ArticlePDF Available

Blocking L-type Voltage-gated Ca2+ Channels with Dihydropyridines Reduces -Aminobutyric Acid Type A Receptor Expression and Synaptic Inhibition

Authors:

Abstract

γ-Aminobutyric acid type A receptors (GABAARs) are the major sites of fast inhibitory neurotransmission in the brain, and the numbers of these receptors at the cell surface can determine the strength of GABAergic neurotransmission. Chronic changes in neuronal activity lead to an adaptive modulation in the efficacy of GABAergic synaptic inhibition, brought about in part by changes in the number of synaptic GABAARs, a mechanism known as homeostatic synaptic plasticity. Reduction in the number of GABAARs in response to prolonged neuronal activity blockade is dependent on the ubiquitin-proteasome system. The underlying biochemical pathways linking chronic activity blockade to proteasome-dependent degradation of GABAARs are unknown. Here, we show that chronic blockade of L-type voltage-gated calcium channels (VGCCs) with nifedipine decreases the number of GABAARs at synaptic sites but not the overall number of inhibitory synapses. In parallel, blockade of L-type VGCCs decreases the amplitude but not the frequency of miniature inhibitory postsynaptic currents or expression of the glutamic acid decarboxylase GAD65. We further reveal that the activation of L-type VGCCs regulates the turnover of newly translated GABAAR subunits in a mechanism dependent upon the activity of the proteasome and thus regulates GABAAR insertion into the plasma membrane. Together, these observations suggest that activation of L-type VGCCs can regulate the abundance of synaptic GABAARs and the efficacy of synaptic inhibition, revealing a potential mechanism underlying the homeostatic adaptation of fast GABAergic inhibition to prolonged changes in activity.
Blocking L-type Voltage-gated Ca
2
Channels with
Dihydropyridines Reduces
-Aminobutyric Acid Type A
Receptor Expression and Synaptic Inhibition
*
Received for publication, July 7, 2009, and in revised form, September 22, 2009 Published, JBC Papers in Press, September 24, 2009, DOI 10.1074/jbc.M109.040071
Richard S. Saliba
, Zhenglin Gu
§
, Zhen Yan
§
, and Stephen J. Moss
‡¶1
From the
Department of Neuroscience, Tufts University, Boston, Massachusetts 02111, the
§
Department of Physiology and
Biophysics, State University of New York at Buffalo, Buffalo, New York 14214, and the
Departments of Neuroscience, Physiology,
and Pharmacology, University College, London WC1E 6BT, United Kingdom
-Aminobutyric acid type A receptors (GABA
A
Rs) are the
major sites of fast inhibitory neurotransmission in the brain,
and the numbers of these receptors at the cell surface can deter-
mine the strength of GABAergic neurotransmission. Chronic
changes in neuronal activity lead to an adaptive modulation in
the efficacy of GABAergic synaptic inhibition, brought about in
part by changes in the number of synaptic GABA
A
Rs, a mecha-
nism known as homeostatic synaptic plasticity. Reduction in the
number of GABA
A
Rs in response to prolonged neuronal activity
blockade is dependent on the ubiquitin-proteasome system.
The underlying biochemical pathways linking chronic activity
blockade to proteasome-dependent degradation of GABA
A
Rs are
unknown. Here, we show that chronic blockade of L-type voltage-
gated calcium channels (VGCCs) with nifedipine decreases the
number of GABA
A
Rs at synaptic sites but not the overall number of
inhibitory synapses. In parallel, blockade of L-type VGCCs
decreases the amplitude but not the frequency of miniature inhib-
itory postsynaptic currents or expression of the glutamic acid
decarboxylase GAD65. We further reveal that the activation of
L-type VGCCs regulates the turnover of newly translated GABA
A
R
subunits in a mechanism dependent upon the activity of the pro-
teasome and thus regulates GABA
A
R insertion into the plasma
membrane. Together, these observations suggest that activation of
L-type VGCCs can regulate the abundance of synaptic GABA
A
Rs
and the efficacy of synaptic inhibition, revealing a potential mech-
anism underlying the homeostatic adaptation of fast GABAergic
inhibition to prolonged changes in activity.
-Aminobutyric acid type A receptors (GABA
A
Rs),
2
the
major sites of action for both benzodiazepines and barbiturates,
are Cl
-selective ligand-gated ion channels that can be assem-
bled from seven subunit classes (
1–6,
1–3,
1–3,
,
,
, and
), providing the structural basis for extensive heterogeneity of
GABA
A
R structure (1–3). A combination of molecular, bio-
chemical, and genetic approaches suggests that, in the brain,
the majority of benzodiazepine receptor subtypes are com-
posed of
,
, and
2 subunits (1).
2-containing receptors are
highly enriched at synaptic sites in neurons and are responsible
for mediating phasic inhibition (46). In contrast, receptors
composed of
,
, and
subunits are believed to form a special-
ized population of extrasynaptic receptors that mediate tonic
inhibition (7). GABA
A
Rs are assembled within the endoplasmic
reticulum (ER) and then transported to the plasma membrane
for insertion, whereas misfolded or unassembled receptor sub-
units are rapidly targeted for ER-associated degradation (5, 6, 8,
9), a process that can be modulated by neuronal activity (9). The
number of GABA
A
Rs on the neuronal cell surface is a critical
determinant for the efficacy of synaptic inhibition and, at steady
state, is determined by the rates of receptor insertion and
removal from the plasma membrane (10, 11). Cell-surface
GABA
A
Rs are dynamic entities that exhibit rapid rates of consti-
tutive endocytosis, with internalized receptors being subject to
rapid recycling or lysosomal degradation (5, 6, 10). Phosphoryla-
tion of the intracellular domains between transmembrane
domains 3 and 4 of the GABA
A
R
and
subunits by serine/
threonine and tyrosine kinases has been shown to alter receptor
function either by a direct effect on receptor properties, such as the
probability of channel opening or desensitization, or by regulating
trafficking of the receptor to and from the cell surface (11, 12).
The activity of neurons in neural circuits is highly regulated,
and when firing frequency either falls below or rises above nor-
mal physiological levels, compensatory mechanisms come into
play to restore normal activity, a process known as homeostatic
plasticity (13, 14). Homeostatic synaptic scaling is one homeo-
static mechanism that involves uniform adjustments in the
strength of all synapses in response to changes in activity while
maintaining the individual weights between synapses (15, 16).
GABA regulates the excitability of neural circuits, and a num-
ber of studies in cultures of dissociated neurons have provided
evidence that chronic changes in neuronal activity can lead to
homeostatic scaling of GABAergic synaptic strength (9, 17–19),
which appears to be dependent on brain-derived neurotrophic
factor and tumor necrosis factor
(19, 20). There is now increas-
ing evidence that homeostatic regulation of GABAergic synaptic
*This work was supported, in whole or in part, by National Institutes of Health
Grants NS047478, NS048045, NS051195, NS056359, and NS054900 from
NINDS (to S. J. M.) and by a fellowship from the American Society for Epi-
lepsy (to R. S. S.).
1
Consultant for Wyeth Pharmaceuticals. To whom correspondence should
be addressed: Dept. of Neuroscience, Tufts University, 136 Harrison Ave.,
Boston, MA 02111. Tel.: 617-636-3976; Fax: 617-636-2413; E-mail: stephen.
moss@tufts.edu.
2
The abbreviations used are: GABA
A
R,
-aminobutyric acid type A receptor;
ER, endoplasmic reticulum; VGCC, voltage-gated calcium channel;
-Bgt,
-bungarotoxin; DIV, days in vitro; PBS, phosphate-buffered saline; NHS,
N-hydroxysuccinimide; RIPA, radioimmune precipitation assay; pH,
pHluorin; BBS, bungarotoxin-binding site; mIPSC, miniature inhibitory
postsynaptic current; TTX, tetrodotoxin; NMDA, N-methyl-D-aspartic acid;
AMPA,
-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 284, NO. 47, pp. 32544 –32550, November 20, 2009
© 2009 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.
