INVITED REVIEW
Ionic mechanisms and Ca2+ regulation in airway
smooth muscle contraction: do the data contradict dogma?
Luke J.
Janssen
Asthma Research Group, Firestone Institute for Respiratory
Health, St. Joseph's Hospital; and Department of Medicine,
McMaster University, Hamilton, Ontario, Canada L8N 4A6
 |
ABSTRACT |
In general,
excitation-contraction coupling in muscle is dependent on
membrane depolarization and hyperpolarization to regulate the opening
of voltage-dependent Ca2+ channels and, thereby, influence
intracellular Ca2+ concentration
([Ca2+]i). Thus Ca2+ channel
blockers and K+ channel openers are important tools in the
arsenals against hypertension, stroke, and myocardial infarction, etc.
Airway smooth muscle (ASM) also exhibits robust Ca2+,
K+, and Cl
currents, and there are elaborate
signaling pathways that regulate them. It is easy, then, to presume
that these also play a central role in contraction/relaxation of ASM.
However, several lines of evidence speak to the contrary. Also, too
many researchers in the ASM field view the sarcoplasmic reticulum as
being centrally located and displacing its contents uniformly
throughout the cell, and they have focused almost exclusively on the
initial single [Ca2+] spike evoked by excitatory
agonists. Several recent studies have revealed complex spatial and
temporal heterogeneity in [Ca2+]i, the
significance of which is only just beginning to be appreciated. In this
review, we will compare what is known about ion channels in ASM
with what is believed to be their roles in ASM physiology. Also, we
will examine some novel ionic mechanisms in the context of
Ca2+ handling and excitation-contraction coupling in ASM.
excitation-contraction coupling; ion channels; membrane
potential
 |
INTRODUCTION |
AIRWAY HYPERREACTIVITY and variable
airflow obstruction are key features of asthma. Indeed, one might say
these are its most clinically relevant features. For this reason, it is
essential to have a good understanding of the mechanisms
underlying excitation-contraction (EC) coupling in airway smooth muscle
(ASM). A great deal of research is being focused on the
electrophysiology of ASM, given the importance of ion channels in
EC coupling in other muscle types. The goal of this review is to
stimulate a reevaluation of the existing literature on the
electrophysiology of ASM as it pertains to EC coupling, with a view to
redirect those research efforts.
Contraction in smooth muscle is a product of the interaction between
actin and myosin, as described by the classic sliding filament theory.
The degree of this interaction is determined by the net level of
phosphorylation of the 20-kDa myosin light chain, which, in turn,
is dependent on the relative activities of myosin light chain kinase
(MLCK) and myosin light chain phosphatase (MLCP). MLCK is activated by
Ca2+/calmodulin. Thus, in general, bronchoconstrictors act
by elevating intracellular Ca2+ concentration
([Ca2+]i) to increase MLCK activity and/or by
decreasing MLCP activity (which effectively increases the
Ca2+ sensitivity of the contractile apparatus).
Bronchodilators, on the other hand, generally produce the opposite
effects. A more complete description of the interactions between
actin, myosin, MLCK, and MLCP is beyond the scope of this review but
can be found elsewhere (87, 209-212).
 |
TRADITIONAL VIEWS |
EC coupling in nonairway muscles: adequate models for ASM?
In striated muscles as well as vascular and gastrointestinal smooth
muscle, EC coupling is largely dependent on membrane depolarization, although for very different reasons.
In cardiac muscle, Na+ channel opening depolarizes the
membrane, resulting in Ca2+ entry via voltage-dependent
("L-type") Ca2+ channels, producing a transient
elevation of [Ca2+]i immediately under the
plasmalemma (17). This initial rise in
[Ca2+]i does not trigger contraction
directly: instead, it activates ryanodine receptors on the sarcoplasmic
reticulum (SR), causing a massive discharge of Ca2+ from
the internal store, resulting in contraction (17).
Voltage-dependent Ca2+ influx also contributes to a number
of other cellular events, including refilling of the SR, activation of
plasmalemmal ion channels, and modulation of various enzyme activities, etc.
In vascular and gastrointestinal smooth muscle, on the other hand,
voltage-dependent Ca2+ influx through L-type channels is
sufficient for contraction (17). Moreover, many
electrophysiological studies of these tissues reveal a "window
current" spanning the physiologically relevant range of membrane
potentials, i.e., a range of potentials more positive than the
threshold for activation of the Ca2+ channels but over
which voltage-dependent inactivation is not complete, giving rise to a
persistent Ca2+ influx. Thus small hyperpolarizations lead
to decreased activation of the channels and a drop in
[Ca2+]i, while small depolarizations increase
Ca2+ channel activation and elevate
[Ca2+]i.
Despite the differences in the mechanisms underlying EC coupling in
these tissues, the central role played by dihydropyridine-sensitive Ca2+ channels in both cases provides the rationale for the
use of Ca2+ channel blockers and K+ channel
agonists in controlling cardiac and smooth muscle contractions in
hypertension, stroke, myocardial infarction, and gastrointestinal motility disorders, etc. More importantly, clinical studies attest to
the efficacy of these tools for these purposes (59, 66, 187).
Agonist-mediated EC coupling in ASM.
Excitation of ASM is similar in many respects to that of vascular or
gastrointestinal smooth muscles. First, it is accompanied by membrane
depolarization (60) mediated primarily by activation of
Cl
and nonselective cation currents as well as
suppression of K+ currents (106, 117, 119, 120,
236). Patch-clamp studies have documented the large
voltage-dependent Ca2+ currents activated by membrane
depolarization (68, 140) (Fig. 1). Finally, these
Ca2+ currents are sufficient to produce contraction, as
manifest in the robust dihydropyridine-sensitive contractions evoked by
potassium chloride (108, 122) or K+ channel
blockers such as tetraethylammonium (TEA), 4-aminopyridine, or
charybdotoxin (45), although these contractions are
generally only a fraction of the size of those evoked by physiological
agonists such as carbachol, histamine, and endothelin, etc.

View larger version (17K):
[in this window]
[in a new window]
|
Fig. 1.
Acetylcholine-evoked depolarizations and Ca2+
currents. Current-voltage relationships in canine tracheal smooth
muscle cells from 2 different laboratories [from Kotlikoff
(140) ( ) and Muraki et al.
(164) ( )] were converted to
activation-voltage relationships by assuming that Ca2+
currents were maximally activated at +30 mV. Bars (left)
indicate membrane potentials recorded in canine trachealis during
stimulation with a range of concentrations of acetylcholine (Ach)
[from Farley and Miles (60)]. Even with maximally
effective concentrations of acetylcholine, membrane potential barely
enters the range at which substantial Ca2+ current is
activated. Similar results have been obtained for every mammalian
species studied (see Fig. 4). VR, resting membrane
potential; I/Imax, fraction of
maximal current.
|
|
Given these similarities with other muscle types, it is understandable
that many treat Ca2+, K+, and Cl
channels as central players in EC coupling in ASM. However, a large
body of data speaks to the contrary.
 |
CHALLENGING THE DOGMA |
Do agonist-evoked contractions in ASM require voltage-dependent
Ca2+ influx?
Many groups have characterized the voltage-dependent Ca2+
currents in ASM (usually of the trachealis) and found these to be almost exclusively L-type in nature (77, 81, 84, 107, 140, 152,
164). It is particularly important to bear in mind a number of
biophysical properties of these currents. 1) All
electrophysiological studies find the threshold potential for these
currents to be in excess of
40 mV, and almost all of them show little
or no Ca2+ current until membrane voltages rise above
20
mV (Fig. 1); peak activation occurs at +10 to +20 mV. 2)
These currents can develop substantial voltage- and
Ca2+-dependent inactivation, and they are also suppressed
by various second messenger signaling pathways (233, 242,
246). 3) L-type channels are selectively and potently
blocked by dihydropyridines (Fig. 2),
with an IC50 in the nanomolar range (158).

View larger version (16K):
[in this window]
[in a new window]
|
Fig. 2.
Contractions in airway smooth muscle (ASM) are seemingly
unaffected by blockers of voltage-dependent Ca2+ channels
or of Cl channels. Despite the ability of niflumic acid
and nifedipine to abolish Cl and Ca2+
currents in ASM (A and B, respectively) [adapted
from Janssen and Sims (117) and Janssen
(107)], these agents have no substantial effect on
agonist-evoked contractions (C).
|
|
Given these properties of the Ca2+ channels, several
observations call into question the view that voltage-dependent
Ca2+ channels play a central role in ASM contraction.
First, many have shown that agonist-evoked contraction of ASM is
seemingly unaffected under conditions in which voltage-dependent Ca2+ influx is prevented by Ca2+ channel
blockers (25, 26, 48, 61) (Fig. 2) by "clamping" the
membrane potential to very negative values far below the threshold for
Ca2+ channel activation (118) (Fig.
3) or even by removal of external Ca2+ (48, 61). Likewise, although
Cl
channels are primarily responsible for the membrane
depolarization (106, 117, 119, 120, 147, 148, 237, 238),
agents such as niflumic acid, which are able to completely block the
Cl
currents in ASM, have essentially no effect on resting
membrane potential (113) or agonist-evoked contractions
(Fig. 2). The only studies that do describe an inhibitory effect of
dihydropyridines on mechanical responses in ASM were carried out under
very nonphysiological conditions (complete depletion of the internal
Ca2+ pool) (8, 24-26, 114, 123, 197, 224)
or used supramaximally effective concentrations of dihydropyridines.
For example, in the case of nifedipine, most groups tend to use
10
6 M, and may even use 10
5 M
(231), even though the IC50 value reported for
this agent is in the nanomolar range (158); submicromolar
concentrations are sufficient to completely block the Ca2+
channels in ASM (77, 84, 243) and to suppress contractions in vascular or gastrointestinal smooth muscle (13, 38, 57, 70,
145). Thus nonspecific effects of the dihydropyridines (191) need to be kept in mind when interpreting such data.

View larger version (10K):
[in this window]
[in a new window]
|
Fig. 3.
Substantial contraction can still be evoked at very negative
membrane potentials. Vertical axis indicates cell length in an ASM cell
held continuously under voltage clamp at 60 mV. Acetylcholine
(10 4 M) was applied at ~3-min intervals
( ). Numerous contractions (to almost 50% of resting
length) can still be evoked at this potential, which is far below
threshold for opening of voltage-dependent Ca2+ channels.
[From Janssen and Sims (118).]
|
|
Second, the range of membrane potentials typically seen in ASM at rest
and during excitation ranges from
70 to
30 mV (1, 27, 40, 46,
60, 85, 95, 96, 99-102, 109, 111, 112, 125, 130, 132, 136, 155,
188, 215, 220), which is well below the range of potentials
required for Ca2+ channel activation (
30 to +20 mV)
(68, 81, 84, 107, 140, 152, 164, 242, 243, 247) (Figs. 1
and 4). More to the point, the voltages
required to only marginally activate voltage-dependent Ca2+
channels (
40 to
30 mV) are attained only with concentrations of
agonist that evoke nearly complete contraction, and Ca2+
currents are maximal at membrane potentials never seen during agonist
stimulation (+10 to +20 mV; see Figs. 1 and 4). Simultaneous electrophysiological and fura 2 fluorimetric recordings in equine ASM
have shown that the Ca2+ currents evoked by voltage step
commands to potentials in the physiologically relevant range produce
elevations in [Ca2+]i of <50 nM
(68), which pales in comparison with
bronchoconstrictor-evoked Ca2+ responses (both the peak and
plateau values typically measure several hundreds of nanomolars)
(249-251).

View larger version (34K):
[in this window]
[in a new window]
|
Fig. 4.
Membrane voltages and Ca2+ currents.
Estimates of membrane potentials at rest ( ) and during
stimulation with various excitatory agents ( ), obtained
using intracellular microelectrode techniques. Numerals in parentheses
indicate reference citation from which data were extracted. Cch,
carbachol; EFS, electric field stimulation; 4-AP, 4-aminopyridine; TEA,
tetraethylammonium; 5-HT, 5-hydroxytryptamine. Also shown are voltages
for threshold and peak ( and ,
respectively) activation of voltage-dependent Ca2+
currents, as determined using patch-clamp electrophysiological
techniques. The voltages at which those currents are half inactivated
are also indicated by the solid vertical bar. Shaded box indicates the
range of potentials at which membrane oscillations (slow waves) are
typically seen in ASM. The purpose of this figure is to highlight the
lack of overlap between the physiologically relevant range of
potentials and those potentials required for significant
voltage-dependent Ca2+ influx. TSM, tracheal smooth
muscle; BSM, bronchial smooth muscle; LTC, leukotriene
C4.
|
|
Third, the Ca2+ currents and changes in
[Ca2+]i described above are overestimations
of those one should expect to see in vivo. Because they were evoked
under experimental conditions in which voltage-dependent inactivation
had not developed to any substantial degree, i.e., the depolarizing
steps used to evoke these currents are generally less than a few
hundred milliseconds in duration and are each separated by several
seconds to minimize inactivation. This inactivation is half-maximal
(meaning the currents are halved in size) whenever membrane potential
approaches
30 mV for more than a few seconds (77, 84, 107, 140,
164) and is nearly complete as membrane potential approaches 0 mV. Thus during slow wave activity, when the membrane potential is
slowly oscillating between
40 and
20 mV for many minutes or even
hours, substantial inactivation of these currents will have occurred.
One group has described a small persistent Ca2+ influx at
the lower limit of this voltage range (referred to as window current)
that can manifest as a detectable elevation of
[Ca2+]i (<100 nM) (67), but the
latter is again too small to account for the robust contractions evoked
by agonists.
Fourth, whenever the tissues are stimulated by bronchoconstrictors such
as cholinergic agonists, voltage-dependent Ca2+ currents
are further suppressed via phosphorylation of the channels by protein
kinase C (246) and/or through Ca2+-induced
inactivation of the Ca2+ channels (233). Thus
the small window current that might exist would be wiped out by many
bronchoconstrictors. Surprisingly,
-agonists have been shown to
augment L-type Ca2+ currents in ASM (242) even
though they are powerful relaxants.
Finally, and perhaps most importantly, clinical studies have found
Ca2+ channel blockers to be ineffective as therapeutic
agents in asthma (14, 65, 76, 86, 162, 189, 202).
This begs the question: Why does ASM exhibit such large
Ca2+ and Cl
currents? We will propose several
answers to this question in this review.
Do K+ channels play a major role in
agonist-evoked relaxations?
In general, K+ channels are divided into four major
classes: Ca2+ dependent (KCa), voltage
dependent (KV), ATP dependent (KATP), and
inward rectifier (KIR) (142, 169, 187), of
which there is substantial direct evidence in ASM for two of these
classes and very limited evidence for the other two classes.
McCann and Welsh (157) were the first to directly record
KCa in ASM, and there have since been innumerable studies
that add to this evidence. These channels are generally of the large
conductance subtype because they are highly sensitive to K+
channel blockers such as TEA, charybdotoxin, and iberiotoxin, with very
little effect of the small conductance blocker apamin (28, 141,
165, 207, 208). Also, direct patch-clamp recordings show these
channels to have unitary conductances of several hundred picoamperes
(207, 208, 216). One of these studies (216)
further found that these channels can undergo random conformational
changes that lead to subconductance states of 17, 33, 41, 52, 63, and 72% of the full conductance. These currents can appear to be very "noisy" with chaotic oscillations and spikes, in part due to marked spatial/temporal changes in [Ca2+]i and the
large unitary conductance of these channels.
KV currents in ASM have also been studied in detail at the
whole cell (69, 141) and single channel (28)
levels. Activation of these channels ensues after somewhat of a delay
(28) (thus they are also referred to as "delayed
rectifier" currents) and occurs much more "smoothly" than
KCa given their smaller unitary conductance (10-15 pS)
and insensitivity to [Ca2+]i relative to
KCa. The channels also exhibit voltage-dependent inactivation, which is roughly half-maximal at the resting membrane potential. KV channels are effectively blocked by
4-aminopyridine (1-5 mM) or dendrotoxin (1-100 nM) but
not by TEA, charybdotoxin, or glybenclamide unless unreasonably high
concentrations of these agents are used (169).
Although there are many studies providing indirect evidence for
KATP in ASM (in that relaxations are evoked by
KATP agonists such as cromakalim) and these mechanical
responses are antagonized by KATP blockers such as
glybenclamide (21, 22, 36, 42, 43, 97, 131, 165, 178),
direct electrophysiological evidence for these channels is essentially
nonexistent. Many groups that have characterized K+
currents in ASM in detail directly using patch-clamp techniques have
not reported a glybenclamide-sensitive component. Rather than an action
on some channel per se, some evidence suggests that KATP
agonists act instead by suppressing phosphodiesterase activity
(178, 203).
Dozens of studies of ASM cells from the larger airways have failed to
identify any inward rectifier K+ currents. However, one
recent study (208) of cells obtained from small human
bronchioles (outer diameter 0.3-1.0 mm) has done so. The
physiological relevance of this possible regional heterogeneity is unclear.
General dogma has it that relaxants act by opening K+
channels and hyperpolarizing the membrane. However, again, several
lines of evidence speak to the contrary.
For example, bronchodilators such as
-agonists and nitric oxide can
still evoke substantial or even complete relaxation in the presence of
K+ channel blockers (6, 7, 45, 102, 116, 127, 128,
165, 219) (Fig. 5). Although a
rightward shift in the concentration-response relationship for the
bronchodilator is sometimes seen (127, 128, 163), this
should be interpreted carefully. Such a parallel shift is a hallmark of
competitive inhibition, yet the K+ channel blocker and
bronchodilator agonist are not competing at a common receptor, and one
should not expect that stimulating the receptor more aggressively (by
using higher concentrations of agonist) would displace the blocker from
the channel and thus unmask the relaxation. Instead, it may be that the
K+ channel blockers are depolarizing nerve endings in the
tissues, causing them to release excitatory agonists (97,
132) that then antagonize the bronchodilator response. Such
functional antagonism can be overcome by using higher concentrations of
bronchodilator agonist (64, 110, 185). Tetrodotoxin is not
a guarantee against this because it only prevents depolarization caused
by Na+ channel activation but not that caused by
suppression of outward K+ currents in the nerve endings.
Instead, agents such as
-conotoxin should be used to prevent the
subsequent Ca2+ influx and neurotransmitter release.

View larger version (17K):
[in this window]
[in a new window]
|
Fig. 5.
Substantial relaxation can still be evoked during blockade of
K+ channels. Relaxations were evoked by the NO donor
soluble N-ethylmaleimide-sensitive factor attachment protein
(SNAP) in ASM tissues pretreated with the K+ channel
blockers charybdotoxin (ChTx; 0.1 µM), TEA (30 mM), or 4-AP (1 mM).
These data suggest that K+ channels are not absolutely
necessary for NO-evoked relaxation. [From Janssen et al.
(116).]
|
|
Also, bronchodilators can often cause relaxations when
voltage-dependent Ca2+ influx has already been abolished
beforehand using Ca2+ channel blockers (22, 102, 130,
218). Others have found that relaxations are not evoked by
artificially imposed hyperpolarizing current (39, 48);
again, bronchoconstrictors can still evoke substantial contraction
during voltage clamp at resting membrane potentials (117,
118) (Fig. 3).
Clearly, then, membrane hyperpolarization alone is neither necessary
nor sufficient for relaxation in ASM. Consistent with this,
K+ channel openers have been found to be ineffective as a
therapy for asthma (41, 62, 135, 204). Thus a better
understanding of agonist-evoked relaxation demands a new emphasis on
mechanisms other than K+ channel activation. Many (perhaps
all) bronchodilators that activate K+ channels are also
known to exert other effects on ASM, including decreased
Ca2+ sensitivity of the contractile apparatus (126,
174, 209, 210, 212), inhibition of
D-myo-inositol 1,4,5-trisphosphate
(IP3) binding to its receptor on the SR (196),
suppression of IP3-induced Ca2+ release
(133, 241), and enhancement of Ca2+
uptake/extrusion (133).
EC coupling in ASM owes much more to voltage-independent
mechanisms.
In contrast to the questionable significance of membrane
voltage-regulated Ca2+ influx, a number of
voltage-independent mechanisms are primarily responsible for
contraction in ASM.
The most widely recognized involves the release of Ca2+
sequestered within the SR. Cholinergic agonists, histamine, endothelin, leukotrienes, and thromboxane A2 activate phospholipase C,
which, in turn, generates the second messengers diacylglycerol and
IP3 (32-34). The latter of these two
messengers activates Ca2+-permeable ion channels on the
membrane of the SR, releasing its store of Ca2+ and,
thereby, triggering contraction (212). The SR also
expresses another group of Ca2+-permeable ion channels that
are activated by Ca2+ itself, caffeine, ryanodine, or by
cyclic ADP ribose (71, 201, 209, 212). The physiological
role of these "ryanodine receptors" is still debated (see
Bronchodilators: is K+
channel activation a causal event or an epiphenomenon?).
Refilling of the SR involves primarily the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA). However, the existence of at least one
other novel refilling pathway has been suggested (25, 26, 118, 186) as summarized in Voltage-dependent
Ca2+ channels and SR
refilling.
More recently, a great deal of attention has been focused on
agonist-induced changes in the sensitivity of the contractile apparatus
to [Ca2+]. One such mechanism that is gaining a great
deal of momentum both in the vascular and ASM fields involves
activation of the monomeric G protein Rho, which in turn translocates
to the membrane and activates Rho kinase (31, 98, 124,
253). The latter phosphorylates and thereby inactivates MLCP,
leading to a net accumulation of phosphorylated myosin light chains,
and thus contraction (211). Others are investigating
mechanisms such as extracellular regulated kinase
(ERK)-mediated phosphorylation of caldesmon and calponin
(72, 75), integrin-mediated tyrosine phosphorylation of
focal adhesion kinase, paxillin, and talin (73, 159, 160, 177,
221, 223, 240), and protein kinase C activation (2, 29,
93). Endothelin-stimulated activation of ERK was recently shown
to be dependent on Ca2+ influx (232),
indicating some "cross talk" between Ca2+-dependent and
Ca2+-independent pathways.
The superficial buffer barrier and spatial/temporal heterogeneity
of [Ca2+] in ASM.
Several recent findings have revolutionized the way Ca2+
handling in ASM should be considered.
First, until recently, the internal Ca2+ pool has usually
been modeled as being more or less centrally located and releasing its
store of Ca2+ roughly uniformly throughout the entire
cytosol. However, we had long been puzzled by the observation that both
acetylcholine and caffeine evoke substantial elevations of
[Ca2+]i as well as membrane currents, but
only the former would reliably evoke contraction (74, 108, 117,
119). That is, caffeine only evokes a contraction when it causes
a massive discharge of Ca2+ (e.g., when applied
instantaneously in high millimolar concentrations), but not when this
discharge occurs gradually (e.g., using lower concentrations and/or
introducing caffeine more slowly via the bath perfusion).
Electronmicroscopy of vascular smooth muscle cells shows the SR to form
sheets around the internal periphery of the cell (170),
thereby dividing the cytosol into two spaces (Fig.
6), the peripheral space immediately
underneath the plasmalemma where ion channels are found (many of them
being regulated by Ca2+), and the deep cytosolic space
where the contractile apparatus is found. Another recent study suggests
that the same anatomical arrangement can be found in ASM
(49). In this way, the cell can dissociate the influence
of [Ca2+] on mechanical and electrical activities. This
"superficial buffer barrier" accounts for the paradoxical effects
of caffeine (108, 124): ryanodine receptors may direct
Ca2+ release preferentially into the peripheral space (and,
thereby, activate ion channels), and the contractile apparatus only
becomes activated when this release is so massive that it "spills
over" into the deep space. It can also explain how subnanomolar
concentrations of acetylcholine can evoke substantial ionic current but
without any change in tension (124). That is, the ionic
currents indicate that Ca2+ is in fact being released
(because they are totally dependent on that process) (108, 117,
118, 121, 143, 238) although this release cannot be discerned
using fluorimetric techniques that measure the average change in
[Ca2+]i throughout the entire cell
(124, 198, 200, 249). However, the anticipated mechanical
response is quelled by the barrier function of the SR. Thus it is
essential to use more refined Ca2+ imaging techniques, ones
that can resolve subcellular regions, for further studies of
Ca2+ handling in ASM.

View larger version (23K):
[in this window]
[in a new window]
|
Fig. 6.
Superficial buffer barrier model. The sarcoplasmic
reticulum (SR) divides the cytosol into 2 spatially and functionally
distinct compartments: the subplasmalemmal space and the deep cytosolic
space. Bottom left: excitatory agonists release
Ca2+ into the deep cytosol to trigger contraction, as well
as into the subplasmalemmal space to activate
Ca2+-dependent Cl channels. Bottom
right: relaxants, on the other hand, promote Ca2+
uptake and Ca2+ extrusion but also trigger Ca2+
release via ryanodine receptors. The latter effect allows for unloading
of the SR, increasing its buffering capacity, without triggering
contraction. It may also lead to activation of
Ca2+-dependent K+ channels and membrane
hyperpolarization.
|
|
Another change in our understanding of Ca2+ handling was
stimulated by advances made in other cell types using highly
sophisticated Ca2+ imaging equipment. In the ASM field,
many focus their attention on the magnitude of the solitary, brief,
spikelike Ca2+ transient evoked by a high concentration of
agonist and relate that to much more persistent cellular events, such
as contraction. In other words, this Ca2+ transient is
being interpreted as a persistent elevation of
[Ca2+]i throughout the entire cell. Instead,
the sustained changes in [Ca2+]i evoked by
lower concentrations of agonist might be more physiologically relevant.
It is now known that agonists can evoke recurring changes in
[Ca2+]i ("oscillations"), which propagate
throughout the cell ("waves") (225), and are further
refined or shaped as they progress through the cytosol
(139). These oscillations convey information within their
amplitudes (peak height as well as their mean or temporally averaged
amplitude) as well as in their frequency, and this information may be
decoded by Ca2+/calmodulin-dependent kinase (19, 20,
30, 35, 50, 52, 55, 184), MLCK (50), the SR
Ca2+ pump (50), calpain (229),
adenylyl cyclase (44), or mitochondria (79).
For example, gene expression of several proinflammatory cytokines is
differentially regulated by Ca2+ oscillations in a
frequency- and amplitude-dependent fashion (54, 146).
Ca2+ oscillations have been reported in ASM from the human
(56), rat (193, 194, 227), guinea pig
(117, 200), pig (171, 180-183), and dog
(117), but their underlying mechanism and their
physiological relevance are still poorly understood.
Finally, a great deal of work has been done in the cardiac and vascular
smooth muscle fields looking at Ca2+ sparks: small and
transient elevations of [Ca2+]i produced by
localized bursts of Ca2+ from a handful of ryanodine
receptors (23, 82, 103, 104, 168). These are proposed to
be the triggers for spontaneous transient ion currents, which, in turn,
are believed to modulate mechanical activity (168, 254).
They also represent the fundamental event underlying Ca2+
oscillations and relaxations in vascular smooth muscle
(168). There have been limited studies of sparks in ASM
(172, 173, 199, 254), but their physiological role(s) is
unclear. Although they may lead to activation of K+ and
Cl
channels (254), electromechanical
coupling is of limited importance in ASM. Instead, their primary
function may be to discharge the SR contents, without evoking
contraction, toward the plasmalemmal Ca2+ pump to increase
the buffering capacity of the SR.
Thus Ca2+ signals are organized into complex temporal and
spatial patterns that are lost using techniques and models that focus solely on the spikelike elevation averaged across the entire ASM cell.
Many previous reports of Ca2+ responses in ASM were
severely limited, because whole cell photometry was used: this approach
averages the changes in [Ca2+]i across the
entire cell. None have yet compared bronchoconstrictor-induced changes
in the deep cytosol vs. the subplasmalemmal space. Also, the sampling
rate of most of these studies (on the order of 1 Hz) was much too slow
to adequately resolve events with time courses on the millisecond
scale, such as Ca2+ oscillations and Ca2+
sparks (199). Finally, the few studies able to resolve
subcellular regions generally used only a maximally effective
concentration of excitatory agonist: we have previously obtained
evidence (using patch-clamp recordings) that concentrations of
acetylcholine that were subthreshold for mechanical or fluorimetric
responses nonetheless evoked substantial membrane currents
(124), suggesting important differences between global and
subplasmalemmal measurements of [Ca2+]. The concentration
of agonist used is also important with respect to the likelihood of
observing Ca2+ oscillations, since models of these
phenomena indicate a critical dependence on variables such as
[IP3] and basal [Ca2+] (199,
225).
 |
NOVEL IONIC MECHANISMS |
Voltage-dependent Ca2+ channels and
SR refilling.
Several studies show voltage-dependent Ca2+ channels in ASM
to be important for refilling and maintenance of the SR (25, 26, 118, 149, 186). For example, we used agonist-evoked
Cl
currents to assess the filling state of the SR, and we
found that we could completely deplete the SR using cyclopiazonic acid (SERCA inhibitor) and then refill the SR in a dihydropyridine-sensitive fashion using a series of depolarizing pulses (Fig.
7) (118). More surprisingly,
though, this refilling occurred in the maintained presence of
cyclopiazonic acid (which was completely sufficient to functionally
deplete the SR), suggesting that this refilling pathway did not involve
SERCA. In other words, Ca2+ was crossing the plasmalemma
and entering the SR without being pumped by the
Ca2+-ATPase.

View larger version (24K):
[in this window]
[in a new window]
|
Fig. 7.
Ca2+ currents may contribute to refilling of
the SR. Acetylcholine-evoked Cl currents
( ) were used to assess the filling state of the SR in
canine tracheal smooth muscle. Exposure to cyclopiazonic acid (CPA)
completely depleted the SR (ii). In the maintained presence
of cyclopiazonic acid, a series of depolarizing pulses led to partial
refilling of the SR (iii). This refilling was enhanced by
the Ca2+ channel agonist BayK8644 and inhibited by
nifedipine (iv). [From Janssen and Sims
(118).]
|
|
These paradoxical findings can be explained using a recently proposed
model that describes a physical interaction between the SR and
plasmalemmal store-operated Ca2+ channels (18, 138,
176, 190, 191, 252). Briefly, agonist-induced depletion of the
internal store triggers activation of tyrosine kinase(s), Ras, and
reorganization of the cytoskeleton in such a way that it directly
couples IP3 receptors on the SR with Ca2+
channels on the plasmalemma. Similarly, a functional interaction between voltage-dependent Ca2+ channels and the nucleus has
recently been described. Dolmetsch et al. (53) found that
gene transcription could be stimulated by Ca2+ influx
through L-type Ca2+ channels but not through N- or P/Q-type
Ca2+ channels (which are also present but seem to
contribute to other physiological responses in these cells). Their data
show that calmodulin, which is tethered at the mouth of the L-type
Ca2+ channels and is activated by Ca2+ influx
through them, signals to the nucleus via a mitogen-activated protein
kinase pathway (53).
Several observations made in ASM are consistent with a physical
coupling between plasmalemmal Ca2+ channels and SR
membranes. First, excitatory stimulation of ASM is accompanied by
activation of tyrosine kinases (206, 228) and Ras/Rho
(58, 80, 83, 226) as well as cytoskeletal rearrangement (78, 83, 226). Second, inhibition of tyrosine kinases
compromises SR refilling (151). Third, ASM depleted of
focal adhesion kinase (which regulates cytoskeleton stability) shows
marked suppression of acetylcholine-evoked Ca2+ transients
and contractions as well as changes in voltage-dependent Ca2+ channel function without any disruptive changes in the
contractile apparatus per se (assessed by addition of Ca2+
to permeabilized strips) (222). Fourth, our observation
that the SR can be refilled by voltage-dependent Ca2+
influx in the maintained presence of cyclopiazonic acid is difficult to
explain otherwise, because there is no other Ca2+ pump on
the SR other than SERCA.
At first glance, this novel model of SR refilling also suffers from the
criticism raised earlier in this review that membrane potential rarely
reaches the threshold for opening of the voltage-dependent Ca2+ channels. However, if there is a very close apposition
or even physical coupling between the plasmalemmal and SR membranes, it is unclear what transmembrane potentials the plasmalemmal
Ca2+ channels would experience. That is, the close
proximity and physical interaction of a large polypeptide (the
IP3 receptor) and the SR membrane with the inner face of
the Ca2+ channel, with the resultant changes in membrane
surface charge and/or induction of conformational changes in the
Ca2+ channel, could easily alter their voltage
characteristics, allowing them to open at physiologically relevant
potentials (167). Also, access of Ca2+ and of
protein kinase C to the inner face of the Ca2+ channels
might be hindered during this interaction, thereby preventing Ca2+ channel inactivation.
Bronchodilators: is K+ channel
activation a causal event or an epiphenomenon?
The traditional view has been that bronchodilators act by decreasing
[Ca2+]i throughout the cell. Surprisingly,
however, we and others have described elevations in
[Ca2+]i in response to relaxant agents
(63, 115, 116, 245). Yamaguchi et al. (245)
resolved the effects of isoproterenol on [Ca2+] in
greater detail and showed that it increases [Ca2+] in the
peripheral regions of the ASM cells and decreases it in their more
central regions. One interpretation of these findings is that
bronchodilators act by triggering Ca2+ release from the SR,
completely contrary to current dogma, but not in the same fashion as
bronchoconstrictors. That is, rather than elevating
[Ca2+]i globally throughout the cell (via
IP3-induced Ca2+ release), the data suggest
that ryanodine receptors are involved and direct Ca2+ into
the subplasmalemmal space where it is extruded from the cell by the
plasmalemmal Ca2+-ATPase. In the process,
Ca2+-dependent K+ channels may or may not be
activated. Although extensive evidence has been given for such a
mechanism in vascular smooth muscle (103, 105, 137, 168),
this model has not been examined in ASM. However, there has been one
report of K+ currents activated by Ca2+ sparks
in ASM (although relaxants were not used in this study) (254), and the isoproterenol-induced elevation of
[Ca2+]i was shown to involve ryanodine
receptors (245). Isoproterenol (171), cAMP
(171), and cGMP (37) suppress the frequency
of cholinergic Ca2+ oscillations in ASM.
It appears, then, that bronchodilators simultaneously trigger uptake of
Ca2+ from the deep cytosol into the SR as well as release
of SR Ca2+ into the subplasmalemmal space, followed by
extrusion of Ca2+ from that peripheral space into the
extracellular space (Fig. 6). Our proposal that relaxants stimulate
Ca2+ release might be counterintuitive, but it should be
expected. The cells must be able to discharge internally sequestered
Ca2+, but without triggering contraction, to
increase/maintain the Ca2+-buffering capacity of the SR.
Data presented in an earlier study (108) suggest that
ryanodine receptors are involved and that this Ca2+ release
is preferentially directed into the subplasmalemmal space. The
important point to be made in all of this is that K+
channels appear to be more like bystanders than key players in the
process of relaxation.
What role do plasmalemmal Cl
channels play in ASM physiology?
There is substantial electrophysiological evidence for a population of
Cl
channels activated during excitatory stimulation
(81, 106, 117, 119, 120, 147, 148, 237, 238). These
channels exhibit a small unitary conductance (far below 20 pS)
(121), and their activation is Ca2+ dependent
(via Gq/G11-stimulated release of
internal Ca2+) (238) but voltage independent
(121). Moreover, these inactivate in a voltage-sensitive
fashion (121) via phosphorylation by
Ca2+/calmodulin-dependent kinase II (237)
(which is in turn triggered by the elevation of
[Ca2+]i that activated the channels in the
first place).
Given that excitatory stimulation is generally associated with such
large Cl
currents that can depolarize the membrane and
are tightly regulated by second messenger signaling events (121,
237, 241, 254), much as is the case in vascular smooth muscle,
it is easy to conclude that the Cl
currents play a key
role in contraction of ASM by depolarizing the membrane and thus
triggering voltage-dependent Ca2+ influx. However, the
truth of the matter is that agonist-evoked contractions are not
affected by Cl
channel blockers (Fig. 2). Also, given
that the equilibrium potential for Cl
is approximately
40 to
30 mV (4, 5), activation of Cl
channels will indeed depolarize the membrane but will essentially clamp
it at potentials barely sufficient for activation of voltage-dependent Ca2+ channels. Again, it bears repeating that substantial
contractions can still be evoked during voltage clamp at very negative
potentials (118) (Fig. 3), indicating that depolarization
is not necessary for contraction anyway. Why, then, are
Cl
currents so prominent in ASM?
Recently, another type of Cl
channel has been isolated
from ASM with properties diametrically opposite to those described above (195). That is, they have a large unitary
conductance (several hundred picosiemens) and their activation is
voltage dependent but Ca2+ independent. All of these
properties are similar to those of the Cl
channels
present on the SR of skeletal and cardiac muscle and facilitate
Ca2+ flux by neutralizing charge buildup on the SR
membranes (3).
This finding prompts us to propose an entirely novel and testable
hypothesis: that agonists activate Cl
currents in the
plasmalemma of the ASM cell to facilitate Ca2+
release/uptake. That is, Ca2+ efflux from the SR leads to a
net negative charge on the inner face of the SR membrane that hinders
Ca2+ release (Fig.
8A) unless alleviated by
compensatory fluxes of Cl
out of the SR (Fig.
8B) (134, 179). However, the accumulation of
Cl
outside the SR opposes further Cl
efflux
from the SR (and thus Ca2+ release; Fig. 8B). A
sudden opening of Cl
channels on the plasmalemma, with
subsequent loss of Cl
from the subplasmalemmal space,
would instantaneously alter the equilibrium potential for
Cl
across the SR membrane, thereby boosting efflux of
Cl
(and Ca2+) from the SR (Fig.
8C). Consistent with this, we have noted anecdotally that
agonist-evoked membrane currents and contractions were lost in cells
studied using a low internal [Cl
] solution (20 mM)
(117).

View larger version (15K):
[in this window]
[in a new window]
|
Fig. 8.
Hypothesized role for membrane Cl channels.
A: release of Ca2+ from the SR results in
accumulation of negative charge ( ) on the membrane of the SR, which
in turn hinders further Ca2+ release. B: efflux
of Cl from the SR compensates for the charge buildup and
allows greater Ca2+ release. However, accumulation of
Cl in the cytosol hinders both processes. C:
opening of Cl channels on the plasmalemma allows a
massive Cl efflux from the cell, instantaneously
increasing the driving force on Cl to exit the SR,
thereby allowing complete dissipation of the charge buildup and thus
facilitating further Ca2+ release.
|
|
Voltage-independent Ca2+ influx
pathways?
Agonists activate a membrane conductance that is nonselective for
various monovalent cations (117, 235, 239). It is unclear whether this conductance is Ca2+ permeable and, thereby,
serves as a source of Ca2+ for contraction as is the case
in vascular and gastrointestinal smooth muscle (51, 90,
94). Elevation of [Ca2+]i is necessary
but not sufficient for activation of these nonselective cation
channels; instead, they are activated by Gi/Go
(activated by M2 muscarinic or H1 histaminergic
receptors). While the Cl
current that is concurrently
activated by these agonists decays within seconds, the nonselective
cation channels remain open as long as the agonists are present,
resulting in a persistent noninactivating inward current.
 |
IONIC MECHANISMS IN ASM PATHOPHYSIOLOGY |
Inflammation plays a central role in asthma and airway
hyperreactivity, and it is becoming increasingly clear that cytokines and inflammatory mediators exert a variety of effects on various aspects of ASM function. Several researchers have sought to examine whether ionic mechanisms are altered in asthma or in animal models of
airway hyperresponsiveness. For example, it might be possible that
membrane potentials are higher or Ca2+ currents greater in
tissues/cells from asthmatics or hyperresponsive animals, as can happen
for vascular smooth muscle cells and hypertension (150,
217). However, immunological stimulation of excised guinea pig
tracheal tissues causes first a small and transient membrane depolarization, followed by a marked and prolonged membrane
hyperpolarization (213, 215). In excised tissues exposed
to allergen in vivo, membrane potentials were slightly more
hyperpolarized (<5 mV) when the guinea pigs had been acutely
sensitized to allergen but markedly depolarized (by >10 mV) when the
animals had been chronically exposed to allergen (156,
214). While canine ASM is normally very polarized (resting
potentials of approximately
60 mV) and does not show spontaneous
phasic electrical activity, ASM from dogs with "aspirin-induced
asthma" exhibited marked membrane depolarization and slow wave
activity (101). This might be related to the suppression of delayed rectifier K+ current (more specifically,
enhanced inactivation of the channels), which is reported to occur in
allergen-sensitized canine bronchial smooth muscle (234).
Despite all these observed changes in electrophysiological activity in
vitro, their significance to airway physiology/pathophysiology is
unclear given that electromechanical coupling is relatively unimportant
in ASM and that Ca2+ channel blockers and K+
channel openers are generally ineffective in the treatment of asthma in
the clinical setting (14, 41, 62, 65, 76, 86, 135, 162, 189, 202,
204).
