Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


For decades it has been known that external K(+) ions are rapid and potent vasodilators that increase CBF. Recent studies have implicated the local release of K(+) from astrocytic endfeet-which encase the entirety of the parenchymal vasculature-in the dynamic regulation of local CBF during NVC. It has been proposed that the activation of KIR channels in the vascular wall by external K(+) is a central component of these hyperemic responses; however, a number of significant gaps in our knowledge remain. Here, we explore the concept that vascular KIR channels are the major extracellular K(+) sensors in the control of CBF. We propose that K(+) is an ideal mediator of NVC, and discuss KIR channels as effectors that produce rapid hyperpolarization and robust vasodilation of cerebral arterioles. We provide evidence that KIR channels, of the KIR 2 subtype in particular, are present in both the endothelial and SM cells of parenchymal arterioles and propose that this dual positioning of KIR 2 channels increases the robustness of the vasodilation to external K(+), enables the endothelium to be actively engaged in NVC, and permits electrical signaling through the endothelial syncytium to promote upstream vasodilation to modulate CBF.

Free full text 


Logo of nihpaAbout Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;
Microcirculation. Author manuscript; available in PMC 2016 Apr 1.
Published in final edited form as:
PMCID: PMC4404517
NIHMSID: NIHMS659596
PMID: 25641345

Vascular Inward Rectifier K+ Channels as External K+ Sensors in the Control of Cerebral Blood Flow

Abstract

For decades it has been known that external potassium (K+) ions are rapid and potent vasodilators that increase cerebral blood flow (CBF). Recent studies have implicated the local release of K+ from astrocytic endfeet—which encase the entirety of the parenchymal vasculature—in the dynamic regulation of local CBF during neurovascular coupling (NVC). It has been proposed that the activation of strong inward rectifier K+ (KIR) channels in the vascular wall by external K+ is a central component of these hyperemic responses; however, a number of significant gaps in our knowledge remain. Here, we explore the concept that vascular KIR channels are the major extracellular K+ sensors in the control of CBF. We propose that K+ is an ideal mediator of NVC, and discuss KIR channels as effectors that produce rapid hyperpolarization and robust vasodilation of cerebral arterioles. We provide evidence that KIR channels, of the KIR2 subtype in particular, are present in both the endothelial and smooth muscle cells of parenchymal arterioles and propose that this dual positioning of KIR2 channels increases the robustness of the vasodilation to external K+, enables the endothelium to be actively engaged in neurovascular coupling, and permits electrical signaling through the endothelial syncytium to promote upstream vasodilation to modulate CBF.

Keywords: KIR channels, neurovascular coupling, cerebral blood flow, functional hyperemia, parenchymal arteriole, capillary, smooth muscle, endothelium, astrocytic endfoot

INTRODUCTION

Brain function is dependent on the ability to match regional neuronal metabolic requirements with an appropriate level of local blood flow. When neurons become active, signaling cascades are initiated that result in hyperemia—a surge of local blood which delivers the energy substrates needed to support ongoing neuronal function. The term ‘functional hyperemia’ arose to describe these neuronally-driven blood flow responses, and the discovery of this phenomenon is attributed to Roy and Sherrington in 1890 [105]. ‘Neurovascular coupling’ (NVC) describes the molecular mechanisms underpinning neuron-to-vasculature communication, and this involves coordinated intra- and intercellular signaling between the cells of the neurovascular unit—neurons, astrocytes, vascular smooth muscle (SM), and likely also endothelial cells (ECs) and pericytes. Work throughout the last 125 years has elucidated some of the mechanisms underlying NVC. However, there exists controversy over the identities of the key mediators of this process, and many significant gaps in our knowledge still remain.

NVC demands the rapid coupling of neuronal activity to blood flow responses; therefore, an ideal mediator of this process would possess the following properties:

  • 1)

    Rapid release and onset of action.

  • 2)

    The ability to evoke robust and sustained vasodilation for the duration of its presence.

  • 3)

    Rapid off-kinetics and clearance.

One factor that fits this profile is potassium (K+) ions. It has been appreciated for many years that computationally active neurons release K+ during the repolarization phase of the action potential, which leads to fluctuations in the K+ concentration of the brain's external milieu. Moreover, it is well known that small elevations in K+ can dilate cerebral arteries and arterioles and increase brain blood flow in vivo [26,27,51,58]. Using this information as a takeoff point, our laboratory has developed a working model of K+-mediated dilation in NVC that satisfies the three criteria outlined above (Figure 1). Although our data, based on ex vivo models, provide strong experimental support for the proposed model and results obtained to date in vivo are generally consistent with this, the precise mechanisms through which K+ induces hyperemia in vivo remain poorly defined. It is also not currently clear how the various proposed vasodilatory mechanisms are integrated to achieve a coordinated hyperemic response. This latter issue is the subject of a number of excellent recent reviews and will not be addressed here [1,10,52,95].

An external file that holds a picture, illustration, etc.
Object name is nihms-659596-f0001.jpg

K+ signaling mechanisms in NVC. K+-mediated dilation (left) begins with neuronal activity which is detected by astrocytic processes adjacent to synapses leading to phospholipase C (PLC)-mediated liberation of inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG) from membrane phosphatidylinositol 4,5-bisphosphate (PIP2) pools, ultimately resulting in an IP3-mediated Ca2+ wave which propagates into astrocytic endfeet enwrapping the cerebral microcirculation [115]. This engages large-conductance calcium (Ca2+)-activated K+ (BK) channels on the endfoot plasma membrane [26,27,97], and their activation leads to an increase in the concentration of K+ in the extracellular nanospace between the endfoot and SM. The rise in external K+ activates strong inward-rectifier K+ (KIR) channels on the smooth muscle, leading to membrane hyperpolarization, closure of voltage-dependent Ca2+ channels, vasorelaxation and increased blood flow [26]. During more intense neuronal activity (right), larger Ca2+ waves promote the release of higher concentrations of K+ from the endfoot, leading to membrane depolarization, VDCC activation and constriction [27]. Under both conditions, the Na+/K+ ATPase likely contributes to K+ clearance and may provide a brief accompanying hyperpolarizing current. Adapted from [67] and [27].

In this article, we explore the concept of K+ as an ideal mediator of NVC and argue that strong inward-rectifier K+ (KIR) channels in the vascular wall are the key sensors of external K+ in the brain. We also place our proposed model—and extensions to it—in the context of the biophysical and electrophysiological properties of the KIR channel to demonstrate how these channels are capable of converting local K+ signals into profound smooth muscle (SM) membrane hyperpolarization to relax arterioles and increase CBF.

REGULATION OF CEREBRAL ARTERY DIAMETER BY EXTERNAL K+: THE VASCULAR SMOOTH MUSCLE AS A K+ ELECTRODE

The molecular mechanisms that control arteriolar diameter differ between distinct segments of the cerebral vasculature and, more broadly, throughout the vascular tree as a whole. Here, our focus is on penetrating cerebral parenchymal arterioles (PAs). PA vascular SM exhibits a steep relationship between membrane potential (Vm) and arteriolar diameter, making control of Vm central to the control of cerebral blood flow. Elevation of intravascular pressure to physiological levels—40 mm Hg for PAs—sets the SM Vm to between -35 and -40 mV [37,89]. The basis of this resting Vm lies in the balance between depolarizing ion conductances that are incrementally activated in response to increasing pressure and hyperpolarizing conductances that counteract them. Depolarization increases the activity of voltage-dependent calcium (Ca2+) channels (VDCCs), leading to an elevation of intracellular Ca2+, which engages the SM contractile machinery and thereby causes constriction. In cerebral arteries, membrane depolarization also activates voltage-dependent K+ (KV) channels [49] and large-conductance Ca2+- and voltage-sensitive K+ (BK) channels [6], which conduct K+ out of the cell and thereby provide a hyperpolarizing influence, acting as a brake on constriction. The result of the balance between depolarizing-contractile and hyperpolarizing-relaxing conductances is the constriction to pressure known as the ‘myogenic response’, and the level of ‘myogenic tone’ sets basal cerebral vessel diameter and resting CBF, which can then be modulated during NVC.

Under experimental conditions with low intravascular pressure (e.g., 5 mm Hg), at which PAs display little myogenic tone, the PA SM Vm is approximately -60 mV which is 43 mV positive to the K+ equilibrium potential (EK) of -103 mV [89], assuming 3 mM extracellular K+ [7] and 140 mM intracellular K+. According to a parallel conductance model [84,85] used to calculate the relative conductance for K+ versus other major ions (i.e., Na+, Ca2+, Cl) under these conditions, these values indicate that the K+ conductance of the membrane dominates over other conductances by about 1.4:1 (see Box 1). If intravascular pressure is then increased to physiological levels (e.g., 40 mm Hg), the SM Vm is depolarized (to between -35 and -40 mV) [37,89], and this ratio decreases to about 0.6:1, reflecting an increase in the conductance of Na+/Ca2+-permeable channels relative to K+ conductances. This pressure-induced depolarization is likely mediated by the activation of transient receptor potential (TRP) channels (e.g., TRPC6 and TRPM4) and VDCCs [28,64], promoting the influx of cations. In pial arteries, the increase in VDCC conductance is offset by activation of KV and BK channels [6,49]. In contrast, in PAs, where there is little BK channel activity at the resting potential of these arterioles [37], the increase in depolarizing conductances is offset by negative-feedback elevation of KV channel activity [116].

Elevation of external K+ depolarizes most cells. Counter-intuitively, however, small elevations of external K+ cause a striking hyperpolarization of SM cells in cerebral arteries [51] and PAs [26]. For example, elevation of external K+ from physiological cerebrospinal fluid levels (3 mM) [7] to 8 mM causes a rapid and near maximal dilation of PAs (Figure 2A) by causing a hyperpolarization from between -35 and -40 mV to close to the new EK of -76 mV (Figure 2B) [26]. To come within even 2 mV of EK under these conditions, the ratio of K+ to Na+/Cl/Ca2+ conductances would have to increase at least 58-fold—from 0.6:1 to 37:1.