32544 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284NUMBER 47• NOVEMBER 20, 2009
by guest on December 28, 2015http://www.jbc.org/Downloaded from
strength is achieved by modulating the numbers of postsynaptic
GABA
A
Rs (9, 17, 19). Furthermore, presynaptic changes underly-
ing activity-dependent scaling of inhibitory synaptic strength have
been reported in which the levels of the GABA-synthesizing
enzyme GAD65, the GABA transporter, and presynaptic GABA
content are modulated by activity (17, 18, 21). Currently, little is
known about the mechanisms underlying the modulation of
synaptic GABA
A
R abundance in response to chronic changes
in neuronal activity. However, it has been recently established
that prolonged alteration in neuronal activity modulates the
ubiquitination and proteasomal degradation of GABA
A
Rs in
the ER, resulting in changes in the number of receptors at syn-
aptic sites and the efficacy of synaptic inhibition (9).
A number of important questions remain. For instance, how
do chronic changes in activity translate into the homeostatic
regulation of GABA
A
R abundance at synaptic sites? In this
study, we have begun to investigate the role of Ca
2
influx
through L-type voltage-gated calcium channels (VGCCs) in the
modulation of activity-dependent expression of GABA
A
Rs and
the efficacy of synaptic inhibition.
EXPERIMENTAL PROCEDURES
Antibodies—Rabbit anti-
3 IgG polyclonal antibodies have
been described previously (22, 23). Rabbit anti-green fluores-
cent protein IgG, mouse anti-synapsin IgG, and rabbit anti-
GAD65 antibodies were purchased from Synaptic Systems.
Peroxidase-conjugated and fluorescent dye-conjugated IgG
secondary antibodies were from Jackson ImmunoResearch
Laboratories. Fluorescently labeled
-bungarotoxin (
-Bgt)
was purchased from Invitrogen.
Neuronal Cell Culture and Transfections—Cultures of hip-
pocampal neurons were prepared from embryonic day 18 rats
(9, 22, 24). Dissociated embryonic day 18 rat cortical neurons
were transfected with 3
g of plasmid DNA/2 10
6
neurons
using the rat neuron Nucleofector
TM
kit (Lonza). 60-mm
dishes were seeded with 0.4 10
6
neurons and used in
experiments after 18–21 days in vitro (DIV).
Biotinylation—Hippocampal neurons were chilled on ice for
5 min and then washed twice with phosphate-buffered saline
(PBS) containing 1 mMCaCl
2
and 0.5 mMMgCl
2
(PBS-CM) at
4 °C. Cells were incubated for 15 min at 4 °C in 1 mg/ml sulfo-
N-hydroxysuccinimide (NHS)-biotin (Pierce) dissolved in PBS-
CM. To quench unreacted biotin, neurons were washed three
times (10 min each wash at 4 °C) with PBS-CM 75 mMglycine
and then washed twice with PBS and lysed in radioimmune
precipitation assay (RIPA) buffer (50 mMTris (pH 8), 150 mM
NaCl, 1% Nonidet P-40, 0.5% deoxycholate, 0.1% SDS, 2 mM
EDTA, and mammalian protease inhibitor mixture (Sigma)).
Protein concentrations were determined using the micro BCA
protein assay kit (Pierce), and equal amounts of solubilized pro-
tein were added to UltraLink-immobilized NeutrAvidin biotin-
binding protein (Pierce) for2hat4°C.Avidin beads were
washed twice for 15 min at 4 °C with high salt (500 mMNaCl)
RIPA buffer, followed by a 15-min wash at 4 °C with RIPA
buffer (150 mMNaCl). Precipitated biotinylated proteins and
total proteins were resolved by SDS-PAGE, and
3 was
detected by immunoblotting with rabbit anti-
3 IgG antibodies
(9, 22, 25), followed by peroxidase-conjugated anti-rabbit IgG
antibodies and detection with ECL. Blots were imaged using the
Fujifilm LAS-3000 imaging system, and bands were quantified
with Fujifilm Multi Gauge software.
Immunocytochemistry—Neurons expressing
pH
3 (where
pH is pHluorin; 18–21 DIV) were fixed in 4% paraformalde-
hyde, stained without membrane permeabilization with rabbit
anti-green fluorescent protein IgG antibodies, and then perme-
abilized with 0.1% Triton X-100 for 4 min. Neurons were
labeled with anti-synapsin IgG antibodies, visualized by confo-
cal microscopy, and analyzed using MetaMorph imaging soft-
ware (Molecular Devices). To quantify the fluorescence inten-
sity of cell-surface
pH
3 synaptic staining, images of neurons
were thresholded to a point at which dendrites were outlined.
Synapsin staining was thresholded to a set value and kept con-
stant for control and test neurons. Next, a 25-
m section along
a given proximal dendrite was selected, and a 1-bit binary image
(exclusive) was made of the synapsin staining in the outlined
dendrite. We then subtracted all
pH
3 staining that did not
co-localize with the binarized synapsin staining. As a result,
only
BBS
3 staining that co-localized with synapsin remained,
and the average fluorescence intensity of these
pH
3 puncta was
determined (9). Data were analyzed from 10–12 neurons for
each condition from at least two to three different cultures. To
quantify the number of
pH
3 synapses, thresholds were set and
kept constant for control and test neurons. Receptor clusters
were defined as being 0.5–2
m in length and 2–3-fold more
intense than background diffuse fluorescence and were co-lo-
calized with synapsin staining (9, 26).
pH
3 synaptic puncta
were counted from 1-bit binary masks. Data were analyzed
from 10–12 neurons for each condition (25
m/dendrite/cell).
Analyses were all performed blind to experimental condition.
GABA
A
R
BBS
3 Insertion Assay—Hippocampal neurons
(18–21 DIV) expressing
BBS
3 were first labeled with 10
g/ml
unlabeled
-Bgt for 15 min at 15 °C to block existing cell-sur-
face receptors. The neurons were then washed three times with
PBS at 15 °C, followed by a 5-min incubation at 37 °C with 1
g/ml Alexa 594-conjugated
-Bgt (9). All incubations were
performed in the presence of 200
Mtubocurarine (Sigma) to
block
-Bgt binding to endogenous acetylcholine receptors (25,
27–29). Cells were fixed in 4% paraformaldehyde. Confocal
images were collected using a 60 objective lens acquired with
Olympus FluoView Version 1.5 software, and the same image
acquisition settings for
BBS
3
with or without nifedipine were
used. These images were analyzed using MetaMorph imaging
software. A three-dimensional reconstruction of an imaged
neuron was made from a series of Zsections, and then the
average fluorescence intensity of Alexa 594-conjugated
-Bgt
staining was measured along 30
m of two proximal dendrites/
neuron after subtraction of background fluorescence.
Metabolic Labeling and Immunoprecipitation—Hippocam-
pal neurons were incubated in methionine-free Dulbecco’s
modified Eagle’s medium for 15 min and then labeled with 500
Ci/ml [
35
S]methionine (PerkinElmer Life Sciences) for 30
min. Neurons were washed and incubated in complete Neuro-
basal medium with an excess of unlabeled methionine (100-
fold) for an additional 04 h. Neurons were lysed in 1% SDS
and 25 mMTris (pH 7.4), and lysates were diluted 10-fold with
RIPA buffer lacking SDS (50 mMTris (pH 8), 150 mMNaCl, 1%
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
NOVEMBER 20, 2009VOLUME 284• NUMBER 47 JOURNAL OF BIOLOGICAL CHEMISTRY 32545
by guest on December 28, 2015http://www.jbc.org/Downloaded from
Nonidet P-40, 0.5% sodium deoxycholate, 4 mMEDTA, and
mammalian protease inhibitor mixture). GABA
A
R
3 subunits
were immunoprecipitated with rabbit anti-
3 IgG antibodies
from equal amounts of solubilized protein as detailed previ-
ously (9, 10, 22, 30). Precipitated material was then subjected to
SDS-PAGE, and
3 band intensities were determined by phos-
phorimage spectrometry (Bio-Rad).
Electrophysiological Recordings—Patch-clamp recordings in
the whole cell mode were used to measure the properties of
miniature inhibitory postsynaptic currents (mIPSCs) in cul-
tured hippocampal neurons. mIPSCs were isolated by the
inclusion of tetrodotoxin (TTX; 0.5
M), D-2-amino-5-phos-
phopentanoic acid (20
M), and 6,7-dinitroquinoxaline-2,3-di-
one (20
M) to block action potentials, N-methyl-D-aspartic
acid (NMDA), and
-amino-3-hydroxy-5-methyl-4-iso-
xazolepropionic acid (AMPA)/kainate receptors, respectively,
as detailed previously (9). The cell membrane potential was
held at 70 mV. A mini analysis program (Synaptosoft, Leonia,
NJ) was used to analyze the spontaneous synaptic events. Sta-
tistical comparisons of the amplitude and frequency of mIPSCs
were made using Student’s ttest. The threshold for detection of
mIPSC amplitude was 15 pA.
RESULTS
Blocking Ca
2
Influx through L-type VGCCs Reduces the
Expression Levels of GABA
A
Rs—The synaptic expression levels
of GABA
A
Rs are dynamically regulated in response to chronic
changes in neuronal activity (9, 17, 19). Furthermore, bidirec-
tional changes in neuronal activity modulate the ubiquitin-de-
pendent proteasomal degradation of GABA
A
Rs (9). Given that
L-type VGCCs in neurons couple membrane depolarization to
numerous processes, including gene expression (31) and
changes in synaptic efficacy (32, 33), we speculated that Ca
2
influx through L-type VGCCs may play a role in regulating
activity-dependent expression of GABA
A
Rs. To determine the
effects of blocking Ca
2
influx through L-type VGCCs on
GABA
A
R expression, we used a class of compounds known as
dihydropyridines, which block L-type VGCCs (34). Cultured
hippocampal neurons (18–21 DIV) were incubated for 24 h
with or without 10
Mnifedipine. We chose a time point of 24 h
given that homeostatic scaling of GABAergic synaptic strength
occurs over hours to days (9, 17, 19). Cell-surface proteins were
labeled with Sulfo-NHS-biotin and isolated using immobilized
avidin, and precipitated cell-surface proteins were then
resolved by SDS-PAGE. GABA
A
R
3 subunits were detected by
immunoblotting with rabbit anti-
3 IgG antibodies. Nifedipine
reduced the cell-surface expression and the total pool of
GABA
A
R
3 subunits by 46 7.3 and 45 3.9%, respectively,
compared with control levels (Fig. 1A).
FIGURE 1. Blockade of L-type VGCCs reduces expression of GABA
A
Rs.
A, hippocampal neurons (18 –21 DIV) were treated with 10
Mnifedipine (Nif)
for 24 h and then biotinylated with Sulfo-NHS-biotin and lysed in RIPA buffer.
Cell-surface proteins were isolated with immobilized avidin. Immunoblots
show the total and cell-surface levels of GABA
A
R
3 subunits as indicated.
Graphs represent quantification of band intensities of cell-surface and total
3 subunits normalized to controls (Ctrl). Data represent the mean S.E.
percentage of control values. *, significantly different from the control (p
0.01; ttest; n4). B, nimodipine reduced the expression of the cell-surface
and total levels of GABA
A
R
3 subunits. Hippocampal neurons were incu-
bated with 15
Mnimodipine for 24 h and then biotinylated with Sulfo-NHS-
biotin and lysed in RIPA buffer. Immunoblots show the cell-surface and total
levels of GABA
A
R
3 subunits as indicated. Graphs represent quantification of
band intensities, and data represent the mean S.E. percentage of control
values. *, significantly different from the control (p0.05; ttest; n3).
FIGURE 2. Blockade of L-type VGCCs reduces the synaptic expression of
3-containing GABA
A
Rs. A, images of hippocampal neurons (18 –21 DIV)
expressing GABA
A
R
pH
3 treated with or without nifedipine (Nif;10
M) for
24 h as indicated. Neurons were fixed, and non-permeabilized cells were
stained with rabbit anti-green fluorescent protein IgG antibodies and Rhoda-
mine Red-X-conjugated anti-rabbit IgG antibodies to label cell-surface
pH
3
subunits (shown in red). pHluorin fluorescence is not shown. Cells were then
permeabilized and incubated with mouse anti-synapsin IgG antibodies and
Cy5-conjugated anti-mouse IgG antibodies to label synaptic sites (shown in
green). The boxed areas in the left panels are magnified in the right panels.
Scale bars 10
m. B, quantification of the fluorescence intensity of
pH
3at
synaptic sites. Data represent the mean S.E. percentage of control (Ctrl)
values. *, significantly different from the control (p0.001; ttest; n10 –12
neurons in two independent experiments). C, quantification of the number of
synaptic sites containing GABA
A
R
pH
3 subunits (no significant difference;
n10 –12 neurons in two independent experiments).
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
32546 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284NUMBER 47• NOVEMBER 20, 2009
by guest on December 28, 2015http://www.jbc.org/Downloaded from
We also determined the effects of another member of the
dihydropyridine class of compounds on GABA
A
R expression
levels, nimodipine, which is cell-impermeable, unlike nifedip-
ine. These experiments revealed that a 24-h treatment with
nimodipine (15
M) reduced the cell-surface and total levels of
GABA
A
R
3 subunits by 42 10.8 and 41 6.2%, respectively,
compared with control levels (Fig. 1B).
To determine whether GABA
A
Rs localized at synaptic sites
were also regulated by Ca
2
influx through L-type VGCCs, we
used a construct containing superecliptic pHluorin encoded
within the N terminus of the GABA
A
R
3 subunit (
pH
3) in our
experiments (9, 26, 29). Embryonic day 18 hippocampal neu-
rons were nucleofected with
pH
3 and cultured for 18 –21 DIV.
Neurons were then incubated with nifedipine (10
M) for 24 h and
fixed in 4% paraformaldehyde. To label only the cell-surface pop-
ulation of
pH
3, rabbit anti-green fluorescent protein IgG antibod-
ies were used on non-permeabilized neurons, and synaptic sites
were labeled with mouse IgG antibodies to the presynaptic marker
synapsin-1 following permeabilization (Fig. 2A). We chose to stain
synapsin because the levels of this synaptic marker are unaltered
by chronic changes in neuronal activity (17). Images were collected
using confocal microscopy, and based on their morphology, only
pyramidal cells were chosen. These studies revealed that nifedip-
ine reduced the levels of
pH
3-containing GABA
A
Rs at synaptic
sites by 47.2 6.15% compared with control levels (Fig. 2B). It
should be noted that the nifedipine-induced reduction in the
expression levels of total and synaptic GABA
A
R
3 subunits is
approximately equal to the reduction observed following chronic
blockade of neuronal activity with TTX (9).
The number of GABAergic inhibitory synapses in cultured
neurons is modulated by activity blockade (9, 17). Therefore, we
determined whether blocking Ca
2
influx had any effect on the
number of synaptic sites containing GABA
A
Rs. We used the
synaptic marker synapsin-1 to label presynaptic sites in neu-
rons expressing
pH
3 following treatment with nifedipine for
24 h. We observed no change in the number of synaptic sites
containing
pH
3 (Fig. 2C). Collectively, these data suggest that
Ca
2
influx through L-type VGCCs modulates the total pool
and the synaptic expression levels of GABA
A
Rs but not the
number of GABAergic synapses.
Blocking Ca
2
Influx through L-type VGCCs Reduces the Effi-
cacy of Synaptic Inhibition—Given the robust effects of block-
ing L-type VGCCs on the synaptic accumulation of GABA
A
R
3 subunits, the effect of nifedipine on the efficacy of synaptic
inhibition was determined. Analysis of mIPSCs in cultured hip-
pocampal neurons (18–21 DIV) was performed following nife-
FIGURE 3. Blockade of L-type VGCCs reduces GABAergic synaptic transmission. Aand B, representative mIPSC traces (A) and cumulative distribution of
mIPSC amplitude (B) from control and nifedipine-treated (10
M, 24 h) hippocampal neurons (18 –21 DIV). The threshold for mIPSC detection was 15 pA.
C, scatter plot of mIPSC amplitudes from control and nifedipine-treated neurons. The mean amplitude in each neuron was the average of all the mIPSC events
above 15 pA. D, bar graph of the average (mean S.E.) mIPSC amplitude and frequency in control (n16) and nifedipine-treated (n16) neurons. *,
significantly different from the control (p0.001; ttest).
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
NOVEMBER 20, 2009VOLUME 284• NUMBER 47 JOURNAL OF BIOLOGICAL CHEMISTRY 32547
by guest on December 28, 2015http://www.jbc.org/Downloaded from
dipine treatment (10
Mfor 24 h) using patch-clamp recording
in the whole cell mode as detailed previously (9). mIPSCs were
recorded from cultured hippocampal pyramidal neurons in the
presence of TTX (500 nM) and in the presence of D-2-amino-5-
phosphonopentanoate and 6,7-dinitroquinoxaline-2,3-dione
to inhibit neuronal depolarization and ionotropic glutamate
receptor activation, respectively. mIPSC amplitudes and fre-
quencies were compared between control neurons and those
treated with nifedipine (Fig. 3, A–D). The data revealed that
mIPSC amplitude was reduced from 57.6 3.6 pA (n16) in
control neurons to 38.3 1.9 pA (n16) in nifedipine-treated
neurons, which represents a 33.5% decrease (Fig. 3, Cand D).
