ArticlePDF Available

Nuclear KATP channels trigger nuclear Ca2+ transients that modulate nuclear function

Authors:

Abstract and Figures

Glucose, the principal regulator of endocrine pancreas, has several effects on pancreatic beta cells, including the regulation of insulin release, cell proliferation, apoptosis, differentiation, and gene expression. Although the sequence of events linking glycemia with insulin release is well described, the mechanism whereby glucose regulates nuclear function is still largely unknown. Here, we have shown that an ATP-sensitive K(+) channel (K(ATP)) with similar properties to that found on the plasma membrane is also present on the nuclear envelope of pancreatic beta cells. In isolated nuclei, blockade of the K(ATP) channel with tolbutamide or diadenosine polyphosphates triggers nuclear Ca(2+) transients and induces phosphorylation of the transcription factor cAMP response element binding protein. In whole cells, fluorescence in situ hybridization revealed that these Ca(2+) signals may trigger c-myc expression. These results demonstrate a functional K(ATP) channel in nuclei linking glucose metabolism, nuclear Ca(2+) signals, and nuclear function.
Content may be subject to copyright.
Nuclear K
ATP
channels trigger nuclear Ca
2
transients
that modulate nuclear function
Ivan Quesada*, Juan M. Rovira, Franz Martin, Enrique Roche, Angel Nadal, and Bernat Soria
Institute of Bioengineering, Miguel Herna´ ndez University, San Juan Campus, 03550 Alicante, Spain
Edited by Ramon Latorre, Center for Scientific Studies, Valdivia, Chile, and approved May 22, 2002 (received for review January 23, 2002)
Glucose, the principal regulator of endocrine pancreas, has several
effects on pancreatic beta cells, including the regulation of insulin
release, cell proliferation, apoptosis, differentiation, and gene
expression. Although the sequence of events linking glycemia with
insulin release is well described, the mechanism whereby glucose
regulates nuclear function is still largely unknown. Here, we have
shown that an ATP-sensitive K
channel (K
ATP
) with similar prop-
erties to that found on the plasma membrane is also present on the
nuclear envelope of pancreatic beta cells. In isolated nuclei, block-
ade of the K
ATP
channel with tolbutamide or diadenosine polypho-
sphates triggers nuclear Ca
2
transients and induces phosphory-
lation of the transcription factor cAMP response element binding
protein. In whole cells, fluorescence in situ hybridization revealed
that these Ca
2
signals may trigger c-myc expression. These results
demonstrate a functional K
ATP
channel in nuclei linking glucose
metabolism, nuclear Ca
2
signals, and nuclear function.
Pancreatic beta cells play a critical role in maintaining a
steady-state level of glucose in the blood and tissues. In-
creased levels of glucose stimulate beta cells to secrete insulin
closing a well established feedback loop. Malfunction of beta
cells causes the widespread pathology, diabetes mellitus. The
signal transduction mechanism leading to insulin release involves
the closure of plasma membrane K
ATP
channels as a result of
glucose metabolism by increasing both the intracellular ATP
ADP ratio and diadenosine polyphosphates (DPs; refs. 1 and 2).
Channel closure leads to membrane depolarization and the
opening of voltage-activated Ca
2
channels (3). The subsequent
cytosolic Ca
2
signal, which is oscillatory (4), triggers a pulsatile
insulin secretion. In most cells, a single second messenger as
Ca
2
is able to provoke different responses depending on its
route of entry, its localization, and a code of amplitude or
frequency of Ca
2
oscillations (4, 5). In pancreatic beta cells,
Ca
2
mediates not only insulin secretion but also a broad range
of other processes such as gene expression (6, 7). Although it is
well established that the nucleoplasmic concentration of free
Ca
2
regulates nuclear function (5), the mechanism whereby
nuclear Ca
2
signals are generated is still unclear. Here, we
report confocal measurements of nuclear Ca
2
concentration
([Ca
2
]
n
) in intact beta cells exposed to glucose. Experiments in
isolated nuclei revealed a K
ATP
channel present on the nuclear
envelope whose blockade results in a [Ca
2
]
n
rise. This increased
[Ca
2
]
n
induces phosphorylation of the transcription factor
cAMP response element binding protein (CREB). We further
demonstrate that [Ca
2
]
n
elevation may result in c-myc expres-
sion in whole cells.
Materials and Methods
Cell Isolation, Culture, and Permeabilization. Islets from adult (8–10
weeks old) Swiss albino male mice (OF1) killed by cer vical
dislocation were isolated and then dispersed into single cells
after a published procedure (3). Isolated cells were cultured in
RPMI medium 1640 for 24 h. Cell permeabilization was per-
formed as described (8).
Spot Confocal Microscopy. Cells were loaded with 2
M Calcium
Green-1AM (Molecular Probes) for 60 min at room temper-
ature (RT) and then were perfused in a modified Krebs–Ringer
buffer [119 mM NaCl4.7 mM KCl1.2 mM MgSO
4
1.2 mM
KH
2
PO
4
25 mM NaHCO
3
2.5 mM CaCl
2
3 mM glucose bub-
bled constantly with a mixture of 95% O
2
5% CO
2
(pH 7.4)].
Ca
2
was measured by using spot confocal microscopy, which
excels in measuring minute Ca
2
transients because of its high
signal-to-noise ratio. The spot illumination-detection configura-
tion has been described (9, 10). Briefly, a laser-illuminated
pinhole (10
m) was focused onto a spot through the objective
on a nucleus equatorial plane. Thus, the predicted detection
volume is about 0.6 0.6 1.1
m
3
. Changes in fluorescence
in this nuclear volume were precisely detected with a photodiode
(HR008; United Detector Technology) which was connected to
an Axopatch-200A amplifier (50 Gfeedback; Axon Instru-
ments, Foster City, CA). The laser illumination time was 30 ms.
Ca
2
-induced fluorescence intensity ratio was plotted as a
function of time (F
t
F
0
).
Nuclei Isolation. While dispersed, islet cells were suspended in a
buffer that mimicked the intracellular medium (125 mM KCl2
mM K
2
P0
4
40 mM Hepes0.1 mM MgCl
2
,pH7.2100 nM Ca
2
,
with 10.2 mM EGTA and 1.65 mM CaCl
2
). The cell membrane
and cytoskeleton were disrupted by brief sonication. Single
isolated nuclei were separated by centrifugation, resuspended in
the intracellular medium, and then allowed to attach onto glass
chambers (11). Ethidium homodimer-1 labeling (Molecular
Probes) indicated that the suspension was 90% enriched in
nuclei. This probe also was applied to identify nuclei after each
electrophysiological and imaging experiment. An antibody to
calnexin allowed an assessment of contamination of the nuclei
preparation with endoplasmic reticulum (ER) fragments (12).
We further performed confocal images and three-dimensional
(3D) reconstruction of nuclei labeled with rhodamine B hexyl
ester to detect the existence of some debris attached to the
surface of some nuclei (10%), which was associated with ER
contamination (13, 14). As this ER debris was completely
detectable by transmitted light or fluorescence microscopy, we
only performed experiments in intact and clean nuclei (free of
debris).
Glibenclamide-Dipyrrometheneboron Difluoride (BODIPY) Staining.
Cultured isolated islet cells were labeled with 40 nM of green-
fluorescent glibenclamide-BODIPY-FL (30 min at 4°C; Molec-
ular Probes) and stained with anti-insulin antibodies (15). Im-
munofluorescence for insulin revealed more than 80% of
pancreatic beta cells in the cultures. Isolated nuclei also were
This paper was submitted directly (Track II) to the PNAS office.
Abbreviations: KATP, ATP-sensitive Kchannel; [Ca2]n, nuclear Ca2concentration; CREB,
cAMP response element binding protein; RT, room temperature; BODIPY, dipyr-
rometheneboron difluoride; nKATP, nuclear KATP channel; ER, endoplasmic reticulum;
ERSR, endo-sarcoplasmic reticulum; RyR, ryanodine receptors.
*Present address: Department of Bioengineering, University of Washington, Seattle, WA
98195.
To whom reprint requests should be addressed at: Institute of Bioengineering, Miguel
Herna´ ndez University, San Juan Campus, Carretera Alicante-Valencia Km 87, 03550 Ali-
cante, Spain. E-mail: bernat.soria@umh.es.
9544–9549
PNAS
July 9, 2002
vol. 99
no. 14 www.pnas.orgcgidoi10.1073pnas.142039299
stained with glibenclamide-BODIPY-FL following the same
protocol. Nuclear envelope was identified with 1
M of rhoda-
mine B hexyl ester (5 min at RT; Molecular Probes; ref. 14);
meanwhile, 1
M of ethidium homodimer-1 (5 min at RT;
Molecular Probes) stained the DNA containing nucleoplasm.
Fluorescence was visualized by using a Zeiss LSM 510 confocal
microscope (63objective, 1.25 N.A.) and 12
m optical slices.
Intensity values were obtained by calculating the average bright-
ness value on the corresponding ring-like staining of the nuclei
and measured on an arbitrary gray scale from 0 (blackest) to 255
(whitest).
Patch-Clamp Experiments. Single-channel currents were recorded
from nuclei positively labeled with ethidium homodimer-1 by
using standard patch-clamp recording procedures (16). The bath
solution contained 140 mM KCl, 1 mM MgCl
2
, 10 mM Hepes,
1 mM EGTA, pH 7.2, and the pipette solution contained 5 mM
KCl, 135 mM NaCl, 10 mM Hepes, 1.1 mM MgCl
2
, pH 7.4.