An external file that holds a picture, illustration, etc.
Object name is nihms-659596-f0002.jpg

Inward-rectifier K+ channels are activated by increases in external K+. (A) In situ, raising K+ to <20 mM causes rapid and substantial vasodilation of pressurized (40 mm Hg) PAs due to KIR channel activation; further increases in K+ drive membrane depolarization and constriction. Trace from [67]. (B) The SM of PAs behaves as a K+ electrode with increasing concentrations of extracellular K+. Experimentally observed membrane potential (Vm) data (from [26,37,89]) are shown versus the potassium equilibrium potential (EK) predicted by the Nernst equation. At 3 mM K+, Vm is depolarized relative to EK due to myogenic inward cation currents. Raising K+ activates KIR channels, and the resultant K+ efflux effectively locks Vm at EK. (C) Hypothetical KIR current-voltage relationship (left) and illustration (right) showing that at basal extracellular K+ (3 mM) the pore of the KIR channel is blocked by Mg2+ or large cationic polyamines. Under these conditions (assuming 140 mM [K+]i), EK is -103, which is highly negative compared to the resting Vm of SM (approximately -35 to -40 mV) at the physiological intravascular pressure of 40 mmHg. The strong driving force for cation efflux under these conditions leads to blockade of the channel pore by the larger cations, resulting in very little channel activity. (D) When [K+]o is elevated to 8 mM (e.g., when released from the astrocytic endfoot during NVC), intracellular blockade of the channel is relieved. Channel unblock allows K+ to exit the cell, driving Vm to the new EK—where it will remain until the extracellular K+ is cleared—and leading to substantial vasodilation.

This increase in K+ conductance is enabled by the activation of KIR channels within the vascular wall (the term ‘vascular wall’ is used where the precise cellular localization of the channel—i.e., smooth muscle and/or endothelium—is not yet known). The KIR channel family consists of seven subfamilies (KIR1-7) comprising a total of 15 subunit isoforms (KIR1.1, KIR2.1-4, KIR3.1-4, KIR4.1-2, KIR5.1, KIR6.1-2, and KIR7.1). Each subunit is a two-transmembrane protein connected by a pore-forming loop with intracellular C- and N-terminal domains. Functional KIR channels are tetramers formed from homo- or heteromeric assemblies of individual subunits. KIR channels can be further separated into four groups on the basis of their functional characteristics: classical, G-protein gated, ATP-sensitive and K+ transport channels [39]. Members of this family are expressed in cells throughout the neurovascular unit (Table 1). Notably, the classical KIR2 channels are expressed in the vascular wall of PAs and cerebral arteries, and exert a strong influence over SM Vm and contractile state [5,26,51,67,99,134], with the KIR2.1 isoform apparently having a prominent role [5,134].

Table 1

KIR channel Ba2+ sensitivity, K+ activation and subunit distribution throughout the neurovascular unit. This table is not exhaustive, as some channels are formed from heteromeric subunit assemblies. Molecular studies for many subunits have not been performed for PAs, although pharmacological data are consistent with the expression of KIR channels that are activated by external K+ and blocked by <100 μM Ba2+ (i.e., KIR2 channels). Ba2+ blockade is voltage-sensitive, and where data are available we have listed IC50 values for negative potentials (note that the highly positive voltages listed for KIR1.1 and 6.1 mean that IC50 values at more negative potentials will be lower).

IsoformRectificationBa2+ IC50 (nM)Voltage (mV)[K+]o (mM)K+-activated?*NVU expression*Refs
KIR1.1Weak43000100Yes Neurons Astrocytes Parenchymal arteriole [3,12,23,48,70,110]
[check] n.d.n.d
KIR2.1Strong3.2−10060Yes [check] [check] [check] [17,26,39,47,65,67,107,108]
KIR2.2Strong0.5−10060Yes [check] [check] [check] [17,26,39,47,65,67,107,108]
KIR2.3Strong10.3−10060Yes [check] [check] n.d.[17,39,47,65,107,108]
KIR2.4Strong390−8096n.d[check] - cranial nerve nuclei x n.d.[17,65,123]
KIR3.1Strong14 (+3.4) a−13090No [check] [check] n.d.[9,47,59,80]
KIR3.2Strongn.d. bNo [check] x n.d.[9,47,80]
KIR3.3Strongn.d.No [check] x n.d.[9,47,80]
K|r3.4Strong92−60100No [check] x n.d.[9,45,47,80]
K|r4.1Intermediate7.1−13050No x [check] n.d.[9,22,38,108]
KIR4.2Intermediaten.d. bYes x x n.d.[38,94]
KIR5.1Strongn.d. bn.d[check] - cell culture [check] n.d.[38,108,121]
KIR6.1Weak89.3+605.4No x [check] x [16,41,122]
KIR6.2Weak29.3 (+SUR1)−1025.6No [check] x x [16,32,81,117,118,122]
K|r7.1Weak1100-1900−110150No [check] [check] n.d[19,55,92]
*K+-activation is defined as an increase in channel conductance in the presence of elevated [K+]o at a constant electrical driving force.
#All brain regions considered, where it is known that expression is highly restricted this is noted. No distinction is made between the smooth muscle and endothelial expression of KIR subunits in parenchymal arterioles due to the current lack of evidence.

[check] = expression confirmed, x = lack of expression observed. n.d. = not investigated to date, to the authors’ knowledge.

aKIR3.1 homomers are non-functional, association with other KIR3.x subunits is required for functional channels [39].
bKIR3.2 [114], 4.2 [94] and 5.1 [121] homomers are all inhibited by Ba2+ ions, but IC50 data are currently lacking.

KIR2 channels display negative slope conductance—meaning that their activity increases upon membrane hyperpolarization [111]—and are also activated by increases in external K+ (Figure 3A). Outward current through these channels decreases with membrane potential depolarization positive to EK. This “inward rectification” occurs because of voltage-dependent intracellular channel blockade by polyvalent cations, such as magnesium (Mg2+), and polyamines, such as spermine, spermidine and putrescine [69,72] (Figure 2C and D). The electrical potential across the membrane appears to be a key driver of this block, forcing one or more Mg2+ ions and/or polyamines into the pore of the channel, where they ultimately bind to the multiple anionic regions therein [13,69,72,101]. It has been suggested that, when external K+ is elevated, blocking particles may be driven out of the pore by electrostatic forces imparted by the increased concentration of extracellular K+ ions, which appear to interact directly with a ‘K+ coordination site’ on the outer mouth of the channel [113].

An external file that holds a picture, illustration, etc.
Object name is nihms-659596-f0003.jpg

(A) KIR channel conductance over a range of Vm values with increasing concentrations of K+. With 3 mM [K+] o at -40 mV (the approximate resting Vm; dotted arrows), KIR channel conductance is very low. Raising K+ greatly increases KIR channel activity at a given Vm; channel activity is also increased by membrane hyperpolarization. Data were plotted according to the eq. 1 in the text. (B) The relationship between external K+ concentration and KIR channel conductance based on the K+- and voltage-dependence of KIR2 channels, calculated using eq. 1 (see text), and measured effects of external K+ on PA SM Vm [26,37]. Elevation of [K+]o from 3 mM to 8 mM causes an enormous, near-maximal increase in KIR channel conductance, with further elevations to 15 and 25 mM causing only small subsequent increases in conductance. (C) Subtracted 100-μM Ba2+-sensitive currents from a freshly dissociated EC from a rat PA in response to a voltage ramp from -140 to 0 mV indicating the presence of functional KIR2 channels in these cells. External K+ was 6 mM in this experiment.

The external K+- and voltage-dependence of KIR2 channel activity have been known for quite some time [63]. Channel conductance is half-maximal at EK [98]; therefore, at a given Vm, elevation of external K+ will increase channel conductance with the rightward shift in EK (Figure 3A). Also, membrane hyperpolarization increases channel conductance e-fold for every 7.4 mV [98], presumably reflecting voltage-dependent removal of polyamine block. These two facets of KIR behavior synergize with an increase in external K+, leading to substantial increases in KIR channel conductance that can be estimated using the equation,

gKIRgKIRmax={1+e[(VV0.5)k]}1,
Eq 1

where gKIR is KIR channel conductance, gKIRmax is the maximal slope conductance for a given concentration of external K+ (proportional to {[K+]o}0.5), V is Vm, V0.5 is the half-maximal activation potential for KIR channels (equal to EK) and k is the steepness factor (equal to 7.4 mV) [98]. According to this equation, when external K+ is raised from 3 mM to 8 mM (causing PA membrane hyperpolarization from about -40 mV to -76 mV), KIR channel conductance increases approximately 4,000-fold. This enormous increase in KIR channel conductance accounts for the hyperpolarization to EK despite the simultaneous loss of KV channel conductance [116]. Increasing K+ beyond 8 mM causes a much smaller relative increase in KIR channel conductance (see Figure 3B).

Thus, the biophysical properties of the KIR2 channel explain how small increases in external K+ can act as a powerful vasodilatory force, directly generating rapid and profound hyperpolarization and dilation of the arterioles that participate in NVC (see Figure 1). Accordingly, when K+ is raised, the SM membrane acts as a K+ electrode, tracking EK (Figure 2B). Gradual increases in K+ eventually lead to membrane depolarization, which also activates voltage-sensitive KV and BK channels, further increasing K+ permeability to keep Vm clamped at EK. Pressurized cerebral arteries and arterioles begin to develop detectable constriction at about -60 mV [50]. At this potential, the open probability of VDCCs is sufficient to deliver Ca2+ to cause constriction [50,106] (see Figure 2A, K+ concentrations above 20 mM). For example, when external K+ is elevated to 25 mM, Vm/EK (approximately -46 mV) is close to the resting Vm measured with 3 mM external K+ (approximately -40 mV). As expected, the degree of constriction under each of these conditions is approximately the same (Figure 2A) [27,67]. Increasing K+ further (e.g., to 30 and 60 mM) depolarizes Vm and promotes more substantial constriction [27,67]. Overall, this leads to the characteristic biphasic relationship between external K+ and arteriolar diameter, with elevations of external K+ to lower concentrations (<20 mM) causing vasodilation and higher concentrations causing constriction. Therefore, elevation of K+ can provide a unified mechanism to account for both PA vasodilations and vasoconstrictions in response to neuronal activity [27,54] (see Figure 1).