Also, nifedipine shifted the entire distribution of amplitudes to
the left, toward smaller values (Fig. 3B), suggesting that there is
a uniform reduction in synaptic inhibition. Nifedipine treat-
ment did not significantly alter mIPSC frequency (control,
1.68 0.17 Hz (n16); nifedipine-treated, 1.35 0.23 Hz (n
16)). Overall, the above data suggest that the efficacy of synaptic
inhibition can be modulated by prolonged blockade of Ca
2
influx through L-type VGCCs.
Previous reports have shown that chronic changes in neuro-
nal activity regulate the expression levels of GAD65 (17, 18), the
main enzyme involved in GABA synthesis, leading to reduced
levels of GABA and decreased synaptic inhibition (18). We
therefore tested whether Ca
2
influx through L-type VGCCs
had any effect on GAD65 expression levels. We treated hip-
pocampal neurons with 10
Mnifedipine for 24 h and deter-
mined the expression levels of GAD65 by Western blotting. We
observed no change in the expression levels of GAD65 (Fig. 4A)
but found reduced steady-state levels of GAD65 when neurons
were treated with 1
MTTX for 24 h (Fig. 4B), in agreement
with previous studies (17, 18). These data suggest that Ca
2
influx through L-type VGCCs does not regulate GAD65
expression levels and consequently suggest that GABA synthe-
sis is unaffected under these conditions.
Ca
2
Influx through L-type VGCCs Regulates the Proteasome-
dependent Turnover and Membrane Insertion of GABA
A
Rs
Given that the steady-state levels of GABA
A
Rs were reduced by
nifedipine, we tested whether Ca
2
influx through L-type
VGCCs directly influenced the turnover of newly synthesized
GABA
A
R
3 subunits within the ER/secretory pathway. To
assess this, we used metabolic labeling with [
35
S]methionine in
pulse-chase experiments. Hippocampal neurons were treated
with nifedipine for 24 h and then labeled with [
35
S]methionine
for 30 min and chased for 4 h (Fig. 5A). Therefore, at the time
point we chose (4 h post-labeling),
3-containing GABA
A
Rs
would not have yet reached the cell surface because they take up
to6htoreach this compartment from the ER (8). Following
lysis of neurons, GABA
A
R
3 subunits were immunoprecipi-
tated with anti-
3 IgG antibodies and resolved by SDS-PAGE.
At 4 h, a 26.7 1.94% decrease in
3 subunits was observed in
FIGURE 4. GAD65 expression is unaltered by L-type VGCC blockade.
A, hippocampal neurons (18 –21 DIV) were treated with 10
Mnifedipine (Nif)
for 24 h and then lysed in RIPA buffer. Western blotting was performed with
rabbit anti-GAD65 IgG antibodies. The graph represents quantification of
band intensities, and data represent the mean S.E. of control (Ctrl) values.
No significant difference from the control was observed (ttest; n4). B, neu-
ronal activity blockade with TTX reduced GAD65 expression. Hippocampal
neurons were treated with 1
MTTX for 24 h and then lysed in RIPA buffer.
Western blotting was performed with rabbit anti-GAD65 IgG antibodies. The
graph represents quantification of band intensities, and data represent
the mean S.E. percentage of control values. *, significantly different from
the control (p0.05; ttest; n3).
FIGURE 5. Blockade of L-type VGCCs increases the turnover of GABA
A
Rs.
A, hippocampal neurons (18 –21 DIV) were treated with 10
Mnifedipine for
24 h, followed by a pulse chase with [
35
S]methionine. Neurons were lysed in
1% SDS and diluted in RIPA buffer, and GABA
A
R
3 subunits were immuno-
precipitated with anti-
3 IgG antibodies or nonspecific IgG antibodies (as
indicated) and subjected to SDS-PAGE. Band intensities were quantified by
phosphorimage spectrometry, and data represent the mean S.E. percent-
age of levels at time 0. *, significantly different from0h(p0.05; ttest; n3).
B, inhibition of proteasome activity blocked the effects of nifedipine on
GABA
A
R
3 expression. Hippocampal neurons (18 –21 DIV) were treated with
or without 10
Mnifedipine for 24 h, and 10
MMG132 was added for the last
8 h of the nifedipine incubation as indicated. Neurons were lysed, and West-
ern blots were probed with anti-
3 IgG antibodies. Data represent the
mean S.E. percentage of control (Ctrl) values. *, significantly different (p
0.05); **, significantly different (p0.01; one-way analysis of variance and
Bonferroni post-test; n3).
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
32548 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284NUMBER 47• NOVEMBER 20, 2009
by guest on December 28, 2015http://www.jbc.org/Downloaded from
neurons previously treated with nifedipine, whereas in control
neurons at 4 h, there was no significant degradation (Fig. 5A).
These data implicate Ca
2
influx through L-type VGCCs in
modulating ER-associated degradation of de novo synthesized
GABA
A
R
3 subunits.
We next tested whether proteasome activity was involved in
the increased turnover of GABA
A
Rs following blockade of
L-type VGCCs in cultured hippocampal neurons. We treated
neurons with nifedipine (10
M) for 24 h and added the protea-
some inhibitor MG132 (10
M) during the last8hofthenife-
dipine treatment. These experiments revealed that the effects
of nifedipine on the steady-state levels of GABA
A
R
3 subunits
were blocked by MG132 (Fig. 5B). Nifedipine decreased the
expression levels of
3 subunits by 38 4.6% compared with
control levels, whereas the addition of MG132 brought
3 sub-
unit levels up to 85 5.77% of control levels (Fig. 5B). Together,
these data suggest that Ca
2
influx through L-type VGCCs can
directly regulate the turnover of newly synthesized GABA
A
Rs
and that proteasome activity is required for this process.
As activity-dependent L-type Ca
2
influx modulates the
turnover of GABA
A
Rs, we speculated that this may influence
the rate of insertion of receptors into the plasma membrane. To
test this, we used a construct that encodes a bungarotoxin-
binding site and pHluorin tag within the N terminus of the
GABA
A
R
3 subunit (
BBS
3) in an insertion assay as described
previously (9, 29). Hippocampal neurons (18 –21 DIV) express-
ing
BBS
3 were incubated with 10
Mnifedipine for 24 h. Neu-
rons were then incubated with excess unlabeled
-Bgt at 15 °C
to block existing receptors and then incubated at 37 °C with
Alexa 594-conjugated
-Bgt for a number of time points (0–10
min) to label newly inserted
BBS
3 at the plasma membrane
(Fig. 6A). Excess unlabeled
-Bgt blocked accumulation of
Alexa 594-conjugated
-Bgt staining on the plasma membrane,
and
-tubocurarine (200
M) was added to block
-Bgt binding
to endogenous acetylcholine receptors. At different time
points, neurons were fixed and imaged by confocal microscopy.
The insertion of
BBS
3 increased linearly over a 10-min time
period, which was reduced by nifedipine treatment (Fig. 6B).
We therefore used a time point of 5 min to assess the influence
of Ca
2
influx on the insertion of
BBS
3 and normalized data to
those seen in untreated neurons, which were assigned a value of
100% (Fig. 6C). These experiments revealed that blocking Ca
2
influx with nifedipine significantly reduced the insertion of
BBS
3-containing GABA
A
Rs into the membrane by 52.7 5%
compared with control levels (Fig. 6C).
DISCUSSION
In this study, we have shown that prolonged blockade of
Ca
2
influx through L-type VGCCs leads to reduced total and
synaptic expression of GABA
A
Rs and diminished GABAergic
synaptic transmission without affecting the levels of GAD65.
Reduced GABA
A
R expression following L-type VGCC block-
ade is a result of increased turnover of de novo synthesized
GABA
A
Rs, which is dependent on the activity of the protea-
some. In addition, chronic blockade of Ca
2
influx reduces the
insertion of GABA
A
Rs into the neuronal membrane. Overall,
prolonged reduction in L-type Ca
2
influx decreases the num-
bers of synaptic GABA
A
Rs and reduces synaptic inhibition,
highlighting a possible role for L-type VGCCs in the activity-
dependent scaling of GABAergic synaptic strength.