Pipette potential was held at 0 mV throughout the record.
Experiments were performed at RT. High resistance seals were
formed (15G), indicating minimal contamination by the
endoplasmic reticulum (ER) membrane (13). Under these cir-
cumstances, it has been proposed that nuclear pores would be
occluded or nonconducting because of experimental conditions
or lack of cytosolic factors (13, 14, 17).
Ca
2
Measurements in Isolated Nuclei. Single isolated nuclei were
loaded as described (11, 18). The membrane impermeant Ca
2
probe Calcium Green-1 dextran (30
gml; 30 min at 4°C;
Molecular Probes) loaded the nucleoplasm while the membrane
permeant Ca
2
probe Fluo-3AM (20
M; 60 min at 4°C)
loaded the nuclear envelope. After loading, the nuclei were
washed twice with the intracellular medium and then were
equilibrated in the same medium supplemented with 1
Mof
ATP and 300 nM Ca
2
for a few minutes to load nuclei with Ca
2
(11, 18). After that, they were washed twice again with the
intracellular buffer (without AT P and Ca
2
). Experiments were
done at RT. No probe leakage was detected during the exper-
iment. Ca
2
imaging in single isolated nuclei was performed by
using a Zeiss LSM 510 confocal microscope with a Zeiss 63X oil
immersion lens, N.A. 1.25. Images were collected at 3 s intervals,
and fluorescence was measured by using the Zeiss LSM software
package. Ca
2
-induced fluorescence intensity ratio (F
t
F
0
) was
plotted as a function of time.
Nuclear Transmembrane Potential (⌬⌿
n
) Measurements. Isolated
nuclei were equilibrated in intracellular medium containing the
ER marker DiOC
6
(3) (200 nM, RT; ref. 19). After 5 min, the
nuclear envelope became stained. DiOC
6
(3) is a fluorescent
cationic lipophyllic dye whose incorporation into lumen is pro-
portional to ⌬⌿.
P-CREB Immunofluorescence. CREB phosphorylation was induced
in isolated nuclei by increasing Ca
2
in an EGTA-buffered
intracellular medium from 65 nM to 1
MCa
2
(20) or by
addition of tolbutamide or AP
4
A for 2 min. After 10 min, nuclei
were fixed with 0.1% paraformaldehyde (wtvol) and perme-
abilized with 0.1% Triton X-100 for 10 min. Nuclei were
preincubated with blocking buffer; then, anti-CREB phospho-
specific rabbit antibodies were applied for 16 h at 4°C (1:200,
Calbiochem). P-CREB was visualized with fluorescein-
conjugated secondary antibodies (1 h, RT, 1:64; Sigma). High
Ca
2
gave the maximum response.
Fluorescence
in Situ
Hybridization. Fragments (237 and 530 bp) of
c-myc and
-actin cDNAs cloned in pBSSK (Stratagene) were
used as templates for digoxigenin-labeled double-stranded DNA
probes, synthesized by PCR using specific primers (21, 22).
Isolated islet cells were cultured at least for 24 h in a culture
medium containing 3 mM of glucose and then were placed in
different stimulating conditions in a KrebsRinger buffer for 10
min. After that, cells were equilibrated in 3 mM of glucose for
45 min. Only when AP
4
A was used as a stimulator, cells were
permeabilized as mentioned above. Then, cells were fixed in a
solution containing 4% (volvol) formaldehyde, 5% (volvol)
acetic acid, and 0.9% NaCl and permeabilized with 1:1,000
Triton X-100 for 10 min. Hybridization was performed under
standard conditions and was revealed by immunof luorescence
detection with f luorescein-conjugated antidigoxigenin (1:500;
Roche, Barcelona, Spain). Nonspecific binding was reduced with
a commercial blocking reagent (Roche, Barcelona, Spain). Cells
were counterstained with 1
M ethidium homodimer for 5 min
before visualization under a Zeiss LSM 510 confocal microscope
(10objective, 0.45 N.A.). The average intensity value from
each stained cell was calculated and expressed on an arbitrary
gray scale from 0 (blackest) to 255 (whitest). Cells whose
fluorescence intensity was above the range of values of unstimu-
lated control cells were scored as activated cells for gene
expression.
Statistical Analysis. Statistical analysis was performed by using
SIGMAPLOT (Jandel, San Rafael, CA). Values are mean SE.
Except where indicated, P0.05 by Studentsttest.
Results
Nuclear Ca
2
Changes. Ca
2
changes were measured in the nuclear
space of isolated pancreatic beta cells by spot confocal micros-
copy (Fig. 1A; refs. 9 and 10). Cells equilibrated in an extracel-
lular medium containing 16.7 mM glucose exhibited a transient
increase of [Ca
2
]
n
(Fig. 1B). Exposure to the sulfonylurea
tolbutamide (20
M), which directly closes K
ATP
channels,
produced a similar increase (Fig. 1B). Both increases of [Ca
2
]
n
were still produced in the absence of extracellular Ca
2
(Fig. 1C).
Increased [Ca
2
]
n
was probably due to direct Ca
2
release from
the nuclear envelope to the nuclear space (11, 14, 23), because
tolbutamide induces [Ca
2
]
n
changes not only in the absence of
extracellular Ca
2
but in digitonin permeabilized cells (8) with
a buffered Ca
2
concentration (Fig. 1D). The tolbutamide-
induced Ca
2
increase was counteracted by the sulfonamide
Fig. 1. Ca2changes in the nuclear space in intact cells revealed by spot
confocal microscopy. (A) Hybrid phase contrast image with a uorescent spot
(see arrow) focused on an intact cell nucleus loaded with Calcium-Green-1
AM. (B) Intact cells were stimulated with different agents at the time indicated
by the line. Values are mean SE and were pooled from six cells [16.7 mM
glucose (G)] and seven cells [20
M tolbutamide (T)]. (C) Beta cells were
stimulated with 16.7 mM G (n5) and 20
MT(n6) in a Ca2-buffered
medium containing 5 mM EGTA. (D) Permeabilized cells were stimulated with
100
M tolbutamide or 100
M tolbutamide plus 100
M diazoxide (n4)
inaCa
2-buffered intracellular medium. G, glucose; T, tolbutamide; D,
diazoxide.
Quesada et al. PNAS
July 9, 2002
vol. 99
no. 14
9545
PHYSIOLOGY
diazoxide, a K
ATP
channel opener (Fig. 1D). These results
strongly suggest that intracellular K
ATP
channels responsible for
eliciting nuclear Ca
2
signals might be present close to the beta
cell nucleus.
Verification of K
ATP
Channels on Nuclei. To verify the existence of a
K
ATP
channel on the nuclear envelope from mouse beta cells, we
used a high-specificity, high-affinity (K
i
s4 nM) binding assay
of glibenclamide-BODIPY-FL to the sulfonylurea receptor
(SUR1; ref. 14). SUR1 is the molecular complement of the
potassium inward rectifier (K
ir
; refs. 24 and 25). The association
of SUR1 and K
ir
6.2 forms the K
ATP
channel in beta cells. Fig. 2
ACshows a confocal optical section of pancreatic beta cells with
an intracellular ring-like labeling, suggesting the binding of
glibenclamide-BODIPY to the nuclear envelope. Similar results
were observed in isolated nuclei (Fig. 2 DF). The gliben-
clamide-BODIPY binding site colocalized with rhodamine B
hexyl ester, a marker of the nuclear membrane (ref. 14; Fig. 2E).
The binding was specific because it was displaced by the non-
fluorescent sulfonylurea tolbutamide in a dose-dependent man-
ner (Fig. 2G). This evidence agrees with previous observations
reporting an intracellular location of K
ATP
channels including
perinuclear sites (2631) in addition to plasma membrane K
ATP
channels.
Excised patch-clamp recordings (16) from the nuclear enve-
lope of isolated nuclei in a 140 mM K
5mMK
solution
exhibited K
-channel activity with conductance of approxi-
mately 25 pS (ref. 25; Fig. 2I). This channel was activated by ADP
and inhibited by ATP in a concentration-dependent manner
(Fig. 2I). Nuclear K
ATP
(nK
ATP
) channels display kinetic prop-
erties similar to channels found on the beta cell plasma mem-
brane, with openings grouped in bursts separated by long closing
periods (data not shown). The pharmacological profile of the
nK
ATP
channel also resembles that of the plasma membrane
K
ATP
channel (25). At concentrations similar to those acting on
the plasma membrane K
ATP
channel, tolbutamide (100
M)
blocks the nK
ATP
channel whereas diazoxide (200
M) opens it
(Fig. 2Iiv) in the presence of ATP. Therefore, this nK
ATP
channel seems remarkably similar to the plasma membrane K
ATP
channel (25).
Blockade of nK
ATP
Channels Elicits Ca
2
Signals in Isolated Nuclei.
Depending upon their lipophylicity, the Ca
2
-sensitive fluores-
cence dyes Fluo-3AM or Calcium Green-1 dextran localized
preferably in the nuclear envelope or the nucleoplasm (11, 14,
18), respectively (Fig. 3 Aand B). In isolated nuclei, confocal
microscopy revealed that tolbutamide induced opposite Ca
2
changes in the nuclear envelope and the nucleoplasm (Fig. 3 A
and B). Either 100
Mor500
M of tolbutamide generated
Ca
2
increases in the nucleoplasm, which paralleled Ca
2
de-
creases observed in the nuclear envelope (Fig. 3 Aand B).