On the other hand, asthma and airway hyperresponsiveness might be
associated with changes in Ca2+ handling. Perhaps basal
levels of [Ca2+]i are higher or
Ca2+ release is greater. Several proinflammatory cytokines
such as interleukin-1
, tumor necrosis factor-
, interferon-
,
platelet-derived growth factor, and eosinophil major basic protein
markedly augment excitatory agonist-evoked Ca2+ transients
(9, 10, 12, 244, 248), phosphoinositide turnover (10, 248), and contractions (11, 154, 175,
244). In some cases, these effects involve mitogen-activated
protein kinase cascades (ras, raf, MEK, Rho, Rho kinase) and induction
of gene expression, protein synthesis, and proliferation (10, 88, 89, 226).
Oxidizing pollutants such as ozone and acrolein induce airway
hyperreactivity. Although these act in part through inflammatory cells,
they can also alter EC coupling and other cellular events in isolated
ASM cells or tissues (15, 16, 91, 92, 153, 192). This
direct action on the ASM per se seems to involve changes in inositol
phosphate metabolism and Ca2+ handling, with induction or
augmentation of Ca2+ oscillations (91, 92, 192,
194).
 |
FUTURE DIRECTIONS |
On the basis of the arguments laid out in this review, we propose
the following recommendations in future studies of ASM physiology.
Ionic mechanisms.
It seems that too many groups are studying ionic mechanisms in ASM in
the same fashion as they would a vascular or gastrointestinal smooth
muscle preparation. There has been far too much emphasis on
voltage-dependent mechanisms, to an extent that is not warranted by
clinical studies of Ca2+ channel blockers or K+
channel agonists. As outlined above, ASM is distinct from skeletal, cardiac, and vascular smooth muscle in many respects pertaining to EC
coupling, and a great deal remains to be learned about its unique
physiology and pathophysiology. There needs to be greater consideration
of other roles for the ion channels that do not depend on, or relate
to, electromechanical coupling.
Ca2+ handling.
The hypothesis that Ca2+ may play an important causal role
in airway hyperreactivity and asthma was proposed more than two decades ago (161, 230) but still has not been explored in
sufficient detail. We believe this has resulted in an overly simplistic
model of Ca2+ handling in ASM and an imbalanced
understanding of electromechanical coupling mechanisms, as outlined
above. A breakdown in the superficial buffer barrier and/or changes in
Ca2+ sparks/Ca2+ oscillations may be more
relevant to airway hyperreactivity than increases in basal
[Ca2+]i or peak magnitudes of agonist-evoked
Ca2+ transients.
EC coupling.
A great deal more attention needs to be focused on recently discovered
EC coupling mechanisms in ASM, particularly Rho/Rho-activated kinase-mediated regulation of MLCP as well as pathways directed at thin
filaments. Moreover, most studies use only a maximally effective
concentration of agonist in their studies. The full range of agonist
concentrations should be examined, from subthreshold to maximally
effective concentrations, since there is now evidence that different EC
coupling mechanisms may contribute to differing degrees depending on
the level of excitation (124, 194). Also, experimental
strategies and tools need to be wielded with a greater degree of
sophistication. For example, we should be increasingly wary of the use
of micromolar concentrations of dihydropyridine blockers, or of
contractions as an index of [Ca2+]i, or of
focusing almost exclusively on the temporally and spatially averaged
spikelike elevations of [Ca2+] evoked by a maximally
effective concentration of agonist. Finally, the vast majority of
studies of airway function are done using tracheal tissues/cells. A
greater use of smaller airways is advocated because evidence is
accumulating for marked regional differences across the airway tree
(47, 124, 208).
In summary, a more imaginative approach to the study of EC coupling in
ASM, one that does not lean so heavily on electromechanical mechanisms,
could lead to major advances in our understanding of the unique
physiology (and pathophysiology) of ASM.
 |
FOOTNOTES |
Address for reprint requests and other correspondence:
L. J. Janssen, L-314, St. Joseph's Hospital, 40 Charlton
Ave. E., Hamilton, ON, Canada L8N 4A6 (E-mail:
janssenl{at}mcmaster.ca).
10.1152/ajplung.00452.2001
 |
REFERENCES |
1.
Abdullah, NA,
Hirata M,
Matsumoto K,
Aizawa H,
Inoue R,
Hamano S,
Ikeda S,
Xie Z,
Hara N,
and
Ito Y.
Contraction and depolarization induced by fetal bovine serum in airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
266:
L528-L535,
1994[Abstract/Free Full Text].
2.
Accomazzo, MR,
Rovati GE,
Vigano T,
Hernandez A,
Bonazzi A,
Bolla M,
Fumagalli F,
Viappiani S,
Galbiati E,
Ravasi S,
Albertoni C,
Di Luca M,
Caputi A,
Zannini P,
Chiesa G,
Villa AM,
Doglia SM,
Folco G,
and
Nicosia S.
Leukotriene D4-induced activation of smooth-muscle cells from human bronchi is partly Ca2+-independent.
Am J Respir Crit Care Med
163:
266-272,
2001[Abstract/Free Full Text].
3.
Ahern, GP,
and
Laver DR.
ATP inhibition and rectification of a Ca2+-activated anion channel in sarcoplasmic reticulum of skeletal muscle.
Biophys J
74:
2335-2351,
1998[Abstract/Free Full Text].
4.
Aickin, CC.
Chloride transport across the sarcolemma of vertebrate smooth and skeletal muscle.
In: Chloride Channels and Carriers in Nerve, Muscle, and Glial Cells, edited by Alvarez-Leefmans FJ,
and Russell JM.. New York: Plenum, 1990, p. 209-249.
5.
Aickin, CC,
and
Vermue NA.
Microelectrode measurement of intracellular chloride activity in smooth muscle cells of guinea-pig ureter.
Pflügers Arch
397:
25-28,
1983[ISI][Medline].
6.
Alioua, A,
Salvail D,
Dumoulin M,
Garon J,
Cadieux A,
and
Rousseau E.
Direct activation of KCa channel in airway smooth muscle by nitric oxide: involvement of a nitrothiosylation mechanism?
Am J Respir Cell Mol Biol
19:
485-497,
1998[Abstract/Free Full Text].
7.
Allen, SL,
Beech DJ,
Foster RW,
Morgan GP,
and
Small RC.
Electrophysiological and other aspects of the relaxant action of isoprenaline in guinea-pig isolated trachealis.
Br J Pharmacol
86:
843-854,
1985[Abstract].
8.
Amoako, D,
Qian Y,
Kwan CY,
and
Bourreau JP.
Probing excitation-contraction coupling in trachealis smooth muscle with the mycotoxin cyclopiazonic acid.
Clin Exp Pharmacol Physiol
23:
733-737,
1996[ISI][Medline].
9.
Amrani, Y,
and
Bronner C.
Tumor necrosis factor alpha potentiates the increase in cytosolic free calcium induced by bradykinin in guinea-pig tracheal smooth muscle cells.
C R Acad Sci III
316:
1489-1494,
1993[ISI][Medline].
10.
Amrani, Y,
Krymskaya V,
Maki C,
and
Panettieri RA, Jr.
Mechanisms underlying TNF-
effects on agonist-mediated calcium homeostasis in human airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
273:
L1020-L1028,
1997[Abstract/Free Full Text].
11.
Amrani, Y,
and
Panettieri RA, Jr.
Cytokines induce airway smooth muscle cell hyperresponsiveness to contractile agonists.
Thorax
53:
713-716,
1998[Free Full Text].
12.
Amrani, Y,
Panettieri RA, Jr,
Frossard N,
and
Bronner C.
Activation of the TNF alpha-p55 receptor induces myocyte proliferation and modulates agonist-evoked calcium transients in cultured human tracheal smooth muscle cells.
Am J Respir Cell Mol Biol
15:
55-63,
1996[Abstract].
13.
Asano, M,
Kuwako M,
Nomura Y,
Suzuki Y,
Shibuya M,
Sugita K,
and
Ito K.
Possible mechanism of the potent vasoconstrictor responses to ryanodine in dog cerebral arteries.
Eur J Pharmacol
311:
53-60,
1996[ISI][Medline].
14.
Barnes, PJ.
Clinical studies with calcium antagonists in asthma.
Br J Clin Pharmacol
20, Suppl2:
289S-298S,
1985[Medline].
15.
Ben Jebria, A,
Marthan R,
Rossetti M,
Savineau JP,
and
Ultman JS.
Effect of in vitro exposure to acrolein on carbachol responses in rat trachealis muscle.
Respir Physiol
93:
111-123,
1993[ISI][Medline].
16.
Ben Jebria, A,
Marthan R,
Rossetti M,
Savineau JP,
and
Ultman JS.
Human bronchial smooth muscle responsiveness after in vitro exposure to acrolein.
Am J Respir Crit Care Med
149:
382-386,
1994[Abstract].
17.
Berridge, MJ.
Inositol trisphosphate and calcium signalling.
Nature
361:
315-325,
1993[ISI][Medline].
18.
Berridge, MJ.
Capacitative calcium entry.
Biochem J
312:
1-11,
1995[ISI][Medline].
19.
Berridge, MJ.
The AM and FM of calcium signalling.
Nature
386:
759-760,
1997[ISI][Medline].
20.
Berridge, MJ,
Bootman MD,
and
Lipp P.
Calcium-a life and death signal.
Nature
395:
645-648,
1998[ISI][Medline].
21.
Berry, JL,
Elliott KR,
Foster RW,
Green KA,
Murray MA,
and
Small RC.
Mechanical, biochemical and electrophysiological studies of RP 49356 and cromakalim in guinea-pig and bovine trachealis muscle.
Pulm Pharmacol
4:
91-98,
1991[ISI][Medline].
22.
Black, JL,
Armour CL,
Johnson PR,
Alouan LA,
and
Barnes PJ.
The action of a potassium channel activator, BRL 38227 (lemakalim), on human airway smooth muscle.
Am Rev Respir Dis
142:
1384-1389,
1990[ISI][Medline].
23.
Bolton, TB,
and
Gordienko DV.
Confocal imaging of calcium release events in single smooth muscle cells.
Acta Physiol Scand
164:
567-575,
1998[ISI][Medline].
24.
Bourreau, JP.
Cross talk between plasma membrane and sarcoplasmic reticulum in canine airway smooth muscle.
Biol Signals
2:
272-283,
1993[Medline].
25.
Bourreau, JP,
Abela AP,
Kwan CY,
and
Daniel EE.
Acetylcholine Ca2+ stores refilling directly involves a dihydropyridine-sensitive channel in dog trachea.
Am J Physiol Cell Physiol
261:
C497-C505,
1991[Abstract/Free Full Text].
26.
Bourreau, JP,
Kwan CY,
and
Daniel EE.
Distinct pathways to refill ACh-sensitive internal Ca2+ stores in canine airway smooth muscle.
Am J Physiol Cell Physiol
265:
C28-C35,
1993[Abstract/Free Full Text].
27.
Boyle, JP,
Davies JM,
Foster RW,
Morgan GP,
and
Small RC.
Inhibitory responses to nicotine and transmural stimulation in hyoscine-treated guinea-pig isolated trachealis: an electrical and mechanical study.
Br J Pharmacol
90:
733-744,
1987[Abstract].
28.
Boyle, JP,
Tomasic M,
and
Kotlikoff MI.
Delayed rectifier potassium channels in canine and porcine airway smooth muscle cells.
J Physiol
447:
329-350,
1992[Abstract].
29.
Bremerich, DH,
Kai T,
Warner DO,
and
Jones KA.
Effect of phorbol esters on Ca2+ sensitivity and myosin light-chain phosphorylation in airway smooth muscle.
Am J Physiol Cell Physiol
274:
C1253-C1260,
1998[Abstract/Free Full Text].
30.
Cartin, L,
Lounsbury KM,
and
Nelson MT.
Coupling of Ca2+ to CREB activation and gene expression in intact cerebral arteries from mouse: roles of ryanodine receptors and voltage-dependent Ca2+ channels.
Circ Res
86:
760-767,
2000[Abstract/Free Full Text].
31.
Chiba, Y,
Sakai H,
and
Misawa M.
Augmented acetylcholine-induced translocation of RhoA in bronchial smooth muscle from antigen-induced airway hyperresponsive rats.
Br J Pharmacol
133:
886-890,
2001[Abstract/Free Full Text].
32.
Chilvers, ER,
Batty IH,
Barnes PJ,
and
Nahorski SR.
Formation of inositol polyphosphates in airway smooth muscle after muscarinic receptor stimulation.
J Pharmacol Exp Ther
252:
786-791,
1990[Abstract].
33.
Chilvers, ER,
Lynch BJ,
and
Challiss RA.
Phosphoinositide metabolism in airway smooth muscle.
Pharmacol Ther
62:
221-245,
1994[ISI][Medline].
34.
Chilvers, ER,
and
Nahorski SR.
Phosphoinositide metabolism in airway smooth muscle.
Am Rev Respir Dis
141:
S137-S140,
1990[ISI][Medline].
35.
Chin, D,
and
Means AR.
Calmodulin: a prototypical calcium sensor.
Trends Cell Biol
10:
322-328,
2000[ISI][Medline].
36.
Chiu, P,
Cook SJ,
Small RC,
Berry JL,
Carpenter JR,
Downing SJ,
Foster RW,
Miller AJ,
and
Small AM.
Beta-adrenoceptor subtypes and the opening of plasmalemmal K+-channels in bovine trachealis muscle: studies of mechanical activity and ion fluxes.
Br J Pharmacol
109:
1149-1156,
1993[Abstract].
37.
Choi, J,
and
Farley JM.
Effects of 8-bromo-cyclic GMP on membrane potential of single swine tracheal smooth muscle cells.
J Pharmacol Exp Ther
285:
588-594,
1998[Abstract/Free Full Text].
38.
Clapham, JC,
and
Wilson C.