The number of GABA
A
Rs is a critical factor in determining
the strength of GABAergic synaptic inhibition (35). It has been
previously reported that prolonged changes in neuronal activity
lead to an adaptive modulation of the numbers of GABA
A
Rs at
synaptic sites and the efficacy of synaptic inhibition (9, 17, 19,
36). Therefore, a critical expression locus of activity-dependent
scaling of GABAergic synaptic strength is the postsynaptic
change in GABA
A
R numbers. Furthermore, the direct ubiquiti-
nation of GABA
A
Rs and the ubiquitin-proteasome system play
a critical role in the activity-dependent modulation of the abun-
dance of these receptors at synaptic sites (9). In this study, we
have shown that blocking Ca
2
influx through L-type VGCCs
in cultured hippocampal neurons mimics the effects of TTX on
GABA
A
R turnover, accumulation at synaptic sites, and efficacy
of synaptic inhibition (9). These observations suggest that
activity-dependent scaling of GABAergic synaptic strength is
mediated in part by Ca
2
influx through L-type VGCCs. Inter-
estingly, a recent report has shown that the increase in AMPA
receptor number brought about by activity blockade with TTX
is mediated by somatic Ca
2
influx and that TTX prevents
these action potential-triggered Ca
2
transients in the cell
soma (37). In addition, the effects of TTX on AMPA receptor
accumulation could be partially mimicked by blocking L-type
VGCCs, but the authors speculate that other VGCCs (T- and
R-type) contribute to this phenomenon (37). Furthermore,
Ca
2
influx through L-type VGCCs is critical for activity-de-
pendent changes in the composition of AMPA receptors (33).
Thus, these studies highlight the importance of L-type Ca
2
FIGURE 6. Insertion of GABA
A
Rs is modulated by Ca
2
influx through
L-type VGCCs. A, images of pHluorin fluorescence and Alexa 594-conjugated
-Bgt staining at 5 min of insertion in the presence or absence of nifedipine (Nif;
10
M) as indicated. Hippocampal neurons expressing
BBS
3 (18 –21 DIV) were
treated with or without nifedipine (10
M) for 24 h. Neurons were then incubated
with 10
g/ml unlabeled
-Bgt to block existing cell-surface
BBS
3, washed, incu-
bated for 5 min with 1
g/ml Alexa 594-conjugated
-Bgt to label newly inserted
BBS
3, and fixed. The boxed areas in the upper right-hand corners show pHluorin
fluorescence. The rectangles in the left panels (Alexa 594-conjugated
-Bgt stain-
ing) are enlarged in the right panels.B, increase in
BBS
3 insertion over time with
or without nifedipine (10
M). The assay was performed as described for A.
C, graph showing quantification of Alexa 594-conjugated
-Bgt fluorescence
intensity at 5 min of insertion for control (Ctrl) and nifedipine (10
M)-treated
neurons. Data represent the mean S.E. percentage of control
BBS
3 values. *,
significantly different from the control (p0.01; ttest; nsix to eight neurons
from two independent cultures).
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
NOVEMBER 20, 2009VOLUME 284• NUMBER 47 JOURNAL OF BIOLOGICAL CHEMISTRY 32549
by guest on December 28, 2015http://www.jbc.org/Downloaded from
influx in the homeostatic scaling of glutamatergic synaptic
strength.
In addition to the changes in abundance of GABA
A
Rs, a
number of presynaptic factors also determine GABAergic syn-
aptic strength. We did not observe a change in GAD65 expres-
sion or the frequency of mIPSCs when L-type Ca
2
influx was
blocked. This is in stark contrast to the presynaptic changes
that have been reported when neuronal activity was blocked
with TTX. Activity blockade in cultured neurons with TTX
induces down-regulation of GAD65 and vesicular inhibitory
amino acid transporter expression (17–19, 21), reduces the
number of GABAergic synapses (9, 17), and results in a smaller
mIPSC frequency (9, 17). Thus, Ca
2
influx through L-type
VGCCs may be responsible for only mediating activity-depen-
dent postsynaptic changes in GABAergic synaptic strength.
However, it appears that other mechanisms regulate presynap-
tic factors governing the homeostatic scaling of GABAergic
synaptic strength. Perhaps activity-dependent changes in the
expression of GAD65 are mediated by Ca
2
influx through
NMDA receptors and not L-type VGCCs, as one report has
shown that NMDA receptor activation dominates Ca
2
influx
in interneurons (38) and that chronic blockade of NMDA
receptors reduces GAD67 expression (39).
This study has begun to delineate the biochemical pathways
underlying the scaling of the strength of GABAergic synaptic
neurotransmission. However, there are a number of unanswered
questions. For instance, we have revealed that Ca
2
influx can
regulate the turnover of GABA
A
Rs, but the exact processes leading
to the modulation of receptor turnover and insertion still remain
to be determined but could possibly involve Ca
2
-dependent
phosphorylation of GABA
A
R subunits and subsequent modula-
tion of receptor ER export/degradation.
In conclusion, this study has demonstrated the importance of
L-type Ca
2
influx in regulating the abundance of GABA
A
Rs at
synaptic sites and the efficacy of synaptic inhibition. Further-
more, this work emphasizes the importance of Ca
2
signaling
in mediating activity-dependent scaling of GABAergic synaptic
strength.