Tolbutamide did not produce any Ca
2
transient in the presence
of the K
ATP
channel opener diazoxide (data not shown). More-
over, the diadenosine polyphosphate AP
4
A (100
M), which
blocks the K
ATP
channel (2), induced a similar Ca
2
release (Fig.
3C). Because AP
4
A is membrane-impermeant, this experiment
also suggests that the regulatory site of this K
channel does not
face the perinuclear space. Thus, the nuclear envelope may act
asaCa
2
reservoir that is mobilized as a result of glucose
metabolism.
All of these results suggested that this Ca
2
pool was sensitive
to changes in the nuclear membrane K
permeability but also
raised the possibility that Ca
2
release may be elicited by means
of a voltage-dependent mechanism. To explore the hypothesis
that blockade of nK
ATP
channels provokes a Ca
2
release from
the nuclear envelope by voltage variations, we changed the
concentrations of extraluminal K
. Reduction of the K
con-
centration (by N-methyl-D-glucamine replacement) and the en-
Fig. 2. KATP channels on the nuclear envelope. (A) Intact islet cells
incubated with glibenclamide-BODIPY-FL showed a strong ring-like label-
ing around the nucleus. (B) Cell in Aidentied as beta cell by immunou-
orescence against insulin. (C) Colocalization of images Aand B. Optical
sections range between 12
m. (Bar 10
m.) Similar results were
obtained in 19 cells from four coverslips from three different cultures. (D)
Glibenclamide-BODIPY-FL ring-like staining of an isolated nucleus. (E)
Nucleus in Dwith rhodamine B hexyl ester staining the nuclear envelope.
(F) Colocalization of images Dand E(n21). (Bar 5
m.) Confocal images
were taken with 1
m optical slices. (G) Glibenclamide-BODIPY competi-
tion with nonuorescent unlabeled tolbutamide reected a displacement
of the binding site at the nuclear envelope. Glibenclamide labeling is
expressed as the percentage of uorescence intensity with respect to the
control condition (0
M tolbutamide). Results (mean SE) were pooled
from 94 nuclei for 0
M tolbutamide, 29 for 1
M, 20 for 10
M, and 24 for
1.000
M from 12 coverslips of 6 different nuclei preparations. (H) Image
showing a patch pipette on an isolated nucleus previously identied using
ethidium homodimer-1. (I) Single-channel records (ltered at 1 KHz) were
obtained from a nucleus membrane excised patch (n5). The pipette
potential was held at 0 mV. Upward currents represent currents going into
the pipette. Single-channel current gives an estimate of 25 pS. A dashed
line represents changes in the external solution. (i) ADP (200
M) increased
the burst length. The change from 200
M ADP to 2 mM ATP rapidly and
completely closed Kchannels. (ii) The Kchannel activity absent in 2 mM
ATP was rapidly restored when ATP was removed. (iii) The change from 40
Mto20
M ATP increased channel activity. (iv) Diazoxide (200
M)
activates the channel in the presence of ATP, whereas tolbutamide (100
M) blocks it.
9546
www.pnas.orgcgidoi10.1073pnas.142039299 Quesada et al.
suing change in the K
electrochemical gradient led to a
corresponding Ca
2
discharge (Fig. 3D). Tolbutamide failed to
produce a Ca
2
transient in nuclei pretreated with 10
Mofthe
K
ionophore valinomycin, further suggesting that the collapse
of K
electrochemical gradient can suppress Ca
2
release (data
not shown). We monitored changes of nuclear transmembrane
potential (⌬⌿
n
) by using DiOC
6
(3), which has been validated as
a potentiometric probe in mitochondria (ref. 19; see Materials
and Methods). DiOC
6
(3) accumulated in the nuclear envelope of
beta cells. Tolbutamide elicited a consistent increase in
DiOC
6
(3) fluorescence, indicating that the perinuclear lumen
became more negative (Fig. 3E). These results pointed to the
existence of a ⌬⌿
n
, which may change as a result of a decrease
in K
permeability. In beta cells, a [Ca
2
]
n
pathway sensitive to
this ⌬⌿
n
may exist.
Several intracellular channels found in the endo-sarcoplasmic
reticulum (ERSR) and nucleus, including inositol-1,4,5-
trisphosphate receptors (InsP
3
-R) and ryanodine receptors
(RyR; refs. 11, 14, and 32) among others, have been shown to be
voltage-sensitive. Ca
2
release induced by tolbutamide de-
creased 81.3 9.3% in isolated nuclei exposed to 10
M
ruthenium red, a RyR blocker (Fig. 3F). Higher ruthenium red
concentrations (100
M) completely blocked the Ca
2
release.
Similar results were observed when isolated nuclei were prein-
cubated with anti-RyR antibodies (1:100, 30 min; data not
shown). Conversely, heparin, a blocker of the InsP
3
-R (15) did
not produce a significant effect (data not shown).
Nuclear Function of nK
ATP
Channels. Several roles have been sug-
gested for nuclear channels, yet very few studies relate these
channels to nuclear functions. As has recently been proposed,
nuclear Ca
2
signals may target the CREB protein, whose
phosphorylation may activate transcription (20, 33). By using
immunofluorescence in isolated functional nuclei (20), we have
observed CREB phosphorylation (Fig. 4) induced by tolbut-
amide and AP
4
A, both K
ATP
blockers that provoked an increase
of [Ca
2
]
n
(Fig. 3). As shown in Fig. 1D, the K
ATP
channel opener
diazoxide counteracted the tolbutamide effects (Fig. 4D). These
observations further suggested that the closure of nK
ATP
chan-
nels initiate the transduction of nuclear signals.
In addition to transcription factors, other specific intranuclear
Ca
2
targets may exist that regulate specific nuclear events, such
as gene induction. These targets might include nuclear kinases,
polymerases and chromatin remodeling, among others (34). To
evaluate the relationship between these nuclear Ca
2
signals and
gene expression, we conducted f luorescence in situ hybridization
experiments in intact beta cells. C-myc expression was chosen as
a model because this immediate early gene is activated by
glucose in a Ca
2
-dependent manner in both cultured cell lines
(21; E.R., unpublished work) and in islet cells (35). Fig. 5 shows
c-myc expression in islet cells under different stimuli. Remark-
ably, tolbutamide and AP
4
A activated gene expression (30 and
20%, respectively) even in the absence of extracellular Ca
2
.In
this condition, both K
ATP
blockers induced Ca
2
changes within
the nuclear space (Figs. 1Cand 3 AC). Diazoxide abolished
these effects when present (data not shown).
Discussion
Exposure of beta cells to glucose results in a Ca
2
-mediated
activation of a broad range of functions (36). However, the link
between glucose-induced Ca
2
signals and specific organelles
remains obscure. Here, we focused on the mechanism of glucose-
regulated [Ca
2
]
n
fluctuations and the effect of these Ca
2
signals on nuclear function. Exposure of intact cells to glucose
and tolbutamide led to a [Ca
2
]
n
rise both in cells equilibrated
in normal KrebsRinger containing 2.5 mM Ca
2
and in Ca
2
-
free KrebsRinger solution (Fig. 1 Band C). These results
suggest that both secretagogues must be acting on intracellular
Ca
2
compartments and are in agreement with previous obser-
vations in beta cells (36) showing that tolbutamide can increase
cytosolic [Ca
2
] in the absence of extracellular Ca
2
. Mitochon-
Fig. 3. Nuclear Ca2signals induced by KATP channel blockade. (A) Nuclei
were loaded with the membrane permeant probe Fluo-3AM, which was
preferentially accumulated in the nuclear envelope (ref. 11; see image).
[Ca2]nwas measured by using confocal microscopy. Tolbutamide applied to
Fluo-3-loaded nuclei produced a Ca2decrease (n3). (B) Calcium Green-1
dextran, a nonpermeant probe, was distributed uniformly in the nucleoplasm
(see image). Tolbutamide provoked a transient Ca2increase in the nucleo-
plasm (n6). Spot confocal methods also were used to observe this tolbut-
amide-induced nuclear Ca2release (n3; data not shown). (Bar 5
m.) (C)
Same experiment described in Bshowing a Ca2release induced by 100
M
AP4A(n3). (D) Replacement of 100 mM Kby NMG(N-methyl-D-
glucamine) led to a similar Ca2transient (n4) in the nucleoplasm. (E)
Nuclear envelope was loaded with DiOC6(3), a voltage-sensitive probe (see
Materials and Methods). Fluorescence was increased upon addition of tolbu-
tamide (n4). (F) Ruthenium red (RR) blocked almost 90% of the tolbut-
amide-induced Ca2release (n4). Two consecutive Ca2transients were
induced in this experiment. Before the second discharge, nuclei were loaded
with Ca2(see Materials and Methods) and then treated with ruthenium red.
When the blocker was not present, the two transients had the same ampli-
tude. Thus, the rst transient was used as a control (C).