Anti-spasmogenic and spasmolytic effects of BRL 34915: a comparison with nifedipine and nicorandil.
J Auton Pharmacol
7:
233-242,
1987[ISI][Medline].
39.
Coburn, RF.
Electromechanical coupling in canine trachealis muscle: acetylcholine contractions.
Am J Physiol Cell Physiol
236:
C177-C184,
1979[Medline].
40.
Coburn, RF,
and
Yamaguchi T.
Membrane potential-dependent and -independent tension in the canine tracheal muscle.
J Pharmacol Exp Ther
201:
276-284,
1977[Abstract].
41.
Cook, NS,
and
Chapman ID.
Therapeutic potential of potassium channel openers in peripheral vascular disease and asthma.
Cardiovasc Drugs Ther
7, Suppl3:
555-563,
1993[ISI][Medline].
42.
Cook, SJ,
Archer K,
Martin A,
Buchheit KH,
Fozard JR,
Muller T,
Miller AJ,
Elliott KR,
Foster RW,
and
Small RC.
Further analysis of the mechanisms underlying the tracheal relaxant action of SCA40.
Br J Pharmacol
114:
143-151,
1995[Abstract].
43.
Cook, SJ,
Small RC,
Berry JL,
Chiu P,
Downing SJ,
and
Foster RW.
Beta-adrenoceptor subtypes and the opening of plasmalemmal K+-channels in trachealis muscle: electrophysiological and mechanical studies in guinea-pig tissue.
Br J Pharmacol
109:
1140-1148,
1993[Abstract].
44.
Cooper, DM,
Mons N,
and
Karpen JW.
Adenylyl cyclases and the interaction between calcium and cAMP signalling.
Nature
374:
421-424,
1995[ISI][Medline].
45.
Corompt, E,
Bessard G,
Lantuejoul S,
Naline E,
Advenier C,
and
Devillier P.
Inhibitory effects of large Ca2+-activated K+ channel blockers on
-adrenergic- and NO donor-mediated relaxations of human and guinea-pig airway smooth muscles.
Naunyn Schmiedebergs Arch Pharmacol
357:
77-86,
1998[ISI][Medline].
46.
Cortijo, J,
Villagrasa V,
Marti-Cabrera M,
Villar V,
Moreau J,
Advenier C,
Morcillo EJ,
and
Small RC.
The spasmogenic effects of vanadate in human isolated bronchus.
Br J Pharmacol
121:
1339-1349,
1997[Abstract].
47.
Croxton, TL,
Fleming C,
and
Hirshman CA.
Expression of dihydropyridine resistance differs in porcine bronchial and tracheal smooth muscle.
Am J Physiol Lung Cell Mol Physiol
267:
L106-L112,
1994[Abstract/Free Full Text].
48.
Daniel, EE,
Jury J,
Serio R,
and
Jager LP.
Role of depolarization and calcium in contractions of canine trachealis from endogenous or exogenous acetylcholine.
Can J Physiol Pharmacol
69:
518-525,
1991[ISI][Medline].
49.
Darby, PJ,
Kwan CY,
and
Daniel EE.
Caveolae from canine airway smooth muscle contain the necessary components for a role in Ca2+ handling.
Am J Physiol Lung Cell Mol Physiol
279:
L1226-L1235,
2000[Abstract/Free Full Text].
50.
Davis, JP,
Tikunova SB,
Walsh MP,
and
Johnson JD.
Characterizing the response of calcium signal transducers to generated calcium transients.
Biochemistry
38:
4235-4244,
1999[ISI][Medline].
51.
Davis, MJ,
Donovitz JA,
and
Hood JD.
Stretch-activated single-channel and whole cell currents in vascular smooth muscle cells.
Am J Physiol Cell Physiol
262:
C1083-C1088,
1992[Abstract/Free Full Text].
52.
De Koninck, P,
and
Schulman H.
Sensitivity of CaM kinase II to the frequency of Ca2+ oscillations.
Science
279:
227-230,
1998[Abstract/Free Full Text].
53.
Dolmetsch, RE,
Pajvani U,
Fife K,
Spotts JM,
and
Greenberg ME.
Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway.
Science
294:
333-339,
2001[Abstract/Free Full Text].
54.
Dolmetsch, RE,
Xu K,
and
Lewis RS.
Calcium oscillations increase the efficiency and specificity of gene expression.
Nature
392:
933-936,
1998[ISI][Medline].
55.
Dupont, G,
and
Goldbeter A.
CaM kinase II as frequency decoder of Ca2+ oscillations.
Bioessays
20:
607-610,
1998[ISI][Medline].
56.
Durand, J,
and
Marmy N.
Arachidonic acid induces [Ca2+]i oscillations in smooth muscle cells from human airways.
Respir Physiol
97:
249-261,
1994[ISI][Medline].
57.
Eglen, RM,
Michel AD,
Sharif NA,
Swank SR,
and
Whiting RL.
The pharmacological properties of the peptide, endothelin.
Br J Pharmacol
97:
1297-1307,
1989[Abstract].
58.
Emala, CW,
Liu F,
and
Hirshman CA.
Gi
but not Gq
is linked to activation of p21(ras) in human airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
276:
L564-L570,
1999[Abstract/Free Full Text].
59.
Faraci, FM,
and
Heistad DD.
Regulation of the cerebral circulation: role of endothelium and potassium channels.
Physiol Rev
78:
53-97,
1998[Abstract/Free Full Text].
60.
Farley, JM,
and
Miles PR.
Role of depolarization in acetylcholine-induced contractions of dog trachealis muscle.
J Pharmacol Exp Ther
201:
199-205,
1977[Abstract].
61.
Farley, JM,
and
Miles PR.
The sources of calcium for acetylcholine-induced contractions of dog tracheal smooth muscle.
J Pharmacol Exp Ther
207:
340-346,
1978[Abstract].
62.
Faurschou, P,
Mikkelsen KL,
Steffensen I,
and
Franke B.
The lack of bronchodilator effect and the short-term safety of cumulative single doses of an inhaled potassium channel opener (bimakalim) in adult patients with mild to moderate bronchial asthma.
Pulm Pharmacol
7:
293-297,
1994[ISI][Medline].
63.
Felbel, J,
Trockur B,
Ecker T,
Landgraf W,
and
Hofmann F.
Regulation of cytosolic calcium by cAMP and cGMP in freshly isolated smooth muscle cells from bovine trachea.
J Biol Chem
263:
16764-16771,
1988[Abstract/Free Full Text].
64.
Fernandes, LB,
Fryer AD,
and
Hirshman CA.
M2 muscarinic receptors inhibit isoproterenol-induced relaxation of canine airway smooth muscle.
J Pharmacol Exp Ther
262:
119-126,
1992[Abstract].
65.
Fish, JE.
Calcium channel antagonists in the treatment of asthma.
J Asthma
21:
407-418,
1984[ISI][Medline].
66.
Fleckenstein, A,
Frey M,
and
Fleckenstein-Grun G.
Antihypertensive and arterial anticalcinotic effects of calcium antagonists.
Am J Cardiol
57:
1D-10D,
1986[Medline].
67.
Fleischmann, BK,
Murray RK,
and
Kotlikoff MI.
Voltage window for sustained elevation of cytosolic calcium in smooth muscle cells.
Proc Natl Acad Sci USA
91:
11914-11918,
1994[Abstract/Free Full Text].
68.
Fleischmann, BK,
Wang YX,
Pring M,
and
Kotlikoff MI.
Voltage-dependent calcium currents and cytosolic calcium in equine airway myocytes.
J Physiol
492:
347-358,
1996[Abstract].
69.
Fleischmann, BK,
Washabau RJ,
and
Kotlikoff MI.
Control of resting membrane potential by delayed rectifier potassium currents in ferret airway smooth muscle cells.
J Physiol
469:
625-638,
1993[Abstract].
70.
Fujiwara, T,
and
Angus JA.
Analysis of relaxation and repolarization mechanisms of nicorandil in rat mesenteric artery.
Br J Pharmacol
119:
1549-1556,
1996[Abstract].
71.
Genazzani, AA,
and
Galione A.
A Ca2+ release mechanism gated by the novel pyridine nucleotide, NAADP.
Trends Pharmacol Sci
18:
108-110,
1997[ISI][Medline].
72.
Gerthoffer, WT.
Regulation of the contractile element of airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
261:
L15-L28,
1991[Abstract/Free Full Text].
73.
Gerthoffer, WT,
and
Gunst SJ.
Invited review: focal adhesion and small heat shock proteins in the regulation of actin remodeling and contractility in smooth muscle.
J Appl Physiol
91:
963-972,
2001[Abstract/Free Full Text].
74.
Gerthoffer, WT,
Murphey KA,
and
Khoyi MA.
Inhibition of tracheal smooth muscle contraction and myosin phosphorylation by ryanodine.
J Pharmacol Exp Ther
246:
585-590,
1988[Abstract].
75.
Gerthoffer, WT,
Yamboliev IA,
Pohl J,
Haynes R,
Dang S,
and
McHugh J.
Activation of MAP kinases in airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
272:
L244-L252,
1997[Abstract/Free Full Text].
76.
Gordon, EH,
Wong SC,
and
Klaustermeyer WB.
Comparison of nifedipine with a new calcium channel blocker, flordipine, in exercise-induced asthma.
J Asthma
24:
261-265,
1987[ISI][Medline].
77.
Green, KA,
Small RC,
and
Foster RW.
The properties of voltage-operated Ca2+ channels in bovine isolated trachealis cells.
Pulm Pharmacol
6:
49-62,
1993[ISI][Medline].
78.
Gunst, SJ,
and
Tang DD.
The contractile apparatus and mechanical properties of airway smooth muscle.
Eur Respir J
15:
600-616,
2000[Abstract/Free Full Text].
79.
Hajnoczky, G,
Csordas G,
Madesh M,
and
Pacher P.
The machinery of local Ca2+ signalling between sarcoendoplasmic reticulum and mitochondria.
J Physiol
529:
69-81,
2000[Abstract/Free Full Text].
80.
Hakonarson, H,
and
Grunstein MM.
Regulation of second messengers associated with airway smooth muscle contraction and relaxation.
Am J Respir Crit Care Med
158:
S115-S122,
1998[Abstract/Free Full Text].
81.
Hazama, H,
Nakajima T,
Hamada E,
Omata M,
and
Kurachi Y.
Neurokinin A and Ca2+ current induce Ca2+-activated Cl
currents in guinea-pig tracheal myocytes.
J Physiol
492:
377-393,
1996[Abstract].
82.
Herrera, GM,
Heppner TJ,
and
Nelson MT.
Voltage dependence of the coupling of Ca2+ sparks to BKCa channels in urinary bladder smooth muscle.
Am J Physiol Cell Physiol
280:
C481-C490,
2001[Abstract/Free Full Text].
83.
Hirshman, CA,
and
Emala CW.
Actin reorganization in airway smooth muscle cells involves Gq and Gi-2 activation of Rho.
Am J Physiol Lung Cell Mol Physiol
277:
L653-L661,
1999[Abstract/Free Full Text].
84.
Hisada, T,
Kurachi Y,
and
Sugimoto T.
Properties of membrane currents in isolated smooth muscle cells from guinea-pig trachea.
Pflügers Arch
416:
151-161,
1990[ISI][Medline].
85.
Honda, K,
Satake T,
Takagi K,
and
Tomita T.
Effects of relaxants on electrical and mechanical activities in the guinea-pig tracheal muscle.
Br J Pharmacol
87:
665-671,
1986[Abstract].
86.
Hoppe, M,
Harman E,
and
Hendeles L.
The effect of inhaled gallopamil, a potent calcium channel blocker, on the late-phase response in subjects with allergic asthma.
J Allergy Clin Immunol
89:
688-695,
1992[ISI][Medline].
87.
Horowitz, A,
Menice CB,
Laporte R,
and
Morgan KG.
Mechanisms of smooth muscle contraction.
Physiol Rev
76:
967-1003,
1996[Abstract/Free Full Text].
88.
Hotta, K,
Emala CW,
and
Hirshman CA.
TNF-
upregulates Gi
and Gq
protein expression and function in human airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
276:
L405-L411,
1999[Abstract/Free Full Text].
89.
Hotta, K,
Hirshman CA,
and
Emala CW.
TNF-
increases transcription of G
(i-2) in human airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
279:
L319-L325,
2000[Abstract/Free Full Text].
90.
Hoyer, J,
Kohler R,
Haase W,
and
Distler A.
Upregulation of pressure-activated Ca2+-permeable cation channel in intact vascular endothelium of hypertensive rats.
Proc Natl Acad Sci USA
93:
11253-11258,
1996[Abstract/Free Full Text].
91.
Hyvelin, JM,
Roux E,
Prevost MC,
Savineau JP,
and
Marthan R.
Cellular mechanisms of acrolein-induced alteration in calcium signaling in airway smooth muscle.
Toxicol Appl Pharmacol
164:
176-183,
2000[ISI][Medline].
92.
Hyvelin, JM,
Savineau JP,
and
Marthan R.
Selected contribution: effect of the aldehyde acrolein on acetylcholine-induced membrane current in airway smooth muscle cells.
J Appl Physiol
90:
750-754,
2001[Abstract/Free Full Text].
93.
Iizuka, K,
Dobashi K,
Yoshii A,
Horie T,
Suzuki H,
Nakazawa T,
and
Mori M.
Receptor-dependent G protein-mediated Ca2+ sensitization in canine airway smooth muscle.
Cell Calcium
22:
21-30,
1997[ISI][Medline].
94.
Inoue, R.
Ion channels involved in responses to muscarinic receptor activation in smooth muscle.
In: Ion Channels of Vascular Smooth Muscle Cells and Endothelial Cells, edited by Sperelakis N,
and Kuriyama H.. New York: Elsevier, 1991, p. 81-91.
95.