REFERENCES
1. Rudolph, U., and Mo¨ hler, H. (2004) Annu. Rev. Pharmacol. Toxicol. 44,
475–498
2. Rudolph, U., and Mo¨ hler, H. (2006) Curr. Opin. Pharmacol. 6, 18–23
3. Sieghart, W., and Sperk, G. (2002) Curr. Top. Med. Chem. 2, 795–816
4. Essrich, C., Lorez, M., Benson, J. A., Fritschy, J. M., and Lu¨ scher, B. (1998)
Nat. Neurosci. 1, 563–571
5. Kittler, J. T., and Moss, S. J. (2003) Curr. Opin. Neurobio.l 13, 341–347
6. Lu¨ scher, B., and Keller, C. A. (2004) Pharmacol. Ther. 102, 195–221
7. Farrant, M., and Nusser, Z. (2005) Nat. Rev. Neurosci. 6, 215–229
8. Gorrie, G. H., Vallis, Y., Stephenson, A., Whitfield, J., Browning, B., Smart,
T. G., and Moss, S. J. (1997) J. Neurosci. 17, 6587–6596
9. Saliba, R. S., Michels, G., Jacob, T. C., Pangalos, M. N., and Moss, S. J.
(2007) J. Neurosci. 27, 13341–13351
10. Kittler, J. T., Thomas, P., Tretter, V., Bogdanov, Y. D., Haucke, V., Smart, T. G.,
and Moss, S. J. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 12736–12741
11. Jacob, T. C., Moss, S. J., and Jurd, R. (2008) Nat. Rev. Neurosci. 9, 331–343
12. Moss, S. J., and Smart, T. G. (2001) Nat. Rev. Neurosci. 2, 240–250
13. Turrigiano, G. G., and Nelson, S. B. (2004) Nat. Rev. Neurosci. 5, 97–107
14. Burrone, J., and Murthy, V. N. (2003) Curr. Opin. Neurobiol. 13, 560–567
15. Turrigiano, G. G., Leslie, K. R., Desai, N. S., Rutherford, L. C., and Nelson,
S. B. (1998) Nature 391, 892–896
16. Turrigiano, G. G. (2008) Cell 135, 422–435
17. Kilman, V., van Rossum, M. C., and Turrigiano, G. G. (2002) J. Neurosci.
22, 1328–1337
18. Hartman, K. N., Pal, S. K., Burrone, J., and Murthy, V. N. (2006) Nat.
Neurosci. 9, 642–649
19. Swanwick, C. C., Murthy, N. R., and Kapur, J. (2006) Mol. Cell. Neurosci.
31, 481–492
20. Stellwagen, D., and Malenka, R. C. (2006) Nature 440, 1054–1059
21. De Gois, S., Scha¨fer, M. K., Defamie, N., Chen, C., Ricci, A., Weihe, E.,
Varoqui, H., and Erickson, J. D. (2005) J. Neurosci. 25, 7121–7133
22. Jovanovic, J. N., Thomas, P., Kittler, J. T., Smart, T. G., and Moss, S. J.
(2004) J. Neurosci. 24, 522–530
23. Brandon, N. J., Jovanovic, J. N., Colledge, M., Kittler, J. T., Brandon, J. M.,
Scott, J. D., and Moss, S. J. (2003) Mol. Cell. Neurosci. 22, 87–97
24. Kittler, J. T., Delmas, P., Jovanovic, J. N., Brown, D. A., Smart, T. G., and
Moss, S. J. (2000) J. Neurosci. 20, 7972–7977
25. Saliba, R. S., Pangalos, M., and Moss, S. J. (2008) J. Biol. Chem. 283,
18538–18544
26. Jacob, T. C., Bogdanov, Y. D., Magnus, C., Saliba, R. S., Kittler, J. T., Hay-
don, P. G., and Moss, S. J. (2005) J. Neurosci. 25, 10469–10478
27. Pedersen, S. E., and Cohen, J. B. (1990) Proc. Natl. Acad. Sci. U.S.A. 87,
2785–2789
28. Sekine-Aizawa, Y., and Huganir, R. L. (2004) Proc. Natl. Acad. Sci. U.S.A.
101, 17114–17119
29. Bogdanov, Y., Michels, G., Armstrong-Gold, C., Haydon, P. G., Lindstrom,
J., Pangalos, M., and Moss, S. J. (2006) EMBO J. 25, 4381–4389
30. Brandon, N. J., Delmas, P., Kittler, J. T., McDonald, B. J., Sieghart, W.,
Brown, D. A., Smart, T. G., and Moss, S. J. (2000) J. Biol. Chem. 275,
38856–38862
31. West, A. E., Griffith, E. C., and Greenberg, M. E. (2002) Nat. Rev. Neurosci.
3, 921–931
32. Grover, L. M., and Teyler, T. J. (1990) Nature 347, 477–479
33. Thiagarajan, T. C., Lindskog, M., and Tsien, R. W. (2005) Neuron 47,
725–737
34. Tsien, R. W., and Tsien, R. Y. (1990) Annu. Rev. Cell Biol. 6, 715–760
35. Nusser, Z., Cull-Candy, S., and Farrant, M. (1997) Neuron 19, 697–709
36. Nusser, Z., Ha´jos, N., Somogyi, P., and Mody, I. (1998) Nature 395,
172–177
37. Ibata, K., Sun, Q., and Turrigiano, G. G. (2008) Neuron 57, 819–826
38. Goldberg, J. H., Yuste, R., and Tamas, G. (2003) J. Physiol. 551, 67–78
39. Kinney, J. W., Davis, C. N., Tabarean, I., Conti, B., Bartfai, T., and Behrens,
M. M. (2006) J. Neurosci. 26, 1604–1615
Dihydropyridines Reduce GABA
A
R Expression/Synaptic Inhibition
32550 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284NUMBER 47• NOVEMBER 20, 2009
by guest on December 28, 2015http://www.jbc.org/Downloaded from
Stephen J. Moss
Richard S. Saliba, Zhenglin Gu, Zhen Yan and
Expression and Synaptic Inhibition
-Aminobutyric Acid Type A Receptor γChannels with Dihydropyridines Reduces
2+
Blocking L-type Voltage-gated Ca
and Biogenesis:
Membrane Transport, Structure, Function,
doi: 10.1074/jbc.M109.040071 originally published online September 24, 2009
2009, 284:32544-32550.J. Biol. Chem.
10.1074/jbc.M109.040071Access the most updated version of this article at doi:
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/284/47/32544.full.html#ref-list-1
This article cites 39 references, 14 of which can be accessed free at
by guest on December 28, 2015http://www.jbc.org/Downloaded from
... Mechanistically, VGCCs can impact downstream receptor trafficking and receptor expression levels. L-VGCCs regulate GABA A R synaptic abundance by reducing proteosomal degradation and enhancing exocytosis of newly translated GABA A R (Saliba et al., 2009) via CaMKII phosphorylation of β3 at S383 (Saliba et al., 2012). On the other hand, L-VGCCs reduce nascent transcription of α1-GABA A Rs with 48-h DZP (Foitzick et al., 2020) and mediate enhanced GABA A R diffusion during chronic depolarization at the axon initial segment (Muir and Kittler, 2014). ...
Article
Full-text available
Synaptic plasticity is a critical process that regulates neuronal activity by allowing neurons to adjust their synaptic strength in response to changes in activity. Despite the high proximity of excitatory glutamatergic and inhibitory GABAergic postsynaptic zones and their functional integration within dendritic regions, concurrent plasticity has historically been underassessed. Growing evidence for pathological disruptions in the excitation and inhibition (E/I) balance in neurological and neurodevelopmental disorders indicates the need for an improved, more “holistic” understanding of synaptic interplay. There continues to be a long-standing focus on the persistent strengthening of excitation (excitatory long-term potentiation; eLTP) and its role in learning and memory, although the importance of inhibitory long-term potentiation (iLTP) and depression (iLTD) has become increasingly apparent. Emerging evidence further points to a dynamic dialogue between excitatory and inhibitory synapses, but much remains to be understood regarding the mechanisms and extent of this exchange. In this mini-review, we explore the role calcium signaling and synaptic crosstalk play in regulating postsynaptic plasticity and neuronal excitability. We examine current knowledge on GABAergic and glutamatergic synapse responses to perturbances in activity, with a focus on postsynaptic plasticity induced by short-term pharmacological treatments which act to either enhance or reduce neuronal excitability via ionotropic receptor regulation in neuronal culture. To delve deeper into potential mechanisms of synaptic crosstalk, we discuss the influence of synaptic activity on key regulatory proteins, including kinases, phosphatases, and synaptic structural/scaffolding proteins. Finally, we briefly suggest avenues for future research to better understand the crosstalk between glutamatergic and GABAergic synapses.
... Chronic changes in neuronal activity can lead to alterations in the number of synaptic GABA A receptors, a phenomenon known as homeostatic synaptic plasticity. Saliba et al. [64] showed that a chronic blockade of L-type VGCCs in rat hippocampal neurons decreases the number of GABA A receptors at the synapses, in parallel with a reduction in the amplitude of mIPSCs. Their results also indicate that calcium influx through L-type VGCCs regulates the insertion of newly translated GABA A receptor subunits into the plasma membrane by a mechanism that involves the activity of the proteasome. ...
Article
Full-text available
GABAA receptors are pentameric ion channels that mediate most synaptic and tonic extrasynaptic inhibitory transmissions in the central nervous system. There are multiple GABAA receptor subtypes constructed from 19 different subunits in mammals that exhibit different regional and subcellular distributions and distinct pharmacological properties. Dysfunctional alterations of GABAA receptors are associated with various neuropsychiatric disorders. Short- and long-term plastic changes in GABAA receptors can be induced by the activation of different intracellular signaling pathways that are triggered, under physiological and pathological conditions, by calcium entering through voltage-gated calcium channels. This review discusses several mechanisms of regulation of GABAA receptor function that result from the activation of L-type voltage gated calcium channels. Calcium influx via these channels activates different signaling cascades that lead to changes in GABAA receptor transcription, phosphorylation, trafficking, and synaptic clustering, thus regulating the inhibitory synaptic strength. These plastic mechanisms regulate the interplay of synaptic excitation and inhibition that is crucial for the normal function of neuronal circuits.
... Given that calcium influx can be mediated through the VGCCs and NMDA receptors (Catterall, 2011;Chen et al. 1999;Leinekugel et al. 1997), we first utilized various drugs in neurons to block calcium influx. Both zinc spikes and calcium spikes vanished when neurons were exposed to the VGCC inhibitor (Helton et al. 2005;Saliba et al. 2009), nifedipine (Figure 3b). Furthermore, we tested if NMDA receptors on the postsynaptic membrane are responsible for maintaining zinc spikes in neurons. ...