Fig. 4. CREB phosphorylation in isolated nuclei. Immunouorescence de-
tection of P-CREB in nuclei treated with different stimuli. (A)Ca
2(1
M). (B)
Tolbutamide (100
M). (C)Ca
2(65 nM). Immunodetection is shown in green,
and ethidium homodimer-1 staining is shown in red. (D) Percentage of labeled
nuclei relative to the maximal response 1
MCa
2(HC; n145) measured in
the following conditions: 100
M tolbutamide (T; n112); 100
M tolbut-
amide plus 200
M diazoxide (TD; n92); 100
MAP
4A(n56); and 65 nM
Ca2(LC; n134).
Quesada et al. PNAS
July 9, 2002
vol. 99
no. 14
9547
PHYSIOLOGY
dria have been suggested as an intracellular target of sulfonyl-
ureas in beta cells (36). However, the intracellular source of this
sulfonylurea-induced Ca
2
release was not univocally identified.
Although Ca
2
discharge from an organelle could diffuse to the
nuclear space, our observations that tolbutamide can increase
[Ca
2
]
n
as well in permeabilized cells equilibrated in Ca
2
-
buffered intracellular medium (Fig. 1D), indicate that in this case
the source of Ca
2
may be located very close to the nucleus. This
tolbutamide effect was readily blocked by diazoxide, a K
ATP
channel opener (Fig. 1D). This observation indicates that K
ATP
channels may be present near the nucleus and their closure may
control [Ca
2
]
n
in beta cells. Excised patches of nuclear mem-
brane confirmed the presence of a nuclear nK
ATP
channel that
exhibits kinetics and pharmacological properties very similar to
the K
ATP
channels found in the plasma membrane (ref. 25; Fig.
2I). Patch scission can provoke membrane remodeling, making
it difficult to know the channel orientation by using the patch-
clamp approach. However, the effect of the membrane-
impermeant K
ATP
channel blocker AP
4
A (2) on Ca
2
release in
isolated nuclei (Fig. 3D) suggests that the regulatory site of this
channel is not facing the perinuclear space.
In addition to this functional characterization we have dem-
onstrated here a specific binding site for sulfonylureas at the beta
cell nuclear envelope (Fig. 2) that was revealed by glibenclamide-
BODIPY-FL binding. This method has been successfully used to
detect the presence of K
ATP
channels in different kinds of cells
including beta cells, neurons, and monocytes (15). Our results
are in agreement with previous work in cells transfected with
green fluorescent protein-tagged constructs linked to SUR1 or
Kir6.2 that showed fluorescence patterns not only in plasma but
in perinuclear membranes, as well as in ER and Golgi (30, 31).
Thus, three independent lines of evidence point to the presence
ofaK
ATP
channel in the nuclear envelope, broadening the idea
that together with the well characterized plasma membrane
K
ATP
channels (24, 25), sulfonylureas might have multiple sites
of action (26), including mitochondria (27),secretory granules
(28, 29), and, as shown here, the nuclear envelope.
Our results show that nK
ATP
channels are involved in beta cell
nuclear Ca
2
signaling because specific K
ATP
blockers such as
tolbutamide (25) and AP
4
A (2) induced Ca
2
release from the
nuclear envelope to the nucleoplasm in isolated nuclei (Fig. 3
AC). Several lines of evidence suggest that the nuclear Ca
2
pathway revealed here depends on a change in K
permeability
and ⌬⌿
n
in beta cell nuclei. First, valinomycin abolished tolbu-
tamide-induced Ca
2
transients. Second, a change in K
elec-
trochemical gradient by NMG
replacement in the extraluminal
medium produced a similar Ca
2
discharge (Fig. 3D). Third,
tolbutamide provoked a transient increase in ⌬⌿
n
(Fig. 3E)as
revealed the voltage-sensor DiOC
6
(3) (19, 37). It is plausible that
a potential difference between perinuclear space and cytosol
exists because the two membranes of the nuclear envelope
surround a lumen that separates different subcellular compart-
ments and contains a diversity of channels with different per-
meabilities (23, 34). The nuclear envelope is a structural exten-
sion of the ERSR network. Thus, similar properties and
molecular composition are shared among these organelles. The
existence of a low resting transmembrane potential described in
the SR (38) is compatible with transient voltage increases during
stimulation similar to those reported here.
Although the understanding of the molecular components of
the nuclear envelope and their regulation have undergone
important progress, their involvement in nuclear Ca
2
signaling
still remains a challenge for exploration (23, 34). A variety of
ERSR and nuclear channels that may contribute to regulate
[Ca
2
]
n
and other ion species have been reported (13, 32,
3944). An important feature is that several ERSR and nuclear
channels are voltage-sensitive. Those include Cl
(13), K
(42),
RyR (32), InsP
3
-R (14), and InsP
3
-insensitive Ca
2
channels (43,
44). Our results are consistent with RyR channels at the nuclear
envelopeas observed in exocrine cells (11) and oocytes (45)
that may control the nuclear Ca
2
release reported here, because
their blockade by ruthenium red or anti-RyR antibodies sup-
pressed the tolbutamide-induced nuclear Ca
2
transients (Fig.
3F). It is noteworthy that RyR channels undergo an active state
of elevated open probability upon changes of voltage (32). Thus,
it is very likely that RyR channels may be affected by tolbut-
amide-induced nuclear voltage changes (Fig. 3E). Their activa-
tion may lead to Ca
2
-release, which may be amplified by a
Ca
2
-induced Ca
2
-release process, as has been observed in ER
of pancreatic beta cells (46).
Our results show that this tolbutamide-induced [Ca
2
]
n
con-
ductance seems to be associated with a change in ⌬⌿
n
as a result
of a decrease in K
permeability (Fig. 3 Dand E). This process
is not exclusive to beta cells. In fact, there is a body of evidence
that establishes a link between K
and Ca
2
conductances in the
ERSR. For instance, Ca
2
release from SR has been induced
not only by K
channel general blockers (47) but also by the
specific K
ATP
channel blocker glibenclamide (48), suggesting the
presence of this K
channel type and its role in intracellular Ca
2
release. Moreover, lowering the extraluminal K
produced a
Ca
2
discharge from SR microsomes of myocytes, suggesting a
voltage change that is not explained well by current ideas about
EC coupling in muscle cells (49). Furthermore, a functional
K
ATP
channel has been proposed to control the K
permeability
and the granular membrane potential in glucagon-containing
secretory granules (29). The presence of nuclear Ca
2
channels
sensitive to changes in transmembrane potential explains
tolbutamide-induced Ca
2
release in beta cell nuclei. None-
theless, we do not exclude the involvement of other voltage-
sensitive pathways that may direct or indirectly trigger the
nuclear Ca
2
release reported here. These possible pathways
may include ion exchange fluxes coupled to Ca
2
(50, 51) or
other mechanisms such as the ER Ca
2
leak pathway (52). In
fact, a remarkable characteristic is that the ER leak pathway is
increased with high concentrations of AT P, a blocker of the
Fig. 5. Gene expression induced by KATP channel blockers in the absence of
extracellular Ca2. Fluorescence in situ hybridization and transmitted light
images of islet cells stimulated in different conditions. (A) Glucose (16 mM). (B)
Tolbutamide (100
M) in the absence of extracellular Ca2(buffered with
EGTA). (C) Glucose (3 mM) as a control condition. The green color shows c-myc
expression of beta cells counterstained with ethidium homodimer (red).
(Bar 10
m.) (D) Percentage of activated cells relative to unstimulated
control cells (3 mM glucose; n360) in different conditions: 40 mM potassium
(K;n290); 16 mM glucose (G; n369); 100
M tolbutamide (T; n329);
100
M tolbutamide in the absence of extracellular Ca2[T(0);n466]; 100
M
AP4A in the absence of extracellular Ca2(n105); and absence of extracel-
lular Ca2[Ca2(0);n211]. Results were pooled from three independent
experiments.
-actin constitutive expression was used as an invariable control
under the same conditions.
9548
www.pnas.orgcgidoi10.1073pnas.142039299 Quesada et al.
K
ATP
channel (52). In summary, we suggest that blockade of
K
ATP
channels at the nuclear envelope may elicit a transient
transmembrane potential rise as a result of K
permeability
decrease, which may trigger a voltage-sensitive Ca
2
discharge,
mainly through RyR channels.
By using immunofluorescence, we have proved that the nu-
clear Ca
2
transients reported here increased the phosphoryla-
tion rate of the transcription factor CREB (Fig. 4) in isolated
nuclei, in agreement with similar results described in hippocam-
pal neuron nuclei (20). Our working model only allows Ca
2
mobilization from the nucleus and unambiguously involves
nK
ATP
channels in pancreatic beta cell nuclear function.
In a whole-cell model, we evaluated the possibility of other
nuclear processes such as gene expression being affected. Tol-
butamide and AP
4
A, which induced Ca
2
release in isolated
nuclei (Fig. 3), triggered c-myc expression in the absence of
extracellular Ca
2
(Fig. 5). In the whole cell, we cannot rule out
the possibility that other Ca
2
release in the cytosol may drive
gene expression in our conditions. However, as shown here, the
nucleus probably accounts for an important part of these intra-
cellular Ca
2
signals (Figs. 1 and 3). Be that as it may, we have
demonstrated that these nuclear Ca
2
signals are sufficient to
drive CREB phosphorylation in isolated nuclei in these cells.
Hence, Ca
2
signals generated at the nuclear envelope have an
important role for nuclear function in beta cells.