Inoue, T,
and
Ito Y.
Pre- and post-junctional actions of prostaglandin I2, carbocyclic thromboxane A2 and leukotriene C4 in dog tracheal tissue.
Br J Pharmacol
84:
289-298,
1985[Abstract].
96.
Inoue, T,
and
Ito Y.
Characteristics of neuro-effector transmission in the smooth muscle layer of dog bronchiole and modifications by autacoids.
J Physiol
370:
551-565,
1986[Abstract].
97.
Isaac, L,
McArdle S,
Miller NM,
Foster RW,
and
Small RC.
Effects of some K+-channel inhibitors on the electrical behaviour of guinea-pig isolated trachealis and on its responses to spasmogenic drugs.
Br J Pharmacol
117:
1653-1662,
1996[Abstract].
98.
Ito, S,
Kume H,
Honjo H,
Katoh H,
Kodama I,
Yamaki K,
and
Hayashi H.
Possible involvement of Rho kinase in Ca2+ sensitization and mobilization by MCh in tracheal smooth muscle.
Am J Physiol Lung Cell Mol Physiol
280:
L1218-L1224,
2001[Abstract/Free Full Text].
99.
Ito, Y.
Pre- and post-junctional actions of procaterol, a beta 2-adrenoceptor stimulant, on dog tracheal tissue.
Br J Pharmacol
95:
268-274,
1988[Abstract].
100.
Ito, Y,
Suzuki H,
Aizawa H,
Hakoda H,
and
Hirose T.
The spontaneous electrical and mechanical activity of human bronchial smooth muscle: its modulation by drugs.
Br J Pharmacol
98:
1249-1260,
1989[Abstract].
101.
Ito, Y,
and
Tajima K.
Spontaneous activity in the trachea of dogs treated with indomethacin: an experimental model for aspirin-related asthma.
Br J Pharmacol
73:
563-571,
1981[Abstract].
102.
Ito, Y,
Takagi K,
and
Tomita T.
Relaxant actions of isoprenaline on guinea-pig isolated tracheal smooth muscle.
Br J Pharmacol
116:
2738-2742,
1995[Abstract].
103.
Jaggar, JH,
Porter VA,
Lederer WJ,
and
Nelson MT.
Calcium sparks in smooth muscle.
Am J Physiol Cell Physiol
278:
C235-C256,
2000[Abstract/Free Full Text].
104.
Jaggar, JH,
Stevenson AS,
and
Nelson MT.
Voltage dependence of Ca2+ sparks in intact cerebral arteries.
Am J Physiol Cell Physiol
274:
C1755-C1761,
1998[Abstract/Free Full Text].
105.
Jaggar, JH,
Wellman GC,
Heppner TJ,
Porter VA,
Perez GJ,
Gollasch M,
Kleppisch T,
Rubart M,
Stevenson AS,
Lederer WJ,
Knot HJ,
Bonev AD,
and
Nelson MT.
Ca2+ channels, ryanodine receptors and Ca2+-activated K+ channels: a functional unit for regulating arterial tone.
Acta Physiol Scand
164:
577-587,
1998[ISI][Medline].
106.
Janssen, LJ.
Acetylcholine and caffeine activate Cl
and suppress K+ conductances in human bronchial smooth muscle.
Am J Physiol Lung Cell Mol Physiol
270:
L772-L781,
1996[Abstract/Free Full Text].
107.
Janssen, LJ.
T-type and L-type Ca2+ currents in canine bronchial smooth muscle: characterization and physiological roles.
Am J Physiol Cell Physiol
272:
C1757-C1765,
1997[Abstract/Free Full Text].
108.
Janssen, LJ,
Betti PA,
Netherton SJ,
and
Walters DK.
Superficial buffer barrier and preferentially directed release of Ca2+ in canine airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
276:
L744-L753,
1999[Abstract/Free Full Text].
109.
Janssen, LJ,
and
Daniel EE.
Characterization of the prejunctional beta adrenoceptors in canine bronchial smooth muscle.
J Pharmacol Exp Ther
254:
741-749,
1990[Abstract].
110.
Janssen, LJ,
and
Daniel EE.
Classification of postjunctional beta adrenoceptors mediating relaxation of canine bronchi.
J Pharmacol Exp Ther
256:
670-676,
1991[Abstract].
111.
Janssen, LJ,
and
Daniel EE.
Depolarizing agents induce oscillations in canine bronchial smooth muscle membrane potential: possible mechanisms.
J Pharmacol Exp Ther
259:
110-117,
1991[Abstract].
112.
Janssen, LJ,
and
Daniel EE.
Pre- and postjunctional effects of a thromboxane mimetic in canine bronchi.
Am J Physiol Lung Cell Mol Physiol
261:
L271-L276,
1991[Abstract/Free Full Text].
113.
Janssen, LJ,
Hague C,
and
Nana R.
Ionic mechanisms underlying electrical slow waves in canine airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
275:
L516-L523,
1998[Abstract/Free Full Text].
114.
Janssen, LJ,
and
Nana R.
Na+/K+ ATPase mediates rhythmic spontaneous relaxations in canine airway smooth muscle.
Respir Physiol
108:
187-194,
1997[ISI][Medline].
115.
Janssen, LJ,
Netherton SJ,
and
Walters DK.
Ca2+-dependent K+ channels and Na+-K+-ATPase mediate H2O2- and superoxide-induced relaxations in canine trachealis.
J Appl Physiol
88:
745-752,
2000[Abstract/Free Full Text].
116.
Janssen, LJ,
Premji M,
Lu-Chao H,
Cox G,
and
Keshavjee S.
NO+ but not NO radical relaxes airway smooth muscle via cGMP-independent release of internal Ca2+.
Am J Physiol Lung Cell Mol Physiol
278:
L899-L905,
2000[Abstract/Free Full Text].
117.
Janssen, LJ,
and
Sims SM.
Acetylcholine activates non-selective cation and chloride conductances in canine and guinea-pig tracheal myocytes.
J Physiol
453:
197-218,
1992[Abstract].
118.
Janssen, LJ,
and
Sims SM.
Emptying and refilling of Ca2+ store in tracheal myocytes as indicated by ACh-evoked currents and contraction.
Am J Physiol Cell Physiol
265:
C877-C886,
1993[Abstract/Free Full Text].
119.
Janssen, LJ,
and
Sims SM.
Histamine activates Cl
and K+ currents in guinea-pig tracheal myocytes: convergence with muscarinic signalling pathway.
J Physiol
465:
661-677,
1993[Abstract].
120.
Janssen, LJ,
and
Sims SM.
Substance P activates Cl
and K+ conductances in guinea-pig tracheal smooth muscle cells.
Can J Physiol Pharmacol
72:
705-710,
1994[ISI][Medline].
121.
Janssen, LJ,
and
Sims SM.
Ca2+-dependent Cl
current in canine tracheal smooth muscle cells.
Am J Physiol Cell Physiol
269:
C163-C169,
1995[Abstract/Free Full Text].
122.
Janssen, LJ,
Walters DK,
and
Wattie J.
Regulation of [Ca2+]i in canine airway smooth muscle by Ca2+-ATPase and Na+/Ca2+ exchange mechanisms.
Am J Physiol Lung Cell Mol Physiol
273:
L322-L330,
1997[Abstract/Free Full Text].
123.
Janssen, LJ,
and
Wattie J.
Non-neurogenic electrically evoked relaxation in canine airway muscle involves action of free radicals on K+ channels.
J Pharmacol Exp Ther
279:
813-821,
1996[Abstract].
124.
Janssen, LJ,
Wattie J,
Lu-Chao H,
and
Tazzeo T.
Muscarinic excitation-contraction coupling mechanisms in tracheal and bronchial smooth muscles.
J Appl Physiol
91:
1142-1151,
2001[Abstract/Free Full Text].
125.
Jing, L,
Inoue R,
Tashiro K,
Takahashi S,
and
Ito Y.
Role of nitric oxide in non-adrenergic, non-cholinergic relaxation and modulation of excitatory neuroeffector transmission in the cat airway.
J Physiol
483:
225-237,
1995[Abstract].
126.
Jones, KA,
Wong GY,
Jankowski CJ,
Akao M,
and
Warner DO.
cGMP modulation of Ca2+ sensitivity in airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
276:
L35-L40,
1999[Abstract/Free Full Text].
127.
Jones, TR,
Charette L,
Garcia ML,
and
Kaczorowski GJ.
Selective inhibition of relaxation of guinea-pig trachea by charybdotoxin, a potent Ca2+-activated K+ channel inhibitor.
J Pharmacol Exp Ther
255:
697-706,
1990[Abstract].
128.
Jones, TR,
Charette L,
Garcia ML,
and
Kaczorowski GJ.
Interaction of iberiotoxin with
-adrenoceptor agonists and sodium nitroprusside on guinea pig trachea.
J Appl Physiol
74:
1879-1884,
1993[Abstract].
129.
Kamei, K,
Nabata H,
and
Kuriyama H.
Effects of KC399, a novel ATP-sensitive K+ channel opener, on electrical and mechanical responses in dog tracheal smooth muscle.
J Pharmacol Exp Ther
268:
319-327,
1994[Abstract].
130.
Kamei, K,
Nabata H,
Kuriyama H,
Watanabe Y,
and
Itoh T.
Effect of KC399, a newly synthesized K+ channel opener, on acetylcholine-induced electrical and mechanical activities in rabbit tracheal smooth muscle.
Br J Pharmacol
115:
1493-1501,
1995[Abstract].
131.
Kamei, K,
Yoshida S,
Imagawa J,
Nabata H,
and
Kuriyama H.
Regional and species differences in glyburide-sensitive K+ channels in airway smooth muscles as estimated from actions of KC128 and levcromakalim.
Br J Pharmacol
113:
889-897,
1994[Abstract].
132.
Kannan, MS,
Jager LP,
Daniel EE,
and
Garfield RE.
Effects of 4-aminopyridine and tetraethylammonium chloride on the electrical activity and cable properties of canine tracheal smooth muscle.
J Pharmacol Exp Ther
227:
706-715,
1983[Abstract].
133.
Kannan, MS,
Prakash YS,
Johnson DE,
and
Sieck GC.
Nitric oxide inhibits calcium release from sarcoplasmic reticulum of porcine tracheal smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
272:
L1-L7,
1997[Abstract/Free Full Text].
134.
Kargacin, GJ,
Ali Z,
Zhang SJ,
Pollock NS,
and
Kargacin ME.
Iodide and bromide inhibit Ca2+ uptake by cardiac sarcoplasmic reticulum.
Am J Physiol Heart Circ Physiol
280:
H1624-H1634,
2001[Abstract/Free Full Text].
135.
Kidney, JC,
Fuller RW,
Worsdell YM,
Lavender EA,
Chung KF,
and
Barnes PJ.
Effect of an oral potassium channel activator, BRL 38227, on airway function and responsiveness in asthmatic patients: comparison with oral salbutamol.
Thorax
48:
130-133,
1993[Abstract].
136.
Kirkpatrick, CT.
Excitation and contraction in bovine tracheal smooth muscle.
J Physiol
244:
263-281,
1975[Abstract].
137.
Knot, HJ,
and
Nelson MT.
Regulation of arterial diameter and wall [Ca2+] in cerebral arteries of rat by membrane potential and intravascular pressure.
J Physiol
508:
199-209,
1998[Abstract/Free Full Text].
138.
Kobayashi, S,
Gong MC,
Somlyo AV,
and
Somlyo AP.
Ca2+ channel blockers distinguish between G protein-coupled pharmacomechanical Ca2+ release and Ca2+ sensitization.
Am J Physiol Cell Physiol
260:
C364-C370,
1991[Abstract/Free Full Text].
139.
Koopman, WJ,
Scheenen WJ,
Errington RJ,
Willems PH,
Bindels RJ,
Roubos EW,
and
Jenks BG.
Membrane-initiated Ca2+ signals are reshaped during propagation to subcellular regions.
Biophys J
81:
57-65,
2001[Abstract/Free Full Text].
140.
Kotlikoff, MI.
Calcium currents in isolated canine airway smooth muscle cells.
Am J Physiol Cell Physiol
254:
C793-C801,
1988[Abstract/Free Full Text].
141.
Kotlikoff, MI.
Potassium currents in canine airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
259:
L384-L395,
1990[Abstract/Free Full Text].
142.
Kotlikoff, MI.
Potassium channels in airway smooth muscle: a tale of two channels.
Pharmacol Ther
58:
1-12,
1993[ISI][Medline].
143.
Kotlikoff, MI,
and
Wang YX.
Calcium release and calcium-activated chloride channels in airway smooth muscle cells.
Am J Respir Crit Care Med
158:
S109-S114,
1998[Abstract/Free Full Text].
144.
Kume, H,
and
Kotlikoff MI.
Muscarinic inhibition of single KCa channels in smooth muscle cells by a pertussis-sensitive G protein.
Am J Physiol Cell Physiol
261:
C1204-C1209,
1991[Abstract/Free Full Text].
145.
Larsson-Backstrom, C,
Arrhenius E,
and
Sagge K.
Comparison of the calcium-antagonistic effects of terodiline, nifedipine and verapamil.
Acta Pharmacol Toxicol (Copenh)
57:
8-17,
1985[Medline].
146.
Li, W,
Llopis J,
Whitney M,
Zlokarnik G,
and
Tsien RY.
Cell-permeant caged InsP3 ester shows that Ca2+ spike frequency can optimize gene expression.
Nature
392:
936-941,
1998[ISI][Medline].
147.
Liu, X,
and
Farley JM.
Acetylcholine-induced Ca2+-dependent chloride current oscillations are mediated by inositol 1,4,5-trisphosphate in tracheal myocytes.
J Pharmacol Exp Ther
277:
796-804,
1996[Abstract].
148.
Liu, X,
and
Farley JM.