Article
Full-text available
Zinc has been suggested to act as an intracellular signaling molecule due to its regulatory effects on numerous protein targets including enzymes, transcription factors, ion channels, neurotrophic factors, and postsynaptic scaffolding proteins. However, intracellular zinc concentration is tightly maintained at steady levels under natural physiological conditions. Dynamic changes in intracellular zinc concentration have only been detected in certain types of cells that are exposed to pathologic stimuli or upon receptor ligand binding. Unlike calcium, the ubiquitous signaling metal ion that can oscillate periodically and spontaneously in various cells, spontaneous zinc oscillations have never been reported. In this work, we made the novel observation that the developing neurons generated spontaneous and synchronous zinc spikes in primary hippocampal cultures using a fluorescent zinc sensor, FluoZin‐3. Blocking of glutamate receptor‐dependent calcium influx depleted the zinc spikes, suggesting that these zinc spikes were driven by the glutamate‐mediated spontaneous neural excitability and calcium spikes that have been characterized in early developing neurons. Simultaneous imaging of calcium or pH together with zinc, we uncovered that a downward pH spike was evoked with each zinc spike and this transient cellular acidification occurred downstream of calcium spikes but upstream of zinc spikes. Our results suggest that spontaneous, synchronous zinc spikes were generated through calcium influx‐induced cellular acidification, which liberates zinc from intracellular zinc binding ligands. Given that changes in zinc concentration can modulate activities of proteins essential for synapse maturation and neuronal differentiation, these zinc spikes might act as important signaling roles in neuronal development. image
Article
The problem of frequency coding is closely related to the studies of inhibitory transmission as a factor of neural network plasticity. The rewiew presents basic mechanisms of inhibitory control of spatio-temporal pattern of neural activity during signal processing. Current views are analyzed in respect of dynamic synapses, their instability and variation within the ongoing activity. The results presented here demonstrate that short-term plasticity operates with the combined contribution of excitatory and inhibitory synapses. The role of GABAergic potentials in modulation of intracellular messenger’s activity is discussed, including those implicated in postsynaptic modifications of excitatory and inhibitory transmission. The main topics concerning the molecular mechanisms centered on the lateral diffusion of GABAA receptors. The data of many reports argue for coordinating role of actin cytoskeleton. It is proposed that postsynaptic mechanisms underlying GABAA plasticity may be activated in result of fast adaptation of actin cytoskeleton and associated proteins to disbalance between excitation and inhibition.
Preprint
GABAergic transmission is influenced by post-translational modifications, like phosphorylation, impacting channel conductance, allosteric modulator sensitivity, and membrane trafficking. O-GlcNAcylation is a post-translational modification involving the O-linked attachment of β-N-acetylglucosamine on serine/threonine residues. Previously we reported an acute increase in O-GlcNAcylation elicits a long-term depression of evoked GABAAR inhibitory post synaptic currents (eIPSCs) onto hippocampal principal cells. Importantly O-GlcNAcylation and phosphorylation can co-occur or compete for the same residue; whether they interact in modulating GABAergic IPSCs is unknown. We tested this by recording IPSCs from hippocampal principal cells and pharmacologically increased O-GlcNAcylation, before or after increasing serine phosphorylation using the adenylate cyclase activator, forskolin. Although forskolin had no significant effect on baseline eIPSC amplitude, we found that a prior increase in O-GlcNAcylation unmasks a forskolin-dependent increase in eIPSC amplitude, reversing the O-GlcNAc-induced eIPSC depression. Inhibition of adenylate cyclase or protein kinase A did not prevent the potentiating effect of forskolin, indicating serine phosphorylation is not the mechanism. Surprisingly, increasing O-GlcNAcylation also unmasked a potentiating effect of the neurosteroids 5α-pregnane-3α,21-diol-20-one (THDOC) and progesterone on eIPSC amplitude, mimicking forskolin. Our findings show under conditions of heightened O-GlcNAcylation, the neurosteroid site on synaptic GABAARs is accessible to agonists, permitting strengthening of synaptic inhibition.
Chapter
Calcium entering via voltage-gated calcium channels is a second messenger key for cellular functions including gene expression, neuronal excitability, neurogenesis, neuronal differentiation, and neurotransmitter release. Alterations in these calcium-dependent processes have been observed in psychiatric disorders. Furthermore, genetic studies have identified associations of risk genetic variants of voltage-gated calcium channel genes (CACNA1A-I) with psychiatric disorders. Thus, these channels are becoming promising targets to treat these pathologies. In this chapter, we will discuss evidence linking calcium to psychiatric disorders, and then we will review key genetic studies that have found strong associations among voltage-gated calcium channel genes and psychiatric disorders. Next, we will examine the role of voltage-gated calcium channels in neurobiological mechanisms linked to psychiatric disorders. We will analyze evidence from animal models that link voltage-gated to behavioral endophenotypes observed in psychiatric disorders. Finally, we will discuss the current view and challenges to target these channels to treat psychiatric disorders.KeywordsVoltage-gated calcium channelsPsychiatric disordersCalciumBehaviorNeurodevelopmental disordersEndophenotypesNeurotransmitter releaseNeurogenesisExcitation/inhibition balancePolymorphismRisk genetic variants
Article
Defects in interneuron migration can disrupt the assembly of cortical circuits and lead to neuropsychiatric disease. Using forebrain assembloids derived by integration of cortical and ventral forebrain organoids, we have previously discovered a cortical interneuron migration defect in Timothy syndrome (TS), a severe neurodevelopmental disease caused by a mutation in the L-type calcium channel (LTCC) Cav1.2. Here, we find that acute pharmacological modulation of Cav1.2 can regulate the saltation length, but not the frequency, of interneuron migration in TS. Interestingly, the defect in saltation length is related to aberrant actomyosin and myosin light chain (MLC) phosphorylation, while the defect in saltation frequency is driven by enhanced γ-aminobutyric acid (GABA) sensitivity and can be restored by GABA-A receptor antagonism. Finally, we describe hypersynchronous hCS network activity in TS that is exacerbated by interneuron migration. Taken together, these studies reveal a complex role of LTCC function in human cortical interneuron migration and strategies to restore deficits in the context of disease.
Article
The recently identified molecule P7C3 has been highlighted in the field of pain research. We examined the effect of intrathecal P7C3 in tissue injury pain evoked by formalin injection and determined the role of the GABA system in the activity of P7C3 at the spinal level. Male Sprague-Dawley rats with intrathecal catheters implanted for experimental drug delivery were studied. The effects of intrathecal P7C3 and nicotinamide phosphoribosyltransferase (NAMPT) administered 10 min before the formalin injection were examined. Animals were pretreated with bicuculline, a GABA-A receptor antagonist; saclofen, a GABA-B receptor antagonist; L-allylglycine, a glutamic acid decarboxylase (GAD) blocker; and CHS 828, an NAMPT inhibitor; to observe involvement in the effects of P7C3. The effects of P7C3 alone and the mixture of P7C3 with GABA receptor antagonists on KCl-induced calcium transients were examined in rat dorsal root ganglion (DRG) neurons. The expression of GAD and the concentration of GABA in the spinal cord were evaluated. Intrathecal P7C3 and NAMPT produced an antinociceptive effect in the formalin test. Intrathecal bicuculline, saclofen, L-allylglycine, and CHS 828 reversed the antinociception of P7C3 in both phases. P7C3 decreased the KCl-induced calcium transients in DRG neurons. Both bicuculline and saclofen reversed the blocking effect of P7C3. The levels of GAD expression and GABA concentration decreased after formalin injection and were increased by P7C3. These results suggest that P7C3 increases GAD activity and then increases the GABA concentration in the spinal cord, which in turn may act on GABA receptors causing the antinociceptive effect against pain evoked by formalin injection.
Article
Full-text available
GABAA receptors are believed to be pentameric hetero-oligomers, which can be constructed from six subunits (alpha, beta, gamma, delta, epsilon, and rho) with multiple members, generating a large potential for receptor heterogeneity. The mechanisms used by neurons to control the assembly of these receptors, however, remain unresolved. Using Semliki Forest virus expression we have analyzed the assembly of 9E10 epitope-tagged receptors comprising alpha1 and beta2 subunits in baby hamster kidney cells and cultured superior cervical ganglia neurons. Homomeric subunits were retained within the endoplasmic reticulum, whereas heteromeric receptors were able to access the cell surface in both cell types. Sucrose density gradient fractionation demonstrated that the homomeric subunits were incapable of oligomerization, exhibiting 5 S sedimentation coefficients. Pulse-chase analysis revealed that homomers were degraded, with half-lives of approximately 2 hr for both the alpha1((9E10)) and beta2((9E10)) subunits. Oligomerization of the alpha1((9E10)) and beta2((9E10)) subunits was evident, as demonstrated by the formation of a stable 9 S complex, but this process seemed inefficient. Interestingly the appearance of cell surface receptors was slow, lagging up to 6 hr after the formation of the 9 S receptor complex. Using metabolic labeling a ratio of alpha1((9E10)):beta2((9E10)) of 1:1 was found in this 9 S fraction. Together the results suggest that GABAA receptor assembly occurs by similar mechanisms in both cell types, with retention in the endoplasmic reticulum featuring as a major control mechanism to prevent unassembled receptor subunits accessing the cell surface.