We propose that signals generated by glucose metabolism not
only inhibit the plasma membrane K
ATP
channel, initiating
insulin release, but also interact with nK
ATP
channels, triggering
nuclear Ca
2
signals that modulate nuclear functions such as
phosphorylation of the transcription factor CREB and likely
gene expression. Our data shows a new signal-transduction
pathway linking nuclear K
ion channels to fluctuations of
[Ca
2
]
n
and nuclear function and also contributes to an under-
standing of the mechanism of action of sulfonylureas, drugs that
are broadly used in the clinic for the treatment of the diabetic
patient.
We thank P. Verdugo and D. Willows for the critical review of the
manuscript. This work was supported in part by Grant PM99-0142 from
Secretarı´a de Estado de Universidades e Investigacio´n, Fundacio´ Marato´
TV3 Grant 99-1210, Fundacio´n Salud 2000, Juvenile Diabetes Foun-
dation Grant 1-2000-575, and Generalitat Valenciana Grant GV99-
139-1-04.
1. Prentki, M. & Matchinsky, F. M. (1987) Physiol. Rev. 67, 1185–1248.
2. Martin, F., Pintor, J., Rovira, J. M., Miras-Portugal, M. C. & Soria, B. (1998)
FASEB J. 12, 1499–1506.
3. Valdeolmillos, M., Nadal, A., Contreras, D. & Soria, B. (1992) J. Physiol. 455,
173–186.
4. Santos, R. M., Rosario, L. M., Nadal, A., Garcia-Sancho, J., Soria, B. &
Valdeolmillos, M. (1991) Pflu¨gers Arch. 418, 417– 422.
5. Berridge, M. J., Bootman, M. D. & Lipp, P. (1997) Nature (London) 386,
759–760.
6. Roche, E. & Prentki, M. (1994) Cell Calcium 16, 331–338.
7. Susini, S., Roche, E., Prentki, M. & Schlegel, W. (1997) FASEB J. 12,
1173–1182.
8. Martin, F., Moya, F., Gutierrez, L. M., Reig, J. A. & Soria, B. (1995)
Diabetologia 38, 860– 863.
9. Quesada, I., Martin, F. & Soria, B. (2000) J. Physiol. 525, 159 –167.
10. Escobar, A. L., Monck, J. R., Fernandez, J. M. & Vergara, J. L. (1994) Nature
(London) 367, 739–741.
11. Gerasimenko, O. V., Gerasimenko, J. V., Belan, P. V. & Petersen, O. H. (1995)
Cell 80, 439– 444.
12. Radominska-Pandya, A., Pokrovskaya, I. D., Xu, J., Little., J. M., Jude, A. R.,
Kurten, R. C. & Czernik, P. J. (2002) Arch. Biochem. Biophys. 399, 37–48.
13. Tabares, L., Mazzanti, M. & Clapham, D. E. (1991) J. Membrane Biol. 123,
49–54.
14. Stehno-Bittel, L., Lu¨ckhoff, A. & Clapham, D. E. (1995) Neuron 14, 163–167.
15. Quesada, I., Nadal, A. & Soria, B. (1999) Diabetes 48, 2390 –2397.
16. Nadal, A., Rov ira, J. M., Laribi, O., Leon-Quinto, T., Andreu, E., Ripoll, C.
& Soria, B. (1998) FASEB J. 12, 1341–1348.
17. Mazzanti, M., Inocenti, B. & Rigatelli, M. (1994) FASEB J. 8, 231–236.
18. Nicotera, P., McConkey, D. J., Jones, D. P. & Orrenius, S. (1989) Proc. Natl.
Acad. Sci. USA 86, 453– 457.
19. Loew, L. M. (1993) in Fluorescent and Luminiscent Probes for Biological Activity,
ed. Mason, W. T. (Academic, San Diego), pp. 150–160.
20. Hardigham, G. E., Arnold, F. J. L. & Bading, H. (2001) Nat. Neurosci. 4,
261–267.
21. Jonas, J. C., Sharma, A., Hasenkamp, W., Ilkova, H., Patane, G., Laybutt, R.,
Bonner-Weir, S. & Weir, G. C. (1999) J. Biol. Chem. 274, 14112–14121.
22. Jiang, F. X., Cram, D. S., DeAizpurua, H. J. & Harrison, L. C. (1999) Diabetes
48, 722–730.
23. Santella, L. & Carafoli, E. (1997) FASEB J. 11, 1091–1109.
24. Aguilar-Bryan, L., Clement, J. P., Gonzalez, G., Kunjilwar, K. & Babenko, A.
(1998) Physiol. Rev. 78, 227–245.
25. Ashcroft, F. M. (1988) Annu. Rev. Neurosci. 11, 97–118.
26. Ozanne, S. E., Guest, P. C., Hutton, J. C. & Hales, C. N. (1995) Diabetologia
38, 277–282.
27. Holmuhamedov, E. L., Jovanovic, S., Dzeja, P. P., Jovanovic, A. & Terzic, A.
(1998) Am. J. Physiol. 275, H1567–H1576.
28. Braun, M., Anderie, I. & The´venod, F. (1997) FEBS Lett. 411, 255–259.
29. Hoy, M., Olsen, H. L., Bokvist, K., Buschard, K., Barg, S., Rorsman, P. &
Gromada, J. (2000) J. Physiol. (London) 527.1, 109–120.
30. John, S. A., Monck, J. R., Weiss, J. N. & Ribalet, B. (1998) J. Physiol. (London)
510.2, 333–345.
31. Makhina, E. N. & Nichols, C. G. (1998) J. Biol. Chem. 273, 3369–3374.
32. Zahradnikova, A. & Meszaros, L. G. (1998) J. Physiol. (London) 509.1, 29–38.
33. Bading, H. (2000) Eur. J. Biochem. 267, 5280 –5283.
34. Rogue, P. J. & Malv iya, A. N. (1999) EMBO J. 18, 5147–5152.
35. Jonas, J. C., Laybutt, D. R., Steil, G. M., Trivedi, N., Pertusa, J. G., Van de
Casteele, M., Weir, G. C. & Henquin, J. C. (2001) J. Biol. Chem. 276,
35375–35381.
36. Smith, P, A., Proks, P. & Moorhouse, A. (1999) Pflu¨gers Arch. 437, 577–588.
37. Raraty, M., Ward, J., Erdemli, G., Vaillant, C., Neoptolemos, J. P., Sutton, R.
& Petersen, O. H. (2000) Proc. Natl. Acad. Sci. USA 97, 13126–13131.
38. Somlyo, A. V., Gonzalez-Serratos, H., Schuman, H., McClellan, G. & Somlyo,
A. P. (1981) J. Cell. Biol. 90, 577–594.
39. Franco-Obregon, A., Wang, H. W. & Clapham, D. E. (2000) Biophys. J. 79,
202–214.
40. Mazzanti, M., DeFelice, L. J., Cohen, J. & Malter, H. (1990) Nature (London)
343, 764–767.
41. Longin, A. S., Mezin, P., Favier, A. & Verdetti, J. (1997) Biochem. Biophys. Res.
Commun. 235, 236–241.
42. Guilhard, G., Proteau, S., Payet, M. D., Escande, D. & Rousseau, E. (2000)
FEBS Lett. 476, 234–239.
43. Martin, C. & Ashley, R. H. (1993) Cell Calcium 14, 427– 438.
44. Schmid, A., Dehlinger-Kremer, M., Schulz, I. & Gogelein, H. (1990) Nature
(London) 346, 374–376.
45. Santella, L. & Kyozuka, K. (1997) Cell Calcium 22, 11–20.
46. Lemmens, R., Larsson, O., Berggren, P. O. & Islam, M. S. (2001) J. Biol. Chem.
276, 9971–9977.
47. Ishida, Y., Honda, H. & Watanabe, T. X. (1992) Br. J. Pharmacol. 106, 764–765.
48. Chopra, L. C., Charles, H. C. T. & Ward, J. P. T. (1992) Br. J. Pharmacol. 105,
259–260.
49. Endo, M. (1985) Cur r. Top. Membr. Transp. 25, 181–230.
50. Patel, J. R., Sukhareva, M., Coronado, R. & Moss, R. (1996) J. Membr. Biol .
154, 81–89.
51. Pozzan, T., Rizzuto, R., Volpe, P. & Meldolesi, J. (1994) Physiol. Rev. 74,
595–636.
52. Hofer, A. M., Curci, S., Machen, T. E. & Schulz, I. (1996) FASEB J. 10, 302–308.
Quesada et al. PNAS
July 9, 2002
vol. 99
no. 14
9549
PHYSIOLOGY
... Hepatocytes infected with adenovirus regarding GCaMP6m (Vector Biolabs, 1909) and NLS-R-GECO for 24 h were subsequently serum-starved for 16 h in Medium 199, after which changes in [Ca 2+ ] i were determined at 488 nm excitation/530 nm emission and 543 nm excitation/560 nm emission using a confocal microscope (Nikon, Japan). Isolated nuclei were loaded as reported 23,24 . The membrane-impermeant Ca 2+ dye Calcium Green Dextran (Thermo Fisher Scientific, C3713) was loaded into the nucleoplasm, while the membrane permeant Ca 2+ dye Fluo-4 AM (Thermo Fisher Scientific, F14201) was loaded into the nuclear envelope. ...