In contrast to the hyperpolarization induced by optimal increases in K+, at lower concentrations of K+ that are insufficient for maximal KIR channel activation it is likely that a ‘tug of war’ between hyperpolarizing and depolarizing forces ensues. Accordingly, as the KIR channel activates in response to K+, the membrane would begin to hyperpolarize, shutting off KV and BK channels, and thereby counteracting the increase in K+ permeability to KIR channel activation. Provided that depolarizing influences remain relatively constant, deactivation of KV and BK channels would tend to depolarize Vm. This repolarization would, in turn, reactivate KV and BK channels, and this process would repeat, manifesting as a bi-stable SM Vm and diameter oscillations. (For example, see Figure 1, panel A in [76], where raising K+ from 6 to 7 mM results in diameter oscillations in rat cerebral arteries.) This phenomenon indicates that there is a critical level of external K+ required for stable and sustained Vm hyperpolarization and vasodilation.

The Na+/K+ ATPase ‘pump’ has also been suggested as a sensor for extracellular K+ in some peripheral vessels, such as the hamster cremaster, and rat mesenteric and tail arteries [8,128,129]. However, considering the comparatively small increase in pump activity in response to K+, and the modest and transient nature of K+-induced vasodilation that is attributable to pump activation in cerebral arteries (see Box 2), it is unlikely that Na+/K+ ATPases are major contributors to K+-induced dilation during NVC. However, it is possible that the transient hyperpolarization provided by the pump could act in synergy with KIR channel activation to ‘kick-start’ robust K+-induced vasodilations. Na+/K+ ATPases likely also aid in rapidly clearing K+ from the perivascular space after NVC (Figure 1).

WHAT ARE THE SITES OF K+ RELEASE IN THE BRAIN?

To exert an effect on the vascular wall KIR channel, K+ ions must be concentrated within an appropriate range in the restricted extracellular space between the SM and the overlying astrocytic endfeet. These endfeet completely encase the cerebral microcirculation [71], and thus represent an anatomical barrier that limits the direct diffusion of K+ ions from their site of release (neurons) to their target of action (cerebral microvessels). However, there is now mounting evidence that astrocytes act as intermediaries in this process [91,119], with the release of K+ from the endfeet in response to intracellular signaling cascades evoked by neuronal activity serving as a key mechanism of NVC (Figure 1).

In the 1980s, the concept of astrocytic K+ spatial buffering/siphoning gained traction as a plausible mechanism for the redistribution of K+ from areas of high concentration, such as around the synapse after neuronal activity, to areas of lower concentration, such as the perivascular space. According to this conceptually appealing idea, K+ was thought to be absorbed by perisynaptic astrocytic processes and then exit the cell via specialized foot processes adjacent to the brain surface and the cerebral microcirculation [56,86,87,90]. Several functions were ascribed to this phenomenon, including the maintenance of ion homeostasis around the synapse to ensure continued normal neuronal function [86,87], and a potential role in neuronal activity-evoked increases in CBF [93]. Follow-on experiments in retina and brain suggested that the molecular entity underlying glial K+ siphoning was the intermediate inward rectifier channel KIR4.1 [42,53,83,120]. However, subsequent work provided strong evidence that efflux of K+ via astrocytic KIR4.1 channels does not contribute to NVC. These experiments, performed in retina, showed that injections of depolarizing current into patch-clamped perivascular astrocytes (to drive K+ efflux through open KIR channels) had no effect on arteriole diameter, and further demonstrated that light-evoked vasodilation was unchanged in KIR4.1-knockout mice [79].

Despite this negative result, recent evidence supports the concept that functional hyperemia does indeed involve the release of K+ from astrocytic endfeet [26,27,67]. However, instead of release via KIR4.1 channels, K+ release from the endfoot is driven by astrocytic inositol 1,4,5-trisphosphate (IP3)-mediated Ca2+ waves initiated by local neuronal activity [115], which engage BK channels on the endfoot plasma membrane [26,27,97]. This increases the concentration of K+ in the extracellular nanospace between the endfoot and SM to activate SM KIR2 channels (see Figure 1). Theoretically, during NVC, a single endfoot BK channel need open for just 200 ms to release sufficient K+ ions to raise K+ in the perivascular space to 10 mM—a concentration high enough to rapidly hyperpolarize the SM Vm to -70 mV and cause substantial vasodilation [26]. In addition to BK channels, intermediate-conductance Ca2+-activated K+ (IK) channels are also expressed in astrocyte endfeet [68], providing an additional Ca2+-dependent pathway through which K+ might contribute to NVC (Box 3).

VASCULAR KIR CHANNELS AS THE EXTERNAL K+ SENSORS IN CEREBRAL BLOOD FLOW CONTROL: CURRENT EVIDENCE AND QUESTIONS ARISING

It has long been known that elevation of external K+ increases pial artery diameter in vivo [58], and that cerebral arteries [24] (and more recently PAs [26]) possess SM inward rectifier currents. Thus, a potential mechanism for increases in CBF is activation of KIR channels in the vascular wall by external K+ released into the perivascular space by astrocytic endfeet in response to neuronal activity, as described in Figure 1. In support of this hypothesis, the KIR channel blocker barium (Ba2+) at 100 μM, a concentration that is selective for KIR2 isoforms (see Table 1), prevents increases in local CBF to whisker stimulation or the metabotropic glutamate receptor agonist trans-ACPD [27]. Ba2+ at this concentration also inhibits pial artery dilation in vivo in response to sciatic nerve stimulation [124]. Further support for the involvement of this K+-to-KIR mechanism is demonstrated by the ability of Ba2+ to inhibit vasodilation evoked by neuronal stimulation in brain slice preparations (by 50-70%) [26,67]. Notably, the salient features of this mechanism have been successfully recapitulated by computer models [131,132]. These studies suggest that KIR channel activity in response to endfoot-derived K+ is indeed a vitally important mechanism for NVC, although the possibility that K+ could act indirectly through Ba2+-sensitive KIR channels in another cell type, such as neurons or astrocytes (Table 1), cannot currently be discounted. A fuller understanding of the molecular details underlying this mechanism awaits an answer to this and a number of other important questions.

One key question with important implications for the operation of the KIR channel-dependent vasodilatory circuitry is whether this channel is localized to the PA SM, endothelium or both. In other vascular beds, KIR channels have been localized to ECs alone (such as in the mesenteric artery [15]), or have been found in both SM and ECs (such as in hamster cremaster arterioles [8,44]). The existence of KIR channels in both SM and ECs may provide a mutually reinforcing amplifier for the hyperpolarization evoked by small elevations in external K+. This reinforcing amplification would, in turn, be predicted to dampen the tug of war between KIR and KV/BK channels that ensues following KIR channel activation. Although KIR channel expression and function in PA endothelium have not been reported in the literature, we have obtained recordings of Ba2+ (100 μM)-sensitive KIR currents in freshly isolated ECs from PAs (Figure 3C), suggesting that functional KIR2 channels are present in this cell type as well as SM cells. While this is not yet definitive evidence (as it is possible that the ion channel properties of acutely isolated cells may differ to those of intact arterioles), it nonetheless sets the stage for a situation in which K+ released from the endfoot could also recruit endothelial KIR channels in addition to activating SM KIR channels, thereby helping to stabilize the hyperpolarized SM Vm and eliminate diameter oscillations, especially at lower levels of elevated K+. In arteries of the peripheral microcirculation, endothelial KIR channels have also been implicated in conducted vasodilatory signaling within and along the vessel wall [31,44]. In this setting, KIR channels transmit EC hyperpolarization to the SM via myo-endothelial gap junctions and along the length of the vessel through endothelial-endothelial gap junctions. Extending this communication paradigm to PAs, it is possible to hypothesize that K+ released during NVC engages both the SM and endothelium through KIR2 channels expressed in these two cellular compartments, leading to a regenerative hyperpolarization along the endothelial lining of PAs up to pial arteries on the brain surface that would simultaneously also be transmitted to the SM through gap junctions. If this hypothesis were confirmed, it would support a major functional contribution of the endothelium to NVC. In support of this concept, Hillman and colleagues recently demonstrated that light-dye disruption of endothelial function halts the propagation of pial artery vasodilation in vivo in response to electrical hindpaw stimulation, thereby blunting increases in surface CBF [11]. Additionally, if brain capillary ECs are found to express KIR channels and exhibit electrical coupling, it would argue that these channels also play a role in conducted signaling from the capillary bed upstream to PAs, as has been suggested to occur in other capillary beds [20], providing a unified mechanism for vasoconduction throughout the entire vascular network of the brain. This overall concept is illustrated in Figure 4.

An external file that holds a picture, illustration, etc.
Object name is nihms-659596-f0004.jpg

Proposed scheme for KIR channels as K+ sensors to control cerebral blood flow. K+ ions released from astrocytic endfeet and neurons activates KIR2 channels present on SM cells and ECs in PAs and (possibly) on capillary ECs. Hyperpolarization caused by KIR engagement may then spread bi-directionally throughout the vascular syncytium, causing further KIR channel activation while at the same time deactivating KV and BK channels and VDCCs in the SM. This locks the membrane potential at EK until K+ is cleared and causes near maximal vasodilation and a substantial increase in cerebral blood flow.

Which KIR channel isoforms are functionally expressed in the SM and endothelium of the PA wall? We have detected transcripts for both KIR2.1 and 2.2 in PA homogenates [67], but to our knowledge, the potential expression of KIR2.3 and 2.4 isoforms in these vessels has yet to be explored. Functional data indicate that the Ba2+ sensitivity of SM inward-rectifier currents match those carried by KIR2.1-2.3 channels [26,67] (Table 1). Further, a study of pial arteries showed that global knockout of KIR2.1 resulted in a complete loss of KIR currents in isolated SM cells and K+-mediated dilations. In contrast, knockout of KIR2.2 had no effect on K+-mediated vasodilation in this vascular bed [134]. These results suggest that the KIR2.1 subunit is a vital component of KIR channels in pial arteries. However, the KIR2.2 subunit does appear to be expressed at the mRNA level in both pial arteries and PAs where, interestingly, transcripts for this subunit are more abundant than those for KIR2.1 [67,111,133]. Accordingly, it has been suggested that SM KIR channels in cerebral vessels may be composed of heteromers of KIR2.1 and 2.2 subunits under normal conditions [133]. In any case, it appears that the KIR2.1 subunit is essential for the formation of functional channels. In pial arteries, Wu et al. [133] observed KIR2.1, 2.2 and 2.4 subunit transcripts in endothelium-intact homogenates, whereas KIR2.4 expression was absent in SM-only homogenates. Whether expression of KIR2.4 mRNA in the endothelium, as implied by these results, translates to a functional role for this channel in these cells was not established by these studies. However, the Ba2+ (100 μM) sensitivity of the currents we observed in ECs from PAs would at least indicate that functional KIR2.4 homomers (Ba2+ IC50 = 390 μM [123]) do not likely contribute to the EC currents recorded from these arterioles.