Article
Full-text available
Information is stored in neural circuits through long-lasting changes in synaptic strengths. Most studies of information storage have focused on mechanisms such as long-term potentiation and depression (LTP and LTD), in which synaptic strengths change in a synapse-specific manner. In contrast, little attention has been paid to mechanisms that regulate the total synaptic strength of a neuron. Here we describe a new form of synaptic plasticity that increases or decreases the strength of all of a neuron's synaptic inputs as a function of activity. Chronic blockade of cortical culture activity increased the amplitude of miniature excitatory postsynaptic currents (mEPSCs) without changing their kinetics. Conversely, blocking GABA (gamma-aminobutyric acid)-mediated inhibition initially raised firing rates, but over a 48-hour period mESPC amplitudes decreased and firing rates returned to close to control values. These changes were at least partly due to postsynaptic alterations in the response to glutamate, and apparently affected each synapse in proportion to its initial strength. Such 'synaptic scaling' may help to ensure that firing rates do not become saturated during developmental changes in the number and strength of synaptic inputs, as well as stabilizing synaptic strengths during Hebbian modification and facilitating competition between synapses.
Article
Full-text available
Changes in synaptic efficacy are essential for neuronal development, learning and memory formation and for pathological states of neuronal excitability, including temporal-lobe epilepsy. At synapses, where there is a high probability of opening of postsynaptic receptors, all of which are occupied by the released transmitter, the most effective means of augmenting postsynaptic responses is to increase the number of receptors. Here we combine quantal analysis of evoked inhibitory postsynaptic currents with quantitative immunogold localization of synaptic GABA(A) receptors in hippocampal granule cells in order to clarify the basis of inhibitory synaptic plasticity induced by an experimental model of temporal-lobe epilepsy (a process known as kindling). We find that the larger amplitude (66% increase) of elementary synaptic currents (quantal size) after kindling results directly from a 75% increase in the number of GABA(A) receptors at inhibitory synapses on somata and axon initial segments. Receptor density was up by 34-40% and the synaptic junctional area was expanded by 31%. Presynaptic boutons were enlarged, which may account for the 39% decrease in the average number of released transmitter packets (quantal content). Our findings establish the postsynaptic insertion of new GABA(A) receptors and the corresponding increase in postsynaptic responses augmenting the efficacy of mammalian inhibitory synapses.
Article
Full-text available
Most fast inhibitory neurotransmission in the brain is mediated by GABAA receptors, which are mainly postsynaptic and consist of diverse alpha and beta subunits together with the gamma 2 subunit. Although the gamma 2 subunit is not necessary for receptor assembly and translocation to the cell surface, we show here that it is required for clustering of major postsynaptic GABAA receptor subtypes. Loss of GABAA receptor clusters in mice deficient in the gamma 2 subunit, and in cultured cortical neurons from these mice, is paralleled by loss of the synaptic clustering molecule gephyrin and synaptic GABAergic function. Conversely, inhibiting gephyrin expression causes loss of GABAA receptor clusters. The gamma 2 subunit and gephyrin are thus interdependent components of the same synaptic complex that is critical for postsynaptic clustering of abundant subtypes of GABAA receptors in vivo.
Article
Full-text available
GABA(A) receptors are critical mediators of fast synaptic inhibition in the brain, and the predominant receptor subtype in the central nervous system is believed to be a pentamer composed of alpha, beta, and gamma subunits. Previous studies on recombinant receptors have shown that protein kinase C (PKC) and PKA directly phosphorylate intracellular serine residues within the receptor beta subunit and modulate receptor function. However, the relevance of this regulation for neuronal receptors remains poorly characterized. To address this critical issue, we have studied phosphorylation and functional modulation of GABA(A) receptors in cultured cortical neurons. Here we show that the neuronal beta3 subunit is basally phosphorylated on serine residues by a PKC-dependent pathway. PKC inhibitors abolish basal phosphorylation, increasing receptor activity, whereas activators of PKC enhance beta3 phosphorylation with a concomitant decrease in receptor activity. PKA activators were shown to increase the phosphorylation of the beta3 subunit only in the presence of PKC inhibitors. We also show that the main sites of phosphorylation within the neuronal beta3 subunit are likely to include Ser-408 and Ser-409, residues that are important for the functional modulation of beta3-containing recombinant receptors. Furthermore, PKC activation did not change the total number of GABA(A) receptors in the plasma membrane, suggesting that the effects of PKC activation are on the gating or conductance of the channel. Together, these results illustrate that cell-signaling pathways that activate PKC may have profound effects on the efficacy of synaptic inhibition by directly modulating GABA(A) receptor function.
Article
Homeostatic synaptic scaling is a form of synaptic plasticity that adjusts the strength of all of a neuron's excitatory synapses up or down to stabilize firing. Current evidence suggests that neurons detect changes in their own firing rates through a set of calcium-dependent sensors that then regulate receptor trafficking to increase or decrease the accumulation of glutamate receptors at synaptic sites. Additional mechanisms may allow local or network-wide changes in activity to be sensed through parallel pathways, generating a nested set of homeostatic mechanisms that operate over different temporal and spatial scales.
Article
Long-term potentiation (LTP) of excitatory synaptic transmission could be a mechanism underlying memory. Induction of LTP requires Ca2+ influx into postsynaptic neurons through ion channels gated by NMDA (N-methyl-D-aspartate) receptors in hippocampus (area CA1 and dentate gyrus) and neocortex. Here we report that a component of LTP not requiring the activation of NMDA receptors can be induced in area CA1. The component is dependent on tetanus frequency, requires increases in postsynaptic intracellular Ca2+ concentrations, and is suppressed by an antagonist of voltage-dependent Ca2+ channels.
Article
The competitive nicotinic antagonist d-[3H]tubocurarine was used as a photoaffinity label for the acetylcholine binding sites on the nicotinic acetylcholine receptor (AcChoR) from Torpedo. Irradiation with 254-nm UV light of AcChoR-rich membranes equilibrated with d-[3H]tubocurarine resulted in covalent incorporation into the alpha, gamma, and delta subunits that could be blocked by alpha-bungarotoxin or by carbamoylcholine. The concentrations of d-[3H]tubocurarine required for half-maximal specific incorporation into the gamma and delta subunits were 40 nM and 0.9 microM, respectively, consistent with the dissociation constants for the high- and low-affinity binding sites (Kd = 35 nM and 1.2 microM). The concentration dependence of incorporation into alpha subunit was biphasic and consistent with labeling of both the high- and low-affinity d-tubocurarine binding sites. The specific photolabeling of each AcChoR subunit was inhibited by carbamoylcholine with appropriate dose dependence. These results establish that, in addition to the alpha subunits, the gamma and delta subunits also contribute directly to the acetylcholine binding sites and that each binding site is at an interface of subunits. Because the AcChoR subunits are homologous and are arranged pseudosymmetrically about a central axis, the photolabeling results are inconsistent with an arrangement of subunits in the AcChoR rosette of alpha beta alpha gamma delta and indicate that either the gamma or delta subunit resides between the alpha subunits.
Article
In many neurons, responses to individual quanta of transmitter exhibit large variations in amplitude. The origin of this variability, although central to our understanding of synaptic transmission and plasticity, remains controversial. To examine the relationship between quantal amplitude and postsynaptic receptor number, we adopted a novel approach, combining patch-clamp recording of synaptic currents with quantitative immunogold localization of synaptic receptors. Here, we report that in cerebellar stellate cells, where variability in GABA miniature synaptic currents is particularly marked, the distribution of quantal amplitudes parallels that of synaptic GABA(A) receptor number. We also show that postsynaptic GABA(A) receptor density is uniform, allowing synaptic area to be used as a measure of relative receptor content. Flurazepam, which increases GABA(A) receptor affinity, prolongs the decay of all miniature currents but selectively increases the amplitude of large events. From this differential effect, we show that a quantum of GABA saturates postsynaptic receptors when <80 receptors are present but results in incomplete occupancy at larger synapses.