... We further examined whether glucagon-induced nuclear Ca 2+ signals were involved with the TRPM2 channel and perinuclear Ca 2+ stores. To this end, we loaded hepatocyte nuclei with either the membrane-permeable Fluo-4 AM Ca 2+ probe to measure Ca 2+ in the perinuclear region or membraneimpermeable Calcium Green Dextran to measure Ca 2+ in the nucleoplasm 18,23 . Intriguingly, confocal microscopy revealed that ADPR-induced reciprocal Ca 2+ exchanges between the perinuclear area and the nucleoplasm, showing a decrease in perinuclear Ca 2+ concentration and an increase in nucleoplasm Ca 2+ concentration (Fig. 4d, e). ...
Article
Full-text available
Hepatic glucose production by glucagon is crucial for glucose homeostasis during fasting, yet the underlying mechanisms remain incompletely delineated. Although CD38 has been detected in the nucleus, its function in this compartment is unknown. Here, we demonstrate that nuclear CD38 (nCD38) controls glucagon-induced gluconeogenesis in primary hepatocytes and liver in a manner distinct from CD38 occurring in the cytoplasm and lysosomal compartments. We found that the localization of CD38 in the nucleus is required for glucose production by glucagon and that nCD38 activation requires NAD+ supplied by PKCδ-phosphorylated connexin 43. In fasting and diabetes, nCD38 promotes sustained Ca2+ signals via transient receptor potential melastatin 2 (TRPM2) activation by ADP-ribose, which enhances the transcription of glucose-6 phosphatase and phosphoenolpyruvate carboxykinase 1. These findings shed light on the role of nCD38 in glucagon-induced gluconeogenesis and provide insight into nuclear Ca2+ signals that mediate the transcription of key genes in gluconeogenesis under physiological conditions.
... They are also implicated in the mechanism of preconditioning, where cells develop resistance to subsequent episodes of ischemia. Mitochondrial K ATP channels orchestrate a number of vital functions including the regulation of mitochondrial matrix volume, matrix calcium levels, oxidative phosphorylation, nuclear calcium transients and the orchestration of gene expression [60]. Importantly, the occurrence of a specific channel, the mitochondrial permeability transition pore (mPTP) located in the inner mitochondrial membrane, is a key event in cell damage during episodes of ischaemic stress [61]. ...
Article
Full-text available
Drug repurposing, also known as repositioning or reprofiling, has emerged as a promising strategy to accelerate drug discovery and development. This approach involves identifying new medical indications for existing approved drugs, harnessing the extensive knowledge of their bioavailability, pharmacokinetics, safety and efficacy. Levosimendan, a calcium sensitizer initially approved for heart failure, has been repurposed for oncology due to its multifaceted pharmacodynamics, including phosphodiesterase 3 inhibition, nitric oxide production and reduction of reactive oxygen species. Studies have demonstrated that levosimendan inhibits cancer cell migration and sensitizes hypoxic cells to radiation. Moreover, it exerts organ-protective effects by activating mitochondrial potassium channels. Combining levosimendan with traditional anticancer agents such as 5-fluorouracil (5-FU) has shown a synergistic effect in bladder cancer cells, highlighting its potential as a novel therapeutic approach. This drug repurposing strategy offers a cost-effective and time-efficient solution for developing new treatments, ultimately contributing to the advancement of cancer therapeutics and improved outcomes for patients. Further investigations and clinical trials are warranted to validate the effectiveness of levosimendan in oncology and explore its potential benefits in a clinical setting.
... Ca 2+ fluorescent dye Cal-520 AM is a methyl acetyl derivative that can penetrate the membrane to combine with free Ca 2+ in the nuclear envelope, while Ca 2+ fluorescent dye Cal-520 Dextran can pass directly through the nuclear pores into the nucleocytoplasm, where it accumulates and remains for a long time to indicate [Ca 2+ ] NC levels [30]. Previous studies have used such Ca 2+ fluorescent dyes (-AM and -Dextran) to detect [Ca 2+ ] NE and [Ca 2+ ] NC in the nuclei of liver cells [30], islet cells [31] and neurons [20]. Notably, the Cal-520 probe shows significant improvements in the signal-to-noise ratio and intracellular retention compared to other dyes. ...
Article
Full-text available
Disuse atrophy of skeletal muscle is associated with a severe imbalance in cellular Ca2+ homeostasis and marked increase in nuclear apoptosis. Nuclear Ca2+ is involved in the regulation of cellular Ca2+ homeostasis. However, it remains unclear whether nuclear Ca2+ levels change under skeletal muscle disuse conditions, and whether changes in nuclear Ca2+ levels are associated with nuclear apoptosis. In this study, changes in Ca2+ levels, Ca2+ transporters, and regulatory factors in the nucleus of hindlimb unloaded rat soleus muscle were examined to investigate the effects of disuse on nuclear Ca2+ homeostasis and apoptosis. Results showed that, after hindlimb unloading, the nuclear envelope Ca2+ levels ([Ca2+]NE) and nucleocytoplasmic Ca2+ levels ([Ca2+]NC) increased by 78% (p < 0.01) and 106% (p < 0.01), respectively. The levels of Ca2+-ATPase type 2 (Ca2+-ATPase2), Ryanodine receptor 1 (RyR1), Inositol 1,4,5-tetrakisphosphate receptor 1 (IP3R1), Cyclic ADP ribose hydrolase (CD38) and Inositol 1,4,5-tetrakisphosphate (IP3) increased by 470% (p < 0.001), 94% (p < 0.05), 170% (p < 0.001), 640% (p < 0.001) and 12% (p < 0.05), respectively, and the levels of Na+/Ca2+ exchanger 3 (NCX3), Ca2+/calmodulin dependent protein kinase II (CaMK II) and Protein kinase A (PKA) decreased by 54% (p < 0.001), 33% (p < 0.05) and 5% (p > 0.05), respectively. In addition, DNase X is mainly localized in the myonucleus and its activity is elevated after hindlimb unloading. Overall, our results suggest that enhanced Ca2+ uptake from cytoplasm is involved in the increase in [Ca2+]NE after hindlimb unloading. Moreover, the increase in [Ca2+]NC is attributed to increased Ca2+ release into nucleocytoplasm and weakened Ca2+ uptake from nucleocytoplasm. DNase X is activated due to elevated [Ca2+]NC, leading to DNA fragmentation in myonucleus, ultimately initiating myonuclear apoptosis. Nucleocytoplasmic Ca2+ overload may contribute to the increased incidence of myonuclear apoptosis in disused skeletal muscle.
... These channels are designated as a "metabolic sensor" that converts metabolic status into electrical activity (51). These channels are found both in the plasma membrane of cells and in the intracellular membranes of the mitochondria, the sarcoplasmic reticulum (52), zymogenic granules and nuclear membranes (53). K ATP channels are generally closed under normal physiological conditions and open in response to changes in altered cellular metabolic states, such as ischemia and hypoxia. ...
Article
Resveratrol (3,5,4'-trihydroxy-trans-stilbene) is a phytoalexin present in a variety of plant species. Resveratrol has a wide spectrum of pharmacologic properties, and it exhibits versatile biological effects on different human and animal models. The studies have shown that potassium (K) channels can be potential targets in the mechanism of resveratrol action. K channels play a crucial role in maintaining membrane potential. Inhibition of K channels causes membrane depo-larization and then contraction of smooth muscles, while the activation leads to membrane hyperpolarization and subsequently, relaxation. Five diverse types of K channels have been identified in smooth muscle cells in different tissue: ATP-sensitive K channels (K ATP), voltage-dependent K channels (Kv), Ca 2+-and voltage-dependent K channels (BK Ca), inward rectifier K channels (Kir), and tandem two-pore K channels (K 2P). The expression and activity of K channels altered in many types of diseases. Aberrant function or expression of K channels can be underlying in pathologies such as cardiac arrhythmia, diabetes mellitus, hypertension, preterm birth, preeclampsia, and various types of cancer. Modulation of K channel activity by molecular approaches and selective drug development may be a novel treatment modality for these dysfunctions in the future. The plant-derived non-toxic polyphenols, such as resveratrol, can alter K channel activity and lead to the desired outcome. This review presents the basic properties, physiological, pathophysiological functions of K channels, and pharmacological roles of resveratrol on the major types of K channels that have been determined in smooth muscle cells.
... Although the classic role of pore-forming subunits is to modulate conduction at the plasma membrane, there is also evidence that ion channels exist in the membranes of intracellular organelles, including the nucleus (Jang et al., 2015). Although the functional roles of nuclear ion channels are largely unknown, they are postulated to include transcription factor activation (Maruyama et al., 1995;Valenzuela et al., 2000;Quesada et al., 2002). The ability of ion channels to regulate gene transcription via modulation of electrical activity at the plasma membrane is termed excitation-transcription coupling. ...