The presence of KIR channels with identical pharmacological profiles in both SM cells and ECs of the cerebral microcirculation presents a technical challenge for accurately dissecting the relative contribution of SM and EC KIR channels to NVC. Furthermore, KIR2 family members are expressed in neurons and astrocytes in the CNS (Table 1) [17,40,47,107,108]. Global KIR-knockout mice are thus not the answer. Even if they were, germline ablation of the gene for KIR2.1—the predominant vascular isoform—leads to the development of a cleft palate; as a result, newborn mice fail to suckle and die shortly after birth [134]. To circumvent this problem, we have generated a floxed KIR2.1 mouse that we have crossed with mice expressing Cre recombinase under the control of EC- or SM cell-specific promoters. This cell-specific knockout approach will ultimately enable us to discern the specific role of the KIR channel in each vascular cell type. These conditional mouse models are also less likely to be influenced by the development of compensatory changes, which are a concern with germline knockouts.

DISRUPTION OF NEUROVASCULAR K+ COMMUNICATION IN DISEASE: LESSONS FROM A CHRONIC STRESS MODEL

Too little or too much K+ signaling is a plausible mechanism for the loss of control of CBF in a range of brain disorders. An interesting recent study demonstrates this principal in glioma tissue. Here, the elevation of perivascular K+ caused by release from cancerous glial cells, which migrate along the vascular wall and displace healthy astrocytic endfeet, results in a loss of control of vascular tone to the invading tumor [127]. Similarly, pathological conditions such as cortical spreading depression are accompanied by an extracellular K+ wave that increases K+ from 3 mM to as high as 60 mM [112], which might contribute to increases or decreases in CBF depending on the concentration of external K+ [2,60,61,74].

Pathologies are also capable of disrupting KIR2 channel expression and function in many different tissues and vascular beds, making this channel a potential therapeutic target in a broad range of diseases. Prominent among these is Andersen-Tawil syndrome, caused by a mutation in KIR2.1 and characterized by cardiac arrhythmias, dysmorphic features and periodic paralysis [96]. Additionally, KIR channel function is impaired in mesenteric arteries [130] and posterior cerebral arteries [76] in hypertension, and KIR channel activity in pial arteries is suppressed by protein kinase C overactivity in streptozotocin-induced diabetes [124]. Very recent evidence from our laboratory suggests that K+ signaling during NVC can be disrupted in PAs at the level of the SM KIR channel in the context of a rodent model of chronic stress [67].

A substantial body of evidence has demonstrated that long-term exposure to stressful stimuli has a negative impact on human health, with links to both psychopathologies and cardiovascular disease [14,18,21,57,66,78,100]. ‘Stress’ is a vague term that has been used in a wide variety of contexts, but here stress refers to any real or psychologically perceived threat to an organism's homeostasis that prompts a stereotyped ‘fight or flight’ response.

The amygdala is one of several brain regions that is engaged during stressful events [62,103], and stressor processing here ultimately leads to an increase in sympathetic nervous system tone and hypothalamic-pituitary adrenal (HPA) axis output, resulting in a “fight or flight response”. This response includes an immediate increase in circulating catecholamines—which act rapidly to stimulate an increase in heart and breathing rate, shut down digestion, increase blood pressure and shunt blood to major muscle groups—and a slower increase in glucocorticoid hormones, principally corticosterone in rodents and cortisol in humans [18]. The neuronal effects of repeated exposure to stressors have been well explored [18]. In contrast, comparatively little work has been devoted to investigating the effects of stress on other cells of the brain, such as astrocytes, and the SM and ECs and of PAs have received even less attention.

With this in mind, we examined the possible impact of stress on cells of the neurovascular unit and on NVC by applying a chronic stress model to male rats. Subjects were exposed to one of five heterotypical stressors per day for a total of seven days. Rats that were exposed to this paradigm developed an anxious behavioral phenotype, in line with previous observations [34,102]. In brain slices prepared from stressed rats, we observed a marked impairment of NVC in the amygdala. Intriguingly, neuronal activity-evoked vasodilation was substantially blunted after stress, whereas the preceding astrocytic endfoot Ca2+ wave was elevated. This enhanced Ca2+ wave is likely the result of stress-induced amygdalar neuron hyperactivity [33,104], a phenomenon in which electrical stimulation may elicit more intense network activity that is transduced into larger Ca2+ signals in local astrocytes. It is also possible that stress-related signaling may alter astrocytic Ca2+ handling, for example by increasing the sensitivity or number of IP3 receptors in the endoplasmic reticulum.

As discussed above (Figure 1), astrocytic Ca2+ waves lead to the release of K+ from the endfoot, which then activates KIR channels on the juxtaposed SM to cause membrane hyperpolarization, leading to relaxation and vasodilation [26]. After stress, not only were isolated arterioles substantially less sensitive to K+, but also Ba2+ no longer inhibited NVC in slices, strongly suggesting a loss of KIR channel function. This was confirmed by experiments examining KIR channel current density in isolated PA SM cells, which indicated fewer functional channels in the membrane after stress, and by molecular experiments, which suggested that the KIR2.1 isoform was downregulated in PAs by stress [67].

How might KIR channel expression/function be reduced by stress? SM cells and ECs of PAs possess abundant glucocorticoid receptors (GRs) and mineralocorticoid receptors (MRs) [29,30,77,88]. Because it readily crosses the blood-brain barrier, corticosterone, which in rats is elevated for prolonged periods after stressor exposure, could plausibly act at these vascular receptors. At basal corticosterone concentrations, MRs are completely occupied, whereas the lower-affinity GRs are available for binding elevated corticosterone after stressor exposure [18]. By inhibiting GRs or exogenously applying corticosterone in the absence of stressors, we were able to protect against or mimic NVC impairment by stress, respectively. These findings support a model in which stress-induced corticosterone-GR signaling causes down-regulation of KIR channel gene expression in the SM, leading to fewer functional channels in the membrane. As a consequence, when a neuronally evoked Ca2+ wave arrives at the astrocytic endfoot and stimulates the release of K+, the SM is no longer equipped to robustly respond, leading to a blunted vasodilatory response during NVC [67]. As we did not examine EC function in this study, we cannot directly rule out a role for EC KIR channels in the stressed phenotype. However, as a blunted K+ induced dilation remains in PAs after stress [67], this is consistent with the concept that that EC KIR channels are present and possibly suggests that they are even protected from stress, although further work is required to examine this possibility.

The consequences of this impairment remain to be elucidated, although it is reasonable to speculate that the loss of SM KIR channels could have a negative impact on neuronal function if it were to remain chronic. Although we did not directly test amygdalar blood flow during NVC in our model, our results would predict a substantial decrease with stress. Thus, the resulting limited availability of oxygen and glucose could contribute to neuronal injury over time.

Because stress is a factor in many brain disorders, including dementia [46], depression [35], schizophrenia [125], anxiety disorders [109] and multiple sclerosis [82], the loss of SM KIR channels and NVC impairment may also be a contributory factor in such pathologies. The loss of KIR channel function may also occur in other disorders through glucocorticoid-independent mechanisms, such as altered levels of intracellular polyamines, suppression by membrane cholesterol [36] or changes in channel phosphorylation by protein kinase C [124], leading to the disruption of NVC and CBF.

CONCLUSIONS

Our understanding of K+-induced cerebral hyperemia has advanced greatly over the past four decades. Recent work has established K+-to-KIR signaling as a rapid and robust mechanism for eliciting vasodilation and increasing blood flow in response to neuronal activity in vivo. Central to this mechanism are the KIR2 channels positioned in the vascular wall (SM and ECs) adjacent to astrocytic endfeet—key sites of K+ release during periods of increased neuronal activity. Cerebrovascular KIR 2 channels respond to small increases in extracellular K+ with a profound increase in conductance, effectively clamping the SM Vm at the new hyperpolarized EK. This hyperpolarization drives substantial vasodilation, leading to a rapid increase in blood flow. The KIR channel is also sensitive to disruption. Indeed, channel expression and function in the vascular wall can be dramatically altered by stress, and may also be disrupted in a broad range of brain pathologies. Thus, targeting KIR2 channels in the vascular wall with the aim of restoring normal hemodynamic function may be a future therapeutic option for such disorders.

Cell-specific genetic ablation models will enable further advances in our understanding of the role of vascular KIR channels in brain in the control of CBF. In the past, the sole pharmacological agent available for testing KIR channel functionality was its pore-blocker, Ba2+; but the recently developed, selective inhibitor ML133 [126] now adds to the pharmacological armamentarium available for functional studies of KIR2 channels. Little is known of the nature and roles of ion channels—in particular KIR channels—in native brain capillary ECs or native brain pericytes (although KIR currents have been observed in retinal pericytes [73]); yet, this information may prove to be critical to a full understanding of the regulation of CBF by K+. The outcome of future studies along these lines may reveal a robust KIR channel-dependent mechanism that allows membrane hyperpolarization to be transmitted from the capillary bed all the way to the brain surface, providing an effective mechanism for engaging the entire vascular tree during NVC.

Box 1. A simple parallel conductance model for the estimation of K+ conductance versus depolarizing ion conductances

To estimate the K+ conductance of SM relative to other major ionic conductances (i.e., Na+, Ca2+ and Cl), we used the following equation [84,85]:

Vm=(GKEK)+(GCEC)(GK+GC)

where GK is the K+ conductance of the membrane and GC represents the combination of other major ionic conductances, with an assumed reversal potential of 0 mV. GK:GC was solved for Vm values derived from experimental data under a range of conditions, as noted in the text.