Acetylcholine-induced chloride current oscillations in swine tracheal smooth muscle cells.
J Pharmacol Exp Ther
276:
178-186,
1996[Abstract].
149.
Liu, X,
and
Farley JM.
Depletion and refilling of acetylcholine- and caffeine-sensitive Ca2+ stores in tracheal myocytes.
J Pharmacol Exp Ther
277:
789-795,
1996[Abstract].
150.
Lozinskaya, IM,
and
Cox RH.
Effects of age on Ca2+ currents in small mesenteric artery myocytes from Wistar-Kyoto and spontaneously hypertensive rats.
Hypertension
29:
1329-1336,
1997[Abstract/Free Full Text].
151.
Madison, JM,
Ethier MF,
and
Yamaguchi H.
Refilling of caffeine-sensitive intracellular calcium stores in bovine airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
275:
L852-L860,
1998[Abstract/Free Full Text].
152.
Marthan, R,
Martin C,
Amedee T,
and
Mironneau J.
Calcium channel currents in isolated smooth muscle cells from human bronchus.
J Appl Physiol
66:
1706-1714,
1989[Abstract/Free Full Text].
153.
Marthan, R,
Roux E,
and
Savineau JP.
Human bronchial smooth muscle responsiveness after in vitro exposure to oxidizing pollutants.
Cell Biol Toxicol
12:
245-249,
1996[ISI][Medline].
154.
Martin, C,
Uhlig S,
and
Ullrich V.
Cytokine-induced bronchoconstriction in precision-cut lung slices is dependent upon cyclooxygenase-2 and thromboxane receptor activation.
Am J Respir Cell Mol Biol
24:
139-145,
2001[Abstract/Free Full Text].
155.
McCaig, DJ.
Electrophysiology of neuroeffector transmission in the isolated, innervated trachea of the guinea-pig.
Br J Pharmacol
89:
793-801,
1986[Abstract].
156.
McCaig, DJ,
and
Souhrada JF.
Alteration of electrophysiological properties of airway smooth muscle from sensitized guinea-pigs.
Respir Physiol
41:
49-60,
1980[ISI][Medline].
157.
McCann, JD,
and
Welsh MJ.
Calcium-activated potassium channels in canine airway smooth muscle.
J Physiol
372:
113-127,
1986[Abstract].
158.
McDonald, TF,
Pelzer S,
Trautwein W,
and
Pelzer DJ.
Regulation and modulation of calcium channels in cardiac, skeletal, and smooth muscle cells.
Physiol Rev
74:
365-507,
1994[Free Full Text].
159.
Mehta, D,
Tang DD,
Wu MF,
Atkinson S,
and
Gunst SJ.
Role of Rho in Ca2+-insensitive contraction and paxillin tyrosine phosphorylation in smooth muscle.
Am J Physiol Cell Physiol
279:
C308-C318,
2000[Abstract/Free Full Text].
160.
Mehta, D,
Wang Z,
Wu MF,
and
Gunst SJ.
Relationship between paxillin and myosin phosphorylation during muscarinic stimulation of smooth muscle.
Am J Physiol Cell Physiol
274:
C741-C747,
1998[Abstract/Free Full Text].
161.
Middleton, E, Jr.
Airway smooth muscle, asthma, and calcium ions.
J Allergy Clin Immunol
73:
643-650,
1984[ISI][Medline].
162.
Middleton, E, Jr.
The treatment of asthma-beyond bronchodilators.
N Engl Reg Allergy Proc
6:
235-237,
1985[ISI][Medline].
163.
Miura, M,
Belvisi MG,
Stretton CD,
Yacoub MH,
and
Barnes PJ.
Role of potassium channels in bronchodilator responses in human airways.
Am Rev Respir Dis
146:
132-136,
1992[ISI][Medline].
164.
Muraki, K,
Imaizumi Y,
Kojima T,
Kawai T,
and
Watanabe M.
Effects of tetraethylammonium and 4-aminopyridine on outward currents and excitability in canine tracheal smooth muscle cells.
Br J Pharmacol
100:
507-515,
1990[Abstract].
165.
Murray, MA,
Berry JL,
Cook SJ,
Foster RW,
Green KA,
and
Small RC.
Guinea-pig isolated trachealis: the effects of charybdotoxin on mechanical activity, membrane potential changes and the activity of plasmalemmal K+-channels.
Br J Pharmacol
103:
1814-1818,
1991[Abstract].
167.
Nakai, J,
Dirksen RT,
Nguyen HT,
Pessah IN,
Beam KG,
and
Allen PD.
Enhanced dihydropyridine receptor channel activity in the presence of ryanodine receptor.
Nature
380:
72-75,
1996[ISI][Medline].
168.
Nelson, MT,
Cheng H,
Rubart M,
Santana LF,
Bonev AD,
Knot HJ,
and
Lederer WJ.
Relaxation of arterial smooth muscle by calcium sparks.
Science
270:
633-637,
1995[Abstract].
169.
Nelson, MT,
and
Quayle JM.
Physiological roles and properties of potassium channels in arterial smooth muscle.
Am J Physiol Cell Physiol
268:
C799-C822,
1995[Abstract/Free Full Text].
170.
Nixon, GF,
Mignery GA,
and
Somlyo AV.
Immunogold localization of inositol 1,4,5-trisphosphate receptors and characterization of ultrastructural features of the sarcoplasmic reticulum in phasic and tonic smooth muscle.
J Muscle Res Cell Motil
15:
682-700,
1994[ISI][Medline].
171.
Nuttle, LC,
and
Farley JM.
Frequency modulation of acetylcholine-induced oscillations in Ca2+ and Ca2+-activated Cl
current by cAMP in tracheal smooth muscle.
J Pharmacol Exp Ther
277:
753-760,
1996[Abstract].
172.
Pabelick, CM,
Prakash YS,
Kannan MS,
and
Sieck GC.
Spatial and temporal aspects of calcium sparks in porcine tracheal smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
277:
L1018-L1025,
1999[Abstract/Free Full Text].
173.
Pabelick, CM,
Sieck GC,
and
Prakash YS.
Invited Review: Significance of spatial and temporal heterogeneity of calcium transients in smooth muscle.
J Appl Physiol
91:
488-496,
2001[Abstract/Free Full Text].
174.
Pabelick, CM,
Warner DO,
Perkins WJ,
and
Jones KA.
S-nitrosoglutathione-induced decrease in calcium sensitivity of airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
278:
L521-L527,
2000[Abstract/Free Full Text].
175.
Parris, JR,
Cobban HJ,
Littlejohn AF,
MacEwan DJ,
and
Nixon GF.
Tumour necrosis factor-alpha activates a calcium sensitization pathway in guinea-pig bronchial smooth muscle.
J Physiol (Lond)
518:
561-569,
1999[Abstract/Free Full Text].
176.
Patterson, RL,
van Rossum DB,
and
Gill DL.
Store-operated Ca2+ entry: evidence for a secretion-like coupling model.
Cell
98:
487-499,
1999[ISI][Medline].
177.
Pavalko, FM,
Adam LP,
Wu MF,
Walker TL,
and
Gunst SJ.
Phosphorylation of dense-plaque proteins talin and paxillin during tracheal smooth muscle contraction.
Am J Physiol Cell Physiol
268:
C563-C571,
1995[Abstract/Free Full Text].
178.
Pocock, TM,
Laurent F,
Isaac LM,
Chiu P,
Elliott KR,
Foster RW,
Michel A,
Bonnet PA,
and
Small RC.
Effects of SCA40 on bovine trachealis muscle and on cyclic nucleotide phosphodiesterases.
Eur J Pharmacol
334:
75-85,
1997[ISI][Medline].
179.
Pollock, NS,
Kargacin ME,
and
Kargacin GJ.
Chloride channel blockers inhibit Ca2+ uptake by the smooth muscle sarcoplasmic reticulum.
Biophys J
75:
1759-1766,
1998[Abstract/Free Full Text].
180.
Prakash, YS,
Kannan MS,
and
Sieck GC.
Nitric oxide inhibits ACh-induced intracellular calcium oscillations in porcine tracheal smooth muscle.
Am J Physiol Lung Cell Mol Physiol
272:
L588-L596,
1997[Abstract/Free Full Text].
181.
Prakash, YS,
Kannan MS,
and
Sieck GC.
Regulation of intracellular calcium oscillations in porcine tracheal smooth muscle cells.
Am J Physiol Cell Physiol
272:
C966-C975,
1997[Abstract/Free Full Text].
182.
Prakash, YS,
Pabelick CM,
Kannan MS,
and
Sieck GC.
Spatial and temporal aspects of ACh-induced [Ca2+]i oscillations in porcine tracheal smooth muscle.
Cell Calcium
27:
153-162,
2000[ISI][Medline].
183.
Prakash, YS,
van der Heijden HF,
Kannan MS,
and
Sieck GC.
Effects of salbutamol on intracellular calcium oscillations in porcine airway smooth muscle.
J Appl Physiol
82:
1836-1843,
1997[Abstract/Free Full Text].
184.
Putney, JW, Jr.
Calcium signaling: up, down, up, down ... what's the point?
Science
279:
191-192,
1998[Free Full Text].
185.
Pyne, NJ,
Grady MW,
Shehnaz D,
Stevens PA,
Pyne S,
and
Rodger IW.
Muscarinic blockade of beta-adrenoceptor-stimulated adenylyl cyclase: the role of stimulatory and inhibitory guanine-nucleotide binding regulatory proteins (Gs and Gi).
Br J Pharmacol
107:
881-887,
1992[Abstract].
186.
Qian, Y,
and
Bourreau JP.
Two distinct pathways for refilling Ca2+ stores in permeabilized bovine trachealis muscle.
Life Sci
64:
2049-2059,
1999[ISI][Medline].
187.
Quayle, JM,
Nelson MT,
and
Standen NB.
ATP-sensitive and inwardly rectifying potassium channels in smooth muscle.
Physiol Rev
77:
1165-1232,
1997[Abstract/Free Full Text].
188.
Richards, IS,
Miller L,
Solomon D,
Kulkarni A,
Brooks S,
and
Sperelakis N.
Azelastine and desmethylazelastine suppress acetylcholine-induced contraction and depolarization in human airway smooth muscle.
Eur J Pharmacol
186:
331-334,
1990[ISI][Medline].
189.
Riska, H,
Stenius-Aaniala B,
and
Arvi AR.
Comparison of the efficacy of an ACE-inhibitor and a calcium channel blocker in hypertensive asthmatics. A preliminary report.
Postgrad Med J
62, Suppl1:
52-53,
1986[ISI][Medline].
190.
Rosado, JA,
and
Sage SO.
The actin cytoskeleton in store-mediated calcium entry.
J Physiol
526:
221-229,
2000[Abstract/Free Full Text].
191.
Rosales, C,
and
Brown EJ.
Calcium channel blockers nifedipine and diltiazem inhibit Ca2+ release from intracellular stores in neutrophils.
J Biol Chem
267:
1443-1448,
1992[Abstract/Free Full Text].
192.
Roux, E,
Guibert C,
Crevel H,
Savineau JP,
and
Marthan R.
Human and rat airway smooth muscle responsiveness after ozone exposure in vitro.
Am J Physiol Lung Cell Mol Physiol
271:
L631-L636,
1996[Abstract/Free Full Text].
193.
Roux, E,
Guibert C,
Savineau JP,
and
Marthan R.
[Ca2+]i oscillations induced by muscarinic stimulation in airway smooth muscle cells: receptor subtypes and correlation with the mechanical activity.
Br J Pharmacol
120:
1294-1301,
1997[Abstract].
194.
Roux, E,
Hyvelin JM,
Savineau JP,
and
Marthan R.
Calcium signaling in airway smooth muscle cells is altered by in vitro exposure to the aldehyde acrolein.
Am J Respir Cell Mol Biol
19:
437-444,
1998[Abstract/Free Full Text].
195.
Salvail, D,
Alioua A,
and
Rousseau E.
Functional identification of a sarcolemmal chloride channel from bovine tracheal smooth muscle.
Am J Physiol Cell Physiol
271:
C1716-C1724,
1996[Abstract/Free Full Text].
196.
Schramm, CM,
Chuang ST,
and
Grunstein MM.
cAMP generation inhibits inositol 1,4,5-trisphosphate binding in rabbit tracheal smooth muscle.
Am J Physiol Lung Cell Mol Physiol
269:
L715-L719,
1995[Abstract/Free Full Text].
197.
Shen, S,
Huang Y,
and
Bourreau JP.
Efficacy of muscarinic stimulation and mode of excitation-contraction coupling in bovine trachealis muscle.
Life Sci
67:
1833-1846,
2000[ISI][Medline].
198.
Shieh, CC,
Petrini MF,
Dwyer TM,
and
Farley JM.
Concentration-dependence of acetylcholine-induced changes in calcium and tension in swine trachealis.
J Pharmacol Exp Ther
256:
141-148,
1991[Abstract].
199.
Sieck, GC,
Kannan MS,
and
Prakash YS.
Heterogeneity in dynamic regulation of intracellular calcium in airway smooth muscle cells.
Can J Physiol Pharmacol
75:
878-888,
1997[ISI][Medline].
200.
Sims, SM,
Jiao Y,
and
Zheng ZG.
Intracellular calcium stores in isolated tracheal smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
271:
L300-L309,
1996[Abstract/Free Full Text].
201.
Sitsapesan, R,
McGarry SJ,
and
Williams AJ.
Cyclic ADP-ribose competes with ATP for the adenine nucleotide binding site on the cardiac ryanodine receptor Ca2+-release channel.
Circ Res
75:
596-600,
1994[Abstract].
202.
Sly, PD,
Olinsky A,
and
Landau LI.
Does nifedipine affect the diurnal variation of asthma in children?
Pediatr Pulmonol
2:
206-210,
1986[ISI][Medline].
203.
