protects against atrial fibrillation. Circ Arrhythm Electrophys. 2010;3:380–390. doi: 10.1161/CIRCEP.110.957407. [DOI] [PubMed] [Google Scholar]
  • 119.Diness JG, Skibsbye L, Jespersen T, Bartels ED, Sorensen US, Hansen RS, Grunnet M. Effects on atrial fibrillation in aged hypertensive rats by Ca2+-activated K+ channel inhibition. Hypertension. 2011;57:1129–1135. doi: 10.1161/HYPERTENSIONAHA.111.170613. [DOI] [PubMed] [Google Scholar]
  • 120.Kohler R. Cardiovascular alterations in KCa3.1/KCa2.3-deficient mice and after acute treatment with KCa3.1/KCa2.3 activators. EDHF 2012 - 10th Anniversary Meeting. J Vas Res. 2012;49:1–54. [Google Scholar]
  • 121.Waeckel L, Beretin F, Clavreul N, Damery T, Desposito D, Kohler R, Sansilvestri-Morel P, Simonet S, Vayssettes-Courchay C, Wulff H, Verbeuren TJ, Feletou M. Chronic treatment with SKA-31, an activator of endothelial K+ channels, in a severe murine model of hypertension associated with stroke, aortic aneurisms and renal failure. EDHF 2012 - 10th Anniversary Meeting. J Vas Res. 2012;49:1–54. [Google Scholar]
  • 122.Damkjaer M, Nielsen G, Bodendiek S, Staehr M, Gramsbergen JB, de Wit C, Jensen BL, Simonsen U, Bie P, Wulff H, Kohler R. Pharmacological activation of KCa3.1/KCa2.3 channels produces endothelial hyperpolarization and lowers blood pressure in conscious dogs. Br. J. Pharmacol. 2012;165:223–234. doi: 10.1111/j.1476-5381.2011.01546.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Koegel H, Kaesler S, Burgstahler R, Werner S, Alzheimer C. Unexpected down-regulation of the hIK1 Ca2+-activated K+ channel by its opener 1-ethyl-2-benzimidazolinone in hACAT keratinocytes. Inverse effects on cell growth and proliferation. J Biol Chem. 2003;278:3323–3330. doi: 10.1074/jbc.M208914200. [DOI] [PubMed] [Google Scholar]
  • 124.Lambertsen KL, Gramsbergen JB, Sivasaravanaparan M, Ditzel N, Sevelsted-Møller LM, Oliván-Viguera A, Rabjerg M, Wulff H, Köhler R. Genetic KCa3.1-deficiency produces locomotor hyperactivity and alterations in cerebral monoamine levels. PLos One. doi: 10.1371/journal.pone.0047744. published 15 October 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES