ArticlePDF Available

Store-operated Ca2+ entry in astrocytes: Different spatial arrangement of endoplasmic reticulum explains functional diversity in vitro and in situ

Authors:
  • Bogomoletz Institute of Physiology, National Academy of Sciences of Ukraine

Abstract and Figures

Ca(2+) signaling is the astrocyte form of excitability and the endoplasmic reticulum (ER) plays an important role as an intracellular Ca(2+) store. Since the subcellular distribution of the ER influences Ca(2+) signaling, we compared the arrangement of ER in astrocytes of hippocampus tissue and astrocytes in cell culture by electron microscopy. While the ER was usually located in close apposition to the plasma membrane in astrocytes in situ, the ER in cultured astrocytes was close to the nuclear membrane. Activation of metabotropic receptors linked to release of Ca(2+) from ER stores triggered distinct responses in cultured and in situ astrocytes. In culture, Ca(2+) signals were commonly first recorded close to the nucleus and with a delay at peripheral regions of the cells. Store-operated Ca(2+) entry (SOC) as a route to refill the Ca(2+) stores could be easily identified in cultured astrocytes as the Zn(2+)-sensitive component of the Ca(2+) signal. In contrast, such a Zn(2+)-sensitive component was not recorded in astrocytes from hippocampal slices despite of evidence for SOC. Our data indicate that both, astrocytes in situ and in vitro express SOC necessary to refill stores, but that a SOC-related signal is not recorded in the cytoplasm of astrocytes in situ since the stores are close to the plasma membrane and the refill does not affect cytoplasmic Ca(2+) levels.
Content may be subject to copyright.
Cell Calcium 43 (2008) 591–601
Store-operated Ca2+ entry in astrocytes: Different spatial arrangement
of endoplasmic reticulum explains functional
diversity in vitro and in situ
Tatjyana Pivneva a,1, Brigitte Haas b,1, Daniel Reyes-Harob, Gregor Laube c,
Ruediger W. Veh c, Christiane Nolteb, Galina Skiboa, Helmut Kettenmannb,
aCytology Department, Bogomoletz Institute of Physiology, Bogomoletz Str. 4, 01024 Kiev, Ukraine
bMax Delbr¨uck Center for Molecular Medicine (MDC) Berlin-Buch, Robert-R¨ossle-Str. 10, 13125 Berlin, Germany
cDepartment of Electronmicroscopy and Molecular Neuroanatomy, Charite University Medicine, Schumannstraße 20/21, 10098 Berlin, Germany
Received 6 July 2007; received in revised form 19 September 2007; accepted 5 October 2007
Available online 3 December 2007
Abstract
Ca2+ signaling is the astrocyte form of excitability and the endoplasmic reticulum (ER) plays an important role as an intracellular Ca2+ store.
Since the subcellular distribution of the ER influences Ca2+ signaling, we compared the arrangement of ER in astrocytes of hippocampus
tissue and astrocytes in cell culture by electron microscopy. While the ER was usually located in close apposition to the plasma membrane in
astrocytes in situ, the ER in cultured astrocytes was close to the nuclear membrane. Activation of metabotropic receptors linked to release of
Ca2+ from ER stores triggered distinct responses in cultured and in situ astrocytes. In culture, Ca2+ signals were commonly first recorded close
to the nucleus and with a delay at peripheral regions of the cells. Store-operated Ca2+ entry (SOC) as a route to refill the Ca2+ stores could be
easily identified in cultured astrocytes as the Zn2+-sensitive component of the Ca2+ signal. In contrast, such a Zn2+-sensitive component was
not recorded in astrocytes from hippocampal slices despite of evidence for SOC. Our data indicate that both, astrocytes in situ and in vitro
express SOC necessary to refill stores, but that a SOC-related signal is not recorded in the cytoplasm of astrocytes in situ since the stores are
close to the plasma membrane and the refill does not affect cytoplasmic Ca2+ levels.
© 2007 Elsevier Ltd. All rights reserved.
Keywords: Astrocytes; Endoplasmic reticulum; Electron microscopy; Ca2+ signaling; Zn2+ ; Capacitative calcium entry
1. Introduction
The astrocyte excitability is based on Ca2+ signaling [1].
Ca2+ signals can propagate as waves over remarkably long
distances through a network of astrocytes both in culture
and in situ [2,3] or can increase locally in response to neu-
ronal activity [4]. Astrocytes express a large repertoire of
metabotropic receptors linked to Ca2+ signaling, e.g. for glu-
tamate or ATP [5]. Their activation is linked to release of
Ca2+ from intracellular stores. The endoplasmic reticulum
Corresponding author at: Max Delbr¨
uck Center for Molecular Medicine
(MDC) Berlin-Buch, Robert-R¨
ossle-Str. 10, 13125 Berlin, Germany.
Tel.: +49 30 9406 3325; fax: +49 30 9406 3819.
E-mail address: kettenmann@mdc-berlin.de (H. Kettenmann).
1Equal contribution.
(ER) is an important Ca2+ storing compartment, and crucially
contributes to the Ca2+ signals that influence cell activity.
Metabotropic receptor induced Ca2+ signals usually com-
prise two components, a rapid release of Ca2+ from the ER
and Ca2+influx from the extracellular space through slowly
activating store-operated channels in the plasma membrane,
also referred to as Ca2+-release activated Ca2+ (CRAC) chan-
nels. Depletion of the Ca2+ stores triggers store-operated
Ca2+ entry (SOC). This ‘capacitative’ Ca2+ entry is a ubiqui-
tous phenomenon, found in both excitable and non-excitable
cells (for review see [6]). While the currents related to
SOC have been well characterized, the molecular identity of
the channel itself and the signal that relays the Ca2+ con-
tent of the stores to the activation of SOC in the plasma
membrane, remain elusive. Several members of the tran-
sient receptor potential (TRP) family of ion channels have
0143-4160/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ceca.2007.10.004
592 T. Pivneva et al. / Cell Calcium 43 (2008) 591–601
been proposed as candidates for the ion channels [7]; for
review [6]), and one member, TRPC1, plays a role in astro-
cyte Ca2+ entry [8]. Recently, Orai1/CRACM1 have been
identified as a component central to the store-operated Ca2+
channel activation [9,10].Ca
2+-release activated Ca2+ influx
not only replenishes depleted Ca2+ stores, but also plays
a role in prolonging intracellular Ca2+ responses, thereby
regulating numerous cellular functions, such as gene tran-
scription, proliferation and cytokine release. In astrocytes, for
instance, intracellular Ca2+ levels control glutamate release
[11,12].
The activity of store-operated Ca2+ channels can be mod-
ulated by extracellular zinc [13,14], which is, besides iron,
the most abundant trace element in the body. In mouse astro-
cytes, Zn2+ can inhibit the capacitative Ca2+ influx, thereby
modulating the intracellular Ca2+ response to metabotropic
agonists [15]. In our previous work we demonstrated a
difference between the impact of Zn2+ on ligand-induced
Ca2+ responses in situ and in culture. While in cultured
astrocytes the activity of SOC is linked to a zinc-sensitive
elevated plateau phase, this zinc-sensitive component was
not recorded in astrocytes in hippocampal tissue [15]. One
possible explanation for the different behavior of cultured
and in situ cells could be different structural arrangement
of stores and Ca2+ entry channels in cultured versus in
situ astrocytes. Although there is evidence suggesting that
the ER in astrocytes consists of spatially distinct compart-
ments, at least in terms of release mechanisms [16], little is
known about the structural organization/localization of the
ER. In the present study, we therefore studied the spatial
arrangement of the ER in cultured astrocytes and astro-
cytes in the tissue and addressed the question whether a
distinct spatial organization would affect Ca2+ signaling in
astrocytes.
2. Materials and methods
2.1. Cell culture
Primary cultures of hippocampal astrocytes were prepared
as described previously [15] with some modifications. In
brief, hippocampi were dissected from brains of newborn
NMRI and eGFP/GFAP mice, and carefully freed from blood
vessels and meninges. Tissue was dissociated by trypsiniza-
tion and gentle trituration in the presence of 0.05% DNAase
(Worthington Biochem. Corp., Freehold, USA). After wash-
ing twice, cells were plated in dishes of 35 mm diameter
containing poly-l-lysine (PLL)-coated glass cover slips using
Basal Medium Eagle’s (BME)/10% fetal calf serum (FCS).
One day later, cultures were washed twice with Hank’s bal-
anced salt solution (HBSS) to remove cellular debris and
maintained for 4 days. When reaching subconfluent state,
cellular debris, microglia cells, oligodendrocytes as well as
their early precursor cells were dislodged by manual shak-
ing and removed by washing with HBSS. The purity of the
astrocytes was routinely determined by immunofluorescence
using a polyclonal antibody against glial fibrillary acidic pro-
tein (GFAP, DAKO, Hamburg, Germany), a specific astrocyte
marker. The cultures typically showed more than 90% cells
positive for GFAP.
2.2. Calcium imaging in cultured astrocytes
Calcium imaging was performed as described previously
[17] with slight modifications. Briefly, cultured astrocytes on
coverslips were dye-loaded for 30 min with 5 M Fluo-4-
acetoxymethylester (Fluo-4-AM, Invitrogen, Karlsruhe) in
HEPES buffer containing (in mM) NaCl (150), KCl (5,4),
MgCl2(1), CaCl2(2), HEPES (5) and d-glucose (10); pH
was adjusted to 7.4. Subsequently, coverslips were trans-
ferred to the stage of an upright microscope (Axioskop;
Zeiss, Oberkochen, Germany) and superfused with buffer. In
Ca2+-free solution, CaCl2was omitted, MgCl2was increased
to 2 mM. A fluorescence imaging system (Till Photonics,
M¨
unchen, Germany) was used for excitation and monitor-
ing the emitted fluorescence. Excitation wavelength was set
to 495 nm by means of a monochromator. Intracellular Ca2+
changes were detected by a cooled CCD camera (Sensi-
Cam, PCO, Kelheim, Germany) mounted to the microscope.