Article
Full-text available
Several integral membrane proteins undergo regulated intramembrane proteolysis (RIP), a tightly controlled process through which cells transmit information across and between intracellular compartments. RIP generates biologically active peptides by a series of proteolytic cleavage events carried out by two primary groups of enzymes: sheddases and intramembrane-cleaving proteases (iCLiPs). Following RIP, fragments of both pore-forming and non-pore-forming ion channel subunits, as well as immunoglobulin super family (IgSF) members, have been shown to translocate to the nucleus to function in transcriptional regulation. As an example, the voltage-gated sodium channel β1 subunit, which is also an IgSF-cell adhesion molecule (CAM), is a substrate for RIP. β1 RIP results in generation of a soluble intracellular domain, which can regulate gene expression in the nucleus. In this review, we discuss the proposed RIP mechanisms of voltage-gated sodium, potassium, and calcium channel subunits as well as the roles of their generated proteolytic products in the nucleus. We also discuss other RIP substrates that are cleaved by similar sheddases and iCLiPs, such as IgSF macromolecules, including CAMs, whose proteolytically generated fragments function in the nucleus. Importantly, dysfunctional RIP mechanisms are linked to human disease. Thus, we will also review how understanding RIP events and subsequent signaling processes involving ion channel subunits and IgSF proteins may lead to the discovery of novel therapeutic targets. SIGNIFICANCE STATEMENT: Several ion channel subunits and immunoglobulin superfamily molecules have been identified as substrates of regulated intramembrane proteolysis (RIP). This signal transduction mechanism, which generates polypeptide fragments that translocate to the nucleus, is an important regulator of gene transcription. RIP may impact diseases of excitability, including epilepsy, cardiac arrhythmia, and sudden death syndromes. A thorough understanding of the role of RIP in gene regulation is critical as it may reveal novel therapeutic strategies for the treatment of previously intractable diseases.
... SUR1 is a component of the K ATP channels of mitochondrial membranes (SUR1/ROMK2) [16,17,23], and also, the nucleus and endoplasmic reticulum membranes (SUR1/Kir6.2), although SUR2 protein was also found in the membranes of the nucleus and reticulum [24,25]. Measurement of Kir6.1 mRNA expression in young SHR rats showed significantly lower expression compared to young Wistar rats. ...
Article
Full-text available
ATP-sensitive potassium (KATP) channels are participants of mechanisms of pathological myocardial remodeling containment. The aim of our work was to find the association of changes in the expression of Kir6.1, Kir6.2, SUR1, and SUR2 subunits of KATP channels with changes in heart function and structure during aging under conditions of the constant increase of vascular pressure. The experiments were carried out on young and old spontaneously hypertensive rats (SHR) and Wistar rats. The expression levels of KATP channels subunits were determined using reverse transcription and quantitative PCR. It is shown that the mRNA expression level of Kir6.1 in young SHR rats is significantly lower (6.3-fold, p = 0.035) than that of young Wistar rats that may be one of the causes of arterial hypertension in SHR. At the same time, mRNA expression of both Kir6.1 and Kir6.2 in old SHR rats was significantly higher (6.8-fold, p = 0.003, and 5.9-fold, p = 0.006, respectively) than in young hypertensive animals. In both groups of old animals, SUR2 expression was significantly reduced compared to young animals, in Wistar rats at 3.87-fold (p = 0.028) and in SHR rats at 48.2-fold (p = 0.033). Changes in SUR1 expression were not significant. Thus, significant changes in the cardiovascular system, including impaired function and structure of the heart in old SHR rats, were associated with a significant decrease in SUR2 expression that may be one of the mechanisms of heart failure decompensation. Therefore, it can be assumed that increased expression of SUR2 may be one of the protective mechanisms against pathological myocardial remodeling.
... There are data which indicate that potassium channels in perinuclear membranes are implicated in the regulation of gene transcription by linking changes of perinuclear membrane potentials, alterations of the nuclear calcium and potassium concentrations, and activation of transcription factors (Matzke et al., 2010;Checchetto et al., 2016). For example, blockade of ATP-sensitive potassium channels was shown to trigger calcium transients in isolated cell nuclei and induces CREB phosphorylation and c-myc expression (Hardingham et al., 2001;Quesada et al., 2002). Kv10.1 channels in perinuclear membranes modulate gene expression and genome stability via alteration of the potassium concentration which affects the stability of transcriptional repressor elements, e.g., in the c-MYC promoter (Sen and Gilbert, 1990;Siddiqui-Jain et al., 2002). ...
Article
Retinal pigment epithelial (RPE) cells express different subtypes of inwardly rectifying potassium (Kir) channels. We investigated whether human and rat RPE cells express genes of strongly rectifying Kir2 channels. We also determined the hypoxic and hyperosmotic regulation of Kir2.1 gene expression in cultured human RPE cells and the effects of siRNA-mediated knockdown of Kir2.1 on VEGFA expression, VEGF secretion, proliferation, and viability of the cells. Extracellular hyperosmolarity was induced by addition of NaCl or sucrose. Hypoxia and chemical hypoxia were produced by cell culture in 0.25% O2 and addition of CoCl2, respectively. Gene expression levels were evaluated by real-time RT-PCR. Rat RPE cells contained Kir2.1, Kir2.2, Kir2.3, and Kir2.4 gene transcripts while human RPE cells contained Kir2.1, Kir2.2, and Kir2.4 transcripts. Immunocytochemical data may suggest that Kir2.1 protein in cultured human cells is expressed in both perinuclear and plasma membranes. Kir2.1 gene expression and Kir2.1 protein level in human cells increased under hypoxic and hyperosmotic conditions. The expression of the Kir2.1 gene was mediated in part by diverse intracellular signal transduction pathways and transcription factor activities under both conditions; the hyperosmotic, but not the CoCl2-induced Kir2.1 gene expression was dependent on intracellular calcium signaling. Autocrine/paracrine activation of purinergic receptors contributed to Kir2.1 gene expression under hyperosmotic (P2Y1, P2Y2, P2X7) and CoCl2-induced conditions (P2Y2, P2X7). Exogenous VEGF, TGF-β1, and blood serum decreased Kir2.1 gene expression. Inhibition of VEGF receptor-2 increased the Kir2.1 gene expression under control conditions and in CoCl2-simulated hypoxia, and decreased it under high NaCl conditions. Knockdown of Kir2.1 by siRNA inhibited the CoCl2-induced and hyperosmotic transcription of the VEGFA gene and caused a delayed decrease of the constitutive VEGFA gene expression while VEGF protein secretion was not altered. Kir2.1 knockdown stimulated RPE cell proliferation under control and hyperosmotic conditions without affecting cell viability. The data indicate that Kir2.1 channel activity is required for the expression of the VEGFA gene and inhibits the proliferation of RPE cells. Under control and hypoxic conditions, the extracellular VEGF level may regulate the production of VEGF via its inhibitory effect on the Kir2.1 gene transcription; this feedback loop may prevent overproduction of VEGF.
Article
Cancer represents a leading cause of death worldwide. Hence, a better understanding of the molecular mechanisms causing and propelling the disease is of utmost importance. Several cancer entities are associated with altered K⁺ channel expression which is frequently decisive for malignancy and disease outcome. The impact of such oncogenic K⁺ channels on cell patho-/physiology and homeostasis and their roles in different subcellular compartments is, however, far from being understood. A refined method to investigate simultaneously metabolic and ionic signaling events on the level of individual cells and their organelles represent genetically encoded fluorescent biosensors, that allow a high-resolution investigation of compartmentalized metabolite or ion dynamics in a non-invasive manner. This feature of these probes makes them versatile tools to visualize and understand subcellular consequences of aberrant K⁺ channel expression and activity in K⁺ channel related cancer research.
Article
The field of mitochondrial ion channels underwent a rapid development during the last decade, thanks to the molecular identification of some of the nuclear-encoded organelle channels and to advances in strategies allowing specific pharmacological targeting of these proteins. Thereby, genetic tools and specific drugs aided definition of the relevance of several mitochondrial channels both in physiological as well as pathological conditions. Unfortunately, in the case of mitochondrial K+ channels, efforts of genetic manipulation provided only limited results, due to their dual localization to mitochondria and to plasma membrane in most cases. Although the impact of mitochondrial K+ channels on human diseases is still far from being genuinely understood, pre-clinical data strongly argue for their substantial role in the context of several pathologies, including cardiovascular and neurodegenerative diseases as well as cancer. Importantly, these channels are druggable targets, and their in-depth investigation could thus pave the way to the development of innovative small molecules with huge therapeutic potential. In the present review we summarize the available experimental evidences that mechanistically link mitochondrial potassium channels to the above pathologies and underline the possibility of exploiting them for therapy.
Article
Full-text available
The First European Conference on Calcium Signalling in the Cell Nucleus took place in Baia Paraelios, Calabria, Italy, from 4-8 October 1997. It was organized by O. Bachs, E. Carafoli, P. Nicotera and L. Santella (local organizers, G. Bagetta and D. Rotiroti) and attended by about 90 specialists. The scientific content was very high and the discussions were particularly intense. Considering that the area is famous for its controversies, one could perhaps have expected aggressive overtones. Instead, in spite of the liveliness of the discussions, the atmosphere was congenial and constructive. The controversies have not disappeared, but a better understanding of some of their origins is now well under way.
Article
Full-text available
A Ca(2+)-sensitive electrode was used to measure the Ca2+ concentration of the medium containing the heavy fraction of the fragmented sarcoplasmic reticulum (SR) prepared from guinea-pig psoas muscle. Among K(+)-channel blockers tested, 4-aminopyridine (4-AP), tetraethylammonium (TEA) and charybdotoxin elicited Ca2+ release from the SR, but apamin and glibenclamide did not. These results suggest that a reduction of SR K+ conductance leads to Ca2+ release from the SR.