Box 2. Potential contribution of the Na+/K+ ATPase to K+-induced dilations during NVC

Functional Na+/K+ ATPase pumps are complexes composed of one of four α-subunit isoforms plus one of three ß-subunits, with the possible addition of an FXYD-domain-containing γ-subunit. The molecular composition of the pump dictates its sensitivity to external K+ (along with [Na+]i and ATP), with α2-containing (K0.5 = 3.6-4.8 mM) and α3-containing (~5.3-6.2 mM) pumps possessing a lower affinity for K+ than α1-containing pumps (~2 mM) [4]. With 3 mM basal [K+]o, Na+/K+ ATPases containing any α isoform could theoretically be recruited by K+ increases. However, in contrast to the robust KIR channel activation that would occur under these circumstances, K+ evokes only a transient activation of the Na+/K+ ATPase, because the rate of pumping is limited by [Na+]i. Thus, increasing [K+]o will increase the rate of pumping, providing a hyperpolarizing current, while [Na+]i will simultaneously fall, thereby curtailing the pump current [25]. Accordingly, a plausible role for the Na+/K+ ATPase during NVC may be to provide a brief accompanying hyperpolarizing current at the beginning of a K+-induced dilation that might ensure the robustness and fidelity of the response; however, sustained pump activation will not occur. Consistent with this idea, in rat cerebral arteries, small, transient dilations attributable to Na+/K+ ATPase activity occur when stepping [K+]o in 1-mM increments from 0 to 4 mM. However, subsequent elevations do not have this effect (but do engage KIR channel activity), suggesting that the native Na+/K+ ATPase is saturated at concentrations above 4 mM [75]. Although the molecular identity of Na+/K+ ATPase subunit isoforms has not been investigated here, these data suggest that α1- containing pumps (K0.5 = ~2 mM [4] and therefore saturated at 4 mM) are the predominant isoform in these arteries. Whether this is also the case for PAs and the brain capillary endothelium awaits investigation.

Box 3. Roles for IK channels in NVC

Of the members of the Ca2+-activated K+ channel family, IK channels are the most Ca2+-sensitive, responding to small increases in intracellular Ca2+ [43] on the order of those that evoke vasodilation in the neurovascular unit [27]. Thus, these channels may also be activated by Ca2+ waves entering the endfoot and could conceivably contribute to K+ release during NVC. Indeed, blockers of this channel inhibit arteriolar dilation in brain slices and functional hyperemia in vivo by 40-50%, suggesting input from IK channels during NVC [68]. While the most straightforward interpretation of these results is that astrocytic IK channels contribute to the neurovascular signaling cascade, this has not yet been unequivocally demonstrated. Thus, it remains possible that IK channels in other cell types may also play a role in NVC. The other major site of IK channel expression in the neurovascular unit is the vascular endothelium, which recent evidence suggests may actively participate in NVC [11]. Indeed, inhibiting IK channels in isolated PAs causes a tonic constriction, suggesting a degree of channel activity under basal conditions. Conversely activation of these channels evokes maximal vasodilation [37], highlighting the fact that endothelial IK channel activity can exert powerful control over vascular diameter. Therefore, if the endothelium were 7 indeed actively engaged during NVC, recruitment of IK channel signaling would be a robust mechanism for influencing the contractile state of the SM.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge David Hill-Eubanks, Nathan Tykocki, Fabrice Dabertrand, Masayo Koide and Albert Gonzales for their insightful discussion and editorial assistance. This work was supported by grants from the American Heart Association (12POST12090001 and 14POST20480144; to TAL), the Totman Medical Research Trust (to MTN), the Fondation Leducq to (MTN), and the National Institutes of Health (P20-RR-16435, P01-HL-095488, R01-HL-044455, R01-HL-098243 and R37-DK-053832; to MTN).

Abbreviations used

Ba2+barium ions
BKlarge-conductance calcium-activated potassium channel
Ca2+calcium ions
CBFcerebral blood flow
Clchloride ions
ECendothelial cell
EKpotassium equilibrium potential
GRglucocorticoid receptor
HPAhypothalamic pituitary adrenal
IKintermediate-conductance calcium-activated potassium channel
IP3inositol 1,4,5-trisphosphate
K+potassium ions
KIRinward rectifier potassium channel
KVvoltage-dependent potassium channel
Mg2+magnesium ions
MRmineralocorticoid receptor
Na+sodium ions
NVCneurovascular coupling
PAparenchymal arteriole
SMsmooth muscle
VDCCvoltage-dependent calcium channel
Vmmembrane potential