Images were sampled at 0.5 Hz using a software developed in
our group combined with the TIDA (HEKA, Lamprecht, Ger-
many) software. F/F0was calculated by averaging the first 15
values of the recording (F0) and dividing all values by this.
Values are given as percentage of increase from baseline. To
display responding cells, the images showing the reaction
were averaged and the background was subtracted.
High resolution calcium imaging in subcellular areas
was performed with a confocal laser scanning microscope
(Noran Odyssey XL confocal microscope Praire Tech-
nologies, WI). The scanner was mounted on the upright
microscope (Axioskop, Zeiss) equipped with a 40×magnifi-
cation, numerical aperture 0.75, water immersion objective.
A wave length of 488 nm was used for optical excitation and
the emission wave length was measured at 530nm. Images
were acquired every 30 ms (33 Hz) for a period of 40 s and
only the raising phase of the response to ATP was considered
for analysis.
To obtain the fluorescence intensity ratio, the resting flu-
orescence value was determined after averaging ten images
of the scan mode. Intensities of all subsequently recorded
images were divided by these resting values. These values
were obtained with the Image Pro Plus 5.0 software and fur-
ther analyzed with the Origin 7.0 software. Astrocyte cultures
that reacted to the application of 10 M ATP were selected
for the experiments.
2.3. Preparation of brain slices, dye-loading and
calcium recordings from astrocytes in situ
All experiments were performed according to the guide-
lines of the German animal protection law. For experiments,
T. Pivneva et al. / Cell Calcium 43 (2008) 591–601 593
10–14 days old NMRI mice were used. Slice preparation,
dye-loading with Fluo-4-AM, and image recording was per-
formed as described in Haas et al. [18]. Briefly, mice were
decapitated and the brains removed. Two hundred and fifty
micrometer thick slices were cut in ice-cold artificial cere-
brospinal fluid (ACSF) using a Vibratome (Microm HM
650 V, Microm International, Germany). ACSF contained (in
mM) 134 NaCl, 2.5 KCl, 2 CaCl2, 1.3 MgCl2, 26 NaHCO3,
1.25 K2HPO4and 10 mM glucose. By continuously gassing
the solution with carbogen (5% CO2, 95% O2), the pH
was adjusted to 7.4. To obtain calcium-free solution, CaCl2
was omitted, MgCl2was increased to 2 mM. The hippocam-
pus was recognized by its specific architecture.Slices were
stored in gassed ACSF for 45 min prior to staining. Dye-
loading with 10 M Fluo-4-AM was performed for 40 min
at room temperature (RT), followed by 10 min incubation
at 37 C.
To selectively label astrocytes, slices were stained with
the red-fluorescent dye Sulforhodamine 101 (SR101; Molec-
ular Probes). We used a staining procedure modified from
Nimmerjahn et al. [23]. Acute slices were immersed in
1 ml standard ACSF containing 30 g (final concentration
50 M) SR101 for 1 min. The solution was quickly removed
and replaced with the conventional staining solution con-
taining Fluo-4-AM as described above. After the staining
period, astrocytes were labeled with both, the red-fluorescent
SR101 and the green fluorescent calcium-indicator dye
Fluo-4.
Hippocampal slices were transferred to the stage of an
upright microscope (Axioskop; Zeiss, Oberkochen, Ger-
many) and intracellular Ca2+ changes were recorded using
a cooled CCD camera as described above. Data analy-
sis and image illustration were performed as described
above.
2.4. Immunocytochemistry
Cultured astrocytes of NMRI or eGFP/GFAP trans-
genic mice were immunostained for calreticulin or Sec61
to show the distribution of endoplasmic reticulum (ER).
Astrocytes on coverslips were rinsed briefly with Dul-
becco’s phosphate buffered saline (DPBS) and then fixed
for 30 min with 2% formaldehyde at RT. Cells were per-
meabilized with 0.1% Triton X-100 in blocking solution
(5% BSA, 5% normal goat serum in 0.1 M PBS) for
30 min at RT. Cells were incubated overnight at 4 C with
anti-calreticulin antibodies (Upstate Biotechnologies, Lake
Placid, USA; 1:50) or anti-Sec61 antibodies (kindly pro-
vided by Thomas Sommer, MDC, Berlin, Germany, 1:1000)
diluted in blocking solution. After extensive rinsing cells
were incubated with secondary antibodies (goat anti-rat
IgG coupled to Cy2 or goat anti-rabbit IgG Alexa 568,
respectively) for 2 h at RT. Cells were rinsed, mounted
with aqua polymount (Polysciences Inc.), and investi-
gated in a fluorescence microscope (Zeiss Axiophot, Zeiss
Germany).
2.5. Electron microscopy
2.5.1. Astrocyte culture
For electron microscopic investigation cultured astrocytes
prepared from newborn NMRI or eGFP/GFAP transgenic
mice were shortly rinsed with DPBS and then fixed for 30 min
with 2% paraformaldehyde, 1.25% glutaraldehyde in 0.1 M
phosphate buffer (PB pH 7.4). After several rinses cells were
postfixed in 1% osmium tetroxide, dehydrated in increasing
series of ethanol, transferred to propylene oxide, infiltrated
with epoxy resin (Plano, Marburg, Germany; Araldite CY
212, DDSA, DMP-30), and flat embedded. Ultrathin sections
were stained with uranyl acetate and lead citrate and studied
with a Ziess 910 electron microscope at 80 kV.
2.5.2. Preparation of tissue
Six weeks old eGFP/GFAP transgenic mice (n= 6) were
anesthetized deeply by pentobarbital and perfused intra-
cardially with saline followed by perfusion with fixative (4%
paraformaldehyde, 0.05% glutaraldehyde and 0.2% of picric
acid in 0.1 M phosphate buffer; pH 7.4) for 20 min. The brains
were removed and 500 m thick, parasagittal slices were
cut with a vibratome (Leica, VT 1000S, Germany). Freeze-
substitution and low temperature embedding in acrylic resins
were carried out as described earlier [19,20]. For cryoprotec-
tion, slices were placed into sucrose solutions (from 0.5 M to
2 M sucrose) in 0.05 M Tris–Maleat buffer. Slices were then
slammed onto copper blocks cooled in liquid N2. This was
followed by freeze-substitution with methanol and embed-
ding in Lowicryl HM 20 resins at 50 C (Chemischer Werke
Lowi GMBH, Germany).
2.5.3. Postembedding immunocytochemistry
Ultrathin sections (75–90 nm thickness) from Lowicryl-
embedded hippocampal area were picked up on formvar-
coated copper grids and were incubated on drops of blocking
solution (50 mM Tris–HCl, pH 7.4, 0.3% NaCl, 10% normal
goat serum (NGS)) for 30 min. Then, sections were floated
over night at 4 C on drops containing the primary antibody,
anti-green fluorescent protein antibodies (GFP; rabbit IgG
fraction, Molecular Probes, Mo Bi Tec.) diluted 1:50 in TBS,
2% NGS. Afterwards, the sections were washed and incu-
bated on drops of goat anti-rabbit IgG coupled to 12 nm
gold particles (Immunotech-Dianova; 1:30) for 2 h at room
temperature.
After extensive rinsing in PB and ultra-pure water, the sec-
tions were contrasted with saturated aqueous uranyl acetate
followed by staining with lead citrate. For negative con-
trol, primary antibodies were either omitted and slices were
incubated in blocking solution or replaced by 5% normal rab-
bit serum. Sections were investigated with a Zeiss electron
microscope (model EM 910; Carl Zeiss MicroImaging, Inc.).
To test for the specificity of immunogold labeling, 20 electron
micrographs (20,000×magnification) from hippocampal
areas were analyzed. Astrocytic versus non-astrocytic areas
were measured using NIH-image (http://rsb.info.nih.gov/nih-
594 T. Pivneva et al. / Cell Calcium 43 (2008) 591–601
Fig. 1. Localization of endoplasmic reticulum (er) in astrocytes in vitro
and astrocytes from hippocampus of eGFP/GFAP transgenic mice in situ.
Astrocytes were prepared for electron microscopic study as described.
(A) Transverse section of cultured astrocytes (in this case prepared from
eGFP/GFAP mice). Both rough and smooth ER (see arrowheads in inset) are
localized in the vicinity of the nucleus (nu). Arrow in inset point to GFAP
fibers. (B) Astrocytes in situ were labeled by eGFP followed by immunogold.
Specificity of the labeling was checked as described in Section 2. For better
image/). The density of labeling was calculated (astrocytes:
6 particles/m2, non-astrocytic areas: 2 particles/m2).
2.5.4. Quantifying the probability of ER localization
ER localization was quantified by putting a rectangular
frame defining the region of interest (ROI, 3mm by 15 mm)
on electron micrographs taken at 12,000×magnification
either from astrocytes in culture or from astrocytes in situ.
At this magnification the ROI equals 0.25 m×1.25 m.
Three reference structures were used: (1) nuclear membrane
with ER structure appearing within the same ROI as a por-
tion of the nuclear membrane; (2) plasma membrane with the
ER structure appearing in the same ROI as a portion of the
plasma membrane and, (3) cytoplasm ER structures being
within the frame with no membrane structures neither close
to nuclear not to the plasma membrane. Care was taken that
the longer dimension of the ROI was oriented in parallel to the
nuclear membrane or plasma membrane. Thus, the distance
between ER and membrane was between 0.05 and 0.125 m.
Micrographs of 41 cultured astrocytes and of 19 astrocytes
in situ were analyzed; the ROI was placed three times on
each reference structure in a given astrocyte by an unbiased
investigator. If ER and the reference structure appeared in the
same ROI, this was counted as positive event. Frequencies of
positive events were calculated for each reference structure,
and mean values determined.
We used Statistica software (v. 5, StatSoft, Tulsa, USA)
and applied the two-tailed Kolmogorov–Smirnov test to
assess the differences between samples [21].P< 0.05 was
considered statistically significant. The mean values were
normalized to the most frequently detected localization and
expressed as % of that.
3. Results
3.1. Differences in ER distribution between astrocytes in
culture and astrocytes in situ
In our previous work we described differences in Ca2+
responses of astrocytes in situ and in culture [15] and hypoth-
esized that one possible explanation for these differences in
shape of the Ca2+ signal after metabotropic receptor stimula-
tion is a different arrangement of the calcium entry and store
sites in cultured astrocytes as compared to astrocytes in situ.
To address this question, we compared the distribution of the
ER between cultured hippocampal astrocytes and astrocytes
visibility gold particles are encircled and astrocytic membrane is marked
by dotted line. (C) Probability of ER localization in different cellular com-
partments in astrocytic cells in situ and in vitro was analyzed as described.
Results were normalized to the most frequent occurrence in a certain com-
partment and are given as %. Two-tailed Kolmogorov–Smirnov test was
applied to assess the differences between samples, and P< 0.05 was con-
sidered as statistically significant. *P< 0.05; **P< 0.01. Bars: A=1m,
B= 0.25 m.
T. Pivneva et al. / Cell Calcium 43 (2008) 591–601 595
Fig. 2. Cytoplasmic expression of eGFP does not affect ER-distribution
in cultured astrocytes. (A) Immunocytochemical labeling of ER in cultured
astrocytes with calreticulin antibodies supports the observation that the ER is
preferentially arranged in the perinuclear area. (B and C) Astrocytic cultures
prepared from two different litters of eGFP/GFAP mice were immunolabeled
with Sec61 antibodies (C) to show the localization of ER. There was com-
parable ER-distribution in eGFP-positive and -negative astrocytes. Arrows
hint at eGFP-positive and -negative cell. Bars: 10 m.
in the intact hippocampus using electron microscopy (EM).
Horizontal sections of astrocyte cultures were analyzed in
the transmission EM and micrographs were taken from a
total of 41 samples showing both, nucleus and cytoplasmic
membrane on the same micrograph (Fig. 1A) at a magnifi-
cation to resolve the ER structures (12,000×). As described
in the method section, we determined the probability of the
ER being either close to the nucleus or to the plasma mem-
brane. The probability ratio between ER localization close
to the plasma versus nuclear membrane is shown in Fig. 1C.
In cultured astrocytes, we most frequently found ER cister-
nae in the areas close to the nuclear membrane, whereas the
probability to detect ER close to the plasma membrane was
only around 30% of that. The difference was found to be sta-
tistically significant (Fig. 1C). Immunocytochemistry with
antibodies to either calreticulin or Sec61 (Fig. 2), proteins
associated with endo(sarco)plasmic reticulum membranes,
confirmed our observation of preferential ER localization in
perinuclear areas. There is strong immunolabeling around the
nucleus whereas in the processes of the cells the labeling is
usually less pronounced (Fig. 2A).
To unambiguously identify the astrocytes’ cytoplasmic
compartments in the hippocampal tissue, we made use of
the eGFP/GFAP transgenic mouse model, in which astro-
cytes express the enhanced green fluorescent protein (eGFP)
under control of the GFAP promoter [22]. Astrocytic com-
partments were identified by postembedding staining with
anti-GFP antibodies. As for cultured astrocytes (see above),
the localization of ER was analyzed in micrographs taken
from 19 different astrocytes. In contrast to cultured astro-
cytes, the ER in astrocytes in situ was preferentially observed
in the vicinity of the plasma membrane while it was found
less frequently close to the nucleus (Fig. 1B). The probability
(after normalization) to detect ER close to the nucleus was
around 20% and we most frequently found ER close to the
plasma membrane (Fig. 1C). Frequencies for nuclear- ver-
sus membrane-associated ER localization were analyzed by
Kolmogorov–Smirnov test as described and turned out to be
significantly different.
To exclude the possibility that the presence of eGFP pro-
tein disturbs the ER arrangement, we analyzed astrocyte
cultures of eGFP/GFAP transgenic animals. Those cultures
normally contain both eGFP-expressing and eGFP-negative
astrocytes (Fig. 2B). Cells were counterstained with anti-
bodies specific for the ER (calreticulin, Ser61). As shown
in Fig. 2B and C, the ER is usually arranged in the perin-
uclear region both in eGFP-negative and eGFP-expressing
astrocytes. These observations also indicate that the cyto-
plasmic abundance of the eGFP protein does not affect the
spatial organization of Ca2+ stores.
3.2. Differences in Ca2+ responses in astrocytes in situ
compared to cultured astrocytes
For comparison of astrocyte Ca2+ responses in culture and
in situ, we used similar experimental paradigms to record
Ca2+ changes. While in our previous study [15] we made use
of the eGFP/GFAP transgenic mouse to identify astrocytes
and the long-wavelength excitation Ca2+ indicator X-rhod-1
(excitation 550 nm), here we used the Ca2+ indicator Fluo-
4 both in hippocampal slices and in hippocampal astrocyte
cultures obtained from NMRI mice. To identify astrocytes,
hippocampal slices were co-loaded with Fluo-4-AM and Sul-
forhodamine 101 (SR101), the latter being a dye which is
specifically taken up by astrocytes [23].Fig. 3A shows that
the cells responding to ATP by an increase in Fluo-4 fluo-
rescence in stratum radiatum of the hippocampus were also
labeled by SR101 (Fig. 2B and C). To confirm the speci-
ficity of SR101 staining in the hippocampus, we also applied
this dye to slices of eGFP/GFAP transgenic mice. Fig. 3D–F
show that virtually all eGFP-positive cells are labeled by the
red SR101 dye. We found, however, that a subpopulation
of SR101 labeled cells were not eGFP-positive. This is in
line with our previous observation that in this eGFP/GFAP
transgenic model not all astrocytes are eGFP-positive [22].
We analyzed Ca2+ responses to ATP and 2-methyl-S-ADP
in astrocytes in cultures. A brief application (30 s) of ATP
(100 M) or 2-methyl-S-ADP (1 M) triggered a biphasic
Ca2+ transient in astrocytes; a rapid increase in intracellular
596 T. Pivneva et al. / Cell Calcium 43 (2008) 591–601
Fig. 3. Fluo-4 loaded cells in slices are identified as astrocytes hippocampal slices obtained from 12 days old NMRI mice were loaded with Fluo-4 and
Sulforhodamine 101 (SR101). (A) Reacting cells at 488 nm illumination in response to the application of ATP. (B) SR101 fluorescence at 578nm (red channel).
(C) The overlay shows that most of the reacting cells are sulforhodamine positive, indicating they are astrocytes. Scale bars in (A–C) correspond to 100 m. (D)
Hippocampal slices from 12 days old eGFP/GFAP mice at 488nm illumination were loaded with SR101, scale bar 50 m. (E) Double-labeled cells (arrows)
are visible in the magnified view in the overlay. (F) Arrowheads indicate SR101 positive cells which do not express eGFP. Scale bar represents 20 m.
Ca2+ followed by a rapid decline to an elevated plateau level.
Ca2+ levels remained elevated for about 2–3 min and then
slowly returned to base level (Fig. 4B and D). The plateau
phase depended on extracellular Ca2+ as it was lacking in the
absence of extracellular Ca2+ (Fig. 4D). We never observed
differences between application of ATP (n= 5, 50–150 cells
each) or 2-methyl-S-ADP (n= 17, 50–100 cells each). Zn2+
reversibly blocked the plateau phase indicative of SOC acti-
vation. Also here, ATP (Fig. 4B, n= 5, 50–100 cells each)
and 2-methyl-S-ADP acted similarly (data not shown, n=3,
50 cells each).
To analyze the involvement of SOC in Ca2+ signals from
astrocytes of hippocampal slices, we used similar paradigms
as in culture. We compared the Ca2+ response to ATP
(100 M, n= 21 slices, 5 animals, Fig. 4F) or 2-methyl-S-
ADP (10 m, n= 3 slices, 2 animals, data not shown) in
the absence or presence of zinc (100 M) in Fluo-4 stained
cells. Fig. 4E displays the reacting cells (background was
subtracted) after ATP application. Fig. 4F gives averaged
responses from 50 cells after control application of ATP and
in the presence of Zn2+. In contrast to responses in cultured
cells, the Ca2+ signal in astrocytes in situ after stimulation
with ATP usually lacked the plateau phase while in cultured
astrocytes it usually lasted for several minutes. Consequently,
Zn2+ had no apparent effect, neither on the amplitude nor
on the time course of the Ca2+ response. We also compared
reactions to 2-methyl-S-ADP in normal versus Ca2+-free
solution similar as we did in cultured hippocampal astro-
cytes (n= 25 slices, 8 animals). Fig. 4H shows average Ca2+
traces obtained from 25 cells in hippocampus responding to
2-methyl-S-ADP in the absence and presence of extracellular
Ca2+.