Small, RC,
Berry JL,
Boyle JP,
Chapman ID,
Elliott KR,
Foster RW,
and
Watt AJ.
Biochemical and electrical aspects of the tracheal relaxant action of AH 21-132.
Eur J Pharmacol
192:
417-426,
1991[ISI][Medline].
204.
Small, RC,
Berry JL,
Burka JF,
Cook SJ,
Foster RW,
Green KA,
and
Murray MA.
Potassium channel activators and bronchial asthma.
Clin Exp Allergy
22:
11-18,
1992[ISI][Medline].
205.
Small, RC,
Foster RW,
Berry JL,
Chapman ID,
and
Elliott KR.
The bronchodilator action of AH 21-132.
Agents Actions Suppl
34:
3-26,
1991[Medline].
206.
Smith, PG,
Garcia R,
and
Kogerman L.
Mechanical strain increases protein tyrosine phosphorylation in airway smooth muscle cells.
Exp Cell Res
239:
353-360,
1998[ISI][Medline].
207.
Snetkov, VA,
Pandya H,
Hirst SJ,
and
Ward JP.
Potassium channels in human fetal airway smooth muscle cells.
Pediatr Res
43:
548-554,
1998[Abstract].
208.
Snetkov, VA,
and
Ward JP.
Ion currents in smooth muscle cells from human small bronchioles: presence of an inward rectifier K+ current and three types of large conductance K+ channel.
Exp Physiol
84:
835-846,
1999[Abstract].
209.
Somlyo, AP,
and
Somlyo AV.
Signal transduction and regulation in smooth muscle.
Nature
372:
231-236,
1994[ISI][Medline].
210.
Somlyo, AP,
and
Somlyo AV.
From pharmacomechanical coupling to G-proteins and myosin phosphatase.
Acta Physiol Scand
164:
437-448,
1998[ISI][Medline].
211.
Somlyo, AP,
and
Somlyo AV.
Signal transduction by G-proteins, Rho-kinase and protein phosphatase to smooth muscle and non-muscle myosin II.
J Physiol
522:
177-185,
2000[Abstract/Free Full Text].
212.
Somlyo, AP,
Wu X,
Walker LA,
and
Somlyo AV.
Pharmacomechanical coupling: the role of calcium, G-proteins, kinases and phosphatases.
Rev Physiol Biochem Pharmacol
134:
201-234,
1999[Medline].
213.
Souhrada, M,
Showell HJ,
and
Souhrada JF.
Differential effects of N-formyl-methionyl-leucyl-phenylalanine and substance P on the electrical and contractile properties of airway smooth muscle cells.
Am Rev Respir Dis
135:
557-561,
1987[ISI][Medline].
214.
Souhrada, M,
and
Souhrada JF.
Reassessment of electrophysiological and contractile characteristics of sensitized airway smooth muscle.
Respir Physiol
46:
17-27,
1981[ISI][Medline].
215.
Souhrada, M,
and
Souhrada JF.
Effect of IgG1 and its fragments on resting membrane potential of isolated tracheal myocytes.
J Appl Physiol
74:
1948-1953,
1993[Abstract].
216.
Stockbridge, LL,
French AS,
and
Man SF.
Subconductance states in calcium-activated potassium channels from canine airway smooth muscle.
Biochim Biophys Acta
1064:
212-218,
1991[ISI][Medline].
217.
Storm, DS,
Turla MB,
Todd KM,
and
Webb RC.
Calcium and contractile responses to phorbol esters and the calcium channel agonist, Bay K 8644, in arteries from hypertensive rats.
Am J Hypertens
3:
245S-248S,
1990[Medline].
218.
Takemoto, M,
Takagi K,
Ogino K,
and
Tomita T.
Comparison of contractions produced by carbachol, thapsigargin and cyclopiazonic acid in the guinea-pig tracheal muscle.
Br J Pharmacol
124:
1449-1454,
1998[Abstract].
219.
Tamaoki, J,
Tagaya E,
Chiyotani A,
Yamawaki I,
and
Konno K.
Role of K+ channel opening and Na+-K+ ATPase activity in airway relaxation induced by salbutamol.
Life Sci
55:
L217-L223,
1994.
220.
Tanaka, H,
Jing L,
Takahashi S,
and
Ito Y.
The possible role of nitric oxide in relaxations and excitatory neuroeffector transmission in the cat airway.
J Physiol
493:
785-791,
1996[Abstract].
221.
Tang, D,
Mehta D,
and
Gunst SJ.
Mechanosensitive tyrosine phosphorylation of paxillin and focal adhesion kinase in tracheal smooth muscle.
Am J Physiol Cell Physiol
276:
C250-C258,
1999[Abstract/Free Full Text].
222.
Tang, DD,
and
Gunst SJ.
Depletion of focal adhesion kinase by antisense depresses contractile activation of smooth muscle.
Am J Physiol Cell Physiol
280:
C874-C883,
2001[Abstract/Free Full Text].
223.
Tang, DD,
and
Gunst SJ.
Selected contribution: roles of focal adhesion kinase and paxillin in the mechanosensitive regulation of myosin phosphorylation in smooth muscle.
J Appl Physiol
91:
1452-1459,
2001[Abstract/Free Full Text].
224.
Tao, L,
Huang Y,
and
Bourreau JP.
Control of the mode of excitation-contraction coupling by Ca2+ stores in bovine trachealis muscle.
Am J Physiol Lung Cell Mol Physiol
279:
L722-L732,
2000[Abstract/Free Full Text].
225.
Toescu, EC.
Temporal and spatial heterogeneities of Ca2+ signaling: mechanisms and physiological roles.
Am J Physiol Gastrointest Liver Physiol
269:
G173-G185,
1995[Abstract/Free Full Text].
226.
Togashi, H,
Emala CW,
Hall IP,
and
Hirshman CA.
Carbachol-induced actin reorganization involves Gi activation of Rho in human airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
274:
L803-L809,
1998[Abstract/Free Full Text].
227.
Tolloczko, B,
Jia YL,
and
Martin JG.
Serotonin-evoked calcium transients in airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
269:
L234-L240,
1995[Abstract/Free Full Text].
228.
Tolloczko, B,
Tao FC,
Zacour ME,
and
Martin JG.
Tyrosine kinase-dependent calcium signaling in airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
278:
L1138-L1145,
2000[Abstract/Free Full Text].
229.
Tompa, P,
Toth-Boconadi R,
and
Friedrich P.
Frequency decoding of fast calcium oscillations by calpain.
Cell Calcium
29:
161-170,
2001[ISI][Medline].
230.
Triggle, DJ.
Calcium, the control of smooth muscle function and bronchial hyperreactivity.
Allergy
38:
1-9,
1983[ISI][Medline].
231.
Vannier, C,
Croxton TL,
Farley LS,
and
Hirshman CA.
Inhibition of dihydropyridine-sensitive calcium entry in hypoxic relaxation of airway smooth muscle.
Am J Physiol Lung Cell Mol Physiol
268:
L201-L206,
1995[Abstract/Free Full Text].
232.
Vichi, P,
Whelchel A,
Knot H,
Nelson M,
Kolch W,
and
Posada J.
Endothelin-stimulated ERK activation in airway smooth muscle cells requires calcium influx and Raf activation.
Am J Respir Cell Mol Biol
20:
99-105,
1999[Abstract/Free Full Text].
233.
Wade, GR,
Barbera J,
and
Sims SM.
Cholinergic inhibition of Ca2+ current in guinea-pig gastric and tracheal smooth muscle cells.
J Physiol
491:
307-319,
1996[Abstract].
234.
Waldron, GJ,
Sigurdsson SB,
Aiello EA,
Halayko AJ,
Stephens NL,
and
Cole WC.
Delayed rectifier K+ current of dog bronchial myocytes: effect of pollen sensitization and PKC activation.
Am J Physiol Lung Cell Mol Physiol
275:
L336-L347,
1998[Abstract/Free Full Text].
235.
Wang, YX,
Fleischmann BK,
and
Kotlikoff MI.
M2 receptor activation of nonselective cation channels in smooth muscle cells: calcium and Gi/Go requirements.
Am J Physiol Cell Physiol
273:
C500-C508,
1997[Abstract/Free Full Text].
236.
Wang, YX,
Fleischmann BK,
and
Kotlikoff MI.
Modulation of maxi-K+ channels by voltage-dependent Ca2+ channels and methacholine in single airway myocytes.
Am J Physiol Cell Physiol
272:
C1151-C1159,
1997[Abstract/Free Full Text].
237.
Wang, YX,
and
Kotlikoff MI.
Inactivation of calcium-activated chloride channels in smooth muscle by calcium/calmodulin-dependent protein kinase.
Proc Natl Acad Sci USA
94:
14918-14923,
1997[Abstract/Free Full Text].
238.
Wang, YX,
and
Kotlikoff MI.
Muscarinic signaling pathway for calcium release and calcium-activated chloride current in smooth muscle.
Am J Physiol Cell Physiol
273:
C509-C519,
1997[Abstract/Free Full Text].
239.
Wang, YX,
and
Kotlikoff MI.
Signalling pathway for histamine activation of non-selective cation channels in equine tracheal myocytes.
J Physiol
523:
131-138,
2000[Abstract/Free Full Text].
240.
Wang, Z,
Pavalko FM,
and
Gunst SJ.
Tyrosine phosphorylation of the dense plaque protein paxillin is regulated during smooth muscle contraction.
Am J Physiol Cell Physiol
271:
C1594-C1602,
1996[Abstract/Free Full Text].
241.
Waniishi, Y,
Inoue R,
Morita H,
Teramoto N,
Abe K,
and
Ito Y.
Cyclic GMP-dependent but G kinase-independent inhibition of Ca2+-dependent Cl
currents by NO donors in cat tracheal smooth muscle.
J Physiol
511:
719-731,
1998[Abstract/Free Full Text].
242.
Welling, A,
Felbel J,
Peper K,
and
Hofmann F.
Beta-adrenergic receptor stimulates L-type calcium current in adult smooth muscle cells.
Blood Vessels
28:
154-158,
1991[ISI][Medline].
243.
Worley, JF,
and
Kotlikoff MI.
Dihydropyridine-sensitive single calcium channels in airway smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
259:
L468-L480,
1990[Abstract/Free Full Text].
244.
Wylam, ME,
Gungor N,
Mitchell RW,
and
Umans JG.
Eosinophils, major basic protein, and polycationic peptides augment bovine airway myocyte Ca2+ mobilization.
Am J Physiol Lung Cell Mol Physiol
274:
L997-L1005,
1998[Abstract/Free Full Text].
245.
Yamaguchi, H,
Kajita J,
and
Madison JM.
Isoproterenol increases peripheral [Ca2+]i and decreases inner [Ca2+]i in single airway smooth muscle cells.
Am J Physiol Cell Physiol
268:
C771-C779,
1995[Abstract/Free Full Text].
246.
Yamakage, M,
Hirshman CA,
and
Croxton TL.
Cholinergic regulation of voltage-dependent Ca2+ channels in porcine tracheal smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
269:
L776-L782,
1995[Abstract/Free Full Text].
247.
Yamakage, M,
Hirshman CA,
and
Croxton TL.
Volatile anesthetics inhibit voltage-dependent Ca2+ channels in porcine tracheal smooth muscle cells.
Am J Physiol Lung Cell Mol Physiol
268:
L187-L191,
1995[Abstract/Free Full Text].
248.
Yang, CM,
Chien CS,
Wang CC,
Hsu YM,
Chiu CT,
Lin CC,
Luo SF,
and
Hsiao LD.
Interleukin-1beta enhances bradykinin-induced phosphoinositide hydrolysis and Ca2+ mobilization in canine tracheal smooth muscle cells: involvement of the Ras/Raf/mitogen-activated protein kinase (MAPK) kinase (MEK)/MAPK pathway.
Biochem J
354:
439-446,
2001[ISI][Medline].
249.
Yang, CM,
Chou SP,
Wang YY,
Hsieh JT,
and
Ong R.
Muscarinic regulation of cytosolic free calcium in canine tracheal smooth muscle cells: Ca2+ requirement for phospholipase C activation.
Br J Pharmacol
110:
1239-1247,
1993[Abstract].
250.
Yang, CM,
Hsia HC,
Hsieh JT,
Ong R,
and
Luo SF.
Bradykinin-stimulated calcium mobilization in cultured canine tracheal smooth muscle cells.
Cell Calcium
16:
59-70,
1994[ISI][Medline].
251.
Yang, CM,
Hsieh JT,
Yo YL,
Ong R,
and
Tsao HL.
5-Hydroxytryptamine-stimulated calcium mobilization in cultured canine tracheal smooth muscle cells.
Cell Calcium
16:
194-204,
1994[ISI][Medline].
252.
Yao, Y,
Ferrer-Montiel AV,
Montal M,
and
Tsien RY.
Activation of store-operated Ca2+ current in Xenopus oocytes requires SNAP-25 but not a diffusible messenger.
Cell
98:
475-485,
1999[ISI][Medline].
253.
Yoshii, A,
Iizuka K,
Dobashi K,
Horie T,
Harada T,
Nakazawa T,
and
Mori M.
Relaxation of contracted rabbit tracheal and human bronchial smooth muscle by Y-27632 through inhibition of Ca2+ sensitization.
Am J Respir Cell Mol Biol
20:
1190-1200,
1999[Abstract/Free Full Text].
254.
ZhuGe, R,
Sims SM,
Tuft RA,
Fogarty KE,
and
Walsh JV.
Ca2+ sparks activate K+ and Cl
channels, resulting in spontaneous transient currents in guinea-pig tracheal myocytes.
J Physiol
513:
711-718,
1998[Abstract/Free Full Text].
Am J Physiol Lung Cell Mol Physiol 282(6):L1161-L1178
1040-0605/02 $5.00
Copyright © 2002 the American Physiological Society