REFERENCES

1. Attwell D, Buchan AM, Charpak S, Lauritzen M, MacVicar BA, Newman EA. Glial and neuronal control of brain blood flow. Nature. 2010;468:232–243. [Europe PMC free article] [Abstract] [Google Scholar]
2. Back T, Kohno K, Hossmann KA. Cortical negative DC deflections following middle cerebral artery occlusion and KCl-induced spreading depression: effect on blood flow, tissue oxygenation, and electroencephalogram. J Cereb Blood Flow Metab. 1994;14:12–19. [Abstract] [Google Scholar]
3. Bhandari S, Hunter M. Biophysical effects of pore mutations of ROMK1. Clin Sci. 2001;101:121–30. [Abstract] [Google Scholar]
4. Blanco G, Mercer RW. Isozymes of the Na-K-ATPase: heterogeneity in structure, diversity in function. Am J Physiol. 1998;275:F633–50. [Abstract] [Google Scholar]
5. Bradley KK, Jaggar JH, Bonev AD, Heppner TJ, Flynn ER, Nelson MT, Horowitz B. Kir2.1 encodes the inward rectifier potassium channel in rat arterial smooth muscle cells. J Physiol. 1999;515:639–651. [Abstract] [Google Scholar]
6. Brayden JE, Nelson MT. Regulation of arterial tone by activation of calcium-dependent potassium channels. Science. 1992;256:532–535. [Abstract] [Google Scholar]
7. Brown PD, Davies SL, Speake T, Millar ID. Molecular mechanisms of cerebrospinal fluid production. Neuroscience. 2004;129:955–968. [Europe PMC free article] [Abstract] [Google Scholar]
8. Burns WR, Cohen KD, Jackson WF. K+-induced dilation of hamster cremasteric arterioles involves both the Na+/K+-ATPase and inward-rectifier K+ channels. Microcirculation. 2004;11:279–293. [Europe PMC free article] [Abstract] [Google Scholar]
9. Butt AM, Kalsi A. Inwardly rectifying potassium channels (Kir) in central nervous system glia: a special role for Kir4.1 in glial functions. J Cell Mol Med. 2006;10:33–44. [Europe PMC free article] [Abstract] [Google Scholar]
10. Cauli B, Hamel E. Revisiting the role of neurons in neurovascular coupling. Front Neuroenerg. 2010;2:1–9. [Europe PMC free article] [Abstract] [Google Scholar]
11. Chen BR, Kozberg MG, Bouchard MB, Shaik MA, Hillman EMC. A critical role for the vascular endothelium in functional neurovascular coupling in the brain. JAHA. 2014;3:e000787–e000787. [Europe PMC free article] [Abstract] [Google Scholar]
12. Choe H, Sackin H, Palmer LG. Permeation and gating of an inwardly rectifying potassium channel. Evidence for a variable energy well. J Gen Physiol. 1998;112:433–446. [Europe PMC free article] [Abstract] [Google Scholar]
13. Clarke OB, Caputo AT, Hill AP, Vandenberg JI, Smith BJ, Gulbis JM. Domain reorientation and rotation of an intracellular assembly regulate conduction in Kir potassium channels. Cell. 2010;141:1018–1029. [Abstract] [Google Scholar]
14. Cohen S, Janicki-Deverts D, Miller GE. Psychological stress and disease. JAMA. 2007;298:1685–1687. [Abstract] [Google Scholar]
15. Crane GJ, Walker SD, Dora KA, Garland CJ. Evidence for a Differential Cellular Distribution of Inward Rectifier K Channels in the Rat Isolated Mesenteric Artery. J Vasc Res. 2003;40:159–168. [Abstract] [Google Scholar]
16. Dabertrand F, Nelson MT, Brayden JE. Acidosis dilates brain parenchymal arterioles by conversion of calcium waves to sparks to activate BK channels. Circ Res. 2012;110:285–294. [Europe PMC free article] [Abstract] [Google Scholar]
17. Day M, Carr DB, Ulrich S, Ilijic E, Tkatch T, Surmeier DJ. Dendritic excitability of mouse frontal cortex pyramidal neurons is shaped by the interaction among HCN, Kir2, and Kleak channels. J Neurosci. 2005;25:8776–8787. [Europe PMC free article] [Abstract] [Google Scholar]
18. de Kloet ER, Joëls M, Holsboer F. Stress and the brain: from adaptation to disease. Nat Rev Neurosci. 2005;6:463–475. [Abstract] [Google Scholar]
19. Derst C, Hirsch JR, Preisig-Müller R, Wischmeyer E, Karschin A, Döring F, Thomzig A, Veh RW, Schlatter E, Kummer W, Daut J. Cellular localization of the potassium channel Kir7.1 in guinea pig and human kidney. Kidney Int. 2001;59:2197–2205. [Abstract] [Google Scholar]
20. Dietrich HH, Tyml K. Capillary as a communicating medium in the microvasculature. Microvasc Res. 1992;43:87–99. [Abstract] [Google Scholar]
21. Dimsdale JE. Psychological stress and cardiovascular disease. J Am Coll Cardiol. 2008;51:1237–1246. [Europe PMC free article] [Abstract] [Google Scholar]
22. Djukic B, Casper KB, Philpot BD, Chin LS, McCarthy KD. Conditional knock-out of Kir4.1 leads to glial membrane depolarization, inhibition of potassium and glutamate uptake, and enhanced short-term synaptic potentiation. J Neurosci. 2007;27:11354–11365. [Europe PMC free article] [Abstract] [Google Scholar]
23. Doi T, Fakler B, Schultz JH, Schulte U, Brandle U, Weidemann S, Zenner HP, Lang F, Ruppersberg JP. Extracellular K+ and intracellular pH allosterically regulate renal Kir1.1 channels. J Biol Chem. 1996;271:17261–17266. [Abstract] [Google Scholar]
24. Edwards FR, Hirst GD, Silverberg GD. Inward rectification in rat cerebral arterioles; involvement of potassium ions in autoregulation. J Physiol. 1988;404:455–466. [Abstract] [Google Scholar]
25. Eisner DA, Lederer WJ, Vaughan-Jones RD. The effects of rubidium ions and membrane potentials on the intracellular sodium activity of sheep Purkinje fibres. J Physiol. 1981;317:189–205. [Abstract] [Google Scholar]
26. Filosa JA, Bonev AD, Straub SV, Meredith AL, Wilkerson MK, Aldrich RW, Nelson MT. Local potassium signaling couples neuronal activity to vasodilation in the brain. Nat Neurosci. 2006;9:1397–1403. [Abstract] [Google Scholar]
27. Girouard H, Bonev AD, Hannah RM, Meredith A, Aldrich RW, Nelson MT. Astrocytic endfoot Ca2+ and BK channels determine both arteriolar dilation and constriction. Proc Natl Acad Sci USA. 2010;107:3811–3816. [Europe PMC free article] [Abstract] [Google Scholar]
28. Gonzales AL, Yang Y, Sullivan MN, Sanders L, Dabertrand F, Hill-Eubanks DC, Nelson MT, Earley S. A PLCγ1-dependent, force-sensitive signaling network in the myogenic constriction of cerebral arteries. Sci Signal. 2014;7:ra49. [Europe PMC free article] [Abstract] [Google Scholar]
29. Goodwin JE, Feng Y, Velazquez H, Sessa WC. Endothelial glucocorticoid receptor is required for protection against sepsis. Proc Natl Acad Sci USA. 2013;110:306–311. [Europe PMC free article] [Abstract] [Google Scholar]
30. Goodwin JE, Zhang J, Geller DS. A critical role for vascular smooth muscle in acute glucocorticoid-induced hypertension. J Am Soc Nephrol. 2008;19:1291–1299. [Europe PMC free article] [Abstract] [Google Scholar]
31. Goto K, Rummery NM, Grayson TH, Hill CE. Attenuation of conducted vasodilatation in rat mesenteric arteries during hypertension: role of inwardly rectifying potassium channels. J Physiol. 2004;561:215–231. [Abstract] [Google Scholar]
32. Gribble FM, Ashfield R, Ammälä C, Ashcroft FM. Properties of cloned ATP-sensitive K+ currents expressed in Xenopus oocytes. J Physiol. 1997;498:87–98. [Abstract] [Google Scholar]
33. Guo YY, Liu SB, Cui GB, Ma L, Feng B, Xing JH, Yang Q, Li XQ, Wu YM, Xiong LZ, Zhang W, Zhao MG. Acute stress induces down-regulation of large-conductance Ca2+− activated potassium channels in the lateral amygdala. J Physiol. 2012;590:875–886. [Abstract] [Google Scholar]
34. Hammack SE, Cheung J, Rhodes KM, Schutz KC, Falls WA, Braas KM, May V. Chronic stress increases pituitary adenylate cyclase-activating peptide (PACAP) and brain-derived neurotrophic factor (BDNF) mRNA expression in the bed nucleus of the stria terminalis (BNST): roles for PACAP in anxiety-like behavior. Psychoneuroendocrinology. 2009;34:833–843. [Europe PMC free article] [Abstract] [Google Scholar]
35. Hammen C. Stress and depression. Annu Rev Clin Psychol. 2005;1:293–319. [Abstract] [Google Scholar]
36. Han H, Rosenhouse-Dantsker A, Gnanasambandam R, Epshtein Y, Chen Z, Sachs F, Minshall RD, Levitan I. Silencing of Kir2 channels by caveolin-1: cross-talk with cholesterol. J Physiol. 2014;592:4025–4038. [Abstract] [Google Scholar]
37. Hannah RM, Dunn KM, Bonev AD, Nelson MT. Endothelial SKCa and IKCa channels regulate brain parenchymal arteriolar diameter and cortical cerebral blood flow. J Cereb Blood Flow Metab. 2010;31:1175–86. [Europe PMC free article] [Abstract] [Google Scholar]
38. Hibino H, Fujita A, Iwai K, Yamada M, Kurachi Y. Differential assembly of inwardly rectifying K+ channel subunits, Kir4.1 and Kir5.1, in brain astrocytes. J Biol Chem. 2004;279:44065–44073. [Abstract] [Google Scholar]
39. Hibino H, Inanobe A, Furutani K, Murakami S, Findlay I, Kurachi Y. Inwardly Rectifying Potassium Channels: Their Structure, Function, and Physiological Roles. Physiol Rev. 2010;90:291–366. [Abstract] [Google Scholar]
40. Howe MW, Feig SL, Osting SMK, Haberly LB. Cellular and subcellular localization of Kir2.1 subunits in neurons and glia in piriform cortex with implications for K+ spatial buffering. J Comp Neurol. 2008;506:877–893. [Abstract] [Google Scholar]
41. Ishida-Takahashi A, Otani H, Takahashi C, Washizuka T, Tsuji K, Noda M, Horie M, Sasayama S. Cystic fibrosis transmembrane conductance regulator mediates sulphonylurea block of the inwardly rectifying K+ channel Kir6.1. J Physiol. 1998;508:23–30. [Abstract] [Google Scholar]
42. Ishii M, Horio Y, Tada Y, Hibino H, Inanobe A, Ito M, Yamada M, Gotow T, Uchiyama Y, Kurachi Y. Expression and Clustered Distribution of an Inwardly Rectifying Potassium Channel, KAB-2/Kir4.1, on Mammalian Retinal Müller Cell Membrane: Their Regulation by Insulin and Laminin Signals. J Neurosci. 1997;17:7725–35. [Europe PMC free article] [Abstract] [Google Scholar]
43. 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. [Europe PMC free article] [Abstract] [Google Scholar]
44. Jackson WF. Potassium Channels in the Peripheral Microcirculation. Microcirc. 2005;12:113–127. [Europe PMC free article] [Abstract] [Google Scholar]
45. Ji S, John SA, Lu Y, Weiss JN. Mechanosensitivity of the Cardiac Muscarinic Potassium Channel: A novel property conferred by Kir3.4 subunit. J Biol Chem. 1998;273:324–1328. [Abstract] [Google Scholar]
46. Johansson L, Guo X, Waern M, Ostling S, Gustafson D, Bengtsson C, Skoog I. Midlife psychological stress and risk of dementia: a 35-year longitudinal population study. Brain. 2010;133:2217–2224. [Abstract] [Google Scholar]