We used the integral of the Ca2+ transient to further
quantify the contribution of the Zn2+-sensitive component.
Therefore the baseline was subtracted and the response was
normalized to the peak. In vitro we found that the peak
area was decreased in the presence of Zn2+ by approx. 50
%, while in situ there was no difference. This substantiates
that the ATP response lacks the Zn2+-sensitive component
in situ.
Fig. 4. Differences in Ca2+ influx upon purinergic receptor stimulation in vitro vs. in situ. (A) The left image shows the Fluo-4 fluorescence in a hippocampal
astrocyte culture. Due to background subtraction only reacting cells are displayed. Traces in (B) show the averaged response of 100 cells to application of
100 M ATP, either in the presence or in the absence of 100 MZn
2+. The delay in the onset of the ATP response is due to a slow perfusion system, not to a
delayed response after ATP application. Note that the plateau of the purinergic Ca2+ signal disappears in the presence of Zn2+ . The maxima of the peaks are
aligned to better display the differences in the delayed component. (C and D) Similar to (A and B) the averaged reaction of 150 cells to application of 1M
2-methyl-S-ADP (MeSADP) compared to Ca2+-free conditions is shown. (E) Fluo-4 fluorescence of the cells in the slice shown in (F) (after ATP application).
Due to background subtraction only reacting cells are displayed. Cells display the typical shape of astrocytes. (F) Traces from the reaction after 30s application
of 100 M ATP and after 2 min application of 100 MZn
2+ prior to the ATP application. The averaged response of 50 cells in a hippocampal slice is shown.
Note that the difference between the traces is more subtle compared to the responses obtained in culture. (G and H) Similar to (A, E and F) the average
response to 2-methyl-S-ADP (30 s, 10M) from 25 cells in a hippocampal slice is shown, in control and nominally Ca2+ -free solution. (I) The integral of the
Ca2+ transient was calculated after baseline subtraction and alignment of the respective peaks. Note that in the presence of Zn2+ the peak area is significantly
decreased in vitro by approx. 50% (P< 0.05 (paired t-test), while there is no difference in situ.
T. Pivneva et al. / Cell Calcium 43 (2008) 591–601 597
598 T. Pivneva et al. / Cell Calcium 43 (2008) 591–601
3.3. Purinergic receptor activation in cultured
astrocytes elicits a faster [Ca2+]iincrease in perinuclear
areas as compared to cell processes
Our ultrastructural studies indicated a concentration of ER
close to the nucleus in cultured astrocytes. To test whether
the Ca2+ increase due to release from internal stores would
differ within a cell, we studied Ca2+ kinetics with sub-
cellular resolution. Recordings were made in low-density
cultures to exclude that calcium signals originate from closely
apposed cells via gap junction coupling. We sampled images
of Fluo-4-fluorescence (F/F0) at 33 Hz while applying the
P2Y-specific ligand 2-Me-S-ADP (1 M). We then compared
the kinetics of the Ca2+ increase indicative of Ca2+ release
from stores close to the nucleus and in more distal regions
remote from the nucleus. The rising phase of the normal-
ized F/F0was fitted to a sigmoidal function and the time to
obtain 50% of the maximal signal (EC50) was calculated; the
EC50 for regions close to the nucleus was 0.41 ±0.04 s, and
0.60 ±0.07 s for regions in the astrocytic processes (n= 32).
In 75% (24/32) of the recorded cells the Ca2+ response
occurred more rapidly in the perinuclear regions as compared
to more distal regions (Fig. 5). Only in 19% of the cells (6
from 32) the increase in [Ca2+]iwas faster in processes than
close the nucleus (not shown). In 2 out of 32 astrocytes we
observed no significant differences between these two areas.
These results indicate that the Ca2+ release is commonly ini-
tiated close to the nucleus and subsequently spreads within
the cell.
4. Discussion
4.1. Ultrastructural differences between astrocytes in
culture and in situ
In the present study, we found significant differences in
the subcellular distribution of the ER in cultured astrocytes as
compared to astrocytes in tissue. In situ, ER is preferentially
localized close to the plasma membrane while in cultured
cells it is often concentrated around the nucleus. The high
density of ER in the perinuclear area of cultured astrocytes
has been reported by Grimaldi et al. [24]. There is ample
evidence that the geometry of a cell influences the architec-
ture of the ER, and a correlation between cell morphology,
cytoskeletal organization and regulation of Ca2+ homeostasis
has been shown for several cell types [25–27]. Astrocytes in
culture usually appear as flattened polygonal cells without
a complex process pattern and thus display a quite different
morphology as their in vivo counterparts. The morphological
phenotype of cultured astrocytes can be affected by prolonged
exposure to cAMP. This treatment results in differentiation,
namely transformation from the flat polygonal form into a
stellate, process-bearing phenotype. With regard to cell mor-
phology, the cAMP-differentiated astrocytes in vitro are very
much comparable to the astrocytes in the tissue. Interestingly,
as reported by Grimaldi et al. [24] the stellate astrocytes
showed a more uniform distribution of the ER including
within processes and close to the plasma membranes, which
is in agreement with our findings from the EM analysis.
4.2. The eGFP/GFAP-mouse is a valuable tool for
morphological studies at the EM level
We used the eGFP/GAFP transgenic animal model for
identifying astrocyte compartments in the ultrastructural
level [22,28]. In contrast to GFAP, which is present only
in the main processes [29], eGFP is spread throughout the
entire astrocytic cytoplasm including fine cellular structures
surrounding synapses. To visualize eGFP distribution on the
EM level, we used an anti-eGFP antibody. Usually, EM-
immunolabeling procedures are counterproductive for a good
ultrastructure preservation of membranes, and details such as
organelles are difficult to detect. Our postembedding pro-
tocol was sufficient to reveal (rough) ER cisternae in the
tissue slices and at the same time preserve immunogenicity
for specific binding of anti-GFP and immunogold antibodies.
This enabled us to unequivocally identify the astrocytic com-
partment and analyze the morphological arrangement of the
major Ca2+ store, the ER in relation to the cytoplasma mem-
brane. Ultrastructure details, such as cytoskeletal elements
connecting ER cisternae with plasma membrane as shown
by Lencesova et al. [30] for muscle cells or neurons were,
however, not detectable.
4.3. Functional differences between astrocytes in culture
and in situ
These structural differences were paralleled by distinct
Ca2+ responses to metabotropic receptor activation when
comparing culture versus in situ astrocytes. To make the data
between culture and in situ comparable, we used the same
types of animals, same dyes and the same ligands to stimulate
cells. Taking cells into culture, however, may trigger expres-
sion of other proteins than in situ resulting, for instance, in
different patterns of purinergic receptors accounting for the
difference in responses to application of ATP. While there is
no convincing evidence for the presence of functional P2X
receptors in astrocytes from hippocampal brain slices, P2X
receptors are expressed in culture [31]. We therefore also
used a specific agonist to P2Y1, P2Y12 and P2Y13 recep-
tors. We found, however, similar differences between culture
and slices. In cultured astrocytes, Ca2+ responses showed a
biphasic time course, a rapid transient followed by a plateau
phase. The rapid transient signal was due to release from
intrinsic Ca2+ stores while the subsequent plateau phase was
due to capacitative Ca2+, as it could be blocked by zinc [15].
This influx of extracellular Ca2+ is most likely via store-
operated Ca2+ channels, also referred to as CRAC [32].Ca
2+
entry through store-operated Ca2+ channels has been shown
for many cell types including cultured glial cells [32–34].
In contrast, astrocytes in situ lack the second, zinc-sensitive
T. Pivneva et al. / Cell Calcium 43 (2008) 591–601 599
Fig. 5. Analysis of Ca2+ transients in subcellular regions of cultured hippocampal astrocyte Astrocytes were loaded with Fluo-4-AM as described and stimulated
with 1 M 2-Me-S-ADP. The internal Ca2+ level rises faster in perinuclear regions. (A) The upper left image shows the spatial uniformity of resting Ca2+ levels
in Fluo-4 loaded cell, 1 and 2 mark the ROIs. Series of ratio images (F/F0) captured at the times indicated were colour coded indicating the level of Ca2+. (B)
Time course of Fluo-4 fluorescence in the two ROIs depicted in the upper left image of panel (A). Arrows indicate start (t= 0.0 s) and end (2.33s) of recording
interval for the false colour images given in (A). Note that the response is faster in the ROI close to the nucleus (1) than in the ROI close to the cytoplasmic
membrane (2). Right graph shows the fit of the data to a sigmoid function.
response. We have, however, shown previously by applying
different protocols that can activate SOC, that astrocytes in
situ express SOC necessary to refill the stores [15]. This indi-
cates that store-operated Ca2+ entry in astrocytes in situ does
not lead to Ca2+ increases in the cytoplasm. We assume that
this is a general principle since we found this difference also
by activating metabotropic glutamate receptors [15].
Grimaldi et al. [24] also compared Ca2+ responses in the
flat versus the cAMP-differentiated, stellate astrocytes. They
found larger responses in the stellate cells and, in addition,
they reported a more pronounced second plateau indicative
of a SOC component recorded in the soma. Although, as
discussed above, the cultured stellate astrocytes are mor-
phologically more reminiscent of astrocytes in situ they
apparently show functional differences with respect to Ca2+
signaling.
4.4. The location of ER could explain functional
differences
The location of ER in relation to the plasma membrane
could provide an explanation for these functional differences
between astrocytes in culture and in tissue. Both, astrocytes
in situ and in vitro express SOC necessary to refill stores,
but a SOC-related signal is not recorded in the cytoplasm
of astrocytes in situ since the stores are close to the plasma
membrane and the Ca2+ is taken up rapidly without affect-
ing cytoplasmic Ca2+ levels. This indicates that Ca2+ entry
channels and stores are closely associated.
Indeed a direct molecular coupling of ER and store-
operated Ca2+ channels has been suggested [35,16]: in such
“signaling microdomains”, also named PlasmaERsomes
[16], the ER membrane can come in very close contact with
the cytoplasmic membrane. Structurally, these microdomains
resemble very much the triads, a coupling between the
sarcoplasmic reticulum and transverse tubule, in the myofib-
rils of skeletal muscle cells. Microdomains/plasmaERsomes
define a restricted space in the cytosol in which Ca2+ ions can
accumulate without getting off into the bulk cytosol. In astro-
cytes, for instance, these microdomains contain clusters of
isoform specific Na+-pumps and Na+/Ca2+-exchangers [36].
Recently, Golovina [8] demonstrated, by using a membrane
tethered Ca2+ indicator to monitor rapid large changes in
[Ca2+]iclose to the plasma membrane and high resolution
immunocytochemistry, that these microdomains also contain
TRPC1-encoded store-operated calcium entry. This result
600 T. Pivneva et al. / Cell Calcium 43 (2008) 591–601
was obtained in cultured astrocytes and our data would indi-
cate that this cooperation of stores and store-operated Ca2+
entry would even be more efficient in tissue astrocytes.
We cannot exclude that other factors may account for the
differences in SOC-related Ca2+ signal: taking cells into cul-
ture drastically changes their phenotype also with regard to
expression of proteins. It is thus conceivable that astrocytes
in vitro have a higher level of expression of SOCs/CRAC
and therefore have an increased driving force for Ca2+ entry.
In turn, astrocytes in situ may extrude Ca2+ out of the cell
more efficiently, both of which would result in differences
in shape of the Ca2+ signal. In addition, mitochondria are
another huge site for Ca2+ storage [37]. The mechanisms how
mitochondria shape the profile of an intracellular Ca2+ signal
in astrocytes and if there exist differences in mitochondrial
status/physiology as well as their distribution and spatial rela-
tionship with the ER in vitro and in situ, are little understood
and remain to be determined.
4.5. Functional consequences of different ER
localization
The localization of the ER Ca2+ stores will influence all
types of intracellular Ca2+ responses. This will also be rel-
evant for Ca2+ waves, a long-range communication form of
astrocytes. Large Ca2+ waves can be easily elicited in cultured
astrocytes, while they are more difficult to elicit in intact tis-
sue and so far have not been observed in vivo under normal
conditions (for review see [38]). Two mechanisms are dis-
cussed for the propagation of these waves: (Ca2+ dependent)
released ATP activating adjacent purinergic receptors or dif-
fusion of Ca2+/IP3 through gap junctions to coupled cells.
Both mechanisms might be intensified by larger cytosolic
Ca2+ elevations and this might be an explanation for spread
over larger distances in culture.
Longer and higher elevated cytosolic Ca2+ concentrations
could also influence other Ca2+-dependent processes such as
the calcium-dependent release of gliotransmitters, e.g. gluta-
mate. The larger Ca2+ responses could facilitate the release
in cultured cells. Indeed, the extend and impact of gluta-
mate release in situ compared to in vitro is still under active
investigation (e.g. [39]).
Acknowledgements
We thank Irene Haupt for excellent technical assistance;
Dr. Thomas Sommer (MDC) for providing antibodies to
Sec61, and Dr. Anja Hoffmann (Berlin) and Alexander G.
Nikonenko (Kiev) for helpful discussions.
References
[1] A. Verkhratsky, H. Kettenmann, Calcium signalling in glial cells,
Trends Neurosci. 19 (1996) 346–352.
[2] A.H. Cornell-Bell, S.M. Finkbeiner, M.S. Cooper, S.J. Smith, Gluta-
mate induces calcium waves in cultured astrocytes: long-range glial
signaling, Science 247 (1990) 470–473.
[3] C.G. Schipke, C. Boucsein, C. Ohlemeyer, F. Kirchhoff, H. Ketten-
mann, Astrocyte Ca2+ waves trigger responses in microglial cells in
brain slices, FASEB J. 16 (2002) 255–257.
[4] J. Grosche, V. Matyash, T. Moller, A. Verkhratsky, A. Reichenbach,
H. Kettenmann, Microdomains for neuron–glia interaction: parallel
fiber signaling to Bergmann glial cells, Nat. Neurosci. 2 (1999) 139–
143.
[5] A. Verkhratsky, R.K. Orkand, H. Kettenmann, Glial calcium: home-
ostasis and signaling function, Physiol. Rev. 78 (1998) 99–141.
[6] A.B. Parekh, J.W. Putney Jr., Store-operated calcium channels, Physiol.
Rev. 85 (2005) 757–810.
[7] Y. Mori, M. Wakamori, T. Miyakawa, M. Hermosura, Y. Hara, M.
Nishida, K. Hirose, A. Mizushima, M. Kurosaki, E. Mori, K. Gotoh,
T. Okada, A. Fleig, R. Penner, M. Iino, T. Kurosaki, Transient recep-
tor potential 1 regulates capacitative Ca(2+) entry and Ca(2+) release
from endoplasmic reticulum in B lymphocytes, J. Exp. Med. 195 (2002)
673–681.
[8] V.A. Golovina, Visualization of localized store-operated calcium entry
in mouse astrocytes. Close proximity to the endoplasmic reticulum, J.
Physiol. 564 (2005) 737–749.
[9] S. Feske, Y. Gwack, M. Prakriya, S. Srikanth, S.H. Puppel, B. Tanasa,
P.G. Hogan, R.S. Lewis, M. Daly, A. Rao, A mutation in Orai1 causes
immune deficiency by abrogating CRAC channel function, Nature 441
(2006) 179–185.
[10] M. Vig, C. Peinelt, A. Beck, D.L. Koomoa, D. Rabah, M. Koblan-
Huberson, S. Kraft, H. Turner,A. Fleig, R. Penner, J.P.Kinet, CRACM1
is a plasma membrane protein essential for store-operated Ca2+ entry,
Science 312 (2006) 1220–1223.
[11] P. Bezzi, G. Carmignoto, L. Pasti, S. Vesce, D. Rossi, B.L. Rizzini,
T. Pozzan, A. Volterra, Prostaglandins stimulate calcium-dependent
glutamate release in astrocytes, Nature 391 (1998) 281–285.
[12] V. Parpura, T.A. Basarsky, F. Liu, K. Jeftinija, S. Jeftinija, P.G. Haydon,
Glutamate-mediated astrocyte-neuron signalling, Nature 369 (1994)
744–747.
[13] C.R. Jan, S.N. Wu, S.C.J. Tseng, Zn2+ increases resting cytosolic Ca2+
levels and abolishes capacitative Ca2+ entry induced by ATP in MDCK
cells, Naunyn Schmiedebergs Arch. Pharmacol. 360 (1999) 249–
255.
[14] D. Choi, J.Y. Koh, Zinc and brain injury, Annu. Rev. Neurosci. 21
(1998) 347–375.
[15] W. Kresse, I. Sekler, A. Hoffmann, O. Peters, C. Nolte, A. Moran, H.
Kettenmann, Zinc ions are endogenous modulators of neurotransmitter-
stimulated capacitative Ca2+ entry in both cultured and in situ mouse
astrocytes, Eur. J. Neurosci. 21 (2005) 1626–1634.
[16] M.P. Blaustein, V.A. Golovina, Structural complexity and functional
diversity of endoplasmic reticulum Ca(2+) stores, Trends Neurosci. 24
(2001) 602–608.
[17] S. Kirischuk, T. Moller, N. Voitenko, H. Kettenmann, A. Verkhratsky,
ATP-induced cytoplasmic calcium mobilization in Bergmann glial
cells, J. Neurosci. 15 (1995) 7861–7871.
[18] B. Haas, C.G. Schipke, O. Peters, G. Sohl, K. Willecke, H. Kettenmann,
Activity-dependent ATP-waves in the mouse neocortex are indepen-
dent from astrocytic calcium waves, Cereb. Cortex 16 (2006) 237–
246.
[19] A. Baude, Z. Nusser, E. Molnar, R.A. Mcllhinney, P. Somogyi, High
resolution immunogold localization of AMPA type glutamate recep-
tor subunits at synaptic and non-synaptic sites in rat hippocampus,
Neuroscience 69 (1995) 1031–1055.
[20] Z. Nusser, S.G. Cull-Candy, M. Farrant, Differences in synaptic
GABAa receptor number underlie variation in GABA mini amplitude,
Neuron 19 (1997) 697–709.
[21] A.G. Nikonenko, G.G. Skibo, Technique to quantify local clustering
of synaptic vesicles using single section data, Microsc. Res. Tech. 65
(2004) 287–291.
T. Pivneva et al. / Cell Calcium 43 (2008) 591–601 601
[22] C. Nolte, M. Matyash, T. Pivneva, C.G. Schipke, C. Ohlemeyer, U.K.
Hanisch, F. Kirchhoff, H. Kettenmann, G FAP promoter-controlled
eGFP-expressing transgenic mice: a tool to visualize astrocytes and
astrogliosis in living brain tissue, Glia 33 (2001) 72–86.
[23] A. Nimmerjahn, F. Kirchhoff, J.N. Kerr, F. Helmchen, Sulforhodamine
101 as a specific marker of astroglia in the neocortex in vivo, Nat.
Method 1 (2004) 31–37.
[24] M. Grimaldi, A. Favit, D.L. Alkon, cAMP-induced cytoskeleton
rearrangement increases calcium transients through the enhancement
of capacitative calcium entry, J. Biol. Chem. 274 (1999) 33557–
33564.
[25] J.A. Rosado, S.O. Sage, The actin cytoskeleton in store-mediated cal-
cium entry, J. Physiol. 526 (2000) 221–229.
[26] C.M. Ribeiro, J. Reece, J.W. Putney Jr., Role of the cytoskeleton in
calcium signaling in NIH 3T3 cells. An intact cytoskeleton is required
for agonist-induced [Ca2+]isignaling, but not for capacitative calcium
entry, J. Biol. Chem. 272 (1997) 26555–26561.
[27] J.R. Holda, L.A. Blatter, Capacitative calcium entry is inhibited in vas-
cular endothelial cells by disruption of cytoskeletal microfilaments,
FEBS Lett. 403 (1997) 191–196.
[28] R. Jabs, T. Pivneva, K. Huttmann, A. Wyczynski, C. Nolte, H. Ketten-
mann, C. Steinhauser, Synaptic transmission onto hippocampal glial
cells with hGFAP promoter activity, J. Cell. Sci. 118 (2005) 3791–
3803.
[29] E.A. Bushong, M.E. Martone, Y.Z. Jones, M.H. Ellisman, Protoplas-
mic astrocytes in CA1 stratum radiatum occupy separate anatomical
domains, J. Neurosci. 22 (2002) 183–192.
[30] L. Lencesova, A. O’Neill, W.G. Resneck, R.J. Bloch, M.P. Blaustein,
Plasma membrane–cytoskeleton–endoplasmic reticulum complexes in
neurons and astrocytes, J. Biol. Chem. 279 (2004) 2885–2893.
[31] W. Walz, G. Gimpl, C. Ohlemeyer, H. Kettenmann, Extracellular ATP-
induced currents in astrocytes: involvement of a cation channel, J.
Neurosci. Res. 38 (1994) 12–18.
[32] A.B. Parekh, R. Penner, Store depletion and calcium influx, Physiol.
Rev. 77 (1997) 901–930.
[33] J.W. Putney Jr., TRP, inositol 1,4,5-trisphosphate receptors, and
capacitative calcium entry, Proc. Natl. Acad. Sci. U.S.A. 96 (1999)
14669–14671.
[34] E.C. Toescu, T. Moller, H. Kettenmann, A. Verkhratsky, Long-term
activation of capacitative Ca2+ entry in mouse microglial cells, Neuro-
science 86 (1998) 925–935.
[35] M.J. Berridge, P. Lipp, M.D. Bootman, Signal transduction—The cal-
cium entry Pas de Deux, Science 287 (2000) 1604–1605.
[36] M. Juhaszova, M.P. Blaustein, Na+pump low and high ouabain affinity
alpha subunit isoforms are differently distributed in cells, Proc. Natl.
Acad. Sci. U.S.A. 94 (1997) 1800–1805.
[37] V.Y. Ganitkevich, The role of mitochondria in cytoplasmic Ca2+
cycling, Exp. Physiol. 88 (2003) 91–97.
[38] E. Scemes, C. Giaume, Astrocyte calcium waves: what they are and
what they do, Glia 54 (2006) 716–725.
[39] T.A. Fiacco, C. Agulhon, S.R. Taves, J. Petravicz, K.B. Casper, X.
Dong, J. Chen, K.D. McCarthy, Selective stimulation of astrocyte cal-
cium in situ does not affect neuronal excitatory synaptic activity, Neuron
54 (2007) 611–626.
... Epileptic variants may affect STIM1/Orai1-mediated SOCE in ER Our previous work indicated that astrocytic knockdown of GLT-1 impairs the calcium signaling pathway [39]. In astrocytes, Ca 2+ in the ER lumen is the major source of intracellular Ca 2+ , and SOCE links ER with plasmalemmal Ca 2+ entry [47,48]. The depletion of ER luminal Ca 2+ triggers STIM, the ER-resident Ca 2+ sensor, to interact with plasmalemmal SOCE channels, including both Orai and transient receptor potential (TRP) families. ...
... SOCE is the main calcium regulator for non-excitatory neural cells, such as astrocytes and microglia [47,48] and astrocytic knockdown of GLT-1 impaired the calcium signaling pathway [39]. Both STIM1 and STIM2 were found to be elevated in the dentate gyrus, CA1 and CA3 regions of chronic epileptic mice and in a hippocampal sample from a patient with medial temporal lobe epilepsy [57]. ...
Article
Full-text available
Epilepsy is a common neurological disorder and glutamate excitotoxicity plays a key role in epileptic pathogenesis. Astrocytic glutamate transporter GLT-1 is responsible for preventing excitotoxicity via clearing extracellular accumulated glutamate. Previously, three variants (G82R, L85P, and P289R) in SLC1A2 (encoding GLT-1) have been clinically reported to be associated with epilepsy. However, the functional validation and underlying mechanism of these GLT-1 variants in epilepsy remain undetermined. In this study, we reported that these disease-linked mutants significantly decrease glutamate uptake, cell membrane expression of the glutamate transporter, and glutamate-elicited current. Additionally, we found that these variants may disturbed stromal-interacting molecule 1 (STIM1)/Orai1-mediated store-operated Ca ²⁺ entry (SOCE) machinery in the endoplasmic reticulum (ER), in which GLT-1 may be a new partner of SOCE. Furthermore, knock-in mice with disease-associated variants showed a hyperactive phenotype accompanied by reduced glutamate transporter expression. Therefore, GLT-1 is a promising and reliable therapeutic target for epilepsy interventions.
... In most electrically non-excitable cells, the depletion of ER Ca 2+ store triggers secondary Ca 2+ influx, known as SOCE, involving plasmalemmal channels [26]. Many studies have identified SOCE as the primary pathway for glial Ca 2+ signaling in all types of glia within the central nervous system [27,28]. However, until recently, little was known about the glial Ca 2+ signaling mechanism in the peripheral nervous system including autonomic and sensory ganglia. ...
Article
Full-text available
Satellite glial cells (SGCs), a major type of glial cell in the autonomic ganglia, closely envelop the cell body and even the synaptic regions of a single neuron with a very narrow gap. This structurally unique organization suggests that autonomic neurons and SGCs may communicate reciprocally. Glial Ca²⁺ signaling is critical for controlling neural activity. Here, for the first time we identified the machinery of store-operated Ca²⁺ entry (SOCE) which is critical for cellular Ca²⁺ homeostasis in rat sympathetic ganglia under normal and pathological states. Quantitative real-time PCR and immunostaining analyses showed that Orai1 and stromal interaction molecules 1 (STIM1) proteins are the primary components of SOCE machinery in the sympathetic ganglia. When the internal Ca²⁺ stores were depleted in the absence of extracellular Ca²⁺, the number of plasmalemmal Orai1 puncta was increased in neurons and SGCs, suggesting activation of the Ca²⁺ entry channels. Intracellular Ca²⁺ imaging revealed that SOCE was present in SGCs and neurons; however, the magnitude of SOCE was much larger in the SGCs than in the neurons. The SOCE was significantly suppressed by GSK7975A, a selective Orai1 blocker, and Pyr6, a SOCE blocker. Lipopolysaccharide (LPS) upregulated the glial fibrillary acidic protein and Toll-like receptor 4 in the sympathetic ganglia. Importantly, LPS attenuated SOCE via downregulating Orai1 and STIM1 expression. In conclusion, sympathetic SGCs functionally express the SOCE machinery, which is indispensable for intracellular Ca²⁺ signaling. The SOCE is highly susceptible to inflammation, which may affect sympathetic neuronal activity and thereby autonomic output.
... The addition of a calcium-free medium with CBX does not result in Ca 2+ signals in hippocampal astrocytes, whereas supplementation with MSC-EV led to the generation of transient Ca 2+ responses or Ca 2+ oscillations in 47±16% of astrocytes without a significant decrease in amplitude ( Figure 9A). The endoplasmic reticulum (ER) is the main Ca 2+ depot in astrocytes [67]. The addition of 1 µM thapsigargin (TG), an inhibitor of sarco/endoplasmic reticulum Ca-ATPase (SERCA), led to the depletion of Ca 2+ from the ER of astrocytes, as evidenced by an increase in [Ca 2+ ] i ( Figure 9B), and the further addition of MSC-EV does not induce the generation of Ca 2+ signals in astrocytes. ...
Article
Full-text available
Mesenchymal stromal cells (MSC) are widely recognized as potential effectors in neuroprotective therapy. The protective properties of MSC were considered to be associated with the secretion of extracellular vesicles (MSC-EV). We explored the effects of MSC-EV in vivo on models of traumatic and hypoxia-ischemia (HI) brain injury. Neuroprotective mechanisms triggered by MSC-EV were also studied in vitro using a primary neuroglial culture. Intranasal administration of MSC-EV reduced the volume of traumatic brain damage, correlating with a recovery of sensorimotor functions. Neonatal HI-induced brain damage was mitigated by the MSC-EV administration. This therapy also promoted the recovery of sensorimotor functions, implying enhanced neuroplasticity, and MSC-EV-induced growth of neurites in vitro supports this. In the in vitro ischemic model, MSC-EV prevented cell calcium (Ca2+) overload and subsequent cell death. In mixed neuroglial culture, MSC-EV induced inositol trisphosphate (IP3) receptor-related Ca2+ oscillations in astrocytes were associated with resistance to calcium overload not only in astrocytes but also in co-cultured neurons, demonstrating intercellular positive crosstalk between neural cells. This implies that phosphatidylinositol 3-Kinase/AKT signaling is one of the main pathways in MSC-EV-mediated protection of neural cells exposed to ischemic challenge. Components of this pathway were identified among the most enriched categories in the MSC-EV proteome.