Article
Aguilar-Bryan, Lydia, John P. Clement IV, Gabriela Gonzalez, Kumud Kunjilwar, Andrey Babenko, and Joseph Bryan. Toward Understanding the Assembly and Structure of K ATP Channels. Physiol. Rev. 78: 227–245, 1998. — Adenosine 5′-triphosphate-sensitive potassium (K ATP ) channels couple metabolic events to membrane electrical activity in a variety of cell types. The cloning and reconstitution of the subunits of these channels demonstrate they are heteromultimers of inwardly rectifying potassium channel subunits (K IR 6.x) and sulfonylurea receptors (SUR), members of the ATP-binding cassette (ABC) superfamily. Recent studies indicate that SUR and K IR 6.x associate with 1:1 stoichiometry to assemble a large tetrameric channel, (SUR/K IR 6.x) 4 . The K IR 6.x subunits form the channel pore, whereas SUR is required for activation and regulation. Two K IR 6.x genes and two SUR genes have been identified, and combinations of subunits give rise to K ATP channel subtypes found in pancreatic β-cells, neurons, and cardiac, skeletal, and smooth muscle. Mutations in both the SUR1 and K IR 6.2 genes have been shown to cause familial hyperinsulinism, indicating the importance of the pancreatic β-cell channel in the regulation of insulin secretion. The availability of cloned K ATP channel genes opens the way for characterization of this family of ion channels and identification of additional genetic defects.
Article
Discovered in the cardiac sarcolemma, ATP-sensitive K+ (KATP) channels have more recently also been identified within the inner mitochondrial membrane. Yet the consequences of mitochondrial KATP channel activation on mitochondrial function remain partially documented. Therefore, we isolated mitochondria from rat hearts and used K+ channel openers to examine the effect of mitochondrial KATP channel opening on mitochondrial membrane potential, respiration, ATP generation, Ca2+ transport, and matrix volume. From a mitochondrial membrane potential of -180 ± 15 mV, K+ channel openers, pinacidil (100 μM), cromakalim (25 μM), and levcromakalim (20 μM), induced membrane depolarization by 10 ± 7, 25 ± 9, and 24 ± 10 mV, respectively. This effect was abolished by removal of extramitochondrial K+ or application of a KATP channel blocker. K+ channel opener-induced membrane depolarization was associated with an increase in the rate of mitochondrial respiration and a decrease in the rate of mitochondrial ATP synthesis. Furthermore, treatment with a K+ channel opener released Ca2+ from mitochondria preloaded with Ca2+, an effect also dependent on extramitochondrial K+ concentration and sensitive to KATP channel blockade. In addition, K+ channel openers, cromakalim and pinacidil, increased matrix volume and released mitochondrial proteins, cytochrome c and adenylate kinase. Thus, in isolated cardiac mitochondria, KATP channel openers depolarized the membrane, accelerated respiration, slowed ATP production, released accumulated Ca2+, produced swelling, and stimulated efflux of intermembrane proteins. These observations provide direct evidence for a role of mitochondrial KATP channels in regulating functions vital for the cardiac mitochondria.
Chapter
Ca release from the sarcoplasmic reticulum (SR) is one of the most important steps in excitation–contraction coupling of skeletal muscle. This chapter describes the physiological release of Ca from the SR, various modes of Ca release from the SR, and the physiological significance of various Ca release mechanisms. Ca ion is the mediator of information of action potentials to the contractile machinery; however, the physiological source of the mediator Ca is not yet unequivocally established. The inhibitors of Ca-induced Ca release––procaine and adenine––were shown not to inhibit contraction of living skeletal muscle fibers induced by the depolarization of the surface membrane. Studies of the ionic composition of the lumen of the SR by electron-probe analysis show that there are no significant differences between the ionic compositions in the lumen of the SR and that in the cytoplasm except for Ca ion. The essential part of the physiological Ca release mechanism is almost entirely unknown; therefore, further studies, especially by using preparations, such as improved cut fibers, that retain the physiological tubule (T)–SR coupling mechanism, but have easy access to sarcoplasm so that its composition can be altered at will, are necessary.
Article
Approximately 60-70% of the total fiber calcium was localized in the terminal cisternae (TC) in resting frog muscle as determined by electron-probe analysis of ultrathin cryosections. During a 1.2 s tetanus, 59% (69 mmol/kg dry TC) of the calcium content of the TC was released, enough to raise total cytoplasmic calcium concentration by approximately 1 mM. This is equivalent to the concentration of binding sites on the calcium-binding proteins (troponin and parvalbumin) in frog muscle. Calcium release was associated with a significant uptake of magnesium and potassium into the TC, but the amount of calcium released exceeded the total measured cation accumulation by 62 mEq/kg dry weight. It is suggested that most of the charge deficit is apparent, and charge compensation is achieved by movement of protons into the sarcoplasmic reticulum (SR) and/or by the movement of organic co- or counterions not measured by energy dispersive electron-probe analysis. There was no significant change in the sodium or chlorine content of the TC during tetanus. The unchanged distribution of a permeant anion, chloride, argues against the existence of a large and sustained transSR potential during tetanus, if the chloride permeability of the in situ SR is as high as suggested by measurements on fractionated SR. The calcium content of the longitudinal SR (LSR) during tetanus did not show the LSR to be a major site of calcium storage and delayed return to the TC. The potassium concentration in the LSR was not significantly different from the adjacent cytoplasmic concentration. Analysis of small areas of I-band and large areas, including several sarcomeres, suggested that chloride is anisotropically distributed, with some of it probably bound to myosin. In contrast, the distribution of potassium in the fiber cytoplasm followed the water distribution. The mitochondrial concentration of calcium was low and did not change significantly during a tetanus. The TC of both tetanized and resting freeze-substituted muscles contained electron-lucent circular areas. The appearance of the TC showed no evidence of major volume changes during tetanus, in agreement with the estimates of unchanged (approximately 72%) water content of the TC obtained with electron-probe analysis.
Article
Cells respond to external signals by transducing the information into internal messengers such as Ca2+. The information-processing capability of the Ca2+ signalling system is enhanced by the well-known engineering principles of amplitude and frequency modulation.
Article
• The intracellular calcium concentration ([Ca2+]i) near the plasma membrane was measured in mouse pancreatic islet cells using confocal spot detection methods. • Whereas small cytosolic Ca2+ gradients were observed with 3 mM glucose, a steeper sustained gradient restricted to domains beneath the plasma membrane (space constant, 0.67 m) appeared with 16.7 mM glucose. • When the membrane potential was clamped with increasing K+ concentrations (5, 20 and 40 mM), no [Ca2+]i gradients were observed in any case. • Increasing glucose concentration (0, 5 and 16.7 mM) in the presence of 100 M diazoxide, a K+ channel opener, plus 40 mM K+ induced steeper [Ca2+]i gradients, confirming the role of membrane potential-independent effects of glucose. • Prevention of Ca2+ store refilling with 30 M cyclopiazonic acid (CPA) or blockade of uniporter-mediated Ca2+ influx into the mitochondria with 1 M carbonyl cyanide m-chlorophenyl hydrazone (CCCP) or 1 M Ru-360 significantly reduced the steepness of the 16.7 mM glucose-induced [Ca2+]i gradients. • Measured values of [Ca2+]i reached 6.74 ± 0.67 M at a distance of 0.5 m from the plasma membrane and decayed to 0.27 ± 0.03 M at a distance of 2 m. Mathematically processed values at 0.25 and 0 m gave a higher [Ca2+]i, reaching 8.18 ± 0.86 and 10.05 ± 0.98 M, respectively. • The results presented indicate that glucose metabolism generates [Ca2+]i microgradients, which reach values of around 10 M, and whose regulation requires the involvement of both mitochondrial Ca2+ uptake and endoplasmic reticulum Ca2+ stores.
Article
1. Intracellular calcium [Ca2+]i and channel activity were simultaneously recorded in single, dissociated mouse beta-cells kept in culture for 1-3 days. [Ca2+]i was estimated from microfluorometric ratio methods using Indo-1. Channel activity was measured using the cell-attached configuration of the patch-clamp technique. 2. At low glucose concentrations (0.3 mM), resting K+ATP channel activity was prevalent. Increasing glucose up to 16 mM, produced a gradual decrease in K+ATP channel activity over a time course of 90-120 s (temperature = 23 degrees C) and an increase in [Ca2+]i. 3. In the majority of experiments, glucose elicited biphasic action currents (action potentials) which preceded the rise in [Ca2+]i. There was a close correlation between spike frequency and the levels of [Ca2+]i. 4. The sulphonylurea tolbutamide (1 mM) blocked K+ATP channels in 10-20 s. K+ATP channel blockade was associated with a quick rise in [Ca2+]i. 5. When K+ATP channel activity was stimulated in the presence of diazoxide (100 microM), increasing the glucose concentration from 3 to 16 mM produced a decrease in [Ca2+]i. Only when diazoxide was removed did glucose produce an increase in [Ca2+]i. 6. In a small population of cells, glucose (16 mM) produced a small decrease in K+ATP channel activity but not an increase in [Ca2+]i. In such cells, tolbutamide blocked K+ATP channels and produced an increase in [Ca2+]i. 7. These results demonstrate a close correlation between K+ATP channel activity and [Ca2+]i in beta-cells. The findings are consistent with the model in which glucose metabolism produces a rise in [Ca2+]i through the blockade of K+ATP channels, membrane depolarization and calcium current activation.