47. Karschin C, Dißmann E, Stühmer W, Karschin A. IRK (1–3) and GIRK (1–4) inwardly rectifying K+ channel mRNAs are differentially expressed in the adult rat brain. J Neurosci. 1996;16:3559–3570. [Europe PMC free article] [Abstract] [Google Scholar]
48. Kenna S, Ho K, Hebert S, Ashcroft SJ, Ashcroft FM. Differential expression of the inwardly-rectifying K-channel ROMK1 in rat brain. Brain Res Mol Brain Res. 1994;24:353–6. [Abstract] [Google Scholar]
49. Knot HJ, Nelson MT. Regulation of membrane potential and diameter by voltage-dependent K+ channels in rabbit myogenic cerebral arteries. Am J Physiol. 1995;269:H348–55. [Abstract] [Google Scholar]
50. Knot HJ, Nelson MT. Regulation of arterial diameter and wall [Ca2+] in cerebral arteries of rat by membrane potential and intravascular pressure. J Physiol. 1998;508:199–209. [Abstract] [Google Scholar]
51. Knot H, Zimmermann P, Nelson MT. Extracellular K+-induced hyperpolarizations and dilatations of rat coronary and cerebral arteries involve inward rectifier K+ channels. J Physiol. 1996;492:419–430. [Abstract] [Google Scholar]
52. Koehler RC, Roman RJ, Harder DR. Astrocytes and the regulation of cerebral blood flow. Trends Neurosci. 2009;32:160–169. [Abstract] [Google Scholar]
53. Kofuji P, Ceelen P, Zahs KR, Surbeck LW, Lester HA, Newman EA. Genetic inactivation of an inwardly rectifying potassium channel (Kir4.1 subunit) in mice: phenotypic impact in retina. J Neurosci. 2000;20:5733–5740. [Europe PMC free article] [Abstract] [Google Scholar]
54. Koide M, Bonev A, Nelson MT, Wellman GC. Inversion of neurovascular coupling by subarachnoid blood depends on large-conductance Ca2+-activated K+ (BK) channels. Proc Natl Acad Sci USA. 2012;109:E1387–E1395. [Europe PMC free article] [Abstract] [Google Scholar]
55. Krapivinsky G, Medina I, Eng L, Krapivinsky L, Yang Y, Clapham DE. A novel inward rectifier K+ channel with unique pore properties. Neuron. 1998;20:995–100. [Abstract] [Google Scholar]
56. Kuffler SW. Neuroglial cells: physiological properties and a potassium mediated effect of neuronal activity on the glial membrane potential. Proc R Soc Lond B Biol Sci. 1967;168:1–21. [Abstract] [Google Scholar]
57. Kulkarni S, O'Farrell I, Erasi M, Kochar MS. Stress and hypertension. WMJ. 1998;97:34–38. [Abstract] [Google Scholar]
58. Kuschinsky W, Wahl M, Bosse O, Thurau K. Perivascular potassium and pH as determinants of local pial arterial diameter in cats. A microapplication study. Circ Res. 1972;31:240–247. [Abstract] [Google Scholar]
59. Lancaster MK, Dibb KM, Quinn CC, Leach R, Lee JK, Findlay JB, Boyett MR. Residues and mechanisms for slow activation and Ba2+ block of the cardiac muscarinic K+ channel, Kir3.1/Kir3.4. J Biol Chem. 2000;275:35831–35839. [Abstract] [Google Scholar]
60. Lauritzen M. Long-lasting reduction of cortical blood flow of the brain after spreading depression with preserved autoregulation and impaired CO2 response. J Cereb Blood Flow Metab. 1984;4:546–54. [Abstract] [Google Scholar]
61. Lauritzen M. Regional cerebral blood flow during cortical spreading depression in rat brain: increased reactive hyperperfusion in low-flow states. Acta Neurologica Scandinavica. 1987;75:1–8. [Abstract] [Google Scholar]
62. LeDoux JE. Emotion circuits in the brain. Annu Rev Neurosci. 2000;23:155–184. [Abstract] [Google Scholar]
63. Leech CA, Stanfield PR. Inward rectification in frog skeletal muscle fibres and its dependence on membrane potential and external potassium. J Physiol. 1981;319:295–309. [Abstract] [Google Scholar]
64. Li Y, Baylie RL, Tavares MJ, Brayden JE. TRPM4 channels couple purinergic receptor mechanoactivation and myogenic tone development in cerebral parenchymal arterioles. J Cereb Blood Flow Metab. 2014;34:1706–1714. [Europe PMC free article] [Abstract] [Google Scholar]
65. Liu GX, Derst C, Schlichthörl G, Heinen S, Seebohm G, Brüggemann A, Kummer W, Veh RW, Daut J, Preisig-Müller R. Comparison of cloned Kir2 channels with native inward rectifier K+ channels from guinea-pig cardiomyocytes. J Physiol. 2001;532:115–126. [Abstract] [Google Scholar]
66. Lombard JH. Depression, psychological stress, vascular dysfunction, and cardiovascular disease: thinking outside the barrel. J Appl Physiol. 2010;108:1025–1026. [Abstract] [Google Scholar]
67. Longden TA, Dabertrand F, Hill-Eubanks DC, Hammack SE, Nelson MT. Stress-induced glucocorticoid signaling remodels neurovascular coupling through impairment of cerebrovascular inwardly rectifying K+ channel function. Proc Natl Acad Sci. 2014;111:7462–7467. [Europe PMC free article] [Abstract] [Google Scholar]
68. Longden TA, Dunn KM, Draheim HJ, Nelson MT, Weston AH, Edwards G. Intermediate-conductance calcium-activated potassium channels participate in neurovascular coupling. Br J Pharmacol. 2011;164:922–933. [Europe PMC free article] [Abstract] [Google Scholar]
69. Lopatin AN, Makhina EN, Nichols CG. Potassium channel block by cytoplasmic polyamines as the mechanism of intrinsic rectification. Nature. 1994;372:366–369. [Abstract] [Google Scholar]
70. Löffler K, Hunter M. Cation permeation and blockade of ROMK1, a cloned renal potassium channel. Pflügers Arch. 1997;434:151–8. [Abstract] [Google Scholar]
71. Mathiisen TM, Lehre KP, Danbolt NC, Ottersen OP. The perivascular astroglial sheath provides a complete covering of the brain microvessels: An electron microscopic 3D reconstruction. Glia. 2010;58:1094–1103. [Abstract] [Google Scholar]
72. Matsuda H. Open-state substructure of inwardly rectifying potassium channels revealed by magnesium block in guinea-pig heart cells. J Physiol. 1988;397:237–258. [Abstract] [Google Scholar]
73. Matsushita K, Puro DG. Topographical heterogeneity of KIR currents in pericyte-containing microvessels of the rat retina: effect of diabetes. J Physiol. 2006;573:483–495. [Abstract] [Google Scholar]
74. Mayevsky A, Weiss HR. Cerebral blood flow and oxygen consumption in cortical spreading depression. J Cereb Blood Flow Metab. 1991;11:829–836. [Abstract] [Google Scholar]
75. McCarron JG, Halpern W. Potassium dilates rat cerebral arteries by two independent mechanisms. Am J Physiol. 1990;259:H902–H908. [Abstract] [Google Scholar]
76. McCarron JG, Halpern W. Impaired potassium-induced dilation in hypertensive rat cerebral arteries does not reflect altered Na+/K+-ATPase dilation. Circ Res. 1990;67:1035–1039. [Abstract] [Google Scholar]
77. McCurley A, Pires PW, Bender SB, Aronovitz M, Zhao MJ, Metzger D, Chambon P, Hill MA, Dorrance AM, Mendelsohn ME, Jaffe IZ. Direct regulation of blood pressure by smooth muscle cell mineralocorticoid receptors. Nat Med. 2012;18:1429–1433. [Europe PMC free article] [Abstract] [Google Scholar]
78. McEwen BS. Mood disorders and allostatic load. Biol Psych. 2003;54:200–207. [Abstract] [Google Scholar]
79. Metea MR, Kofuji P, Newman EA. Neurovascular coupling is not mediated by potassium siphoning from glial cells. J Neurosci. 2007;27:2468–2471. [Europe PMC free article] [Abstract] [Google Scholar]
80. Meuth SG, Budde T, Kanyshkova T, Broicher T, Munsch T, Pape HC. Contribution of TWIK-related acid-sensitive K+ channel 1 (TASK1) and TASK3 channels to the control of activity modes in thalamocortical neurons. J Neurosci. 2003;23:6460–6469. [Europe PMC free article] [Abstract] [Google Scholar]
81. Miki T, Nagashima K, Seino S. The structure and function of the ATP-sensitive K+ channel in insulin-secreting pancreatic beta-cells. J Mol Endocrinol. 1999;22:113–123. [Abstract] [Google Scholar]
82. Mohr DC, Hart SL, Julian L, Cox D, Pelletier D. Association between stressful life events and exacerbation in multiple sclerosis: a meta-analysis. BMJ. 2004;328:731. [Europe PMC free article] [Abstract] [Google Scholar]
83. Nagelhus EA, Horio Y, Inanobe A, Fujita A, Haug FM, Nielsen S, Kurachi Y, Ottersen OP. Immunogold evidence suggests that coupling of K+ siphoning and water transport in rat retinal Müller cells is mediated by a coenrichment of Kir4.1 and AQP4 in specific membrane domains. Glia. 1999;26:47–54. [Abstract] [Google Scholar]
84. Nelson MT, Quayle JM. Physiological roles and properties of potassium channels in arterial smooth muscle. Am J Physiol Cell Physiol. 1995;268:C799–C822. [Abstract] [Google Scholar]
85. Nelson MT, Patlak JB, Worley JF, Standen NB. Calcium channels, potassium channels, and voltage dependence of arterial smooth muscle tone. Am J Physiol. 1990;259:C3–18. [Abstract] [Google Scholar]
86. Newman EA. High potassium conductance in astrocyte endfeet. Science. 1986;233:453–454. [Europe PMC free article] [Abstract] [Google Scholar]
87. Newman EA, Frambach DA, Odette LL. Control of extracellular potassium levels by retinal glial cell K+ siphoning. Science. 1984;225:1174–1175. [Europe PMC free article] [Abstract] [Google Scholar]
88. Nguyen Dinh Cat A, Griol-Charhbili V, Loufrani L, Labat C, Benjamin L, Farman N, Lacolley P, Henrion D, Jaisser F. The endothelial mineralocorticoid receptor regulates vasoconstrictor tone and blood pressure. FASEB J. 2010;24:2454–2463. [Abstract] [Google Scholar]
89. Nystoriak MA, O'Connor KP, Sonkusare SK, Brayden JE, Nelson MT, Wellman GC. Fundamental increase in pressure-dependent constriction of brain parenchymal arterioles from subarachnoid hemorrhage model rats due to membrane depolarization. Am J Physiol Heart Circ Physiol. 2011;300:H803–H812. [Europe PMC free article] [Abstract] [Google Scholar]
90. Orkand RK, Orkand PM, Tang CM. Membrane properties of neuroglia in the optic nerve of Necturus. J Exp Biol. 1981;95:49–59. [Abstract] [Google Scholar]
91. Otsu Y, Couchman K, Lyons DG, Collot M, Agarwal A, Mallet J, Pfrieger FW, Bergles DE, Charpak S. Calcium dynamics in astrocyte processes during neurovascular coupling. Nat Neurosci. 2014 10.1038/nn.3906. (Epub ahead of print) [Europe PMC free article] [Abstract] [Google Scholar]
92. Papanikolaou M, Virginia B, Lewis A, Butt AM. Immunohistochemical evidence for neuronal and glial expression of the inward rectifying potassium channel subtype Kir7.1 in the mouse brain. Proc Physiol Soc. 2012;27:PC69. [Google Scholar]