... The location and geometry of endoplasmic reticulum influence calcium signaling (Berridge et al., 2003). According to Pivneva et al. (2008), the endoplasmic reticulum locates in vicinity of the nucleus and is also dispersed in the cytosol, near the plasma membrane. In astrocytes in situ, the investigators have found most of reticulum in vicinity of plasma membrane, that is, in the branches. ...
Preprint
Full-text available
Most of the functions performed by astrocytes in brain information processing are related to calcium waves. Experimental studies involving calcium waves present discrepant results, leading to gaps in the full understanding of the functions of these cells. The use of mathematical models help to understand the experimental results, identifying chemical mechanisms involved in calcium waves and the limits of experimental methods. The model is diffusion-based and uses receptors and channels as boundary conditions. The computer program developed was prepared to allow the study of complex geometries, with several astrocytes, each of them with several branches. The code structure allows easy adaptation to various experimental situations in which the model can be compared. The code was deposited in the ModelDB repository, and will be under number 266795 after publication. A sensitivity analysis showed the relative significance of the parameters and identifies the ideal range of values for each one. We showed that several sets of values can lead to the same calcium signaling dynamics. This encourages the questioning of parameters to model calcium signaling in astrocytes that are commonly used in the literature, and it suggests better experimental planning. In the final part of the work, the effects produced by the endoplasmic reticulum when located close to the extremities of the branches were evaluated. We conclude that when they are located close to the region of the glutamatergic stimulus, they favor local calcium dynamics. By contrast, when they are located at points away from the stimulated region, they accelerate the global spread of signaling.
... Importantly, this was performed in non-permeabilized cells, to label only NMDAR located at the plasma membrane (PM). In these experiments, we observed that the ER and mitochondria accumulate mainly in the perinuclear region, as it has been reported previously for the ER, 34 although both organelles are also present throughout some regions of the cell soma ( Figure 4A,B). We also found that GluN1 tends to accumulate in the PM of the perinuclear region as we described in our previous work. ...
Article
Full-text available
Glutamate N‐methyl‐D‐aspartate (NMDA) receptor (NMDAR) is critical for neurotransmission as a Ca²⁺ channel. Nonetheless, flux‐independent signaling has also been demonstrated. Astrocytes express NMDAR distinct from its neuronal counterpart, but cultured astrocytes have no electrophysiological response to NMDA. We recently demonstrated that in cultured astrocytes, NMDA at pH6 (NMDA/pH6) acting through the NMDAR elicits flux‐independent Ca²⁺ release from the Endoplasmic Reticulum (ER) and depletes mitochondrial membrane potential (mΔΨ). Here we show that Ca²⁺ release is due to pH6 sensing by NMDAR, whereas mΔΨ depletion requires both: pH6 and flux‐dependent NMDAR signaling. Plasma membrane (PM) NMDAR guard a non‐random distribution relative to the ER and mitochondria. Also, NMDA/pH6 induces ER stress, endocytosis, PM electrical capacitance reduction, mitochondria‐ER, and ‐nuclear contacts. Strikingly, it also produces the formation of PM invaginations near mitochondria along with structures referred to here as PM‐mitochondrial bridges (PM‐m‐br). These and earlier data strongly suggest PM‐mitochondria communication. As proof of the concept of mass transfer, we found that NMDA/pH6 provoked mitochondria labeling by the PM dye FM‐4‐64FX. NMDA/pH6 caused PM depolarization, cell acidification, and Ca²⁺ release from most mitochondria. Finally, the MCU and microtubules were not involved in mΔΨ depletion, while actin cytoskeleton was partially involved. These findings demonstrate that NMDAR has concomitant flux‐independent and flux‐dependent actions in cultured astrocytes.
... In practice however, IP 3 diffusion between astrocyte somata could be more complicated. This is because connections between astrocytes through GJCs are mostly at the cell distal processes ) whose complex morphology and narrow intracellular space (Witcher et al., 2007;Pivneva et al., 2008) could considerably hinder IP 3 diffusion from/to somata. Moreover, GJCs cluster at discrete sites of these processes (Nagy and Rash, 2000), thereby constraining the diffusion pathway of IP 3 from one cell to another. ...
Preprint
Full-text available
Astrocytes organize in complex networks through connections by gap junction channels that are regulated by extra‐ and intracellular signals. Calcium signals generated in individual cells, can propagate across these networks in the form of intercellular calcium waves, mediated by diffusion of second messengers molecules such as inositol 1,4,5-trisphosphate. The mechanisms underpinning the large variety of spatiotemporal patterns of propagation of astrocytic calcium waves however remain a matter of investigation. In the last decade, awareness has grown on the morphological diversity of astrocytes as well as their connections in networks, which seem dependent on the brain area, developmental stage, and the ultrastructure of the associated neuropile. It is speculated that this diversity underpins an equal functional variety but the current experimental techniques are limited in supporting this hypothesis because they do not allow to resolve the exact connectivity of astrocyte networks in the brain. With this aim we present a general framework to model intercellular calcium wave propagation in astrocyte networks and use it to specifically investigate how different network topologies could influence shape, frequency and propagation of these waves.
... These cells represent a good model system for astrocytes due to various merits outlined earlier [28][29][30][31][32] . These cells undergo differentiation and are shown to propagate calcium ion waves, called astrocyte excitability 56 , in the brain as well as under in vitro conditions 57,58 . Treatment with dibutyryl cAMP 59,60 or taxol 54 enabled these cells to differentiate, giving typical neuronal morphology. ...
Article
Full-text available
Methamphetamine (METH) is a powerfully addictive psychostimulant that has a pronounced effect on the central nervous system (CNS). The present study aimed to assess METH toxicity in differentiated C6 astroglia-like cells through biochemical and toxicity markers with acute (1 h) and chronic (48 h) treatments. In the absence of external stimulants, cellular differentiation of neuronal morphology was achieved through reduced serum (2.5%) in the medium. The cells displayed branched neurite-like processes with extensive intercellular connections. Results indicated that acute METH treatment neither altered the cell morphology nor killed the cells, which echoed with lack of consequence on reactive oxygen species (ROS), nitric oxide (NO) or inhibition of any cell cycle phases except induction of cytoplasmic vacuoles. On the other hand, chronic treatment at 1 mM or above destroyed the neurite-like processors and decreased the cell viability that paralleled with increased levels of ROS, lipid peroxidation and lactate, depletion in glutathione (GSH) level and inhibition at G0/G1 phase of cell cycle, leading to apoptosis. Pre-treatment of cells with N-acetyl cysteine (NAC, 2.5 mM for 1 h) followed by METH co-treatment for 48 h rescued the cells completely from toxicity by decreasing ROS through increased GSH. Our results provide evidence that increased ROS and GSH depletion underlie the cytotoxic effects of METH in the cells. Since loss in neurite connections and intracellular changes can lead to psychiatric illnesses in drug users, the evidence that we show in our study suggests that these are also contributing factors for psychiatric-illnesses in METH addicts.
... Based on experimental evidence on Ca 2+ compartmentalization, we treated each astrocytic process as an independent unit with high (50%) ER-to-cytosol volume ratio 90 . Calcium responses at these tiny domains . ...
Preprint
Full-text available
Accumulation of amyloid-β peptide (Aβ), a hallmark of Alzheimer's disease (AD), is associated with synchronous hyperactivity and dysregulated Ca2+ signaling in hippocampal astrocytes. However, the consequences of altered Ca2+ signaling on the temporal dynamics of Ca2+ and gliotransmitter release events at astrocytic microdomains are not known. We have developed a detailed biophysical model of microdomain signaling at a single astrocytic process that accurately describes key temporal features of Ca2+ events and Ca2+-mediated kiss-and-run and full fusion exocytosis. Using this model, we ask how aberrant plasma-membrane Ca2+ pumps and mGluR activity, molecular hallmarks of Aβ toxicity that are also critically involved in Ca2+ signaling, modify astrocytic feedback at a tripartite synapse. We show that AD related molecular pathologies increase the rate and synchrony of Ca2+ and exocytotic events triggered by neuronal activity. Moreover, temporal precision between Ca2+ and release events, a mechanism indispensable for rapid modulation of synaptic transmission by astrocytes, is lost in AD astrocytic processes. Our results provide important evidence on the link between AD-related molecular pathology, dysregulated calcium signaling and gliotransmitter release at an astrocytic process.
Article
Ethnopharmacological relevance Activated astrocytes are involved in the progression of neurodegenerative diseases. Traditionally, Ailanthus altissima (Mill.) Swingle, widely distributed in East Asia, has been used as a medicine for the treatment of fever, gastric diseases, and inflammation. Although A. altissima has been reported to play an anti-inflammatory role in peripheral tissues or cells, its role in the central nervous system (CNS) remains unclear. Aim of the study In the present study, we investigated the anti-inflammatory effects and mechanism of action of A. altissima in primary astrocytes stimulated by lipopolysaccharide (LPS). Materials and methods A nitrite assay was used to measure nitric oxide (NO) production, and the tetrazolium salt 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) assay was performed to determine cytotoxicity. The expression levels of inducible nitric oxide synthase (iNOS), cyclooxygenase-2 (COX-2), and mitogen-activated protein kinase (MAPK) were determined with western blotting. Reverse-transcription PCR was used to assess the expression of inflammatory cytokines. The levels of reactive oxygen species were measured using 2,7-dichlorodihydrofluorescein diacetate. Luciferase assay and immunocytochemistry were used for assessing nuclear factor-kappa B (NF-κB) transcription and p65 localization, respectively. Memory and social interaction were analyzed using the Y-maze and three-chamber tests, respectively. Results The ethanol extract of A. altissima leaves (AAE) inhibited iNOS and COX-2 expression in LPS-stimulated astrocytes. Moreover, AAE reduced the transcription of various proinflammatory mediators, hindered NF-κB activation, and suppressed extracellular signal-regulated kinase (ERK) and c-Jun N-terminal kinase (JNK) activation without p38 activation. Ultra-high performance liquid chromatography with mass spectrometry analysis revealed that AAE comprised ethyl gallate, quercetin, and kaempferol, along with luteolin, which has anti-inflammatory properties, and repressed LPS-induced nitrite levels and the nuclear translocation of p65. Finally, oral administration of AAE attenuated LPS-induced memory and social impairment in mice and repressed LPS-induced ERK and JNK activation in the cortices of mice. Conclusion AAE could have therapeutic uses in the treatment of neuroinflammatory diseases via suppression of astrocyte activation.
Article
Full-text available
Protoplasmic astrocytes are increasingly thought to interact extensively with neuronal elements in the brain and to influence their activity. Recent reports have also begun to suggest that physiologically, and perhaps functionally, diverse forms of these cells may be present in the CNS. Our current understanding of astrocyte form and distribution is based predominately on studies that used the astrocytic marker glial fibrillary acidic protein (GFAP) and on studies using metal-impregnation techniques. The prevalent opinion, based on studies using these methods, is that astrocytic processes overlap extensively and primarily share the underlying neuropil. However, both of these techniques have serious shortcomings for visualizing the interactions among these structurally complex cells. In the present study, intracellular injection combined with immunohistochemistry for GFAP show that GFAP delineates only ∼15% of the total volume of the astrocyte. As a result, GFAP-based images have led to incorrect conclusions regarding the interaction of processes of neighboring astrocytes. To investigate these interactions in detail, groups of adjacent protoplasmic astrocytes in the CA1 stratum radiatum were injected with fluorescent intracellular tracers of distinctive emissive wavelengths and analyzed using three-dimensional (3D) confocal analysis and electron microscopy. Our findings show that protoplasmic astrocytes establish primarily exclusive territories. The knowledge of how the complex morphology of protoplasmic astrocytes affects their 3D relationships with other astrocytes, oligodendroglia, neurons, and vasculature of the brain should have important implications for our understanding of nervous system function.
Article
Full-text available
We have generated transgenic mice in which astrocytes are labeled by the enhanced green fluorescent protein (EGFP) under the control of the human glial fibrillary acidic protein (GFAP) promoter. In all regions of the CNS, such as cortex, cerebellum, striatum, corpus callosum, hippocampus, retina, and spinal cord, EGFP-positive cells with morphological properties of astrocytes could be readily visualized by direct fluorescence microscopy in living brain slices or whole mounts. Also in the PNS, nonmyelinating Schwann cells from the sciatic nerve could be identified by their bright green fluorescence. Highest EGFP expression was found in the cerebellum. Already in acutely prepared whole brain, the cerebellum appeared green-yellowish under normal daylight. Colabeling with GFAP antibodies revealed an overlap with EGFP in the majority of cells. Some brain areas, however, such as retina or hypothalamus, showed only low levels of EGFP expression, although the astrocytes were rich in GFAP. In contrast, some areas that were poor in immunoreactive GFAP were conspicuous for their EGFP expression. Applying the patch clamp technique in brain slices, EGFP-positive cells exhibited two types of membrane properties, a passive membrane conductance as described for astrocytes and voltage-gated channels as described for glial precursor cells. Electron microscopical investigation of ultrastructural properties revealed EGFP-positive cells enwrapping synapses by their fine membrane processes. EGFP-positive cells were negative for oligodendrocyte (MAG) and neuronal markers (NeuN). As response to injury, i.e., by cortical stab wounds, enhanced levels of EGFP expression delineated the lesion site and could thus be used as a live marker for pathology. GLIA 33:72–86, 2001. © 2000 Wiley-Liss, Inc.
Article
Calcium influx in nonexcitable cells regulates such diverse processes as exocytosis, contraction, enzyme control, gene regulation, cell proliferation, and apoptosis. The dominant Ca2+ entry pathway in these cells is the store-operated one, in which Ca2+ entry is governed by the Ca2+ content of the agonist-sensitive intracellular Ca2+ stores. Only recently has a Ca2+ current been described that is activated by store depletion. The properties of this new current, called Ca2+ release-activated Ca2+ current (ICRAC), have been investigated in detail using the patch-clamp technique. Despite intense research, the nature of the signal that couples Ca2+ store content to the Ca2+ channels in the plasma membrane has remained elusive. Although ICRAC appears to be the most effective and widespread influx pathway, other store-operated currents have also been observed. Although the Ca2+ release-activated Ca2+ channel has not yet been cloned, evidence continues to accumulate that the Drosophila trp gene might encode a store-operated Ca2+ channel. In this review, we describe the historical development of the field of Ca2+ signaling and the discovery of store-operated Ca2+ currents. We focus on the electrophysiological properties of the prototype store-operated current ICRAC, discuss the regulatory mechanisms that control it, and finally consider recent advances toward the identification of molecular mechanisms involved in this ubiquitous and important Ca2+ entry pathway.
Article
The careful regulation of calcium ion entry by the cell is essential for activation of numerous signal transduction cascades. It is well-established that the depletion of internal calcium ion stores such as those of the endoplasmic reticulum (ER) results in the opening of plasma membrane channels (SOCs) allowing calcium ions to flow into the cell. However, the mechanism by which calcium ion store depletion and the opening of SOCs are coupled is not clear. In a Perspective, [Berridge and colleagues][1] discuss new findings ([ Ma et al. ][2]) demonstrating that the inositol trisphosphate receptor in the ER membrane forms a physical connection with the SOCs in the plasma membrane. Inhibitors that block activation of inositol trisphosphate receptors prevent opening of SOCs in response to depletion of internal calcium stores. [1]: http://www.sciencemag.org/cgi/content/full/287/5458/1604 [2]: http://www.sciencemag.org/cgi/content/full/287/5458/1647
Article
Capacitative Ca2+ entry (CCE) activated by release/depletion of Ca2+ from internal stores represents a major Ca2+ influx mechanism in lymphocytes and other nonexcitable cells. Despite the importance of CCE in antigen-mediated lymphocyte activation, molecular components constituting this mechanism remain elusive. Here we demonstrate that genetic disruption of transient receptor potential (TRP)1 significantly attenuates both Ca2+ release-activated Ca2+ currents and inositol 1,4,5-trisphosphate (IP3)-mediated Ca2+ release from endoplasmic reticulum (ER) in DT40 B cells. As a consequence, B cell antigen receptor–mediated Ca2+ oscillations and NF-AT activation are reduced in TRP1-deficient cells. Thus, our results suggest that CCE channels, whose formation involves TRP1 as an important component, modulate IP3 receptor function, thereby enhancing functional coupling between the ER and plasma membrane in transduction of intracellular Ca2+ signaling in B lymphocytes.
Article
Considerable evidence, including recent direct observations, suggest that endoplasmic reticulum (ER) Ca2+ stores in neurons, glia, and other cell types, consists of spatially-distinct compartments that can be individually loaded and unloaded. In addition, sub-plasmalemmal (‘junctional’) components of the ER (jER) are functionally coupled to the overlying plasmalemmal (PL) microdomains in PL-jER units named ‘PLasmERosomes’. The PL microdomains and the jER contain clusters of specific transport proteins that regulate Na+ and Ca2+ concentrations in the tiny cytosolic space between the PL and jER. This organization helps the ER to produce the many types of complex local and global Ca2+ signals that are responsible for the simultaneous control of numerous neuronal and glial functions.
Article
Cytoplasmic calcium serves as a ubiquitous signal for acute cellular activation and for regulation of important cellular processes such as cell growth, division, differentiation, and even cell death and apoptosis. Increases in cytoplasmic calcium or calcium signals can be generated either by release of calcium from intracellular stores or by influx of calcium across the plasma membrane, but commonly by both means. The release of intracellular calcium comes from the endoplasmic reticulum or its specialized counterpart in muscle, the sarcoplasmic reticulum, and is generally signaled by the formation or influx of small mediators such as inositol 1,4,5-trisphosphate (IP3) (1), cyclic ADP ribose (2), and even calcium itself (3). Mechanisms for controlling calcium influx are somewhat more varied (4). Calcium influx can be more directly controlled by the membrane potential (5) or by the binding of extracellular neurotransmitters directly to the channels (6). Calcium influx can be signaled by second messengers such as cyclic nucleotides (7); however, in the case of receptors coupled to the phospholipase C pathway, the classical messengers of this pathway, IP3 and diacylglycerol, do not appear to be the primary mediators of activated calcium entry. Rather, in most instances, the signal for entry is somehow derived from the IP3-mediated depletion of calcium from intracellular stores, a process called “capacitative calcium entry” or “store-operated calcium entry” (8, 9).