93. Paulson OB, Newman EA. Does the release of potassium from astrocyte endfeet regulate cerebral blood flow? Science. 1987;237:896–898. [Europe PMC free article] [Abstract] [Google Scholar]
94. Pearson WL, Dourado M, Schreiber M, Salkoff L, Nichols CG. Expression of a functional Kir4 family inward rectifier K+ channel from a gene cloned from mouse liver. J Physiol. 1999;514:639–653. [Abstract] [Google Scholar]
95. Petzold GC, Murthy VN. Role of astrocytes in neurovascular coupling. Neuron. 2011;71:782–797. [Abstract] [Google Scholar]
96. Plaster NM, Tawil R, Tristani-Firouzi M, Canún S, Bendahhou S, Tsunoda A, Donaldson MR, Iannaccone ST, Brunt E, Barohn R, Clark J, Deymeer F, George AL, Jr, Fish FA, Hahn A, Nitu A, Ozdemir C, Serdaroglu P, Subramony SH, Wolfe G, Fu YH, Ptácek LJ. Mutations in Kir2.1 Cause the Developmental and Episodic Electrical Phenotypes of Andersen's Syndrome. Cell. 2001;105:511–519. [Abstract] [Google Scholar]
97. Price DL, Ludwig JW, Mi H, Schwarz TL, Ellisman MH. Distribution of rSlo Ca2+− activated K+ channels in rat astrocyte perivascular endfeet. Brain Res. 2002;956:183–193. [Abstract] [Google Scholar]
98. Quayle JM, Dart C, Standen NB. The properties and distribution of inward rectifier potassium currents in pig coronary arterial smooth muscle. J Physiol. 1996;494:715–726. [Abstract] [Google Scholar]
99. Quayle JM, McCarron JG, Brayden JE, Nelson MT. Inward rectifier K+ currents in smooth muscle cells from rat resistance-sized cerebral arteries. Am J Physiol Cell Physiol. 1993;265:C1363–C1370. [Abstract] [Google Scholar]
100. Richardson S, Shaffer JA, Falzon L, Krupka D, Davidson KW, Edmondson D. Meta-analysis of perceived stress and its association with incident coronary heart disease. Am J Cardiol. 2012;110:1711–1716. [Europe PMC free article] [Abstract] [Google Scholar]
101. Robertson JL, Palmer LG, Roux B. Multi-ion distributions in the cytoplasmic domain of inward rectifier potassium channels. Biophys J. 2012;103:434–443. [Europe PMC free article] [Abstract] [Google Scholar]
102. Roman CW, Lezak KR, Kocho-Schellenberg M, Garret MA, Braas K, May V, Hammack SE. Excitotoxic lesions of the bed nucleus of the stria terminalis (BNST) attenuate the effects of repeated stress on weight gain: evidence for the recruitment of BNST activity by repeated, but not acute, stress. Behav Brain Res. 2012;227:300–304. [Europe PMC free article] [Abstract] [Google Scholar]
103. Roozendaal B, McEwen BS, Chattarji S. Stress, memory and the amygdala. Nat Rev Neurosci. 2009;10:423–433. [Abstract] [Google Scholar]
104. Rosenkranz JA, Venheim ER, Padival M. Chronic stress causes amygdala hyperexcitability in rodents. BPS. 2010;67:1128–1136. [Europe PMC free article] [Abstract] [Google Scholar]
105. Roy CS, Sherrington CS. On the regulation of the blood-supply of the brain. J Physiol. 1890;11:85–158. [Abstract] [Google Scholar]
106. Rubart M, Patlak JB, Nelson MT. Ca2+ currents in cerebral artery smooth muscle cells of rat at physiological Ca2+ concentrations. J Gen Physiol. 1996;107:459–472. [Europe PMC free article] [Abstract] [Google Scholar]
107. Schröder W, Seifert G, Hüttmann K, Hinterkeuser S, Steinhäuser C. AMPA Receptor-mediated modulation of inward rectifier K+ channels in astrocytes of mouse hippocampus. Mol Cell Neurosci. 2002;19:447–458. [Abstract] [Google Scholar]
108. Seifert G, Hüttmann K, Binder DK, Hartmann C, Wyczynski A, Neusch C, Steinhäuser C. Analysis of astroglial K+ channel expression in the developing hippocampus reveals a predominant role of the Kir4.1 subunit. J Neurosci. 2009;29:7474–7488. [Europe PMC free article] [Abstract] [Google Scholar]
109. Shin LM, Liberzon I. The neurocircuitry of fear, stress, and anxiety disorders. Neuropsychopharm. 2009;35:169–191. [Europe PMC free article] [Abstract] [Google Scholar]
110. Shuck ME, Piser TM, Bock JH, Slightom JL, Lee KS, Bienkowski MJ. Cloning and characterization of two K+ inward rectifier (Kir) 1.1 potassium channel homologs from human kidney (Kir1.2 and Kir1.3). J Biol Chem. 1997;272:586–593. [Abstract] [Google Scholar]
111. Smith PD, Brett SE, Luykenaar KD, Sandow SL, Marrelli SP, Vigmond EJ, Welsh DG. KIR channels function as electrical amplifiers in rat vascular smooth muscle. J Physiol. 2007;586:1147–1160. [Abstract] [Google Scholar]
112. Somjen GG. Mechanisms of spreading depression and hypoxic spreading depression-like depolarization. Physiol Rev. 2001;81:1065–1096. [Abstract] [Google Scholar]
113. Stanfield PR, Nakajima S, Nakajima Y. Constitutively active and G-protein coupled inward rectifier K+ channels: Kir2.0 and Kir3.0. Rev Physiol Biochem Pharmacol. 2002;92:47–179. [Abstract] [Google Scholar]
114. Stevens EB, Woodward R, Ho IH, Murrell-Lagnado R. Identification of regions that regulate the expression and activity of G protein-gated inward rectifier K+ channels in Xenopus oocytes. J Physiol. 1997;503:547–562. [Abstract] [Google Scholar]
115. Straub SV, Bonev AD, Wilkerson MK, Nelson MT. Dynamic inositol trisphosphate-mediated calcium signals within astrocytic endfeet underlie vasodilation of cerebral arterioles. J Gen Physiol. 2006;128:659–669. [Europe PMC free article] [Abstract] [Google Scholar]
116. Straub SV, Girouard H, Doetsch PE, Hannah RM, Wilkerson MK, Nelson MT. Regulation of intracerebral arteriolar tone by KV channels: effects of glucose and PKC. Am J Physiol: Cell Physiol. 2009;297:C788–C796. [Europe PMC free article] [Abstract] [Google Scholar]
117. Sun HS, Feng ZP, Miki T, Seino S, French RJ. Enhanced neuronal damage after ischemic insults in mice lacking Kir6.2-containing ATP-sensitive K+ channels. J Neurophysiol. 2005;95:2590–2601. [Abstract] [Google Scholar]
118. Takano M, Ashcroft FM. The Ba2+ block of the ATP-sensitive K+ current of mouse pancreatic beta-cells. Pflügers Arch. 1996;431:625–631. [Abstract] [Google Scholar]
119. Takano T, Tian GF, Peng W, Lou N, Libionka W, Han X, Nedergaard M. Astrocyte-mediated control of cerebral blood flow. Nat Neurosci. 2005;9:260–267. [Abstract] [Google Scholar]
120. Takumi T, Ishii T, Horio Y, Morishige K, Takahashi N, Yamada M, Yamashita T, Kiyama H, Sohmiya K, Nakanishi S. A novel ATP-dependent inward rectifier potassium channel expressed predominantly in glial cells. J Biol Chem. 1995;270:16339–16346. [Abstract] [Google Scholar]
121. Tanemoto M, Fujita A, Higashi K, Kurachi Y. PSD-95 mediates formation of a functional homomeric Kir5.1 channel in the brain. Neuron. 2002;34:387–397. [Abstract] [Google Scholar]
122. Thomzig A, Wenzel M, Karschin C, Eaton MJ, Skatchkov SN, Karschin A, Veh RW. Kir6.1 is the principal pore-forming subunit of astrocyte but not neuronal plasma membrane K-ATP channels. Mol Cell Neurosci. 2001;18:671–690. [Abstract] [Google Scholar]
123. Töpert C, Döring F, Wischmeyer E, Karschin C, Brockhaus J, Ballanyi K, Derst C, Karschin A. Kir2.4: a novel K+ inward rectifier channel associated with motoneurons of cranial nerve nuclei. J Neurosci. 1998;18:4096–4105. [Europe PMC free article] [Abstract] [Google Scholar]
124. Vetri F, Xu H, Paisansathan C, Pelligrino DA. Impairment of neurovascular coupling in type 1 diabetes mellitus in rats is linked to PKC modulation of BKCa and Kir channels. Am J Physiol Heart Circ Physiol. 2012;302:H1274–H1284. [Europe PMC free article] [Abstract] [Google Scholar]
125. Walker EF, Diforio D. Schizophrenia: a neural diathesis-stress model. Psychol Rev. 1997;104:667–685. [Abstract] [Google Scholar]
126. Wang HR, Wu M, Yu H, Long S, Stevens A, Engers DW, Sackin H, Daniels JS, Dawson ES, Hopkins CR, Lindsley CW, Li M, McManus OB. Selective inhibition of the Kir2 family of inward rectifier potassium channels by a small molecule probe: the discovery, SAR, and pharmacological characterization of ML133. ACS Chem Biol. 2011;6:845–856. [Europe PMC free article] [Abstract] [Google Scholar]
127. Watkins S, Robel S, Kimbrough IF, Robert SM, Ellis-Davies G, Sontheimer H. Disruption of astrocyte-vascular coupling and the blood-brain barrier by invading glioma cells. Nat Commun. 2014;5:4196. [Europe PMC free article] [Abstract] [Google Scholar]
128. Webb C, Bohr DF. Potassium-induced relaxation as an indicator of Na+-K+ ATPase activity in vascular smooth muscle. J Vasc Res. 1978;15:198–207. [Abstract] [Google Scholar]
129. Weston AH, Richards G, Burnham M, Feletou M, Vanhoutte P, Edwards G. K+-induced hyperpolarization in rat mesenteric artery: identification, localization and role of Na+/K+-ATPases. Br J Pharmacol. 2002;136:918–926. [Europe PMC free article] [Abstract] [Google Scholar]
130. Weston A, Porter E, Harno E, Edwards G. Impairment of endothelial SKCa channels and of downstream hyperpolarizing pathways in mesenteric arteries from spontaneously hypertensive rats. Br J Pharmacol. 2010;160:836–843. [Europe PMC free article] [Abstract] [Google Scholar]
131. Witthoft A, Karniadakis GE. A bidirectional model for communication in the neurovascular unit. J Theor Biol. 2012;311:80–93. [Abstract] [Google Scholar]
132. Witthoft A, Filosa JA, Karniadakis GE. Potassium buffering in the neurovascular unit: models and sensitivity analysis. Biophys J. 2013;105:2046–2054. [Europe PMC free article] [Abstract] [Google Scholar]
133. Wu BN, Luykenaar KD, Brayden JE, Giles WR, Corteling RL, Wiehler WB, Welsh DG. Hyposmotic challenge inhibits inward rectifying K+ channels in cerebral arterial smooth muscle cells. Am J Physiol Heart Circ Physiol. 2006;292:H1085–H1094. [Abstract] [Google Scholar]
134. Zaritsky JJ, Eckman DM, Wellman GC, Nelson MT, Schwarz TL. Targeted disruption of Kir2.1 and Kir2.2 genes reveals the essential role of the inwardly rectifying K+ current in K+-mediated vasodilation. Circ Res. 2000;87:160–166. [Abstract] [Google Scholar]

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1111/micc.12190

Supporting
Mentioning
Contrasting
3
144
1

Article citations


Go to all (77) article citations

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


Funding 


Funders who supported this work.

American Heart Association (2)

Fondation Leducq

    NCRR NIH HHS (2)

    NHLBI NIH HHS (6)

    NIDDK NIH HHS (2)

    National Institutes of Health (5)

    Totman Medical Research Trust